Sie sind auf Seite 1von 17

Article

pubs.acs.org/IECR

Single-Event Microkinetics for Methanol to Olens on HZSM5


Pravesh Kumar, Joris W. Thybaut,*, Stian Svelle, Unni Olsbye, and Guy B. Marin

Laboratory for Chemical Technology, Ghent University, Krijgslaan 281S5, Ghent B-9000, Belgium
inGap Center of Research Based innovation, Department of Chemistry, University of Oslo, P.O. Box 1033, Blindern, 0315 Oslo,
Norway

ABSTRACT: A single-event microkinetic (SEMK) model was developed for the conversion of methanol to olens (MTO) and
used in the assessment of experimental data obtained on H-ZSM-5 with a Si/Al ratio of 200. The experiments were performed at
temperatures from 643 to 753 K, space times between 0.5 and 6.5 kgcats mol1 and at atmospheric pressure. Dimethyl ether
(DME) and primary olens formation through aromatic hydrocarbon pool and higher olens formation via the alkene
homologation cycle, was implemented in terms of elementary steps. The single-event concept, in combination with
thermodynamic constraints allowed a signicant reduction of the number of adjustable parameters. A further reduction was
achieved by calculation of the single-event pre-exponential factors based on statistical thermodynamics. Physicochemical
constraints along with Boudarts criteria were used to limit the parameter space. Twenty one activation energies of kinetically
signicant reaction families and eight protonation enthalpies corresponding to methanol, DME and olens were estimated via
regression to the experimental data. The SEMK model well describes the product distribution, relying on model parameters with
a precise physical meaning. The trends in activation energies obtained, are as could be expected from the considered reaction
family, and the type of carbenium ions involved as reactant and product. Olen protonation enthalpies decrease from 11.2 kJ/
mol for ethene to 70.3 kJ/mol for heptene. A reaction path analysis established that ethene originates exclusively from the
aromatic hydrocarbon pool, while propene is formed both via the aromatic hydrocarbon pool and the alkene homologation cycle.

1. INTRODUCTION
The ever increasing cost of crude oil and depleting oil reserves
force us to look for alternatives to meet the demand for
chemicals and fuels. The zeolite-catalyzed methanol to olens
(MTO) process constitutes one potentially viable, future route
in the transformation of natural gas, coal, or biomass into
synthetic fuels and chemicals.1 Established processes for natural
gas transformation into synthetic fuels require large investments, which are prohibitive for the exploitation of
approximately one-third of the worlds natural gas reserves
that are considered as stranded.2 In addition, large natural gas
reserves found in remote areas have further enhanced the
research in this area.3
The mechanism of the MTO reaction and the rst CC
bond formation is a matter of debate over the years.3 The direct
mechanisms consider the conversion of methanol or DME into
a species with a rst CC bond. Surface methoxy species
decompose, yielding an intermediate of a carbene or ylide
nature that is susceptible to CC bond formation via
methylation. Subsequently, higher aliphatics, such as propane
and isobutane, and aromatics, such as polymethylbenzene can
be formed.4 However, theoretical calculations result in very
high energy barriers for such reactions.5,6 More recently,
evidence has emerged710 that the MTO reaction proceeds
through a so-called aromatic hydrocarbon pool mechanism in
which polymethylbenzenes acts as active centers for primary
olens formation.
An aromatic hydrocarbon pool-type mechanism for MTO
was rst proposed when Mole et al.11 observed cocatalytic
eects when cofeeding toluene with methanol on H-ZSM-5.
Langer et al.12 reported an 18-fold reduction in the induction
period when the feed contained 0.0035% cyclohexanol
2012 American Chemical Society

providing further evidence for an aromatic hydrocarbon pooltype mechanism. Dahl et al.13 proposed that methanol reacted
on a hydrocarbon pool of intermediates to form olens;
however, the type of hydrocarbon pool species were not
identied. Mikkelsen et al.14 observed isotopic scrambling in
olenic products when cofeeding toluene and 13C methanol
over H-ZSM-5, also suggesting the involvement of aromatic
hydrocarbon pool type species.
Kinetic modeling is an indispensable tool in the design and
optimization of industrial processes. Kinetic models should
represent the molecular interactions that occur when chemical
bonds are broken and restructured resulting in new chemical
compounds.15 The traditional approach when dealing with
complex mixtures is to dene lumped kinetic models and
obtain global kinetics. Even though the corresponding rate
equations may be useful in reactor design, they have a very
limited use in catalyst design. They only reect the global
reactions and, hence, barely account for the detailed chemistry.
As a result, the corresponding rate and equilibrium coecients
typically depend on reaction conditions and feed composition.
On the other hand, in fundamental, microkinetic models, the
reaction mechanism is considered in terms of elementary steps,
which describe the reactions taking place at the catalyst surface
at a molecular level. The kinetic and catalyst descriptors
determined based on such models have a clear physical
meaning, which can be exploited in the design of improved and
innovative catalysts by relating the catalyst descriptors to the
Received:
Revised:
Accepted:
Published:
1491

June 13, 2012


December 12, 2012
December 28, 2012
December 28, 2012
dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

composition and synthesis procedure.16,17 The complexity of


the product pattern in the MTO reaction forced many
researchers to lump reactants as well as products.1826
Nevertheless, fundamental kinetic modeling eorts2729 have
also been reported for MTO. Chen et al.18 proposed a kinetic
model over ZSM-5 by considering the autocatalytic reaction
between the oxygenates and the produced olens. The
reactants were lumped as oxygenates and the products as
olens and aromatics/parans. Chang et al.19 added a
bimolecular step accounting for carbene insertion into primary
olens in the scheme of Chen et al.18 and proposed a four lump
kinetic model. The seven lump model of Schoenfelder et al.20
on H-ZSM-5 involving oxygenate ethene, propene, butene,
parans, methane, carbon monoxide, hydrogen, and water
explicitly takes into account ethene formation from dehydration
of methanol and higher olens by reaction between lower
olens and methanol. Bos et al.21 developed a model consisting
of 12 reactions involving 6 lumps and coke. The reaction
mechanism of Bos et al.21 was further modied, and a kinetic
model on SAPO-34 was proposed, which takes into account the
individual formation of ethene, propene, butene, and a lump for
parans.22 More recently, a uidized bed reactor model for a
ZSM-5 catalyst has been published in which the active species
for product formation was represented as a hydrocarbon of high
molecular weight.25 Mihail et al.26 proposed a kinetic model in
which dimethyl ether (DME) generates a carbene species
which reacts with oxygenates to form the rst CC bond, that
is, a light olen, that subsequently reacts with other light olens
to form higher olens. The higher olens are assumed to
produce aromatics and parans via carbenium ion intermediates. As indicated above, due to the product lumping in the
mentioned models, the resulting parameters are only valid for a
specic feed composition and operating conditions.
Among the fundamental models, Park and Froment27,28
modeled the MTO kinetics on an H-ZSM-5 catalyst using a
detailed mechanistic reaction scheme. The surface oxonium
ylide mechanism was found to be most suitable for describing
primary product formation, and formation of higher olens was
proposed to occur via carbenium ion chemistry. Alwahabi and
Froment29 implemented this reaction mechanism on SAPO-34
and also included catalyst deactivation. However, none of the
above-discussed models considers primary product formation
via the aromatic hydrocarbon pool mechanism, which is, at
present, the most accepted mechanism for the formation of the
rst CC bond MTO.710,30 These fundamental models
account for all elementary steps via the concept of reaction
families. As a result, a maximum number of elementary steps in
a complex reaction mechanism can be described via a minimum
number of parameters. The model parameters have a precise
physical meaning, and the values that are obtained should be
physically sound and are independent of feed composition and
operating conditions.
In the present study, for the rst time, a SEMK model for
MTO has been developed using the aromatic hydrocarbon
pool-type mechanism on a H-ZSM-5 catalyst. The reaction
mechanism for DME, primary olens, and higher olens
formation was implemented in terms of elementary steps. The
single-event concept allowed us to dene unique rate
coecients for the considered reaction families of elementary
steps. Pre-exponential factors were calculated using statistical
thermodynamics, while thermodynamic consistency is guaranteed by applying the principle of microscopic reversibility.
Activation energies and protonation enthalpies were obtained

by regression with experimental data. Physicochemical


constraints along with Boudarts criteria were implemented to
limit the parameter space.

2. PROCEDURES
2.1. Operating Conditions, Catalyst, and Products.
Methanol-to-hydrocarbon conversion, the experimental data of
which were used in the present work, was carried out in a
tubular, xed-bed reactor on an H-ZSM-5 zeolite with a Si/Al
ratio of 200. The experiments were performed in a temperature
range from 360 to 480 C using space times between 0.5 and
6.5 kgcats mol1 at atmospheric pressure.28 This space time
corresponds to a weightly hourly space velocity (WHSV) of
0.00490.064 kgMeOHkgcat1s 1. The reactor tube has a length
of 0.27 m with an internal diameter of 21.4 mm. The catalyst
powder was pelletized by pressing it into wafers under a
pressure of 36.8 MPa. The wafers were subsequently crushed
and screened to have particle diameters in the range from 0.5 to
1.0 mm. The surface area was on the order of 400 m2/g. The
reactor was loaded with 1 g of catalyst and the catalyst bed was
diluted with inert oxide beads in a volumetric ratio of 5/1 (5
vol. of beads inert/1 vol. of catalyst). Additionally, inert beads
were placed on top of catalyst bed to serve as a preheating zone.
The catalyst bed itself was supported by a stainless steel screen.
A thin layer of glass wool and inert beads were placed between
the catalyst bed and the screen to avoid any inuence of the
screens on the ow pattern through the catalyst bed. Heat and
mass transfer limitations have been veried by calculating the
usual criteria31 and were found to be insignicant, mainly
because of the limited particle size of the catalyst and the
catalyst bed dilution. The temperature inside the reactor was
measured at three dierent heights along the reactor axis by a
thermocouple that slides inside a well with an external diameter
of 2.5 mm. To ensure a good temperature control, the reactor
was immersed in a molten salt bath during reaction. For catalyst
conditioning, however, it was lifted by an oil system into an
infrared-heated furnace. Aromatics and parans yields started
to be non-negligible at 70% conversion of methanol and always
remained below 20 mol % even at complete methanol
conversion. Below 70% of methanol conversion the olens
yield rapidly increases with the space time, the main product
being propene, while the ethene yield levels o at higher space
times. Methane was also observed as a product, albeit only in
minor amounts, while no formaldehyde was observed because
of its further decomposition.25 To be able to focus on the
implementation of the aromatic hydrocarbon pool mechanism
only the data with a methanol conversion below 70% have been
used in the regression.
2.2. Parameter Estimation. Parameter estimation was
performed by minimization of the following objective function,
which is dened on the basis of the weighted squared residuals
between the experimentally observed and model calculated
outlet ow rates for the dierent responses:
nresp nresp

SSQ =

nob

jk (Fij Fij )(Fik Fik )minimum


j=1 k=1

i=1

(1)

There are 10 responses related to the observable products:


methanol, water, DME, ethene, propene, butene, pentene,
hexene, heptene, and methane. The calculation of the molar
outlet ow rates follows from the reactor model described in
section 2.3 combined with the rate equations discussed in
1492

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

Table 1. Rate and Equilibrium Coecients in DME and Methane Formation:


DME and methane formation
elementary step

rate or equilibrium coecients

+ KPr(MeOH)

(i)

MeOH + H MeOH+2

(ii)

MeOH+2
R1+ + H 2O
+

(iii)

R1+ + MeOH
DMO+
+

(iv)

DMO+ DME + H+

(v)a

R1+ + MeOH CH4 + HCHO + H+

KPr(MeOH)

kF(R1+)

kF(R1+), k C(R1+)

k C(R1 )

kF(DMO+)

kF(DMO+), k C(DMO+)

k C(DMO )

{KPr(DME)}1

{KPr(DME)}1

kF(CH 4)

kF(CH4)

HCHO further decomposes in C and H2O25.

Figure 1. Formation of primary olens from the aromatic hydrocarbon pool mechanism-exocyclic methylation route7,10.

ingly, the continuity equation for a gas phase component j over


an innitesimal catalyst mass element can be written as

section 4.3. A combination of a Rosenbrock and a Levenberg


Marquardt algorithm32 was used to minimize this objective
function. An in-house developed code for the Rosenbrock
algorithm33 was used to nd an adequate direction leading to a
possible global optimum. For the LevenbergMarquardt
method, the ordinary least-squares option of ODRPACK
version 2.01 from Netlib34 was used. Some additional code
was added to ODRPACK to retrieve statistical information.
The weighting factors correspond to the diagonal elements of
the inverse of the covariance matrix of the experimental errors
of the responses as determined from replicate experiments.
When no replicate experiments were available, weighing factors
were calculated from the following approximate formula:

dFj
dW

Fj(0) = Fj 0

nresp

when

W=0

(4)

A mass balance, resulting in a dierential equation, is


constructed for all bulk phase species considered, that is, 50
olens as well as DME and methane. The outlet ow rate of
methanol and water is calculated based on the carbon and
hydrogen balance, respectively. The continuity equations are
numerically integrated up to the catalyst mass W for each
experiment.
These dierential equations are complemented by rate
equations for the net rate of formation of surface species
involved in the aromatic hydrocarbon pool mechanism and the
surface methoxy species resulting in a set of non-linear algebraic

(i =ob1 Fij)1
n

j = 1 (i =ob1 Fij)1

(3)

with, as initial condition,

jj =

= Rj

(2)

2.3. Reactor Model. A pseudohomogeneous, 1-dimensional, ideal plug ow reactor model is used to simulate the
experimental data acquired on the H-ZSM5 catalyst. Accord1493

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

bound surface methoxy or a more weakly adsorbed methanol


molecule. On the basis of the experimental evidence presented
in the literature,4043 surface methoxy species are considered in
the methylation reactions of aromatics and alkenes accounted
for in the SEMK model. The formation of higher olens also
takes place via alkylation reactions in the alkene homologation
cycle. These higher olens can recrack to lower olens.
Reaction of the surface methoxy species with methanol results
in a minor amount of methane formation as well. The
elementary steps involved in the reaction mechanism are
further explained in detail in the next sections.
3.1. Formation of Dimethyl Ether and Methane.
Methanol is converted into an equilibrium mixture of methanol,
water and DME at typical MTO operating conditions,
including a surface methoxy species which further participates
in olen formation as the central intermediate. Magic-anglespinning (MAS) and nuclear-magnetic-resonance (NMR)
studies have shown that methanol is reversibly adsorbed
through its alcoholic hydrogen to the bridging hydroxyl on the
Bronsted sites of the zeolite, step (i) in Table 1. The
protonation is very fast and can be considered in quasiequilibrium.44 Quantum chemical calculations using density
functional theory revealed that the formation of an adsorbed
complex out of methanol or DME does not exhibit an
appreciable activation barrier.45 The methoxonium ion thus
formed is subject to dehydration to form a surface methoxy
species, step (ii) in Table 1, which is covalently bonded to the
lattice oxygen of the zeolite. Experimental studies have
indicated that a chemical equilibrium exists in the formation
of a surface methoxy species from methanol or DME.4 The
surface methoxy species again reacts with methanol to form a
dimethyl oxonium ion, step (iii) in Table 1. Deprotonation of
the dimethyl oxonium ion yields DME, step (iv) in Table 1.
The formation of methane in the MTO process may also take
place via this surface methoxy species, step (v) in Table 1
reacting with methanol. Under typical MTO reaction
conditions methane can be considered as not reactive and,
hence, its formation is considered irreversible.
3.2. Formation of Primary Olens. In primary olen
formation through the aromatic hydrocarbon pool mechanism,
polymethylbenzenes, trapped inside the zeolite pores are
considered to act as active centers for the formation of the
light olens. These polymethylbenzenes are believed to
originate either from impurities in the methanol or the carrier
gas fed, from incomplete combustion of organics during
calcination or from surface methoxy species during an
induction period.46,47 Considering the dierent views on the
true origin of these aromatic hydrocarbon pool species, the
formation of these species is not explicitly described in the
present study, but they are considered as an inherent catalyst
property similar to the active sites. The amount of aromatic
hydrocarbon pool species depends on the chosen catalyst and
the operating conditions. In the SEMK model, however, a
single average value has been considered for this total
concentration and determined by regression. The role of
polymethylbenzenes as a major aromatic hydrocarbon pool
species appears to be independent of the zeotype catalyst
chosen. On the contrary, the exact nature of the aromatic
hydrocarbon pool species is found to depend on the catalyst
type and the reaction conditions.10
As a part of the aromatic hydrocarbon pool mechanism, two
alternatives have been reported for grouping several methyl
substituents into a single, larger substituent that is susceptible

equations expressing the pseudosteady state approximation


applied to all surface species:
R k(T , pj , C hp) = 0

(5)

where Rk is the net production rate of surface species. These


nonlinear algebraic equations are solved simultaneously with
continuity equations for gas phase components (eq 3).
The partial pressure of a component j in the gas phase is
calculated as
pj =

Fj
nresp pt
I = 1 Fj

(6)

The set of dierential algebraic equations is numerically


integrated using the subroutine DASPK also available from
the Netlib software library.34 DASPK uses variable step size
backward dierentiation formulas applying either direct linear
methods or a preconditioned Krylov iterative method. In the
present work, the direct method was used. When using DASPK,
the integration must start with a consistent set of initial
conditions. The initial concentration of the gas phase
components is known, vide eq 4. However, for the initial
concentration of the surface species associated with algebraic
equations, the numerical subroutine DNSQE,34 is used to solve
a set of nonlinear algebraic equations by implementation of a
hybrid Powell method.

3. REACTION SCHEME
The MTO process consists of three major groups of elementary
steps, that is, DME, primary olens, and higher olens
formation. Table 1 and Figure 1 give the elementary steps
and corresponding rate coecients in the DME and primary
olens formation, while Figure 2 gives an overview of the

Figure 2. Elementary steps in the formation of higher olens.

elementary steps considered in the formation and interconversion of higher hydrocarbons. Methanol is converted to a surface
methoxy species which is subsequently converted to DME. The
formation of this surface methoxy species is also observed when
using DME as feed instead of MeOH.35 The surface methoxy
species further participates in the methylation reactions in the
aromatic hydrocarbon pool and alkene homologation cycle to
form primary and higher olens, respectively.36 It may be noted
that there are dierent views3739 about the intermediate
involved in methylation reactions, that is, a strongly chemically
1494

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

to dealkylation, i.e., the paring mechanism48 and the exocyclic


methylation route.11 In both mechanisms, several methylations
of the aromatic ring and (de)protonation steps take place
followed by dealkylation to split o olens. In the paring
mechanism, methylation followed by rapid ring contraction and
expansion takes place, to extend an alkyl side chain, and
subsequent olen elimination. It typically results in the
formation of propene and, mainly, isobutene and the
interchange of carbon atoms between the ring and its
substituents. On the other hand, in the exocyclic methylation
route proposed by Mole et al.11 and later rened by Haw et al.,7
polymethylbenzenes are involved in several consecutive
methylation and (de)protonation steps as shown in Figure 1.
The ring structure remains intact and, hence, no interchange of
carbon atoms occurs between the ring and the substituents
which are split o as ethene and propene.
There is still an ongoing debate10,4951 about the preferred
reaction route, however. Since ethene is the major primary
product in the data used in the present work, the exocyclic
methylation route is used to describe primary olen formation.
Moreover, the data set used does not allow the determination
of whether carbon atom exchange between the ring and the
substituents has occurred. In this route, see Figure 1, the lowermethylbenzenes, represented as p-xylene, undergo methylation
with a surface methoxy to form the 1,1,4-trimethylbenzenium
cation (TMeB+), step (i) in Figure 1 in which two geminal
methyl groups are attached to the aromatic ring.52 The latter
subsequently undergoes deprotonation, losing a proton from
the isolated methyl group forming 3,3-dimethyl 6-methylene
1,4-cyclohexadiene (DMeMCHDE), step (ii) in Figure 1. The
exocyclic double bond is then methylated with a surface
methoxy, resulting in an ethyl substituent on the aromatic ring,
that is, 1,1-dimethyl-4-ethylbenzenium cation (DMeEtB+), step
(iii) in Figure 1, which subsequently can dealkylate to produce
ethene and yield the 1,1-dimethylbenzenium cation (DMeB+),
step (iv) in Figure 1.
The ethyl group on the aromatic ring is also susceptible to
deprotonation forming 3,3-dimethyl 6-ethylene 1,4 cyclohexadiene (DMeEtCHDE), step (v) in Figure 1, which can
subsequently be methylated to form a propyl substituent, 1isopropyl-4,4-dimethylbenzenium cation (PDMeB+), step (vi)
in Figure 1. After dealkylation the latter yields propene and 1,1dimethylbenzenium cation (DMeB+), step (vii) in Figure 1.
Hence, propene formation is possible from this reaction
scheme, while heavier olens formation is believed not to
occur via the aromatic hydrocarbon pool mechanism, for
example, due to steric hindrance. Moreover, the alkene
homologation cycle, as occurring in the formation of heavier
alkenes, cfr. infra, is much more kinetically favored from butene
onward because of the more stable types of carbenium ions
involved. After dealkylation, DMeB+ undergoes a methyl shift
to form PX+, the deprotonation of which is required to
regenerate the original species and close the catalytic cycle, step
(viii) in Figure 1.
3.3. Formation of Higher Hydrocarbons. The formation
of higher hydrocarbons2729,53 is interpreted in terms of
carbenium ion chemistry, in which gas-phase olens interact
with the Bronsted acid sites of catalyst resulting in (de)protonation reactions. Whether carbenium ions are reaction
intermediates or highly unstable transition states has been a
matter of debate in the past decade and beyond. Quantum
chemical calculations on the interaction of short alkyl chains
and aluminosilicate clusters tend to indicate that covalently

bonded alkoxy species are more stable,54,55 especially for short


chain hydrocarbons while for longer chain hydrocarbons
carbenium ions are more favored. In the present SEMK
model, C1 surface intermediates are represented as covalently
bound to the zeolite lattice, while C2 surface intermediates are
interpreted in terms of carbenium ions. The latter carbenium
ions are subject to rearrangements and alkylation/cracking
steps as shown in Figure 2.
Methylation and alkylation elementary steps increase the
carbon number of these carbenium ions while -scission can
recrack them again into lighter ones. Hydride shift, methyl shift,
and protonated cyclopropane (PCP) branching lead to isomers
and, hence, leave the carbon number unaected. Under typical
MTO reaction conditions, thermodynamic equilibrium is
generally established among olen isomers.56 The carbenium
ions stability, which depends on type and carbon number, is
explicitly considered in the model. Primary carbenium ions, that
are inherently unstable in nature are nevertheless retained as
product carbenium ions when they are potentially more stable
than the reactant carbenium ions because of their higher carbon
number and when no alternative pathway involving a secondary
or tertiary carbenium ion exists; for example, in ethene
methylation a primary propyl carbenium ion is allowed.
Propene methylation to a butyl primary carbenium ion is
excluded from the reaction network because an alternative
pathway involving secondary/tertiary carbenium ions for the
formation of C4 carbenium ions from C3 carbenium ions is
possible. It results in three types of elementary steps in
methylation reactions based on carbenium ion stability;
primaryprimary, primarysecondary, and primarytertiary
types of methylation reactions are considered.
In alkylation reactions, based on the stability of the
carbenium ion, three types of elementary steps, that is,
secondarysecondary, secondarytertiary, and tertiarysecondary, are involved, except for ethene alkylation, to form butyl
primary carbenium ions, which constitutes a primaryprimary
type of reaction. Only the formation of secondary and tertiary
carbenium ions in cracking is retained. As in the case of primary
carbenium ion formation during cracking the more stable
carbenium ion would be converted into a less stable ion.27,57
Since only species with a carbon number up to 7 are accounted
for in the reaction network, no alkylation/-scission between
tertiary carbenium ions (t, t) occur in the considered network.

4. SINGLE-EVENT MICROKINETIC MODEL


4.1. Reaction Network. The MTO process comprises a
vast number of elementary steps involved in DME formation, in
primary olens formation via the aromatic hydrocarbon pool
mechanism, and in higher olens formation via the alkene
homologation route comprising typical acid catalyzed reactions.
The network for such a complex process in terms of elementary
steps was generated using an in-house reaction network
generation program (ReNG). 58 The total number of
elementary steps and species involved, extending up to carbon
number 7, is given in Table 2. The representation of the
hydrocarbon molecules in ReNG is based on their carbon
skeleton, accounting for the bond type and potential
heteroatoms. Given the number of carbon atoms out of
which the aromatic hydrocarbon pool species exist and the
limited modications undergone by these species in the light
olen formation, see Figure 1, a way of representing these
species has been developed requiring a single atom only. The
aromatic hydrocarbon pool species are, hence, treated as an
1495

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

Table 2. Species and Number of Elementary Steps Involved


in the Reaction Network Considering Olens up to Carbon
Number 7
DME
formation
(cyclic)olens/
DME/H2O
carbenium ions
aromatics
total
protonation
deprotonation
hydride shift
methyl shift
PCP branching
methylation
demethylation
alkylation
Dealkylation
-scission
hydration
dehydration
total

primary olens
formation

higher olens
formation

Number of Species
3
4

50

41

5
1
6
10
Number of Elementary Steps
3
3
3
3

3
3
2
2

The number of single events, that is, the ratio between the
global symmetry numbers of the reactant and the transition
state, provides information about the number of symmetrically
equivalent pathways between the reactant and the transition
state of the corresponding elementary step in the reaction
family.27,60,61
The reparameterized form of single event rate coecient was
used to reduce correlation between the pre-exponential factors
and activation energies:

91
71
71
40
15
54
22

E 1
1
k = A exp a

R T
Tm

4.3. Rate Equations. The rate of every step is explicitly


accounted for in the microkinetic model. The net rate of
formation of every species is obtained by summation of the
rates of the elementary steps in which these species are formed
or consumed.
4.3.1. Rate Equations for DME and Methane Formation in
Terms of Elementary Steps. The surface methoxy species
formed by adsorption and subsequent dehydration of methanol
on the acid site of the zeolite again reacts with gas-phase
methanol to form a dimethyloxonium ion (DMO+), step (iii) in
Table 1. Because the (de)protonation of DMO+ to DME, step
(iv) in Table 1 is potentially very fast, DME and DMO+ can be
considered in quasi-equilibrium, and, hence the DME net rate
of formation can be obtained as follows:

15
6

1
1
8

16

294

active site, distinct from the acid sites involved in the higher
hydrocarbon formation.
4.2. Single-Event Concept. In principle, the number of
rate coecients in the MTO reaction equals the number of
elementary steps and, hence, is very large. The single-event
methodology allows a denition of unique rate parameters
within a given reaction family and is used to reduce this number
of adjustable parameters in the model.58 Transition state
theory59 allows a relation of the rate coecient of an
elementary step to the corresponding single-event rate
coecient.
The standard entropy and enthalpy dierence between the
reactant and transition state species governs this rate coecient
as follows according to the Eyring equation:59
k=

S o
Ho
kBT
exp
exp

h
R
RT

R(DME) = R(DMO+)
= kF(DMO+)C R1+pMeOH k C(DMO+)C DMO+
(12)
+

Rather than producing DMO , the reaction of a surface


methoxy and methanol can also yield methane and formaldehyde and the regeneration of the acid site, step (v) in Table
1. However, it is not reactive under MTO reaction conditions
and, hence, its formation is considered as irreversible. The
methane net rate of formation based on an elementary step can
be written as

(7)

R(CH4) = kF(CH4)C R1+pMeOH

Symmetry eects can be factored out from the standard


entropy dierence between the reactant and the activated
complex of the elementary step considered:
S = ln gl + S

(13)

4.3.2. Rate Equations for Aromatic Hydrocarbon Pool


Species in Terms of Elementary Steps. Primary olen
formation in the MTO process is modeled through the
aromatic hydrocarbon pool mechanism as shown in Figure 1.
The mechanism starts with the methylation of p-xylene to form
TMeB+. The net rate of formation of TMeB+ in terms of
elementary steps, corresponding to steps (i)(ii)) in Figure 1,
can be written as

(8)

These symmetry eects are essentially related to rotational


degrees of freedom of the considered species and comprise an
intrinsic term and the external symmetry number. Additionally,
the number of chiral centers also has to be accounted for so
that the global symmetry number can be calculated as follows:

gl = ext n int
2

(11)

R(TMeB+) = 0 = kMe(PX)C R1+C PX


kDem(TMeB+)C TMeB+

(9)

The rate coecient of an elementary step can, hence, be


rewritten as the product of the number of single events, ne, and
single event rate coecient k:

+ kPr(DMeMCHDE)C DMeMCHDEC H+
kDe(TMeB+)C TMeB+
1496

(14)

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

formation, that is, a carbon number exceeding three, are based


on the elementary steps as shown in Figure 2. Methylation and
alkylation steps are involved in the formation of the higher
olens.
For methylation three single-event rate coecients are
considered depending on the type of product carbenium ion,
that is, primary, secondary or tertiary:

Similar rate expressions are obtained for the other intermediates (DMeMCHDE, DMeEtB+, PDMeB+, DMeEtCHDE,
PX+) involved in the aromatic hydrocarbon pool.
The surface methoxy, net rate of formation, is obtained by
summation of elementary steps in which it is involved. It is
formed by dehydration of the methoxonium ion, and is again
consumed in the reverse reaction, that is, hydration, step (ii) in
Table 1. It is also consumed in the formation of DMO+ and
formed in the reverse reaction, step (iii) in Table 1. It is also
involved in the formation of methane, step (v) in Table 1. It
reacts in the aromatic hydrocarbon pool as a methylating agent,
steps (i), (iii), and (vi) in Figure 1 as well as in the alkene
homologation route, see eq 18 in section 4.3.4.

rMe(j) = kMe(i)C R1+pO(j)

(18)

The rate coecients for alkylation reactions depend on the


nature of the carbenium ions that participate as reactant and
product, that is, secondarysecondary, secondarytertiary,
tertiarysecondary but not on the olen that is being
consumed.

R(R1+) = 0 = kF(R1+)C R1+CMeOH+2 k C(R1+)C R1+C R1+pH O


2

rAlk(j) = kAlk(i)C R k+pO(j)

+ k C(DMO+)C DMO+ kF(DMO+)C R1+pMeOH

To further reduce the number of parameters to be estimated,


cracking is considered as the reverse of alkylation and, hence,
the corresponding rate coecients are calculated from the
alkylation rate coecients by applying thermodynamic
consistency and principles of microscopic reversibility. Hydride
shift, methyl shift, and PCP branching reactions yield olen
isomers and are typically equilibrated at MTO reaction
conditions.56 As a result, no corresponding rate coecients
are required in the model.
4.4. (De)protonation Equilibrium Relationship. The
concentration of carbenium ions with carbon numbers between
2 and 7 as required in eqs 1619, are obtained from the
protonation equilibrium and the corresponding olen which
leads to this carbenium ion:

kF(CH4)C R1+pMeOH + kDem(TMeB+)C TMeB+


kMe(PX)C R1+C PX + kDem(DMeEtB+)C DMeEtB+
kMe(DMeMCHDE)C R1+C DMeMCHDE
+ kDem(PDMeB+)C PDMeB+
kMe(DMeEtCHDE)C R1+C DMeEtCHDE
C R1+ kMe(i)pO

(15)

4.3.3. Rate equations for Primary Olens Formation in


Terms of Elementary Steps. Ethene is formed from the
aromatic hydrocarbon pool by DMeEtB+ dealkylation and is
consumed in the reverse reaction, step (iv) in Figure 1. Ethene
is also consumed in methylation (eq 18) and alkylation (eq 19)
reactions.

C R k+ = K prot(Oj )pO C H+
j

R(O2 ) = kDealk(DMeEtB )C DMeEtB+


kAlk(O2 , PX +)pO C PX+ rMe(O2 reactant)
2

(16)

Propene is formed via the aromatic hydrocarbon pool by


PDMeB+ dealkylation and is consumed in the reverse reaction,
step (vii) in Figure 1. Gas-phase propene is also involved in
various methylation (eq 18) and alkylation (eq 19) elementary
steps as reactant olen and in cracking elementary steps as a
product olen.

The number of unknown single-event protonationdeprotonation equilibrium coecients was reduced27,60,62 by expressing
them as a product of the single-event protonation/deprotonation equilibrium coecient of a well chosen reference alkene
Or, K prot(Or), and the single-event isomerization equilibrium
coecient between the alkene Oj and this reference alkene Or,
K isom (Oj; Or)

R(O3) = kDealk(PDMeB+)C PDMeB+ kAlk(O3 , PX +)pO

K prot(Oj ) = K isom(Oj ; Or )K prot(Or )

C PX+ rMe(O3reactant) + rCr(O3product)


rAlk(O3reactant)

(20)

Similar to the rate coecients of the elementary steps,


symmetry contributions in the protonation/deprotonation
equilibrium coecients were accounted for via symmetry
numbers of the alkene and the alkylcarbenium ion involved.
Oj
K prot(Oj ) =
K prot(Oj )
R k+
(21)

rAlk(O2 reactant)

(19)

(22)

The K prot(Or) comprises both the physical adsorption of alkene


and subsequent protonation on the acid site. The physisorption
is accounted for according to the experimental observations
made by Denayer et al.63 and only the protonation enthalpies of
the olens are estimated.
As a result, the expression for the alkylcarbenium ion
concentration becomes
Oj
C R k+ =
K isom(Oj ; Or )K prot(Or )pO C H+
j
R k+
(23)

(17)

Two types of propyl carbenium ions may occur, i.e., a


primary one formed by ethene methylation (eq 18) and a
secondary one involved in all other elementary steps, that is,
alkylation (eq 19) and cracking. Since the primary carbenium
ion is highly unstable compared to the secondary one, it is
reasonable to assume that it will be instantaneously converted
into secondary carbenium ion. Therefore, ethene methylation is
considered to contribute directly to the formation of the
secondary propyl carbenium ion.
4.3.4. Rate Equations for Higher Hydrocarbons in Terms
of Elementary Steps. The kinetic expressions for higher olens

K isom (Oj;Or) was calculated using thermodynamic data


generated with Bensons group contribution method.
1497

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

Table 3. Reaction Entropy Change and Forward and Reverse Single-Event Pre-exponential Factors for Elementary Steps/
Reaction Families Involved in Methanol to Olens
pre-exponential factors
(s1 or s1 bar1)
step/
equation

elementary step/reaction family


+
+ kDealk(DMeEtB )

(ii)a

DMeEtB + O2 + DMeB+

(iii)a

MeOH + R1+
DMO+
+

(v)a

MeOH + R1+ CH4 + HCHO + H+

(i)b

PX + R1+
TMeB+
+

(iii)b

DMeMCHDE + R1+
DMeEtB+
+

(vi)b

DMeEtCHDE + R1+
PDMeB+
+

(ii)b

TMeB+ DMeMCHDE

(v)b

DMEEtB+ DMeEtCHDE

(viii)b

DMeB+ PX

(iv)b

DMeEtB+ + O2 + DMeB+

(vii)b
eq18 and
eq 19
a

chemisorptionc

transition state
formationc

S
(J mol1 K1)

S (forward)
(J mol1 K1)

forward

reverse

51.8

0.0

1.1 1013

2.6 1010

54.3

54.3

1.9 1010

1.3 1013

164.5

3.3 104

176.5

115.1

1.3 107

2.1 1016

161.6

99.7

8.2 107

2.3 1016

173.0

110.8

2.1 107

2.4 1016

+1

193.7

61.4

2.1 1016

1.6 106

+1

191.5

61.9

2.3 1016

2.3 106

+1

197.8

60.9

2.3 1016

9.4 105

+1

141.0

61.9

2.3 1016

9.8 108

PDMeB+
O3 + DMeB+
+

+1

158.2

62.3

2.3 1016

1.3 108

methylation/alkylation

kAlk(O2,DMeB )

kF(DMO+)

kC(DMO )

kF(CH 4)

kMe(PX)

kDem(TMeB )

kMe(DMeMCHDE)
kDem(DMeEtB )

kMe(DMeEtCHDE)
kDem(PDMeB )

kDe(TMeB+)

kPr(DMeMCHDE)

kDe(DMeEtB+)

kPr(DMeEtCHDE)

kDe(DMeB+)
kPr(PX)

kDealk(DMeEtB+)
kAlk(O2,DMeB )

kDealk(PDMeB+)
kAlk(O3,DMeB )

2.1 107

See Table 1 for corresponding step. bSee Figure 1 for corresponding step. cLoss/gain of translational degree of freedom.

4.6. Acid Site Balance. The concentration of available acid


sites, CH+ can be obtained by applying a balance over the acid
sites.

The carbenium ion concentration is a function of the


concentration of available acid sites CH+, the partial pressure of
gas-phase species, and equilibrium coecient of the corresponding (de)protonation reaction.
Since, protonation of MeOH and DME are in equilibrium
also, the MeOH+2 and DMO+ surface concentrations can be
obtained from the corresponding (de)protonation equilibrium
relationships:
CMeOH+2 = KPr(MeOH)pMeOH C H+

(24)

C DMO+ = KPr(DME)pDME C H+

(25)

C Ht+ = C H+ + CMeOH+2 + C DMO+ +

+
k

(27)

whereas concentration of the oxonium and carbenium ions with


carbon numbers between 2 and 7 can be obtained from the
protonation equilibrium relationship, as these steps are quasiequilibrated (see section 4.4). The concentration of CR1+ has to
be determined by the pseudosteady state approximation as
mentioned in section 4.3.2 and, hence, the balance needs to be
written as follows:

4.5. Aromatic Hydrocarbon Pool Active Site Balance.


The total concentration of aromatic hydrocarbon pool species,
that is, polymethylbenzenes, involved in the formation of
primary olens is considered as a single adjustable parameter
and is determined by regression. The concentration of
intermediate aromatic hydrocarbon pool species and surface
methoxy species, which is also involved in the methylation
reaction of aromatic hydrocarbon pool species is obtained by
applying the pseudosteady state approximation as explained in
section 4.3.2. The aromatic hydrocarbon pool active sites
balance can be written as

C Ht+ = C H+ + CMeOH+2 + C DMO+ + C R1+ +

CR
k1

+
k

(28)

C H+ =
C Ht+ C R1+
1 + KPr(MeOH)pMeOH + KPr(DME)pDME + j KPr(Oj )pO

(29)

The total concentration of acid sites CtH+, is calculated as 8.3


102 mol H kgcat1 catalyst taking into account the Si/Al ratio
of 200.

(C PX )t = C TMeB+ + C DMeMCHDE + C DMeEtB+ + C DMeEtCHDE


+ C PDMeB+ + C PX

CR

(26)
1498

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

5. MODEL PARAMETER DETERMINATION


The SEMK model parameters can be divided into single-event
pre-exponential factors and activation energies/protonation
enthalpies. To keep the number of adjustable parameters within
tractable limits, the single-event pre-exponential factors were
calculated a priori based on statistical thermodynamics as
shown in Table 3. Activation energies and protonation
enthalpies were estimated by model regression to experimental
data. The total concentration of aromatic hydrocarbon pool
species, (CPX)t, is considered as an adjustable parameter as
discussed in section 4.5, while the concentration of free acid
sites is obtained by applying the acid-site balance as given in
section 4.6.
5.1. Single-Event Pre-exponential Factors Calculation.
The single-event pre-exponential factor for an elementary step j
is given by
A jfor =

S 0,
kBT
j

exp

h
R

0
Strans

0,

nreact


= STS

(30)

i
iSsurf,

(31)

i=1

,
The single-event standard entropy of the transition state, S0TS
and the single-event standard entropy of the surface species i
are calculated from the single-event standard entropy of
0
associated molecules in the gas phase, Sgas,i
, and their
corresponding single-event standard chemisorption entropy
change due to their chemisorption on the catalyst, S0chem,i:
0

i = Sgas,
i + Schem,
i
Ssurf,

(32)

The single-event pre-exponential factor for the reverse


elementary step is calculated from the forward single-event
pre-exponential factor and the single-event reaction entropy,
following the principle of microscopic reversibility:

A jrev =

A jfor
Sr 0,j
exp R

(33)

The single-event entropy change of an elementary reaction j,


Sr,j0, is given by
0

Sr , j =

n prod

i=1

iSi

nreact

k=1

k Sk

3/2
kBT
5
+ R
2
2
h

( )
Mw
NA

(35)

For DME formation, an entropy loss of two translational


degrees of freedom upon chemisorption is assumed for ionic
molecules, step (ii) in Table 3, while for neutral molecules an
entropy loss of a single translational degree of freedom is
assumed, steps (iii) and (v) in Table 3, as the motion of an
ionic molecule is expected to be more signicantly reduced
upon chemisorption as compared to neutral molecules. In
transition state formation, no entropy change is considered in
dehydration, step (ii) in Table 3 while for methoxonium ion
formation an entropy loss of a single translational degree of
freedom is assumed, step (iii) in Table 3. Doing so, the
obtained single-event pre-exponential comes in the range of
pre-exponential factors previously reported for this type of
elementary step67 and close to values calculated using DFT.6
The formation of methane, step (v) in Table 3, being a slow
step, an entropy loss equal to two translational degrees of
freedom is assumed in the transition state.
For the aromatic hydrocarbon pool species, an entropy loss
corresponding to three translational degrees of freedom was
assumed upon chemisorption for ionic species (steps (ii), (iv),
(v), (vii), and (viii) in Table 3), while for neutral molecules an
entropy loss corresponding to a single translational degree of
freedom was assumed (steps (i), (iii), and (vi) in Table 3).
For the formation of transition state species, an entropy loss
or gain corresponding to one degree of translational freedom
was assumed depending on the type of reaction. The singleevent pre-exponential obtained for dealkylation, steps (iv) and
(vii) and for deprotonation, steps (ii), (v), and (viii) in Table 3,
for the forward direction are comparable to the single-event
pre-exponential used for exocyclic beta-scission and deprotonation, that is, 2.0 1016 for catalytic cracking of (cyclo)alkanes.68
For methylation and alkylation reactions the single-event
entropy change between reactant and transition state is
calculated based on the translational entropy of the reference
olen. It is assumed that the entropy change is equal to the loss
of one translational degree of freedom of the reference olen.
In summary, the loss of translational freedom upon
chemisorption and gain/loss upon transition state formation
are considered dependent on the type of molecules involved in
the elementary steps to have reasonable pre-exponential factor
values, within the range of values reported for same type of
elementary step.67 The variation of standard entropies of
species with their molecular mass is similar for both the
reactant and product in elementary steps belonging to the same
reaction family and, hence, in agreement with the single-event
concept, the corresponding activation and reaction entropies
are kept constant at an average value.
5.2. Protonation Enthalpies and Activation Energy
Estimation. The total number of adjustable parameters in the
SEMK model for MTO amounts to 29, i.e., 21 activation
energies and 8 protonation enthalpies. Out of 21 activation
energies, 4 are for DME formation, steps (ii) and (iii) in Table
1, 1 for methane formation, step (v) in Table 1, and 10 for
aromatic hydrocarbon pool reactions, steps (i)(viii) in Figure
1. It may be noted that elementary steps of same type, i.e.,
deprotonation of TMeB+ and DMeEtB+, which both form an

The single-event standard activation entropy for an elementary

step j, S0,
j , between the reactant and corresponding transition
state is calculated as follows:
Sj

2
RT
= R ln o
p NA

(34)

The standard entropies of the associated molecules in the gas


, are obtained directly from databases64 or calculated
phase, S0gas,i
using group contribution methods65 or density functional
theory.66
It has been assumed that the entropy change associated to a
chemisorption step is dominated by the loss of translational
entropy.67 The translational entropy of a gas phase molecule
has been approximated using the Sackur-Tetrode equation:
1499

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

Table 4. Estimated Forward and Reverse Activation Energies and Protonation Enthalpiesa
forward
kJ mol1

activation energy

reverse
kJ mol1

Surface Methoxy, and DME Formation


222 8
170 15
138 7
163.7 14
Methane Formation
methane formation (step (v) of Table 1)
121 18
Aromatic Hydrocarbon Pool Species Formation
methylation of p-xylene (step (i) of Figure 1)
101 1
123 17
deprotonation of TMeB+ and DMeEtB+ (step ii and v of Figure 1)
156 9
123 16
methylation of DMeMCHDE and DMeEtCHDE (steps (iii) and (vi) of Figure
75 16
151 17
1)
84 8
14 2
dealkylation of DMeEtB+ and PDMeB+ (steps (iv) and (vii) of Figure 1)
deprotonation of PX+ (steps (viii) of Figure 1)
143 20
123 16
Higher Olens Formation
methylation (pp) (eq 18)
131 28
methylation (ps) (eq 18)
92 10
methylation (pt) (eq 18)
54 9
alkylation (ss) (eq 19)
138 9.4
alkylation (st) (eq 19)
119 17
alkylation (ts) (eq 19)
167 28
dehydration (step (ii) of Table 1)
protonation with MeOH (step (iii) of Table 1)

protonation enthalpy

kJ mol1

methanol (step i of Table 1)


DME (step iv of Table 1)

61 3
42 8

ethene (eq 20)


propene (eq 20)
butene (eq 20)
pentene (eq 20)
hexene (eq 20)
heptene (eq 20)

11
42
53
61
67
70

0.1
7
9
15
2
12

Values with a 95% condence interval obtained by regression of experimental data at T = 360480 C, Pt = 1.04 bar and W/FMeOH = 0.56.5 kgcat s
mol1 with the pre-exponential factors from Table 3 by integration of eq 3 and simultaneously solving eq 5.

exocyclic double bond on the aromatic ring, steps (ii) and (v)
in Figure 1, methylation of DMeMCHDE and DMeEtCHDE,
steps (iii) and (vi) in Figure 1, in which the exocyclic double
bond on the aromatic ring undergoes methylation and
dealkylation of DMeEtB+ and PDMeB+, steps (iv) and (vii)
in Figure 1, which split o ethene and propene, are grouped
together and assigned a single parameter for each reaction type.
There are six activation energies related to methylation and
alkylation reactions (eq 18 and eq 19) based on the type of
carbenium ions involved as reactant and product in the
elementary steps. In methylation a surface methoxy reacts with
a gas phase olen to form a product carbenium ion of either
primary, secondary or tertiary nature, that is, (p,p), (p,s), and
(p,t). Hence three activation energies based on product
carbenium ion formed are considered for methylation reactions.
In alkylation, a surface carbenium ion reacts with a gas phase
olen forming a product carbenium ion, three additional types
of kinetic coecients are distinguished depending on the type
of reactant and product carbenium ion, that is, (s, s), (s, t), and
(t, s). Hence, three activation energies based on the type of
reactant and product carbenium ion involved are considered.
Protonation enthalpies are required for methanol, step (i) in
Table 1, DME, step (iv) in Table 1, and each of the olens
from carbon number 2 to 7 (eq 22). Some additional
physicochemical constraints and Boudarts criteria69 have
been used to limit the parameter space. In agreement with
what is typically reported with respect to carbenium ion
stability,62 the protonation enthalpies of the reference olens
have been imposed to satisfy the following relationship:

A SEMK model regression resulted in the parameter estimates


and corresponding 95% individual condence intervals
reported in Table 4. The F value for the global signicance
of regression amounts to 290, which exceeds the tabulated
value of 2.79. All calculated t values are considerably higher
than the tabulated value of 1.96. The maximum absolute value
among the binary correlation coecients was 0.71 indicating
that there is no strong binary correlation between the
parameters. The parity diagrams of MTO products for the
complete range of experimental conditions used for parameter
estimation are presented in Figure 3, while the agreement
between model calculation and experimental data in terms of
yield is shown in Figure 47. The agreement between model
simulations and experiments is remarkable, taking into account
the complex product pattern encountered in MTO.
5.2.1. Activation Energies and Protonation Enthalpy for
DME Formation. The obtained activation energy for DME
formation is reported in Table 4 and is in agreement with the
literature reported values. The methanol and DME protonation
enthalpies obtained amount to 62 and 43 kJ mol1 which
are also close to the values obtained previously,28 that is, 69.3
and 40.1 kJ mol1, respectively. The methane formation is
slow and requires an activation energy of 121 kJ mol1 as
compared to 122.12 kJ mol1 obtained previously.28 The
activation energies obtained for methoxonium ion dehydration,
222 and 170 kJ mol1 for the forward and the reverse step, are
close to the values calculated using DFT,6 223.6 and 176.2 kJ
mol1. The estimated activation energies for methoxonium ion
protonation with MeOH to form dimethyloxonium ion amount
to 138 and 163 kJ mol1 for forward and reverse steps and are
also close to the values of 141 and 187.6 kJ mol1 reported
using DFT.6
5.2.2. Activation Energies for Aromatic Hydrocarbon Pool
Mechanism. The estimated activation energies for the aromatic
hydrocarbon pool species follow the trend of activation
enthalpies for transition state formation from the reactive
intermediates as obtained from theoretical calculations.52 The
activation energy obtained for p-xylene methylation is 25 kJ

HPr 0(O7r ) < HPr 0(O6r ) < ... < HPr 0(O3r )
< HPr 0(O2r )

(36)

For each of the protonation steps, Boudarts criterion69 denes


a rigorous constraint for the standard protonation enthalpy
(HPr0) and entropy (SPr0):
41.8 < SPr 0 < 51.04 + 1.4 103 (HPr 0)

(37)
1500

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

Figure 3. Parity diagram of experimental vs calculated outlet MTO products ow rates at T = 360480 C, pt = 1.04 bar and W/FMeOH = 0.56.5
kgcat s mol1: (line) experimental; () obtained from model regression with the pre-exponential factors from Table 3 and the set of parameter
estimates given in Table 4 by integration of eq 3 and simultaneously solving eq 5 with the corresponding net production rates eqs 12, 13, 16, 17, 18
and 19.
1501

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

kJ mol1 for ethene as its protonation exclusively leads to a


primary carbenium ion to 70 kJ mol1 for heptene as shown
in Table 4. The values obtained are in line with those reported
for n-hexane hydroconversion on ZSM-5, in which hexene
protonation71 to a secondary carbenium ion was found to
exhibit a protonation enthalpy of 65.4 2.3 kJ mol1 as
compared to 68 kJ mol1 obtained in the present work. The
increment in standard protonation enthalpy decreases as the
carbon number increases because additional stabilization eects
provided by CC bonds to alkyl carbenium ions gradually
become less pronounced.62
5.3. Comparison between Experiment and Model
Simulation. The SEMK model adequately simulates the
various MTO products with physically realistic parameter
values. The estimated value of the total concentration of
aromatic hydrocarbon pool species amounts to 3.47 102 mol
kgcat1 which is about half of the total acid site concentration,
that is, 8.3 102 mol h kgcat1, Figure 4 and Figure 5 show

Figure 4. Experimental and model calculated yield of MTO products


at T = 400 C and pt = 1.04 bar. Symbols: experimentally observed
values, ethene (), propene (), butene (). Lines: calculated by
integration of eq 3 and simultaneously solving eq 5 with the
corresponding net production rates eqs 16, 17, 18 and 19 with preexponential factors from Table 3 and parameter estimates given in
Table 4.

mol1 higher than that for methylation of DMeMCHDE and


DMeEtCHDE because the former involves methylation of the
aromatic ring as compared to methylation of the exocyclic
double bond in the case of DMeMCHDE and DMeEtCHDE.
For deprotonation, TMeB+ and DMeEtB+ require a slightly
higher activation energy than PX+ because the former involve
the deprotonation of the aromatic ring, whereas the latter
involves side chain deprotonation.
5.2.3. Activation Energies for Higher Olens Formation.
The activation energies obtained for the elementary steps
involved in higher olen formation were in line with the
stability of the carbenium ions involved as reactants and
products, that is, less stable reactants and more stable products
lead to lower activation energies. For example, in the
methylation reaction, where a surface methoxy reacts with a
gas phase olen to form a product carbenium ion of either
primary, secondary, or tertiary nature, the activation energy of
the reaction leading to a primary carbenium ion, 132 kJ mol1,
is higher than that of the reaction leading to a secondary
carbenium ion, 93 kJ mol1, which, in turn, is higher than the
one leading to a tertiary carbenium ion, 55 kJ mol1. The
intrinsic activation energy determined for ethene methylation
on H-ZSM-5 catalyst70 amounted to 135 kJ mol1 which is
close to values obtained in this work, 131.9 kJ mol1. Among
the alkylation reactions, in which a surface carbenium ion reacts
with a gas phase olen, three types of elementary steps are
distinguished, (s, s), (s, t), and (t, s). The activation energy for
(s, s), 138 kJ mol1, is higher than for (s, t), 120 kJ mol1, as
the formation of a tertiary carbenium ion requires less energy
because it is more stable than a secondary ion. The activation
energy required for formation of a secondary carbenium ion
from tertiary carbenium ion is the highest, 168 kJmol1, as a
more stable reactant is converted to less stable product.
5.2.4. Alkene Protonation Enthalpies. The formation of
carbenium ions from gas phase alkenes is composed out of two
steps: a physical adsorption, denoted in short as physisorption,
inside the catalyst pore structure and subsequent protonation of
the double bond at an acid site. The physisorption of alkenes is
accounted for according to the experimental observations made
by Denayer et al.63 The protonation enthalpies are obtained by
regression to experimental data. The estimated protonation
enthalpies exhibited a carbon number dependence which is
most pronounced at the lowest carbon numbers. The
protonation enthalpies of the reference olens vary from 11

Figure 5. Experimental and model calculated yield of MTO products


at T = 400 C and pt = 1.04 bar. Symbols: experimentally observed
values, pentene (), hexene (), heptene (), methane (). Lines:
calculated by integration of eq 3 and simultaneously solving eq 5 with
the corresponding net production rates eqs 13, 18 and 19 with preexponential factors from Table 3 and parameter estimates given in
Table 4.

olen yields versus space time at 400 C. A generally increasing


trend of the olen yields with the space time is experimentally
obtained and simulated. The lower the carbon number is, the
less pronounced this increase is. For ethene, the olen with the
lowest possible carbon number, the experimentally obtained
and simulated yield even level o at higher space times,
indicative of the participation of ethylene in consecutive
reactions.
Figure 6 shows the ethene and propene yield versus
temperature at 65% methanol conversion. It can be seen
that as the temperature increases the yield of ethene and
propene increases. However, the increase of the propene yield
is much more pronounced as compared to ethene. This can be
explained by the increase in the contribution of alkene
homologation cycle at higher temperatures for product
formation as compared to the aromatic hydrocarbon pool,
since ethene is exclusively formed via the aromatic hydrocarbon
pool mechanism, while propene is formed both via the aromatic
hydrocarbon pool as well as the alkene homologation cycle, see
also section 6.
Figure 7 shows the experimentally observed and model
calculated yield of dierent olenic products at a space time of
4.28 kgcat s mol1 and 360 C. The yield of primary olens,
ethene and propene, largely exceeds that of the higher carbon
1502

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

from cracking as compared to ethene formation. The ratio of


propene to butene also decreases because the increase in yield
of olenic products is more pronounced for higher carbon
number olens as the space time increases because the alkene
homologation cycle plays a more dominant role.
A preliminary assessment of hexene cofeeding in MTO and
cracking25 has also been performed. Lower amounts of
methane, ethene, and propene were obtained in favor of
butene and higher alkenes. This can be attributed to a reduced
contribution of the hydrocarbon pool mechanism in the overall
reaction because of the lower methanol partial pressure. In turn,
the amounts of higher olens increase due to hexene cracking
and its subsequent involvement in methylationalkylation and
cracking via the alkene homologation cycle. Hence, upon
cofeeding hexene the alkene homologation cycle becomes more
dominant. When simulating the behavior of a pure hexene feed
on H-ZSM-5, the conversion is signicantly underestimated
because typical catalytic cracking reaction families such as
hydride transfer and protolytic cracking are not accounted for
in the present MTO model. These reaction families appear not
to be determining the catalyst activity when methanol/
methylating agents are present. The reactions determining the
selectivity in both cases, that is, hexane cofeeding and cracking,
however, seem to be similar and, hence, result in product
distributions that resemble each other.

Figure 6. Experimental and model calculated yield of primary olens


at methanol conversion = 65% and pt = 1.04 bar. Symbols:
experimentally observed values, ethene (), propene (). Lines:
calculated by integration of eq 3 and simultaneously solving eq 5 with
the corresponding net production rates eqs 16, 17, 18 and 19 with preexponential factors from Table 3 and parameter estimates given in
Table 4.

6. CONTRIBUTION ANALYSIS
A dierential contribution analysis was carried out to
understand the relative importance of the various elementary
steps involved in MTO. Figure 9 shows the contribution of
Figure 7. Experimental and model calculated yield of olenic products
at T = 360 C, pt = 1.04 bar and W/FMeOH = 4.28 kgcat s mol1.
Calculated by integration of eq 3 and simultaneously solving eq 5 with
the corresponding net production rates eqs 13, 16, 17, 18 and 19 with
pre-exponential factors from Table 3 and parameter estimates given in
Table 4.

number olenic products, as the catalyst used in present study


is most selective for light olens formation.28
Figure 8 shows the ratio of dierent olenic products vs
space time. As the space time increases the ratio of ethene to
propene decreases because of enhanced propene formation

Figure 9. Contribution analysis for conversion of methanol into olens


on H-ZSM-5 catalyst; T = 400 C; W/FMeOH = 2.21 kgcat s mol1; pt =
1.04 bar; calculated by integration of eq 3 and eq 5 with the
corresponding net production rates eqs 16, 17, 18 and 19 with preexponential factors from Table 3 and the parameter estimates given in
Table 4.

dierent elementary steps in olens formation at 400 C and a


space time amounting to 2.21 kgcat s mol1. The numbers
indicate the percentage contribution of reaction families to the
formation or disappearance of the species involved in the
reaction mechanism. At these conditions, a conversion of 66.9%
is obtained. As already indicated above, ethene is exclusively
formed from the aromatic hydrocarbon pool, as primary
carbenium ion formation from cracking is not considered. Its
consumption in methylation (58%) is slightly higher than
alkylation (42%) reactions because for alkylation it requires

Figure 8. The ratio O2/O3 and O3/O4 versus space time at T = 480 C
and pt = 1.04 bar. Symbols: experimentally observed values O2/O3
(), O3/O4 (). Lines: calculated by integration of eqs 3 and
simultaneously solving eq 5 with the corresponding net production
rates eqs 16, 17, 18 and 19 with pre-exponential factors from Table 3
and the parameter estimates given in Table 4.
1503

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

higher activation energies as compared to methylation reactions


and, hence, are favored at higher temperature. Focusing on
propene formation from ethane methylation and the aromatic
hydrocarbon pool only, that is, excluding propene formation by
cracking from heavier alkenes, it is evident from the numbers
cited above that the aromatic hydrocarbon pool is contributing
more signicantly as compared to ethane methylation at higher
temperatures. The observations are in line with the
experimentally70,72,73 studied methylation reaction in the
MTO on ZSM-5 type catalyst.

formation of a primary carbenium ion which is energetically


more dicult. Propene formation occurs via both the aromatic
hydrocarbon pool (22%), and alkene homologation route
(78%). The latter production route comprises ethene
methylation (66%) and cracking of heavier olens (12%).
Just like ethene, propene is almost equally consumed in
methylation (47%) and alkylation (53%) reactions, although
consumption via alkylation is somewhat more pronounced for
propene than for ethene, because its protonation results in a
more stable secondary carbenium ion. Butene is the lightest
olen that is no longer formed via the aromatic hydrocarbon
pool mechanism, but via methylation (67%), alkylation (29%),
and cracking (4%). As butene formation via alkylation will
require an ethyl primary carbenium ion, its formation from
methylation is much more pronounced as compared to
alkylation. Its consumption via methylation (57%) and
alkylation (43%) is similar to that of ethene. The olens
from pentene onward show a clear trend to be formed more by
alkylation than by methylation as they no longer require a
primary carbenium ion in alkylation. Pentene is formed (57%)
from alkylation as compared to (42%) from methylation and
consumed in methylation (62%) and alkylation (38%). Hexene
is also formed more prominently from alkylation (73%) than
from methylation (27%). On the other hand, it is consumed in
methylation (17%) and cracking (83%) to form lower olens.
Heptene is formed from methylation (32%) and alkylation
(68%) and consumed only in cracking reactions.
Figure 10 shows the contribution analysis at 480 C and at a
space time of 1.40 kgcat s mol, corresponding to a conversion of

7. CONCLUSIONS
The MTO kinetics on H-ZSM-5 can be adequately described
using the single-event microkinetic (SEMK) methodology. The
single-event concept along with thermodynamic constraints
allows a drastic reduction of the number of adjustable
parameters. The pre-exponential factors can be calculated
using statistical thermodynamics based on logical assumptions
with respect to the gain or loss of translational degrees of
freedom. More unstable species exhibit more entropy loss upon
chemisorption compared to more stable, neutral species. The
estimated activation energies and protonation enthalpies reect
the trends as expected considering the type of the carbenium
ions involved. Olen protonation enthalpies become more
negative with the carbon number; which is related to an
increased potential number of carbon atoms in the position of
the charge. Ethene exclusively forms from the aromatic
hydrocarbon pool while propene forms from both the aromatic
hydrocarbon pool and alkene homologation cycle. The higher
the carbon number of the considered olen is, the lower the
contribution of methylation is to their formation compared to
alkylation. At higher temperatures, the contribution of the
aromatic hydrocarbon pool to propene formation is proportionally larger than that of ethene methylation. However, the
enhancement of the more highly activated alkylation-cracking
steps by the temperature and their corresponding contribution
to alkene formation, including propene, is even more
pronounced.

AUTHOR INFORMATION

Corresponding Author

*Fax:+32 9 264 49 49. E-mail: Joris.Thybaut@UGent.be.


Notes

The authors declare no competing nancial interest.

Figure 10. Contribution analysis for conversion of methanol into


olens on H-ZSM-5 catalyst; T = 480 C, W/FMeOH =1.40 kgcat s
mol1, pt = 1.04 bar, calculated by integration of eq 3 and eq 5 with the
corresponding net production rates eqs 16, 17, 18 and 19 with preexponential factors from Table 3 and the parameter estimates given in
Table 4.

ACKNOWLEDGMENTS
This paper reports work undertaken in the context of the
project OCMOL, Oxidative Coupling of Methane followed by
Oligomerization to Liquids. OCMOL is a Large Scale
Collaborative Project supported by the European Commission
in the seventh Framework Programme (GA No. 228953). For
further information about OCMOL see http://www.ocmol.eu
or http://www.ocmol.com.

73.6%. Compared to the analysis at 400 C, the methylation


reactions are much less dominant, for example, 58% ethene
consumption via methylation at 400 C compared to 43% at
480 C. Likewise, the alkylation-cracking reactions contribute
more signicantly to the product formation, 53% propene
consumption via alkylation and 12% propene formation from
cracking at 400 C, compared to 79% and 39% at 480 C. The
propene formation from methylation decreases from 66% to
42% and from aromatic hydrocarbon pool from 22% to 19%
indicating that as the temperature increases the alkylation/
cracking reaction becomes more important. This can be
attributed to the fact that the alkylation reactions are having

NOMENCLATURE

Roman Symbols

A = single event pre-exponential factor (s1 or s1 bar1)


C = concentration (mol kgcat1)
CtH+ = total concentration of acid sites (mol kgcat1)
CH+ = concentration of available acid sites (mol kgcat1)
Ea = activation energy of reaction (kJ mol1)
F = molar ow rate (mol s1)

1504

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

H = enthalpy (kJ mol1)


k = rate coecient of an elementary step (s1 or s1 bar1)
k = single event rate coecient (s1 or s1 bar1)
Kisom = equilibrium coecient for isomerization between
alkene (kgcat mol1)
Kprot = equilibrium coecient for protonation of alkene j
(kgcat mol1)
ne = number of single events
nresp = number of responses
Oj = olen with carbon number j
pt = total pressure (bar)
PCP = protonated cyclopropane
Rk+ = carbenium ion k
Rk = net production rate of surface species (mol kgcat1 s1)
Rj = net production rate of component j (mol kgcat1 s1)
p = partial pressure (bar)
R1+ = surface methoxy or methoxonium ion
S = entropy (J mol1 K1)
SSQ = sum of squares of residuals between observed and
calculated response values (mol2 s2)
T = temperature (K)
W = catalyst mass (kg)

PX+ = 1,4-dimethylbenzenium cation


resp = responses
surf = surface
t = total
trans = translational
TMEB+ = 1,1,4-trimethylbenzenium ion
TS = transition state

REFERENCES

(1) Olsbye, U.; Svelle, S.; Bjrgen, M.; Beato, P.; Janssens, T. V. W.;
Joensen, F.; Bordiga, S.; Lillerud, K. P. Conversion of Methanol to
Hydrocarbons: How Zeolite Cavity and Pore Size Controls Product
Selectivity. Angew. Chem. Int. Ed. 2012, 51, 5810.
(2) Economides, M. J.; Mokhatab, S. Compressed natural gas
Another solution to monetize stranded gas. Hydrocarbon Process. 2007,
86, 59.
(3) Stocker, M. Methanol-to-hydrocarbons: Catalytic materials and
their behavior. Microporous Mesoporous Mater. 1999, 29, 3.
(4) Wang, W.; Buchholz, A.; Seiler, M.; Hunger, M. Evidence for an
initiation of the methanol-to-olefin process by reactive surface
methoxy groups on acidic zeolite catalysts. J. Am. Chem. Soc. 2003,
125, 15260.
(5) Lesthaeghe, D.; Van Speybroeck, V.; Marin, G. B.; Waroquier, M.
Understanding the failure of direct CC coupling in the zeolitecatalyzed methanol-to-olefin process. Angew. Chem., Int. Ed. 2006, 45,
1714.
(6) Lesthaeghe, D.; Van Speybroeck, V.; Marin, G. B.; Waroquier, M.
The rise and fall of direct mechanisms in methanol-to-olefin catalysis:
An overview of theoretical contributions. Ind. Eng. Chem. Res. 2007,
46, 8832.
(7) Haw, J. F.; Song, W. G.; Marcus, D. M.; Nicholas, J. B. The
mechanism of methanol to hydrocarbon catalysis. Acc. Chem. Res.
2003, 36, 317.
(8) Bjorgen, M.; Olsbye, U.; Petersen, D.; Kolboe, S. The methanolto-hydrocarbons reaction: Insight into the reaction mechanism from
[C-12]benzene and [C-13]methanol coreactions over zeolite H-beta. J.
Catal. 2004, 221, 1.
(9) Bjorgen, M.; Joensen, F.; Lillerud, K. P.; Olsbye, U.; Svelle, S.
The mechanisms of ethene and propene formation from methanol
over high silica H-ZSM-5 and H-beta. Catal. Today 2009, 142, 90.
(10) Olsbye, U.; Bjorgen, M.; Svelle, S.; Lillerud, K. P.; Kolboe, S.
Mechanistic insight into the methanol-to-hydrocarbons reaction. Catal.
Today 2005, 106, 108.
(11) Mole, T.; Whiteside, J. A. Conversion of methanol to ethylene
over ZSM-5 zeolite in the presence of deuterated water. J. Catal. 1982,
75, 284.
(12) Langner, B. E. Reactions of methanol on zeolites with different
pore structures. Appl. Catal. 1982, 2, 289.
(13) Dahl, I. M.; Kolboe, S. On the reaction-mechanism for propene
formation in the MTO reaction over Sapo-34. Catal. Lett. 1993, 20,
329.
(14) Mikkelsen, O.; Ronning, P. O.; Kolboe, S. Use of isotopic
labeling for mechanistic studies of the methanol-to-hydrocarbons
reaction. Methylation of toluene with methanol over H-ZSM-5, Hmordenite and H-beta. Microporous Mesoporous Mater. 2000, 40, 95.
(15) Froment, G. F.; Bischo, K. B. Chemical Reactor Analysis and
Design, 2nd ed.; Wiley: New York, 1990.
(16) Thybaut, J. W.; Narasimhan, C. S. L.; Marin, G. B. Single event
microkinetics (SEMK) as a tool for catalyst and process design. Stud.
Surf. Sci. Catal. 2006, 159, 5560.
(17) Thybaut, J. W.; Sun, J. J.; Olivier, L.; Van Veen, A. C.;
Mirodatos, C.; Marin, G. B. Catalyst design based on microkinetic
models: Oxidative coupling of methane. Catal. Today 2011, 159, 29.
(18) Chen, N. Y.; Reagan, W. J. Evidence of auto-catalysis in
methanol to hydrocarbon reactions over zeolite catalysts. J. Catal.
1979, 59, 123.
(19) Chang, C. D. Kinetic-model for methanol conversion to
hydrocarbons. Chem. Eng. Sci. 1980, 35, 619.

Constants

h = Plank constant (6.626068 1034 m2 kg s1)


kB = Boltzmann constant (1.380653 1023 m2 kg s1 K1)

Greek Symbols

= vector of parameters
= global symmetry number
jk = inverse of the covariance between the experimental
errors associated with measurements of the jth and kth
responses
Superscript

= model calculated values


= transition state
= single event
r = reactant
ref = reference
Subscript

0 = standard state
Alk = alkylation
C = consumption
chem = chemisorption
cr = cracking
De = deprotonation
Dealk = dealkylation
Dem = demethylation
DMeB+ = 1,1-dimethylbenzenium cation
DMeEtB+ = 1,1-dimethyl-4-ethylbenzenium cation
DMeEtCHDE = 3,3-dimethyl-6-ethylene-1,4-cyclohexadiene
DMeMCHDE = 3,3-dimethyl 6-methylene-1,4-cyclohexadiene
ext = external
F = formation
gl = global
int = internal
m = mean
Me = methylation
ob = observations
ox = oxonium ion
PDMeB+ = 1-isopropyl-4,4-dimethylbenzenium cation
Pr = protonation
PX = para-xylene
1505

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

(20) Schoenfelder, H.; Hinderer, J.; Werther, J.; Keil, F. J. Methanol


to olefins-prediction of the performance of a circulating fluidized-bed
reactor on the basis of kinetic experiments in a fixed-bed reactor.
Chem. Eng. Sci. 1994, 49, 5377.
(21) Bos, A. N. R.; Tromp, P. J. J.; Akse, H. N. Conversion of
methanol to lower olefins - kinetic modeling, reactor simulation, and
selection. Ind. Eng. Chem. Res. 1995, 34, 3808.
(22) Gayubo, A. G.; Aguayo, A. T.; del Campo, A. E. S.; Tarrio, A.
M.; Bilbao, J. Kinetic modeling of methanol transformation into olefins
on a SAPO-34 catalyst. Ind. Eng. Chem. Res. 2000, 39, 292.
(23) Chen, D.; Grlnvold, A.; Moljord, K.; Holmen, A. Methanol
conversion to light olefins over SAPO-34: Reaction network and
deactivation kinetics. Ind. Eng. Chem. Res. 2007, 46, 4116.
(24) Hu, H.; Ying, W. Y.; Fang, D. Y. Mathematical modeling of
multibed adiabatic reactor for the methanol-to-olefin process. React.
Kinet. Mech. Catal. 2010, 101, 49.
(25) Kaarsholm, M.; Rafii, B.; Joensen, F.; Cenni, R.; Chaouki, J.
Patience, G. S. Kinetic modeling of methanol-to-olefin reaction over
ZSM-5 in fluid bed. Ind. Eng. Chem. Res. 2010, 49, 29.
(26) Mihail, R.; Straja, S.; Maria, G.; Musca, G.; Pop, G. A kineticmodel for methanol conversion to hydrocarbons. Chem. Eng. Sci. 1983,
38, 1581.
(27) Park, T. Y.; Froment, G. F. Kinetic modeling of the methanol to
olefins process. 1. Model formulation. Ind. Eng. Chem. Res. 2001, 40,
4172.
(28) Park, T. Y.; Froment, G. F. Kinetic modeling of the methanol to
olefins process. 2. Experimental results, model discrimination, and
parameter estimation. Ind. Eng. Chem. Res. 2001, 40, 4187.
(29) Alwahabi, S. M.; Froment, G. F. Single event kinetic modeling of
the methanol-to-olefins process on SAPO-34. Ind. Eng. Chem. Res.
2004, 43, 5098.
(30) Ilias, S.; Bhan, A. Tuning the selectivity of methanol-tohydrocarbons conversion on H-ZSM-5 by co-processing olefin or
aromatic compounds. J. Catal. 2012, 290, 186.
(31) Berger, R. J.; Stitt, E. H.; Marin, G. B.; Kapteijn, F.; Moulijn, J.
A. Eurokin. Chemical reaction kinetics in practice. Cattech 2001, 5, 30.
(32) Marquardt, D. W. An algorithm for least-squares estimation of
nonlinear parameters. J. Soc. Ind. Appl. Math. 1963, 11, 431.
(33) Rosenbrock, H. H. An automatic method for finding the greatest
or least value of a function. Comput. J. 1960, 3, 175.
(34) Nelib. http://netlib.org (accessed Jan. 14, 2013).
(35) Chang, C. D. Hydrocarbons from methanol. Catal. Rev. 1983,
25, 1.
(36) Svelle, S.; Visur, M.; Olsbye, U. Saepurahman; Bjorgen, M.
Mechanistic aspects of the zeolite catalyzed methylation of alkenes and
aromatics with methanol: A review. Top. Catal. 2011, 54, 897.
(37) Ivanova, I. I.; Corma, A. Surface species formed and their
reactivity during the alkylation of toluene by methanol and dimethyl
ether on zeolites as determined by in situ C-13 MAS NMR. J. Phys.
Chem. B 1997, 101, 547.
(38) Vos, A. M.; Nulens, K. H. L.; De Proft, F.; Schoonheydt, R. A.;
Geerlings, P. Reactivity descriptors and rate constants for electrophilic
aromatic substitution: Acid zeolite catalyzed methylation of benzene
and toluene. J. Phys. Chem. B 2002, 106, 2026.
(39) Svelle, S.; Kolboe, S.; Olsbye, U.; Swang, O. A theoretical
investigation of the methylation of methylbenzenes and alkenes by
halomethanes over acidic zeolites. J. Phys. Chem. B 2003, 107, 5251.
(40) Ono, Y.; Mori, T. Mechanism of methanol conversion into
hydrocarbons over ZSM-5 zeolite. J. Chem. Soc. Farad. T. 1 1981, 77,
2209.
(41) Forester, T. R.; Howe, R. F. In situ FTIR studies of methanol
and dimethyl ether in ZSM-5. J. Am. Chem. Soc. 1987, 109, 5076.
(42) Forester, T. R.; Wong, S. T.; Howe, R. F. In situ Fouriertransform IR observation of methylating species in ZSM-5. J. Chem.
Soc. Chem. Comm. 1986, 1611.
(43) Ivanova, I. I.; Pomakhina, E. B.; Rebrov, A. I.; Hunger, M.;
Kolyagin, Y. G.; Weitkamp, J. Surface species formed during aniline
methylation on zeolite H-Y investigated by in situ MAS NMR
spectroscopy. J. Catal. 2001, 203, 375.

(44) Anderson, M. W.; Barrie, P. J.; Klinowski, J. H-1 magic-anglespinning NMR-studies of the adsorption of alcohols on molecularsieve catalysts. J. Phys. Chem.-US 1991, 95, 235.
(45) Blaszkowski, S. R.; van Santen, R. A. Theoretical study of CC
bond formation in the methanol-to-gasoline process. J. Am. Chem. Soc.
1997, 119, 5020.
(46) Song, W. G.; Marcus, D. M.; Fu, H.; Ehresmann, J. O.; Haw, J.
F. An oft-studied reaction that may never have been: Direct catalytic
conversion of methanol or dimethyl ether to hydrocarbons on the
solid acids H-ZSM-5 or HSAPO-34. J. Am. Chem. Soc. 2002, 124,
3844.
(47) Wang, W.; Hunger, M. Reactivity of surface alkoxy species on
acidic zeolite catalysts. Acc. Chem. Res. 2008, 41, 895.
(48) Sullivan, R. F.; Sieg, R. P.; Langlois, G. E.; Egan, C. J. A new
reaction that occurs in hydrocracking of certain aromatic hydrocarbons. J. Am. Chem. Soc. 1961, 83, 1156.
(49) Sassi, A.; Wildman, M. A.; Ahn, H. J.; Prasad, P.; Nicholas, J. B.;
Haw, J. F. Methylbenzene chemistry on zeolite HBeta: Multiple
insights into methanol-to-olefin catalysis. J. Phys. Chem. B 2002, 106,
2294.
(50) McCann, D. M.; Lesthaeghe, D.; Kletnieks, P. W.; Guenther, D.
R.; Hayman, M. J.; Van Speybroeck, V.; Waroquier, M.; Haw, J. F. A
complete catalytic cycle for supramolecular methanol-to-olefins
conversion by linking theory with experiment. Angew. Chem., Int. Ed.
2008, 47, 5179.
(51) Lesthaeghe, D.; Horre, A.; Waroquier, M.; Marin, G. B.; Van
Speybroeck, V. Theoretical insights on methylbenzene side-chain
growth in ZSM-5 zeolites for methanol-to-olefin conversion. Chem.
Eur. J. 2009, 15, 10803.
(52) Arstad, B.; Nicholas, J. B.; Haw, J. F. Theoretical study of the
methylbenzene side-chain hydrocarbon pool mechanism in methanol
to olefin catalysis. J. Am. Chem. Soc. 2004, 126, 2991.
(53) Dessau, R. M. On the H-ZSM-5 catalyzed formation of ethylene
from methanol or higher olefins. J. Catal. 1986, 99, 111.
(54) Haw, J. F.; Nicholas, J. B.; Xu, T.; Beck, L. W.; Ferguson, D. B.
Physical organic chemistry of solid acids: Lessons from in situ NMR
and theoretical chemistry. Acc. Chem. Res. 1996, 29, 259.
(55) Kazansky, V. B.; Frash, M. V.; van Santen, R. A. Quantumchemical study of the isobutane cracking on zeolites. Appl .Catal. AGen. 1996, 146, 225.
(56) Quann, R. J.; Green, L. A.; Tabak, S. A.; Krambeck, F. J.
Chemistry of olefin oligomerization over ZSM-5 catalyst. Ind. Eng.
Chem. Res. 1988, 27, 565.
(57) Bjrgen, M.; Svelle, S.; Joensen, F.; Nerlov, J.; Kolboe, S.;
Bonino, F.; Palumbo, L.; Bordiga, S.; Olsbye, U. Conversion of
methanol to hydrocarbons over zeolite H-ZSM-5: On the origin of the
olefinic species. J. Catal. 2007, 249, 195.
(58) Froment, G. F. Modeling of the kinetics of complex processes
based on elementary steps. Abstr. Pap. Am. Chem. Soc. 1991, 201, 26.
(59) Eyring, H. The activated complex and the absolute rate of
chemical reactions. Chem. Rev. 1935, 17.
(60) Baltanas, M. A.; Vanraemdonck, K. K.; Froment, G. F.;
Mohedas, S. R. Fundamental kinetic modeling of hydroisomerization
and hydrocracking on noble-metal-loaded faujasites. 1. Rate
parameters for hydroisomerization. Ind. Eng. Chem. Res. 1989, 28, 899.
(61) Martinis, J. M.; Froment, G. F. Alkylation on solid acids. Part 2.
Single-event kinetic modeling. Ind. Eng. Chem. Res. 2006, 45, 954.
(62) Thybaut, J. W.; Marin, G. B.; Baron, G. V.; Jacobs, P. A.;
Martens, J. A. Alkene protonation enthalpy determination from
fundamental kinetic modeling of alkane hydroconversion on Pt/H(US)Y-zeolite. J. Catal. 2001, 202, 324.
(63) Denayer, J. F.; Souverijns, W.; Jacobs, P. A.; Martens, J. A.;
Baron, G. V. High-temperature low-pressure adsorption of branched
C-5-C-8 alkanes on zeolite beta, ZSM-5, ZSM-22, zeolite Y, and
mordenite. J. Phys. Chem. B 1998, 102, 4588.
(64) Macdonald, F.; Lide, D. R. CRC handbook of chemistry and
physics: From paper to web. Abstr. Pap. Am. Chem. Soc. 2003, 225.
1506

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Industrial & Engineering Chemistry Research

Article

(65) Benson, S. W.; Cruicksh, Fr; Golden, D. M.; Haugen, G. R.;


Oneal, H. E.; Rodgers, A. S.; Shaw, R.; Walsh, R. Additivity rules for
estimation of thermochemical properties. Chem. Rev. 1969, 69, 279.
(66) Frisch, A.; Frisch, M. J.; Trucks, G. W. Gaussian 03 Users
Reference; Gaussian: Carnagie, PA, 2003.
(67) Dumesic, J. A. The Microkinetics of Heterogeneous Catalysis;
American Chemical Society: Washington, DC, 1993.
(68) Quintana-Solorzano, R.; Thybaut, J. W.; Marin, G. B. A singleevent microkinetic analysis of the catalytic cracking of (cyclo)alkanes
on an equilibrium catalyst in the absence of coke formation. Chem.
Eng. Sci. 2007, 62, 5033.
(69) Boudart, M. Physical limitations to values of parameters used in
rate equations for reactions catalyzed by solids. Ind. Chim. Belg. 1966,
31, P74.
(70) Svelle, S.; Ronning, P. A.; Kolboe, S. Kinetic studies of zeolitecatalyzed methylation reactions 1. Coreaction of [C-12]ethene and
[C-13]methanol. J. Catal. 2004, 224, 115.
(71) Thybaut, J. W.; Choudhury, I. R.; Martens, J. A.; Denayer, J. F.
M.; Marin, G. B. Single event microkinetic modeling of n-hexane
hydroisomerization on Pt/H-ZSM-5. Annual AIChE Meeting 2009,
Nashville, Tennessee, USA, November 813, 2009.
(72) Svelle, S.; Ronning, P. O.; Olsbye, U.; Kolboe, S. Kinetic studies
of zeolite-catalyzed methylation reactions. Part 2. Co-reaction of [C12]propene or [C-12]n-butene and [C-13]methanol. J. Catal. 2005,
234, 385.
(73) Van Speybroeck, V.; Van der Mynsbrugge, J.; Vandichel, M.;
Hemelsoet, K.; Lesthaeghe, D.; Ghysels, A.; Marin, G. B.; Waroquier,
M. First principle kinetic studies of zeolite-catalyzed methylation
reactions. J. Am. Chem. Soc. 2011, 133, 888.

1507

dx.doi.org/10.1021/ie301542c | Ind. Eng. Chem. Res. 2013, 52, 14911507

Das könnte Ihnen auch gefallen