Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

ch1.

1
Introduction

1.1
Principles of Heterogeneous Catalysis1 hidden or unknown variables. Accordingly, principles
of heterogeneous catalysis are typically formulated from
James A. Dumesic∗ , George W. Huber and Michel Boudart studies of model catalysts in ideal reactors with simplified
reactants under mild pressure conditions (e.g., 1 bar),
rather than from catalytic performance data obtained with
1.1.1 commercial catalysts in complex reactors using mixed
Introduction feed streams under industrial reaction conditions. The
principles derived from these more simplified studies
Heterogeneous catalysis is of vital importance to the advance the science of heterogeneous catalysis, and they
world’s economy, allowing us to convert raw materi- guide the researcher, inventor, and innovator of new
als into valuable chemicals and fuels in an economical, catalysts and catalytic processes.
efficient, and environmentally benign manner. For ex-
ample, heterogeneous catalysts have numerous industrial 1.1.2
applications in the chemical, food, pharmaceutical, au- Definitions of Catalysis and Turnover
tomobile and petrochemical industries [1–5], and it has
been estimated that 90% of all chemical processes use het- The definition of a catalyst has been discussed many
erogeneous catalysts [6]. Heterogeneous catalysis is also times [19]. For example, a catalyst is a material that
finding new applications in emerging areas such as fuel converts reactants into products, through a series of
cells [7–9], green chemistry [10–12], nanotechnology [13], elementary steps, in which the catalyst participates while
and biorefining/biotechnology [14–18]. Indeed, contin- being regenerated to its original form at the end of each
ued research into heterogeneous catalysis is required to cycle during its lifetime. A catalyst changes the kinetics
allow us to address increasingly complex environmental of the reaction, but does not change the thermodynamics.
and energy issues facing our industrialized society. Another definition is that a catalyst is a substance
Discussing the principles of heterogeneous catalysis is that transforms reactants into products, through an
difficult, because catalysts are used for a wide range of uninterrupted and repeated cycle of elementary steps
applications, involving a rich range of surface chemistries. in which the catalyst participates while being regenerated
Moreover, the field of heterogeneous catalysis is highly to its original form at the end of each cycle during its
interdisciplinary in nature, requiring the cooperation lifetime [20].
between chemists and physicists, between surface The main advantage of using a heterogeneous catalyst
scientists and reaction engineers, between theorists and is that, being a solid material, it is easy to separate from
experimentalists, between spectroscopists and kineticists, the gas and/or liquid reactants and products of the overall
and between materials scientists involved with catalyst catalytic reaction. The heart of a heterogeneous catalyst
synthesis and characterization. Furthermore, industrial involves the active sites (or active centers) at the surface
catalysts are complex materials, with highly optimized of the solid. The catalyst is typically a high-surface area
chemical compositions, structures, morphologies, and material (e.g., 10–1000 m2 g−1 ), and it is usually desirable
pellet shapes; moreover, the physical and chemical to maximize the number of active sites per reactor volume.
characteristics of these materials may depend on Identifying the reaction intermediates – and hence the

1 A list of abbreviations/acronyms used in the text is provided at the


end of the chapter.
∗ Corresponding author. References see page 14
ch1.1

2 1.1 Principles of Heterogeneous Catalysis

mechanism – for a heterogeneous catalytic reaction is has been observed quantitatively at temperatures as low
often difficult, because many of these intermediates are as 78 K and as high as 1500 K; at pressures between
difficult to detect using conventional methods (e.g., gas 10−9 and 103 bar; with reactants in the gas phase or in
chromatography or mass spectrometry) because they do polar or non-polar solvents; with or without assistance
not desorb at significant rates from the surface of the of photons, radiation or electron transfer at electrodes;
catalyst (especially for gas-phase reactions). with pure metals as unreactive as gold and as reactive as
Heterogeneous catalysts typically contain different sodium; with multicomponent and multiphase inorganic
types of surface sites, because crystalline solids exhibit compounds and acidic organic polymers; and at STYs
crystalline anisotropy. Equilibrated single crystals expose as low as 10−5 s−1 (one turnover per day) and as high
different faces with different atomic structures so as 109 s−1 (gas kinetic collision rate at 10 bar). TOFs of
as to minimize total surface energy. It would be commonly used heterogeneous catalysts are commonly
surprising, in fact, if different crystallographic planes on the order of one per second. The life of the catalyst can
exposing sites with different coordination environments be defined as the number of turnovers observed before
possessed identical properties for chemisorption and the catalyst ceases to operate at an acceptable rate. Clearly,
catalytic reactions. Moreover, most catalytic solids are this number must be larger than unity, otherwise the
polycrystalline. Furthermore, in order to achieve high substance used is not a catalyst but a reagent. Catalyst
surface areas, most catalysts contain particles with sizes life can either be short, as in catalytic cracking of oil, or
in the nanometer length scale. The surfaces of these very long, corresponding to as many as 109 turnovers in
nanoscopic particles contain sites associated with terraces, ammonia synthesis.
edges, kinks, and vacancies [21]. If the catalyst contains
more than one component (as is generally the case), the 1.1.3
surface composition may be different from that of the bulk Steps in a Heterogeneous Catalytic Reaction
and differently so for each exposed crystallographic plane.
Solids normally contain defects of electronic or atomic During an overall catalytic reaction, the reactants
nature; in addition, they contain impurities which are and products undergo a series of steps over the
either known or unknown in the bulk, but are mostly catalyst, including:
unknown at the surface. Finally, the surface atomic
structure and composition may change with time-on- 1. Diffusion of the reactants through a boundary layer
stream as the catalytic reaction proceeds. In short, it surrounding the catalyst particle.
is normal to expect that a catalytic surface exposes a 2. Intraparticle diffusion of the reactants into the catalyst
variety of surface sites, in contrast to displaying a single pores to the active sites.
type of active site. Indeed, it is so normal today to 3. Adsorption of the reactants onto active sites.
expect such complexity that it seems surprising that, 4. Surface reactions involving formation or conversion
in 1925, when Taylor formulated his principle of active of various adsorbed intermediates, possibly including
sites or active centers, the report created so much surface diffusion steps.
attention and remains one of the most often cited in 5. Desorption of products from catalyst sites.
heterogeneous catalysis [22]. The relative importance of 6. Intraparticle diffusion of the products through the
surface structure – as influenced by crystalline anisotropy, catalyst pores.
surface defects, and surface composition – underlines the 7. Diffusion of the products across the boundary layer
difficulty of identifying the active sites, either simple surrounding the catalyst particle.
or complex, that are responsible for turning over the
catalytic cycle. The identification and counting of active Accordingly, different regimes of catalytic rate control
sites in heterogeneous catalysis became the ‘‘Holy Grail’’ can exist, including: (i) film diffusion control (Steps 1
of heterogeneous catalysis in 1925, and the situation and 7); (ii) pore diffusion control (Steps 2 and 6); and
remains the same today. (iii) intrinsic reaction kinetics control (Steps 3 to 5) of
The activity of a catalyst is defined by the number of catalyst performance. In addition to mass transfer effects,
revolutions of the catalytic cycle per unit time, given in heat transfer effects can also occur in heterogeneous
units of turnover rate (TOR) or turnover frequency (TOF). catalysis for highly exothermic or endothermic reactions
In cases where the rate is not uniform within the catalytic (especially in combustion or steam reforming).
reactor or within the catalyst pellets, it is useful to report Figure 1 shows a general effect of temperature on
the rate as a site time yield (STY), defined as the overall rate the reaction rate for a heterogeneous catalyst. At low
of the catalytic reaction within the reactor normalized by temperatures, diffusion through the film and pores is
the total number of active sites within the reactor, again fast compared to rates of surface reactions, and the
in units of reciprocal time. Catalysis by solid materials overall reaction rate is controlled by the intrinsic reaction
ch1.1

1.1.4 Desired Characteristics of a Catalyst 3

Slope = 3-5 kJ/mol


of catalysts, and we refer the reader to other articles for
further discussion on transport effects in heterogeneous
Slope = Ea/2R
catalysis [23–31].
Film diffusion
controlled regime
1.1.4
ln (rate)

Pore diffusion
Slope = Ea/R Desired Characteristics of a Catalyst
controlled regime

The following list provides several of the key attributes of


a good catalyst:
Intrinsic
regime
Increasing
• The catalyst should exhibit good selectivity for produc-
Temperature tion of the desired products and minimal production of
undesirable byproducts.
1/ Temperature • The catalyst should achieve adequate rates of reaction
at the desired reaction conditions of the process
Fig. 1 General effects of temperature on catalytic activity. The (remembering that achieving good selectivity is usually
intrinsic activation energy is equal to Ea , and R is the gas constant. more important than achieving high catalytic activity).
• The catalyst should show stable performance at reaction
conditions for long periods of time, or it should be
kinetics. As the temperature is increased, the rates of possible to regenerate good catalyst performance by
surface reactions typically increase more rapidly than the appropriate treatment of the deactivated catalyst after
rates of diffusion, and the overall rate of the catalytic short periods.
process becomes controlled by intraparticle diffusion. • The catalyst should have good accessibility of reactants
The apparent activation energy in this regime is equal and products to the active sites such that high rates can
to the intrinsic activation energy divided by two. As the be achieved per reactor volume.
temperature is increased further, mass transfer through
the external boundary layer becomes the controlling The first three key attributes of a good catalyst are
step. The onset of diffusion limited regimes can be influenced primarily by the interactions of the catalyst
altered by changing the reactor design, the catalyst pore surface with the reactants, products, and intermediates
structure, the catalyst particle size, and the distribution of the catalytic process. In addition, other species may
of the active sites in the catalyst particles. Values form on the catalyst surface (e.g., hydrogen-deficient
of various dimensionless groups can be calculated to carbonaceous deposits denoted as coke) that are not
estimate the extents to which transport phenomena directly part of the reaction scheme (or mechanism) for
may control catalytic performance for specific operating the overall catalytic process.
conditions [23–31]; however, these calculations are most The principle of Sabatier states that a good heteroge-
reliable for cases where the intrinsic reaction kinetics are neous catalyst is a material that exhibits an intermediate
known. In these cases, it is possible to make catalysts strength of interaction with the reactants, products, and
with structures designed to provide adequate rates of intermediates of the catalytic process [37, 38]. Interactions
diffusion and yet offering high surfaces areas, leading of the catalyst surface with the various adsorbed species
to high rates of reaction per reactor volume, such as of the reaction mechanism that are too weak lead to
the design of specific pore size distributions (e.g., bi- high activation energies for surface reactions and thus
modal distributions containing large pores leading to low catalytic activity, whereas interactions of the catalyst
high accessibility of the active sites within the interior of with adsorbed species that are too strong lead to excessive
the catalytic pellet, and small pores that branch from the blocking of surface sites by these adsorbed species, again
larger pores leading to high surface areas), the formulation leading to low catalytic activity.
of unique pellet shapes (that lead to high accessibility of The principle of Sabatier is elegant in its simplicity
the active sites but do not cause large pressure drops and generality, but it is deceptively difficult to use in
through the catalytic reactor), and the synthesis of catalyst practice. In particular, this principle applies to a catalyst
pellets containing a spatial distribution of the active in its working state, and the nature of the catalyst surface
material within the catalyst pellet [32]. In some cases, can be expected to be dependent on the nature of the
transport effects can be used to improve the selectivity of catalytic reaction conditions. For example, one may begin
a catalyst, such as in the case of shape-selective catalysis the catalytic reaction with the heterogeneous catalyst
in zeolites [33–36]. In the following sections, we focus on
various factors controlling the intrinsic reaction kinetics References see page 14
ch1.1

4 1.1 Principles of Heterogeneous Catalysis

in a given oxidation state (e.g., containing zero-valent of sites producing a prior non-uniformity, and effects
metal particles following treatment of the catalyst in of surface coverage causing induced non-uniformity).
H2 at elevated temperature); however, the nature of the At a selected set of reaction conditions, an optimal
surface can be changed dramatically upon interaction set of surface binding energies exists that satisfy the
with strongly adsorbed species, such as the formation of principle of Sabatier (as discussed below). Accordingly,
carbonaceous deposits (coke), and formation of oxides, the performance of a heterogeneous catalyst with a
carbides, nitrides, or sulfides upon interaction with O, non-uniform surface will be dominated by the subset
C, N, or S species, respectively [39–44]. In this case, of the sites having surface binding energies closest to
the interactions of these oxide, carbide, nitride, or the optimal values. At higher temperatures, other sites
sulfide surfaces with the adsorbed species enter into having stronger binding energies with adsorbed species
the reaction mechanism. Of even greater complexity will become the dominant contributors to the observed
is the fact that a variety of different types of sites are catalytic activity, whereas sites having weaker binding
typically present on a catalyst surface (e.g., sites having energies with adsorbed species will control catalytic
different coordination and/or chemical composition), and activity at lower temperatures. Thus, while the effects of
a majority of the observed catalytic activity may be surface non-uniformity make it more difficult to predict
caused by the contributions from a small fraction of the performance of a heterogeneous catalyst from a
the sites present on the catalyst surface. In this case, molecular-level understanding, these effects may serve
the adsorbed species interact with these special surface to broaden the range of reaction conditions over which
sites (e.g., steps and defect sites on a metal nanoparticle, the catalyst can operate effectively. In this respect, our
or sites present at the metal–support interface of a desire to design catalysts having very high selectivity
supported metal catalyst). Another factor that complicates is guided by the synthesis of uniform catalysts, where
catalyst design is that the strengths of interaction of each site has the optimal properties for production of
the surface with adsorbed species typically depend on the desired reaction product. This strategy leads to the
the surface coverages by adsorbed species. For example, idea of highly selective, single-site catalysis as discussed
the interaction of a transition metal surface with adsorbed by Thomas et al. [51]. In contrast, the design of catalysts
CO may be very strong at low surface coverages (e.g., that operate over a wide range of reaction conditions is
binding energy of nearly 200 kJ mol−1 ), suggesting that guided by the synthesis of non-uniform catalysts, such
these surfaces would be completely covered and thus that different subsets of sites control catalyst performance
poisoned by adsorbed CO at moderate pressures and at different reaction conditions. The disadvantage of using
temperatures; however, these surfaces may carry out non-uniform catalysts, however, is that different sites may
catalytic reactions in the presence of gaseous CO at these display different selectivities for the production of various
pressures and temperatures because the differential heat products, and control over catalytic selectivity may thus
of CO adsorption decreases significantly (e.g., to binding be limited [22].
energies near 100 kJ mol−1 ) as the surface coverage by
adsorbed CO increases [45, 46]. Accordingly, there is a 1.1.5
relationship between activity and the interaction of the Reaction Schemes and Adsorbed Species
surface with adsorbed species at the surface coverage
regime appropriate for the catalytic reaction conditions. We now explore further the principle of Sabatier using
The aforementioned complications caused by the a specific example: water-gas shift over a metal catalyst
presence of different types of sites on the surface, and (e.g., Cu). This reaction (CO + H2 O ← −−−−−
→− CO2 + H2 ) is
the effects on the surface binding energies caused by of importance for the production of H2 from steam-
changes in surface coverages, clearly make it difficult to reforming of fossil fuels, and for controlling the CO : H2
interpret the performance of a heterogeneous catalyst ratio in synthesis gas mixtures used in methanol and
in quantitative detail. Tools are certainly available to Fischer–Tropsch synthesis processes. For this example,
address these complications, such as kinetic Monte we consider the reaction scheme shown in Fig. 2, where
Carlo simulations combined with results from density * represents a surface site. The stoichiometric numbers,
functional theory (DFT) calculations [47–50]. Yet, from σi,1 and σi,2 , indicate the number of times that step i
a different point of view, the presence of different occurs to give the overall reaction for reaction schemes 1
types of sites and the effects of surface coverage may and 2, respectively.
well contribute to the robustness of the heterogeneous In this sequence of steps, the water-gas shift reaction
catalyst for operation over a wide range of reaction can take place via the formation of carboxyl species
conditions. In general, the presence of different types (COOH) or through the formation of formate species
of sites and the effects of surface coverage both (HCOO) [52]. In the absence of a catalyst, the rate of water-
contribute to surface non-uniformity (different types gas shift via this mechanism is negligible, because the
ch1.1

1.1.5 Reaction Schemes and Adsorbed Species 5

si,1 si,2 does not participate in the overall reaction. In this latter
1. CO + ∗ CO∗ 1 1 case, the spectator species inhibits the overall reaction
by blocking surface sites, and it serves no useful role in
2. H2O +∗ H2O∗ 1 1
the overall reaction scheme. For purposes of elucidating
3. H2O ∗ + ∗ OH ∗+H ∗ 2 1 catalytic reaction schemes it is essential to distinguish
between reactive intermediates and spectator species. This
4. CO ∗ + OH ∗ COOH ∗+ ∗ 1 0 distinction is of paramount importance in spectroscopic
CO2∗ + H2O ∗
studies of adsorbed species on catalyst surfaces, where
5. COOH ∗ + OH ∗ 1 0
the detection of a specific adsorbed species using a
6. CO2∗ CO2 + ∗ 1 1 spectroscopic method (e.g., the detection by infra-red
(IR) studies of adsorbed ethylidyne species on platinum
7. 2H ∗ H2 + 2∗ 1 1
surfaces during ethylene hydrogenation [53]) does not
8. CO ∗ + OH ∗ HCOO ∗∗ 0 1 guarantee that this species is a reactive intermediate.
Instead, these spectroscopic studies must be conducted
9. HCOO∗∗ CO2∗ + H ∗ 0 1 under dynamic conditions (e.g., so-called operando
CO + H2O CO2 + H2
measurements, where spectroscopic and reaction kinetics
overall
data are collected simultaneously) to determine that
the time constant for the formation or disappearance
Fig. 2 Assumed reaction mechanism for water-gas shift reaction
over Cu. Adapted from Ref. [52]. of the surface species is the same as the time constant for
the overall catalytic reaction [54, 55].
The overall catalytic reaction is given by a linear
reaction intermediates (e.g., OH∗ , H∗ , COOH∗ , HCOO∗ ) combination of elementary steps, and the enthalpy change
are at very low concentrations in the gas phase. For for the overall reaction, H , is given by:
example, the enthalpy change for step 3 in the gas phase

is approximately 500 kJ mol−1 . However, adsorption of H = σi Hi (1)
the reaction intermediates onto the catalyst surface allows i
these steps to take place with small enthalpy changes. In
the case of copper, the binding energies of H and OH where Hi are the enthalpy changes for elementary
are approximately 250 and 280 kJ mol−1 on Cu(111), such steps i. From the principle of Sabatier, it is now clear
that the enthalpy change for step 3 on the catalyst surface that the overall value of H should be composed of
is now slightly exothermic. According to the principle of approximately equal contributions from each of the
Sabatier, a good catalyst is a material that adsorbs reaction values of Hi , giving rise to a relatively flat potential
intermediates with intermediate strength. However, we energy diagram of energy versus reaction coordinate in
now must distinguish between reactive intermediates and moving from reactants, through adsorbed intermediates,
spectator species on the catalyst surface. to products. Specifically, any value of Hj that is very
In the above reaction scheme, we see that the water- negative must be balanced by a value of Hk that is
gas shift reaction can take place through adsorbed very positive, such that the surface will become highly
carboxyl species or formate species. Results from DFT covered (and poisoned) by the adsorbed species produced
calculations indicate that adsorbed formate species have in step j , and the activation energy for step k will be high.
lower energy compared to adsorbed carboxyl species on Both of these effects lead to low catalytic activity. We note
copper surfaces, suggesting that path 2 for water-gas that the reaction mechanism can certainly contain steps
shift (σi,2 ) would be favored versus path 1 (σi,1 ) based with positive values of Hi , because the intermediates
on thermodynamic arguments. However, the activation produced in such a step can be consumed by following
energy for step 4 is considerably lower than that for steps having negative values of Hi . This situation is
step 8, and the primary path for water-gas shift over termed ‘‘kinetic coupling’’, where the conversion of an
copper involves the formation and subsequent reaction of unfavorable step is increased by its combination with a
adsorbed carboxyl species. Accordingly, the most stable favorable step that consumes the unstable products of the
species on the catalyst surface are not necessarily the first step. The highest value of Hi for a surface reaction
most reactive species. This idea leads us to distinguish that can be tolerated can be estimated from transition
between a most abundant surface intermediate (MASI) and state theory. The value of Hi,max depends on the overall
a most abundant reactive intermediate (MARI). In certain rate of the reaction (TOF), the surface coverage by the
cases, the MASI and the MARI may be the same species, surface species that reacts in this step (θA ), a frequency
but in other cases (such as in this case of water-gas
shift on copper), the MASI is a spectator species that References see page 14
ch1.1

6 1.1 Principles of Heterogeneous Catalysis

factor ν (of the order of 1013 s−1 ), and the temperature T , for the reaction conditions of interest. Unfortunately,
as given by: at the outset of research on a given catalyst process,
  we usually do not know which rate constants will be
Hi,max kinetically significant. Accordingly, an important objective
TOF = ν exp − θA (2)
RT of research into a given catalytic process is to identify
which steps are kinetically significant, such that further
For a reaction operating at 500 K with a TOF of 1 s−1 ,
research can focus on altering the nature of the catalyst
the maximum value of Hi that can be tolerated for a
and the reaction conditions to enhance the rates of these
species with high surface coverage (θA approaching unity)
kinetically controlling steps. This situation is illustrated
corresponds to 125 kJ mol−1 , which is still a rather high
in Fig. 3 for the above case of water-gas shift involving
value. In practice, the highest value of Hi that could be
carboxyl species, according to which the rate is controlled
tolerated would be lower than this value of 125 kJ mol−1 ,
by steps 3 and 4, whereas steps 1, 2, 5, 6, and 7 are
because the surface coverage by the reactive intermediate
quasi-equilibrated.
would typically be lower than unity and the above analysis
The net rate of step 3 in Fig. 3, is twofold faster than
assumes that the activation energy for the reverse of
the net rates of all other steps, because the stoichiometric
step i (i.e., the exothermic direction for this step) is equal
number of step 3 is equal to 2 whereas all other stoichio-
to zero. This situation where the overall enthalpy change
metric numbers are equal to 1. Importantly, the net rate of
is shared fairly equally between the various steps of the
reaction scheme is a necessary condition for high catalytic each step divided by its stoichiometric number is equal to
activity, but it is not a sufficient condition, because we the net rate of the overall catalytic reaction. This equality
have not yet considered the transition states for the various is due to the principle of kinetic steady state, as stated by
elementary steps. Bodenstein (see Ref. [38]), according to which determin-
The aforementioned reaction scheme for water-gas shift ing the rate of one single reaction (typically the overall
involving the formation of carboxyl species contains seven reaction) allows one to calculate the net rates of all the
steps, thereby requiring the determination (or estimation) other individual reactions. The Bodenstein principle is an
of 13 rate constants to describe the reaction kinetics important foundation of our thinking about how catalytic
completely; that is, a forward and reverse rate constant for cycles turn over. This principle also shows that the notion
each step (kfor,i and krev,i ) constrained by the relationship of a ‘‘slow’’ step in a catalytic cycle at the steady state is a
that these rate constants must give the proper equilibrium misnomer, because all steps proceed at the same net rate.
constant for the overall reaction, Keq , as given below:
1.1.6
  kfor,i σi Conditions for Catalyst Optimality
= Keq (3)
krev,i
i
It can be shown that the net rate of the overall catalytic
However, it is a rare case that all of these rate reaction is controlled by kinetic parameters which depend
constants are kinetically significant. Thus, while we only on the properties of the transition states for the
generally have the desire to know the values for as kinetically significant steps relative to the reactants (and
many rate constants as possible, we typically need to possibly the products) of the overall reaction [56]. The
know only the values of a limited number of these rate overall rate is also controlled by an additional kinetic
constants to describe the performance of the catalyst parameter for each surface species that is abundant on

1. CO + ∗ CO∗
2. H O +∗
2 H O∗
2

3. H2O ∗+ ∗ OH ∗ + H ∗
4. CO ∗+OH ∗ COOH ∗+ ∗
5. COOH ∗+OH ∗ CO2∗+H2O ∗
6. CO2∗ CO2 + ∗
7. 2H ∗ H + 2∗
2

Fig. 3 Rates of forward and reverse steps in the water-gas shift reaction on Cu.
ch1.1

1.1.6 Conditions for Catalyst Optimality 7

the catalyst surface. Specifically, the net rate of the overall of adsorption of one of the reactants [57], the heat of
reaction is determined by the kinetic parameters as well formation of a bulk compound that can be correlated
as by the fraction of the surface sites, θ ∗ , that is available with a heat of adsorption [58], the position of the catalytic
for the formation of the transition states; the value of θ ∗ element along a horizontal series in the Periodic Table,
is determined by the extent of site blocking by abundant an electronic property of the catalyst such as Pauling’s
surface species. d-band character of the metal [59], or the d-band center of
To illustrate how to determine the optimal activity the metal [60]. The optimal catalyst can thus characterized
of a catalyst, we consider an example in which the by the following relationship:
reaction scheme contains a single rate-controlling step  o‡
 o‡
and a single abundant surface species. According to H1 dH1
−C1 exp −
results obtained using DeDonder relations (discussed drnet RT RT dx
in Section 5.2.1.10) [56], we may write this reaction =0=   o 2
dx H2
scheme in terms of a quasi-equilibrated step involving 1 + C2 exp −
the transition state for the rate-controlling step, TSi , and RT
   
a second equilibrated step involving the formation of the H1
o‡
H2o dH2o
most abundant surface species, A∗ , as given below and: 2C1 exp − C2 exp −
RT RT RT dx
+    (11)
1. Reactants + 2∗ −
←−−
−−
→− T Si (4) H2o 3
1 + C2 exp −
2. A+∗ − ←−−
−−
→−A

(5) RT

The overall rate of the reaction, rnet , as will be discussed This relationship may be simplified to give:
in Section 5.2.1.12, is now given by:  
H2o dH2o
  o‡ 2C2 exp −
o‡ o‡ dH1 RT dx dH2o
ν‡ S1 H1 1/σ =    = 2θA
rnet = exp − F (ai )θ∗2 (1 − ztot 1 ) dx H2o dx
σ1 R RT 1 + C2 exp −
RT
(6) (12)
1 Thus, for the optimal catalyst, the surface coverage by
θ∗ =   (7)
S2o H2o the most abundant surface species is equal to:
1 + exp − aA
R RT o‡
dH1

where F (ai ) is a function of the activities (ai ) of the reac- dx ω
θA = o = 1 (13)
tants and/or products of the overall reaction. Neglecting dH2 2ω2
entropy effects, as we change the nature of the catalyst 2
dx
for constant reaction conditions, the primary items in the
o‡
above equations that change are H1 and H2o . (Note, where the values of ωi are defined as:
we implicitly assume that the reaction mechanism does o‡
‡ dH1
not change.) Accordingly, the overall rate of the reaction ω1 = (14)
dx
for different catalysts is given by:
  dH2o
o‡ ω2 = (15)
H1 dx
C1 exp −
RT ‡
rnet =   2 (8) In the above derivation, we assume that ω1 and ω2
H2o have the same sign, such that variations in x change
1 + C2 exp −
RT the enthalpy of the transition state and the MASI in
  o‡
the same direction. We also assume that (d 2 H1 )/dx 2
ν‡ S1
o‡  
1/σ o‡
C1 = exp F (ai ) 1 − ztot 1 (9) and (d 2 H2 )/dx 2 are small or zero. This assumption
σ1 R
is valid if we are searching for improved catalysts over
 
S2o a small range of x, which typically occurs when testing
C2 = exp aA (10) catalysts. In fact, when we vary x over a large range, then
R
the mechanism of the catalytic reaction would probably
We next consider that the surface properties of the change.
catalyst are described in terms of some fundamental
catalyst property, x. This property x could be a heat References see page 14
ch1.1

8 1.1 Principles of Heterogeneous Catalysis

Case 1: w‡1 >> w2 Ratenet is maximum at w‡1 = 2qMASIw2


Ratenet
Intermediate energies

Intermediate energies
is maximum as
∆H‡tran → 0 Case 2: w‡1 ≈ w2 ∆HMASI
∆H‡tran

Rate

Rate
∆H‡tran
∆HMASI

x (Fundamental catalyst parameter) x (Fundamental catalyst parameter)

Ratenet is maximum at w‡1 = 2qMASIw2


Intermediate energies

Case 3: w‡1 << w2


Rate

∆H‡tran
∆HMASI

x (Fundamental catalyst parameter)

Fig. 4 Reaction rates and energies of transition state and most abundant surface intermediate (MASI) as functions of fundamental
catalyst parameter ‘‘x’’.

First, we will consider Case 1 shown in Fig. 4, where increases for these three cases, the number of vacant sites

ω1  ω2 . In this case, a maximum rate does not exist on the catalyst increases.
and the parameter x should be adjusted to its lowest Cases 2 and 3 clearly illustrate the principle of Sabatier,
possible value (i.e., the strongest bonding to the surface) in which a maximum rate occurs at some moderate level
which would decrease the enthalpy of the transition of interaction of the catalyst surface with the intermediates
state as much as possible. In this case the optimal and adsorbed species. While Case 1 appears to contradict

catalyst operates at high surface coverage. We now Sabatier’s principle, the situation where ω1  ω2 is highly

consider Case 2, where ω1 ≈ ω2 ; that is, x changes unlikely. In particular, this situation corresponds to the
the enthalpies of the transition state and the MASI case where the catalyst interacts more strongly with the
by similar amounts. This situation is probably more transition state than with any of the reactive intermediates.
physically realistic, and if the MASI is in fact a reactive However, if the activation energies for the elementary
intermediate (i.e., if it is a MARI), then the family of steps of the mechanism are positive, then the reactants
catalysts described by the variation of x follows the and/or products of the elementary step involving the rate-
Brønsted–Evans–Polanyi–Semenov relationship, which controlling transition state are more strongly adsorbed
relates the thermodynamics and kinetics of the system. on the surface than is the transition state, leading to the
Here, there is a clear maximum in the rate versus x, situation described by Case 2 or 3.
as shown in Fig. 4. The plot of rates versus x appears We may generalize the above expression for catalyst
as a volcano-type curve which decreases symmetrically optimality to the case where the surface contains several

on both sides. In the case where ω1 is equal to ω2 , the abundant surfaces species, A∗ , B∗ , C∗ , and D∗ , leading to
surface coverage by the MASI on the optimal catalyst is the following expression:

equal to 0.5. For Case 3, ω1  ω2 , corresponding to the

situation where x changes the enthalpy of the MASI more ω1 = 2(ωA θA + ωB θB + ωC θC + ωD θD ) (16)
significantly than the transition state. The optimal catalyst
in this case has a low surface coverage by the MASI. A In this case, the nature of the optimal catalyst is
maximum rate occurs in this case, as shown in Fig. 4, controlled by the change in the binding energy of the

provided that ω1 > 0; however, the plot of rate versus x

transition state with respect to x (ω1 ) compared to the
is not symmetric with respect to x. We note that as x changes in the bindings energies of species A, B, C, and
ch1.1

1.1.6 Conditions for Catalyst Optimality 9

D (ωA , ωB , ωC , ωD ) weighed by their respective surface We now define the standard enthalpy for the formation
coverages at the steady state. of the transition state from gaseous species A as:
A bridge between the thermodynamics and kinetics o‡
of a reaction is provided by the Brønsted–Evans–Pol- H1 = Eact + B.E.A = E0 + α(Hgas )
anyi–Semenov relationship, which states that there is a + α(B.E.B + B.E.C ) + (1 − α)B.E.A (20)
linear relationship between the activation energy Eact of
an elementary step and the heat of reaction if entropy and we differentiate the enthalpy of the transition state
effects are neglected [38]: with respect to x, leading to the following result:
 
Eact = E0 + αH (17) dH1
o‡
dB.E.B dB.E.C dB.E.A
=α + + (1 − α)
where H is the enthalpy of reaction, α is the transfer dx dx dx dx
(21)
coefficient that varies between zero and one, and E0
is a constant. In other words, if we neglect entropic This relationship shows how the change of the
effects, the activation energy of an elementary step in ‡
transition state enthalpy with respect to x(ω1 ) is related to
the exothermic direction is lower when the heat of the changes in the binding energies B.E. of the adsorbed
reaction becomes more favorable (i.e., H becomes species (ω2 ). If α is equal to zero, then the change of the
more negative). DFT calculations have recently shown transition state enthalpy depends only on the change of
that Brønsted–Evans–Polanyi–Semenov relationships ‡
the binding energy of A, and ω1 = ω2 . In this case, the
are generally upheld in chemical reactions on catalyst
transition state is an early transition state that chemically
surfaces [61–64].
looks similar to A [65]. If α is equal to 1, then the transition
We now consider the following catalytic reaction:
state is a late transition state that resembles the products
A∗ −−−→ B∗ + C∗ (18) of the reaction, and the change of the transition state
enthalpy is related to changes of the binding energies of B
If we use the gas phase A species as the zero energy and C. If (dB.E.A )/dx = (dB.E.B )/dx = (dB.E.C )/dx,
level (as shown in Fig. 5) we can define the activation then the late transition state gives rise to the situation
energy as: ‡
where ω1 = 2ω2 .
The above examples show how it is possible to maximize
Eact = E0 + α(HBgas + B.E.B + HCgas + B.E.C
the activity of the catalyst. However, it is often more
− HAgas − B.E.A ) = E0 + α(B.E.B important to optimize the selectivity of the catalyst.
Similar types of analyses can be carried out for these cases.
+ B.E.C − B.E.A ) + αHgas (19)
In general, these situations are classified as being series
where the B.E. terms are the binding energies of the selectivity challenges, such as A → B → C, where B is
various species on the surface. the desired product, and/or parallel selectivity challenges,
such as A → B coupled with A → C. In these cases, if we
want to optimize the selectivity of the catalyst, we search
dB.E.A dB.E.B dB.E.C
Agas → Bgas + Cgas Assume: = = Bgas + Cgas for some catalyst property, x, that decreases the enthalpy
dx dx dx
of the transition state for the desired reaction more than
it decreases the enthalpies of the transition states for the
Agas ∆H‡ undesired reactions.
0
Eact
An undesired reaction that leads to progressive
a=0 blocking of surface sites leads to deactivation of the
B.E.A
catalyst with respect to time-on-stream. More generally,
A* a=1 various mechanisms exist by which a catalyst can
undergo deactivation, such as: (i) poisoning; (ii) thermal
A* B* + C* degradation (sintering); (iii) leaching of the active site;
and (iv) attrition [66]. The first three of these mechanisms
are chemical in nature, whereas the last mechanism
B* + C* is physical (e.g., the catalyst pellet breaks apart). Some
catalysts do not show any measurable deactivation over
Fig. 5 Schematic potential energy diagram of energy versus
periods of years, such as in ammonia synthesis. However,
reaction coordinate, showing the relationship between the energy other catalysts lose an important fraction of their activity
of transition state, H= , and changes in the energies of adsorbed
species. References see page 14
ch1.1

10 1.1 Principles of Heterogeneous Catalysis

after less than a minute of contact with feed, as in catalytic involves close collaboration between experts in such areas
cracking. In the latter case, if deactivation is caused by as catalyst synthesis, catalyst characterization, surface
coking, the catalyst must be regenerated by continuous spectroscopy, chemical kinetics, chemical reaction engi-
regeneration in an oxidizing atmosphere. neering and, most recently, in theoretical calculations of
catalyst structure and performance using density func-
1.1.7 tion theory. These broad studies can be grouped into
Catalyst Design three levels, as shown in Fig. 6.
All studies of heterogeneous catalysis begin at the
Given that the performance of a catalyst is controlled Materials Level. High-surface area catalytic materials
by a limited number of kinetic parameters, it is unclear must be synthesized with specific structures and textures,
why it is so difficult to design a catalyst from molecular- the latter referring to such features as the sizes of
level concepts. As noted above, during the early stages the various phase domains and the details of the pore
of research into a catalytic process, first we do not know structure. Clearly, the synthesis of catalytic materials
which steps in the reaction mechanism are kinetically must be guided by detailed characterization studies to
significant, and which species are most abundant on the determine the structures, compositions, and textures
catalyst surface under reaction conditions. Second, we of the materials that have been prepared. These
do not often know the structure of the active site and characterization studies should be conducted after the
its dependence on the nature of the reaction conditions. catalyst has been subjected to various treatment steps
Third, we do not usually know how the activity and (such as those treatments employed during activation of
selectivity for the catalytic reaction depend on the structure the catalytic material), and it is most desirable to carry
of the active sites. Fourth, we do not typically know during out characterization studies of the catalyst under the
these early stages the rates of various modes of catalyst actual reaction conditions of the catalytic process. Indeed,
deactivation (e.g., sintering, phase changes, deposition the properties of a heterogeneous catalyst are inherently
of carbonaceous deposits on the surface, etc.), and we dynamic in nature, and these properties often change
do not know whether the catalyst can be regenerated dramatically with changes in the reaction conditions
following deactivation. Finally, we must ensure that the (e.g., phase changes, surface reconstructions, changes
texture of the catalyst and the geometry of the reactor are in surface versus bulk composition, etc.) [67].
designed in such a way that mass transport of reactants The central level of research and development of het-
and products to and from the active sites is sufficiently erogeneous catalysts involves the quantification of catalyst
rapid that high rates of reaction per unit volume of reactor performance (this is known as the Catalyst Performance
can be achieved. Level). These studies can be carried out in a prelimi-
Because of these difficulties, the field of heteroge- nary fashion over a wide range of catalytic materials (e.g.,
neous catalysis is highly interdisciplinary in nature, and high-throughput studies) to identify promising catalysts

Materials level
Characterization studies: Materials synthesis:
catalyst structure, composition, catalytic materials
& texture, (ideally under with specific structures
reaction conditions) & textures

Catalyst performance level


Exploratory studies: Reaction kinetics studies:
promising leads for new activity, selectivity &
catalytic materials & new stability for various
catalytic reactions reaction conditions

Elucidation level
Surface studies: Experimental studies: Theoretical studies:
surface composition and nature interactions of probe stability & reactivity
of surface sites, (ideally under molecules with of species on
reaction conditions) well-defined sites well-defined sites

Fig. 6 Levels of study in heterogeneous catalysis research.


ch1.1

1.1.8 Catalyst Development 11

and reaction conditions for further studies. The perfor- a probe because it has an advantageous feature for spec-
mance of the catalyst is then documented in greater detail troscopic identification (e.g., CO for infrared studies, a
by determining catalytic activity, selectivity, and stability 13 C-containing molecule for NMR studies). These studies

with respect to time-on-stream for various reaction con- of the interaction of probe molecules with surfaces are
ditions. These measurements must be made at various designed to determine the surface concentrations of dif-
conversions when multiple reaction pathways exist, be- ferent types of surface site, to determine the nature of the
cause catalytic selectivities in these cases are different, adsorbed species formed on the surface sites, and to de-
depending on whether the desired products are formed termine the reactivities of the surface sites by monitoring
in primary versus secondary reactions, or in series versus the adsorbed species on the surface versus time, versus
parallel pathways. We note here that various definitions temperature or, most commonly, during a temperature
of catalytic activity are used, depending on the nature of ramp (e.g., temperature-programmed desorption).
the study. For practical studies, catalytic activities can be The third pillar of studies at the Elucidation Level
reported as rates per gram of catalyst or per unit surface involves the use of DFT calculations to assess the
area. However, for more detailed studies or for research structures, stabilities, and reactivity of species adsorbed
purposes, it is often desirable to report catalytic activities onto the surface sites (with the sites being composed of
as rates per surface site (i.e., TOFs), with the number of clusters of atoms or as periodic slabs of atoms) [77–81].
surface sites measured most often by selective adsorption These studies are used to help interpret the results
measurements (e.g., adsorption of H2 or CO to titrate obtained from spectroscopic studies of catalyst surfaces
metal sites, adsorption of ammonia or pyridine to titrate (e.g., to predict the vibrational spectra of species adsorbed
acid sites). In some cases it is possible to report catalytic in different orientations on different sites), to calculate
activity as rate per active site (also called TOF), when it is heats of adsorption for various intermediates in a reaction
possible to distinguish active sites from the larger num- mechanism (e.g., to predict which species are expected to
ber of surface sites using special probe molecules (e.g., be abundant on the catalyst under reaction conditions),
dissociative adsorption of N2 to titrate sites for ammo- to estimate the energy changes for possible steps in a
nia synthesis [68]; selective poisoning by adsorbates that reaction mechanism (thereby eliminating from further
compete with the reactants of the catalytic reaction [69]); consideration steps with very positive energy changes),
or by transient isotopic tracing [70]. and to determine activation energy barriers for steps
For the purposes of catalyst development, it is prob- that are suspected as being kinetically significant in the
ably sufficient to work at the Materials Level and the reaction scheme. Indeed, a key feature of these theoretical
Catalyst Performance Level. However, research into het- studies is the ability to predict how the surface properties
erogeneous catalysis is dominated by studies conducted are expected to change as the nature of the surface is
at a third level – the Elucidation Level – where the aim altered (e.g., by changing the surface structure, or by
is to identify the fundamental building blocks of knowl- adding possible promoters). This in turn will provide
edge which can be assembled to build a molecular-level feedback to the Materials Level with regards to new
understanding of catalyst performance in order to guide materials that should be synthesized and which are likely
further investigations to improve catalyst performance. to lead to an improved catalyst performance. In addition,
At the Elucidation Level the studies are designed to deter- these theoretically based studies provide information
mine the surface composition and nature of the surface about highly reactive intermediates which might be
sites on the catalyst [71–74]. Clearly, these investigations difficult to obtain by direct experimental measurements.
must be conducted with the catalyst under controlled con- Most importantly, studies conducted at the Elucidation
ditions (e.g., under ultra-high vacuum, after treatment Level provide a scientific basis about the working catalysis
with H2 , after calcination, etc.) and, where possible, such that may, in future, be used to design different reaction
measurements should be made with the catalyst under pathways.
reaction conditions. Moreover, the studies may be carried
out on real catalytic materials and on more well-defined 1.1.8
surfaces (e.g., single crystals, or model samples formed Catalyst Development
by depositing known amounts of materials onto well-
defined supports) [73, 75, 76]. Most measurements at the Catalyst development typically involves testing a large
Elucidation Level involve studies of the interactions of number of catalysts with a feedback loop, as it is currently
specific probe molecules with the catalyst surface. These difficult to design catalysts a priori. In this respect, catalyst
probe molecules may be the reactants, intermediates, or development studies involve examining a large number
products of the catalytic reaction, or they may be more of catalysts, for which recent advances in high-throughput
simple species chosen to monitor a specific functionality
of the surface. Alternatively, a molecule may be used as References see page 14
ch1.1

12 1.1 Principles of Heterogeneous Catalysis

testing have attracted considerable attention [82–87]. 1.1.9


Catalyst development through the testing of a wide range Bridging Gaps in Heterogeneous Catalysis
of materials was first practiced in 1909 by Mittasch at
BASF who, according to Timm [88], issued the following The above description of research and development into
directive to his team who at the time were developing the heterogeneous catalysis as being interdisciplinary in na-
synthesis of ammonia: ture, involving studies at the levels of materials, catalyst
performance and elucidation, can also be cast in the form
• The search for a suitable catalyst necessitates carrying of building bridges between various types of studies and
out experiments with a number of elements, together different types of material. As depicted in Fig. 7, we often
with numerous additives. talk about bringing together the field of surface science
• The catalytic substances must be tested at high (which traditionally is focused on studies of single crystal
pressures and temperatures, just as in the case of surfaces at low pressures) with the field of heterogeneous
Haber’s experiments. catalysis (which traditionally is focused on studies of
• A very large number of tests will be required. high-surface area catalytic materials surfaces under high-
pressure reaction conditions). More recently, we have
talked about ‘‘bridging the materials gap’’, as we have
Ten years later, the number of tests conducted had
attempted to use experimental results from studies of well-
exceeded 10 000, and more than 4000 catalysts had been
defined model materials to interpret the performance of
studied. This extraordinary effort was also extraordinarily
more complex, high-surface area catalytic materials. Tra-
successful. What has changed since then, however, is
ditionally, these model materials have been single crystals,
the way in which the systematic search is assisted.
cut at various angles to expose surfaces containing differ-
Today, armed with an arsenal of principles, concepts,
ent types of sites, such as surfaces with different symme-
instrumentation and computers, it is possible to identify
tries and atoms present at terraces, steps, and kinks [89].
and to improve new catalytic materials in a much
More recently, however, these model materials have be-
shorter time and with a smaller number of trial samples,
come highly sophisticated, such as the deposition of
especially with the possibility of advanced characterization
nanoparticles with specific sizes and geometries on well-
methods (especially in-situ techniques) and insights from
defined support surfaces (e.g., metal nanoparticles sup-
theoretical calculations (e.g., DFT calculations). The
ported on thin films of oxides deposited on single crystal
practical merit of this ‘‘assisted catalyst design’’ is clear,
metal surfaces, or non-metallic nanoparticles supported
while its scientific dividend is the possibility of learning
directly on single crystal metal surfaces) [73, 75, 76]. We
as the design proceeds, with the building of a data bank
also talk about ‘‘bridging the pressure gap’’, as we at-
of rate constants and the formulation of more precise
tempt to use experimental results from studies conducted
models of active sites. With new theoretical insights
at low pressures (less than 10−6 Torr) to interpret the
or principles, quantitative bases of catalyst preparation
performance of catalysts under high-pressure reaction
and reproducibility of catalyst behavior, the future of
heterogeneous catalysis still looks very bright.
The path to the design of an optimal catalytic process
would be clear if the activity, selectivity and stability High m2
of the catalyst were to move in the same direction
upon an increase in a single process variable, such
as temperature. However, this simple behavior is not Heterogeneous
typically observed, and choices must be made in every catalysis
instance. For example, while the activity of a catalyst
Materials gap

may increase with temperature, its stability usually


decreases with temperature. In addition, the relationship Low P Pressure gap High P
between catalytic activity and selectivity is typically very
complex, and is not understood in detail until the surface Surface
chemistry of the catalytic process has been elucidated. science
Accordingly, selectivity, stability and activity must be
considered together, and trade-offs may have to be
negotiated, perhaps by using multi-functional reactors Low m2
with catalytic distillation or catalytic membranes. Success
in heterogeneous catalysis begins with chemistry, but Fig. 7 Bridging the gap between surface science and heteroge-
ends with catalytic reaction engineering. neous catalysis.
ch1.1

1.1.10 A Philosophical Note 13

conditions. The origin for this pressure gap comes from techniques such as scanning tunneling microscopy that
the fact that, whereas some spectroscopic techniques can allow atomic-scale imaging of materials at elevated tem-
be employed to study the surface and bulk properties of peratures and pressures) [73, 97]. However, as an as
catalysts under high-pressure reaction conditions (e.g., increasing number of research groups become involved
FTIR, Raman, XRD, EXAFS, Mössbauer spectroscopy), in nanotechnology, it is possible that an ‘‘applications
other spectroscopic and characterization techniques (e.g., gap’’ will be created in heterogeneous catalysis, where
XPS, TEM) are most easily conducted with the sample new materials are formed without clear applications for
at low pressures (e.g., <10−6 Torr) [90]. These latter tech- catalytic processes. Clearly, this gap can be bridged by re-
niques are typically associated with use of electrons to alizing that research and development into heterogeneous
probe the sample, with the electrons interacting strongly catalysis involves the combination of studies at the levels
with molecules in the gas phase. This pressure gap can be of materials, catalyst performance, and elucidation. As ad-
bridged directly by designing advanced instrumentation, vances in nanotechnology allow us to create new materials
such that the distance traversed by the electrons in the (the Materials Level) and to characterize these materials
gas phase is minimized [91, 92]. In addition, the pressure in greater detail (Materials and Elucidation Levels), we
gap can be bridged indirectly by using molecular-based are positioned to take full advantage of these advances by
models (e.g., kinetic Monte Carlo calculations, micro- conducting studies at the Catalyst Performance Level.
kinetic models), first to describe the experimental results
obtained at low pressures, and then to extrapolate this 1.1.10
information to high-pressure reaction conditions. A Philosophical Note
The past few years have witnessed an explosion in
the area of nanotechnology, in which researchers have Today, we live in an era in which it is possible to employ
learned – and are continuing to learn – how to engineer a vast range of advanced experimental techniques and
materials at the nanometer length scale. The field of het- theoretical methods to elucidate in detail the surface
erogeneous catalysis has been involved in the synthesis chemistry of catalytic processes. This situation, with
of nanomaterials for many years (e.g., the synthesis of respect to the hierarchy of theoretical methods that
zeolites). Indeed, the scheme depicted in Fig. 6 shows can be employed to describe the reaction kinetics for
that essentially all studies of heterogeneous catalysis a catalytic process, is illustrated schematically in Fig. 8. At
begin at the Materials Level. Recent advances in nan- the lowest level, we use empirical rate expressions to fit
otechnology offer new routes for catalyst synthesis (e.g., reaction kinetics data over a range of process conditions;
atomic layer deposition, self-assembly methods) [93–96]
and, importantly, also for catalyst characterization (e.g., References see page 14

Elucidation well

Empirical rate expressions

Simplified rate expressions based


on assumed surface chemistry
Sufficient insight gained

new catalytic processes


to guide the search for

Micro-kinetic rate expressions based


Elucidation

on results from DFT calculations

Kinetic monte-carlo analyses based


on results from DFT calculationsns

Quantum molecular dynamics calculations


for surface reactions

Discovery Future methods …..

Desire to know more


about surface chemistry

Fig. 8 The catalyst ‘‘elucidation well’’ and catalyst discovery.


ch1.1

14 1.1 Principles of Heterogeneous Catalysis

however, these models typically have questionable success STY site time yield
in predicting catalyst performance outside the range of TEM transmission electron microscopy
experimental conditions used to fit the model. TOF turnover frequency
We then move to rate expressions based on assumed TOR turnover rate
surface chemistry. These models should have a better XPS X-ray photoelectron spectroscopy
predictive value, although it is often difficult to determine XRD X-ray diffraction
which types of assumption should be made. Accordingly,
we turn to results from DFT calculations to build micro- References
kinetic models that describe catalyst performance without
the need to make prior assumptions about which steps are 1. J. M. Thomas, W. J. Thomas, Principles and Practice of Hetero-
kinetically significant and which species are abundant on geneous Catalysis, VCH, Weinheim, 1997, 669 pp.
the catalyst surface. These micro-kinetic models, however, 2. J. N. Armor, Appl. Catal., A 2001, 222, 407.
are typically based on the mean-field approximation, and 3. I. Chorkendorff, J. W. Niemantsverdriet, Concepts of Mod-
ern Catalysis and Kinetics, Wiley-VCH, Weinheim, 2003,
they thus make simplified assumptions about (or neglect)
452 pp.
the effects of surface coverage and lateral interactions 4. R. J. Farrauto, C. H. Bartholomew, Fundamentals of Indus-
between adsorbed species. Accordingly, these restrictive trial Catalytic Processes, Chapman & Hall, London, 1997,
assumptions are relaxed when using kinetic Monte Carlo 754 pp.
methods to describe reaction kinetics based on binding 5. R. A. van Santen, P. W. N. M. v. Leeuwen, J. A. Mooulijn,
energies and lateral interaction terms determined from B. A. Averill, Catalysis: An Integrated Approach, Elsevier
DFT calculations. Indeed, today’s research groups are Science B.V., Amsterdam, 1999, 574 pp.
6. National Research Council Panel on New Directions in
beginning to combine quantum mechanics and molecular Catalytic Sciences and Technology, Catalysis Looks to the Future,
dynamics calculations to describe a variety of surface National Academy Press, Washington D.C., 1992, p. 1.
processes. Who knows what new computational methods 7. W. Vielstich, A. Lamm, H. Gasteiger, Handbook of Fuel Cells:
are on the horizon? Fundamentals, Technology, Applications, Wiley, Chichester,
The sequential use of the aforementioned methods to 2003, 2690 pp.
describe reaction kinetics in greater detail is depicted 8. S. Park, J. M. Vohs, R. J. Gorte, Nature 2000, 404, 265.
9. S. Ha, R. Larsen, R. I. Masel, J. Power Sources 2005, 144, 28.
in Fig. 8, as the digging of an ‘‘elucidation well’’. As
10. A. Corma, H. Garcia, Chem. Rev. 2003, 103, 4307.
we dig deeper by using more sophisticated methods, we 11. G. Centi, Catal. Today 2003, 77, 287.
learn more about the details of the surface chemistry. 12. R. A. Sheldon, Green Chem. 2005, 7, 267.
Indeed, we are driven to dig deeper by our desire to learn 13. A. Borgna, L. Balzano, J. E. Herrera, W. E. Alvarez, D. E.
as much as possible about the fundamental principles Resasco, J. Catal. 2001, 204, 131.
that control catalyst performance. However, this desire 14. H. van Bekkum, P. Gallezot, Top. Catal. 2004, 27, 1.
15. G. W. Huber, J. W. Shabaker, J. A. Dumesic, Science 2003,
to know as much as possible must be balanced by our
300, 2075.
need to discover new catalysts and new catalytic processes. 16. G. W. Huber, J. N. Chheda, C. J. Barrett, J. A. Dumesic, Sci-
Clearly, as we learn more about the fundamentals of the ence 2005, 308, 1446.
catalytic process (i.e., as we dig deeper), we should have 17. I. K. Mbaraka, B. H. Shanks, J. Catal. 2005, 229, 365.
better insight to guide our search for better catalysts. 18. S. Varadarajan, D. J. Miller, Biotechnol. Progr. 1999, 15, 845.
Luckily, we need not dig to the deepest levels to begin the 19. E. K. Rideal, H. S. Taylor, Catalysis in Theory and Practice,
Macmillan, London, 1919, Chapter 2.
search for better catalysts. As noted at the beginning
20. M. Boudart, in Perspectives in Catalysis, J. M. Thomas,
of this chapter, our industrialized society is facing K. I. Zamaraev (Eds.), Blackwell, Oxford, 1992, p. 183.
increasingly complex environmental and energy issues 21. G. A. Somorjai, Introduction to Surface Chemistry and Catalysis,
for sustained growth. Thus, while our scientific curiosity John Wiley, New York, 1994, 667 pp.
for fundamental knowledge drives us to dig deeper toward 22. H. S. Taylor, Proc. Roy. Soc. (London) 1925, A108, 105.
detailed elucidation of catalytic phenomena, we must also 23. J. B. Anderson, Kagaku Kogaku (Chem. Eng. Jpn.) 1962, 147,
continue to look horizontally as we use our newfound 191.
24. J. J. Carberry, AICHE 1961, 7, 350.
insight to develop new catalysts and catalytic processes
25. G. F. Froment, K. B. Bischoff, Chemical Reactor Analysis and
for the benefit of society. Design, Wiley, New York, 1990, 664 pp.
26. C. N. Satterfield, Mass Transfer in Heterogeneous Catalysis, MIT
List of Abbreviations Press, Cambridge, MA, 1970, 267 pp.
27. D. E. Mears, Ind. Eng. Chem. Proc. Des. Dev. 1971, 10, 541.
28. D. E. Mears, J. Catal. 1971, 20, 127.
DFT density functional theory 29. J. M. Smith, J. Chem. Eng. Japan 1973, 6, 191.
EXAFS extended X-ray absorption fine structure 30. P. B. Weisz, Z. Phys. Chem. 1954, 11, 1.
FTIR Fourier transform infrared 31. P. B. Weisz, C. D. Prater, Adv. Catal. 1957, 6, 143.
ch1.1

References 15

32. R. Aris, in Catalyst Design: Progress and Perspectives, 68. H. Topsøe, J. A. Dumesic, N. Topsøe, H. Bohlbro, in
L. L. Hegedus (Ed.), Wiley Interscience, New York, 1987, Proceedings of the Seventh International Congress in Catalysis,
Chapter 7. T. Seiyama, K. Tanabe (Eds.), Elsevier, Amsterdam, 1981,
33. P. A. Jacobs, J. A. Martens, J. Weitkamp, H. K. Beyer, Faraday p. 247.
Discuss. Chem. Soc. 1981, 72, 353. 69. H. Knözinger, Adv. Catal. 1976, 25, 183.
34. W. O. Haag, R. M. Lego, P. B. Weisz, Faraday Discuss. Chem. 70. J. G. Goodwin Jr., S. Kim, W. D. Rhodes, Catal. 2004, 17, 320.
Soc. 1981, 72, 317. 71. D. W. Goodman, J. Catal. 2003, 216, 213.
35. E. G. Derouane, P. Dejaifve, Z. Gabelica, Faraday Discuss. 72. G. A. Somorjai, K. R. McCrea, J. Zhu, Top. Catal. 2002, 18,
Chem. Soc. 1981, 72, 331. 157.
36. J. M. Thomas, G. R. Millward, S. Ramdas, L. A. Busil, 73. H.-J. Freund, M. Baumer, J. Libuda, T. Risse, G. Rupprechter,
M. Audier, Faraday Discuss. Chem. Soc. 1981, 72, 345. S. Shaikhutdinov, J. Catal. 2003, 216, 223.
37. P. Sabatier, La catalyse en chimie organique, Bérange, Paris, 74. G. Ertl, J. Mol. Catal. A: Chem. 2002, 182–183, 5.
1920, 388 pp. 75. J. V. Lauritsen, M. Nyberg, J. K. Norskov, B. S. Clausen,
38. M. Boudart, Kinetics of Chemical Processes, Blackwell, Oxford, H. Topsoe, E. Laegsgaard, F. Besenbacher, J. Catal. 2004,
Stoneham, MA, 1991, 246 pp. 224, 94.
39. J. V. Lauritsen, M. Nyberg, J. K. Norskøv, J. Catal. 2004, 224, 76. M. S. Chen, D. W. Goodman, Science 2004, 306, 252.
94. 77. J. Greeley, J. K. Norskøv, M. Mavrikakis, Annu. Rev. Phys.
40. B. Hinnemann, J. K. Norskøv, H. Topsøe, J. Phys. Chem. B Chem. 2002, 53, 319.
2005, 109, 2245. 78. M. Neurock, J. Catal. 2003, 216, 73.
41. H. H. Hwu, J. G. Chen, Chem. Rev. 2005, 105, 185. 79. B. Hammer, J. K. Norskøv, Adv. Catal. 2000, 45, 71.
42. R. B. Levy, M. Boudart, Science 1973, 181, 547. 80. S. Linic, H. Piao, K. Adib, M. A. Barteau, Angew. Chem. Int.
43. M. K. Neylon, S. Choi, H. Kwon, K. E. Curry, L. T. Thompson, Ed. 2004, 43, 2918.
Appl. Catal., A 1999, 183, 253. 81. M. T. M. Koper, R. A. van Santen, M. Neurock, in Catalysis
44. S. T. Oyama, Catal. Today 1992, 15, 179. and Electrocatalysis at Nanoparticle Surfaces, A. Wieckowski,
45. S. G. Podkolzin, J. Shen, J. J. D. Pablo, J. A. Dumesic, J. Phys. E. R. Savinova, C. G. Vayenas (Eds.), Marcel Dekker, Inc., New
Chem. B 2000, 104, 4169. York, 2003, p. 1.
46. R. M. Watwe, B. E. Spiewak, R. D. Cortright, J. A. Dumesic,
82. A. Hagemeyer, P. Strasser, J. Anthony, F. Volpe, High-
Catal. Lett. 1998, 51, 139.
Throughput Screening in Chemical Catalysis: Technologies,
47. C. G. M. Hermse, A. P. van Bave, A. P. J. Jansen, L. A. M. M.
Strategies and Applications, Wiley-VCH Verlag, Weinheim,
Barbosa, P. Sautet, R. A. van Santen, J. Phys. Chem. B 2004,
Germany, 2004, 339 pp.
108, 11035.
83. A. Hagemeyer, R. Borade, P. Desrosiers, S. Guan, D. M. Lowe,
48. M. Neurock, S. A. Wasileski, D. Mei, Chem. Eng. Sci. 2004, 59,
D. M. Poojary, H. Turner, H. Weinberg, X. Zhou, R. Armbrust,
4703.
G. Fengler, U. Notheis, Appl. Catal., A 2002, 227, 43.
49. S. Raimondeau, P. Aghalayam, A. B. Mhadeshwar, D. G.
84. R. J. Hendershot, C. M. Snively, J. Lauterbach, Chem. Eur. J.
Vlachos, Ind. Eng. Chem. Res. 2003, 42, 1174.
2005, 11, 806.
50. F. J. Garcia, E. E. Wolf, Chem. Eng. Sci. 2004, 59, 4723.
85. J. M. Serra, E. Guillon, A. Corma, J. Catal. 2005, 232, 342.
51. J. M. Thomas, C. R. A. Catlow, G. Sankar, Chem. Commun.
2002, 2921. 86. S. Senkan, Angew. Chem. Int. Ed. 2001, 40, 312.
52. A. A. Gokhale, Water–Gas Shift Reaction and Fischer–Tropsch 87. J. W. Saalfrank, W. F. Maier, Angew. Chem. Int. Ed. 2004, 43,
Synthesis on Transition Metal Surfaces, Phd thesis, University 2028.
of Wisconsin, 2005. 88. B. Timm, Proceedings of the 8 th International Congress on
53. P. S. Cremer, X. Su, Y. R. Shen, G. A. Somorjai, J. Am. Chem. Catalysis Berlin 1984, Verlag Chemie, Frankfurt, 1984, Vol. I,
Soc. 1996, 118, 2942. p. 7.
54. H. Topsøe, J. Catal. 2003, 216, 155. 89. P. L. J. Gunter, J. W. Niemantsverdriet, F. H. Ribeiro, G. A.
55. I. E. Wachs, Catal. Today 1996, 27, 437. Somorjai, Cat. Rev. - Sci. Eng. 1997, 39, 77.
56. J. A. Dumesic, J. Catal. 1999, 185, 496. 90. J. W. Niemantsverdriet, Spectroscopy in Catalysis, Wiley, Wein-
57. O. Beeck, Disc. Faraday Soc. 1950, 8, 118. heim, Germany, 1993, 288 pp.
58. W. J. M. Rootsaert, W. M. H. Sachtler, Z. Physik. Chem. 1960, 91. P. L. Hansen, J. B. Wagner, S. Helveg, J. R. Rostrup-Nielsen,
26, 16. B. S. Clausen, H. Topsøe, Science 2002, 295, 2053.
59. J. H. Sinfelt, Bimetallic Catalysts, Wiley, New York, 1983 p. 14. 92. D. Teschner, A. Pestryakov, E. Kleimenov, M. Haevecker,
60. M. Boudart, J. Am. Chem. Soc. 1950, 72, 1050. H. Bluhm, H. Sauer, A. Knop-Gericke, R. Schloegl, J. Catal.
61. T. Bligaard, J. K. Norskøv, S. Dahl, J. Matthiesen, C. H. 2005, 230, 186.
Christensen, J. Sehested, J. Catal. 2004, 224, 206. 93. A. Corma, F. Rey, J. Rius, M. J. Sabater, S. Valencia, Nature
62. J. K. Norskøv, T. Bligaard, A. Logadottir, S. Bahn, L. B. Hansen, 2004, 431, 287.
M. Bollinger, H. Bengaard, B. Hammer, Z. Sljivancanin, M. 94. Y. Yin, R. M. Rioux, C. K. Erdonmez, S. Hughes, G. A.
Mavrikakis, Y. Xu, S. Dahl, C. J. H. Jacobsen, J. Catal. 2002, Somorjai, A. P. Alivisatos, Science 2004, 304, 711.
209, 275. 95. M. J. Pellin, P. C. Stair, G. Xiong, J. W. Elam, J. Birrell,
63. V. Pallassana, M. Neurock, J. Catal. 2000, 191, 301. L. Curtiss, S. M. George, C. Y. Han, L. Iton, H. Kung, M.
64. Z. P. Liu, P. J. Hu, J. Chem. Phys. 2001, 114, 8244. Kung, H. H. Wang, Catal. Lett. 2005, 102, 127.
65. R. A. van Santen, J. W. Niemantsverdriet, Chemical Kinetics 96. A. Gervasini, P. Carniti, J. Keranen, L. Niinisto, A. Auroux,
and Catalysis, Plenum Press, New York, 1995, p. 233. Catal. Today 2004, 96, 187.
66. C. H. Bartholomew, Appl. Catal., A 2001, 212, 17. 97. F. Besenbacher, E. Laegsgaard, I. Stensgaard, Mater. Today
67. G. A. Somorjai, Annu. Rev. Phys. Chem. 1994, 45, 721. 2005, 8, 26.

You might also like