Você está na página 1de 36

CH05CH26-Marin ARI 20 May 2014 19:23

Experimental and Theoretical


Methods in Kinetic Studies of
Heterogeneously Catalyzed
Reactions
Marie-Francoise Reyniers and Guy B. Marin
Laboratory for Chemical Technology (LCT), Ghent University, B-9052 Ghent, Belgium;
email: guy.marin@ugent.be
Annu. Rev. Chem. Biomol. Eng. 2014. 5:56394
The Annual Review of Chemical and Biomolecular
Engineering is online at chembioeng.annualreviews.org
This articles doi:
10.1146/annurev-chembioeng-060713-040032
Copyright c 2014 by Annual Reviews.
All rights reserved
Keywords
microkinetics, zeolite acidity, catalyst design, catalytic cracking, ethanol
conversion
Abstract
This review aims to illustrate the potential of kinetic analysis in general
and microkinetic modeling in particular for rational catalyst design. Both
ab initio calculations and experiments providing intrinsic kinetic data allow
us to assess the effects of catalytic properties and reaction conditions on
the activity and selectivity of the targeted reactions. Three complementary
approaches for kinetic analysis of complex reaction networks are illustrated,
using select examples of acid zeolitecatalyzed reactions from the authors
recent work. Challenges for future research aimed at dening targets for
synthesis strategies that enable us to tune zeolite properties are identied.
563
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
Click here for quick links to
Annual Reviews content online,
including:
Other articles in this volume
Top cited articles
Top downloaded articles
Our comprehensive search
Further
ANNUAL
REVIEWS
CH05CH26-Marin ARI 20 May 2014 19:23
INTRODUCTION
The ultimate goal of chemical kinetics is to unravel the mechanism of a chemical reaction and
to predict the rate of formation of desired and undesired products, both of which are essential
to maximize the energy efciency and minimize the environmental impact of chemical processes.
Approximately 90% of all chemical processes worldwide (1) make use of catalysts, and there is a
pressing need for more efcient solid catalysts to allow sustainable production of fuels and bulk
chemicals. At present, catalyst development is dominated by experimental research and is focused
mainly on optimization of a given catalyst with respect to given criteria (2). The ability to predict
how the catalysts composition and structure, and hence its physicochemical properties, control
the rate of key steps in the catalytic mechanism is essential to come to a rational design of new
and more efcient catalysts.
Microkinetic models are uniquely suited to identify which (surface) species or surface reac-
tion paths control the observed product spectrum; thus, they provide a tool to dene the process
conditions to obtain the maximum yield of the desired product (3). Moreover, microkinetic mod-
els have the unique ability to provide a unied, quantitative understanding of heterogeneously
catalyzed reactions by incorporating knowledge of a series of catalysts obtained from various sur-
face science techniques (46), experimental kinetic studies (7, 8), quantum chemical studies, and
statistical thermodynamics (918). Hence, they have the potential not only to incorporate (10,
11, 1922) but also to develop structure-activity relationships and/or identify catalyst descriptors,
which makes them an ideal tool for in silico catalyst design (10, 11, 23). Therefore, kinetic studies
in heterogeneous catalysis now aim mainly to identify the detailed reaction mechanism, develop
an appropriate mathematical description of the inuence of the reaction conditions and catalyst
properties on the elementary reaction rates, and determine the unknown parameters in the rate
expressions. Kinetic analysis of the reaction network then allows us to identify which elementary
steps control the reaction rates (2428), how their rates are affected by the reaction conditions,
and the catalyst properties, all of which are crucial to provide guidelines for the design of bet-
ter catalysts (29). To obtain the necessary information and/or to validate rate coefcients and
mechanisms obtained from quantum chemical calculations, steady-state and non-steady-state or
transient kinetic experiments can be used (24).
We begin with a brief survey of the necessary requirements to extract model-free intrinsic re-
action rates from dedicated experiments. Comparison of experimentally observed rates with those
obtained based on quantum chemical calculations is often biased by overlooking these necessary
requirements. Next, we illustrate the potential of microkinetic modeling for catalyst design for
three complex acid zeolitecatalyzed reactions taken from our work. Zeolites are microporous,
crystalline, inorganic oxides that are widely used as solid Brnsted acid catalysts in a variety of
chemical processes for the production of fuels, energy carriers, and bulk and ne chemicals. The
catalytic properties of zeolites are related mainly to their structural and compositional character-
istics, such as pore opening, channel and void dimensions, and Si/Al ratio. Typically, zeolites are
excellent catalysts owing to their high surface area, high adsorption capacity, and well-dened
micropores that can induce shape selectivity (3037). Historically, the emphasis on the size and
shape of zeolite voids and their connecting apertures in explaining shape selectivity has led to
the development of zeolite descriptors based on reactions of molecules of different sizes, such
as the constraint index (38) and the spaciousness index (39). These methods provide an indi-
rect assessment of zeolite activity when it is controlled by molecular transport or size exclusion.
Although these methods allow us to screen the immense diversity of shape, size, and connectiv-
ity of voids and pores in hypothetical (10
6
in number) (4042) and known (10
2
in number)
(http://www.iza-structure.org/databases/) zeolites, they do not provide insight into how the
564 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
physicochemical properties of the zeolite, such as acid strength and solvation ability of conning
voids, inuence zeolite reactivity and selectivity (43). Given that only approximately ten families of
zeolites are commonly applied in industrial processes (44), there is a clear need for a more rational
design of zeolite materials with desired properties for a targeted catalytic reaction. Three different
but complementary approaches for kinetic analysis of zeolite-catalyzed reactions are presented to
illustrate their ability to provide physicochemical insight into the inuence of zeolite properties on
the observed product yields under working conditions. We conclude by identifying challenges for
future research that should help to further advance the rational design of heterogeneous catalysts
in general and of zeolites in particular.
KINETIC EXPERIMENTS
Experimental Requirements
Usually, experimental kinetic studies are performed in laboratory-scale stirred tank reactors
(STRs) or plug ow reactors (PFRs). Product yields of heterogeneously catalyzed processes per-
formed in these types of reactor not only depend on the rate of the chemical reactions but also
are inuenced by physical transport phenomena, such as heat and mass exchange between the
catalyst and the uid phase (so-called external heat and mass transfer) and heat and mass transport
within the catalyst pellet, so-called internal transport. To extract intrinsic kinetic information
(i.e., kinetic data that pertain to the chemical reactions only), the typical strategy in experimental
kinetic studies is to eliminate the effects of mass and heat transport as much as possible. Ideally,
the kinetic experiment must fulll two requirements: isothermicity and uniformity of chemical
composition. In dedicated laboratory setups, this can be accomplished by, for example, diluting
the reactive medium by using special mixing devices, either internally (impellers) or externally by
recirculation. Several experimental and theoretical (4548) diagnostic tools are available to iden-
tify the range of reaction conditions and catalyst particle diameters that allow us to ensure that
in a given reactor conguration the inuence of inter- and intraparticle transport on the reaction
rates is eliminated. In so-called intrinsic kinetic studies, these requirements are fullled. In the
case of irreducible transport phenomena, the transport regime in the reactor must be well dened,
and its mathematical description must be reliable (24). Failure to comply with these experimental
requirements inevitably produces kinetic data that are biased by ill-dened transport phenomena.
This leads to biased values of tted rate coefcients and to misleading conclusions when compar-
ing experimentally determined turnover frequencies and/or rate-controlling steps from different
literature sources or with those obtained from ab initiobased kinetic model simulation.
Obtaining Model-Free Kinetic Data
Importantly, in kinetic experiments, the reaction rate can seldom be measured directly. The
experimentally determined characteristic in most experimental kinetic studies is the net rate of
production of a component, R
i
, which can be dened as a linear combination of the rates of all
the elementary steps, r
j
, in which the component is consumed or formed. The coefcients in this
linear combination are the components stoichiometric coefcients,
ij
, in each of the steps:
R
i
=

i j
r
j
. 1.
Both continuously stirred tank reactors (CSTRs) and PFRs are used frequently to obtain steady-
state kinetic information. Steady-state kinetic data are usually related to the rate-controlling step(s)
www.annualreviews.org Heterogeneously Catalyzed Reactions 565
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
STEADY-STATE ISOTOPIC TRANSIENT KINETIC ANALYSIS
Steady-State Isotopic Transient Kinetic Analysis (SSITKA) is based on the determination of isotopic labels of a
component in the reactor efuent versus time following a stepwise change in the isotopic labeling of one of the
reactants in the reactor feed. SSITKA can be used to determine the amount and rate of formation of surface species;
in other words, it allows us to characterize the steady state of the catalyst surface under working conditions.
of a complex catalytic reaction. Toobtainkinetic informationonthe individual elementary reaction
steps, non-steady-state experiments are required.
Ideally, kinetic information should be obtained in a model-free manner, which means that
during the experimental procedure and the subsequent data processing no bias is introduced
by making assumptions about, for example, the reaction mechanism, the kinetic model, or the
reactor model. Some techniques, such as a combination of CSTR or PFR with isotopic step-
response experiments, allow the model-free determination of formation rates by Steady-State
Isotopic Transient Kinetic Analysis (SSITKA) (49) (see sidebar, Steady-State Isotopic Transient
Kinetic Analysis).
Non-steady-state kinetic data can also be based on the technique of Temporal Analysis of
Product (TAP) (24, 50, 51). The earlier three- and one-zone congurations of the TAP reactor
could not provide model-free kinetic information. In contrast, the Y-procedure (52), a new theo-
retical method for the thin-zone TAP reactor, enables values of reaction rates and, hence, kinetic
coefcients to be obtained from exit ow rates without any assumptions regarding the detailed
kinetic model (see sidebar, Temporal Analysis of Products).
CSTRandPFRexperiments are performedat temperatures andpressures relevant for industrial
conditions. For highly exothermic reactions, CSTRs are usually operated at low conversions to
ensure isothermicity. The TAP system can be used in a wide domain of operation conditions;
however, to ensure the Knudsen diffusion regime (i.e., a well-dened transport regime that can
be reliably described mathematically), the preferred window of pressures ranges from 10
2
to
10
1
Pa. The latter pressure is the upper boundary of the surface science domain; hence, TAP
experiments provide a bridge between surface science and more traditional applied kinetics.
KINETIC ANALYSIS OF COMPLEX ACID
ZEOLITECATALYZED REACTIONS
Three different but complementary approaches for the kinetic analysis of complex acid zeolite
catalyzed reaction networks are presented. All three approaches aim to obtain a quantitative
TEMPORAL ANALYSIS OF PRODUCTS
The main idea of Temporal Analysis of Products (TAP) is to inject a series of narrow gas pulses into one end of a
microreactor and continuously evacuate the other end. The intensity of each pulse is small relative to the amount
of catalyst used to ensure that the surface state of the catalyst is not signicantly perturbed by each pulse and that
the Knudsen diffusion regime is guaranteed. The change in the composition of the pulsed mixture is monitored
by a precise mass-spectroscopic technique. Upon applying a large series of such narrow pulses, the surface state of
the catalyst can be changed signicantly in a controlled manner. Thus, a sequence of innitesimal steps is used to
produce a nite change of the catalyst activity; hence, the technique can also be named chemical calculus.
566 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
BOTTOM-UP APPROACH TO KINETIC ANALYSIS
In the Bottom-Up Approach to kinetic analysis, a detailed analysis of the observed product distribution over
a broad range of reaction conditions is complemented with kinetic ngerprinting and reaction path analysis to
resolve the so-called minimal reactionnetwork. Subsequently, thermodynamic-consistent Langmuir-Hinshelwood-
Hougen-Watson-type rate equations are constructed that contain both kinetic and catalyst descriptors. The catalyst
descriptors that reect the effect of the catalyst on the stability of intermediates and transition states are identied
based on a kinetic analysis of the experimental observations and enable a physicochemical interpretation of the
inuence of the catalyst properties.
understanding of the inuence of the zeolite properties, such as Si/Al ratio (i.e., number and
strength of acid sites) and framework structure, on its catalytic behavior under working conditions.
The liquid-phase alkylation of benzene with 1-octene over a series of Y zeolites (22, 53) ex-
emplies the Bottom-Up Approach (see sidebar, Bottom-Up Approach to Kinetic Analysis). Al-
ternatively, in the Top-Down Approach, insight into the chemistry of the catalytic process is
translated into chemical rules that are implemented as algorithms in computer codes, allowing us
to generate detailed reaction networks consisting of several thousands of elementary steps (27, 28,
5456). This approach is illustrated for the catalytic cracking of hydrocarbons over a series of Y
and ZSM-5 zeolites (21, 57, 58) (see sidebar, Top-Down Approach to Kinetic Analysis).
In the third approach, molecular modeling is used to construct a minimal reaction network
for ethanol dehydration to ethylene over H-FAU, H-MOR, H-ZSM-5, and H-ZSM-22 together
with their corresponding kinetic coefcients (59, 60). Comparing simulated and experimentally
observed product yields over a broad range of reaction conditions then allows us to identify
important elementary steps that are missing from the network and/or must be found/calculated
more accurately by, for example, molecular dynamics.
Bottom-Up Approach: Liquid-Phase Alkylation of Benzene with 1-Octene
Alkylation of benzene with linear long-chain alkenes is used to synthesize linear alkyl benzenes
(LABs) as building blocks for the production of detergents. Conventional processes are based on
highly corrosive acid catalysts, such as aluminumtrichloride and hydrouoric acid (61). Therefore,
an efcient, recyclable, and environmentally friendly solid-acid catalyst is highly desired (62
64). Among various zeolites tested, Y zeolites have been shown to be suitable catalysts for the
alkylation of benzene with long-chain olens because they combine two important features: strong
TOP-DOWN APPROACH TO KINETIC ANALYSIS
In the Top-Down Approach to kinetic analysis, a detailed reaction network is automatically generated a priori based
on so-called reaction rules. The advantage of this approach is that all surface intermediates are explicitly accounted
for and no a priori reaction pathways are assumed. Databases, based either on experiment or on quantum chemical
calculations for model components, provide values for kinetic descriptors. If unavailable, estimates of kinetic and/or
catalyst descriptors can be obtained by regression of experimental data. Numerical integration or kinetic Monte
Carlo simulations then allowus to evaluate the relative importance of various reaction paths and the effect of catalyst
properties on the product yields and to dene optimal values for catalyst descriptors.
www.annualreviews.org Heterogeneously Catalyzed Reactions 567
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
Table 1 Properties of the Y zeolites used in the liquid-phase alkylation of benzene with 1-octene
CBV712 CBV720 CBV760
Bulk Si/Al ratio 6 13 30
Framework Si/Al ratio 21.2 26.2 75.8
Acid site concentration (mol kg
1
) 0.69 0.62 0.24
Al (IV) concentration (mol kg
1
) 0.75 0.61 0.22
BET surface area (m
2
g
1
) 730 780 720
Micropore volume (cm
3
g
1
) 0.25 0.27 0.25
Mesopore volume (cm
3
g
1
) 0.15 0.16
Unit cell size, a
o
(nm) 24.36 24.34 24.25
Total Al per unit cell 28.2 13.7 6.2
Framework Al per unit cell 8.6 7.1 2.5
acidic character and the ability to accommodate large molecules, such as alkylbenzenes, inside
their pores (65, 66). De Almeida et al. (65) reported that the selectivity to 2-phenyldodecane
increased with increasing acid strength in the alkylation of benzene with 1-dodecene over Y
zeolites performed in a batch-slurry reactor. However, Y zeolites are prone to deactivation, and
operation in batch mode does not allow us to exclude the inuence of coke deposition on the
product distribution. Hence, Cruciun et al. (53) studied, as a model reaction, the liquid-phase
alkylation of benzene with 1-octene over three commercial Y zeolites with bulk Si/Al ratios of 6,
13, and 30 (see Table 1) in a Robinson-Mahoney reactor operating in a continuous-ow (CSTR)
mode. This allowed extrapolation to zero time-on-stream and thus enabled evaluation of the
inuence of reaction conditions and acid properties on activity and selectivity in the absence of
deactivation. Temperatures ranged from 343 K to 373 K, benzene/1-octene feed molar ratios
ranged from 1 to 10, and 1-octene conversions were between 10% and 99%.
Denition of the minimal reaction network: kinetic ngerprinting and reaction path anal-
ysis. Representative proles of conversion versus site time (mol
H+
mol
1-octene
1 s) and selectivity
versus 1-octene conversion are shown in Figure 1. Clearly, the catalytic activity per acid site
increases with increasing Si/Al ratio, indicating that the acid strength is important in determining
the reactivity, whereas the selectivities are not affected. Also, for all catalysts, the experimental
data indicate that product selectivities are not inuenced by temperature.
Two main groups of components were obtained as reaction products: octene isomers
(2-octene, 3-octene, and 4-octene) and phenyloctanes (2-phenyloctane, 3-phenyloctane, and 4-
phenyloctane). Traces of di-alkylated benzenes were observed at low benzene-to-1-octene molar
feed ratios only. Branched octene isomers were not observed, indicating that the octene isomers
result from double-bond migration. As illustrated in Figure 2, kinetic ngerprinting using the
delplot technique (67) revealed that, when the feed consists of only 1-octene and benzene, 2-octene
and 2-phenyloctane are the only primary products, whereas 3-octene and 3-phenyloctane are sec-
ondary products and 4-octene and 4-phenyloctane are tertiary products. Hence, intramolecular
hydride shifts could be discarded as elementary steps in the reaction network because their occur-
rence would allow for the formation of all the reaction products as primary products. Therefore,
we can conclude that the formation of octene isomers occurs through a sequence of consecutive
protonation-deprotonation steps. Intermolecular hydride shift reactions could be eliminated as
well fromthe minimal reaction network because octane, octadienes, and components derived from
them are not observed under the mild conditions investigated.
568 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
0
0
20
20
40
40
60
60
80
CBV712 6
CBV720 13
CBV760 30
Si/Al
80
100
100
0 20 40 60 80 100
1
-
O
c
t
e
n
e

c
o
n
v
e
r
s
i
o
n

(
%
)
N
H+
/F
o,1-oct
(s)
0
20
40
60
80
100
S
e
l
e
c
t
i
v
i
t
y

(
%
)
S
e
l
e
c
t
i
v
i
t
y

(
%
)
1-Octene conversion (%)
0 20 40 60 80 100
1-Octene conversion (%)
0
10
20
30
40
2-Phenyloctane
3-Phenyloctane
2-Octene
3-Octene
4-Octene
4-Phenyloctane
a
b c
Figure 1
(a) Conversion of 1-octene as a function of site time and (b) octene isomer distribution and (c) phenyloctane
isomer distribution as a function of conversion over three Y zeolite catalysts at 373 K and a benzene/1-octene
feed molar ratio of 5. Experimental points and model simulations (lines) shown (adapted from Reference 53).
Figure 2a,b also illustrates that, although double-bond isomerization is faster compared with
benzene alkylation, the internal equilibrium distribution of the octenes is not reached. Conse-
quently, the double-bond isomerization and the alkylation take place on a comparable timescale;
hence, the two processes must be described as occurring simultaneously rather than as two con-
secutive, decoupled processes.
From the high values of the equilibrium coefcients (K

= 10
9
M
1
) for the formation of
phenyloctanes from benzene and octenes, we can conclude that the alkylation is practically
irreversible at the experimental conditions investigated. Hence, the minimal reaction network
shown in Figure 3 could be presented for the alkylation of benzene with octenes over Y zeolites.
The underlying mechanism is of the EleyRideal type, with benzene reacting from the liquid
phase and olens reacting as adsorbed species on the surface of the catalyst. EleyRideal-type
mechanisms have been proposed previously for the alkylation of benzene over large pore zeolites,
such as Y and beta (68, 69). Also, preliminary isothermal model discrimination allowed us to
eliminate models considering benzene adsorption (22).
Althoughthe nature of the protonatedhydrocarbonspecies inzeolite catalyststhat is, whether
they are carbenium ions or alkoxidesis the subject of ongoing investigation (7074), the mech-
anism is formulated based on carbenium ion chemistry that is assumed to occur on the catalyst
surface because the expressions for the rate equations of the elementary steps do not depend on
www.annualreviews.org Heterogeneously Catalyzed Reactions 569
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
0
20
40
60
80
100
S
e
l
e
c
t
i
v
i
t
y

(
%
)
2-Octene
3-Octene
4-Octene
2-Octene
3-Octene
4-Octene
2-Phenyloctane
3-Phenyloctane
4-Phenyloctane
2-Phenyloctane
3-Phenyloctane
4-Phenyloctane
0
5
10
15
20
25
S
e
l
e
c
t
i
v
i
t
y

(
%
)
0
0.5
1
1.5
2
2.5
0 20 40 60 80 100
1-Octene conversion (%)
S
e
l
e
c
t
i
v
i
t
y
/
c
o
n
v
e
r
s
i
o
n
0
0.05
0.1
0.15
0.2
0.25
0 20 40 60 80 100
1-Octene conversion (%)
S
e
l
e
c
t
i
v
i
t
y
/
c
o
n
v
e
r
s
i
o
n
0 20 40 60 80 100
1-Octene conversion (%)
0 20 40 60 80 100
1-Octene conversion (%)
Figure 2
First-order (top) and second-order (bottom) delplot method charts for octene isomers (left) and phenyloctanes
(right) for data over CBV760 at 343 K and benzene-1-octene molar feed ratio of 5. Markers indicate
experimental data, and solid lines are added to guide the eye. Dotted lines correspond to theoretical
selectivities calculated for equilibrium between octene isomers, respectively, between 2-, 3-, and
4-phenyloctane isomers (adapted from Reference 53).
the nature of the intermediates involved, and hence, the kinetic model is independent of the nature
of the reactive intermediates.
Kinetic analysis: denition of catalyst descriptors. As explainedinearlier work (53), the ratioof
the initial selectivities (i.e., in the limit of zero 1-octene conversion) of 2-octene to 2-phenyloctane
is related to the ratio of the rate coefcient for the deprotonation step leading to the formation of
2-octene to the rate coefcient for alkylation leading to 2-phenyloctane (Scheme 1).
Considering the benzene concentrations used in the kinetic experiments, the observed higher
initial selectivity for 2-octene as compared with that for 2-phenyloctane indicates that the rate
coefcient for deprotonation is higher than the rate coefcient for alkylation. Because the selec-
tivity of the primary products, 2-octene and 2-phenyloctane, is independent of temperature, we
can conclude that the activation energy for the deprotonation step leading to 2-octene is similar
to that for the alkylation step leading to 2-phenyloctane, as illustrated in the enthalpy diagram in
Figure 4, and that entropy favors double-bond isomerization over alkylation.
Achange in acid strength can lead to a change in the protonation enthalpy of the olens and the
phenyloctanes. Because H
iso(liq)
between the olens is independent of the catalyst, it follows from
570 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
+ + +
+ + +
+ H
+
+ H
+
+ H
+
+
H
+
+
H
+
+
H
+
+
H
+
+
+
+
1
7 8 9
2 3 4 5 6
10 11 12
Figure 3
Reaction network of elementary steps for the alkylation of benzene with octenes over ultrastable Y zeolites (22).
thermodynamic considerations that the effect of the catalyst on the protonation of the octenes can
be captured by the introduction of a single
cat
(H
pr,O
), where
cat
presents a change in going
from one catalyst to another. Hence,
cat
(H
pr,O
) can be seen as an average measure of the effect
of the catalyst on the stability of the intermediate octyl carbenium ions.
Also, activation energies of the elementary steps could be affected by a change in catalyst.
However, the observation that product selectivities are not affected by the catalyst indicates that
the effect on the activation energies for deprotonation and alkylation is similar; in other words,

cat
E
a,al k
=
cat
E
a,de pr,O
, 2.
indicating that the stability of all transition states (smaller ones for protonation/deprotonation
and larger ones for alkylation) are inuenced to the same extent by the zeolite. By applying the
EvansPolanyi principle (75), the change in activation energy of the alkylation step can be related
to the enthalpy change of the surface alkylation step,
cat
H
alk
, as follows:

cat
E
a,al k
=
al k

cat
H
al k
. 3.
Given that H
r(liq)
is independent of the catalyst, thermodynamic considerations allow us to
express
cat
H
alk
as

cat
H
al k
=
cat
H
pr,Phe O

cat
H
pr,O
. 4.
www.annualreviews.org Heterogeneously Catalyzed Reactions 571
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
+
+
+
k
pr,0
k
depr,1
k
depr,2
k
alk
k
pr,0
H
+
+ H
+
+
H
+
R
0
2PO
k
alk
K
pr,0
C
t
C
0
1oct
C
0
b
1 + K
iso, I
+ K
pr,0
C
0
1oct
+ K
pr,0
C
0
b
=
k
alk
k
pr
R
0
2O
k
depr,2
K
pr,0
C
t
C
0
1oct
1 + K
iso,I
+ K
pr,0
C
0
1oct
+ K
pr,0
C
0
b
=
k
alk
k
pr
S
0
2PO
S
0
2O
k
alk
C
0
b
k
depr,2
=
Scheme 1
Initial reaction scheme and initial rate equations.
Because the transition states for deprotonation and alkylation are similarly inuenced by the
catalyst, the extent to which the intermediate phenyloctane arenium ions are inuenced likely is
the same as that for the octyl carbenium ions; in other words, it can be assumed that

cat
H
pr,PheO
=
cat
H
pr,O
=
cat
H
pr
. 5.
E
n
t
h
a
l
p
y
Reaction coordinate
E
a,pr
H
pr,O
E
a,alk
E
a,de-pr,2
2C
+
8 ads
2C
=
8
E
a,alk
E
a,de pr,2
2PhO
2PhO
+
ads
1C
=
8
H
pr,PO
H
r
Figure 4
Energy diagram for the alkylation of benzene with octenes over a Y zeolite.
572 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
40 20 20 40 60 80 100
C
a
l
c
u
l
a
t
e
d

(
%
)
Experimental (%)
1-Octene
2-Octene
3-Octene
4-Octene
0
5
10
15
20
25
30
0 5 10 15 20 25 30
C
a
l
c
u
l
a
t
e
d

(
%
)
Experimental (%)
Benzene
2-Phenyloctane
3-Phenyloctane
4-Phenyloctane
40
20
0
20
40
60
80
100
0
Figure 5
Parity plots for the 1-octene and benzene conversions and for the molar yields of the individual octene and phenyloctane isomers in the
alkylation of benzene with octenes over CBV760, CBV720, and CBV712 in the temperature range 343373 K and with
benzene-to-olen molar ratios 1:110:1 (22).
Consequently,

cat
H
al k
= 0, 6.
and applying the EvansPolanyi principle (75) to the activation energy of the deprotonation
reaction leads to

c at
E
a,al k
=
c at
E
a,de pr,O
= 0 = (
pr
1)
c at
H
pr,O
, 7.
and hence to

pr
= 1. 8.
This indicates a late transition for protonation (75), which is generally accepted to be the case for
olen protonation in zeolites (7678). Also, because
cat
E
a,pr,O
=
cat
E
a,depr,O
+
cat
H
pr,O
, it can
be seen that

c at
E
a, pr,O
=
c at
H
pr
. 9.
Because the liquid-phase alkylation of benzene is performed at lowtemperatures (343 Kto 373 K),
enthalpy factors likely dominate over entropy factors (43), and entropy factors likely are indepen-
dent of the catalyst. Kinetic and thermodynamic analyses thus indicate that the effect of the catalyst
on the stability of the intermediates and transition states, and hence on the reaction kinetics, can
be captured by a single parameter, i.e.,
cat
H
pr
.
Estimates of the kinetic coefcients and the catalyst descriptor
cat
H
pr
referenced to
CBV760 are obtained by regression to the experimental data and are provided in Supplemen-
tal Table 1 (follow the Supplemental Material link from the Annual Reviews home page at
http://www.annualreviews.org). Figure 5 shows excellent agreement between simulated and
observed product yields.
Although the differences in protonation enthalpy between the three Y zeolites are within only
a few kJ mol
1
of each other, these differences have a signicant effect on the activity of the
catalyst because they also induce a change in the olen protonation activation energy. According
to Equation 9, an increase in the activation barrier for the olen protonation with 5 kJ mol
1
for CBV720 and 7 kJ mol
1
for CBV712 leads to a vefold and tenfold decrease, respectively,
of the protonation rate coefcient as compared with CBV760. At the same time, the 1-octene
protonation enthalpy is decreased from 33 kJ mol
1
for CBV760 to 28 kJ mol
1
for CBV720
www.annualreviews.org Heterogeneously Catalyzed Reactions 573
Supplemental Material
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
0.25
0.1
0.08
0.06
0.04
0.02
0
20
1
3
5
7
40
60
80
100
Total C
+
8
2-C
+
8
0.2
0.15
F
r
a
c
t
i
o
n
a
l

c
o
v
e
r
a
g
e

(

)
A
c
id
it
y
d
if
e
r
e
n
c
e
(
k
J
/
m
o
l)
S
i
t
e
t
i
m
e
(
s
)
0.1
0.05
0
a
0.15
0.1
0.08
0.06
0.04
0.02
0
20
1
3
5
7
40
60
80
100
0.12
0.09
F
r
a
c
t
i
o
n
a
l

c
o
v
e
r
a
g
e

(

)
A
c
id
it
y
d
if
e
r
e
n
c
e
(
k
J
/
m
o
l)
S
i
t
e
t
i
m
e
(
s
)
0.06
0.03
0
b
0.08
0.1
0.08
0.06
0.04
0.02
0
20
1
3
5
7
40
60
80
100
3-C
+
8
4-C
+
8
0.06
0.04
F
r
a
c
t
i
o
n
a
l

c
o
v
e
r
a
g
e

(

)
A
c
id
it
y
d
if
e
r
e
n
c
e
(
k
J
/
m
o
l)
S
i
t
e
t
i
m
e
(
s
)
0.02
0
c
0.08
0.1
0.08
0.06
0.04
0.02
0
20
1
3
5
7
40
60
80
100
0.06
0.04
F
r
a
c
t
i
o
n
a
l

c
o
v
e
r
a
g
e

(

)
A
c
id
it
y
d
if
e
r
e
n
c
e
(
k
J
/
m
o
l)
S
i
t
e
t
i
m
e
(
s
)
0.02
0
d
Figure 6
Model prediction for acid site coverage at 373 K and feed benzeneto-1-octene ratio of 5:1 as a function of the acid strengthrelated
parameter, H
pr
, and the variation of the coverage with the site time: (a) total octyl carbenium ion coverages, (b) 2-octyl cation
coverages, (c) 3-octyl cation coverages, and (d ) 4-octyl cation coverages (22).
and to 26 kJ mol
1
for CBV712. Thus, on CBV760, the protonation steps occur more easily
than on the other two Y zeolites. This results in a higher coverage of octyl carbenium ions on the
surface of CBV760, which in turn leads to higher alkylation rates. The fractional site coverage of
the various octyl carbenium ions, for different values of
cat
H
pr
, is illustrated in Figure 6 as a
function of site time.
Physicochemical signicance of
cat
H
pr
. To interpret the physicochemical signicance of

cat
H
pr
in terms of zeolite properties, the thermodynamic cycle presented in Figure 7 is fre-
quently used (43, 7981) because the protonation enthalpy of, for example, an octene can then be
expressed as
H
pr
= DPE +PA
ol e f
+ E
s tab
, 10.
with DPE (deprotonation energy of the zeolite) dened as the energy required to remove a proton
from the solid acid (increases for weaker acids), implying that acid strength differences reect
differences in stability between the protonated zeolites (ZO-H) and their conjugate bases (ZO

).
Ingeneral, acidity increases withincreasingpolarity of the ZO-Hbondandwithincreasingstability
of the conjugate base (i.e., with increasing ability of the conjugate base to accept and delocalize
574 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
H
H
+
+
+
H
prot
PA
1-octene
E
stab
DPE
Al Si
Si
O O

Al Si
Si
O O

Al Si
Si
O O

Al Si
Si
O O
Figure 7
Thermodynamic cycle explaining the contribution of the zeolite deprotonation energy (DPE), the alkene
proton afnity, and the stabilization energy (E
stab
) owing to connement of the free-gas-phase carbenium ion
in the zeolite to the protonation enthalpy of the olen.
the negative charge). PA
ol e f
is the proton afnity of the olen, and E
s tab
is the stabilization energy
owing to connement of the free-gas-phase octyl carbenium ion in the zeolite. The extent to
which guest molecules are solvated by connement in the zeolite frame depends on one hand on
the structure and composition of the zeolite and on the other hand on the size, shape, functional
groups, and charge of the guest molecule. These factors determine the number and nature of host-
guest interactions, such as, for instance, the weak induced-dipole interactions between alkanes and
the Brnsted acid site (43, 82, 83), H-bonding interactions between the hydrogen of the hydroxyl
group in alcohols and framework oxygens (84, 85), long-range electrostatic interaction between
charged species and the zeolite conjugate base, and van der Waals interactions (8486). Attractive
van der Waals interactions between the zeolite framework atoms and polarizable electron clouds
of the guest molecule typically increase when the conning space becomes smaller, up to the point
where Pauli repulsion takes over and the interaction becomes repulsive (i.e., sterical hindrance
comes into play). In general, E
stab
can be decomposed into a long-range electrostatic interaction
energy, E
elec
, and stabilizing van der Waals interactions, E
vdW
. Similar equations can be written
for the protonation of the phenyloctanes and also for the activation energies.
Hence,
cat
H
pr
can be rewritten as

c at
H
pr
=
c at
DPE +
c at
E
s tab
=
c at
DPE +
c at
E
el ec
+
c at
E
vdW
, 11.
www.annualreviews.org Heterogeneously Catalyzed Reactions 575
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
because the proton afnity of the olen is independent of the catalyst and cancels out.
cat
DPE
represents the difference in acid strength between two zeolites, whereas
cat
E
stab
represents the
difference in solvation ability of the considered conned species between the two zeolites. Clearly,
these two factors can contribute to the value of
cat
H
pr
. Recently, it was suggested that the
main factor responsible for activity differences between dealuminated Y zeolites is differences in
solvation ability,
cat
E
stab
, whereas differences in the Brnsted acid strength would play only a
minor role (43, 80). Thus, activity differences among dealuminated Y zeolites were attributed to
changes in the effective size of the supercages owing to the presence of varying amounts of extra
framework alumina. An increase in activity over dealuminated Y zeolites would then be related to
increased van der Waals interaction in supercages with smaller void space. If so, then the effect
of a change in catalyst is expected to be more pronounced on the more bulky conned hydro-
carbon species. However, as discussed above, in benzene alkylation, experimental observations
indicate that intermediates and transition states, which strongly differ in size and shape, are all
affected to the same extent in going from one Y zeolite to the other. Hence, for a given conned
hydrocarbon species, differences in van der Waals interactions with the large supercages of the
Y zeolites can be neglected; in other words, for a given intermediate,
cat
E
vdW

= 0. Given that
the charge on the intermediates (i.e., octyl carbenium ions and protonated phenyloctanes) is not
expected to change signicantly in going from one Y zeolite to the other, E
elec
is not expected to
change signicantly for a given hydrocarbon species; in other words, for a given intermediate,

cat
E
elec

= 0. Our analysis thus indicates that activity differences in benzene alkylation are mainly
related to acidity differences between the Y zeolites, in agreement with recent theoretical (87, 88)
and experimental studies on protolytic and catalytic cracking (89) and previous work on hydro-
cracking (20, 23).
Whether the physical origin of
cat
H
pr
is related to a change in acid strength, to a change in
van der Waals interactions, or to a combination of both factors is of less importance if the interest
remains restricted to modeling the effect of the catalyst on the product yields because both factors
are captured by the same catalyst descriptor. However, to suggest targets for synthesis strategies
that would enable us to tune the zeolite properties, a clear distinction between these two factors
is needed. This requires rst-principles methods that enable us to assess the contribution of
dispersive interactions to the thermodynamics of adsorbed species and transition states in zeolites.
Top-Down Approach: Catalytic Cracking of Hydrocarbons
Catalytic cracking of hydrocarbons over a solid acid catalyst is one of the most important oil
rening processes, and high-performance catalysts are needed to ensure optimal use of feed and
energy resources. In uid catalytic cracking (FCC), dealuminated Y zeolites and H-ZSM-5 are
mainly used as catalysts.
Although catalytic cracking of hydrocarbons proceeds via a complex reaction mechanism, the
thousands of reactions that occur can be grouped into a limited number of elementary reaction
families, including protolytic scission, (de)protonation, -scission/alkylation, hydride transfer,
and isomerization reaction steps. Hence, a kinetic analysis based on the automated generation of
a detailed reaction network is ideally suited to evaluate the inuence of reaction conditions and
catalyst properties on the product yields.
Detailed reaction network generation. Various algorithms and techniques are in use for the
automated generation of complex chemistries, and several excellent reviews on the topic are avail-
able (27, 28, 5456). We used a computer algorithm in which species are represented by vec-
tors and reactions are implemented as operations on Boolean relation matrices (90, 91). Specic
576 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
Table 2 Summary of the detailed generated reaction network for catalytic cracking of
2,2,4-trimethylpentane
Hydrocarbon species Elementary reaction steps
35 Alkanes 179 Protolytic scissions
78 Hydride transfers
94 Alkenes 137 Protonations
102 Carbenium ions 88 Hydride shifts
36 Methyl shifts
193 PCP-isomerizations
21 -Scissions
6 Alkylations
161 Deprotonations
231 Total 899 Total
reaction rules for each elementary reaction family are dened; hence, all possible reaction path-
ways, intermediates, and products are generated starting from a specied feed molecule. Known
kinetic coefcients can be provided from databases based on kinetic experiments or on quantum
chemical calculations for model components. Missing kinetic and/or catalyst descriptors can be t-
tedtokinetic experiments usingmodel molecules. Asummary of the resultingdetailedreactionnet-
work for 2,2,4-trimethylpentane is given in Table 2 to illustrate that even during cracking of a sin-
gle molecule a huge number of elementary reactions occur simultaneously: In total, 899 reactions
consisting of 179 protolytic scissions, 78 hydride transfers, 137 protonations, 88 hydride transfers,
36 methyl shifts, 193 PCP-isomerizations, 21 -scissions, 6 alkylations, and 161 deprotonations
between 231 species consisting of 35 alkanes, 94 alkenes, and 102 alkylcarbenium ions occur (21).
Kinetic and catalyst descriptors. The effect of the catalyst properties on the product yields was
evaluated for the cracking of two model hydrocarbons, 2,2,4-trimethylpentane (57) and methylcy-
clohexane (58), over a series of commercially available Yand MFI zeolites with varying Si/Al ratios.
The properties of the catalysts used and the experimental conditions investigated are provided in
Supplemental Tables 2 and 3. Dedicated kinetic experiments were performed in a reactor that
operates gradientlessly at high conversion because of the use of external recirculation (92).
Representative proles of conversion versus site time (mol
H+
mol
1-octene
1 s) and selectivity
versus conversion are shown in Figures 8 and 9. At equal conditions, the site time yields obtained
when cracking 2,2,4-trimethylpentane over the MFIs are a factor 34 lower than over the FAUs,
whereas for methylcyclohexane the site time yields are similar, showing that the lower activity
displayedby MFI for crackingof 2,2,4-trimethylpentane results fromreactant diffusionlimitations.
As illustrated in Figure 9, a unique relationship was found between the product selectivities
and conversion within a given framework type, independent of the Si/Al ratio of the zeolite.
This is valid for FAU as well as for MFI and for methylcyclohexane cracking as well as for
2,2,4-trimethylpentane cracking. Changing the framework alters the product selectivity versus
conversion prole owing to shape selectivity.
The main reaction route on FAU in 2,2,4-trimethylpentane cracking was found to be hydride
transfer followed by -scission leading to mainly C4 species, whereas on MFI protolytic scission
it is responsible for the formation of high amounts of C1C3 species, pointing toward the occur-
rence of transition-state shape selectivity hampering the bimolecular hydride transfer reaction.
Reactant shape selectivity can be ruled out because the kinetic diameter of 2,2,4-trimethylpentane
www.annualreviews.org Heterogeneously Catalyzed Reactions 577
Supplemental Material
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
80
60
40
20
0
0 50 100 150 200
C
o
n
v
e
r
s
i
o
n

[
m
o
l
%
]
Site time [mol H
+
s mol
1
]
80
60
40
20
0
0 50 100 150 200
C
o
n
v
e
r
s
i
o
n

[
m
o
l
%
]
Site time [mol H
+
s mol
1
]
80
60
40
20
0
0 25 50 75 100
C
o
n
v
e
r
s
i
o
n

[
m
o
l
%
]
Site time [mol H
+
s mol
1
]
80
60
40
20
0
0 20 40 60 80
C
o
n
v
e
r
s
i
o
n

[
m
o
l
%
]
Site time [mol H
+
s mol
1
]
a i-Octane b Methylcyclohexane
CBV500
CBV720
CBV760
LZ-Y20
Y62
CBV3020
E
CBV5524
G
CBV8014
Figure 8
Conversion of (a) i-octane and (b) methylcyclohexane on ve FAU zeolites (top) and three MFI zeolites (bottom) as a function of the site
time at 748 K and 7 kPa partial pressure. Solid lines represent trend lines obtained by tting the data series to a logarithmic curve
(adapted from References 57 and 58).
is 0.62 nm (93), whereas the effective catalytic pore size of MFI at 573 K lies between 0.662
and 0.727 nm and increases to 0.764 nm at 643 K (94). In methylcyclohexane cracking on FAU,
methylcyclohexane isomerization, followed by ring opening and subsequent cracking, is the main
reaction pathway, whereas on MFI, protolytic scission followed by cracking is predominant,
indicating that here too the bimolecular hydride transfer reaction is hampered and transition-state
shape selectivity is operative.
So far, only a fewkinetic models that explicitly account for the effect of the catalyst acidity onthe
product yields have been reported (95107). To the best of our knowledge, only one microkinetic
modeling study accounting for the effect of transition-state shape selectivity in hydroconversion
of n-octane on the Pt-H-ZSM-22 zeolite has been reported (108).
We have used the single-event microkinetic modeling approach (20, 23) to describe the con-
version of the hydrocarbons based on carbenium ion chemistry that is assumed to occur on the
catalyst surface (see sidebar, Single-Event Microkinetic Modeling Approach).
To account for the effect of variations in acidity on the observed catalytic behavior, a change
in protonation enthalpy,
cat
H
pr
, referenced to LZ-Y20 is sufcient because the absence of
selectivity differences when cracking alkanes or cycloalkanes on FAUor MFI zeolites indicates that
each reaction pathway is affected in the same way by a change in acid strength; in other words, all
transition states and surface intermediates are affected to the same extent. This results in a change
in activation energy of the so-called initiation reactions (i.e., protonation and protolytic scission),
whereas the activation energies of the surface reactions and deprotonations remain unaffected by
the zeolite acid strength. A variation in protonation enthalpy of 29 kJ mol
1
was found for the
578 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
SINGLE-EVENT MICROKINETIC MODELING APPROACH
In the single-event microkinetic modeling approach, the global symmetry numbers of the reactant and transition
state are factored out from the rate coefcient and calculated from statistical thermodynamics, leading to the
so-called single-event rate coefcient,

k:
k =

reac t
gl

gl

k,
in which
reac t
gl
and

gl
are the global symmetry numbers of the reactant and the transition state, respectively.
The single-event rate coefcient,

k, is assumed to depend only on the reaction family and the type of carbenium
ions (primary, secondary, or tertiary) involved. These assumptions allow for a drastic reduction in the number of
adjustable parameters to be tted to the experimental data.
ve faujasites (21), whereas for the three MFI zeolites, a variation in protonation enthalpy of 10 kJ
mol
1
was found (109).
To describe the effect of transition-state shape selectivity on the rates of hydride transfer of
gas-phase alkanes and cycloalkanes when cracking methylcyclohexane on MFI, it was sufcient to
40
30
20
MFI
MFI
MFI
MFI
FAU
FAU
FAU
MFI
FAU
FAU
10
0
S
e
l
e
c
t
i
v
i
t
y

[
m
o
l
%
]
0 20 40 60 80
Conversion [mol%]
CBV500
CBV720
CBV760
LZ-Y20
Y62
CBV3020
E
CBV5524
G
CBV8014
60
45
30
15
0
S
e
l
e
c
t
i
v
i
t
y

[
m
o
l
%
]
0 20 40 60 80
Conversion [mol%]
80
60
40
20
0
S
e
l
e
c
t
i
v
i
t
y

[
m
o
l
%
]
0 20 40 60 80
Conversion [mol%]
50
30
40
20
10
0
S
e
l
e
c
t
i
v
i
t
y

[
m
o
l
%
]
0 20 40 60 80
Conversion [mol%]
a Ethylene b Propane
c Propylene d Isobutane
Figure 9
Representative product selectivities as a function of methylcyclohexane conversion on ve FAU zeolites and three MFI zeolites at
748 K and 7 kPa partial pressure. The solid lines represent trend lines obtained by tting all FAU and MFI data to a second-order
polynomial curve (adapted from References 57 and 58).
www.annualreviews.org Heterogeneously Catalyzed Reactions 579
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
PROTOLYTIC CRACKING
Protolytic cracking is a monomolecular reaction and involves proton transfer from the zeolite to a C-C or C-H
of a physisorbed alkane. Assuming equilibrium for the physisorption step, the relationship between the observed
cracking rate, the intrinsic protolytic cracking rate coefcient, and the physisorption enthalpy and entropy can be
expressed as
r = kK
phys
p
i
and
r = Ae xp

E
a
RT

e xp

S
0
phys
R

e xp

H
0
phys
R

p
i
.
This leads to
E
c o mp
= E
a
+H
0
phys
and
A
c o mp
= A+

S
0
phys
R

for the composite activation energy, E


comp
, and the composite preexponential factor, A
comp
.
introduce a single shape-selectivity descriptor in the model:
E
a,MFI,ht f
= E
a,FAU,ht f
+
c at
E
a,h f t
. 12.
On MFI, an increase in activation energy of hydride transfer of 10 kJ mol
1
relative to FAU
was found (109), reecting that in MFI, Pauli repulsion comes into play; in other words, sterical
hindrance between the hydride transfer transition state and the zeolite wall becomes important in
the smaller MFI voids. The entropic contributions were assumed to be similar for both framework
types. Narasimhan et al. (108) also reported that the effect of shape selectivity in ZSM-22 can be
captured by a change in activation energy only.
However, Bhan et al. (110) recently emphasized the important role of the activation entropy
in protolytic cracking of alkanes in zeolites (see sidebar, Protolytic Cracking).
Earlier studies on protolytic cracking reported a similar intrinsic activation energy for pro-
tolytic scission on all zeolites, and observed differences in catalytic activity were explained by
differences in physisorption enthalpies between zeolites. Bhan et al. (110) pointed out that at typi-
cal protolytic cracking conditions, differences in physisorption enthalpies alone cannot account for
the observed differences in activity and selectivity among different zeolite framework types owing
to the compensation effect between adsorption enthalpy and entropy. Differences in activity were
attributed to differences in intrinsic preexponential factors as well as differences in intrinsic activa-
tion energies. From a dispersion-corrected periodic density functional theory (DFT-D) study of
adsorption equilibrium coefcients of n-alkanes in various zeolites, De Moor et al. (83) concluded
that changes in monomolecular cracking activity with carbon number and with zeolite framework
cannot be attributed solely to changes in the physisorption equilibrium coefcient. Given that
reported intrinsic activation energies are rather independent of the zeolite and the length of the
alkane (43, 111115), these authors concluded that the intrinsic preexponential factors, and thus
the intrinsic activation entropies, also play an important role. In addition, Iglesia and coworkers
580 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
(43, 110, 115) reported that the location of the Brnsted acid sites in a given framework type may
also inuence the observed product yields in monomolecular alkane cracking. With an increasing
fraction of protons located in the 8-MR pockets of MOR, a systematic increase in turnover rates
for cracking (C-C protonation) and hydrogenation (C-H protonation) of propane, butane, and
isobutane was observed, whereas the cracking-to-dehydrogenation ratio was found to decrease for
linear alkanes (C
3
H
8
, n-C
4
H
10
) and to increase for branched alkanes (iso-C
4
H
10
). In the smaller
8-MR pockets, the n-alkanes are only partially conned, and the observed behavior is reminiscent
to reactant shape selectivity. To interpret the observed reactant shape-selectivity effect in terms of
zeolite properties, it is instructive to consider the following thermodynamic cycles. The composite
activation energies for protolytic cracking and dehydrogenation of a given alkane at the 8-MRand
12-MR sites in MOR can be expressed as
E
a,C-C,8
+H
fys,8
= DPE
8
+PA
C-C
+ E
stab,TSC-C,8
13.
E
a,C-C,12
+H
fys,12
= DPE
12
+PA
C-C
+ E
stabTSC-C,12
14.
E
a,C-H,8
+H
f ys ,8
= DPE
8
+PA
C-H
+ E
stabTSC-H,8
15.
E
a,C-H,12
+H
f ys ,8
= DPE
12
+PA
C-H
+ E
stabTSC-H,12
. 16.
Here, E
a,C-X,8
and E
a,C-X,12
present the intrinsic activation energy for protolytic cracking and
dehydrogenation at the 8-MR or 12-MR site; H
fys,i
is the physisorption enthalpy of the alkane at
the 8-MR or 12-MR site; DPE
i
is the deprotonation energy of protons located at the 8-MR or 12-
MR sites of MOR; PA
C-C
and PA
C-H
are the gas-phase afnities for protonation of the alkane C-C
and C-H bond, respectively; and E
stab,TSC-X,i
is the stabilization energy owing to connement of
the respective transition states in the 8-MR pockets or in the 12-MR channels of MOR relative to
their gas-phase analogs. Similar equations can be considered for the intrinsic activation entropies.
Because composite activation parameters on the 8-MR and 12-MR sites are expressed relative
to the same gas-phase alkane, differences incomposite activationGibbs free energies,
site
G
comp,C-X
,
reect differences in free energy between the transition states conned in the 8-MR pockets and
the 12-MR channels. Given that both the n-alkane and the transition states are only partially
conned within the 8-MR pockets, the observed higher turnover frequencies on the 8-MR sites
indicate that the enthalpic penalty resulting frompartial connement of the transition states within
the 8-MR pockets is compensated for by a smaller loss in entropy, leading to a lower free energy
of the transition states in the 8-MR pockets and, hence, to higher composite activation energies
but to less negative composite activation entropies.
Because alkane cracking and dehydrogenation occur through parallel paths, differences in
composite activation energies of the competing steps,
r
G
comp,i
, reect differences in free energy of
the transition states for the two reactions because they are referenced to the same gas-phase alkane.
The following equations can be used to interpret the enthalpic effect of the zeolite properties on
the observed cracking-to-dehydrogenation ratio of a given alkane on the 8-MR and 12-MR sites:

r
E
a,8
= (PA
C-C
PA
C-H
) +(E
s tabTSC-C,8
E
stabTSC-H,8
), 17.
and

r
E
a,12
= (PA
C-C
PA
C-H
) +(E
stabC-C,12
E
stabTSC-H,12
), 18.
because H
fys,i
cancels out from the equations. Also, for a given acid site, the DPE term rigor-
ously cancels out, and in this particular case, acidity differences clearly are of no importance. As
Gounder & Iglesia (43) pointed out, the difference in intrinsic activation energy between crack-
ing and hydrogenation in a given conning space can thus be related mainly to the difference in
stability between the protonated gas-phase species because the transition states for cracking and
www.annualreviews.org Heterogeneously Catalyzed Reactions 581
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
hydrogenation are similar in size, shape, and composition, and it is reasonable to assume that they
are solvated to the same extent within a given conning space. Hence,

r
E
a,8

=
r
E
a,12

= (PA
C-C
PA
C-H
), 19.
indicating that the difference in activation energy between cracking and hydrogenation is inde-
pendent of the zeolite environment. Given the observed difference in cracking-to-hydrogenation
ratio between 8-MR and 12-MR sites in MOR, entropy differences induced by connement of
the two transition states in the different surrounding voids clearly can play a role in explaining the
observed product yields; in other words,

r
S
a,8
= (S
stabTSC-C,8
S
stabTSC-H,8
) =
r
S
a,8
= (S
stabTSC-C,12
S
stabTSC-H,12
). 20.
To explain the observed selectivity behavior with varying fractions of 8-MR sites in MOR,
Gounder & Iglesia (43, 110, 115) thus emphasized the dominant role of entropy in the stabiliza-
tion of adsorbed species and transition states within different conning spaces in a given zeolite
when competing reactions proceed via transition states of similar size, shape, and composition;
therefore, differences in the enthalpy factors more or less cancel out in Equations 17 and 18.
The effect of the surrounding environment (8-MR or 12-MR sites) on a given reaction (i.e.,
the shape-selectivity effect) can, however, also be interpreted using the expressions

s i te
E
a,C-C
=
s i te
DPE +(E
stabTSC-C,12
E
stabTSC-C,8
) (H
f ys ,12
H
f ys ,8
) 21.
and

s i te
E
a,C-H
=
s i te
DPE +(E
stabTSC-H,12
E
stabTSC-H,8
) (H
f ys ,12
H
f ys ,8
), 22.
because the PA term cancels out from the equations. Usually, van der Waals contributions in the
physisorbed alkanes and transition states are similar; hence,

s i te
E
a,C-C
=
s i te
DPE +
s i te
E
el ec TSC-C
23.
and

s i te
E
a,C-H
=
s i te
DPE +
s i te
E
el ec TSC-H
. 24.
In the particular case considered (i.e., competing reactions of a given alkane in differently shaped
surroundings in a given zeolite with a given Si/Al ratio but in which the protons on the 8-MR sites
have been selectively replaced by monovalent sodium ions), it can be expected that the 8-MR and
12-MR sites have more or less equal strength and that the long-range electrostatic stabilization of
the charged transition states by the MOR framework is rather similar too.
Inthe more general case (i.e., for a givenreactionina different zeolite framework witha different
Si/Al ratio), Equation 23 or 24, together with its counterpart for the intrinsic activation entropy,
clearly shows that, in principle, changes in the intrinsic activation energy and/or entropy of a given
reaction occurring on sites that are located in a different zeolite framework with a different Si/Al
ratio can be induced by differences in long-range electrostatic stabilization between the zeolites,
but also by differences in acid strength. Moreover, differences in physisorption enthalpy and
entropy of the alkane inthe different conning spaces canbe expected to contribute to the observed
shape-selectivity effect as well. Thus, in general, depending on the particular reaction and on the
particular zeolite frameworks under consideration, one or the other factor can become dominant
in determining enthalpy and/or entropy contributions to explain shape-selectivity effects.
Moreover, Tranca et al. (116) recently reported on a molecular mechanicsMonte Carlo study
of adsorption of light alkanes in H-ZSM-5, which suggested that the physisorption entropy sig-
nicantly depends on temperature. At typical cracking temperatures (700 K), the calculated alkane
adsorption entropy was found to be signicantly less negative than at the temperatures usually
582 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
employed for adsorption measurements (<350 K), and adsorption entropy losses were found to
decrease with increasing temperature. The authors reported intrinsic activation entropies that
are less negative by 5560 J mol
1
K
1
when using their calculated temperature-dependent ad-
sorption parameters. Based on the intrinsic rate parameters obtained from static periodic DFT-D
calculations at 0 K, these authors also found that the intrinsic enthalpy barrier for monomolec-
ular cracking is practically independent of the alkane chain length, whereas the intrinsic en-
tropy of activation increases with the carbon number. Therefore, the use of available measured
temperature-independent adsorption entropies to estimate intrinsic activation entropies from ex-
perimental kinetic data could lead to biased values and, hence, to erroneous interpretations of the
physicochemical origin of shape-selectivity effects. Currently, temperature-dependent measured
adsorption data are not available to validate the calculated adsorption entropies.
Zimmerman et al. (117) suggested another complicating factor in extracting kinetic coef-
cients from experimental data via a recent dynamic rst-principles quantum mechanics/molecular
mechanics (QM/MM) study of n-pentane cracking in ZSM-5. These authors used quasi-classical
trajectory simulations that begin from a given transition state to identify the pathways via which
products are formed at a working reaction temperature and to determine the distribution of prod-
ucts formed from the considered transition state. Cracking transition states for n-pentane were
found to lead to a metastable intermediate with the proton shared between two hydrocarbon frag-
ments. The barriers to escape fromthis local metastable minimumto lower lying minima are small,
and one-picosecond quasi-classical trajectory simulations performed at 773 K indicated that the
dynamic paths are substantially more diverse than the static potential energy paths (at 0 K). The
authors reported a nearly tenfold change in the branching ratio between C2/C3 cracking chan-
nels upon inclusion of post-transition-state dynamics as compared with static electronic structure
calculations. The results of this study suggest that the product selectivity would be determined
mainly by the dynamics after passing through the rate-limiting transition state for C-H or C-C
protonation and not by the potential energy barriers alone. Selectivity in zeolite catalysis thus
seems to be determined mainly by high-temperature pathways that differ signicantly from the
0 K potential surfaces. Thus, dynamic effects too can contribute to the shape selectivity induced
by zeolites.
Clearly, further systematic joined experimental and theoretical studies on adsorptionand crack-
ing of various hydrocarbons over well-dened and well-characterized zeolites are needed to un-
ravel the physicochemical origin of the shape-selectivity effect of zeolites and to obtain quantitative
insight into the effect of zeolite properties on their cracking behavior.
Molecular Modeling: Ethanol Dehydration
Owing to its large scale of production, bioethanol can be viewed as a potential renewable feedstock
for the production of other chemicals, in the same way that naphtha is used today (118). In partic-
ular, zeolite-catalyzed conversion of bioethanol offers promising perspectives for the sustainable
production of ethylene (119, 120), butenes (121), and higher liquid hydrocarbons (122).
Depending on the reaction conditions, ethanol conversion over zeolites leads to the formation
of diethyl ether, ethylene, or gasoline. At low temperatures (323373 K), dehydration of ethanol
produces mainly diethyl ether, whereas at higher temperatures (373473 K), ethylene is the
dominant product (118). The dehydration reaction is proposed to proceed via two competitive
reaction paths (123). One is the monomolecular dehydration of ethanol to ethylene and water, and
the other is the bimolecular dehydration of ethanol to diethyl ether and water. Also, a consecutive
path in which diethyl ether is further converted to ethylene and ethanol has been reported (124).
From temperatures above 423 K, ethylene oligomerization, leading to the formation of higher
www.annualreviews.org Heterogeneously Catalyzed Reactions 583
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
Al
O O
H
O
H
Al
O
O O
O
O O
O
H
O
H
O
EtOH
g
D
1
D
2
M
1
C
1
M
2
Ethoxy Ethene
ads
DEE
ads
DEE
g
DEE
g
EtOH
g
EtOH
g
H
2
O
g
H
2
O
g
Al
H
Al
H
Al
O
O
H
Al
O
H
O
O
Al
O
H
O
O
H
H
O
Al
O O
O
H
H
H
Al
O O
O
H
H
Al
H
O
H
Al
O O
H
EtOH
g
EtOH
g
EtOH
g
H
2
C = CH
2
H
2
C = CH
2
H
2
C = CH
2g
H
2
C = CH
2g
1
2
3
4
5
6
7 8
9
10
11
Figure 10
Reaction network of elementary steps for the zeolite-catalyzed dehydration of ethanol (adapted from Reference 59). The
rate-determining step along each reaction path is indicated in red. For the quasi-equilibrated steps, forward and backward arrows are
indicated, whereas only the forward arrow is indicated for the rate-determining step along each reaction path. For all elementary steps,
forward and backward steps have been considered explicitly in the microkinetic simulations.
hydrocarbons, also occurs, even on these monofunctional acid catalysts. And at temperatures
above 473 K, the higher hydrocarbons constitute the dominant products (118). In analogy with
the methanol-to-gasoline process, a hydrocarbon pool mechanism has been suggested to explain
the production of the higher hydrocarbons (125).
Because the number of elementary reaction steps involved in ethanol dehydration remains
limited, it is feasible to compute thermodynamics and kinetics for all elementary steps in various
types of zeolites using quantum chemical calculations. Therefore, we have chosen ethanol
dehydration as a probe reaction to obtain a basic understanding of the inuence of the zeolite
properties on its reactivity and selectivity behavior in bioalcohol conversion over various types of
zeolites using a rst-principles-based microkinetic modeling approach. Relevant thermodynamic
and kinetic parameters for the elementary steps involved in ethanol dehydration over H-FAU, H-
MOR, H-ZSM-5, and H-ZSM-22 were obtained fromstatic dispersioncorrected periodic DFT-
Dcalculations at 0 Kand statistical thermodynamics and fed to a microkinetic model to investigate
the inuence of reaction conditions and zeolite properties on reactivity and product selectivity (59,
60). The reaction scheme considered in the microkinetic modeling is illustrated in Figure 10,
584 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
Table 3 Elementary steps and the three reaction paths for the assumed mechanism for the
dehydration of ethanol
a
Step Elementary step

A B C
1 C
2
H
5
OH
(g)
+

M
1
1 1 0
2 M
1
M
2
1 0 0
3 M
2
Ethoxy + H
2
O
(g)
1
0 0
4 Ethoxy Ethene
ads
1 0 0
5 Ethene
ads
Ethene
(g)
+

1 0 1
6 M
1
+C
2
H
5
OH
(g)
D
1
0 1 0
7 D
1
D
2
0 1 0
8 D
2
DEE
ads
+ H
2
O
(g)
0
1
0
9 DEE
ads
DEE
(g)
+

0 1 1
10 DEE
ads
A = r
2
C
1
0 0
1
11 C
1
Ethene
ads
+C
2
H
5
OH
(g)
0 0 1
Path A C
2
H
5
OH
(g)
Ethene
(g)
+ H
2
O
(g)
Path B 2 C
2
H
5
OH
(g)
DEE
(g)
+ H
2
O
(g)
Path C DEE
(g)
Ethene
(g)
+ C
2
H
5
OH
(g)
a
The rate-determining step along each reaction path is indicated in red. For the quasi-equilibrated steps, forward and
backward arrows are indicated, whereas only the forward arrow is indicated for the rate-determining step along each path.
For all elementary steps, forward and backward steps have been considered explicitly in the microkinetic simulations.
and an overview of the three reaction paths is given in Table 3. Kinetic experiments were
performed over H-ZSM-5 in a broad range of conditions and were used for validation purposes.
Our periodic DFT-D calculations indicate that the monomolecular pathway (see Path
A in Table 3) occurs via an intermediate chemisorbed monomer M1 (step 1) that rear-
ranges to a physisorbed monomer M2 (step 2), which undergoes water elimination to form
surface-bound ethoxide (step 3). The water elimination from M2 occurs through activation
of the alcohol C

-O bond by the zeolite proton for attack of the carbon atom by the
aluminum-bound oxygen adjacent to the acid site, resulting in a [CH
3
CH
2
OH
2
]
+
cationic
transition state. This mechanismis consistent with the secondary kinetic isotopic effects measured
for ethylene synthesis rates using C
2
D
5
OH, which implies that the kinetically relevant step
involves the cleavage of the C

O bond via a carbenium-ion-like transition state (123). Other


mechanisms were found to be signicantly more activated. Ethoxide is then decomposed
into physisorbed ethene (Ethene

) (step 4) that desorbs with regeneration of the active site


(step 5).
In the bimolecular path (see Path B in Table 3), an additional ethanol molecule coadsorbs
with M1 to form a chemisorbed dimer D1 (step 6), in which the proton is shared between the two
ethanol molecules. Rearrangement of D1 to the physisorbed dimer D2 (step 7) is required for water
elimination to generate chemisorbed diethyl ether (DEE

) (step 8). Diethyl ether then desorbs


from the zeolite (step 9) with regeneration of the acid site. The bimolecular elimination toward
DEE

involves activationof the C

-Obondof one alcohol molecule by the zeolite protonfor attack


of its carbon atom by the OH of the other alcohol molecule, leading to a [C
2
H
5
OHC
2
H
5

H
2
O]
+
cationic transition state. The second ethanol molecule in the transition state functions as
a solvating agent to stabilize the carbenium-ion-like C
2
H
5
+
fragment (80, 123, 126). Reaction of
M1 or M2 with gas-phase ethanol via an Eley-Rideal mechanism was found to be signicantly
more activated.
www.annualreviews.org Heterogeneously Catalyzed Reactions 585
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
Table 4 Rate coefcients (s
1
) at 473 K of the rate-determining step along the monomolecular,
the bimolecular, and the consecutive path in the dehydration of ethanol over zeolites H-FAU,
H-MOR, H-ZSM-5, and H-ZSM-22
H-FAU H-MOR
a
H-ZSM-5 H-ZSM-22
Reaction k (s
1
) k (s
1
) k (s
1
) k (s
1
)
M2 Ethoxy + H
2
O
(g)
2.1 10
2
7.3 10
2
1.0 3.6 10
1
D2 DEE

+ H
2
O
(g)
3.7 10
2
7.3 10
1
2.4 10
2
8.0 10
4
DEE

C1 1.2 10
4
2.6 10
5
5.3 10
4
2.6 10
5
a
12-MR channels.
In the consecutive path (see Path C in Table 3), a very late and loose transition state is formed
via attack of the -hydrogen of adsorbed diethyl ether by the aluminum-bound oxygen adjacent to
the acid site, leading to simultaneous C

-H and C

-O bond scission and resulting in physisorbed


ethylene and ethanol (step 10). Desorption of ethanol (step 11) and ethylene (step 5) then restores
the acid site. Several other mechanisms for the consecutive path were investigated and found to
be signicantly more activated.
Microkinetic simulations in a broad range of conditions revealed that steps 3, 8, and 10 are the
rate-determiningsteps alongthe monomolecular, bimolecular, andconsecutive paths, respectively.
All other reaction steps were found to be quasi-equilibrated. Rate coefcients at 473 K for these
rate-determining steps are provided in Table 4. As expected from the values of the rate coef-
cients, Figure 11 shows that at 473 K, over the four zeolites, diethyl ether is the dominant
reaction product and, hence, that the bimolecular path is by far the dominant reactant path. Also,
the 10-MR zeolites, H-ZSM-5 and H-ZSM-22, are found to be much more reactive than the
12-MR zeolites, H-MOR and H-FAU (60). H-MOR, H-ZSM-5, and H-ZSM-22 have higher
DPE values (i.e., lower acid strengths) in comparison with H-FAU, but they provide higher sta-
bilization energies owing to connement of the bimolecular transition state in the zeolite pores.
Among the four zeolites, H-ZSM-5 and especially H-ZSM-22 effectively stabilize the cationic
transition state by exerting larger dispersive, hydrogen-bonding and/or electrostatic interactions
than H-FAU and H-MOR, leading to signicantly higher reactivity, by several orders of magni-
tude, of H-ZSM-5 and H-ZZSM-22 (see Figure 12).
As illustrated in Supplemental Figure 1, a good agreement between simulated and exper-
imental conversions on H-ZSM-5 is obtained at temperatures ranging from 470 K to 520 K.
However, with increasing conversion, a systematic underprediction of the ethylene selectivity and
overprediction of the diethyl ether selectivity can be noticed. Analysis of the surface coverages
(see Supplemental Figure 2) revealed that at higher conversion, adsorbed diethyl ether becomes
the most abundant surface intermediate, indicating that the observed deviation with the experi-
mental data most likely results from an underestimation of the preexponential of step 8 (i.e., the
rate-determining step of the consecutive path) and/or the occurrence of a reaction step for the
conversion of diethyl ether to ethylene and ethanol that has not been considered yet in the current
reaction network.
Because we have calculated the entropic contributions based on the harmonic oscillator approx-
imation, in view of the rather late and loose transition state of reaction 8, the approximation likely
performs poorly for the calculation of the activation entropy of this step. For loosely bonded
complexes in zeolites where there are many soft degrees of freedom, entropy losses calculated
based on the harmonic oscillator approximation are overestimated (117, 127). Therefore, the cal-
culated free-energy barrier of step 8 could be too high owing to the use of the harmonic oscillator
586 Reyniers

Marin
Supplemental Material
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
0
0 1 2 3 4
10
20
30
C
o
n
v
e
r
s
i
o
n

(
%
)
Site time (N
H
+/F
EtOH,0
)
H-ZSM-22
H-MOR
H-ZSM-5
H-FAU
0
0 1 2 3 4
10
20
30
D
E
E

y
i
e
l
d

(
%
)
Site time (N
H
+/F
EtOH,0
)
H-ZSM-22
H-MOR
H-ZSM-5
H-FAU
0 1 2 3 4
C
2
H
4

y
i
e
l
d

(
%
)
Site time (N
H
+/F
EtOH,0
)
H-ZSM-22
H-MOR
H-ZSM-5
H-FAU
0
0.4
0.2
0.6
0.8
1
a
b c
Figure 11
(a) Simulated ethanol conversion and (b) diethyl ether and (c) ethylene yield as a function of site time over H-FAU (Si/Al = 47),
H-MOR (Si/Al = 95), H-ZSM-5 (Si/Al = 95), and H-ZSM-22 (Si/Al = 35) at 473 K and inlet partial pressure of ethanol of 10 kPa.
approximation. An increase of the preexponential of step 8 with a factor 10, corresponding to a
decrease in activation entropy of approximately 20 J mol
1
K
1
(in the order of the accuracy that
can be expected fromthe harmonic oscillator approximation), results in very good agreement with
the experimental data, as shown in Figure 13. However, further exploration of the free-energy
surface using more advanced, ab initio dynamic simulations is needed to identify possible missing
steps in the network and/or to obtain a more accurate calculation of reaction free-energy barriers.
CONCLUSIONS AND FUTURE OUTLOOK
Chemical kinetics seeks to unravel the reaction mechanism at the molecular level and to provide a
quantitative description of the reaction rates. Undoubtedly, microkinetic models are indispensable
for chemical process optimization and design. In addition, their ability to incorporate a wealth of
knowledge available from surface science and operando studies, experimental kinetic studies, and
quantum chemistry makes them an ideal tool in the design of high-performance catalysts. In this
review, we have presented three alternative but complementary approaches for the kinetic analysis
of zeolite-catalyzedreactions toillustrate the potential of microkinetic modelingfor rational design
of zeolite catalysts with improved activity and selectivity.
Although rst-principles calculations can already provide valuable insights into the reaction
mechanismand help to unravel the inuence of the catalyst properties on the kinetics of individual
elementary steps, full-edged, rst-principles-based microkinetic simulations remain out of
www.annualreviews.org Heterogeneously Catalyzed Reactions 587
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
H-FAU 7.0
6.3
5.6
4.9
4.2
3.5
2.8
2.1
1.4
0.7
0.0
H-MOR (12-MR)
H-ZSM-5 H-ZSM-22
H-FAU H-MOR
H-ZSM-5 H-ZSM-22
H
y
d
r
o
g
e
n

b
o
n
d
i
n
g
E
l
e
c
t
r
o
s
t
a
t
i
c

p
o
t
e
n
t
i
a
l

(
e
V
)
E
comp,ZeOH
E
stab,ZeOH
DPE + PA
E
comp,H-FAU
E
stab,H-FAU
+ DPE + PA

cat
E
comp
= E
comp,ZeOH
E
comp,H-FAU
E
comp
= E
stab,ZeOH
E
stab,H-FAU
H
ads,D2
DPE
PA
E
stab
= E
vdW
+ E
stab,DFT
E
stab,vdW
= E
comp,vdW
DPE PA
E
stab,DFT
= E
comp,DFT
DPE PA
E
stab
TS
B
D2
E
comp
E
a,8
ZeO

+ [C
2
H
5
OH---C
2
H
5
---H
2
O]

(g)
ZeOH + 2C
2
H
5
OH
(g)
ZeO

+ H
+
2C
2
H
5
OH
(g)
kJ/mol E
comp
DPE E
stab
E
stab,vdW
E
stab,DFT
(E
comp
) (DPE) (E
stab
) (E
stab,vdW
) (E
stab,DFT
)
H-FAU 60 1,181 462 71 391 0 0 0 0 0
H-MOR 63 1,211 495 83 412 3 30 33 13 20
H-ZSM-5 100 1,210 531 97 434 40 29 69 27 42
H-ZSM-22 109 1,210 540 111 429 49 29 78 41 37
O
1
O
1
O
1
O
1
O
2
O
2
O
2
O
2
C
1
C
1
C
1
C
1
C
3
C
3
C
3
C
3
C
4
C
4
C
4
C
4
C
2
C
2
C
2
C
2
Figure 12
Thermodynamic cycle (top) explaining the dispersion (E
vdW
) and long-range electrostatic (E
DFT
) contribution to the stabilization
energy (E
stab
) (middle) of transition state TS
B
(bottom left) of the rate-determining step of the bimolecular reaction path (reaction 8 in
Figure 10) owing to connement in H-FAU (Si/Al = 47) and 12-MR channels of H-MOR (Si/Al = 95), H-ZSM-5 (Si/Al = 95), and
H-ZSM-22 (Si/Al = 35). Deprotonation energies are calculated using QMPot(MP2//B3LYP:GULP). Electrostatic potential (bottom
right) for TS
B
mapped on a plane that crosses the O
1
, C
1
, and O
2
atoms. The atoms of the acid site do not belong to the plane and are
not shown (adapted from Reference 59).
reach in the foreseeable future. However, the traditional trial-and-error experimental approach
to heterogeneous catalyst design has become far too expensive and time consuming. From the
three approaches presented in this review, suggesting targets for synthesis strategies that enable
us to tune the zeolite properties clearly will require a very close intertwinement of rst-principles
theoretical and experimental kinetic studies. In particular, systematic joined experimental and
theoretical studies on adsorption and reaction of various hydrocarbons and oxygenates over
588 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
C
o
n
v
e
r
s
i
o
n
/
s
e
l
e
c
t
i
v
i
t
y
Temperature (K)
100
80
60
40
20
0
470 480 490 500 510 520
X (%)
S-DEE (%)
S-C
2
H
4
(%)
Figure 13
Ethanol conversion (X), diethyl ether (S-DEE), and ethylene (S-C
2
H
4
) selectivity as a function of
temperature at space time W
cat
/F
EtOH,0
= 6.5 kg s mol
1
and P
EtOH,0
= 24 kPa over H-ZSM-5.
Experimental data are indicated with their 95% condence interval. Model simulations were obtained by
adjusting (see text) the ab initiocalculated preexponential of reaction 11 in Figure 10.
well-dened and well-characterized zeolites are needed to unravel the physicochemical origin of
the shape-selectivity effect of zeolites and to obtain quantitative insight into the effect of zeolite
properties on their catalytic behavior under working conditions.
DISCLOSURE STATEMENT
The authors are not aware of any afliations, memberships, funding, or nancial holdings that
might be perceived as affecting the objectivity of this review.
ACKNOWLEDGMENTS
This work was carried out using the Stevin Supercomputer Infrastructure at Ghent University and
is supported by the Long-Term Structural Methusalem Funding from the Flemish Government
and the Interuniversity Attraction Poles Programme Belgian State Belgian Science Policy.
LITERATURE CITED
1. Bartholomew CH, Farrauto RJ. 2005. Fundamentals of Industrial Catalytic Processes. 2nd ed. New York:
Wiley-AIChE
2. Keil FJ. 2013. Complexities in modeling of heterogeneous catalytic reactions. Comput. Math. Appl.
65:167497
3. Marin GB, ed. 2007. Chemical Engineering Kinetics. Adv. Chem. Eng. 32. London: Academic
4. Ertl G. 2009. Reactions at Solid Surfaces. New York: Wiley & Sons
5. Chorkendorff I, Niemantsverdriet JW. 2003. Concepts of Modern Catalysis and Kinetics. Weinheim: Wiley-
VCH
6. Campbell CT. 1989. Studies of model catalysts with well-dened surfaces combining ultrahigh-vacuum
surface characterization with medium-pressure and high-pressure kinetics. Adv. Catal. 36:154
www.annualreviews.org Heterogeneously Catalyzed Reactions 589
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
7. Dumesic JA, Rudd DF, Aparicio LM, Rekoske JE, Trevino AA. 1993. The Microkinetics of Heterogeneous
Catalysis. Washington, DC: Am. Chem. Soc.
8. Masel RI. 2001. Chemical Kinetics and Catalysis. New York: Wiley-Intersci.
9. Bell AT, Head-Gordon M. 2011. Quantummechanical modeling of catalytic processes. Annu. Rev. Chem.
Biomol. Eng. 2:45377
10. Salciccioli M, Stamatakis M, Caratzoulas S, Vlachos DG. 2011. A review of multiscale modeling of
metal-catalyzed reactions: mechanism development for complexity and emergent behavior. Chem. Eng.
Sci. 66:431955
11. Nrskov JK, Bligaard T, Rossmeisl J, Christensen CH. 2009. Towards the computational design of solid
catalysts. Nat. Chem. 1:3746
12. Hansen N, Keil FJ. 2013. Multiscale approaches for modeling hydrocarbon conversion reactions in
zeolites. Chem. Ing. Tech. 85:41319
13. Sabbe MK, Reyniers MF, Reuter K. 2012. First-principles kinetic modeling in heterogeneous catalysis:
an industrial perspective on best-practice, gaps and needs. Catal. Sci. Technol. 2:201024
14. van Santen RA, Sautet P, eds. 2009. Computational Methods in Catalysis and Material Science. Weinhem,
Ger.: Wiley-VCH
15. van Santen RA, Neurock M. 2006. Molecular Heterogeneous Catalysis: A Conceptual and Computational
Approach. Weinheim, Ger.: Wiley-VCH
16. Smit B. 2008. Molecular simulations of zeolites: adsorption, diffusion, and shape selectivity. Chem. Rev.
108:412584
17. Kim J, Smit B. 2012. Efcient Monte Carlo simulations of gas molecules inside porous materials.
J. Chem. Theory Comput. 8:233643
18. Gubbins KE, Liu YC, Moor JD, Palmera JC. 2011. The role of molecular modeling in conned systems:
impact and prospects. Phys. Chem. Chem. Phys. 13:5885
19. Cheng J, Hu P. 2008. Utilization of the three-dimensional volcano surface to understand the chemistry
of multiphase systems in heterogeneous catalysis. J. Am. Chem. Soc. 130:1086869
20. Martens GG, Marin GB, Martens JA, Jacobs PA, Baron GV. 2000. A fundamental kinetic model for
hydrocracking of C
8
to C
12
alkanes on Pt/US-Y zeolites. J. Catal. 195:25367
21. Van Borm R, Reyniers MF, Marin GB. 2012. Catalytic cracking of alkanes on FAU: single-event mi-
crokinetic modeling including acidity descriptors. AIChE J. 58:220215
22. Craciun I, Reyniers MF, Marin GB. 2012. Liquid-phase alkylation of benzene with octenes over Y
zeolites: kinetic modeling including acidity descriptors. J. Catal. 294:13650
23. Thybaut JW, Marin GB, Baron GV, Jacobs PA, Martens JA. 2001. Alkene protonation enthalpy de-
termination from fundamental kinetic modeling of alkane hydroconversion on Pt/H-(US)Y-zeolite.
J. Catal. 202:32439
24. Marin GB, Yablonsky GS. 2011. Kinetics of Chemical Reactions: Decoding Complexity. Weinheim, Ger.:
Wiley-VCH
25. Uhe A, Kozuch S, Shaik S. 2010. Automatic analysis of computed catalytic cycles. J. Comput. Chem.
32:97885
26. Stegelmann C, Andreasen A, Campbell CT. 2009. Degree of rate control: how much the energies of
intermediates and transition states control rates. J. Am. Chem. Soc. 131:807782
27. Green WH. 2007. Predictive kinetics: a new approach for the 21st century. In Chemical Engineering
Kinetics, ed. GB Marin, pp. 150. Adv. Chem. Eng. 32. London: Academic
28. Vinu R, Broadbelt LJ. 2012. Unraveling reaction pathways and specifying reaction kinetics for complex
systems. Annu. Rev. Chem. Biomol. Eng. 3:2954
29. Campbell CT. 2004. Towards tomorrows catalysts. Nature 432:28283
30. Cejka J, Corma A, Zones S, eds. 2010. Zeolites and Catalysis. Weinheim, Ger.: Wiley-VCH (2 volumes)
31. Corma A. 1997. From microporous to mesoporous molecular sieve materials and their use in catalysis.
Chem. Rev. 97:2373420
32. Corma A. 2003. State of the art and future challenges of zeolites as catalysts. J. Catal. 216:298312
33. Cundy CS, Cox PA. 2003. The hydrothermal synthesis of zeolites: history and development from the
earliest days to the present time. Chem. Rev. 103:663701
590 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
34. Cejka J, Mintova S. 2007. Perspectives of micro/mesoporous composites in catalysis. Catal. Rev. Sci. Eng.
49:457509
35. Davis ME. 2002. Ordered porous materials for emerging application. Nature 417:81321
36. Weisz PB, Frilette VJ. 1960. Intracrystalline and molecular-shape-selective catalysis by zeolite salts.
J. Phys. Chem. 64:382
37. Weisz PB, Frilette VJ, Maatman RW, Mower EB. 1962. Catalysis by crystalline aluminosilicates II.
Molecular-shape selective reactions. J. Catal. 1:30712
38. Martens JA, Tielen M, Jacobs PA, Weitkamp J. 1984. Estimation of the void structure and pore dimen-
sions of molecular-sieve zeolites using the hydroconversion of normal-decane. Zeolites 4:98107
39. Weitkamp J, Ernst S, Kumar R. 1986. The spaciousness index: a novel test reaction for characterizing
the effective pore width of bifunctional zeolite catalysts. Appl. Catal. 27:20710
40. Deem MW, Pophale R, Cheeseman PA, Earl DJ. 2009. Computational discovery of new zeolite-like
materials. J. Phys. Chem. C 113:2135360
41. Pophale R, Cheeseman PA, Deem MW. 2011. A database of new zeolite-like materials. Phys. Chem.
Chem. Phys. 13:1240712
42. Treacy MMJ, Rivin I, Balkovsky E, Randall KH, Foster MD. 2004. Enumeration of periodic tetrahedral
frameworks. II. Polynodal graphs. Microporous Mesoporous Mater. 74:12132
43. Gounder R, Iglesia E. 2013. The catalytic diversity of zeolites: connement and solvation effects within
voids of molecular dimensions. Chem. Commun. 49:3491509
44. Niwa M, Katada N, Okumura K. 2010. Characterization and Design of Zeolite Catalysts. Berlin: Springer-
Verlag
45. Berger RJ, Stitt EH, Marin GB, Kapteijn F, Moulijn JA. 2001. Eurokin. Chemical reaction kinetics in
practice. CATTECH 5:3660
46. Mears DE. 1971. Diagnostic criteria for heat transport limitations in xed bed reactors. J. Catal. 20:127
31
47. Mears DE. 1971. Tests for transport limitations in experimental catalytic reactors. Ind. Eng. Chem. Proc.
Des. Dev. 10:54147
48. Berger RJ, Kapteijn F, Moulijn JA, Marin GB, De Wilde J, et al. 2008. Dynamic methods for catalytic
kinetics. Appl. Catal. A 342:328
49. Shannon SL, Goodwin JG. 1995. Characterization of catalytic surfaces by isotopic-transient kinetics
during steady-state reaction. Chem. Rev. 95:67795
50. Gleaves JT, Ebner JR, Kuechler TC. 1988. Temporal analysis of products (TAP)a unique catalyst
evaluation system with submillisecond time resolution. Catal. Rev. Sci. Eng. 30:49116
51. Gleaves JT, Yablonsky GS, Phanawadee P, Schuurman Y. 1997. TAP-2: an interrogative kinetics ap-
proach. Appl. Catal. A 160:5588
52. Yablonsky GS, Constales D, Gleaves JT. 2002. Multi-scale problems in the quantitative characterization
of complex catalytic materials. Syst. Anal. Model. Simul. 42:114366
53. Craciun I, Reyniers MF, Marin GB. 2007. Effects of acid properties of Y zeolites on the liquid-phase
alkylation of benzene with 1-octene: a reaction path analysis. J. Mol. Catal. A Chem. 277:114
54. Ranzi E, Dente M, Goldaniga A, Bozzano G, Faravelli T. 2001. Lumping procedures in detailed kinetic
modeling of gasication, pyrolysis, partial oxidation and combustion of hydrocarbon mixtures. Prog.
Energy Combust. Sci. 27:99139
55. Clymans PJ, Froment GF. 1984. Computer-generationof reactionpaths andrate equations inthe thermal
cracking of normal and branched parafns. Comput. Chem. Eng. 8:13742
56. Vandewiele NM, Van Geem KM, Reyniers MF, Marin GB. 2012. Genesys: kinetic model construction
using chemo-informatics. Chem. Eng. Sci. 207:52638
57. Van Borm R, Aerts A, Reyniers MF, Martens JA, Marin GB. 2010. Catalytic cracking of 2,2,4-
trimethylpentane on FAU, MFI, and bimodal porous materials: inuence of acid properties and pore
topology. Ind. Eng. Chem. Res. 49:681523
58. Van Borm R, Reyniers MF, Martens JA, Marin GB. 2010. Catalytic cracking of methylcyclohexane on
FAU, MFI, and bimodal porous materials: inuence of acid properties and pore topology. Ind. Eng.
Chem. Res. 49:1048695
www.annualreviews.org Heterogeneously Catalyzed Reactions 591
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
59. Nguyen CM, Reyniers MF, Marin GB. 2012. Reactions of bioalcohols in H-FAU, H-MOR, H-ZSM-5 and
H-ZSM-22. Presented at AIChE Annu. Meet., Pittsburgh, PA
60. Nguyen CM, Alexopoulos K, Reyniers MF, Marin GB. 2013. Ab initio based microkinetic modelling of
ethanol dehydration in zeolites. Presented at EuropaCat XI, Sept. 16, Lyon, France
61. Berna JL, Cavalli L, Renta C. 1995. A life-cycle inventory for the production of linear alkylbenzene
sulphonates in Europe. Tenside Surfactants Deterg. 32:12227
62. Corma A, Garcia H. 2003. Lewis acids: from conventional homogeneous to green homogeneous and
heterogeneous catalysis. Chem. Rev. 103:430766
63. Yadav GD, Siddiqui MINI. 2009. UDCaT-5: a novel mesoporous superacid catalyst in the selective
synthesis of linear phenyldodecanes by the alkylation of benzene with 1-dodecene. Ind. Eng. Chem. Res.
48:108039
64. Clark JH. 1999. Green chemistry: challenges and opportunities. Green Chem. 1:18
65. De Almeida JLG, Dufaux M, Ben Taarit Y, Naccache C. 1994. Effect of pore-size and aluminumcontent
on the production of linear alkylbenzenes over HY, H-ZSM-5 and H-ZSM-12 zeolites: alkylation of
benzene with 1-dodecene. Appl. Catal. A Gen. 114:14159
66. Cao Y, Kessas R, Naccache C, Ben Taarit Y. 1999. Alkylation of benzene with dodecene. The activity and
selectivity of zeolite type catalysts as a function of the porous structure. Appl. Catal. A Gen. 184:23138
67. Bhore NA, Klein MT, Bischoff KB. 1990. The delplot technique: a new method for reaction pathway
analysis. Ind. Eng. Chem. Res. 29:31316
68. Smirniotis PG, Ruckenstein E. 1995. Alkylation of benzene or toluene with MeOHor C
2
H
4
over ZSM-5
or zeolite: effect of the zeolite pore openings and of the hydrocarbons involved on the mechanism of
alkylation. Ind. Eng. Chem. Res. 34:151728
69. Namuangruk S, Pantu P, Limtrakul J. 2004. Alkylation of benzene with ethylene over faujasite zeolite
investigated by the ONIOM method. J. Catal. 225:52330
70. Rosenbach N Jr, dos Santos APA, Franco M, Mota CJA. 2010. The tertbutyl cation on zeolite Y: a
theoretical and experimental study. Chem. Phys. Lett. 485:12428
71. Boronat M, Corma A. 2008. Are carbenium and carbonium ions reaction intermediates in zeolite-
catalyzed reactions? Appl. Catal. A 336:210
72. Tuma C, Sauer J. 2005. Protonated isobutene in zeolites: Tert-butyl cation or alkoxide? Angew. Chem.
Int. Ed. 44:476971
73. Haw JF. 2002. Zeolite acid strength and reaction mechanisms in catalysis. Phys. Chem. Chem. Phys.
4:543141
74. Haouas M, Fink G, Taulelle F, Sommer J. 2010. Low-temperature alkane C-H bond activation by
zeolites: an in situ solid-state NMR H/D exchange study for a carbenium concerto. Chemistry 16:9034
39
75. van Santen RA, Neurock M, Shetty SG. 2010. Reactivity theory of transition-metal surfaces: a Brnsted-
Evans-Polanyi linear activation energy-free-energy analysis. Chem. Rev. 110:200548
76. Rozanska X, van Santen RA, Demuth T, Hutschka F, Hafner J. 2003. A periodic DFTstudy of isobutene
chemisorptioninproton-exchangedzeolites: dependence of reactivity onthe zeolite framework structure.
J. Phys. Chem. B 107:130915
77. Boronat M, Viruela PM, Corma A. 2004. Reaction intermediates in acid catalysis by zeolites: prediction
of the relative tendency to form alkoxides or carbocations as a function of hydrocarbon nature and active
site structure. J. Am. Chem. Soc. 126:33009
78. Nguyen CM, De Moor BA, Reyniers MF, Marin GB. 2012. Isobutene protonation in H-FAU, H-MOR,
H-ZSM-5, and H-ZSM-22. J. Phys. Chem. C 116:1823649
79. Macht J, Carr RT, Iglesia E. 2009. Consequences of acid strength for isomerization and elimination
catalysis on solid acids. J. Am. Chem. Soc. 131:655465
80. Carr RT, Neurock M, Iglesia E. 2011. Catalytic consequences of acid strength in the conversion of
methanol to dimethyl ether. J. Catal. 278:7893
81. Macht J, Carr RT, Iglesia E. 2009. Functional assessment of the strength of solid acid catalysts. J. Catal.
264:5466
82. De Moor BA, Reyniers MF, Sierka M, Sauer J, Marin GB. 2008. Physisorption and chemisorption of
hydrocarbons in H-FAU using QM-Pot(MP2//B3LYP) calculations. J. Phys. Chem. C 112:11796812
592 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
83. De Moor BA, Reyniers MF, Gobin OC, Lercher JA, Marin GB. 2011. Adsorption of C2-C8 n-alkanes
in zeolites. J. Phys. Chem. C 115:120419
84. Nguyen CM, Reyniers MF, Marin GB. 2010. Theoretical study of the adsorption of C1C4 primary
alcohols in H-ZSM-5. Phys. Chem. Chem. Phys. 12:948193
85. Nguyen CM, Reyniers MF, Marin GB. 2011. Theoretical study of the adsorption of the butanol isomers
in H-ZSM-5. J. Phys. Chem. C 115:865869
86. Nguyen CM, De Moor BA, Reyniers MF, Marin GB. 2011. Physisorption and chemisorption of lin-
ear alkenes in zeolites: a combined QM-Pot(MP2//B3LYP:GULP)-statistical thermodynamics study.
J. Phys. Chem. C 115:2383147
87. Chu Y, Han B, Fang H, Zheng A, Deng F. 2012. Inuence of acid strength on the reactivity of alkane
activation on solid acid catalysts: a theoretical calculation study. Microporous Mesoporous Mater. 151:241
49
88. Niwa M, Suzuki K, Morishita N, Sastre G, Okumura K, Katada N. 2013. Dependence of cracking
activity on the Brnsted acidity of Y zeolite: DFT study and experimental conrmation. Catal. Sci.
Technol. 8:191927
89. Almutairi SMT, Mezari B, Filonenko GA, Magusin PCMM, Rigutto MS, et al. 2013. Inuence of ex-
traframework aluminumonthe Brnsted acidity and catalytic reactivity of faujasite zeolite. ChemCatChem
5:45266
90. Feng W, Vynckier E, Froment GF. 1993. Single event kinetics of catalytic cracking. Ind. Eng. Chem. Res.
32:29973005
91. Vynckier E, Froment GF. 1991. Modeling of the kinetics of complex processes based upon elementary
steps. In Kinetic and Thermodynamic Lumping of Multicomponent Mixtures, Vol. 10, ed. G Astaritra, SI
Sandler, pp. 13161. Amsterdam: Elsevier Sci.
92. Beirnaert HC, Vermeulen R, Froment GF. 1994. Arecycle electrobalance reactor for the study of catalyst
deactivation by coke formation. Stud. Surf. Sci. Catal. 88:97112
93. Funke HH, Argo AM, Falconer JF, Noble RD. 1997. Separations of cyclic, branched, and linear hydro-
carbon mixtures through silicalite membranes. Ind. Eng. Chem. Res. 36:13743
94. Webster CE, Drago RS, Zerner MC. 1999. A method for characterizing effective pore sizes of catalysts.
J. Phys. Chem. B 103:124249
95. Costa C, Lopes JM, Lemos F, Ram oa Ribeiro F. 1999. Activityacidity relationship in zeolite Y. Part 3.
Application of Brnsted type equations. J. Mol. Catal. A Chem. 144:23338
96. Costa C, Dzikh IP, Lopes JM, Lemos F, Ram oa Ribeiro F. 2000. Activityacidity relationship in zeolite
ZSM-5. Application of Brnsted-type equations. J. Mol. Catal. A Chem. 154:193201
97. Lemos F, Lemos MANDA, Wang X, Ramos Pinto R, Borges P, et al. 2002. Analysis and modelling
of multi-site acid catalysts. In Principles and Methods for Accelerated Catalyst Design and Testing, ed. EG
Derouane, V Parmon, F Lemos, F Ram oa Ribeiro, pp. 21743. Dordrecht, Neth.: Kluwer Acad. Publ.
98. Borges P, Ramos Pinto R, Lemos MANDA, Lemos F, V edrine JC, et al. 2005. Activityacidity rela-
tionship for alkane cracking over zeolites: n-hexane cracking over HZSM-5. J. Mol. Catal. A Chem.
229:12735
99. Borges P, Ramos Pinto R, Oliveira P, Lemos MANDA, Lemos F, et al. 2009. Contributions for the
study of the acid transformation of hydrocarbons over zeolites. J. Mol. Catal. A Chem. 305:6068
100. Fonseca N, Lemos F, Laforge S, Magnoux P, Ram oa Ribeiro F. 2010. Inuence of acidity on the H-Y
zeolite performance in n-decane catalytic cracking: evidence of a series/parallel mechanism. React. Kinet.
Mech. Catal. 100:24963
101. Oliveira P, Borges P, Ramos Pinto R, Lemos MANDA, Lemos F, et al. 2010. Light olen transformation
over ZSM-5 zeolites with different acid strengthsa kinetic model. Appl. Catal. A 384:17785
102. Yaluris G, Madon RJ, Rudd DF, Dumesic JA. 1994. Catalytic cycles and selectivity of hydrocarbon
cracking on Y-zeolite-based catalysts. Ind. Eng. Chem. Res. 33:291323
103. Yaluris G, Rekoske JE, Aparicio LM, Madon RJ, Dumesic JA. 1995. Isobutane cracking over Y-zeolites.
I. Development of a kinetic model. J. Catal. 153:5464
104. Yaluris G, Madon RJ, Dumesic JA. 1999. Catalytic ramications of steam deactivation of Y zeolites: an
analysis using 2-methylhexane cracking. J. Catal. 186:13446
www.annualreviews.org Heterogeneously Catalyzed Reactions 593
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05CH26-Marin ARI 20 May 2014 19:23
105. Watson BA, Klein MT, Harding RH. 1996. Mechanistic modeling of n-heptane cracking on HZSM-5.
Ind. Eng. Chem. Res. 35:150616
106. Watson BA, Klein MT, Harding RH. 1997. Catalytic cracking of alkylbenzenes: modeling the reaction
pathways and mechanisms. Appl. Catal. A 160:1339
107. Watson BA, Klein MT, Harding RH. 1997. Mechanistic modeling of a 1-phenyloctane/n-hexadecane
mixture on rare earth Y zeolite. Ind. Eng. Chem. Res. 36:295463
108. Narasimhan CSL, Thybaut JW, Marin GB, Jacobs PA, Martens JA, et al. 2003. Kinetic modeling of
pore mouth catalysis in the hydroconversion of n-octane on Pt-H-ZSM-22. J. Catal. 220:399413
109. Van Borm R. 2012. Single-event microkinetics of hydrocarbon cracking on zeotype catalysts: effect of acidity and
shape selectivity. PhD thesis, Ghent Univ.
110. Bhan A, Gounder R, Macht J, Iglesia E. 2008. Entropy considerations in monomolecular cracking of
alkanes on acidic zeolites. J. Catal. 253:22124
111. Narbeshuber TF, Vinek H, Lercher JA. 1995. Monomolecular conversion of light alkanes over H-ZSM-
5. J. Catal. 157:38895
112. Babitz SM, Williams BA, Miller JT, Snurr RQ, Haag WO, Kung HH. 1999. Monomolecular cracking
of n-hexane on Y, MOR, and ZSM-5 zeolites. Appl. Catal. A Gen. 179:7186
113. Ramachandran CE, Williams BA, van Bokhoven JA, Miller JT. 2005. Observation of a compensation
relation for n-hexane adsorption in zeolites with different structures: implications for catalytic activity.
J. Catal. 233:1008
114. Xu B, Sievers C, Hong SB, Prins R, van Bokhoven JA. 2006. Catalytic activity of Brnsted acid sites in
zeolites: intrinsic activity, rate-limiting step, and inuence of the local structure of the acid sites. J. Catal.
244:16368
115. Gounder R, Iglesia E. 2009. Catalytic consequences of spatial constraints and acid site location for
monomolecular alkane activation on zeolites. J. Am. Chem. Soc. 131:195871
116. Tranca DC, Hansen N, Swisher JA, Smit B, Keil FJ. 2012. Combined density functional theory and
Monte Carlo analysis of monomolecular cracking of light alkanes over H-ZSM-5. J. Phys. Chem. C
116:2340817
117. Zimmerman PM, Tranca DC, Gomes J, Lambrecht DS, Head-Gordon M, Bell AT. 2012. Ab initio
simulations reveal that reaction dynamics strongly affect product selectivity for the cracking of alkanes
over H-MFI. J. Am. Chem. Soc. 134:1946876
118. Taarning E, Osmundsen CM, Yang X, Voss B, Andersen SI, Christensen CH. 2011. Zeolite-catalyzed
biomass conversion to fuels and chemicals. Energy Environ. Sci. 4:793804
119. Haro P, Ollero P, Trippe F. 2013. Technoeconomic assessment of potential processes for bio-ethylene
production. Fuel Process. Technol. 114:3548
120. Alvarenga RAF, Dewulf J. 2013. Plastic vs. fuel: Which use of the Brazilian ethanol can bring more
environmental gains? Renew. Energy 59:4952
121. Angelici C, Weckhuysen BM, Bruijnincx PCA. 2013. Chemocatalytic conversion of ethanol into buta-
diene and other bulk chemicals. ChemSusChem 6:121
122. Tretyakov VF, Makara YI, Tretyakov KV, Frantsuzovaa NA, Talyshinskiia RM. 2010. The catalytic
conversion of bioethanol to hydrocarbon fuel: a review and study. Catal. Ind. 2:40220
123. Chiang H, Bhan A. 2010. Catalytic consequences of hydroxyl group location on the rate and mechanism
of parallel dehydration reactions of ethanol over acidic zeolites. J. Catal. 271:25161
124. Zhang M, Yu Y. 2013. Dehydration of ethanol to ethylene. Ind. Eng. Chem. Res. 52:950514
125. Johansson R, Hruby SL, Rass-Hansen J, Christensen CH. 2009. The hydrocarbon pool in ethanol-to-
gasoline over HZSM-5 catalyst. Catal. Lett. 127:16
126. Blaszkowski SR, van Santen RA. 1997. Theoretical study of CC bond formation in the methanol-to-
gasoline process. J. Am. Chem. Soc. 119:502027
127. De Moor BA, Ghysels A, Reyniers MF, Van Speybroeck V, Waroquier M, Marin GB. 2011. Normal
mode analysis in zeolites: toward an efcient calculation of adsorption entropies. J. Chem. Theory Comput.
7:1090101
594 Reyniers

Marin
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05-FrontMatter ARI 5 May 2014 12:43
Annual Review of
Chemical and
Biomolecular
Engineering
Volume 5, 2014 Contents
Plans and Detours
James Wei p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Simulating the Flow of Entangled Polymers
Yuichi Masubuchi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 11
Modeling Chemoresponsive Polymer Gels
Olga Kuksenok, Debabrata Deb, Pratyush Dayal, and Anna C. Balazs p p p p p p p p p p p p p p p p p p p 35
Atmospheric Emissions and Air Quality Impacts from Natural Gas
Production and Use
David T. Allen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 55
Manipulating Crystallization with Molecular Additives
Alexander G. Shtukenberg, Stephanie S. Lee, Bart Kahr,
and Michael D. Ward p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 77
Advances in Mixed-Integer Programming Methods for Chemical
Production Scheduling
Sara Velez and Christos T. Maravelias p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 97
Population Balance Modeling: Current Status and Future Prospects
Doraiswami Ramkrishna and Meenesh R. Singh p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 123
Energy Supply Chain Optimization of Hybrid Feedstock Processes:
A Review
Josephine A. Elia and Christodoulos A. Floudas p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 147
Dynamics of Colloidal Glasses and Gels
Yogesh M. Joshi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 181
Rheology of Non-Brownian Suspensions
Morton M. Denn and Jeffrey F. Morris p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 203
Factors Affecting the Rheology and Processability of Highly
Filled Suspensions
Dilhan M. Kalyon and Seda Akta s p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 229
vi
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05-FrontMatter ARI 5 May 2014 12:43
Continuous-Flow Differential Mobility Analysis of Nanoparticles
and Biomolecules
Richard C. Flagan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 255
From Stealthy Polymersomes and Filomicelles to Self
Peptide-Nanoparticles for Cancer Therapy
N uria Sancho Oltra, Praful Nair, and Dennis E. Discher p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 281
Carbon Capture Simulation Initiative: A Case Study in Multiscale
Modeling and New Challenges
David C. Miller, Madhava Syamlal, David S. Mebane, Curt Storlie,
Debangsu Bhattacharyya, Nikolaos V. Sahinidis, Deb Agarwal, Charles Tong,
Stephen E. Zitney, Avik Sarkar, Xin Sun, Sankaran Sundaresan, Emily Ryan,
Dave Engel, and Crystal Dale p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 301
Downhole Fluid Analysis and Asphaltene Science for Petroleum
Reservoir Evaluation
Oliver C. Mullins, Andrew E. Pomerantz, Julian Y. Zuo, and Chengli Dong p p p p p p p p p p 325
Biocatalysts for Natural Product Biosynthesis
Nidhi Tibrewal and Yi Tang p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 347
Entangled Polymer Dynamics in Equilibrium and Flow Modeled
Through Slip Links
Jay D. Schieber and Marat Andreev p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 367
Progress and Challenges in Control of Chemical Processes
Jay H. Lee and Jong Min Lee p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 383
Force-Field Parameters from the SAFT- Equation of State for Use in
Coarse-Grained Molecular Simulations
Erich A. M uller and George Jackson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 405
Electrochemical Energy Engineering: A New Frontier of Chemical
Engineering Innovation
Shuang Gu, Bingjun Xu, and Yushan Yan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 429
A New Toolbox for Assessing Single Cells
Konstantinos Tsioris, Alexis J. Torres, Thomas B. Douce,
and J. Christopher Love p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 455
Advancing Adsorption and Membrane Separation Processes for the
Gigaton Carbon Capture Challenge
Jennifer Wilcox, Reza Haghpanah, Erik C. Rupp, Jiajun He, and Kyoungjin Lee p p p p p 479
Toward the Directed Self-Assembly of Engineered Tissues
Victor D. Varner and Celeste M. Nelson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 507
Ionic Liquids in Pharmaceutical Applications
I.M. Marrucho, L.C. Branco, and L.P.N. Rebelo p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 527
Contents vii
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
CH05-FrontMatter ARI 5 May 2014 12:43
Perspectives on Sustainable Waste Management
Marco J. Castaldi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 547
Experimental and Theoretical Methods in Kinetic Studies of
Heterogeneously Catalyzed Reactions
Marie-Fran coise Reyniers and Guy B. Marin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 563
Indexes
Cumulative Index of Contributing Authors, Volumes 15 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 595
Cumulative Index of Article Titles, Volumes 15 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 598
Errata
An online log of corrections to Annual Review of Chemical and Biomolecular Engineering
articles may be found at http://www.annualreviews.org/errata/chembioeng
viii Contents
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.
ANNUAL REVIEWS
Its about time. Your time. Its time well spent.
ANNUAL REVIEWS | Connect With Our Experts
Tel: 800.523.8635 (US/CAN) | Tel: 650.493.4400 | Fax: 650.424.0910 | Email: service@annualreviews.org
New From Annual Reviews:
Annual Review of Statistics and Its Application
Volume 1 Online January 2014 http://statistics.annualreviews.org
Editor: Stephen E. Fienberg, Carnegie Mellon University
Associate Editors: Nancy Reid, University of Toronto
Stephen M. Stigler, University of Chicago
The Annual Review of Statistics and Its Application aims to inform statisticians and quantitative methodologists, as
well as all scientists and users of statistics about major methodological advances and the computational tools that
allow for their implementation. It will include developments in the feld of statistics, including theoretical statistical
underpinnings of new methodology, as well as developments in specifc application domains such as biostatistics
and bioinformatics, economics, machine learning, psychology, sociology, and aspects of the physical sciences.
Complimentary online access to the frst volume will be available until January 2015.
TABLE OF CONTENTS:
What Is Statistics? Stephen E. Fienberg
A Systematic Statistical Approach to Evaluating Evidence
from Observational Studies, David Madigan, Paul E. Stang,
Jesse A. Berlin, Martijn Schuemie, J. Marc Overhage,
Marc A. Suchard, Bill Dumouchel, Abraham G. Hartzema,
Patrick B. Ryan
The Role of Statistics in the Discovery of a Higgs Boson,
David A. van Dyk
Brain Imaging Analysis, F. DuBois Bowman
Statistics and Climate, Peter Guttorp
Climate Simulators and Climate Projections,
Jonathan Rougier, Michael Goldstein
Probabilistic Forecasting, Tilmann Gneiting,
Matthias Katzfuss
Bayesian Computational Tools, Christian P. Robert
Bayesian Computation Via Markov Chain Monte Carlo,
Radu V. Craiu, Jefrey S. Rosenthal
Build, Compute, Critique, Repeat: Data Analysis with Latent
Variable Models, David M. Blei
Structured Regularizers for High-Dimensional Problems:
Statistical and Computational Issues, Martin J. Wainwright
High-Dimensional Statistics with a View Toward Applications
in Biology, Peter Bhlmann, Markus Kalisch, Lukas Meier
Next-Generation Statistical Genetics: Modeling, Penalization,
and Optimization in High-Dimensional Data, Kenneth Lange,
Jeanette C. Papp, Janet S. Sinsheimer, Eric M. Sobel
Breaking Bad: Two Decades of Life-Course Data Analysis
in Criminology, Developmental Psychology, and Beyond,
Elena A. Erosheva, Ross L. Matsueda, Donatello Telesca
Event History Analysis, Niels Keiding
Statistical Evaluation of Forensic DNA Profle Evidence,
Christopher D. Steele, David J. Balding
Using League Table Rankings in Public Policy Formation:
Statistical Issues, Harvey Goldstein
Statistical Ecology, Ruth King
Estimating the Number of Species in Microbial Diversity
Studies, John Bunge, Amy Willis, Fiona Walsh
Dynamic Treatment Regimes, Bibhas Chakraborty,
Susan A. Murphy
Statistics and Related Topics in Single-Molecule Biophysics,
Hong Qian, S.C. Kou
Statistics and Quantitative Risk Management for Banking
and Insurance, Paul Embrechts, Marius Hofert
Access this and all other Annual Reviews journals via your institution at www.annualreviews.org.
A
n
n
u
.

R
e
v
.

C
h
e
m
.

B
i
o
m
o
l
.

E
n
g
.

2
0
1
4
.
5
:
5
6
3
-
5
9
4
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

U
n
i
v
e
r
s
i
t
y

o
f

O
k
l
a
h
o
m
a

-

N
o
r
m
a
n

o
n

0
7
/
1
1
/
1
4
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

Você também pode gostar