Sie sind auf Seite 1von 13

Ind. Eng. Chem. Res.

2009, 48, 1107–1119 1107

Fundamental Kinetic Modeling of Catalytic Reforming


Rogelio Sotelo-Boyás† and Gilbert F. Froment*
Artie McFerrin Department of Chemical Engineering, Texas A&M UniVersity, College Station, Texas 77843-3122

A fundamental kinetic model for the catalytic reforming process has been developed. The complex network
of elementary steps and molecular reactions occurring in catalytic reforming was generated through a computer
algorithm characterizing the various species by means of vectors and Boolean relation matrices. The algorithm
is based on the fundamental chemistry occurring on both acid and metal sites of a Pt-Sn/Al2O3 catalyst. The
number of rate coefficients for the transformations occurring on the metal sites was reduced by relating them
to the nature of the involved carbon atoms. The single event concept was applied in the development of rate
expressions for the elementary steps on the acid sites. This approach allows obtaining rate coefficients that
are independent of the feedstock, owing to their fundamental chemical nature. The Levenberg-Marquardt
algorithm was used to estimate the rate coefficients. The estimation was based on data reported from a previous
naphtha reforming study in a fixed bed reactor with Pt-Sn/Al2O3 as a catalyst. The agreement between the
experimental and estimated yields is excellent. The statistical tests were also satisfied. The kinetic model was
used in pseudo-homogeneous and heterogeneous reactor models simulating an industrial three-bed adiabatic
catalytic reformer with centripetal radial flow.

1. Introduction on the basis of a new experimental program. Quann and Jaffe10


have addressed this problem by developing structure oriented
Recent environmental legislations established by the Clean
lumping. A large number of key molecules are generated by
Air Act1 and the Federal Reformulated Gasoline Program
assembling 22 structural groups in various ways. This synthetic
demand the reduction in emissions of volatile, toxic, and
feed has to satisfy the observable characteristics, both chemical
polluting components in gasoline, such as sulfur, benzene, and
and physical. Klein and co-workers11 reduced the complexity
olefins. One of the most affected refining units by this
of the feed by introducing a number of representative pseudocom-
reformulation is the catalytic reformer. The main objective of
ponents by Monte-Carlo simulation and generated the reaction
catalytic reforming is the transformation of low-octane straight
network of this synthetic feed by computer using graph theory.
run naphtha consisting of C5 to C12 hydrocarbons with a research
The approach was applied to catalytic reforming.12 The draw-
octane number of 50-60 into gasoline with an octane number
back of the two approaches mentioned here is the use of
of 90-105. This is achieved by transforming paraffins and
lumpssbe they called pseudo- or key-componentssthat inevi-
naphthenes into the corresponding isoparaffins and aromatics.
tably lead to rate coefficients depending on the feedstock
The reformer contributes for about 50 vol % to the gasoline
composition.
pool in a refinery. The process is typically carried out in three
In the present paper the kinetic modeling is based upon a
or four adiabatic catalytic beds in which reactions such as
detailed description of the fundamental chemistry of the
dehydrogenation, isomerization, hydrocracking, dehydrocycliza-
transformation of each individual hydrocarbon. The rates of the
tion, hydrogenolysis, and coking are taking place. It is carried
reactions on the metal sites are systematized by relating them
out over bifunctional catalysts (metal/acid) with platinum alone
to the nature of the C-atoms linked to a double bond in the
or platinum combined with rhenium, iridium, or tin, dispersed
product or to the single bond to be broken in the reactant. The
on acidic alumina. The Pt-Sn/Al2O3 catalyst used in the present
rates of the transformations on the acid sites are described in
study is less subject to coking than Pt- and Pt-Re catalysts
terms of the elementary steps of carbenium ion chemistry. By
and permits operation at a lower pressure, which is beneficial
doing so the number of elementary steps definitely becomes
for the yield of aromatics and, consequently, for the octane
huge, but they belong to a limited number of types. It is possible
number.
to formulate their rates using only a reasonable number of
Advanced optimization of such a complex process requires
independent kinetic parameters by applying the single event
a detailed mathematical model capable of accurately predicting
concept introduced by Froment et al.13-19 The single-event
the reformate composition, the product quality, and the catalyst
concept has been applied previously to the kinetic modeling of
life cycle over a wide range of operating conditions. The
thermal cracking,14,20 catalytic cracking,17,21,22 hydrocracking,23
catalytic reforming process has been modeled until now by
alkylation,24 hydroisomerization,13,25,26 and the methanol-to-
lumping the spectrum of naphtha components ((200 hydro-
olefins27,28 process. The single-event rate parameters are truly
carbons at a reasonable concentration level) into a small number
invariant with respect to the feedstock composition, and the
of pseudocomponents.2-9 With such an approach the kinetic
prediction of the reformate composition is much more accurate
modeling is simple and the estimation of the kinetic parameters
and reliable, even for a wide range of operating conditions.
is a relatively easy task. The drawback is that these parameters
depend on the composition of the feedstock, so that for a
2. Chemistry of the Process
different feed the kinetic parameters have to be re-estimated
Catalytic reforming requires a bifunctional catalyst containing
* To whom correspondence should be addressed. E-mail:
gilbert.froment@skynet.be. a metal and an acid function.29 The metal function is responsible

Present address: Instituto Politécnico Nacional, ESIQIE, México for the dehydrogenation of paraffins and naphthenes, the
D.F. 07738, México. hydrogenation of olefins and aromatics, and the hydrogenolysis
10.1021/ie800607e CCC: $40.75  2009 American Chemical Society
Published on Web 09/10/2008
1108 Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009

of paraffins, that is, the demethylation (DEM) and deethylation


(DET) producing methane and ethane. The transformations on
the metal sites are considered at a molecular level; the molecule
is considered as a whole.
The transformations on the acid alumina sites are described
in terms of elementary steps of carbenium ion chemistry. The
paraffins are first dehydrogenated on the metal sites into olefins,
which are then protonated on the acid sites of the catalyst to
produce carbenium ions. These isomerize through hydride shift
(HS), methyl and ethyl shift (MS and ES) and branching (PCP-
isomerization), but also undergo cyclization and β-scission. The
ions thus generated (also the cyclic ones) are deprotonated into
the corresponding olefins, which are then hydrogenated on the
metal sites, thus producing new paraffins and naphthenes.
The reforming of naphthenes involves dehydrogenation,
isomerization, and cracking steps. First, the naphthenes are
dehydrogenated into cyclic olefins, which are protonated into
cyclic carbenium ions. Four types of isomerization steps were
considered for these cyclic ions. The first is an intraring alkyl
shift (IRAS) and leads to ring contraction or ring expansion
without altering the degree of branching. The ring expansion is
typical for naphthene isomerization. The conversion of methyl
cyclopentane (MCP) into benzene, for example, first proceeds
through ring expansion into cyclohexane, which is then dehy-
drogenated into benzene. The second type is the formation of a
protonated cyclic propane ionsa cyclic PCP stepsleading to
ring contraction or ring expansion while altering the degree of
branching. The third type is an IRAS in which the relative
positions of the substituents on the ring are changed without
altering the degree of branching. The fourth is an acyclic PCP
isomerization, in the alkyl side chain. The cracking of the cyclic
ions proceeds through endo- and exocyclic β-scissions. Hydro-
genolysis on the side chain of cycloalkanes is also accounted
for, but not their ring opening on the metal sites. This reaction
is difficult under reforming conditions and does not significantly
affect the product distribution.30
The aromatics are produced by dehydrocyclization of paraffins
or by dehydrogenation of naphthenes. As in the dehydroisomer- Figure 1. Molecular reactions and elementary step of carbenium ions in
the catalytic reforming of paraffins.
ization of naphthenes, there are intermediate steps between the
paraffin and the aromatic, involving both metal and acid sites
of the catalyst. Direct dehydrocyclization of paraffins into 3. Computer Algorithm for the Generation of the
aromatics is not included in this work. Instead, as shown in Reaction Network
Figure 1, the paraffins dehydrogenate into olefins which undergo The network of molecular reactions and elementary steps for
an additional dehydrogenation producing diolefins. These are paraffins, naphthenes, aromatics, and all types of olefins and
protonated into olefinic carbenium ions that are transformed into carbenium ions was generated by computer. Some simplifica-
cyclic carbenium ions and subsequently into naphthenes, which tions and rules were imposed:
are dehydrogenated to produce aromatics. For cyclization to 1. The maximum carbon number in the network is C12.
occur, a paraffin with at least a six-carbon straight chain is 2. Methyl and primary carbenium ions are very unstable so
needed. that their contribution to the process can be neglected.
Alkyl aromatics may undergo dehydrogenation and hy- 3. The acid-active sites of the catalyst are of the Brönsted
drodealkylation in their side chain to produce aromatic olefins type.
and smaller aromatics, respectively. A previous study31 has 4. There is no direct transformation of paraffins into aromatics.
shown that dealkylation of propylbenzene can proceed on both 5. Hydrogenolysis of naphthenes and aromatics occurs only
acid and metal sites. In this work hydrodealkylation of aromatics on the side chain.
follows a bifunctional mechanism. On the acid sites, the 6. Bimolecular reactions do not occur.
dealkylation of aromatics occurs via exocyclic β-scission of 7. Dicyclic components are not generated.
aromatic carbenium ions. Although a naphtha feedstock may 8. The maximum number of double bonds in a molecule is
contain hydrocarbons up to C12, dicyclic components were not two. Only conjugated diolefins are considered.
included in the network, based upon conclusions by Davis30 Starting from these rules the network involving all possible
according to whom the use of a Pt/Sn catalyst drastically reduces transformations was generated by means of a computer program
the extent of a second cyclization producing a dicyclic aromatic. following the sequence described in Figure 4. The program is
Figures 1, 2, and 3 illustrate molecular reactions and based on the algorithm developed by Clymans and Froment14
elementary steps in the reforming of paraffins, naphthenes, and and Baltanas and Froment.13,26 Boolean relation matrices are
aromatics, respectively. used to describe the structure of each species and generate the
Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009 1109
4. Single Event Kinetic Modeling
Figure 5 illustrates that the catalytic reforming of naphtha
involves thousands of species and thousands of individual elemen-
tary steps. Fortunately, in catalytic reforming as in other acid-
catalyzed hydrocarbon processes, the elementary steps belong to
only a few types: protonation, deprotonation, alkyl shifts, branching
reactions, cyclization, and β-scission. Nevertheless, the influence
of the structure of the members of a given homologous series
undergoing a certain type of elementary step still leads to a huge
number of parameters. The only way to reduce the number of
rate coefficients is to model them. Froment et al.15,26 have made
use of the transition state theory for this purpose:

k)
k BT
h
exp ( ) (
∆Ŝo‡
R
exp -
∆Ho‡
RT ) (1)

The effect of structure on the rate coefficient is related to the


ratio of the global symmetry numbers of the reactant and
r ‡
activated complex, σgl and σgl . This ratio, called the number of
single events, ne, is factored out of the rate coefficient

k)
( )(
σrgl kBT
σ‡
gl
h ) ( ) (
exp
∆Ŝo‡
R
exp -
∆Ho‡
RT ) (2)

The remaining part of k is called the single event rate coefficient


and is represented by k̃. The rate coefficient of the elementary
step, k, is a multiple of k̃:

k ) nek̃ (3)
Figure 2. Molecular reactions and elementary steps of carbenium ions in Equation 2 also reveals that it is possible to define a “single
the catalytic reforming of naphthenes. event” frequency factor that does not depend upon the structure
of the reactant and activated complex and is unique for a given
type of elementary step:19

à )
kBT
h
exp ( )
∆Ŝo‡
R
(4)

By factoring out the number of single events from the rate


coefficient or frequency factor of an elementary step, the effect
of structure on the change of entropy is explicited. To calculate
the number of single events, it is necessary to know the
configuration of the reactant and the activated complex.
Quantum chemical packages (Gaussian, Gamess, Mopac, etc.)
yield reliable results for this. Because of the extremely large
number of intermediates the determination of the transition
structure for each of them is an overwhelming task. Therefore,
on the basis of previous experience in quantum chemical
Figure 3. Molecular reactions and elementary steps of carbenium ions in
the catalytic reforming of aromatics. calculations, formal rules have been established for the deter-
mination of the single event numbers.13
Recently, in their work on the hydrocracking of a mixture of
molecular reactions and elementary steps. Although the Boolean paraffins, Kumar and Froment32 used the linear free-energy-
matrices are rather sparse they occupy a lot of memory. type relationship of Evans-Polanyi to obtain the activation
Therefore, they are stored in the form of vectors and recon- energy of each elementary step belonging to a given type,
structed from these to generate a transformation of a molecule starting from two parameters related to a reference step for that
or ion in the network. type. For a very complex mixture like vacuum gas oil, however,
they considered only four activation energies per type, defined
In the reforming of C10 hydrocarbons, the total number of by the nature (secondary, tertiary) of both the reacting and
components (P, O, N, A) generated amounts to 4748 and the produced carbenium ions and independent of the chain length
total number of carbenium ions to 6202 (Table 1). Figure 5 and other structural aspects.32 Calculations of Park and Fro-
shows how the number of components and carbenium ions ment33 about methylation, oligomerization, and β-scission in
increases with the carbon number for three types of hydrocar- the methanol to olefins process (MTO) indicated this to be the
bons. For C11 hydrocarbons the number of components and ions main determining factor for E. As a consequence and given the
is much higher than for C10 hydrocarbons so that the time unique value of à only a maximum of four single event rate
required for generating the network for C11 hydrocarbons is coefficients have to be considered per type of elementary step:
about three times that for C10. k̃(s,s), k̃(s,t), k̃(t,s), and k̃(t,t).
1110 Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009

Figure 4. General sequence followed in the network generation for naphtha reforming: (---) (de)hydrogenation reactions on metal sites; (s) elementary steps
on acid sites.

Table 1. Number of Molecular Components and Carbenium Ions in 5. Rate Equations for the Elementary Steps on the Acid
the Network for C10 Hydrocarbons Reforming Sites
component no.
5.1. Derivation. Rates have been formulated for the elemen-
Hydrocarbons tary steps of paraffins, naphthenes, aromatics, and olefins. The
paraffins 125 olefins are adsorbed on the acid sites and are converted after
naphthenes 165 protonation into carbenium ions. Taking into account the steps
aromatics 36
in which the olefins participate (protonation/deprotonation and
olefins 584
diolefins 1085 cracking), their net rates of formation on the acid sites can be
cyclic olefins 963 written as

∑ (n ) ∑ (n )
cyclic diolefins 1757
aromatic olefins 33 ROij ) +-
e depk̃dep(mik, Oij)CRik e prk̃pr(mik)CH+POij +
total HCs: 4748 k k

Carbenium Ions ∑ ∑ (n ) e crk̃cr(mlo, mik, Oij)CRlo


+(5)

paraffinic carbenium ions 454 l o


cyclic carbenium ions 801 The concentrations of the carbenium ions on the surface of the
aromatic carbenium ions 32 catalyst are not directly accessible, but by applying the pseu-
olefinic carbenium ions 1835
dosteady state approximation the net rate of formation of
cyclic olefinic carbenium ions 3080
total ions: 6202 carbenium ions can be written as
RRik+ ) ∑ (n )
e prk̃pr(mik)CH+POij - ∑ (n ) +-
e depk̃dep(mik, Oij)CRik
j j

∑ ∑ (n ) ++
e isomk̃isom(mik, mlo))CRik
l o

∑ ∑ (n ) +-
e isomk̃isom(mlo, mik)CRlo
l o

∑ ∑ (n ) ++
e crk̃cr(mik, mlo, OuV)CRik
l o

∑ ∑ (n ) +)0
e crk̃cr(mlo, mik, OuV)CRlo (6)
l o
The isomerization term involves all types of isomerizations that
the acyclic carbenium ions can undergo: ethyl-, methyl-, hydride
shift, and PCP-branching.
(De)protonation steps are known to be extremely rapid and
to reach equilibrium.16 Equating the rates of deprotonation and
protonation and solving for CRik+ leads to:

∑ (n ) *
e prk̃pr(mik)CH+POij
j
C R* ik+ ) (7)
∑ (n ) e depk̃dep(mik, Oij)
Figure 5. Number of paraffins, naphthenes and aromatics in the network j
up to C11 hydrocarbons. with
Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009 1111

*
CR* i+ )
CRi+
Ct
and CH* + )
CH+

CH+ represents the concentrations of free acid sites relative to


Ct
(8) CR*′1+ )
( CR* 1+

K̃pr(Oref, m) ) ∑(
)
j
(ne)pr
)
K̃ (O , O )C* P
(ne)dep isom j ref H+ Oj
(14)
the total concentration of acid sites. They relate to CR* i+ by the
following equation: In the rate equations the single-event rate coefficients, k̃, are
replaced by the corresponding composite rate coefficients, k̃*:
CH* + + ∑∑C *
+)1
Rik (9)
k̃* ) CtK̃pr(Oref, m)k̃ (15)
i k

The use of relative concentrations avoids the estimation of the These composite single event parameters are then used in the
total acid sites concentration, Ct. final formulation of the rates of the elementary steps on the
5.2. Thermodynamic Constraints. The single event rate acid sites. For example, the rate equation for the elementary
coefficients for deprotonation depend on the nature of the step PCP-isomerization of a secondary carbenium ion R+1 (s) into
carbenium ion and of the olefin. This leads to an overwhelming another secondary carbenium R2+(s) is written
number of deprotonation coefficients. Baltanas et al.26 were able
to reduce the number of required deprotonation parameters by Rpcp(s, s) ) (ne)pcpCtk̃*pcp(s, s)CR*′1+(s) (16)
using the following thermodynamic constraint:
The rate coefficients are considered to depend only on the nature
k̃dep(m, Oref) of the reacting and produced carbenium ions and not on the
K̃isom(Oj, Oref) ) (10) produced olefin.15 The composite parameters are the ones that
k̃dep(m, Oj) are estimated from experimental data.
or 5.4. Accounting for Incomplete Analysis of Feed and
Products. For a complex feedstock like the naphtha used in
k̃dep(m, Oref) catalytic reforming, a gas chromatographic analysis that quanti-
k̃dep(m, Oj) ) (11) fies all the individual isomers is not available. The process model
K̃isom(Oj, Oref) has to account for this. That is why groups of isomers (GOI) of
where Oref is a reference olefin, and K̃isom(Oj,Oref) is the homologous hydrocarbons with a same degree of branching were
equilibrium constant for the isomerization of the olefin j into introduced. This also reduces the number of continuity equations
the reference olefin. This leads to the selection of one reference to be integrated along the reactor in the kinetic analysis of
olefin per carbon-number so that only two independent depro- integral reactor data and in design simulations. If the composi-
tonation rate coefficients per carbon number are left, k̃dep(s,Oj) tion of the GOI of the feedstock is unknown, it is safe to assume
and k̃dep(t,Oj). These two coefficients are then included by means that the isomers are in equilibrium. Their evolution through the
of a reference protonation equilibrium constant, K̃pr. For reactor is, of course, accounted for through the network
instance, for a secondary (s) carbenium ion R+ 1 that deprotonates generation and kinetic equations.
into an olefin Oj, the concentration is given by K̃isom(Oj,Oref): Paraffins were subdivided into normal paraffins (nP), single-
branched paraffins (SBP), double-branched paraffins (DBP), and
∑ (n ) *
e prk̃pr(s)CH+POj
triple-branched paraffins (TBP). The same was done with the
olefins. Naphthenes were subdivided into cyclopentanes (5Ni)
j
CR* 1+ ) (12)
∑ (n ) e depk̃dep(s, Oj)
and cyclohexanes (6Ni), while aromatics were grouped accord-
ing to their number of carbon atoms. A naphtha containing
j
hydrocarbons up to C10 then contains 27 paraffinic, 25 olefinic,
Equation 12 is further simplified by introducing the thermody- 11 naphthenic, and 5 aromatic components and GOI. Including
namic constraint for the deprotonation step (eq 11): hydrogen, the kinetic model contains 69 components and GOI.

( )(ne)pr The rate coefficients in such a model can be constructed from


CR* 1+ ) ∑ K̃ (O , s)K̃isom(Oj, Oref)CH* +POj
(ne)dep pr ref
(13) the single event rate coefficients, accounting for the full network
j of elementary steps. It should be stressed that for the rate of
Equation 13 is the concentration of carbenium ions produced
by protonation of olefins. By application of the pseudosteady
state approximation to the net rates of formation for each type
of ion, similar equations can be obtained for calculating the
concentration of other types of carbenium ions, such as those
produced from diolefins, cyclic olefins, and aromatic olefins.
These equations, together with eq 9, form a set of nonlinear
equations that can be easily solved to obtain CR* ik+ and CH+*
.
5.3. Composite Single Event Parameters. The substitution
of the relative concentrations of the carbenium ions into eq 5
requires the inclusion of the total concentration of active acid
sites, Ct, into the single-event rate coefficients k̃ so that new
single-event rate coefficients have to be defined. Moreover,
because of the equilibrium of the (de)protonation steps, the
single-event rate coefficients and the single-event protonation
equilibrium constant cannot be determined separately, so that
they appear in a composite parameter. For these reasons, the
relative concentrations of the carbenium ions are replaced by
the relative concentrations divided by K̃pr, that is: Figure 6. LC for PCP(s,s) isomerization of normal olefins at 733 K.
1112 Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009

transformation of GOI1 into GOI2, all possible transformations Scheme 1


of the components of GOI1 into the components of GOI2 are
considered. Hence, the reaction rate is the sum of the rates of
all the elementary steps that convert all the carbenium ions and
molecules of GOI1 into carbenium ions and molecules of GOI2.
For instance, the rate of reaction for the PCP isomerization of
GOI1 into GOI2 can be written as

RpcpL TL )
1 2
∑ (n ) * *′
+,m -
e pcpk̃pcp(m1, m2)CRik 1
k

∑ (n ) * *′
e pcpk̃pcp(m2, m1)CRik
+,m (17)
2
k

The relative carbenium ion concentrations are calculated as


described by eq 14. Substitution into the first term of eq 17 and the olefin. The mole fractions of the components of the
leads to GOI are determined from ratios of isomerization equilibrium
constants calculated by the group contribution method of Benson
RpcpL fL ) ((LC)pcp(s, s)k̃*pcp(s, s) + (LC)pcp(s, t)k̃*pcp(s, t) + et al.34
1 2
6.2. Hydrogenolysis of Paraffins: Demethylation and
(LC)pcp(t, s)k̃*pcp(t, s) + (LC)pcp(t, t)k̃*pcp(t, t))CH* +POj(18) De-ethylation. The hydrogenolysis of paraffins on the metal
sites involves a C-C bond scission with formation of methane
where POj are the partial pressures of the olefinic GOIj and LC and ethane. Several authors36-40 have studied the hydrogenolysis
are the so-called lumping coefficients. They are defined by the of paraffins. A generally accepted mechanism for hydrogenolysis
following equation: on metals assumes that a paraffin must lose several H atoms

( )
(ne)pr prior to the C-C bond scission. That requires the formation of
(LC)pcp(m1, m2) ) ∑ n y K̃ (O , O ) (19)
(ne)dep epcp i,Lj isom j ref
multiple carbonsmetal bonds.41,42
m1,m2
H2 chemisorption: KH
where yi,Lj is the mole fraction of the component i in the GOIj. H2 + 2L h 2HL (I)
5.5. Calculation of the LC-Coefficients. The LC are
functions of temperature only and do not depend on the paraffin dehydrogenation: Kp
feedstock composition. Their values, calculated for temper- CmHn + (n + 1 - x)L h CmHxL + (n - x)HL (II)
ature intervals of 5 K and stored in the computer memory,
scission of the dehydrogenated species (rds) k1
are utilized in the integration of the ODEs in the adiabatic
CmHxL + HL f CpHyL + Cm-pHx-y+1L (III)
reactor simulations. The LC do not contain the rate coef-
ficient. They vary with the number of carbon atoms and hydrogenation and desorption of products: kj
slightly with temperature. To calculate the equilibrium CpHyL + (2p + 2 - y)HL f CpH2p+2 + (2p + 3 - y)L (IV)
constants that appear in the LC requires the thermodynamic
properties cp, Hf, and Sf of the saturated and unsaturated where n ) m + 2. It is the number of hydrogen atoms in the
species. These properties were calculated using group reacting paraffin; x is the number of H atoms remaining on the
contribution methods developed by Benson et al.34 hydrocarbon species; x < n + 1, y < x + 1.
Figure 6 shows the LC for PCP(s,s) isomerization of n-olefins With step III as the rate determining step, the rate equation
at 733 K as a function of the carbon number. The increase in takes the form
the value of LC with respect to the carbon number results from
k1Kppp(KHpH)(n+1-x) ⁄2
the increase in the number of isomerization steps with increasing rhls ) (21)
carbon number. (Kppp + (KHpH)(n-x) ⁄2 + (KHpH)(n+1-x) ⁄2)2
From the four parameters (k1, Kp, KH, and x), the most
6. Rate Expressions for Reactions on Metal Sites significant is Kp, which is the one that controls the concentra-
The main types of molecular reactions that occur on the metal tion of the reacting paraffin. Because the dehydrogenation
sites under reforming conditions are dehydrogenation of paraf- is endothermic Kp increases with temperature. Bond et al.42-44
fins and naphthenes and hydrogenolysis of paraffins, as well as have shown that the value of x does not significantly affect
the reversible reactions involved, that is, hydrogenations of the rate. A similar behavior was found in this work and
olefins, cyclic olefins, and aromatics. therefore a constant value for each paraffinic GOI was set
6.1. Dehydrogenation of Paraffins and Hydrogenation for x. This implies that whatever the paraffin and hydrogen
of Olefins. In the present work, a dehydrogenation mechanism pressures there is only one reactive intermediate for each
similar to that proposed by Dumez and Froment35 was used paraffinic GOI.
(Scheme 1). 6.3. Dehydrogenation of Alkyl Cyclohexanes and Hydro-
When the surface reaction is the rate determining step (rds), genation of Aromatics. The dehydrogenation of cyclohexanes
the following rate equation is derived: has been extensively studied45-48 over different types of catalyst.
Van Trimpont et al.48 studied this reaction on a Pt-Re/Al2O3

rP-O )
(
kP-OK1 · pP -
pOpH
Keq(P-O) ) (20)
catalyst. They found that a model based on a single site
mechanism reproduced the observed data better than a model
with a dual site mechanism. Verstraete and Froment47 have also
(1 + K1pP + K3pO + K4pH)2
studied this reaction over a Pt-Sn/Al2O3 catalyst. After
The rate coefficient kP-O and the adsorption constants K1, K3, discrimination between three different mechanisms they retained
K4 depend on the nature and molecular weight of the paraffin a single site mechanism:
Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009 1113
N + L h NL 7. Reduction of the Number of Parameters
NL h CNL + H2
The formulation of the rate equations for the elementary steps
CNL h CDNL + H2 (rds) and molecular reactions leads to a kinetic model containing for
CDNL h AL + H2 isothermal conditions 56 parameters, belonging to the types
AL h A + L listed in Table 2.
where CNL ) cyclic olefin, CDNL ) cyclic diolefin. When In Table 2 the parameters for the elementary steps on the
the surface reaction is the rate determining step the rate equation acid sites are the true single-event rate parameters. The use of
becomes the composite parameters, together with the application of
thermodynamic constraints, allows a reduction in the number
kN-A · (pN - pApH3 ⁄ KN-A) of parameters. The thermodynamic constraint is applied to the
rN-A ) (22) isomerization steps, as first derived by Baltanas et al.26 in terms
pH(1 + KNpN + KApA)
of the single-event protonation equilibrium constants:
The equilibrium constant for the dehydrogenation, KN-A, is +
(OrefTRs )
calculated from Benson’s group contribution method. k̃isom(t, s) K̃pr
6.4. Modeling of the Metal Site Parameters. 6.4.1. Kinetic ) +
(25)
(OrefTRt )
k̃isom(s, t) K̃pr
Parameters for the Dehydrogenation of Paraffins. The kinetic
parameters of equations (20) are valid for the dehydrogenation By substituting one of the rate coefficients into eq 25, it is readily
of a particular paraffin. The large number of paraffins in naphtha found that
leads to a very large number of rate parameters. Only a
systematic approach can reduce that number. The energy
*
k̃isom(t, s) ) k̃isom
*
(s, t) (26)
required to remove a hydrogen atom depends on the nature of Because the thermodynamic constraint applies to each type of
the parent carbon atom: primary, secondary, or tertiary.49 The isomerization, the number of parameters for PCP-branching,
activation energies for the dehydrogenation of paraffins then intraring alkyl shift, and ring contraction/expansion is reduced
depend on the location of the double bond in the product olefin, from four to three for each of them. A second reduction stems
that is, on the nature of the carbon atoms linked to the double from the similarity between acyclic β-scission and exocyclic
bond.32 As a result, all the rate coefficients, kP-O in eq 21, belong β-scission on the side chain.
to one of the following five types: kdeh_p(p-s), kdeh_p(p-t), Further reduction is possible by observing that cyclization is
kdeh_p(s-s), kdeh_p(s-t) and kdeh_p(t-t). The first one, for example, the reverse of endocyclic β-scission, so that only one set of
indicates that the dehydrogenation of a given paraffin produces parameters and the thermodynamic equilibrium constant are
an olefin with a double bond between primary and secondary required. The cyclization of 1-heptene into methyl cyclohexane
carbon atoms. is given here as an example:
6.4.2. Kinetic Parameters for the Hydrogenolysis of
Paraffins. In this case no olefin is formed so that the nature of
the C-atoms linked to the bond where the scission occurs is
taken to be determining. Consequently, for both types of
hydrogenolysis considered in this work, the modeling is based
only on three types of kinetic parameters, k(s-p), k(t-p),
k(q-p), for demethylation, and k(s-s), k(t-s), k(q-s), for de-
ethylation. For example, k(q-p) is the rate constant for the
scission of a bond between quaternary and primary carbon
atoms.
6.4.3. Kinetic Parameters for the Dehydrogenation of The overall equilibrium constant for the cyclization of the olefin
Naphthenes. In this case only one kinetic constant is used for into the naphthene can be expressed as the product of the
each type of aromatic produced, that is, one for benzene, one equilibrium constants of the steps involved in that reaction
for toluene, one for xylenes, etc., up to C10. sequence:

( )(
6.4.4. Adsorption Constants of the Hydrocarbons. In their
study of the adsorption of alkanes on zeolites, Denayer et al.50,51 + kOcycifNj + -1
showed that the physisorption enthalpy and entropy vary linearly KOcycihNj ) KOdehihDOjKpr⁄de
DOjhORim
Kdeh )
COjhCRim NjhCOi
Kpr⁄de
with the carbon number of the physisorbed alkane. kNendo
jfOi

(27)
ln Ki ) aCNi + b (23)
so that
The a and b parameters account for the contribution of the
+
physisorption entropy and enthalpy, and are considered to be KOdehihDOjKpr⁄de
DOjhORim

dependent on the type of hydrocarbon. By relating the adsorption kNendo


jfOi
) +
kOcycifNj (28)
COjhCRim NjhCOi OihNj
constant of a given hydrocarbon, Ki, with that of a reference Kpr⁄de Kdeh Kcyc
hydrocarbon, K0, of the same type (same value for b), the The protonation/deprotonation equilibrium constants cancel out
following equation is derived: when the rate coefficients are written in terms of the composite
single-event rate coefficients. The dehydrogenation and cycliza-
Ki ) K0 ea(CNi-CN0) (24)
tion equilibrium constants are calculated from the thermody-
Therefore, only one independent adsorption constant (K0) for namic properties of the corresponding species by means of
each type of hydrocarbon (P, O, N, A) is included in the model. Benson’s group contribution method.34
Denayer et al.50,51 showed that the parameters a and b in eq 23 Summarizing, the number of parameters has been reduced
strongly depend on the catalyst type and pore size. In the present by 11 so that for isothermal conditions the model contains 45
work a is assumed to have a unique value at a given temperature. independent parameters: 21 composite single-event rate coef-
1114 Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009

Table 2. Types and Numbers of Parameters in the Kinetic Model nresp nresp nexp

parameter description no. SSQ ) ∑ ∑ w ∑ (y jl ij - y


^ij)(yil - y^il) (30)
j)1 l)1 i)1
kpcp PCP isomerization of acyclic R+ 4
kβ β-scission of acyclic R+ 4 The weighting factors used in the estimation were calculated
kiras intraring alkyl shift of cyclic R+ 4 from the elements of the main diagonal of the inverse of the
krce ring contraction/expansion of cyclic R+ 4 error covariance matrix (nrespectively × nrespectively).

(∑ )
kcyc cyclization of olefinic R+ 4
kendoβ endocyclic β-scission of cyclic R+ 4 nexp -m
kexoβ exocyclic β-scission type I (on side chain of cyclic R+) 4 yij
kexoβ exocyclic β-scission type II (of cyclic R+) 4

∑ (∑ )
wjj ) σ-1
i)1
kdeh–p dehydrogenation of paraffins 5 jj ) n resp nexp -m
(31)
kdem demethylation of paraffins 3
kdeet de-ethylation of paraffins 3 yik
kdeh–N dehydrogenation of naphthenes 5 k)1 i)1
KP-deh adsorption of paraffins for dehydrogenation 1
The power m expresses the relative importance of the responses.
KP-hyls adsorption of paraffins for hydrogenolysis 1
KO adsorption of olefins 1 For m ) 1, the real weight of each response is used in the
KN adsorption of naphthenes 1 minimization, while for m ) 0, all the responses are equally
KA adsorption of aromatics 1 weighted. Previously, Feng53 and Park54 have found values
KH2deh adsorption of hydrogen 1 between 0.1 and 0.3 to be adequate. In this work, values of m
KH2hyls adsorption of hydrogen in hydrogenolysis 1
a parameter in eq 24 1
ranging from 0.3 to 0.6 were chosen. A value of m equal to 0.6
total 56 is used for olefins, which are present in very low concentrations.
The Marquardt55-Levenberg56 algorithm was used to mini-
Table 3. Experimental Conditions (Verstraete and Froment47) mize the objective function. Parameter estimates were first
conditions 673 K 713 K 733 K determined per temperature. Then, using appropriate reparam-
etrization57 and considering the complete set of experimental
total pressure, bar 7 5, 7, and 9 5, 7, and 9
molar ratios H2/HC 10 10 10, 15.4 data, the parameters (Ea and Ao, or ∆H and ∆S) were
space time interval, 0-103 0-92 0-92 simultaneously estimated at the three experimental temperatures.
(gcat h mol-1 HC fed) All parameters were significant and showed narrow confidence
number of experiments 8 8 8 intervals and high t-values. The F-tests were consistently high
number of responses 15 40 56
at the three temperatures.
The rate coefficients of the molecular reactions on the metal
ficients for the elementary steps on acid sites and 24 parameters sites generate trends observed in industrial reforming. For
for the reactions on the metal sites. These 45 parameters were example, the rate coefficient of naphthene dehydrogenation
estimated from the experimental data. increases with the C-number and the number of substituents in
the ring and largely exceeds that of paraffin dehydrogenation.
8. Experimental Data and Their Simulation In the latter, k(s-s) is favored. Combined with the relevant
concentrations it becomes clear that intrinsically these rates on
The present modeling work was based on data collected in a
the metal sites are much higher than those on the acid sites, a
comprehensive experimental study of Verstraete and Froment.47
situation aimed for in catalytic reforming. The rate coefficients
Their data were obtained from experimentation in a fixed-bed
for de-ethylation exceed those for demethylation. In the former
integral reactor loaded with a Pt-Sn/Al2O3 catalyst. The reactor
the (q-s) type is the fastest, in the latter the (q-p) type is.
was divided into eight beds, with sampling in between them so
The values for the single-event rate coefficients follow a trend
that values of the yields of each response were obtained at eight
typical for carbenium ion chemistry. In PCP steps k̃*(s,s) >
different space times. To eliminate internal mass transfer
k̃*(s,t) > k̃*(t,t). In acyclic and exocycyclic β-scission k̃*(t,t)
limitations, the catalyst was crushed to a size between 0.5 and largely exceed the others. The same trend is observed in
0.7 mm. To avoid catalyst deactivation by coke deposition, the endocyclic β-scission, but the k̃* values are lower. In ring
experiments were carried out at temperatures not exceeding 733
K and higher partial pressures of hydrogen than the one applied
in industrial operation. Three different temperatures and three
total pressures were investigated in the experimental program.
A summary of the operating conditions is given in Table 3.
For a plug flow reactor the fluid field continuity equations
for the various species can be written as

dFi
) Ri (29)
dWcat

Because of the very fast dehydrogenation of paraffins and


naphthenes, the set of differential equations can be very stiff,
especially at the entrance of the first reactor. Gear’s method52
was used to solve the set of the coupled nonlinear continuity
equations for the components and GOI.
The optimal set of parameters was estimated by minimizing
an objective function based upon the weighted sum of squares
of residuals which are the differences between experimental and Figure 7. Calculated (s) and experimental ((,2,b) yields for normal, single-
calculated yields of the responses: branched, and multibranched paraffins at 733 K and 7 bar.
Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009 1115

Figure 8. Calculated (s) and experimental (b, 9) yields for single-branched Figure 10. Caculated (s) and experimental (b, 9, 2) yields for the BTX
and multibranched C7 paraffins at 733 K and 7 bar. aromatics at 733 K and 7 bar.

spheres with a diameter of 3.18 mm, and the reactor loading


density was 1211 kgcat m3r-1. In what follows the results of the
simulations based upon a heterogeneous model are compared
with those obtained by a pseudohomogeneous model.59 The
heterogeneous reactor model explicitly accounts for diffusion
limitations inside the catalyst particle by additional transport
equations, to be integrated in each increment used for the
integration of the fluid field conservation equations. Interparticle
transport limitations are negligible in industrial catalytic reform-
ers because of the high flow rates. The homogeneous reactor
model ignores these transfer limitations and considers only the
fluid field continuity-, energy-, and pressure drop-equations.
Figure 11 shows the temperature profiles in each catalyst bed.
A steep temperature drop is observed in the inlet zone of the
first bed, caused by the very rapid naphthene dehydrogenations.
For the heterogeneous model, the total temperature drop in the
first bed amounts to 60 K. In the second bed additional
Figure 9. Calculated (s) and experimental (() yields for naphthenes at
733 K and 7 bar. dehydrogenation of naphthenes and paraffins occur, causing a
temperature drop of about 55 K. In the third bed dehydrocy-
contraction and expansion k̃*(t,t) > k̃*(s,t) > k̃*(s,s). The LC clization and the exothermic hydrocracking occur. The temper-
for PCP are two or three orders larger than those for β-scission. ature drop in the third bed is 33 K. The total space time was
It may be safely concluded that the rate of PCP-branching limited to 56.1 Kgcat h K mol-1 HC fed to avoid excessive
isomerization is larger than that of β-scission, as also observed hydrocracking and the resulting loss in reformate yield, which
in hydrocracking16 and MTO.58 is defined as the summation of yields of C5+ hydrocarbons.
In Figures 7-10 the values of the yields calculated by the Figure 12 shows the composition profiles of the PINA
model at 733 K are compared with the experimental data for fractions, that is, C5+ n- and i-paraffins, naphthenes, and
the various types of hydrocarbons. The model also predicts aromatics, and again the difference between the two model
values of the individual members of the different types, of predictions is pronounced. The diffusion limitations inside the
course, as evidenced by Figure 10. catalyst particle cannot be neglected.
The yield of the nP and naphthene fractions predicted by
the heterogeneous model is lower than that predicted by
9. Simulation of a Multibed Adiabatic Reformer with
the homogeneous model. The components of these fractions are
Radial Flow
reactants and in the presence of diffusion limitations the
A reforming unit consisting of three adiabatic beds with radial concentration of reactants is lower in the center of a particle
centripetal flow was simulated using the detailed kinetic model than at the surface so that the rate at which they react is lower.
described above. The reactor model contained continuity equa- It is the other way around with the aromatic fraction that is
tions for 69 components and GOI, an energy equation, and an made up to a large extent of products. The difference in the iP
equation for the pressure drop. The diffusional limitations inside yields of both models reflects the production of iP components
the catalyst particles were accounted for either through the in the first bed and their conversion at higher space times, that
Stefan-Maxwell or Wilke equations. The naphtha feed amounted is, in beds 2 and 3. At the end of the second bed, the paraffin
to 120 000 kg h-1. Its composition was that used in the isomers are almost at equilibrium among themselves. In the third
experimental program discussed above. The hydrogen recycle bed dehydrocyclization of paraffins into aromatics takes place,
ratio was maintained at 5.0 mol H2 mol-1 HC, and the total along with hydrocracking by β-scission on the acid sites and
pressure was 7 bar. The three beds operate at equal inlet hydrogenolysis on the metal sites, so that the concentration of
temperatures of 510 °C (783.15) K. The catalyst particles were i-paraffins keeps decreasing.
1116 Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009

Figure 11. Comparison of the simulated temperature profiles through the catalytic beds for the pseudohomogeneous and heterogeneous models. Inlet conditions:
783.15 K, 5.0 mol H2/mol HC, and 7 bar total pressure. The difference between the two model predictions is significant.

Figure 13. Evolution of the effectiveness factor for the dehydrogenation


Figure 12. Comparison of the simulated PINA yield profiles through the of n-octane: solution by the Wilke-type approximation with three collocation
catalyst beds. points.

Figure 13 shows the low effectiveness factors for the


dehydrogenation of n-octane resulting from the steep internal
concentration profiles. These factors are even lower for the very
fast naphthene dehydrogenations: they evolve from 0.15 to 0.30.
The η-values increase toward the end of each bed where the
lower temperature slows down the reaction rate. Figure 13 also
illustrates that the detailed kinetic model used here obviously
allows the prediction of the compositions of the various fractions
in great detail, that is, up to the individual components.
The hydrogen yield profile through the catalyst beds is shown
in Figure 14. The highest rate of production of hydrogen is
observed in the first bed, due to the fast dehydrogenation of
naphthenes. Because hydrogen is a product its yield predicted
by the pseudohomogeneous model is too high.
Olefins are undesirable in the reformate because of their
tendency to produce gums. Figure 16 illustrates that for the
chosen operating conditions the total olefin (C2) to C10)) yield
Figure 14. Hydrogen yield profiles.
at the reactor exit is not negligible. The olefins are mainly
produced by the dehydrogenations on the metal function of the
catalyst. Their yield rises very sharply at the high inlet Figure 16 shows the evolution of the RON and reformate
temperature of the three beds and then decreases as they are yield. The RON was determined by using the correlation
consumed by the acid catalyzed steps. obtained by Petroff et al.60
Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009 1117

Figure 16. Simulated profiles of the reformate yield and research octane
Figure 15. Simulated total olefins yield profiles, from C2) to C10). number.

10. Conclusions absence of these in the reactor model by expressing the kinetic
model in terms of so-called effective rate parameters.
The very detailed model for the catalytic reforming of naphtha
developed from its fundamental chemistry provides a thorough Acknowledgment
insight into the process and a wide predictive potential but it
contains a huge number of parameters. The application of the Rogelio Sotelo-Boyás acknowledges the National Council of
single-event approach to the modeling of the rate constants and Science and Technology of Mexico, CONACYT, for the
thermodynamic constraints allowed the reduction of the number financial support in the development of this work. The authors
of independent parameters for the steps on the acid sites to only also address special acknowledgments to Dr. Jan Verstraete for
21. providing the experimental data, as well as to Dr. Rayford
Current reforming models impose equilibrium for the hydro- Anthony for stimulating discussions.
genations and dehydrogenations. This is not the case in the
present model. By focusing on moieties in the molecules, in Appendix
this case the C-atoms on the produced double bonds or those List of Symbols
linked to the bond where the scission occurs, it was possible to A ) aromatic
reduce the number of rate and adsorption parameters for the a ) first contribution parameter of physisorption ∆H and ∆S
reactions on the metal function to 24. The total number of b ) second contribution parameter of physisorption ∆H and ∆S
parameters of the kinetic model then amounts to 45. These can Ct ) surface concentration of total acid active sites, mol sites gcat-1
be estimated from a well-designed experimental program. CR+ ) surface concentration of carbenium ions, mol R+ gcat-1
Because of their fundamental nature the parameters are invariant CH ) surface concentration of free acid sites, mol sites gcat-1
with respect to the feedstock. CR* + ) relative concentration of acid sites with R+ on the surface
For the commercial catalyst used in this investigation the CR*′+ ) composite relative concentration of acid sites with R+ on
values of the rate parameters of the dehydrogenation reactions the surface
on the metal function were found to be very large compared CH* + ) relative concentration of free acid sites on the surface
with those of the single-event rate coefficients, so that the rate- deh ) dehydrogenation
determining steps for the reforming reactions occur on the acid dep ) deprotonation
alumina function. That leads to a reformate with optimal Fi ) flow rate of component i, mol h-1
composition. Because of its fundamental nature the model is GOI ) group of isomers
also applicable to other reforming catalysts, like those of the H+ ) free or vacant acid sites
Pt/Re type, provided that the appropriate rate parameters are h ) Plank’s constant, 6.626068 × 10-34 m2 kg h-1
hyd ) hydrogenation
available.
K ) equilibrium adsorption constant
Despite the size of the kinetic model, the application to the K̃isom ) equilibrium single-event isomerization constant, dimension-
simulation of industrial reformers with radial flow was straight- less
forward and smooth. The reactor model contains 69 continuity K̃pr ) equilibrium single-event protonation constant, bar
equations for the external fluid field and the corresponding Keq ) equilibrium constant, dimensionless
number of equations for diffusion and reaction inside the pores. KDH ) equilibrium dehydrogenation constant, bar
The results proved the reliability of the model and illustrated k̃cr ) single-event rate coefficient for cracking, h-1
its performance as a tool for the prediction of the reformer k̃dep ) single-event rate coefficient for deprotonation, h-1
behavior and the optimization of the product yields. In this way, k̃pr ) single-event rate coefficient for protonation, bar h-1
the model significantly reduces the amount of pilot plant k̃*dep ) composite deprotonation single-event rate coefficient, mol
experimentation. It can also be used in the investigation of new gcat-1 h
catalysts. k̃*pr ) composite protonation single-event rate coefficient, mol gcat-1
The simulations have shown in an unambiguous way that bar h
diffusion limitations are important in catalytic reforming and k̃*pcp ) composite single-event rate coefficient for PCP, mol gcat-1
have to be accounted for. There is no way to compensate the bar h
1118 Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009

k̃*cr ) composite single-event rate for cracking, mol gcat-1 bar h (7) Hu, S. Y.; Zhu, X. X. Molecular Modeling and Optimization for
kN-A ) rate constant for the dehydrogenation of naphthenes Catalytic Reforming. Chem. Eng. Commun. 2004, 191, 500–512.
L ) active site (8) Padmavathi, G.; Chaudhuri, K. K. Modeling and Simulation of
Commercial Catalytic Naphtha Reformers. Can. J. Chem. Eng. 1997, 75,
Li ) lump i 930–937.
LC ) lumping coefficient, dimensionless (9) Turpin, L. E. Modeling Commercial Reformers. In Catalytic Naphtha
N ) naphthene Reforming; Antos, G. J., Aitani, A. M., Parera, J. M., Eds.; Marcel Dekker:
n ) chiral centers 1995; Chapter 14, p 437.
(10) Quann, R. J.; Jaffe, S. B. Building Useful Models of Complex
ne ) single-event number in an elementary step
Reaction Systems in Petroleum Refining. Chem. Eng. Sci. 1996, 51, 1615–
nexp ) number of experiments 1635.
nrespectively ) number of responses (11) Wei, W.; Bennett, C. A.; Tanaka, R.; Hou, G.; Klein, M. T. J.;
O ) olefin Klein, M. T. Computer Aided Kinetic Modeling with KMT and KME. Fuel
P ) paraffin Process. Technol. 2008, 89, 350–363.
pro ) protonation (12) Wei, W.; Bennett, C. A.; Tanaka, R.; Hou, G.; Klein, M. T. J.;
Klein, M. T. Detailed Kinetic Models for Catalytic Reforming. Fuel Process.
rds ) rate determining step Technol. 2008, 89, 344–349.
T ) absolute temperature, K (13) Baltanas, M. A.; Froment, G. F. Computer-Generation of Reaction
W ) weight of catalyst, Kg Networks and Calculation of Product Distributions in the Hydroisomeriza-
wjl ) elements of the inverse of the covariance matrix tion and Hydrocracking of Paraffins on Pt-Containing Bifunctional Catalysts.
Y ) response in terms of yields, g of i/100 g HC fed Comput. Chem. Eng. 1985, 9, 71–81.
(14) Clymans, P. J.; Froment, G. F. Computer-Generation of Reaction
Greek Letters Paths and Rate-Equations in the Thermal-Cracking of Normal and Branched
β ) refers to β-scission step Paraffins. Comput. Chem. Eng. 1984, 8, 137–142.
σi ) symmetry number of component i (15) Vynckier, E. ; Froment, G. F. Modeling of the Kinetics of Complex
Superscripts Processes Based Upon Elementary Steps In Kinetics and Thermodynamic
m ) power relative to the weight factor of responses Lumping of Multicomponent Mixture; Astarita, G., Sandler, S. I., Eds.;
Elsevier Science Publishers: Amsterdam, The Netherlands, 1991; p 131.
+ ) indicates charge of carbenium ion
(16) Svoboda, G. D.; Vynckier, E.; Debrabandere, B.; Froment, G. F.
* ) used to describe a composite single-event rate coefficient, i.e., Single-Event Rate Parameters for Paraffin Hydrocracking Oil a Pt/US-Y
k* Zeolite. Ind. Eng. Chem. Res. 1995, 34, 3793–3800.
^ ) calculated value, e.g., ŷ is the calculated yield (17) Feng, W.; Vynckier, E.; Froment, G. F. Single-Event Kinetics of
o ) initial, e.g., initial flow: Fo Catalytic Cracking. Ind. Eng. Chem. Res. 1993, 32, 2997–3005.
Subscripts (18) Froment, G. F. Kinetic Modeling of Acid-Catalyzed Oil Refining
Processes. Catal. Today 1999, 52, 153–163.
A ) aromatic
(19) Froment, G. F. Single Event Kinetic Modeling of Complex Catalytic
cr ) cracking Processes. Catal. ReV.-Sci. Eng. 2005, 47, 83–124.
dep ) deprotonation (20) Reyniers, G. C.; Froment, G. F.; Kopinke, F. D.; Zimmermann, G.
deh ) dehydrogenation Coke Formation in the Thermal-Cracking of Hydrocarbons. 4. Modeling
hyd ) hydrogenation of Coke Formation in Naphtha Cracking. Ind. Eng. Chem. Res. 1994, 33,
2584–2590.
hls ) hydrogenolysis
(21) Moustafa, T. M.; Froment, G. F. Kinetic Modeling of Coke
iras ) intraring alkyl shift Formation and Deactivation in the Catalytic Cracking of Vacuum Gas Oil.
M ) general notation for type of carbenium ion: s or t Ind. Eng. Chem. Res. 2003, 42, 14–25.
N ) naphthene, also naphtha in xN (22) Dewachtere, N. V.; Santaella, F.; Froment, G. F. Application of a
O ) olefin Single-Event Kinetic Model in the Simulation of an Industrial Riser Reactor
P ) paraffin for the Catalytic Cracking of Vacuum Gas Oil. Chem. Eng. Sci. 1999, 54,
3653–3660.
pro ) protonation (23) Arroyo, J. A. M.; Martens, G. G.; Froment, G. F.; Marin, G. B.;
pcp ) isomerization via a protonated cyclopropane intermediate Jacobs, P. A.; Martens, J. A. Hydrocracking and Isomerization of N-Paraffin
Q ) quaternary carbon atom Mixtures and a Hydrotreated Gasoil on Pt/ZSM-22: Confirmation of Pore
Ri+ ) carbenium ion Mouth and Key-Lock Catalysis in Liquid Phase. Appl. Catal. A-Gen. 2000,
rce ) ring contraction o expansion 192, 9–22.
(24) Martinis, J. M.; Froment, G. F. Alkylation on Solid Acids. Part 2.
ref ) reference, e.g., reference olefin: Oref Single-Event Kinetic Modeling. Ind. End. Chem. Res. 2006, 45, 954–967.
s ) secondary carbenium ion (25) Debrabandere, B.; Froment, G. F. Influence of the Hydrocarbon
t ) tertiary carbenium ion Chain Length on the Kinetics of the Hydroisomerization and Hydrocracking
of N-Paraffins. Hydrotreat. Hydrocracking Oil Fractions 1997, 106, 379–
389.
Literature Cited (26) Baltanas, M. A.; Vanraemdonck, K. K.; Froment, G. F.; Mohedas,
S. R. Fundamental Kinetic Modeling of Hydroisomerization and Hydroc-
(1) U.S. Code 7521; Clean Air Act, 2005; Title 42, Chapter 85, racking on Noble-Metal-Loaded Faujasites 0.1. Rate Parameters for
Subchapter II, Part A. Hydroisomerization. Ind. Eng. Chem. Res. 1989, 28, 899–910.
(2) Smith, R. B. Kinetic Analysis of Naphtha Reforming with Platinum (27) Alwahabi, S. M.; Froment, G. F. Single Event Kinetic Modeling
Catalyst. Chem. Eng. Prog. 1959, 55, 76–80. of the Methanol-to-Olefins Process on Sapo-34. Ind. Eng. Chem. Res. 2004,
(3) Ramage, M. P.; Graziani, K. R.; Schipper, P. H.; Krambeck, F. J.; 43, 5098–5111.
Choi, B. C. KINPTR (Mobil’s Kinetic Reforming Model): A Review of (28) Park, T. Y.; Froment, G. F. Kinetic Modeling of the Methanol to
Mobil’s Industrial Process Modeling Philosophy. AdV. Chem. Eng. 1987, Olefins Process. 1. Model Formulation. Ind. Eng. Chem. Res. 2001, 40,
13, 193–266. 4172–4186.
(4) Ancheyta, J.; Villafuerte-Macias, E.; Diaz-Garcia, L.; Gonzalez- (29) Mills, G. A.; Heinemann, H.; Milliken, T. H.; Oblad, A. G. Catalytic
Arredondo, E. Modeling and Simulation of Four Catalytic Reactors in Series Mechanism. Ind. End. Chem. 1953, 45, 134–137.
for Naphtha Reforming. Energy Fuels 2001, 15, 887–893. (30) Davis, B. H. Alkane Dehydrocyclization Mechanism. Catal. Today
(5) Ancheyta, J.; Villafuerte-Macias, E. Kinetic Modeling of Naphtha 1999, 53, 443–516.
Catalytic Reforming Reactions. Energy Fuels 2000, 14, 1032–1037. (31) Toppi, S.; Thomas, C.; Sayag, C.; Brodzki, D.; Le Peltier, F.;
(6) Hu, Y. Y.; Xu, W. H.; Hou, W. F.; Su, H. Y.; Chu, J. Dynamic Travers, C.; Djega-Mariadassou, G. Kinetics and Mechanisms of N-
Modeling and Simulation of a Commercial Naphtha Catalytic Reforming Propylbenzene Hydrodealkylation Reactions over Pt(Sn)/SiO2 and (Cl-)
Process. Chin. J. Chem. Eng. 2005, 13, 74–80. Al2O3 Catalysts in Reforming Conditions. J. Catal. 2002, 210, 431–444.
Ind. Eng. Chem. Res., Vol. 48, No. 3, 2009 1119
(32) Kumar, H.; Froment, G. F. A Generalyzed Mechanistic Kinetic Alumina and Platinum-Rhenium Alumina Catalysts. Ind. Eng. Chem. Fund.
Model for the Hydroisomerization and Hydrocracking of Long-Chain 1986, 25, 544–553.
Paraffins. Ind. End. Chem. Res. 2007, 46, 4075–4090. (47) Verstraete, J. Kinetic Modeling of Catalytic Reforming on Pt, Pt-
(33) Park, T. Y.; Froment, G. F. Analysis of Fundamental Reaction Rates Re and Pt-Sn Catalysts. Ph.D. Dissertation, Gent University, Belgium, 1998.
in the Methanol-to-Olefins Process on ZSM-5 as a Basis for Reactor Design (48) Van Trimpont, P. A.; Marin, G. B.; Froment, G. F. Kinetics of the
and Operation. Ind. Eng. Chem. Res. 2004, 43, 682–689. Reforming of C-7 Hydrocarbons on a Commercial PtRe/Al2O3 Catalyst.
(34) Benson, S. W. Thermochemical Kinetics; Methods for the Estima- Appl. Catal. 1986, 24, 53–68.
tion of Thermochemical Data and Rate Parameters, 2nd ed.; Wiley: New (49) Pines, H. The Chemistry of Catalytic Hydrocarbon ConVersions;
York, 1976. Academic Press, Inc.: London, 1981; Chapter 1.
(35) Dumez, F. J.; Froment, G. F. Dehydrogenation of 1-Butene into (50) Laxmi Narasimhan, C. S.; Thybaut, J. W.; Marin, G. B.; Martens,
ButadienesKinetics, Catalyst Coking, and Reactor Design. Ind. Eng. Chem. J. A.; Denayer, J. F.; Baron, G. V. Pore Mouth Physisorption of Alkanes
Proc. Des. DeV. 1976, 15, 291–301. on ZSM-22: Estimation of Physisorption Enthalpies and Entropies by
(36) Garin, F.; Gault, F. G.; Maire, G. Hydrogenolysis and Isomerization Additivity Method. J. Catal. 2003, 218, 135.
of Hydrocarbons on Metals 0.1. Hydrogenolysis of Pentanes on a (Pt-Al2O3- (51) Denayer, J. F.; Baron, G. V.; Martens, J. A.; Vanbutsele, G.; Jacobs,
0.2-Percent) Catalyst. NouV. J. Chim. 1981, 5, 553–561. P. A. Modeling of Adsorption and Bifunctional Conversion of N-Alkanes
(37) Sinfelt, J. H. Catalytic Hydrogenolysis on Metals. Catal. Lett. 1991, on Pt/H-ZSM-22 Zeolite Catalyst. Chem. Eng. Sci. 1999, 54, 3553.
9, 159–172. (52) Gear, C. W. Numerical Initial Value Problems in Ordinary
(38) Sinfelt, J. H. Kinetics of Ethane Hydrogenolysis. J. Catal. 1972, Differential Equations; Prentice-Hall: Englewood Cliff, N. J., 1971.
27, 468–471. (53) Feng, W. Kinetic Modeling of Catalytic Cracking of Paraffins and
(39) Zimmer, H.; Zoltan, P.; Tetenyi, P. Hydrogenolysis of N-Alkanes Aromatics. Ph.D. Dissertation, Universiteit Gent, Belgium, 1992.
over Platinum Catalyst. Acta Chim. Hung. 1987, 124, 13. (54) Park, T. Y. Kinetic Modeling of MTO Process. Ph.D. Dissertation,
(40) Cortright, R. D.; Matwe, R. M.; Dumesic, J. A. Ethane Hydro- University of Gent, Belgium, 1998.
genolysis over Platinum. Selection and Estimation of Kinetic Parameters. (55) Marquardt, D. W. An Algorithm for Least-Squares Estimation of
J. Mol. Catal. A 2000, 163, 91–103. Nonlinear Parameters. J. Soc. Indust. Appl. Math. 1963, 11, 431.
(41) Shang, S. B.; Kenney, C. N. Steady-State and Transient Kinetic (56) Levenberg, K. A Method for the Solution of Certain Problems in
Studies of Ethane Hydrogenolysis over Ru/Al2O3. J. Catal. 1992, 134, 134– Least Squares. Q. Appl. Math 1944, 2, 164–168.
150. (57) Kittrell, J. R. Mathematical Modeling of Chemical Reactions. AdV.
(42) Bond, G. C.; Hooper, A. D.; Slaa, J. C.; Taylor, A. O. Kinetics of Chem. Eng. 1970, 8, 97.
Metal-Catalyzed Reactions of Alkanes and the Compensation Effect. J. (58) Park, T. Y.; Froment, G. F. Kinetic Modeling of the Methanol to
Catal. 1996, 163, 319. Olefins Process. 2. Experimental Results, Model Discrimination, and
(43) Adam, S.; Bauer, A.; Timpe, O.; Wild, U.; Mestl, G.; Bensch, W.; Parameter Estimation. Ind. Eng. Chem. Res. 2001, 40, 4187–4196.
Schlogl, R. The Origin of the Positive Effect of Cadmium Acetate on the (59) Froment, G. F.; Bischoff, K. B. Chemical Reactor Analysis and
Action of Supported Palladium Catalysts. Chem.sEur. J. 1998, 4, 1458– Design, 2nd ed.; John Wiley & Sons: New York, 1990; Chapter 7.
1469. (60) Petroff, N.; Boscher, Y.; Durand, J. P. Determination Automatique
(44) Bond, G. C.; Cunningham, R. H. Alkane Transformations on de L’indice D’octane et de la Composition des Reformats par Chromatog-
Supported Platinum Catalysts. J. Catal. 1997, 166, 172. raphie en Phase Gazeuse. ReV. Inst. Fr. Pet. 1988, 43, 259.
(45) Sinfelt, J. H. The Turnover Frequency of Methylcyclohexane ReceiVed for reView April 15, 2008
Dehydrogenation to Toluene on a Pt Reforming Catalyst. J. Mol. Catal. A ReVised manuscript receiVed June 30, 2008
2000, 163, 123–128. Accepted July 2, 2008
(46) Van Trimpont, P. A.; Marin, G. B.; Froment, G. F. Kinetics of
Methylcyclohexane Dehydrogenation on Sulfided Commercial Platinum IE800607E

Das könnte Ihnen auch gefallen