You are on page 1of 11

Chemical Engineering Journal 146 (2009) 466476

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Steady-state particle size distribution modeling of polypropylene produced


in tubular loop reactors
Zheng-Hong Luo , Pei-Lin Su, Xiao-Zi You, De-Pan Shi, Jin-Cheng Wu
Department of Chemical and Biochemical Engineering, College of Chemistry and Chemical Engineering, Xiamen University, Shiming Street 422,
Xiamen 361005, Peoples Republic of China

a r t i c l e i n f o a b s t r a c t

Article history: A comprehensive steady-state population balance model is developed for the particle size distribution
Received 31 July 2008 (PSD) of the polypropylene produced in tubular loop reactors. The model considers the ow type, the
Received in revised form 23 October 2008 polymer particle dynamics, the particle growth and the attrition. Moreover, an empirical single particle
Accepted 26 October 2008
model is used to describe the particle growth under internal and external heat and mass transfer limita-
tions totally. The predicted PSD data obtained under steady-state polymerization conditions agree well
Keywords:
with the actual data collected from industrial scale plant. The model are also used to predict the effects
Polypropylene
of some operational parameters on the polymer PSD produced under steady-state conditions. The results
Tubular loop reactor
PSD
show that the polymer PSD are greatly affected by the feed catalyst size, the feed catalyst PSD, and the
Population balance polymerization temperature.
Steady state 2008 Elsevier B.V. All rights reserved.

1. Introduction predicted the effects of the feed catalyst PSD on the polyolen
PSD produced in a FBR. Soares et al. [14] developed a mathe-
Polypropylene can be produced in various types of reactors, such matical model to analyze the effects of the ideal and non-ideal
as autoclave, continuous stirred tank, uidized-bed reactor (FBR) or reactor residence time distributions on the polymer PSD. Zacca et
tubular loop reactor. The last one is certainly the most important at al. [2] presented a population balance approach modeling of multi-
present [1]. In the tubular loop reactor, small catalyst particles (e.g. stage olen polymerization processes using the catalyst residence
20100 m) react with monomers to form polymer particles in a time as main coordinate. Their work allowed the consideration of
size range of 1005000 m in a liquid phase and the polymer parti- particle size selection effects within the process in a FBR. Khang
cles are produced as a solid suspension in the liquid stream [28]. In et al. [3] developed a population balance approach to model the
addition, the polymerization rate, the cost of post-treatment after effect of non-ideal mixing behavior of solid particles on the poly-
polymeric process and the polymer properties are inuenced by olen PSD in a FBR. Hatzantonis et al. [4] developed a generalized
the polymer particle size distribution (PSD) [68]. Therefore, the steady-state population balance model rigorously accounting for
polymer PSD modeling may be useful for the understanding of the the combined effects of particle growth, attrition, elutriation and
propylene polymerization in tubular loop reactors. Furthermore, a agglomeration in a gas-phase FBR. Yiannoulakis et al. [5] employed
mathematical model is necessary in order to predict the effects of a polymeric ow model to describe the growth rate of a sin-
operation parameters, especially the feed catalyst on the product gle particle under internal and external heat and mass transfer
PSD in tubular loop reactors. limitations in a FBR and this single particle model was solved
In propylene polymerization eld, most of papers published together with a steady-state population balance model to pre-
were concerned with the modeling of heat and mass transfer inside dict polymer PSD. Mattos et al. [15] developed and implemented
the polymer particles and of the reaction mechanism [912]. The a mathematical model for steady-state slurry and bulk propy-
overall polymerization process combined with the polymer PSD in lene polymerization process. The model was capable of predicting
the reactor was usually neglected. the polymer molecular structure and morphology, including PSD.
Choi et al. [13] put forward a population balance model for Using a mixing cell approach and detailed polymerization kinet-
a continuous gas-phase olen polymerization reactor. The model ics combined with population balance equation, Harshe et al. [6]
developed a comprehensive model for propylene polymerization in
a FBR. The need for coupling the reaction-engineering model with
Corresponding author. Tel.: +86 592 2187190; fax: +86 592 2187231. the population balance equation was also demonstrated in their
E-mail address: luozh@xmu.edu.cn (Z.-H. Luo). work.

1385-8947/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2008.10.023
Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476 467

PSD produced in the tubular loop reactors. Based on the reaction


Nomenclature parameters estimation and the population balance equation, the
model takes into account the ow type, the particle dynamics and
A parameter for the ne powders size distribution the particle growth, as well as, the attrition. Particular attention is
C* concentration of activate site (mol/kg) paid to the effects of the operational parameters on the polymer
D particle diameter (m) PSD and to the comparison of PSD between the predicted and the
E(t) residence time density function (h1 ) actual data.
F owrate (kg/h)
F(t) residence time accumulation function
G(D) particle diameter growth rate (m/h) 2. Process for propylene polymerization in tubular loop
k reaction kinetic constant reactors
m particle mass (kg)
M monomer concentration (mol/m3 ) In the present study, Himont Spheripol Loop technology with
Mm monomer molecular weight (kg/mol) the continuous tubular loop reactors as its heart is studied. To illus-
n parameter for single particle model trate the concepts discussed in this paper, a typical schematic ow
P(D) mass density function (m1 ) sheet is given in Fig. 1.
r reaction rate There are many papers concerning the loop technology [2,16,17].
Ra particle diameter attrition rate (m/h) Comprehensive reviews have also been reported [18,19]. Some
t residence time (h) points related to this work are listed in the following contents.
T reaction temperature ( C) First, the loop reactors, including the prepolymerization reactor
w weight fraction of solid phase in the reactor (R200, 0.46 m3 ) and the main polymerization reactor (R201, 56 m3 ),
W total particles mass in tubular loop reactor (kg) are of most importance in the Himont Spheripol Loop technol-
ogy. Secondly, R201 consists of three tubular reactors connected
Greek letters and symbols in sequence with each other, in practice, the tubular loop reactor
 efciency is a closed tube as a whole, wherein the reacting slurry driven by
 mathematical constant a recycling pump circulates with high-recycle rates, as depictured
 particle density (kg/m3 ) in Fig. 1. Under these conditions, it is possible and reasonable to
 standard deviation assume the tubular loop reactor as a continuous stirred-tank reac-
tor (CSTR) with constant volume [2,15,20,21], but slurry density.
Subscripts and superscripts The reacting slurry is supposed to be a mixture of liquid phase
cat catalyst property (monomer and hydrogen) and a solid phase (polymer and cat-
d deactivation reaction alyst). Finally, the catalyst is assumed to undergo the common
DONOR electronic donor for ZieglerNatta catalyst activationpolymerizationdecay pathway.
e efciency
f polymer ne powders 3. Mathematical model
H2 hydrogen
max maximum size 3.1. Introduction
min minimum size
p propagation reaction During the solid-catalyzed liquid-phase propylene polymeriza-
pp polypropylene tion in the tubular loop reactors, the small catalyst particles are
prop propylene continuously fed into R200 at a constant rate. The catalyst frag-
TEAL triethyl aluminum ments into a large number of small particles or layers under mild
0 initial stage reaction conditions in R200 [22], accordingly, the polymerization
1 inlet stream reaction is initiated, and the catalyst particles are well lmed to
2 outlet stream prevent the fragments from separating from the catalyst parti-
200 prepolymerization reactor R200 cles. Subsequently, the small catalyst particles coated with the thin
201 main polymerization reactor R201

Based on above discussion, it becomes clear that the early mod-


eling efforts in this eld are made to account for the detailed aspect
of the population balance model and the particle residence time
in reactors separately. In practice, the two are coupled with each
other. It is also clear that the PSD models established before mainly
focused on the FBR. Though there is no difference between them
when both FBR and tubular loop reactor are treated as a continu-
ous stirred tank reactor (CSTR) in simulation, they are still different
in plant operation and solution methodology of the population
balance equation. Moreover, it is important that most of the PSD
models proposed are not validated with any actual industrial data.
Nevertheless, they are still helpful as reference when dealing with
the PSD model of the polypropylene produced in tubular loop reac- Fig. 1. Himont Spheripol Loop process (Fujian Petrochemical Company of SINOPEC):
Acatalyst; Bpropylene; Chydrogen; Dcoolant in; Ecoolant out; 18control
tors.
sites; 9E201, heat exchanger; 10E203, heat exchanger; 11P200, pump;
In the present work, a comprehensive steady-state population 12P201, pump; 13Z203, mixer; 14R200, prepolymerization reactor; 15R201,
balance model is developed for the prediction of the polypropylene main polymerization reactor; 16D202, buffer tank.
468 Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476

polymer lm in R200 are transported into R201 and react with polymerization. Thus, Eq. (5) can possibly be improved and modi-
the incoming monomers. Simultaneously, the particles may either ed empirically as shown below:
grow in size due to the polymerization reaction or rupture into  n
small ones caused by mechanical or rheological attrition. There- Dcat
rp (t) =  ekd t rp,0 (6)
fore, the polymer particle is inuenced by the original size of the Dpp
catalyst particle, the polymerization rate, the attrition rate, and
wherein Dcat and Dpp are the catalyst and the polymer particle diam-
the residence time [15]. Moreover, due to the original catalyst PSD
eter, respectively;  represents the efciency factor, which is a liner
and residence time distribution (RTD) inside the reactors, the nal
function of Dcat shown in Eq. (7), describing the specic surface
product PSD can be very broad. Accordingly, the population bal-
and the active sites buried in the support material of different size.
ance model should take into account of the following three basic
Notice that Dpp is also a function of the residence time and the
aspects, namely, the propylene polymerization kinetics, the particle
catalyst particle diameter Dcat , which is described in Eq. (8):
dynamics and the residence time distribution.
In the present study, attention is specially paid to the particle  = ke Dcat + 1 (7)
size of the solid phase. For the sake of simplicity, the polymer par-
ticle is assumed to be of spherical shape with constant density. Dpp = Dpp (t, Dcat ) (8)
According to Refs. [4,23], the polymer PSD in the reactors (R200
and R201) can be described via the statistical density function, P(D), 3.3. Particle dynamics
which can be formulated with the following equation:
 Dmax According to Refs. [2,3,5,6,13,15], the following assumptions for
P(D) dD = 1 (1) particle dynamics are made in this paper:
Dmin
(1) all catalyst active sites have been activated simultaneously,
where P(D) dD is the mass fraction of the polymer particles in the
(2) there are no radial concentration gradients within the particles,
size range of (D, D + dD). Note that due to the well-mixed assump-
(3) the temperature gradients in the particles are ignored.
tion for the particles in the reactors, the particle density function
in the product steam is identical to that in the reactors.
Assuming a polymer particle formed base on a catalyst particle,
then the following mass balance equation can be obtained:
3.2. Polymerization kinetics
 t
 3  3 3
To describe the propylene polymerization kinetics over a D cat rp (t) dt = (D Dcat )pp (9)
6 cat 0
6 pp
ZieglerNatta catalyst, a simple kinetic model is employed [6,24].
The polymerization kinetic scheme comprises a series of ele- In addition, Eq. (10) can be obtained via derivation of Eq. (9):
mentary reactions, namely, site activation, propagation, site  2

3pp Dpp dDpp
deactivation, site transformation and chain transfer reactions. rp (t) = (10)
3
cat Dcat dt
Among them, the propagation reaction and the site deactivation
play an important role in the particle growth rate [6,22]. Based on
Based on the constant value of Dcat for a single polymer particle,
the above discussion and the pseudo-kinetic rate constant method,
the correlation of Dpp and t can also be obtained and shown in Eq.
the following kinetic equations are suggested and adopted:
(11) by incorporating Eq. (6) into Eq. (10):
Propagation rate:
 n  2

Dcat 3pp Dpp dDpp
 ekd t rp,0 = (11)
Dpp 3
cat Dcat dt
rp = kp Mc Mm (2)
Site deactivation rate: Accordingly, Eqs. (12) and (13) can be obtained by integrating
Eq. (11):
rd = kd c (3)
Dpp = Dpp (t, Dcat )
Combination of Eqs. (2) and (3) leads to  1/(3+n)
  (3 + n)  rp,0 cat Dcat
(3+n)
(1 ekd t )
dc rp =
(3+n)
+ Dcat
rd = = kd (4) 3 pp kd
dt kp M Mm
(12)
Integrate rp with the isothermal condition and rp as a function
of time, t, thus Eq. (5) can be obtained:  
3+n 3+n
1 3(Dpp Dcat ) pp kd
rp = rp,0 ekd t (5) t = t(Dpp , Dcat ) = ln 1 (13)
kd 3+n
(3 + n)  rp,0 cat Dcat
where it is important to point out that the feed impurity can lead
to the catalyst deactivation. Besides, the catalyst can also deacti- From the viewpoint of statistical methodology, it is obvious that
vate spontaneously [3]. Therefore, kd is supposed to represent the a polymer particle with diameter of Dpp in the size range of (Di , Di+1 )
above two deactivation routes totally, which are independent and has the following possibility:
dependent, respectively.  Di+1  Dcat,max  j
t
i+1
When referring to the polymer particles, the inter- and
Ppp (Dpp ) dDpp = Epp (t) dt Pcat (Dcat ) dDcat
intra-particle heat and mass transfer limitations do affect the poly- Di Dcat,min t
j
i
merization rate. Though slow reaction rate considered in Ref. [25], (14)
some aspects of polymer particles growth are similar, thus we
assume that the polymerization rate are more likely to be affected where Pcat (Dcat ) is the catalyst PSD density function; Epp (t) is the
by the amount of the polymer produced on the surface of the cata- residence time distribution based on the polymer particles for each
j j
lyst particle compared with the residence time in the early stage of diameter of catalyst particles; ti and ti+1 are both functions of Dpp
Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476 469

and Dcat as discussed in Eq. (13). In addition, as for the discrete per particle mass due to the attrition can be described in Eq. (18):
situation, Eq. (14) can be rewritten as Eq. (15):
(dmpp (Dpp )/dt) 3Ra (Dpp )
 = (18)

ncat j
t
i+1
mpp (Dpp ) Dpp
Wpp,i = mcat,j Epp (t) dt (15)
j
t
j=1 i 3.4.2. Formulation of population balance equation
Following the population balance method described in Fig. 2
where ncat represents the number of catalyst particle diameter size
[4], the steady-state mass balance equation for particles in the size
cuts; mcat,j represents the mass fraction of catalyst particles in each
range of (D, D + D) is given by Eq. (19):

(19)

It is necessary to mention that for a particle size distributed


size cuts; Wpp,i represents the mass fraction of polymer particles
feed, the rst term in Eq. (19) represents the particle inow in the
with the diameter lying between Di and Di+1 .
size range of (D, D + D), which is the case of R201 when the out-
For the sake of estimation for kinetic parameters from plant data,
ow of R200 enters into with size-distributed polymer particles.
a chi-square merit function is dened and used to determine best-
On the other hand, for a uniform-size catalyst feed as in R200, this
t parameters by its minimization:
term is only taken into account for the rst particle cut. Substitute
npp  
 (D)pp,i
W (D)pp,i W
each of the related terms in Eq. (19) for mathematical expression
2 = (16) using the particle density function, particle growth rate and attri-
(D)i
i=1 tion rate, as well as the corresponding mass ow rate, then we
obtain:
where npp is the number of plant data points and (D) is the stan- F F
dard deviation of the plant data points [26]. 1 2
P1 (Di ) D P2 (Di ) D + P2 (Di1 )G(Di1 )
W W
3.4. Steady-state population balance model 3G(Di )
+ P2 (Di+1 )Ra (Di+1 ) + P2 (Di ) D P2 (Di )(G(Di )
Di
3.4.1. Particle growth and attrition
For the sake of the simplicity and keeping in consistence with 3Ra (Di )
+ Ra (Di )) P2 (Di ) D = 0 (20)
the particle dynamics presented above, a simplication of the mor- Di
phology model is introduced and the growth rate of the polymer
particle diameter is given as follows, which is derived from Eq. (10) All of the terms of Eq. (20) are divided by D and let D
[4,8]: approaches to 0; accordingly, the continuous form of the popula-
  3  tion balance equation for the case of uniform-size catalyst feed is
dDpp rp (t) Dcat cat
G(D) = = (17) obtained as given below:
dt 2
3pp Dpp
g F F d d
1 2
P1 (D) P2 (D) [P2 (D)G(D)] + [P2 (D)Ra (D)]
Mattos et al. [15] and Freitas et al. [27] assumed and testi- W W dD dD
ed, respectively, that particle coalescence and breakage were not
3
important for PSD evolution in liquid-phase polymerization reac- + P2 (D) [G(D) Ra (D)] = 0 (21)
D
tors. However, polymer particle attritions may occur under the
conditions of high density and relatively fast circular velocity in the
tubular loop reactors. Then, polymer nes may generate due to the According to Ref. [4], when catalyst particles of size Dcat enter
attrition forces resulting from particle collision in the reactors, and into a loop reactor, once they get contact with the liquid phase
actually, large amount of nes are found in the local plant. Under monomer, polymerization occurs and their sizes increase instan-
the assumption of a continuous size reduction of the polymer par- taneously by at least a very small increment before they exit the
ticles [23], the size of the subtle product can be considered to be reactor. Therefore, the particle size distribution appears to be dis-
innitely small, accordingly, the amount of active sites contained continuous at Dpp = Dcat , or otherwise:
can be ignored. Thus, the subtle product attrited makes little con-

0 Dpp = Dcat
tribution to the overall polymerization rate [4]. Denoting Ra (D) as Ppp (D) = (22)
Ppp (D0 ) Dpp = D0
the particle attrition rate, the average rate of particle mass change
470 Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476

Fig. 2. Schematic representation of particle size variation with respect to particle growth and attrition mechanism.

where D0 is an arbitrary value close to Dcat , but not exactly equals to introducing Eq. (13) and (24):
Dcat . For the case of size distributed catalyst feed, the particle den- 
t
sity function will be given by the weighted integral of the particle d Ppp [Dpp (t, Dcat )] (dDpp /dt) dt
dFpp (t) 0
size distributions resulting from the various discrete uniform size Epp (t) = = (24)
catalyst fractions: dt dt

 Dcat,max
Because of the catalysts short residence time in R200 and its
mild reaction conditions, there is only a little effect of the R200
Ppp (Dpp ) = Ppp (Dpp , Dcat )Pcat (Dcat ) dDcat (23)
Dcat,min
on the total polymer PSD, thus, in the estimation procedure, only
R201 is taken into account. With the residence time distribu-
tion of the polymer particles in the tubular loop reactors and the
3.5. Solution methodology plant data including PSDs of the polymer and catalyst, the kinetic
parameters of kd , ke and n can be tted iteratively through the
The present study has considered R200 and R201 as shown in procedure discussed in Section 3.3. On the one hand, the esti-
Fig. 1. The corresponding solution methodology is introduced as mation procedure is necessary when lots of parameters are to
follows. be conrmed or sorts of parameters are available; on the other
First, the input is read through the Read Input module. The input hand; we also emphasize that the catalyst, especially, its sup-
is validated and the initialization of the model variables is done. port material have important effects on the catalyst fragmentation
The overall reactors (R200, R201) simulation is calculated in the and the intra-particle mass and heat limitations. Papers published
external Reactors Module (not discussed in this paper). This exter- on the polymer particles and catalyst fragmentations have clearly
nal module is incorporated to supply the total particle mass, W; shown the importance of the support material of the catalyst [25].
the polymer particles output rate, F1 and F2 , in the tubular loop Therefore, it is necessary to describe an empirical single polymer
reactors as expressed in Eq. (21), besides, the polymer properties particle model as accurate as possible through parameter estima-
including the molecular weight distribution and melt index. For tion.
more information, readers are encouraged to refer Refs. [1,12,28].
According to the external module, we can easily predict both the 4. Results and discussion
polymer PSD and properties, and discuss the effects of the oper-
ational parameters on the polymer PSD. Then the program goes 4.1. Parameters estimation
to the next modules solving the population balance equations for
R200 and R201, in sequence. To solve these steady-state population The catalyst used in the plant is an advanced fourth generation
balance equations, shown in Eq. (21), the whole range of varia- of the ZieglerNatta catalyst (TiCl4 /MgCl2 + PEEB + AlR3 ), and was
tion of particle diameter is divided into a suite of equally spaced sampled and characterized by means of microscope to obtain the
diameters. Subsequently, the differential Eq. (21) is approximated particle accumulation curve, and then the particle density curve
by a system of linear recursive algebraic equations written at dis- was also obtained by derivation. Both the corresponding curves are
crete diameters, and then this system of linear algebraic equations shown in Fig. 3. The polymer powder obtained from the plant was
is recursively solved for each catalyst diameter cuts starting from analyzed by using a sizer and was also numerically tackled, the
each catalyst size, to calculate the PSD in the reactors in sequence same as the catalyst, and the result is given in Fig. 4.
[4]. It is important to point out that in present study, Ppp (D0 ) is The estimation of kinetic parameters was accomplished as dis-
rearranged to make Eqs. (1) and (23) satised. This procedure is cussed in Section 3.5 using the above plant data. Accordingly, the
different from Refs. [4,8], wherein the value of the product out- values of kd , ke and n were obtained successfully. Moreover, accord-
ow rate is rearranged to make Eqs. (1) and (23) satised given the ing to the value of kd , the value of kd,0 is also updated via the
case of the constant bed weight in the FBR. Herein, the two PSDs Arrhenius law. It can be seen from Fig. 5 that the predicted data
of the two reactors have been calculated; moreover, if plant data is in good agreement with the plant data. Hence, the single particle
are available, it is interesting that the selective route of parame- model was available as accurate as possible with updated kinetic
ters estimating can be chosen to estimate kd , ke and n introduced parameters through estimation procedure. Parts of parameters esti-
previously. Besides, there is a RTD for polymer module wherein mated or used here are tabulated in Table 1. The following shows
the polymer particles residence time distribution is obtained by the employment of this single particle model.
Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476 471

Fig. 3. Catalyst PSD from Plant.

Fig. 6. Polymerization rate decays with time at the initial stage of polymerization.

Table 1
Simulated parameters in the study.

Parameter Value Unit Ref.

Ep 5.04 104 J/mol [1,28]


Ed 5.04 104 J/mol [1,28]
kp,0 1.789 108 m3 /(mol h) [1,28]
kd,0 6.313 105 h1 Estimated
ke 3.15 103 m1 Estimated
n 9.1264 102 Estimated
Fig. 4. Polymer PSD from Plant.
C* 25 mol/kg-cat Plant data
pp 910 kg/m3 Plant data
cat 600 kg/m3 Plant data

As shown in Fig. 6, the polymerization rate decreases greatly


in the initial stage of polymerization due to the strong decrease
of Dcat /Dpp ratio in the early stage of the polymer/catalyst parti-
monomer concentration between the active site and the bulk uid-
cle formed. For catalyst particles with different efciency factor
phase.
 according to Eq. (7), the polymerization rate decays according
Fig. 7 shows the polymerization rate of different catalyst parti-
to the corresponding initial polymerization rate. Pater et al. [25]
cles decays in same trend as a whole. That it is, all polymerization
have put forward an argument that the above result is ascribed
rates rst decay strongly in the small diameter range due to the
to the phase transition in the growing particle. That it is, ini-
mass or heat limitations, and then decrease gently for a short res-
tially, the catalyst forms the continuous phase, within which the
idence time before the polymerization rates decay severely due to
polymer is distributed. Then after the phase transition, the poly-
the active site deactivation in the large diameter range, namely, the
mer forms the continuous phase in which catalyst fragments are
long residence time in the reactors. Besides, it is important to point
distributed. And this change causes a change in monomer con-
out that catalyst particles with different diameters take different
centration at the active sites, resulting in lower reaction rates.
residence time to reach the same diameter.
Besides, Hutchinson et al. [29] also notice the difference of the

Fig. 5. Comparison of the particle weight fraction between the predicted data and
the plant data. Fig. 7. Polymerization rate decays with particle growth.
472 Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476

Fig. 8. Particle size distribution predicted for R200: T200 = 15.5; T201 = 70;
Fig. 10. Residence time distribution for the catalyst and polymer particles.
FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14;
Fprop,200 = 1500.03.

Eq. (12) into Eq. (25), the residence time distribution for the catalyst
4.2. Comparison between the plant data and the simulated data particles can be obtained and shown in Eq. (26):
  3
F E (t) dt
The polymer PSD data predicted by the population balance W Ppp (Dpp ) dDpp = 3
Dpp pp D cat
cat cat
model are compared with the plant data. Because the R200 reactor 6 6 cat 3 
(/6)Dcat cat
is relatively small in volume and the corresponding polymerization (25)
rate is also small with its mild conditions. In addition, the polymer
particles in R200 are difcult to sample. Therefore, only the poly-
3 
W Dcat
mer particles in R201 were sampled and their PSD data are used in cat dDpp
Ecat (t) = 3  3 
Ppp (Dpp ) (26)
present study, but both of the PSDs predicted of the two reactors (Dpp pp Dcat cat )Fcat dt
are shown in Figs. 8 and 9.
Based on the population balance equations for uniform-size cat-
Fig. 9 shows that there is a very good agreement between the
alyst feed, residence time distributions for some of the catalyst cuts
predicted data and the plant data. Therefore, the present work can
are shown in Fig. 10. In addition, the residence time distributions
be applied to predict the polymer PSD in the steady-state tubular
of the polymer particles discussed previously are also shown in
loop reactors; though Fig. 9 shows that the predicted data has more
Fig. 10, which are all identical for each of the catalyst cuts. Fig. 10
ne powder and less large particles than the plant data. The rea-
shows that the residence time distributions of the polymer and the
sons for these differences may be the neglects of the agglomeration
catalyst are different even though they are both suspending in the
effect during the polymerization and the electrostatic effect during
liquid phase and dragged through the tubular loop reactors with
the experimental analysis of the polymer samples, which may alter
the liquid phase. The reason for this difference is that the polymer
the shape of the PSD more or less. Whats more, the assumption
particles grow with the increase of the residence time but the cat-
of well mixed in tubular loop reactor also affects on the predicted
alyst particles remains constant in mass with the polymerization
results.
continuing. Therefore, the residence time distribution of the cata-
For polymer particles with diameter of Dpp in a size cut of dDpp ,
lyst seems to the same as that of the liquid phase, which assumed
their total mass can be expressed in Eq. (25), where Ecat (t) is the
as the perfect mixed.
residence time distribution based on the catalyst. By incorporating
In addition, Fig. 10 also shows that catalyst particle with differ-
ent size has different residence time distribution, and the above
difference is caused by Eq. (21), wherein, for each single-size cat-
alyst feed, the weight of the total solid in the reactor, W, resulting
from the external module, remains constant, but each catalyst par-
ticle is different in polymerization rate. Thus, to reach the same W,
the residence time for larger catalyst particles have to be extended
according to Eq. (7).
Its interesting that the population balance model proposed here
and the models proposed by Zacca et al. [2] concerning the loop
reactor both characterize the catalysts residence time distribution
as the type of CSTR.

4.3. Single-size catalyst feed

4.3.1. Effects of the feed catalyst size on the polymer PSD


The effects of the feed catalyst particle size on the polymer PSD
are shown in Figs. 11 and 12, which show that the average diam-
Fig. 9. Comparison between the predicted data and the plant data: T200 = 15.5; eter of the polymer particles increases and there is a clear shift to
T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14; broader PSD with the increase of the diameter of the catalyst par-
Fprop,200 = 1500.03. ticles. The increase of the reaction rate at larger catalyst leads to
Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476 473

Fig. 11. Effects of the catalyst size on the polymer PSD in R200: T200 = 15.5; Fig. 14. Effects of the attrition rate constant on the polymer PSD in R201: T200 = 15.5;
T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14; T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14;
Fprop,200 = 1500.03. Fprop,200 = 1500.03.

Fig. 13 illustrates the effect of the catalyst deactivation con-


stant on the polymer PSD in R201. For rst-order site deactivation
kinetics and a given catalyst feed rate, the active site concentra-
tion in the polymer particles decreases greatly with the increase
of the deactivation rate constant. Accordingly, the polymerization
rate decreases with the particles residence time and the polymer
PSD curves become sharp in large particles. Fig. 13 shows that for
faster deactivation constant, the product PSD shifts to lower particle
size.
Figs. 14 and 15 show the effects of the attrition rate constant
on the polymer PSD. Unlike previously published work [4]; the
attrition rate is redened in Eq. (27):

0
 Dpp < 1 mm
Fig. 12. Effects of the catalyst size on the polymer PSD in R201: T200 = 15.5; Ra = w 7 (27)
10 Dpp 1 mm
T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14; 0.55
Fprop,200 = 1500.03.
where the attrition rate of polymer particles in the diameter range
the increase of the polymer particles diameter, and its also well of 01 mm is assumed to be 0, and w is the weight fraction of the
explained according to Eq. (17). solid phase in the reactor. It suggests that when the density of the
material within the loop reactor gets higher, the attrition rate will
4.3.2. Effects of the polymerization parameters on the polymer increase, so does the ne powder.
PSD According to Ref. [30], the particle size distribution of the poly-
For propylene polymerization in industry, the polymer PSD is mer nes, given in Eqs. (28)(30) in the reactor and outow streams
affected by the polymerization parameters including the catalyst
deactivation constant and the attrition rate constant.

Fig. 13. Effects of the catalyst deactivation constant on the polymer PSD in Fig. 15. Fine powders produced with the increase of the attrition rate: T200 = 15.5;
R201: T200 = 15.5; T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14;
Fprop = 24616.14; Fprop,200 = 1500.03. Fprop,200 = 1500.03.
474 Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476

Fig. 18. Effects of the propylene feed rate on the polymer PSD in R201: T200 = 15.5;
Fig. 16. Effects of the reaction temperature on the polymer PSD in R201: T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop,200 = 1500.03.
T200 = 15.5; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14;
Fprop,200 = 1500.03.
particles with the increase of the polymerization temperature. The
observed shift of the PSD curve to large particles is due to the
are improved by following a conic function other than a delta func- increase of the polymerization rate at high temperature.
tion: Figs. 17 and 18 show the effects of the catalyst feed rate and the
 Df max 
np propylene feed rate on the polymer PSD in R201, respectively.
3Ra (Dpp,i )
F2 Ppp (Df ) dDf = W Ppp (Dpp ) dDpp (28) With the increase of the catalyst feed rate, the total polymeriza-
Dpp,i
Df min
i=1 tion rate increases. Hence, the content of the solid in the reactor
increases, leading the increase of the density of the mixture in
Ppp (Df ) = A(Df Dfmin )(Df Df max ) (29) R201, thus, increase the particle residence time, though, this slightly
affects the product PSD, as shown in Fig. 17. With the increase of
Df min Df Df max (30) the propylene feed rate, as can be seen from Fig. 18 the catalyst
average residence time in R201 decreases greatly, which leads to
With the attrition rate constant increases, the polymer PSD the decrease of the average polymer particle size.
becomes narrower and the weight fraction of the larger particles The effect of the polymer PSD in R200 on the polymer PSD in
decreases slightly as showed in Fig. 14, though the ne powders R201 is showed in Figs. 19 and 20. There is nearly the same polymer
increase evidently in Fig. 15. PSD in R201 showed in Fig. 20 even though three case of polymer
particles feed with different PSD from R200 into R200. It proves
4.3.3. Effects of the operational parameters on the polymer PSD that the effect of the polymer PSD in R200 on the polymer PSD in
For propylene polymerization in industry, the polymer PSD is R201 is weak.
inuenced by the polymerization temperature, the catalyst feed
rate, and the propylene feed rate generally. These inuences are 4.4. Multi-size catalyst feed
predicted in our work.
Fig. 16 illustrates the effect of the liquid-phase polymerization Based on the single-size catalyst feed model and Eq. (23), the
temperature on the polymer PSD in R201. As can be seen from effect of the catalyst PSD on the polymer PSD in R201 can also be
Fig. 16, the distribution width of the polymer PSD increases with obtained via the above model. In present study, the initial catalyst
the increase of the polymerization temperature in R201. Simulta- PSD is divided into a number of size cuts and the overall PSD in
neously, there is a great shift in the PSD curve to larger polymer R201 is obtained as a sum of PSDs resulted from the individual

Fig. 17. Effects of the catalyst feed rate on the polymer PSD in R201: T200 = 15.5; Fig. 19. Effects of the propylene feed rate on the polymer PSD in R200: T200 = 15.5;
T201 = 70; FTEAL = 3.37; FDONOR = 0.33; FH2 = 0.7; Fprop = 24616.14; Fprop,200 = 1500.03. T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14.
Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476 475

Fig. 20. Effects of the prepolymerization in R200 on main polymerization of Fig. 23. Effects of the catalyst particle size multi-distribution on the polymer PSD
R201: T200 = 15.5; T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; in R201: T200 = 15.5; T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7;
Fprop = 24616.14. Fprop = 24616.14; Fprop,200 = 1500.03.

in Fig. 22:

2
1 (ln Dcat ln Dcat )
Pcat (Dcat ) = exp (31)
2Dcat  2 2

Fig. 21 shows that the three catalysts fed with the same Dcat
of 60 m are different in  of 0.3, 0.4 and 0.5 based on Eq. (31).
Moreover, the three corresponding polymer PSDs are also obtained
via our model and illustrated in Fig. 22. In addition, Fig. 23 illus-
trates that the bimodal polymer PSD can be obtained by using a
catalyst feed consisting of two types of catalyst particles with dis-
tinct diameter, and the shape of the overall polymer PSD directly
depends on the diameter and mass fraction of the two catalysts.
With the increase of the fraction of the large catalyst particles, the
area of the large polymer particles increases in the outlet polymer
Fig. 21. Catalyst size distribution with different standard deviations.: T200 = 15.5; PSD. Therefore, it is possible to produce special polymer particles
T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14;
with certain PSD type by varying the feed catalyst PSD.
Fprop,200 = 1500.03.

5. Conclusions
catalyst cuts. Hatzantonis et al. [4] assumed that the feed catalyst
PSD has a log-normal distribution and is shown in Eq. (31). Accord-
A comprehensive steady-state population balance model is
ingly, the feed catalyst size distribution is illustrated in Fig. 21,
developed for the prediction of the polypropylene PSD produced
and the corresponding predicted polymer PSD is also shown
in tubular loop reactors. The model takes into account the ow
type, the polymer particle dynamics and the particle growth. Some
actual data of the steady-state polymerization collected from a local
plant are supplied to verify the model. The predicted polymer PSD
produced under steady-state polymerization condition is found to
agree well with the plant data.
The effects of the kinetic and operational parameters on the
polymer PSD are also investigated. The simulated results show that
the average diameter of the polymer particles increases and there
is a shift to a broader PSD with the increase of the diameter of the
catalyst particles when the single-size catalyst feeds. Moreover, as
to the single-size catalyst applied, the polymer PSD becomes nar-
rower with the increases of both the deactivation rate constant
and the attrition rate, but becomes broader with the increase of
the polymerization temperature. Simultaneously, we also nd that
the average diameter of the polymer particles decreases with the
increase of the propylene feed rate. Furthermore, the simulated
results show that a bimodal PSD can be obtained by using a cat-
Fig. 22. Effects of catalyst size distribution on the PSD in R201: T200 = 15.5; alyst feed consisting of two catalysts with distinct diameter, the
T201 = 70; FTEAL = 3.37; FDONOR = 0.33; Fcat = 0.5775; FH2 = 0.7; Fprop = 24616.14; shape of the overall polymer PSD directly depends on the diameter
Fprop,200 = 1500.03. and the mass fraction of this two catalysts.
476 Z.-H. Luo et al. / Chemical Engineering Journal 146 (2009) 466476

Acknowledgments [14] J.B.P. Soares, A.E. Hamielec, Effect of residence time distribution on the size
distribution of polymer particles made with heterogeneous ZieglerNatta and
supported metallocene catalysts. A generic mathematical model, Macromol.
The authors thank National Natural Science Foundation of China Theor. Simul. 4 (1995) 10851104.
(no. 20406016) for nancial support and Fujian Petrochemical [15] A.G. Mattos Neto, J.C. Pinto, Steady-state modeling of slurry and bulk propylene
Company of SINOPEC for plant data and permission to publish it. polymerizations, Chem. Eng. Sci. 56 (2001) 40434057.
[16] Pdero Henrique Hermes Araujo, Jose Carlos de la Cal, Jose Maria Asua, Jose
Authors also thank the anonymous referees for comments on this Carlos Pinto, Modeling particle size distritbution (PSD) in emulsion copoly-
manuscript. merization reactions in a continuous loop reactor, Macromol. Theor. Simul. 10
(2001) 769779.
[17] Z.H. Luo, Z.K. Cao, N.J. Zhu, an analysis on model of a loop reactor for polypropy-
References lene under a steady operating condition, J. Xiamen Univ. (Nat. Sci.) 44 (2005)
534537 (in Chinese).
[1] Z.H. Luo, Y. Zheng, Z.K. Cao, S.H. Wen, Mathematical modeling of the molecular [18] K.Y. Choi, W.H. Ray, Recent developments in transition metal catalysed olen
weight distribution of polypropylene produced in a loop reactor, Polym. Eng. polymerizationa sury. I. Ethylene polymerization, J. Macromol. Sci. C: Polym.
Sci. 47 (2007) 16431649. Rev. 25 (1985) 155.
[2] J.J. Zacca, J.A. Debling, W.H. Ray, Reactor residence time distribution effects [19] K.Y. Choi, W.H. Ray, Recent developments in transition metal catalysed olen
on the multistage polymerization of olens. I. Basic principles and illustrative polymerizationa sury. II. Propylene polymerization, J. Macromol. Sci. C:
examples, polypropylene, Chem. Eng. Sci. 51 (1996) 48594886. Polym. Rev. 25 (1985) 5797.
[3] D.Y. Khang, H.H. Lee, Particle size distribution in uidized beds for catalytic [20] B. Cui, Q.R. Zhang, Y.N. Li, C.Q. Cao, Z.Y. Zhu, Modeling and parameter distribu-
polymerization, Chem. Eng. Sci. 52 (1997) 421431. tions in a loop reactor employed for propylene polymerization, Chem. Ind. Eng.
[4] H. Hatzantonis, A. Goulas, C. Kiparissides, A comprehensive model for the 13 (1996) 18 (in Chinese).
prediction of particle-size distribution in catalyzed olen polymerization [21] B. Cui, Q. Ma, C.Q. Cao, Z.Y. Zhu, Q.R. Zhang, J.Z. Liu, H.T. Wang, Kinetic charac-
uidized-bed reactors, Chem. Eng. Sci. 53 (1998) 32513267. teristics of the tubular loop reactor for propylene polymerization, Chem. React.
[5] H. Yiannoulakis, A. Yiagopoulos, C. Kiparissides, Recent developments in the Eng. Technol. 12 (1996) 263269 (in Chinese).
particle size distribution modeling of uidized-bed olen polymerization reac- [22] T.M.P. Jochem, W. Gunter, W.P.M. van Swaaij, Polymerization of liquid propy-
tors, Chem. Eng. Sci. 56 (2001) 917925. lene with a 4th generation ZieglerNatta catalyst: inuence of temperature,
[6] M.H. Yogesh, P.U. Ranjeet, V.R. Vivek, A computational model for predicting par- hydrogen and monomer concentration and prepolymerization method on poly-
ticle size distribution and performance of uidized bed polypropylene reactor, merization kinetics, Chem. Eng. Sci. 57 (2002) 34613477.
Chem. Eng. Sci. 59 (2004) 51455156. [23] O. Levenspiel, D. Kunii, T. Fitzgerald, The processing of solids of changing size
[7] B.Z. Yang, J.D. Wang, Y.R. Yang, Polyethylene particle dynamics of gas phase in bubbling uidized beds, Powder Technol. 69 (1968) 8796.
ethylene polymerization in uidized bed reactor, Petrochem. Technol. 33 [24] M. Caracotsios, Theoretical modeling of Amocos gas phase horizontal stirred
(2004) 11301132 (in Chinese). bed reactor for the manufacturing of polypropylene resins, Chem. Eng. Sci. 47
[8] B.Z. Yang, W. Jiang, J.D. Wang, Y.R. Yang, Simulation of particle size distribution (1992) 25942596.
in gas phase ethylene polymerization uidized-bed reactor, J. Chem. Eng. Chin. [25] J.T.M. Pater, G. Weickert, J. Loos, W.P.M. van Swaaij, High precision propolymer-
Univ. 19 (2005) 461467 (in Chinese). ization of propylene at extremely low reaction rateskinetics and morphology,
[9] T.F. Mckenna, J. Dupuy, R. Spitz, Modelling of transfer phenomena on het- Chem. Eng. Sci. 56 (2001) 41074120.
erogeneous Ziegler catalysts: differences between theory and experiment, an [26] J.B.P. Soares, A.E. Hamielec, Deconvolution of chain-length distributions of
introduction, J. Appl. Polym. Sci. 57 (1995) 371384. linear polymers made by multiple-site-type catalyst, Polymer 11 (1995)
[10] T.F. Mckenna, J. Dupuy, R. Spitz, Modelling of transfer phenomena on het- 22572263.
erogeneous Ziegler catalysts. Part 3. Modelling of intraparticle mass transfer [27] M.F. Freitas, Modelagem e Simulacao de Reatores de Polimerizacao de Etileno
resistance, J. Appl. Polym. Sci. 63 (1997) 315322. em Leito de Lama, DSc thesis, PEQ/COPPE, Universidade Federal do Rio de
[11] Z.H. Luo, J. Li, X.L. Zhan, X.B. Yao, Monte Carlo simulation of solid phase graft Janeiro, Rio de Janeiro, 1998.
polymerization of acrylic acid onto polypropylene, J. Chem. Eng. Jpn. 37 (2004) [28] Y. Zheng, Modeling for Propylene Polymerization in a Liquid Phase Loop Reactor,
737743. Master Thesis, Xiamen University, P.R. China, 2006.
[12] Z.H. Luo, Z.K. Cao, Y.T. Su, Monte Carlo simulation of propylene polymeriza- [29] R.A. Hutchinson, C.M. Chen, W.H. Ray, Polymerization of olens through het-
tion. (I) Effects of impurity on propylene polymerization, Chin. J. Chem. Eng. 14 erogeneous catalysis X: modeling of particle growth and morphology, J. Appl.
(2006) 194199. Polym. Sci. 44 (1992) 13891414.
[13] K.Y. Choi, X. Zhao, S. Tang, Population balance modeling for a continuous gas [30] T.P. Chen, S.C. Saxena, Solids population balance for attrition in uidized beds,
phase olen polymerization reactor, J. Appl. Polym. Sci. 53 (1994) 15891594. Powder Technol. 18 (1977) 279281.

You might also like