Sunteți pe pagina 1din 10

Fuel 88 (2009) 169178

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

A comprehensive estimation of kinetic parameters in lumped catalytic cracking reaction models


Jos Roberto Hernndez-Barajas a, Richart Vzquez-Romn b,*, Ma.G. Flix-Flores c
a

Universidad Jurez Autnoma de Tabasco, Divisin Acadmica de Ciencias Biolgicas, Carretera Villahermosa-Crdenas Km 0.5, 68039 Villahermosa, Mexico Instituto Tecnolgico de Celaya, Departamento de Ingeniera Qumica, Av. Tecnolgico y G, Cubas s/n, 38010 Celaya, Mexico c Universidad Autnoma de Zacatecas, Unidad Acadmica de Ciencias Qumicas, Carretera a Cd. Cuauhtmoc Km 0.5, Guadalupe, Zacatecas 98600, Mexico
b

a r t i c l e

i n f o

a b s t r a c t
This work presents a comprehensive approach to estimate kinetic parameters when the involved reactions contain lumped chemical species. This approach is based on representing rate constants with a continuous probability distribution function. In particular, the beta function is used to estimate kinetic parameters in catalytic cracking reactions. Thus, several kinetic models for the catalytic process containing different number of lumps are selected and a discretization procedure is carried out to estimate the corresponding kinetic parameters. The kinetic representation based on the probability distribution substantially reduces the computational and experimental effort involved in the numerical evaluation of kinetic parameters. Published by Elsevier Ltd.

Article history: Received 12 November 2007 Received in revised form 16 July 2008 Accepted 17 July 2008 Available online 19 August 2008 Keywords: Catalytic cracking Lumping Kinetic modeling Distribution function

1. Introduction The catalytic cracking process converts heavy oil cuts into products of better quality such as olens and high octane gasoline using a zeolite-type catalyst. The uid catalytic cracking process (FCC) is in fact the key step of conversion in petroleum reneries. The FCC unit employed as a reference in this work is capable of processing 30,000 barrels per day as given in [1]. Catalytic reactions take place in the riser consisting of a vertical pipe with a diameter between 0.5 and 1.5 m and a height between 30 and 50 m. The heavy feedstock is instantaneously vaporized when it mixes with the hot and regenerated catalyst (catalyst residence time in the riser is around 5 s) and then the reaction proceeds in rather little time because of the high catalyst activity. During the reaction, some FCC products and a hydrogen-decient compound named coke is formed and deposited on catalyst surface. As a result, the coke deposition temporarily deactivates the catalyst. FCC products are then disengaged via cyclones while spent catalyst is passed through a stripper to separate the FCC products adsorbed on catalyst surface. Finally, coke is burnt in the regenerator and converted into CO, CO2, H2O, sulfur compounds and nitro compounds. This action restores the catalyst activity though fresh catalyst must be added periodically

* Corresponding author. Tel.: +52 461 61 17575x153; fax: +52 461 61 17744. E-mail addresses: roberto.hernandez@dacbiol.ujat.mx (J.R. Hernndez-Barajas), richart@iqcelaya.itc.mx (R. Vzquez-Romn), g_felixores@hotmail.com (Ma.G. Flix-Flores). 0016-2361/$ - see front matter Published by Elsevier Ltd. doi:10.1016/j.fuel.2008.07.023

to keep the desired activity (catalyst residence time in the dense bed is around 9 min). Several studies have been focused on modeling of cracking reactions kinetics. Some of them proposed kinetic models based on empirical correlations with a limited range of applicability [27]. Most of kinetic studies consider the lumping technique, which establishes that various chemical species can be lumped according to similar characteristics (e.g. the boiling point). The rst model based on this technique considered two lumps [8]: gasoil as the reactant, and a product containing gasoline, gases and coke. A three-lump model consisting in gasoil, gasoline, and light gases and coke was developed [9]. The gasoil cracking is represented with second-order kinetics although coupled to rst-order activity decay, whereas gasoline conversion is represented with rst-order kinetics. Using different gasoil streams, this model detected a signicant effect of the feedstock quality on reaction rate constants. Others have [10,11] separated the third lump in Naces model to produce a four-lump model to ratify that rate constants are strongly dependent on feedstock quality. Several kinetic models considering ve lumps have been published: Larocca et al. [12] split oil feed into parafn, naphtha and aromatic (PNA) fractions; Corella et al. [13] separate oil feed in two gasoil fractions. A six-lump model that was also applied to residual oil cracking [14]. Then, the lumping approach has been extended to include more lumps: eight lumps [15], 11 lumps [16], 13 lumps [17], and 19 lumps [18]. One of the most widely used lumping models is the 10-lump model proposed by Jacob et al. [19]. This model takes into account PNA fractions for two feed fractions as well

170

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178

Nomenclature A b A j Aji aji BU C C0 C Pg C Pl C Ps C Pv C Pv CATOIL Eji Fg Fs Fv Fv Fj fj T Bi GA HCO HN DHF DHv kij ; kji kv kvo ^ k j kHCO,BU kHCO,GA kHCO,HN kHCO,LN kHN,BU kHN,GA kHN,LG kLCO,BU riser cross section area, m2 overall pre-exponential factor dened by Eq. (18), [m3 of gas/(kg of catalyst s)][m3 of gas/kg mol]n1 pre-exponential factor for reaction constants, [m3 of gas/(kg of catalyst s)][m3 of gas/kg mol]n1 terms ruled by Eqs. (16)(21) butane-butylenes (C4) concentration, kg mol/m3 of gas gas concentration at z =0, kg mol/m3 of gas gas feed average heat capacity, kJ/(kg K) liquid feed average heat capacity, kJ/(kg K) catalyst average heat capacity, kJ/(kg K) dispersion steam average heat capacity, kJ/(kg K) coke average heat capacity, kJ/(kg K) catalyst/oil ratio, kg of catalyst/kg of oil activation energy, kJ/kg mol feed oil owrate, kg/s catalyst owrate, kg/s steam owrate, kg/s coke owrate, kg/s cumulative distribution function probability density function as a function of T Bi , fj T Bi  aji gases (butanes C4 + light gases C1C3) heavy cyclic oil heavy naphta heat of formation, kJ/kg mol heat of feed vaporization, kJ/kg reaction rate constant, [m3 of gas/(kg of catalyst s)][m3 of gas/kg mol]n1 Rate constant of coke formation, kg/kg pre-exponential factor in Arrhenius-type equation for coke formation, kg/kg overall cracking constant for lump j kinetic constant for cracking from heavy cyclic oil to butane-butylenes in a six-lump model kinetic constant for cracking from heavy cyclic oil to gases in a six-lump model kinetic constant for cracking from heavy cyclic oil to heavy naphta in a six-lump model kinetic constant for cracking from heavy cyclic oil to light naphta in a six-lump model kinetic constant for cracking from heavy naphta to butane-butylenes in a six-lump model overall cracking constant from heavy naphta to gases (light gases + butanes) in a six-lump model kinetic constant for cracking from heavy naphta to light gases in a six-lump model kinetic constant for cracking from light cyclic oil to butane-butylenes in a six-lump model kinetic constant for cracking from light cyclic oil to gases in a six-lump model kLCO,HN kinetic constant for cracking from light cyclic oil to heavy naphta in a six-lump model kinetic constant for cracking from light cyclic oil to light kLCO,LN naphta in a six-lump model overall cracking constant from light naphta to gases kLN,GA (light gases + butane-butylenes) in a six-lump model kinetic constant for cracking from light naphta to bukLN,BU tane-butylenes in a six-lump model kinetic constant for cracking from light naphta to light kLN,LG gases in a six-lump model LCO light cyclic oil LG light gases LN light naphta M molecular weight, kg/kg mol n apparent order of reaction N number of lumps in a kinetic model NESS normalized error sum of squares number of overall lumps, NO = 4 NO number of kinetic parameters involved in a cracking Npar model heat of reaction, kJ/(m3 of gas s) QR R reaction rate, kg mol/(m3 of gas s) reaction rate, kg mol/(kg of catalyst s) R0 universal constant of gases Ru boiling point temperature of oil fraction, K TB boiling point temperature of FCC feed, K T Bf T Bi; min ; T Bi;max minimum and maximum boiling point temperatures for the lump formed i, K preheat feed temperature, K Tf mix temperature in the vaporization zone, K T mix riser temperature, K T rs hot catalyst temperature, K Ts steam temperature, K Tv gas velocity, m/s ug x experimental or calculated values employed in non-linear regression overall conversion, kg/kg XO y mole fraction, kg mol/kg mol z axial position, m kLCO,GA Greek symbols scale parameters of the beta distribution coke on catalyst, kg of coke/kg of catalyst e void fraction, m3 of gas/m3 total / deactivation function m stoichiometric coefcient qg ; qs gas and catalyst densities, kg/m3 W slip factor, non-dimensional

a; b vrs

as heavy and light cyclic oils and the authors have indicated that rate constants are independent of the feedstock quality. The main disadvantage of all models based on lumping is that a large number of experiments must be carried out to predict their kinetic constants. In last years, novel contributions have taken into account adsorption and intra-crystalline diffusion effects in kinetic modeling [2022]. As a result, non-linear mass balances account for both diffusion constraints experienced by hydrocarbon species while evolving in the zeolite pore network and the effect of the system pressure on the apparent adsorption selectivity.

Another possible approach to model cracking reactions is based on the kinetic representation at molecular level. These models use vectors to include the representation of most chemical species involved in complex mixtures [23]. In a similar study, kinetic models are based on carbocation chemistry [2426]. The main advantage of these models is that they provide a detailed knowledge of the performance of catalytic reactions. Unfortunately, these techniques require a much larger database of kinetic information than the database required in the lumping technique and this information is unavailable in the open literature. Recently, Gupta et al. [27] proposed an elementary reaction

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178

171

scheme whereas the kinetic parameters are estimated using a semi-empirical approach based on normal probability distribution. Reaction rate constants are typically calculated from experimental data. Considering that catalytic reactions occurs in irreversible reaction steps [19,28], then the maximum number of reaction pathways and, consequently, of rate constants, is

Number of pathways

N
2

N! NN 1 2!N 2! 2

bles, respectively. Catalyst heat capacity depends linearly on alumina composition and its function is proprietary, provided by manufacturer. A conclusion achieved in several studies on ow-pattern models indicates that the plug ow reactor model for both gas and solid phases provides good approximations in the riser unit [13,3941]. As a result of this assumption, steady-state mass balance for all chemical species involved as reactants or products is written as

ug

The computational effort to determine the rate constants increases when all reaction pathways are considered. For instance, the maximum number of rate constants in the 10-lump model is calculated as 45. However, the number of kinetic parameters is reduced because some reaction pathways are considered as improbable and its contribution in the reaction network is neglected [19,10]. For example, the 10-lump model proposed by Jacob et al. [19] contains no more than 20 reaction pathways. Furthermore, Arbel et al. [29] consider that some kinetic rate constants from the Jacobs model are numerically identical and the number of constants can thus be reduced to 14. Even under the above consideration, the number of rate constants remains large. Thus, a comprehensive kinetic model is proposed in this paper to estimate kinetic constants when cracking reactions representation is based on the lumping technique. This approach characterizes the FCC products in detail, prevents excessive reduction of rate constants and reduces computational effort in the numerical evaluation of kinetic parameters. 2. The riser model A model of the riser is developed in this section since the cracking reactions are carried out within it. The riser is fed at the bottom of the unit with an atomized gasoil, the regenerated catalyst, and steam streams. The gasoil is almost instantaneously vaporized and mixed with the other streams. In order to model this stage, two basic assumptions are adopted here: instantaneous vaporization occurs as a result of an intimate gassolid mixing and chemical reactions in the vapor phase due to thermal cracking can be neglected at low temperatures [3033]. The heat carried in the riser via the regenerated catalyst is consumed to vaporize the feedstock and to sustain the cracking reactions. Both vaporization and cracking processes are endothermic so that the hot catalyst provides energy to make the oil achieve the boiling point, vaporize and, in the vapor phase, get the mix temperature. Thus, the following equation can be obtained from an overall energy balance at the feeding point:

dC j qs 1 e/ 0 Rj ; e dz 1 with e q Fs 1 w qg F g
s

j 1; 2; . . . ; N

3a 3b 3c

ug

Fg Aq g e

where ug is the gas velocity, C j is concentration of compound j, qs is the catalyst density, qg is the gaseous mixture density e is void fraction, w is the ratio between gas and solid velocities in the riser (slip factor, in this work, no slip between phases is considered: w = 1), A is the riser cross section area, / is a deactivation function, R0j is the reaction rate for chemical species j, and z is the axial position. More details related to riser modeling have been provided in a previous work [1]. The deactivation function is modeled here via the exponential law proposed in [42]. The boundary condition for the above equation is

C j 0 C O yj ;

j 1; 2; . . . ; N

where CO represents the gas concentration at the riser inlet and mole fractions yi, which are evaluated employing the ASTMD1160 distillation curve of the FCC feed. The ASTM-D1160 is converted into TBP curve according to an algorithm based on procedures detailed in [35] and recently reviewed by Ahmed [43]. The reaction rates and the number of lumps involved in catalytic cracking reactions depend on the selected kinetic model. In a generic sense this reaction rate is

R0j

j 1 X i1

2 vij kij C n i

N X ij1

1 kji C n j ;

j 1; 2; . . . N

where

vij

Mi Mj

T mix

F s C Ps T s F v C Pv T s F g C Pg T Bf F v C Pv T v F g C Pl T Bf T f F g DHv F s C Ps F v C Pv F g C Pg F v C Pv 2

where F s , F v , F g and F v are the catalyst owrate, coke owrate, feed owrate, and the steam owrate, respectively; C Ps , C Pv , C Pg , C Pl , and C Pv represent the average heat capacities for the catalyst, coke, liquid feed oil, gas feed oil and the dispersion steam, respectively; T s , T Bf , T v and T f are the temperatures for catalyst, feed oil at the boiling point, steam and the feed oil leaving the pre-heater; and DHv is the heat of vaporization of gasoil fed to the unit. Heat capacities for hydrocarbons were calculated via empirical correlations based on thermophysical databases [34,35]. Molecular weights were calculated by using classical correlations [3638]. Gasoil heat of vaporization was calculated employing the original PengRobinson EOS. Heat capacities for steam and coke, were obtained from saturated steam and pure carbon (coke is assumed as graphite) ta-

being kij the rate constants, vij the stoichiometric coefcients and M the molecular weight. It should be mentioned here, kinetic constants dened above are no intrinsic rates because of assumptions taken in this study make difcult to distinguish between diffusion, adsorption or selective catalyst deactivation phenomena. As a result, this kinetic approach is useful for calculating observed kinetic constants. Eq. (5) indicates that lump j is being generated and cracked as a function of the postulated reaction network with an apparent order of reaction n, whereas the rate constants depend exclusively on reaction temperature. The temperature effect on kinetic constants is represented by an Arrhenius-type expression [10,11,14,19]. Thus, a reaction rate constant, where a lump j is cracked into a lump i, can be written as

kji Aji e RT

Eji

where Aji is the pre-exponential Arrhenius factor, Eji is the activation energy and R is the universal constant of ideal gases. In particular, coke is a solid with pseudo-graphite nature so that it cannot be described in terms of the boiling point and a semi-

172

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178

empirical relationship based on experimental observations is preferred. Thus, coke is related to the overall gasoil conversion by [44]:

vrs

  kv XO CATOIL 1 X O

where vrs is the coke in the riser, kv is the rate constant for the coke formation, CATOIL is the catalyst-oil ratio, and X O is the overall conversion. The rate constant for coke formation can be represented via an Arrhenius function to take into account the temperature dependence [12,45]. This expression assumes catalyst deactivation is no selective for each lump. Considering the above-described facts the derivative of Eq. (8) with respect to z can be written as
Ev

equation are expressed here by a probability distribution function. The independent variable is selected here as the boiling point since oil feeds are generally dened according to boiling point cuts. Considering N lumps where the rst lump remains as the heaviest lump and the Nth-lump is the lightest compound, the rst lump has no generation side for the reaction and the Nth-lump has no disappearance term. Expanding Eq. (5) and considering the existence generation and disappearance terms, the following table can be generated:
Lump j Generation R1 1 R2 m1;2 k1;2 C n 1 . . . . . . nN 2 1 RN1 m1;N1 k1;N1 C n 1 ... mN 2;N 1 kN 2;N 1 C N 2 n1 nN 1 RN m1;N k1;N C 1 mN1;N kN1;N C N1 Disappearance 1 k1;2 ... k1;N C n 1 2 k2;3 ... k2;N C n 2 . . . N 1 kN1;N C n N 1 15

kv eRu T dvrs /vrs o dz CATOIL 1 X O 2

where Ev is the energy of activation for coke formation. This parameter depends on feedstock and catalyst type. The endothermic nature of cracking reactions produces an appreciable temperature drop along the riser length. To assess this temperature drop, an energy balance for an adiabatic operation is established using similar assumptions adopted for species balances with respect to the ow-pattern, e.g. plug ow-pattern:

The set of rate constants for the disappearance term on the right-hand side is

^  k j

N X ij1

kji

16

qg C Pg ug

dT rs Q R dz

10

where subscript i refers to the product formed from lump j. The summation above is called the overall cracking constant for lump j. This term is equal to zero for the Nth-lump due to the no-existence of any lump with lower molecular weight. Using the Arrhenius-type of equation in Eq. (16) the following equation is obtained:

where QR represents the heat required for catalytic cracking that can be calculated via the heat of formation DHF for each reaction pathway according to

^ k j

N X ij1

Aji e

Eji Ru T

17

QR

N X i1

R i DH F i

11

The heat of formation for each lump is calculated with the standard enthalpy of formation and heat capacity of the lumps involved in the reaction pathway i. Heat of formation for hydrocarbons were calculated employing empirical correlations based on thermophysical databases from [35]. Combining Eqs. (5), (10), and (11), the following equation is obtained:

It should be noted that activation energies remain constant for each lump and they are in fact known parameters. Thus, the pre-exponential factors are represented with a probability distribution function. Dening the overall pre-exponential factor as,

bj  A

N X ij1

bj Aji  A

N X ij1

aji

18

then the summation becomes equal to unity so that each value can be referred as a proportion, i.e.
N X ij1

qg C Pg ug
with,

N 1 N X X dT rs 1 e n C i i kij DHF ji qs / e dz i 1 ji1

12

aji 1

19 20

b j aji Aji  A 13

DHF ji mij DHF j DHF i


and boundary condition,

The above proportion is dened as a cumulative distribution for continuous variables:


N X ij1

T rs 0 T mix

14

aji F j T Bi

21

It should be observed that the mixing temperature and other variables evaluated in the vaporization zone such as density, void fraction and gas velocity also constitute boundary conditions at the riser inlet. 3. The kinetic model This novel approach for the kinetic model is supported on the assumption that intrinsic kinetic constants could be represented by a continuous distribution function by selecting an independent stochastic variable. This idea is supported on previous works where rate constants for desulphurization of diesel fuel have been expressed in terms of the C distribution [46] and normal distribution function [27]. Hence frequency factors for Arrhenius-type

A cumulative distribution is equivalent to the area under the distribution function f dened in the positive range T Bi; min 6 T Bi 6 T Bi; max :

F j T Bi ; T Bi; min ; T Bi; max

TB

i; max

TB

fj T Bi dT Bi

22

i; min

Integration is recognized as the summation of the distribution function


N X ij1

aji

N X ij1

f j T Bi

23 24

aji fj T Bi

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178 Table 1 Kinetic parameters to be estimated with and without a function distribution Lumps Kinetic parameters using Eq. (1) 6 15 190 595 Kinetic parameters using the distribution function Npar = (N 1) + 2 5 7 21 36 Experimental values 44 66 220 385 NESS

173

4 6 20 35

0.24 0.35 0.23 0.22

Combining Eqs. (17), (20) and (23), (24):

b j aji e Ru T A b j fj T B e Ru T kji A i

Eji

Eji

25

the distribution function would be a function for continuous variables dened in a given positive range T Bi; min 6 T B 6 T Bi;max . In this paper, the beta distribution function has been selected because it provides large exibility for data tting and it has been successfully used in several science elds [47]. The beta distribution is a twoshape parameter continuous function with an interval dened by a minimum and maximum value and a polynomial nature, which is given by

fj T Bi ; aj ; bj

1 T Bmax T Bmin T Bi T Bi;min Caj bj Caj Cbj T Bi;max T Bi;min

!aj 1

T Bi;max T Bi T Bi;max T Bi;min

!bj 1

26
in the range T Bmin 6 T B 6 T Bmax . The cumulative function will be the result of using Eq. (26) in the integration according to the selected range for the boiling point. It has mentioned above that Eq. (1) can give an estimation of the number of kinetic parameters required to model the cracking reactions. Using this approach, the number of parameters required can be calculated by

Npar N 1 2

27

since there are (N 1) pre-exponential factors and two parameters for the distribution function. The advantage of using the approach in this paper becomes greater when the number of lumps increases, Table 1. In cracking, reaction rates are generally considered as pseudorst-order for light lumps and pseudo-second-order for heavy lumps [8,48]. However, pseudo-rst-order is adopted here when more than 20 lumps are used in the model because the expected behavior is like elementary reactions rather than using multiple steps. The operating conditions used in the evaluation of rate constants for the three-lump kinetic model, gasoil lump is cracked into gasoline, light gases and coke [9], are similar to those used in [8]. The utilization of an apparent reaction order equal to 2 is feasible because various elementary steps are represented when the overall conversion of FCC feed is considered. Summarizing, the selection of apparent reaction orders may be founded on previous studies but reasonable assumptions may properly justify any other choice. 4. Case studies denition Several kinetic models with a marked degree of complexity are used to demonstrate the advantages of this novel approach. The choice of a kinetic model depends on particular requirements. For example, a kinetic model containing a large number of lumps could permit a detailed characterization and, consequently, a bet-

ter representation of the products quality. It should be recognized that a kinetic model with these features implies a greater computational effort. This could be unnecessary if our target is to determine responses in unsteady state of some of the state variables. A more detailed kinetic model is required when it is used for optimization purposes such maximizing olen production or the octane number of gasoline. In reneries, each lump quality is governed by the operating conditions and characteristics of the main fractionators. As a consequence, the number of lumps established in most kinetic models is commonly related with the number of boiling point cuts considered in the fractionators. In this section, a modication for the four-lump model published by Yen [10] and a modication for the six-lump model presented by Takatsuka et al. [14] are studied. In addition, two more complex models are also proposed here: a 20-lump model and a 35-lump model. In chemical kinetics, it has been considered that lumps boiling up to the boiling temperature of n-C12 are cracked with an apparent order equal to 2. On the other hand, lighter lumps will be cracked with rst-order kinetics. The modied four-lump model consists of: unconverted gasoil C 12 , heavy naphtha (approx. C13C14), light naphtha (approx. C5C12), and light gases (C1C4). The modied six-lump model is dened by: heavy cycle oil C 21 , light cycle oil (approx. C15C21), heavy naphtha (approx. C13C14), light naphtha (approx. C5C12), butanes-butylenes (C4) and light gases (C1C3). These denitions are referred to approximate boiling point cuts though the range of boiling points depends on each operating case. The 20-lump model is an extension of the modied six-lump model proposed here. This 20-lump model takes into account four subdivisions for each FCC liquid product: heavy cycle oil (HCO), light cycle oil (LCO), heavy naphtha (HN), and light naphtha (LN). This subdivision is based on respective distillation curve cut points. The Iump called light gases (LG) is subdivided into three lumps: methane, ethaneethylene and propanepropylene. The 20th lump is butanes-butylenes (C4). In the 35-lump model it is considered that a single compound, in this work an n-alkane, represents a single lump. The result is that the reactive mixture contains 35 lumps, each one represented by a corresponding n-alkane. Lumping for 35-lump model was performed via distillation curve splitting as a function of the boiling points of respective n-alkane compound. Finally, coke is not directly considered in the reaction network and Eq. (9) rules its formation rate. 5. Results and discussion Numerical experiments were carried out to establish important issues related to the FCC kinetic models. These experiments include parameters estimation for selected kinetic models and the effect of particular variables such as changes in the feedstock and operating conditions. 5.1. Kinetic parameter estimation Kinetic parameters are rstly estimated to detect the tting capability of lumping models. A database containing experimental values based on the real operation of a FCC unit is used in this work. Because of the plant database is proprietary, only two operating cases are shown here (Table 4). Typical feedstock characteristics, dimensions and operating conditions of the FCC unit are given in Table 2. The general regression package GREG [49] was employed for all parameter estimations. In addition, the public domain FORTRAN code DASSL [50] was used to solve the set of differential-algebraic equations. More details related to computational algorithm have been given in a previous work [1].

174

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178

Table 2 Typical feedstock characteristics, dimensions and operating conditions of main FCC equipment Characteristic Value Min Feedstock (Maya crude oil) Specic gravity @ 15/15 C Molecular weight, kg/kg mol Mean average boiling point, K Sulfur content, wt% Item FCC unit dimensions Riser Length, m Diameter, m Stripper-reactor Height, m Diameter, m Catalyst holdup, Tons Regenerator Height, m Diameter, m Catalyst holdup, Tons Variable 0.880 280 651 1.7 Value Max 0.923 360 690 2.3

40 1.3 3.1 6 42 12 10.2 198 Value Min Max 490 12.0 810 973

Operating conditions Preheat feed temperature, K Catalyst/oil ratio, kg/kg Riser outlet temperature, K Dense bed temperature, K

430 7.5 790 912

Table 3 Activation energies employed in this study Pathway Heavy cyclic oil-light cyclic oil Heavy cyclic oil-gases Gasoil(cyclic oils)-naphtas Naphtas-light gases Source [11] [51] [51] [51] Value (kJ/kg mol) 44,000 89,000 68,250 52,700

Table 4 FCC operating cases used in this study Light gasoil Physical properties Specic gravity (15/15 C) ASTM D 1160, K T10 T30 T50 T70 T90 Molecular weight Refractive index Operating conditions Riser inlet temperature, K Preheat feed temperature, K Catalyst/oil ratio Feed owrate, kg/s 0.865 530 592 651 702 742 296 1.492 815 486 8.7 51 Heavy gasoil 0.911 617 660 702 743 782 356 1.510 815 486 8.7 51

Models containing four, six, 20 and 35 lumps were selected in this rst set of numerical experiments. Plant data include eleven operating cases so that the number of observations, i.e. experimental values, is equivalent to eleven times the number of lumps, Table 1. In all minimizations, the energies of activation employed in the pre-exponential factors estimation were adapted from Lee et al.

[11], Table 3. To establish a similar basis of comparison, results for the four models were grouped so that four overall lumps were congured as: gasoil (GOL, C 12 ), light naphtha (LN, C5C12), gases (GA, C1C4) and coke. In this form, the minimization procedure yields similar values in the normalized errors of the sum of squares for both four-lump and 20-lump models, Table 1 and Fig. 1. Results indicate no meaningful difference in the predicted values though it is obvious that using 20 lumps will provide more details of the FCC reactions. A better prediction of gasoline and gasoil production is observed when the 35-lump model is used whereas prediction of gases and coke is better with the 6 and 20-lump models. The distribution function strongly depends on the interval of boiling temperatures for each type of feedstock. Hence it is expected that values of kinetic constants will also depend on the type b j have remained almost of feedstock (Eq. (26)); however, a, b, and A unchanged in all cases studied. This apparent invariability is only valid for a specic FCC catalyst. For two types of feedstock considered in this analysis (Table 4) and using a catalyst with selectivity to gasoline, it was observed that the difference in numerical values of kinetic constants is negligible since these differences represented 0.10.6%. However, this apparent non-difference in the kinetic constants is sufcient to generate different axial proles of FCC products in different feedstock types. Average values for kinetic constants at 815 K and 1% of condence are shown in Table 5 where it can be observed that cracking ratios for models of four and six-lumps are very similar. Physical units depend on the order of reaction. The cracking constants kHN,GA and kLN,GA behave as overall cracking constants since kHN,GA % kHN,BU + kHN,LG and kLN,GA % kLN,BU + kLN,LG. On the other hand, in the model containing six lumps, the relationships kHCO,BU/kHCO,GA = 1.87, kLCO,BU/kLCO,GA = 1.82 and kHN,BU/kHN,GA = 1.77 indicate that the rate of the cracking of heavy species to light species is practically a constant. The relationship kHCO,HN/kLCO,HN = 1.8 indicates that HCO strongly contributes in producing HN but the relation kHCO,LN/kLCO,LN = 1.0 means that both cyclic oils contribute to produce LN. A similar analysis shows that LCO contributes in larger proportion to produce light species. Pre-exponential factors of the kinetic constants were calculated for the heavy gasoil. Fig. 2a shows the numerical values for 595 kinetic constants for the model with 35 lumps. Distribution functions, denoted with letter C, refer to the frequency factors for light lumps. It is appreciated in this case that lumps having less than 10 carbons tend to be disintegrated to produce lumps with less than ve carbons. Letter B in Fig. 2a, stands for the heaviest lumps and it is observed that they are distributed within a larger range of carbon numbers. Letter A indicates the reactive behavior of intermediate lumps, which tend to crack into close lumps that can also be overcracked to produce light lumps. In Fig. 2b, the prediction capability of the novel approach is demonstrated. Predicted yields for lumps with carbon number greater than ve agree reasonably well with plant data, especially those lumps constituting the light gasoline (C5C12). However, the lumping technique employing a 35-lump kinetic model offers reasonable predictions for lumps with carbon number lower than ve but it fails signicantly in prediction of C4 yield. Axial proles of FCC products were calculated using two kinetic models: a 20-lump model and a six-lump model. A light gasoil was used for the simulation and conversion predictions indicate less that 0.3 wt% at 10 m high. The conversion at the riser outlet becomes 69.3 wt% in both cases though the 20-lump model improves gases prediction. This improvement is because the 20-lump model disaggregates the C1C3 lump of six-lump model to have methane, ethaneethylene and propanepropylene. It should be observed the double functionality of heavy naphtha since it starts as a product of the cracking reactions until a maximum yield is achieved so that heavy naphtha behaves as a reactant.

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178

175

a) 4 lumps
50

b) 6 lumps
50

40 Plant Yields [% wt]

40 Plant Yields [% wt]

Gasoil Gasoline Gases Coke

30

30

20

20

10

10

0 0 10 20 30 40 50 Calculated Yields [% wt]

0 0 10 20 30 40 50 Calculated Yields [% wt]

c) 20 lumps
50

d) 35 lumps
50

40 Plant Yields [% wt] Plant Yields [% wt]

40

Gasoil Gasoline Gases Coke

30

30

20

20

10

10

0 0 10 20 30 40 50 Calculated Yields [% wt]

0 0 10 20 30 40 50 Calculated Yields [% wt]


C 12 , gasoline or light naphta (C5C12), and gases (C1C4).

Fig. 1. Minimization results for selected kinetic models. Gasoil

Table 5 Averaged kinetic constants of four-lump model (a = 0.2 0.01, b = 1.20 0.02) and six-lump model (a = 4.6 0.04, b = 1.6 0.02) Four lumps kGOL,HN = 5.60E2 kGOL,LN = 3.01E1 kGOL,GA = 8.64E2 Six lumps kHCO,LCO = 8.24E1 kHCO,HN = 3.01E1 kHCO,LN = 3.78E1 kHCO,BU = 7.22E4 kHCO,LG = 3.87E4

kHN,LN = 6.37E3 kHN,GA = 1.24E1

kLN,GA = 5.26E4

kLCO,HN = 1.67E1 kLCO,LN = 3.81E1 kLCO,BU = 8.67E4 kLCO,LG = 4.76E4

kHN,LN = 5.42E1 kHN,BU = 1.09E1 kHN,LG = 6.15E2

kLN,BU = 2.61E4 kLN,LG = 3.11E4

kBU,LG = 2.51E2

5.2. Effects of feedstock and operating conditions A second set of experiments were carried out to establish the effect of several important variables on the performance of the models. One of the most essential goals in simulation of FCC process is the knowledge of the effect of feedstock on FCC yields. Hence two feedstock exhibiting considerable physical differences

have been studied and their impact on the FCC process is summarized in Table 6 where both the 20-lump and the 35-lump models were considered. The most relevant difference found in this simulation is that the overall conversion achieved with the heavy gasoil is greater than that with light gasoil. In contrast, gases and coke yields are greater for light gasoil. As a consequence, the degree of cracking severity should be diminished when light gasoil is pro-

176

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178

a) Pre-exponential factors
1.0 Pre-exponential factor, aji 0.8 0.6 0.4 0.2 0.0 5 10 15 20 Carbon number 25 30 B LN HN LCO HCO

b) Comparison between results of this work and plant data


0.10 Plant data This work

C
Weigth fraction

0.08 0.06 0.04 0.02 0.00 0 5 10

15 20 25 Carbon number

30

35

Fig. 2. Results of 35-lump model. (a = 5.6 0.04, b = 0.5 0.02).

Table 6 Simulation results using two feedstocks Results Experimental data Kinetic model 20 Lumps Light Riser outlet temperature, K FCC yields, wt% Coke Gasoil Gasoline Gases 797 6.1 27.2 46.4 20.3 Heavy 794 5.0 29.0 48.4 17.6 Light 804 6.0 30.7 45.2 18.1 Heavy 800 5.3 27.3 49.2 18.2 35 Lumps Light 804 6.3 33.2 40.9 19.6 Heavy 801 5.3 28.4 48.4 17.9

a) Light gasoil
40 35 30 Riser Length [m]

b) Heavy gasoil
40 35 30 Riser Length [m] 25 20 15 10 5 0

wt / wt
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10

25 20 15 10 5 0 50 100 150 200 250 300 350 Molecular Weight [kg/kg mole]

50

100

150

200

250

300

350

Molecular Weight [kg/kg mole]

Fig. 3. Weight distribution of hydrocarbons for a 20-lump model.

cessed. The risk of overcracking could then be reduced. At the bottom of the riser, most of light feed oil is distributed from 200 to 280 kg/kg mol and it is gradually cracked into gasoline and light gases for the 20-lump model, Fig. 3. Gasoline yield is typically observed in the range of 80120 kg/kg mol and light gases in the range of 2080 kg/kg mol. The heaviest feed oil is dispersed from

250 to 400 kg/kg mol and, in this case, cracking reactions lead quickly to more valuable products, particularly light gasoline. On the other hand, Fig. 4 depicts the weight distribution using a 35lump model where differences in performance of the 20-lump and 35-lump models are evident. A signicant portion of light feed oil close to 200 kg/kg mol could not be cracked into valuable prod-

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178

177

a) Light gasoil
40 35 30

c) Heavy gasoil
40

wt / wt
35 30

25 20 15 10 5 0 50 100 150 200 250 300 350 400

25 20 15 10 5 0 50 100 150 200 250 300 350 400

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10

Riser Length [m]

Molecular Weight [kg/kg mole]

Riser Length [m]

Molecular Weight [kg/kg mole]

Fig. 4. Weight distribution of hydrocarbons for a 35-lump model.

6. Conclusions
Light naphta [% wt] 48 46 44 42 40 38

wt %

38 40 42 44 46 48

Pr

eh ea 500 tT em 400 14 pe 10 12 ra 300 8 6 tu 4 re g] tio [kg/k [K t/Oil Ra ] Catalys

It is always possible to propose a model in accordance to the available kinetic information but experimental data is usually limited. The proposed representation has been capable to successfully predict the performance of cracking reactions for gasoils in a wide range of feedstock and operating conditions. Thus, the novel approach developed in this work shows to be highly exible to include kinetic models exhibiting different degree of complexity. According to the results, it is possible to obtain a detailed kinetic model without a large kinetic database of experiments but using commonly available information in reneries. In addition, the kinetic representation based on a probability distribution has permitted to reduce substantially the computational effort involved in the numerical evaluation of intrinsic kinetic parameters and it prevents excessive reduction of reaction pathways. Acknowledgments

Fig. 5. The effect of riser operating conditions on light naphta yields. Heavy gasoil using a six-lump model.

ucts. This fact is reected in the FCC yields shown in Table 6. It is worth mentioning that heavy gasoil is adequately converted into gasoline and gases. In a nal set of numerical simulations, the impact of riser operating conditions on FCC yields in steady-state with constant regenerator operating conditions is studied. A six-lump model is used in order to illustrate our results for a heavy gasoil. The effects of both feed preheat temperature and catalyst/oil ratio on light gasoline yields is given in Fig. 5. A maximum on gasoline yield appears when the feed preheat temperature is 343 K and catalyst/oil ratio is 10 to make gasoline yield going up to almost 50%. The position of this point depends markedly on the feedstock, type of catalyst, regenerator operating conditions and quality of the available experimental information. Unfortunately, a suitable comparison between kinetic constants showed in this work and other constants available in the open literature cannot be carried out since several factors affect the quality and reproducibility of experimental results. However, some valuable studies have been published related to the effect of operating conditions on FCC yields [29,52]. The weight distribution of gasoil cracking obtained in this work reasonably agrees with those results published by Liguras and Allen [53]. Concentration proles obtained in this work also agree with those presented by several authors [24,29,40,41,54].

The authors would thank to the CONACyT and COSNET for the economical support for this research. References
[1] Hernndez-Barajas JR, Vzquez-Romn R, Salazar-Sotelo D. Multiplicity of steady states in FCC units: effect of operating conditions. Fuel 2006;85: 84959. [2] Reif HE, Kress RF, Smith JS. How feeds affects cat cracker yields? Petrol Rener 1961;40:237. [3] Kurihara H. Optimal control of uid catalytic cracking processes, PhD dissertation, MIT, USA, 1967. [4] Ewell RB, Gadmer G. Design cat crackers by computer. Hydrocarbon Process 1978:12534. [5] Lin TD. FCCU advanced control and optimization. Hydrocarbon Process 1993:10714. [6] McFarlane RC, Reinman RC, Bartee JF, Georgakis C. Dynamic simulation for model IV uid catalytic cracking unit. Comput Chem Eng 1993;17:275300. [7] Lappas AA, Iatridis DK, Vasalos I. Production of reformulated gasoline in the FCC unit. Effect of feedstock type on gasoline composition. Catal Today 1999;50:7385. [8] Blanding FH. Reaction rates in catalytic cracking of petroleum. Ind Eng Chem 1953;45:118697. [9] Nace DM, Voltz SE, Weekman VW. Application of a kinetic model for catalytic cracking. Ind Eng Chem Process Des Develop 1971;10:5307. [10] Yen LC. Kinetic modelling of uid catalytic cracking. AIChE Spring Nat Meet, April 26, Houston, Texas, 1987. [11] Lee LS, Chen YW, Huang TN, Pan WY. Four-lump kinetic model for uid catalytic cracking process. Canadian J Chem Eng 1989;67:6159.

178

J.R. Hernndez-Barajas et al. / Fuel 88 (2009) 169178 [33] Mirgain C, Briens C, Pozo MD, Loutaty R, Bergougnou M. Modeling of feed vaporization in uid catalytic cracking. Ind Eng Chem Res 2000;39:43929. [34] Maxwell JB. Data book on hydrocarbons application to process engineering. Princeton, New Jersey: Van Nostrand; 1950. [35] American Petroleum Institute (API). Technical data book petroleum rening. New York: EUA; 1997. [36] Sim WJ, Daubert TE. Prediction of vaporliquid equilibria for undened mixtures. Ind Eng Chem Process Des Dev 1980;19:356. [37] Riazi M, Daubert T. Simplify property predictions. Hydrocarbon Process 1980;59:1156. [38] Riazi M, Daubert T. Prediction of thermophysical properties of petroleum fractions. Ind Eng Chem Res 1987;26:7559. [39] Froment GF, Bischoff KB. Chemical reactor analysis and design. John Wiley & Sons; 1990. [40] Theologos KN, Markatos NC. Advanced modeling of uid catalytic cracking riser-type reactors. AIChE J 1993;39:100717. [41] Theologos KN, Nikou ID, Lygeros AI, Markatos NC. Simulation and design of uid catalytic cracking riser-type reactors. AIChE J 1997;43:48694. [42] Froment GF. The modeling of catalyst deactivation by coke formation. In: Bartholomew CH, Butt JB, editors. Catalyst deactivation 1991. Elsevier; 1991. p. 5383. [43] Ahmed T. Equations of state and PVT analysis. Applications for improved reservoir modeling. Houston, USA: Gulf Publishing; 2007 [Chapter 2]. [44] Krambeck FJ. Continuous mixtures in uid catalytic cracking and extensions. Mobil workshop on chemical reaction of complex mixtures. New York: Van Nostrand; 1991. [45] Corella J, Bilbao R, Molina JA, Artigas A. Variation with time of the mechanism, observable order and activation energy of the catalyst deactivation by coke in the FCC process. Ind Eng Chem Process Dev 1985;24:62536. [46] Inoue S, Takatsuka T, Wada Y, Hirohama S, Ushida T. Distribution function model for deep desulfurization of diesel fuel. Fuel 2000;79:8439. [47] Gupta PL, Gupta RC. The monotonicity of the reliability measures of the beta distribution. Appl Math Lett 2000;13:59. [48] Weekman VW. A model of catalytic cracking conversion in xed, moving and uid bed reactors. Ind Eng Chem Process Des Develop 1968;7:905. [49] Stewart WE, Caracotsios M, Srensen JP. Parameter estimation from multiresponse data. AIChE J 1992;38:64150. [50] Brenan KE, Campbell SL, Petzold LR. Numerical solution of initial-value problems in differential-algebraic equations. New York: North-Holland; 1989. [51] Ancheyta JJ. Estudio cintico de las reacciones de desintegracin cataltica de gasleos, PhD thesis, Universidad Autnoma Metropolitana, 1998. [52] Han IS, Chung CB, Riggs JB. Modeling of a uidized catalytic cracking process. Comput Chem Eng 2000;24:16817. [53] Liguras DK, Allen DT. Structural models for catalytic cracking. 2. Reactions of simulated oil mixtures. Ind Eng Chem Res 1989;28:67483. [54] Pitault I, Nevicato D, Forissier M, Bernard J. Kinetic model on a molecular description for catalytic cracking of vacuum gas oil. Chem Eng Sci 1994;49: 424962.

[12] Larocca M, Ng S, de Lasa H. Fast catalytic cracking of heavy gas oils: modeling coke deactivation. Ind Eng Chem Res 1990;29:17180. [13] Corella J, Francs E. Analysis of the riser reactor of a uid cracking unit. Fluid catalytic cracking II, ACS symposium series, vol. 452, 1991. p. 16582. [14] Takatsuka T, Sato S, Morimoto Y, Hashimoto H. Resid FCC reaction model, AIChE Nat Meet, 1984. [15] Sugungun MM, Kolesnikov IM, Vinogradov VM, Kolesnikov SI. Kinetic modeling of FCC process. Catal Today 1998;43:31525. [16] John T, Wojciechowsky BW. On identifying of the primary and secondary products of the catalytic cracking of neutral distillates. J Catal 1975;37: 24050. [17] Gao J, Xu C, Lin S, Yang G, Guo Y. Advanced model for gassolid turbulent ow and reaction in FCC riser reactors. AIChE J 1999;45:232156. [18] Derouin C, Nevicato D, Forissier M, Wild G, Bernard JR. Hydrodynamics of riser units and their impact on FCC operation. Ind Eng Chem Res 1997;36:450415. [19] Jacob SM, Gross B, Voltz SE, Weekman VW. A lumping and reaction scheme for catalytic cracking. AIChE J 1976;22:70113. [20] Al-Khattaf S, de Lasa H. The role of diffusion in alkyl-benzenes catalytic cracking. Appl Catal A: Gen 2002;226:13953. [21] Avila AM, Bidabehere CM, Sedran U. Diffusion and adsorption selectivities of hydrocarbons over FCC catalysts. Chem Eng J 2007;132:6775. [22] Al-Sabawi M, Atias JA, de Lasa H. Kinetic modeling of catalytic cracking of gas oil feedstocks: Reaction and diffusion phenomena. Ind Eng Chem Res 2006;45:158393. [23] Quann RJ, Jaffe SB. Structure-oriented lumping: describing the chemistry complex hydrocarbon mixtures. Ind Eng Chem Res 1992;31:248397. [24] Dewachtere NV, Santaella F, Froment GF. 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;54:365360. [25] Froment GF. Single event kinetic modeling of complex catalytic processes. Catal Rev 2005;47:83124. [26] Kumar H, Froment GF. A generalized mechanistic kinetic model for the hydroisomerization and hydrocracking of long-chain parafns. Ind Eng Chem Res 2007;46:407590. [27] Gupta RK, Kumar V, Srivastava VK. A new generic approach for the modeling of uid catalytic cracking (FCC) riser reactor. Chem Eng Sci 2007;62:451028. [28] Paraskos JA, Shah YT, McKinney JD, Carr NLA. Kinematic model for catalytic cracking in a transfer line reactor. Ind Eng Chem Process Des Develop 1976;15: 1659. [29] Arbel A, Huang Z, Rinard IH, Shinnar R, Sapre AV. Dynamic and control of uidized catalytic crackers. 1. Modeling of the current generation of FCCs. Ind Eng Chem Res 1995;34:122843. [30] Martin MP. Contribution a ltude des lvateurs de Craquage Catalytique. Effects lies a lInjection de la Charge, PhD dissertation, Institut National Polytechnique de Lorraine, France, 1990. [31] Buchanan JS. Analysis of heating and vaporization of feed droplets in uidized catalytic cracking risers. Ind Eng Chem Res 1994;33:310411. [32] Ali H, Rohani S. Dynamic modeling and simulation of a riser-type uid catalytic cracking unit. Chem Eng Technol 1997;29:11830.

S-ar putea să vă placă și