You are on page 1of 6

Antonio Espuña, Moisès Graells and Luis Puigjaner (Editors), Proceedings of the 27th European

Symposium on Computer Aided Process Engineering – ESCAPE 27


October 1st - 5th, 2017, Barcelona, Spain © 2017 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/B978-0-444-63965-3.50058-1

A Comprehensive Model for the Simulation of


Ethylene Decomposition in High-Pressure LDPE
Autoclaves
Prokopios Pladisa, Apostolos Baltsasa and Costas Kiparissidesa,b,*
a
Chemical Process & Energy Resources Institute (CPERI), Centre for Research and
Technology Hellas (CERTH), Thessaloniki 57001, Greece
b
Department of Chemical Engineering, Aristotle University of Thessaloniki (AUTH),
Thessaloniki 54124, Greece
costas.kiparissides@cperi.certh.gr
Abstract
In the present study, a comprehensive mathematical model was developed to simulate the
dynamic behaviour of multi-zone, multi-feed high pressure ethylene polymerization
autoclaves and assess the risk of ethylene decomposition under different scenarios. To
describe the complex flow patterns occurring in low density polyethylene (LDPE)
autoclaves, a user-specified multi-segment, multi-recycle model representation of the actual
multi-zone reactor is established. A suitable micro-mixing model has been applied taking
into account initiator segregation phenomena due to inefficient mixing conditions at the
initiator injection points. A general reaction mechanism (that includes a comprehensive
ethylene decomposition kinetic scheme) is employed to represent the kinetics of ethylene
polymerization. The present model is capable of predicting accurately the dynamic
behaviour of industrial LDPE autoclaves (reactor temperature profile, polymerization rate,
monomer conversion, molecular weight properties and MWD) and, thus, can be employed
in the design, optimization and control of these reactors. Decomposition phenomena due to
operation condition changes (e.g., feed impurities and disturbances, excess initiator,
controller failure, poorly tuned controller, etc.) have been analyzed and simulated.
Keywords: LDPE, high-pressure, decomposition, mixing, modelling.

1. Introduction
The high pressure ethylene polymerization is an industrial process of significant economic
importance. Two reactor technologies, namely, tubular and autoclaves, are currently
employed in the production of low density polyethylene (LDPE). High-pressure LDPE
reactors typically operate at high temperatures (150 - 330 oC) and pressures (1200 - 3500
atm). An autoclave reactor is a constantly mixed vessel made up of two or more reaction
zones in series, separated by disks and stirred by a vertical stirrer shaft (see Figure 1).
Despite the large specific power input to the reacting system (20-100 kW/m3) effected by
high agitation rates, the zones are not perfectly mixed due to the very fast reaction kinetics.
The polymerization of ethylene in autoclaves is practically carried out in an adiabatic way.
Cooling of the reaction mixture is effected by the introduction of cold monomer at several
side-feed points along the reactor height. The reaction temperature in a zone is controlled
by manipulating the corresponding initiator feed rate. It is important to point out that
cautious reactor control is required because the reactor usually operates at an open-loop
unstable steady state (Topalis, et al, 1995). The flow behaviour in these reactors deviates
338 P. Pladis et al.

significantly from ideal flow (e.g., perfectly mixed conditions). Thus, it is important to
employ a non-ideal flow mixing model needs to approximate the macro- and micro-mixing
flow behaviour in the reactor.
The main reason for the non-ideal flow in LDPE autoclaves is the lack of spatial
homogeneity in the initiator concentration in a reaction zone which is primarily due to
the very fast reaction kinetics and design of agitation/injection system. It is important to
point out that the characteristic reaction time constant associated with the initiator
decomposition is of the order of 0.1 sec. Note that spatial homogeneity in the initiator
concentration in a reaction zone is brought about by mass exchange at two different
mixing scales, namely, (i) macroscopic scale relative to individual molecules, whereby
large-scale relative motions (e.g., eddies) result in the so-called macro-mixing of
reactive species and (ii) microscopic scale relative molecular motion (e.g., small-scale
eddies and molecular diffusion), which effects mixing at the molecular scale. For all
practical reasons, macro-mixing can be associated with the residence time distribution
(RTD) in a reaction zone, whereas micro-mixing may be linked with the state of fluid
aggregation.

ŽŶĞŝ Ͳϭ

ŽŶĞŝ
Z

ƚŚLJůĞŶĞ sϭ sϮ
н
/ŶŝƚŝĂƚŽƌ

ŽŶĞŝ нϭ

ŽŶĞŝ нϮ

Figure 1: Schematic of an LDPE autoclave and the segregation-backmixing model


A great deal of work has been done on the effect of macro/micro-mixing in chemical
reactors. This is especially important for LDPE polymerization because cold monomer
and initiator are fed into the hot reaction mixture, forming a plume near the inlet, which
has a high initiator concentration. At temperatures above ~300°C, a fast exothermic
reaction can decompose ethylene, leading to localized dark spots in the polymer or
global reactor runaway (Wells and Ray, 2005).
Simulation of Ethylene Decomposition in High-Pressure LDPE Autoclaves 339

At the typical operating temperatures for LDPE reactors, the initiator half-life is shorter
than the mixing times and adequate mixing becomes an important issue. It has also been
suggested that imperfect micro-mixing may lead to local hot spots that can lead to
thermal runaway of the entire reactor. These hot spots can be caused by a number of
incidents, e.g., inadequate mixing, controller failure, poorly tuned controller, feed
temperature disturbance, feed impurities, etc.
In the present study a non-ideal mixing model are being examined, namely, the
segregation–back mixing model (see Figure 1). The model developments were driven
by a number of industrial requirements regarding (i) the application to industrial-scale
multi-zone autoclaves using a user-specified (based on CFD calculations) multi-
compartment, multi-recycle model representation of the actual multi-zone reactor
configuration for process monitoring, process optimization and control (ii) prediction of
the varying specific initiator consumption rate per reaction zone for different initiators
and LDPE grades, (iii) prediction of ethylene decomposition onset in an industrial plant.
To model the molecular and compositional developments in an LDPE autoclave, a
comprehensive kinetic mechanism describing the free-radical polymerization of
ethylene is considered. To describe the complex flow patterns occurring in low density
polyethylene (LDPE) autoclaves, a user-specified multi-segment, multi-recycle model
representation of the actual multi-zone reactor is established. Dynamic mass, molecular
species, macromolecular properties and energy balance equations are derived for each
volume segment of the multi-zone autoclave reactor to simulate the transient reactor
behaviour. The method of moments is employed to reduce the infinite number of
molecular species balances into a low-order system of moment differential equations.
The present comprehensive model is capable of describing the complex macro-mixing
phenomena occurring in an LDPE autoclave and quantifying the effects of process
conditions (e.g. ethylene feed rate, temperature, initiator concentration) and macro-
mixing parameters (e.g. number of compartments per zone and recycle parameters) on
the ethylene conversion and molecular properties (e.g., number and weight average
molecular weights, long chain branching and short chain branching as well as the
bivariate molecular weight-long chain branching distribution) of the polymer. In
addition, the present model is capable to analyze and simulate decomposition phenomena
due to operation condition changes (e.g., feed impurities and disturbances, excess initiator,
controller failure, poorly tuned controller, etc.).

2. Kinetic Mechanism
The following elementary reactions are considered to describe the kinetic mechanism
for ethylene polymerization:
௞೏೔
Initiation ‫ܫ‬௜ ሱሮ ʹܴ‫; כ‬ i = 1, 2 … Ni (1)
௞಺
‫כ‬
Chain Initiation ܴ ൅ ‫ ܯ‬՜  ܴଵ (2)
௞೛
Propagation ܴ௫ ൅ ‫ ܯ‬ሱሮ ܴ௫ାଵ (3)
௞೟೘
Chain Transfer to Monomer ܴ௫ ൅ ‫ ܯ‬ሱۛሮ  ‫ܦ‬௫ ൅ ܴଵ (4)
௞೟ೞ
Chain Transfer to Solvent ܴ௫ ൅ ܵ ሱሮ  ‫ܦ‬௫ ൅ ܴଵ (5)
௞೟೛
Chain Transfer to Polymer ܴ௫ ൅ ‫ܦ‬௬  ሱሮ ‫ܦ‬௫ା ܴ௬ (6)
௞್
Intramolecular Transfer (SCB) ܴ௫  ሱሮ ܴ௫ (7)
340 P. Pladis et al.

௞ഁమ

ȕ-Scission of Internal Radicals ܴ௫ ൅ ‫ܦ‬௬ ሱሮ ‫ܦ‬௫ ൅ ܴ௭ ൅ ‫ܦ‬௬ି௭ (8)
௞೟೏
Termination by Combination ܴ௫ ൅ ܴ௬ ሱሮ ‫ܦ‬௫ା௬ (9)
௞೟೏
Termination by Disproportionation ܴ௫ ൅ ܴ௬ ሱሮ ‫ܦ‬௫ ൅ ‫ܦ‬௬ (10)
௞భ
Ethylene Decomposition Reactions ʹ‫ܥ‬ଶ ‫ܪ‬ସ ՜ ‫ܥ‬ଶ ‫ܪ‬ଷ ή ൅‫ܥ‬ଶ ‫ܪ‬ହ ή (11)
௞మ
‫ܥ‬ଶ ‫ܪ‬ହ ή՜ ‫ܥ‬ଶ ‫ܪ‬ସ ൅ ‫ ܪ‬ή (12)
௞మ
‫ܥ‬ଶ ‫ܪ‬ହ ή՜ ‫ܥ‬ଶ ‫ܪ‬ସ ൅ ‫ ܪ‬ή (13)
௞య
‫ܥ‬ଶ ‫ܪ‬ହ ή  ൅‫ܥ‬ଶ ‫ܪ‬ସ  ՜ ‫ܥ‬ଶ ‫ ଺ܪ‬൅ ‫ܥ‬ଶ ‫ܪ‬ଷ ή (14)
௞ర
‫ܥ‬ଶ ‫ܪ‬ସ ൅ ‫ ܪ‬ή՜ ‫ܥ‬ଶ ‫ܪ‬ଷ ή ൅‫ܪ‬ଶ (15)
௞ఱ
‫ܥ‬ଶ ‫ܪ‬ଷ ή՜ ‫ ܥ‬൅ ‫ܪܥ‬ଷ ή (16)
௞ల
‫ܪܥ‬ଷ ή ൅‫ܥ‬ଶ ‫ܪ‬ସ  ՜ ‫ܪܥ‬ସ ൅ ‫ܥ‬ଶ ‫ܪ‬ଷ ή (17)
௞ళ
ʹ‫ܪܥ‬ଷ ή՜ ‫ܥ‬ଶ ‫଺ܪ‬ (18)
௞ఴ
‫ܥ‬ଶ ‫ܪ‬ଷ ή  ൅‫ܪܥ‬ଷ ή՜ ‫ܥ‬ଶ ‫ܪ‬ଶ ൅ ‫ܪܥ‬ସ (19)
௞వ
ʹ‫ܥ‬ଶ ‫ܪ‬ଷ ή՜ ‫ܥ‬ଶ ‫ܪ‬ଶ ൅ ‫ܥ‬ଶ ‫଺ܪ‬ (20)
Let Rx and Dx be the concentration of “live” and “dead” polymer chains, and ”ୖ౮ and
”ୈ౮ the corresponding net rates of production of “live” and “dead” polymer chains,
respectively. Based on the postulated kinetic scheme one can derive the following
expressions for ”ୖ౮  and ”ୈ౮ by combining the individual rates of generation and
consumptions of “live” and “dead” polymer chains (Pladis and Kiparissides, 1998).
Rate function of “live” polymer chains:
λ

‫  ݔܴݎ‬ൌ  ൝݇‫ܫ‬ ܴ‫ܯ כ‬ ൅  ሺ݇‫ ܯ ݉ݐ‬൅ ݇‫ܵ ݏݐ‬ሻ ෍ ܴ‫ ݔ‬ൡ ߜሺ‫ ݔ‬െ ͳሻ ൅ ݇‫ ݌‬ሺܴ‫ݔ‬െͳ െ ܴ‫ ݔ‬ሻ‫ ܯ‬െ ሺ݇‫ ܯ ݉ݐ‬൅ ݇‫ܵ ݏݐ‬ሻܴ‫ݔ‬
‫ݔ‬ൌͳ
λ λ λ λ

൅ ݇‫ ݔܦݔ ݌ݐ‬൭෍ ܴ‫ ݔ‬൱ െ ݇‫ ݔܴ ݌ݐ‬෍ ‫ ݔܦݔ‬െ ݇‫ ݔܴ ܿݐ‬෍ ܴ‫ ݔ‬െ ݇‫ ݔܴ ݀ݐ‬෍ ܴ‫ݔ‬


‫ݔ‬ൌͳ ‫ݔ‬ൌʹ ‫ݔ‬ൌͳ ‫ݔ‬ൌͳ
λ λ λ

െ ݇ߚ ܴ‫ ݔ‬෍ ‫ ݔܦݔ‬൅ ݇ߚ ෍ ܴ‫ ݔ‬෍ ‫ݕܦ‬


‫ݔ‬ൌʹ ‫ݔ‬ൌͳ ‫ݕ‬ൌ‫ݔ‬൅ͳ (21)
Rate function of “dead” polymer chains:
λ ‫ ݔ‬െͳ λ λ
ͳ
‫ ݔܦݎ‬ൌ ሺ݇‫ ܯ ݉ݐ‬൅ ݇‫ܵ ݏݐ‬ሻܴ‫ ݔ‬൅ ݇‫ ݔܴ ݀ݐ‬෍ ܴ‫ ݔ‬൅ ݇‫ ܿݐ‬෍ ܴ‫ݕܴ ݕ‬െ‫ ݔ‬൅ ݇‫ ݔܴ ݌ݐ‬෍ ‫ ݔܦݔ‬െ ݇‫ ݔܦݔ ݌ݐ‬෍ ܴ‫ݔ‬
ʹ
‫ ݔ‬ൌͳ ‫ ݕ‬ൌͳ ‫ݔ‬ൌʹ ‫ݔ‬ൌͳ
λ λ λ λ

൅ ݇ߚ ܴ‫ ݔ‬෍ ‫ ݔܦ ݔ‬൅ ݇ߚ ቌ෍ ܴ‫ ݔ‬ቍ ෍ ‫ ݕܦ‬െ ݇ߚ ‫ ݔܦݔ‬෍ ܴ‫ݔ‬


‫ ݔ‬ൌʹ ߯ൌͳ ‫ ݕ‬ൌ‫ ݔ‬൅ͳ ‫ ݔ‬ൌͳ (22)
where į(x) is the Kronecker delta function. Based on the above definitions of rate
functions, one can easily derive a large set of deferential equations to describe the
molecular developments in a polymerization reactor. However, instead of solving the
large system of differential equations, a lower order system of differential equations is
usually obtained by using the method of moments. This method is based on the
statistical representation of the molecular properties of the polymer in terms of the
leading moments of the number of chain length distributions of the “live” and “dead”
polymer chains. The single moments of “live” and “dead” polymer chains are given by:
Simulation of Ethylene Decomposition in High-Pressure LDPE Autoclaves 341

ஶ ஶ

ߣ௡ ൌ  ෍ ‫ܴ ݔ‬௫ Ǣߤ௡ ൌ  ෍ ‫ ݔ‬௡ ‫ܦ‬௫ 


௫ୀଵ ௫ୀଶ (23)

Thus, the rate functions of the nth moment is given by


ஶ ஶ

‫ݎ‬ఒ೙ ୀ ෍ ‫ ݔ‬௡ ‫ݎ‬ோೣ Ǣ‫ݎ‬ఓ೙ୀ ෍ ‫ ݔ‬௡ ‫ݎ‬஽ೣ


௜ୀଵ ௜ୀଵ (24)

Using the above definitions one can obtain the corresponding moments equations by
summing the resulting expressions over the total variation of x. The kinetic rate
parameter values were taken from previous studies (Pladis and Kiparissides, 1998;
Wells and Ray, 2005).

3. The Segregation–Backmixing Model


To introduce the basic concept a simple model will be considered first. Let assume that
after entering the mixing vessel the feed stream is broken up by the action of a stirrer or
by turbulence. The lumps of fluid mix with their surroundings. In addition, such a lump
can grow by dilution with the surrounding fluid. Before reaction with the second
component (which is resident in the bulk liquid of the reactor) is possible, a number of
successive mixing steps must be taken: Firstly, on entering the reactor, the feed stream
will be transported through the bulk of the reactor by means of the circulation flow,
induced by the impeller. The turbulent whirls will break the feed stream into aggregates
of smaller length scales. However reaction is not taking place yet, since the lumps of
fluid are still completely segregated from the bulk liquid. This process is called macro-
mixing. The time that is necessary for this is usually in the order of a few seconds (a
little shorter in lab-scales, a little longer on plant scale). However, next to the macro-
mixing also micro-mixing has to occur. With micro-mixing we mean the mixing of the
feed with the bulk to a molecular scale. Once this is started, the reaction can take place.
In Figure 1, the proposed segregation-backmixing model is shown for one single zone
of the autoclave. The model assumes that a part of initiator is in segregated form and
interacts with the rest of the reaction medium with a rate that is determined by a mixing
time, tm. The model includes five parameters the volume V1, the recycle stream R, the
initiator mixing parameters tm1 and tm2 and the backmixing parameter į. It can be shown
that the mixing parameters tm1 and tm2 can be calculated in terms of kinematic viscosity
and local energy dissipation rate, the V1 parameter from CFD results and the R and į
from on-line temperature and initiator feed rate measurements in a zone.

4. Simulation Results
The simulation capabilities of both non-ideal mixing models proposed in this work were
tested using the CFD results of Wells and Ray (2005). Figure 2 shows a comparison of
the simulation results using the segregation-backmixing model on the effect of initiator
feed concentration on the final temperature in a LDPE reactor, using a series of three
CSTRs, a model consisting of 100-CSTRs and the original CFD results of Wells-Ray
(2005). The three lines show the original calculations of Wells-Ray. The black solid
squares represent the calculations of this study. The simulation parameters were: R =
10, V1 = 0.1Vtot, fo = 1. The temperatures shown in the plot are the steady-state
temperatures of the second reactor. The mixing time parameters were set equal to 10
342 P. Pladis et al.

times the engulfment time. All other conditions were similar to the paper of Wells and
Ray (2005). It is apparent that the predictions of the segregation-backmixing model
agree with the results presented by Wells and Ray (continuous lines), for 100-
compartments and CFD simulations shown also in the original paper.
Figure 3 shows the simulation of ethylene decomposition in an industrial LDPE
autoclave. The imposed operation changes were exactly the same as the plant operation
changes before decomposition. It is important to point out that in the particular zone
where the decomposition occurs there is no feed of initiator which means that the
decomposition is due to initiator that exists in segregated form that is entrained from
another zone. This example shows the importance of specifying the state of mixing of
the initiator at any zone of the reactor. In addition it is also worth mentioning that the
initiator efficiency at any zone is not an input variable but is calculated based on the
actual radical production rate in each model compartment.
300
290
280
270
260
250
Temperature
Temperature

240
230
CFD Simulations
220
100 Compartments
210 3 Compartments
200 Seggregation-Backmixing model
First Compartment (V1 ) in Zone i
190
Second Compartment (V2 ) in Zone i
180
170
160
0 30 60 90 120 150
20 30 40 50 60 70 80
Initiator feed (ppm)
Time (min)

Figure 2: Comparison of simulation results Figure 3: Simulation of ethylene decomposition


with literature models. in an industrial LDPE autoclave.

5. Conclusions
In the present study, a comprehensive mathematical model is developed to simulate the
dynamic behaviour of multi-zone, multi-feed high pressure ethylene polymerization
autoclaves. A micro-mixing model has been applied taking into account initiator
segregation phenomena due to inefficient mixing conditions at the initiator injection points.
A general reaction mechanism (include a comprehensive ethylene decomposition kinetic
scheme) is employed to represent the kinetics of ethylene polymerization. The present
model is capable of predicting accurately the dynamic behaviour of industrial LDPE
autoclaves (polymerization rate, monomer conversion, molecular weight properties and the
reactor temperature profile) and, thus, can be employed in the design, optimization and
control of these reactors. Decomposition phenomena due to operation condition changes
(e.g., feed impurities and plant disturbances, excess initiator, controller failure, poorly tuned
controller, etc.) have been analyzed and simulated.

References
P. Pladis, C. Kiparissides, 1998, Chem. Eng. Sci.,, 53, 3315
P. Pladis, C. Kiparissides, 1999, Journal of Applied Polymer Science,73, 2327-2348.
V.Topalis, P. Pladis, C. Kiparissides and I. Goossens, 1996, Chem. Eng. Sci., 51, 2461-2470
G. J. Wells, W. H. Ray, 2005, AIChE J., 51, 3205–3218.

You might also like