Sie sind auf Seite 1von 9

120

Ind. Eng. Chem. Res. 2006, 45, 120-128

Modeling and Simulation of an Industrial Fluid Catalytic Cracking Riser Reactor


Using a Lump-Kinetic Model for a Distinct Feedstock
Celia Araujo-Monroy and Felipe Lo pez-Isunza*
Departamento de Ingeniera de Procesos e Hidra ulica, UniVersidad Auto noma Metropolitana-Iztapalapa,
AVenida San Rafael Atlixco 186, Col. Vicentina Me xico 09340, D.F. Mexico

A process model for an industrial fluid catalytic cracking reactor is an important tool for predicting the flexibility
to operate with different feedstocks within the expected range of conversion, yield of gasoline, and coke
production. In this work, a steady-state process model has been developed to describe a two-phase transportedbed riser reactor that incorporates the stripping section and uses a lump-kinetic scheme to account for the
cracking of different feedstocks characterized by the paraffin, aromatic, and naphthene contents. The dependency
of the kinetic parameters on the type of feedstock is accounted for by expressing frequency factors as a
function of the aromatic/naphthenic ratio in the feedstock. The stripping section, where the recovery of adsorbed
hydrocarbons takes place, is modeled as a continuous stirred tank reactor where mainly the thermal cracking
of the adsorbed species takes place. The process model is used to predict product yields from the catalytic
cracking of three different gas oils, and the comparison with industrial data shows a good prediction.
Introduction
Fluid catalytic cracking (FCC) is one of the most important
processes in the petroleum refining industry, converting heavy
hydrocarbon fractions into gasoline, distillates, and light olefins.
FCC reactors may operate to maximize gasoline production,
but the continuous reduction in the reservoir of light petroleum
fractions and the need for diesel and petrochemicals (olefins
and aromatics) will force refiners to process heavier feeds and
more residue.1 These will generate an increasing need to develop
detailed process models for industrial FCC reactors as a tool to
test the reactor flexibility to process a wide range of heavier
feeds. To compare the predictions of the actual product
distribution, the process model should include the separation
of catalyst from vapor products in the stripping section and the
catalyst regeneration section. On the other hand, a compromise
should be reached about the complexity of the kinetic expressions to be used: either the use of those based on carbenium
ion chemistry or the use of lump schemes, which may account
for families of different types of products.
Kinetic Modeling. Since the pioneering works of Voorhies,2
Blanding,3 and Weekman and Nace,4 there has been important
progress in the kinetic modeling of the catalytic cracking of
gas oils. Kinetic models based on lump schemes have been
defined in terms of their boiling point range and chemical family.
This has allowed the description of catalytic cracking reactions
performed in microactivity and industrial reactors. These
schemes normally consider that the cracking of gas oils follows
a second-order law,3 whereas gasoline cracks according to firstorder kinetics.4 In most cases, catalyst deactivation is described
by an exponential decay expression in terms of the contact time2
to account for catalyst deactivation due to coke deposition. The
early three-lump model,4 which has been useful in the past for
simulations of the steady state and the dynamic behavior of FCC
reactors, has a main drawback because it lumps together the
light vapor products (C1-C4) with coke. Modifications to the
three-lump model have led to models with an increasing number
* To whom correspondence should be addressed. E-mail felipe@
xanum.uam.mx.

of lumps, separation of coke from light gases, and the proposal


of different criteria to group the reaction products into light cycle
oil (LCO), gasoline, liquefied petroleum gas (LPG), light gas,
and coke. A successful 10-lump kinetic model was proposed
by Jacob et al.5 based on a reaction scheme that introduces
functional hydrocarbon groups such as paraffins, naphthenes,
aromatic rings, and alkyl aromatics to produce LCO, gasoline,
and coke, although it ignores important groups such as LPG
and lumps coke together with light gases. Different derivations
of the above scheme have been used by several researchers
because of its capacity to predict the cracking of a wide range
of feedstock compositions and operating conditions but requiring the tuning of the kinetic parameters to fit the experimental data.
A different class of kinetic models, based on fundamentals,
has allowed a detailed prediction of the product distribution of
families from the cracking reactions of gas oils. Although
different in the conceptual approach, two of these are the singleeVent kinetics6 (SEK), which considers the elementary steps
based on carbenium ion chemistry, and the structure-oriented
lumping,7 based on the cracking chemistry of a homologous
series of hydrocarbons, however preserving the lump concept.
In the SEK model, the kinetic parameters for the elementary
steps are estimated from an extensive experimental program of
key hydrocarbons and the number of single events calculated
for all of the molecules contained in the feedstock. However,
computational limitations appear because of the analytical
complexity of the hundreds of reactions involved in the
generation of the reaction network and in the offline calculation
of the lumping coefficients.6 In the second case,7 3000 molecular
species are employed using 60 reaction rules, which results in
a pathway with more than 30 000 elementary chemical reactions.
Again, the limitations are the lack of reactivity relationships
and the estimation of kinetic parameters based on a comprehensive data set for a wide range of gas oils, catalysts, and
operating conditions.10 Watson et al.11,12 have performed the
mechanistic modeling of cracking reactions, where reactants and
product molecules are organized into compound classes (paraffins, olefins, and aromatics) and their cracking reactions,

10.1021/ie050503j CCC: $33.50 2006 American Chemical Society


Published on Web 12/08/2005

Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006 121

described in terms of fundamental reaction steps involving


carbocation intermediates, are ruled by quantitative structurereactivity relationships determined from pure-component experiments. Liguras and Allen13,14 have developed structural models
that consider a hydrocarbon molecule in a petroleum mixture
as a collection of carbon centers, and the cracking behavior is
deduced from data of the cracking of model compounds.
Another approach to describe the complex kinetics of the
cracking reactions is the model by Pitault et al.8 and Landeghem
et al.;9 this is based on the molecular description of the cracking
and hydrogen-transfer reactions, together with the use of lumps
for families of alkanes, alkenes, cycloalkanes, alkenic cycloalkanes, and aromatics, in an attempt to describe an industrial
feedstock, however using only a limited number of reactions.
The kinetic parameters are estimated from a set of cracking
experiments using different types of gas oils in a microactivity
fixed-bed reactor.
Reactor Modeling. Several models have been developed for
the simulation of industrial FCC reactors covering different
aspects ranging from steady-state modeling9,15,16 to dynamics17-20
and control studies and its relevance21-24 or those dealing with
the detailed modeling of the complex hydrodynamics of twoand three-phase transported-bed cracking reactors using computational fluid dynamics methods,25-28 to mention just a few
of them. Most of FCC reactor models describe the riser reactor
as a two-phase plug-flow transported bed with constant slip
between the two phases, using lump-kinetic schemes with
different numbers of lumps. Some of them include the catalystvapor separation section, and others incorporate the catalyst
regeneration reactor into the FCC model. The aim in most cases
has been the description of the steady-state operation and the
prediction of product yields and temperature profiles to fit the
corresponding measurements at the exit of the industrial unit.
The models that incorporate the description of the hydrodynamics of the two- or three-phase system use again lump kinetics
with different numbers of lumps and are aimed at investigating
the contribution of radial density profiles of solids or the effect
of the position and number of feed nozzles, and the liquid drop
size on the liquid-vapor-catalyst contact along the riser reactor,
and predicting product yields.
Types of Feedstock. It has been shown from experiments29
in an isothermal microactivity reactor, using cuts with different
boiling point ranges and chemical compositions, that product
selectivity is mainly determined by the family type of the
feedstock, but it is only slightly affected by differences in the
boiling point range. However, an important aspect to be
considered in the development of a process model for an
industrial FCC reactor is the heat balance, for the proper
comparison of the units performance when using different types
of feedstock. The model should account for the total amount of
heat needed to fully vaporize the different types of feed, which
affect the required heat supply for the (mainly endothermic)
cracking reactions and, moreover, the amount of coke deposited
on the catalyst. This leads to the modification of operational
variables such as the gas oil feed temperature or the catalyst
circulation rate to obtain similar gas oil conversion or gasoline
yields.
In this work, a steady-state process model is developed to
describe a two-phase transported-bed reactor and the stripping
section. Our aim is to take into account the different types of
families within each lump that are considered in the kinetic
model to describe the cracking reactions of different types of
gas oils (characterized by their contents of paraffins, aromatics,
and naphthenes). Also, a second lump-kinetic scheme is used

Figure 1. Schematic diagram of the industrial riser reactor and stripping


section.

to describe the thermal cracking reactions of the adsorbed


species taking place in the stripping section, which is modeled
as a continuous stirred tank reactor (CSTR), because they have
been reported to be dominant in this unit. It is assumed that an
almost instantaneous vaporization of gas oil takes place, in less
than 0.2 s, based on previous published works.30 Then, the model
is used to predict the product distribution from the cracking of
three different gas oils, and the predictions of the product yields
from the steady-state operation are compared with industrial
data31 from a FCC unit in a Mexican refinery (Pemex).
Process Model for the Riser Reactor and the Stripping
Section
In this work, only the riser reactor and the separation section
are included in the steady-state model, as shown in Figure 1.
When a partially vaporized feedstock contacts the hotter catalyst
at the bottom of the riser, the vaporization of the liquid feed
and the cracking reactions taking place generate an increase in
the volumetric vapor flow rate that transports the catalyst along
the riser, with both traveling in nearly plug flow. The adsorbed
hydrocarbons on the catalyst leaving the riser must be stripped
off before entering the regenerator, and this is achieved in the
stripper section by using a steam flow rate in the range of 2-5
kg/1000 kg of catalyst recirculation. It has been reported that
the use of steam above this range has almost no effect, but below
this range, the temperature in the regenerator may increase.32
Kinetic Model for the Cracking Reactions. The lump
kinetics for the cracking of gas oils is based on a sequential
series of reaction pathways,8,9 in which the gas oil reacts to
form primary products contained in lumps like LCO, gasoline
(G), LPG, light gases (LG), and coke, as shown in Figure 2a.
Secondary products from the cracking of LCO give rise to
fractions in lumps: G, LPG, LG, and coke. Gasoline produces
LPG, LG, and coke; finally, LPG may produce important
amounts of LG and even coke.
To describe the cracking reactions within each lump, it is
assumed that the jth lump is constituted by (PONA) families
of paraffins (Pj), olefins (Oj), naphthenes (Nj), and aromatics
(Aj). The mass fraction of the ith family in the jth lump in the
gas phase (Xi,j) must adsorb first on the catalyst surface, and
then the adsorbed species (i,j) can react to produce the PONA
families in lump j + 1 in the gas phase, with mass fraction of
products Xi,j+1. Parts b-e of Figure 2 show the series of reaction
pathways from gas oil to LPG, assuming that they follow firstorder kinetics;4 this scheme has been based on the catalytic
cracking reactions reported in the literature.33 In what follows,

122

Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006

of paraffins (i ) 1) in the gasoline lump (j ) 3), the net rates


are

Rg1,3 ) k151,2 + k212,2 + k253,2 - k32(X1,3 - 1,3/K32) (4)


The first three terms on the right-hand side correspond to the
formation of paraffins from the PON components in lump 2,
and the last term gives its consumption for P4 and O4. The
corresponding expression for the net adsorption of species 1,3
is

R1,3 ) k32(X1,3 - 1,3/K32) - (k33 + k34)1,3

(5)

Second, for olefins (i ) 4) in the LCO lump (j ) 2), the net


rates are

Rg4,2 ) k31,1 + k82,1 - k20(X4,2 - 4,2/K20)

(6)

The corresponding expression for the net adsorption of species


4,2 is

R4,2 ) k20(X4,2 - 4,2/K20) - (k21 + k22 + k23)4,2


Figure 2. Lump-kinetic model for catalytic cracking of gas oil. P1, A1,
and N1 are the paraffinic, naphthenic, and aromatic contents of the gas oil,
and P1, N1, and A1 are their respective adsorbed species on the catalyst
surface. The following lumps are as indicated: LCO ) light cycle oil;
GASOLINE ) gasoline; LPG ) liquid petroleum gas; LG ) light gas,
and COKE.

the number of families (NF) is 4, and the number of lumps


(NL) is 6, although gas oil is only formed by three families
(PNA).
Riser Model. The model for the riser reactor describes a plugflow two-phase transported bed with slip between vapor and
solid phases, operating with residence times in the range of
1-10 s. It contains the mass and energy balances to give the
axial variation of temperature and mass fraction of all lumps in
the gas and solid phases along the riser reactor, expressing
catalyst deactivation by coke deposition in terms of the contact
time.2
Mass Balance Equations. The mass balances are given for
the PONA families in each lump, for the gas phase, and for the
adsorbed species on the catalyst. Then, the mass balance for
species from the ith family in the jth lump in the gas phase,
Xi,j, is

ugFg

dXi,j
) FbRgi,j
dz

(1)

The mass balance for adsorbed species on the catalyst, from


the ith family in the jth lump, i,j, is

(1 - )usFs

di,j
) FbRi,j
dz

(2)

In the above expressions, the catalyst deactivation function


is given in terms of the Voorhies expression:2

(7)

For any gas oil, it is assumed that the frequency factor for
the mth reaction step, A0m, is given by an empirical expression
in terms of the aromatic/naphthenic ratio contained in each
lump,34,35 multiplied by an additional frequency factor parameter,
R0m, which is fitted to the data from the industrial unit.

A0m ) R0m[25(A/N)-0.42]

(8)

Then, Arrhenius rate constants for any feedstock for cracking


reactions of gas oil, LCO, and gasoline are given by eq 9.

km ) A0me-EAm/RgT

(9)

Energy Balance Equations. The energy balances describe


an adiabatic riser reactor, considering that the catalyst supplies
the heat necessary to carry out the cracking reactions and to
vaporize the liquid feed, which involves as much as 60-70%
of the total amount of heat supplied by the incoming catalyst.32,36
This process is accounted for in terms of a modified interfacial
heat-transfer coefficient (see the Appendix). It has been assumed
that the vaporization of the feed is a very fast process and that
the contact between the solid and fluid phases has a very high
rate of interfacial heat transport.30
The energy balance for the fluid phase considers the transport
of heat between the gas and solid phases and is given by

ugFgCpg

dTg
) heffav(Ts - Tg)
dz

(10)

The energy balances for the solid phase consider the interfacial
heating of the gas phase and the heat supplied for the cracking
reactions:

(1 - )usFsCps

dTs
dz

NF NL NR

(r)R
i)1 j)1 r)1

) heffav(Tg - Ts) - Fb

i,j

(11)
) e-Rtc

(3)

To illustrate some of the rate expressions for the different


species in the gas phase, according to the kinetic scheme shown
in Figure 2b-e, two cases are presented: first, for the cracking

The last term accounts for all of the heat supply to the mainly
endothermic cracking reactions for all families in all lumps (NR
is the total number of reactions), neglecting the heat involved
in the adsorption of species; av is the ratio of the external surface
area per volume of the catalyst sphere.

Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006 123


Table 1. Physical Properties of Gas Oils
gas oil

Figure 3. Lump-kinetic model for thermal cracking and hydrocarbon


desorption in the stripping section. P1, N1, and A1 are the paraffinic,
aromatic, and naphthenic adsorbed species in the gas oil. P2, O2 N2, and
A2 are the paraffinic, olefinic, aromatic, and naphthenic contents of the
LCO. P3, O3 N3, and A3 are their respective adsorbed species for gasoline
on the catalyst surface.

The pertinent initial conditions at z ) 0 require that

Xi,j(0) ) X0i,j

for all i, and j ) 1

composition
paraffins (wt %)
aromatics (wt %)
naphthenes (wt %)
Conradson carbon
molecular weight avearge (kg/mol)
API
specific gravity (20/4 C)
aniline point (C)
K (UOP)
refraction index (0/20)
average boiling point (C)
distillation ASTM D-1160 (vol %/C)
initial boiling point
5/10
30/50
70/90
final boiling point

54.1
15.3
30.6
0.430
354
23.8
0.9112

60.5
22.5
17.0
0.110
387
22.94
0.9162
80.1
11.75
1.5131
434

56.9
22.2
20.9
0.230
290
24.82
0.9052
69
11.55
1.5055
361

226
322/361
408/432
456/494
539

223
243/253
307/409
475/515
548

335/343
412/430
478/510

(12a)

Xi,j(0) ) 0

for all i, and j > 1

(12b)

The mass balance in the stripper section for species from the
ith family in the jth lump in the gas phase, Yi,j, is

i,j(0) ) 0

for all i, and j g 1

(12c)

FgYi,j ) FgY0i,j + FbVR(rdesi,j - radsi,j)

(13)

coke(0) ) 0coke

(12d)

The corresponding mass balance for the adsorbed species in


s
the stripper i,j
is

Tg(0) ) Tg0

(12e)

s
) Fcat0s i,j - FbVR(ri,j + rdesi,j - radsi,j)
Fcati,j

Ts(0) ) Ts0

(12f)

Equations 12a, 12e, and 12f refer to the gas-oil composition


and fluid and catalyst feed temperatures, whereas eqs 12b and
12c state that no other lumps, either in the fluid phase or
adsorbed species, exist in the feed. Equation 12d gives the
residual coke on the catalyst coming from the regenerator that
enters into the riser reactor.
Stripping Section. The stripping section completes the
description of the system under consideration (see Figure 1),
allowing the prediction of yield products as a function of the
residence time in the stripper. In this section, mainly nonselective thermal cracking of adsorbed species (gas oil, LCO,
gasoline, and LPG) is assumed to take place, producing light
gases and coke, according to the kinetic scheme describing the
reactions in the stripping section. It is also assumed that the
stripping section has a larger catalyst inventory compared to
that in the riser reactor, and this section is modeled as an
isothermal CSTR. The contribution to the heat balance, corresponding to the feed of 2.5 kg of steam per ton of catalyst
is rather a small quantity as compared to that of the catalyst
(0.4% of the total heat); therefore, it is justified to assume an
isothermal operation. The residence time of the vapor products
and the catalyst in this section is typically in the range of 4-10
min.
Kinetics of the Thermal Cracking and the Mass Balances.
The lump-kinetic scheme to describe the thermal cracking
reactions of the adsorbed species on the catalyst entering into
the stripper is given in Figure 3a-d. Here, part of the different
components adsorbed on the catalyst are recovered to the vapor
phase; others are involved in the production of light gases and
coke. Olefins and aromatics produced mainly light gases and
coke, whereas part of the paraffins produced light gases, and
the naphthenic fractions desorbed. Within the LPG lump, olefins
produce coke and paraffins form light gases.

(14)

where rdesi,j and radsi,j are the rates for desorption and adsorption
in the kinetic scheme given in Figure 3a-d, whereas ri,j is the
rate of thermal cracking of the adsorbed species.
In the stripper section, conditions at the feed are for all i
and j

Yi,j(0) ) Xi,j(Lriser)

(15a)

s
i,j
(0) ) i,j(Lriser)

(15b)

scoke(0) ) coke(Lriser)

(15c)

Equation 15a specifies that the gas-phase compositions leaving


the riser are equal to those entering the stripper, whereas eqs
15b and 15c define the quantities adsorbed on the catalyst and
coke, respectively, at the risers exit.
Model Simulations of the Type of Feedstock. Model
simulations of the steady-state operation of the FCC reactor were
performed for three different vacuum gas oils; their characteristics are given in Table 1. Gas oils A and B are industrial
feedstocks, whereas gas oil C is a synthetic blend prepared at
Instituto Mexicano del Petroleo37 for catalyst testing and ranking,
containing different proportions of light and heavy gas-oil cuts
from Maya crude oil. Gas oil A has a larger naphthenic fraction
and is less aromatic than B, which, in turn, has a larger paraffinic
fraction than A.
To compare model predictions and test the kinetic scheme,
model simulations were performed to predict the yield of
products from the cracking of gas oil A, which comes from
that used in an industrial FCC unit.31 Table 2 lists the operating
conditions and reactor dimensions used in the simulations. Also,
model simulations of the cracking of gas oils B and C were
performed using a constant value for the enthalpy of vaporization, according to their true boiling point (TBP) curves. In all

124

Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006

Table 2. Reactor and Stripper Dimensions and Operating


Conditions Used in the Simulations
feed mass flow rate
catalyst mass flow rate
specific gravity
density of the catalyst
feed temperature
catalyst temperature
diameter
length
diameter
volume
catalyst mass flow rate
steam mass flow rate
steam/catalyst ratio
steam temperature

Riser Reactor
kg/s
kg/s
kg/m3
K
K
m
m

39.7
354.21
0.911
1500
487
943
1.32
38.9

Stripping Section
m
m3
kg/s
kg/s
kg/ton
K

6.9
117.9
354.21
0.885
2.5
483

simulations, the Conradson carbon for each feedstock was taken


into account in the deactivation expression.
Prediction of the Product Yields for Feedstock A. Values
of the activation energies used in the simulations were taken
from the study of the cracking of gas oils B and C performed
in a fluidized-bed microreactor (ACE-R) unit.38 Frequency
factors, given by eq 8 in terms of the ratio aromatics/
naphthenes34 in each lump, were obtained for gas oil A by a
trial-and-error fitting procedure of the frequency factor parameters R0m to predict the industrial yields (as a lump) at the exit
of the stripping unit.31 These also required the tuning of the
kinetic parameters in the stripping section, which were kept
constant during all simulations because of the nonselective
nature of thermal cracking. The values of the kinetic parameters
obtained for the cracking of gas oil A are summarized in Tables
3-5. Model predictions of the industrial data, at the exit of the
stripper, are very good, as can be seen in Table 6. Figure 4
gives the typical predicted profiles showing on the left-hand

side of this figure the axial mass yield profiles for all lumps
and on the right-hand part the yields at the exit of the stripping
unit. Figure 5 shows the corresponding axial temperature profiles
of solid and vapor phases along the riser, using an average value
for the enthalpy of vaporization of the feed. Figure 6 shows
the axial profiles of adsorbed hydrocarbons and coke produced
along the riser reactor, as well as the recovery of hydrocarbons
to the vapor phase after being stripped off from the catalyst in
the separation section.
An analysis of the results given in Table 6 shows that the
predicted recovery in the stripping section was 23.0% for gas
oil, 7.7% for LCO, and 1.95% for gasoline. The net loss of
LCO was 11.22% to light gas by thermal cracking, which agrees
well with the experimental data from the stripping stage in a
microreactor unit.37 As a consequence of this, the light gas lump
was increased by 15.9%, also in agreement with these observations.37 The extra amount of coke produced by thermal cracking
in the stripper was 1.4% of the total.
One of the aims for developing this kinetic model is to predict
the distribution of paraffins, olefins, aromatics, and naphthenes
within each lump. Figures 7-10 show the prediction of these
fractions in the conversion of gas oil and the PONA composition
of LCO, gasoline, and LPG. It can be observed that LCO
contains larger amounts of aromatics; gasoline is formed by
significant quantities in the order aromatics, olefins, paraffins,
and naphthenes. LPG contains more paraffins than olefins. The
product distribution and PONA composition agree well with
those reported in the literature, which suggests that paraffins
and aromatics in the gas oil produced mainly the lighter species
in LCO and gasoline, whereas the tendency of naphthenes is to
produce LPG and light gas.5,34,35
Model Predictions for the Cracking of Different Feedstocks. The prediction of reactor operation under different
feedstocks is a very important aspect that may decide the

Table 3. Kinetic Parameters Used in the Simulations for the Cracking of Three Gas Oils
path
no.
k1
k2
k3
k4
k5
k6
k7
k8
k9
k10
k11
k12
k13
k14
k15
k16
k17
k18
k19
k20
k21
k22
k23
k24
k25
k26
k27
k28
k29
k30
k31
k32

P1 f P2
P1 f O2
P1 f N2
P1 f A2
P1 f C
N1 f O2
N1 f A2
N1 f C
A1 f A2
A1 f C
P2 f P3
P2 f O3
P2 f N3
P2 f A3
P2 f C
O2 f P3
O2 f A3
O2 f C
N2 f P3
N2 f A3
N2 f GS
A2 f A3
A2 f GS
A2 f C
P3 f P4
P3 f O4
O3 f P4
O3 f C
N3 f O4
N3 f GS
A3 f GS
A3 f C

frequency
factor R0m

activation
energy EAl
(kcal/kg)

2.00 105
3.00 105
5.00 104
3.00 105
2.00 103
4.00 105
6.00 102
2.00 104
5.00 105
5.00 103
2.70 102
6.29 105
5.69 104
1.20 102
1.20 104
5.99 105
5.99 105
4.79 10
5.36 105
5.24 103
3.00 100
3.00 100
7.49 100
1.08 101
1.29 106
3.75 105
3.87 103
1.65 10-1
1.24 103
7.49 104
1.38 106
1.65 100

9500
10 100
10 200
10 300
10 400
11 000
11 100
11 200
12 000
12 100
10 500
9500
9500
10 500
10 400
10 500
10 500
11 000
11 000
11 000
13 000
9200
13 000
12 800
10 000
10 000
10 980
14 000
9900
14 500
13 500
14 500

rate coefficient
ki,j (kggas/kgcats)
B

6.67 106
2.95 102
4.43 101
2.40 102
1.44 10
1.55 102
2.10 10-1
6.30 10
6.89 101
6.22 10-1
1.75 10-1
1.15 103
1.04 102
7.78 10-2
8.63 10
3.89 102
3.89 102
1.86 10-3
2.08 102
2.03 100
1.47 10-4
7.45 10-3
3.68 10-4
6.51 10-4
1.40 103
4.08 102
1.53 100
2.88 10-6
1.50 10
7.81 10-1
4.03 101
1.72 10-5

4.44 106
1.96 102
2.95 101
1.59 102
9.59 10-1
1.03 102
1.40 10-1
4.20 100
4.59 101
4.14 10-1
1.17 10-1
7.65 102
6.92 101
5.18 10-2
5.75 100
2.59 102
2.59 102
1.24 10-3
1.38 102
1.35 100
9.79 10-5
4.96 10-3
2.45 10-4
4.33 10-4
9.34 102
2.72 102
1.02 100
1.92 10-6
1.00 10
5.20 10-1
2.68 101
1.14 10-5

4.87 106
2.15 102
3.23 101
1.75 102
1.05 100
1.13 102
1.53 10-1
4.60 100
5.04 101
4.54 10-1
1.28 10-1
8.39 102
7.59 101
5.68 10-2
6.30 100
2.84 102
2.84 102
1.36 10-3
1.52 102
1.48 100
1.07 10-4
5.44 10-3
2.69 10-4
4.75 10-4
1.02 103
2.98 102
1.12 100
2.10 10-6
1.10 10
5.70 10-1
2.94 101
1.25 10-5

Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006 125


Table 4. Adsorption and Desorption Parameters Used in the
Simulations for the Cracking of Three Gas Oils
path no.

adsorption constant
(mgas3/kgfree sitess)

desorption constant
(kggas/kgcats)

k1
k2
k3
k4
k5
k6
k7
k8
k9
k10
k11

1.0 103
2.0 101
3.0 101
1.5 102
5.0 101
1.0 102
1.5 102
7.5 101
8.0 101
8.5 101
7.5 101

1.0 10-3
2.0 10-4
4.0 10-4
1.0 10-4
8.0 10-5
2.0 10-4
2.0 10-4
9.0 10-5
9.0 10-5
1.0 10-4
1.0 10-4

Table 5. Adsorption and Desorption Parameters Used in the


Simulations for the Stripping Section
path
no.

adsorption
constant
(mgas3/kgfree sitess)

desorption
constant
(kggas/kgcats)

k2
k3
k4
k7
k8
k9
k10
k13
k15
k16
k17
k20
k22

2.0 101
6.8 100
4.5 101
5.3 101
1.1 101
1.5 101
6.8 100
4.9 101
1.5 101
3.8 100
3.8 10-1
7.5 100
5.3 10-1

1.35 10-1
4.5 10-2
3.0 10-3
3.5 10-1
7.0 10-2
9.5 10-2
4.5 10-2
3.2 10-1
1.0 10-1
2.5 10-2
2.5 10-3
5.0 10-2
3.5 10-3

path
no.

rate
coefficient
(kggas/kgcats)

k1
k5
k6
k11
k12
k14
k18
k19
k21

1.6 10-7
4.1 10-6
7.4 10-6
9.4 10-8
2.6 10-6
1.0 10-7
4.1 10-6
7.4 10-6
9.4 10-8

Figure 4. Axial profiles in the riser section and time evolution in the
stripping section of product yields.

Table 6. Comparison of the Product Yields of the Experimental


Data from the Pemex Refinery with the Predicted Values for Gas
Oil A at the Exit of the Riser Reactor and from the Stripping
Section
lump

riser outlet
(L ) 38.9 m)

stripper outlet
(t ) 5.05 min)

industrial
data

gas oil
LCO
gasoline
LPG
light gas
coke

5.19
13.32
49.83
14.61
8.43
5.68

6.40
14.35
50.58
12.97
9.77
5.76

6.50
15.00
49.70
13.30
9.90
6.00

revenue of a FCC unit. Changes in the contents of paraffins,


aromatics, or naphthenes in the gas oil will impact product
distribution and coke formation, which, in turn, will also have
a strong impact on the heat balance of the FCC unit. To test
the kinetic scheme, the simulation of the reactor performance
compares now the cracking of gas oils A-C with characteristics
given in Table 1. The kinetic parameters used in these
simulations are those given in Table 3, but the frequency factor
parameters R0m obtained for gas oil A were corrected by eq 6
for the simulations of the cracking of gas oils B and C. Table
7 gives the comparison of the gas-oil conversion and product
yields using the same operating conditions as those in the FCC
unit. It can be observed that gas oil A produces more gasoline,
LPG, and coke than B or C, as shown in Table 7. These results
can be explained in terms of the decreasing order of reactivity
of the different families in the gas oil, with the aromatics being
the less reactive, whereas the largest reactivity is associated with
paraffins followed by naphthenes.34,35 As shown in Table 1, gas
oil B has more aromatics than gas oil A. The above has also
been reflected in the magnitude of the kinetic parameters for
the different gas oils, as shown in Table 3. One of the important
results considering the simulations given in Table 7, when
comparing the cracking of gas oils A and B, is the prediction

Figure 5. Axial temperature variations for both gas and solid phases.

Figure 6. Coke formation and adsorption-desorption of hydrocarbons in


the riser and stripping sections.

of a larger amount of gasoline (6.6%) from the cracking of gas


oil A, although the amount of coke increases in 0.8%; this has
to be balanced with the economical benefits for processing gas

126

Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006

Figure 7. Conversion of gas oil A along the riser reactor in terms of the
contents of paraffins, aromatics, and naphthenes.

Figure 9. Gasoline yield in terms of the contents of paraffins, olefins,


aromatics, and naphthenes, from the cracking of gas oil A.

Figure 8. LCO yield in terms of the contents of paraffins, olefins, aromatics,


and naphthenes, from the cracking of gas oil A.

Figure 10. LPG yield in terms of the contents of paraffins and olefins,
from the cracking of gas oil A.

oil A, which forces one to operate the unit with a larger amount
of coke on the catalyst.
Finally, a simulation was performed considering the variations
in the gas-oil enthalpy of vaporization due to their differences
in the TBP-distillation curves, although a rigorous calculation
was not carried out. According to Nelson,39 the enthalpy of
vaporization decreases as the molecular weight increases,
depending also on the amount of paraffins, aromatics, and
naphthenes contained in the gas oil. The predicted temperatures
for the three gas oils at the exit of the riser reactor showed that
their differences were negligible; however, the main impact of
processing different types of feedstocks is on the heat balance
of the FCC unit because of the differences in the amount of
coke on the catalyst produced.

Table 7. Comparison of Product Conversion and Yields at the Exit


of the Riser Obtained from Model Simulations of the Cracking of
Three Gas Oils

Conclusions
A steady-state process model has been developed to simulate
the operation of an industrial FCC riser reactor when different
types of feedstocks are processed. The model describes a twophase transported-bed riser reactor connected to a CSTR
stripping section and uses a 13-lump kinetic scheme that

feedstock

gas oil
LCO
gasoline
LPG
light gas
coke

5.19
13.32
49.83
14.61
8.43
5.68

7.81
28.41
43.25
4.85
5.87
4.88

7.32
27.23
44.70
5.34
5.81
5.01

considers each gas oil in terms of its paraffin, olefin, aromatic,


and naphthene contents. The dependency of the kinetic parameters on the type of feedstock is accounted for in the frequency
factors, which are given as a function of the ratio aromatics/
naphthenes in the gas oil. On the other hand, a set of reactions
that take place in the stripping section describe the thermal
cracking of all adsorbed species on the catalyst during the
recovery of the adsorbed hydrocarbons. The process model was
used to predict the product yields from the catalytic cracking
of three different gas oils. The model successfully predicted
conversion and yield from the cracking of gas oil A when
compared to industrial data. The simulation study showed that

Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006 127

the reaction rate constants increase as the aromatics/naphthenes


ratio decreases and the molecular weight of the feedstock
increases; however, the selectivity to gasoline and coke is a
function of the gas-oil PONA composition of each lump.

interval where vaporization took place is

FGCpGTG ) z

3vFLKev
3h
(T - TG) +
RP s
2R 2
d

(A.9)

Appendix
On the Vaporization of a Liquid Drop. It is considered
that right at the entrance of the reactor the liquid vaporizes in
a period of time no longer than 0.05-0.2 s,30 which represents
a very short distance from the entrance, z. The approximation
used to account for the amount of heat spent in vaporizing the
feed is based on the following assumptions:
1. Assume that the heat to vaporize a single drop comes
mainly from the gas-solid mixture that surrounds the drop (the
feedstock enters the riser partially valorized), at T.
2. The heat balance for a single drop may also include the
heat transferred by radiation from the solid, but its contribution
has been shown to be small.
3. The vaporization is instantaneous, and all of the heat that
has been transferred to the drop for its vaporization is transferred
to the gas phase, where no reactions take place (thermal cracking
is negligible in comparison with catalytic cracking).
4. The temperature of the drop is constant and equal to that
in its external surface.
5. A heat balance at some time before vaporization finishes,
assuming that the increase in the sensible amount of heat in the
drop is much smaller that the latent heat, leads to

qvap ) qconv

(A.1)

where

qvap ) Lv

heat flux due to vaporization (A.2)

qvap ) h(T - TL)

convective heat flux

(A.3)

We can express the mass flux due to the vaporization process


as

L ) FL

dRd
dt

(A.4)

where Rd is the diameter of the droplet at some time t and L


is the rate of mass vaporization. If the vaporization process obeys
the R 2 law,30 then

dRd2/dt ) Kev

(A.5)

Therefore, substituting eq A.4 into eq A.5 gives

L ) FLKev/2Rd

(A.6)

Substituting eq A.6 into eq A.1 gives

h(T - TL) ) vap L ) vapFL

dRd vapFLKev
(A.7)
)
dt
2Rd

Finally, the constant term for vaporization is given by

Kev )

2hRd
T
vapFL

(A.8)

If we consider that the heat spent in vaporizing the liquid


feed is transferred at once after the vaporization has been
finished, the heat balance for the gas phase along the length

Substituting eq A.8 into eq A.9, we obtain

FGCpG

TG 3h
3h
(Ts - TG) + (Ts - TG) )
z
RP
Rd

( )

Rd
3h
(Ts - TG) 1 +
(A.10)
Rd
RP
We define an effective interfacial heat-transfer coefficient that
takes into account the heat supplied to vaporize the liquid
feedstock as

( )

heff ) h 1 +

Rd
RP

RP is the radius of the catalyst particle, and the expression for


h is taken from the correlation developed by Ranz and Marshall
(Chem. Eng. Prog. 1952, 48, 141 and 173).
Nomenclature
Cpg ) specific heat capacity of the gas phase, kcal/kgK
Cps ) specific heat capacity of the solid phase, kcal/kgK
Fg ) gas flow rate in the stripper, kggas/s
Fcat ) flow rate of the catalyst in the stripper, kgcat/s
kads ) adsorption rate constants, m3gas/kgfree sitess
kdes ) desorption rate constants, kggas/kgcats
km ) kinetic rate constants for the adsorbed phase in the riser,
kggas/kgcats
h ) interphase heat-transfer coefficient, kcal/m2sK
LR ) reactor length, mR
Qvap ) heat of vaporization, kcal/m3s
rgi,j ) reaction rate for the gas phase in the stripper, kgi,j/kgcats
ri,j ) reaction rate for the solid phase in the stripper, kgi,j/kgcat
s
Rgi,j ) reaction rate for the gas phase in the riser, kgi,j/kgcats
Ri,j ) reaction rate for the solid phase in the riser, kgi,j/kgcats
tc ) contact time s
T0 ) gas temperature at the feed, K
Tg ) gas temperature, K
TReg ) solid temperature at the feed, K
Ts ) solid temperature, K
ug ) gas velocity, mR/s
us ) solid velocity, mR/s
VR ) stripper volume, m3
Xi,j ) weight fraction in the gas phase, kgi,j/kgtotal
z ) axial coordinate, m
Symbols
R ) deactivation parameter, s-1
Hr ) enthalpy of reaction, kcal/kgi,j
Hvap ) enthalpy of vaporization, kcal/kggas
 ) void fraction in the riser, m3gas/m3R
) deactivation function
i,j ) weight fraction in the solid phase, kgi,j/kgtotal
L ) fraction of free active sites, kg/kgtotal
i,j ) fraction of adsorbed species, kg/kgtotal
Fb ) bulk density, kg/mR3

128

Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006

Fg ) gas density, kggas/m3gas


Fs ) catalyst density, kgcat/m3cat
Acknowledgment
The authors gratefully acknowledge the financial support from
Instituto Mexicano del Petroleo and Consejo Nacional de
Ciencia y Tecnologa (Mexico) for the realization of this study.
We are indebted to R. Quintana-Solorzano and Dr. F. J.
Hernandez-Beltran for providing the experimental data from the
microreactor plant and for helpful discussions and to Drs. D.
Salazar and R. Maya for providing the industrial data. We also
thank Dr. Noemi Moreno for providing the kinetic parameter
data and Dr. Ir. Antonio Munoz-Arroyo for many helpful
suggestions.
Literature Cited
(1) Letzsch, W. S. Fluid catalytic cracking in the new millenium. NPRA
1999 Annual Meeting, San Antonio, TX, Mar 21-23, 1999; Paper AM99-15.
(2) Voorhies, A. Carbon formation in catalytic cracking. Ind. Eng. Chem.
1945, 37, 318.
(3) Blanding, F. Reaction Rates in Catalytic Cracking of Petroleum. Ind.
Eng. Chem. 1953, 45, 1186.
(4) Weekman, V. W., Jr.; Nace, D. M. Kinetics of catalytic cracking
selectivity in fixed, moving and fluid bed reactors. AIChE J. 1970, 16, 397.
(5) Jacob, S. M.; Gross, B.; Voltz, S. E.; Weekman, V. W., Jr. A lumping
and reaction scheme for catalytic cracking. AIChE J. 1976, 22, 701.
(6) Feng, W.; Vynckier, E.; Froment, G. F. Single-event kinetics of
catalytic cracking. Ind. Eng. Chem. Res. 1993, 32, 2997.
(7) Quann, R. J.; Jaffe, S. B. Structure-oriented lumping: describing
the chemistry of complex hydrocarbon mixtures. Ind. Eng. Chem. Res. 1992,
31, 2483.
(8) Pitault, I.; Nevicato, D.; Forissier, M.; Bernard, J. R. Kinetic Model
Based on a Molecular Description for Catalytic Cracking of Vacuum Gas
Oil. Chem. Eng. Sci. 1994, 49 (24A), 4249.
(9) Landeghem, F. V.; Nevicato, D.; Pitault, I.; Forissier, M.; Turlier,
P.; Derouin C.; Bernard, J. R. Fluid catalytic cracking: modelling of an
industrial riser. Appl. Catal. A 1996, 138 (2), 159.
(10) Christensen, G.; Apelian, M. R.; Hickey, K. J.; Jaffe, S. B. Future
directions in modeling the FCC processes: an emphasis on product quality.
Chem. Eng. Sci. 1999, 54, 2753.
(11) Watson, B. A.; Klein, M. T.; Harding, R. H. Mechanistic modeling
for n-heptane cracking on HZSM-5. Ind. Eng. Chem. Res. 1996, 35, 1506.
(12) Watson, B. A.; Klein, M. T.; Harding, R. H. Mechanistic modeling
of a 1-phenyloctane and n-hexadecane mixture on rare earth Y zeolite. Ind.
Eng. Chem. Res. 1996, 36, 2954.
(13) Liguras, D. K.; Allen, D. T. Structural models for catalytic cracking.
1. Model compound reactions. Ind. Eng. Chem. Res. 1989, 28, 665.
(14) Liguras, D. K.; Allen, D. T. Structural models for catalytic cracking.
2. Reactions of simulated oil mixtures. Ind. Eng. Chem. Res. 1989, 28,
674.
(15) Kumar, S.; Chadha, A.; Gupta, R.; Sharma, R. CATCRAK: a
process simulator for an integrated FCC-regenerator system. Ind. Eng. Chem.
Res. 1995, 34, 3737.
(16) Dewachtere, N. V.; Santella, F.; Froment, G. F. Application of a
single event kinetic model in the simulation of an industrial riser reactor
for the catalytic cracking of vacuum gas oil. Chem. Eng. Sci. 1999, 54,
3653.
(17) Lopez Isunza, F. Dynamic Modelling of an industrial fluid catalytic
cracking unit. Comput. Chem. Eng. 1992, 16 (Suppl.), 139.
(18) McFarlane, R. C.; Reineman, R. C.; Bartee, J. F.; Georgakis, C.
Dynamic simulator for a model IV fluid catalytic cracking unit. Comput.
Chem. Eng. 1993, 17, 275.

(19) Ali, H.; Rohani, S. Dynamic modeling and simulation of a risertype fluid catalytic cracking unit. Chem. Eng. Technol. 1997, 20, 118.
(20) Arandes, J. M.; Azkoiti, M. J.; Bilbao, J.; de Lasa, H. I. Modelling
FCC Units under Steady and Unsteady-State Conditions. Can. J. Chem.
Eng. 2000, 78, 111.
(21) Lopez-Isunza, F.; Alvarez-Ramirez, J.; Aguilar, R. A strategy to
regulate temperature in FCC units with poor knowledge of chemical kinetics.
Comput. Chem. Eng. 1996, 20 (Suppl.), 859.
(22) Alvarez Ramirez, J.; Aguilar, R.; Lopez-Isunza, F. Robust regulation
of temperature in reactor-regenerator fluid catalytic cracking units. Ind. Eng.
Chem. Res. 1996, 35, 1652.
(23) Ali, H.; Rohani, S.; Corriou, J. P. Modelling and control of a riser
type fluid catalytic cracking (FCC) unit. Trans. Inst. Chem. Eng. 1997, 75A,
401.
(24) Lee, W.; Weekman, V. W., Jr. Advanced control practice in the
chemical process industry: a view from industry. AIChE J. 1976, 22, 27.
(25) Theologos, K. N.; Markatos, N. C. Advanced modeling of fluid
catalytic cracking riser-type reactors. AIChE J. 1993, 39, 1007.
(26) Theologos, K. N.; Nikou, I. D.; Lygeros, A. I.; Markatos, N. C.
Simulation and design of fluid catalytic-cracking riser-type reactors. AIChE
J. 1997, 43, 486.
(27) Das, A. K.; Boudrez, E.; Marin, G.; Heynderickx, G. J. Threedimensional simulation of a fluid catalytic cracking riser reactor. Ind. Eng.
Chem. Res. 2003, 42, 2602.
(28) Gao, J.; Xu, C.; Lin, S.; Yang, G.; Guo, Y. Feedstock atomization
effects on FCC riser reactors selectivity. AIChE J. 2001, 47, 677.
(29) Harding, Z.; Qian, R. Fluid catalytic cracking selectivities of gas
oil boiling point and hydrocarbon fractions. Ind. Eng. Chem. Res. 1996,
35, 2651.
(30) Buchanan, J. S. Analysis of Heating and Vaporization of Feed
Droplets in Fluidized Catalytic Cracking Risers. Ind. Eng. Chem. Res. 1994,
33, 3104.
(31) Salazar, D.; Maya, R. Personal communication, Maya Crude Oil
Programme, Instituto Mexicano del Petroleo, 2003.
(32) Sadeghbeigi, R. Fluid Catalytic Cracking Handbook, 2nd ed.; Gulf
Professional Publishing: Houston, TX, 2000.
(33) Gates, B. C.; Katzer, J. R.; Schuit, G. C. Chemistry of Catalytic
Processes; McGraw-Hill: New York, 1979.
(34) Voltz, S. E.; Nace, D. M.; Weekman, V. W., Jr. Application of
Kinetic Model for Catalytic Cracking: some correlations of rate constants.
Ind. Eng. Chem. Process Des. DeV. 1971, 10, 538.
(35) Nace, D. M.; Voltz, S. E.; Weekman, V. W., Jr. Application of
Kinetic Model for Catalytic Cracking. Effects of charge stocks. Ind. Eng.
Chem. Process Des. DeV. 1971, 10, 530.
(36) Pekediz, A.; Kraemer, D.; Blasetti, A.; de Lasa, H. I. Heats of
Catalytic Cracking. Determination in a Riser Simulator Reactor. Ind. Eng.
Chem. Res. 1997, 36, 4516.
(37) Quintana-Solorzano, R.; Hernandez-Beltran, F. Study of the effect
of synthetic blends of Maya feedstock on the stripping of gasoline and LPG
adsorbed on equilibrium FCC catalysts. Internal Report (in Spanish),
Instituto Mexicano del Petroleo, June 2000.
(38) Moreno-Montiel, N. Evaluation of catalytic cracking catalysts by
modeling of micro-activity reactors: catalytic cracking and deactivation
kinetic models. (in Spanish). Ph.D. Thesis, Universidad Autonoma Metropolitana-Iztapalapa, Mexico, 2004.
(39) Nelson, W. L. Petroleum Refinery Engineering, 4th ed.; McGrawHill: New York, 1958.

ReceiVed for reView April 27, 2005


ReVised manuscript receiVed October 11, 2005
Accepted October 18, 2005
IE050503J

Das könnte Ihnen auch gefallen