You are on page 1of 322

ARISTOTLE UNIVERSITY OF THESSALONIKI (AUTh)

SCHOOL OF ENGINEERING

DEPARTMENT OF CHEMICAL ENGINEERING

SECTION OF ANALYSIS, DESIGN AND CONTROL OF CHEMICAL

PROCESSES & PLANTS

MATHEMATICAL MODELLING AND SIMULATION

OF AN INDUSTRIAL α-OLEFINS CATALYTIC SLURRY PHASE

LOOP-REACTOR SERIES

Dissertation Submitted to the Department of Chemical Engineering

for the Degree of

DOCTOR OF PHILOSOPHY

in Chemical Engineering

Vassileios Touloupides

Chemical Engineer

Thessaloniki, November 2010


Thessaloniki, November 22, 2010
© Βασίλειος Δ. Τουλουπίδης
© Α.Π.Θ.

MATHEMATICAL MODELLING AND SIMULATION

OF AN INDUSTRIAL α-OLEFINS CATALYTIC SLURRY PHASE

LOOP-REACTOR SERIES

 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 

‘Η έγκριση της παρούσης ∆ιδακτορικής ∆ιατριβής από το Τμήμα Χημικών Μηχανικών του
Αριστοτελείου Πανεπιστημίου Θεσσαλονίκης δεν υποδηλώνει αποδοχή των γνωµών του
συγγραφέως’ (Ν. 5343/1932, άρθρο 202, παρ. 2)
Acknowledgements

I am heartily thankful to my supervisor, Professor C.


Kiparissides for giving me the chance to graduate from his
excellent research group. His encouragement, guidance and
support from the initial to the final level, enabled me to develop
a better understanding on chemical engineering and
polymerization process modelling.

The present work is a part of two industrial projects


entitled ‘Mathematical Modelling of Industrial Polyethylene
Slurry Loop Reactos(s) for Ziegler-Natta Catalysts’ and ‘Multi-
scale Multi-zone Mathematical Modelling of Industrial
Polyethylene Slurry Loop Reactor(s) for Ziegler–Natta
Catalysts’. I gratefully acknowledge the technical support and
financial assistance of Total Petrochemicals Research at Feluy
(TPRF), Belgium. More specifically, I would like to thank Denis
Mignon, Pierre Van Grambezen and Alvaro Fernandez for our
helpful discussions throughout this research project.

Furthermore, I would like to acknowledge the financial


support of Chemical Process Engineering Research Institute/
Centre for Research and Technology Hellas (CPERI/ CERTH)
throughout all five-year period of my PhD research.

I would also like to thank all my fellow colleagues at the


Laboratory of Polymer Reaction Engineering of CPERI/
CERTH. More specifically, I would like to thank Dr. Pladis and
Dr. Kanellopoulos for their invaluable help throughout this
research project.

Last but not least, I would like to thank, my mother


Aphrodite and my sister Zina.
Στον παππού μου Νίκο και τη γιαγιά μου Ζηνοβία.
‘If you don't do the best you can with what you happen to have,

you'll never do the best you might have done with what you should have had’

Aris Rutherford*

*
Author’s note: If a job's worth doing, it's worth doing well.
*http://www.wordle.net/
i

Έκθεση Πρωτοτυπίας

Στα πλαίσια της διδακτορικής διατριβής, αναπτύχθηκε ένα ολοκληρωμένο


μαθηματικό μοντέλο για την προσομοίωση της δυναμικής συμπεριφοράς της βιομηχανικής
διεργασίας παραγωγής συμπολυμερούς πολυαιθυλενίου-1-εξενίου σε δύο αντιδραστήρες
τύπου βρόχου υγρής φάσης σε σειρά. Στο μοντέλο της διεργασίας ενσωματώνονται όλα τα
φυσικοχημικά φαινόμενα που λαμβάνουν χώρα σε διαφορετικές κλίμακες μεγέθους και
χρόνου, περιλαμβάνοντας την κινητική των αντιδράσεων πολυμερισμού, τη
θερμοδυναμική ανάλυση όλων των φάσεων της διεργασίας και τα ολικά ισοζύγια μάζας
και ενέργειας. Το μοντέλο που αναπτύχθηκε έχει τη δυνατότητα πρόβλεψης όλων των
μεταβλητών που άπτονται στη λειτουργία της διεργασίας και υπολογισμού των τελικών
μοριακών (μέσο κατά αριθμό και κατά βάρος μοριακό βάρος, κατανομή μοριακού
βάρους), ρεολογικών (ιξώδες και δείκτης τήγματος) και μορφολογικών (κατανομή
μεγέθους σωματιδίων) ιδιοτήτων του παραγόμενου πολυμερούς. Οι προβλέψεις του
μοντέλου συγκρίθηκαν με πειραματικές μετρήσεις της αντίστοιχης βιομηχανικής μονάδας
παραγωγής πολυαιθυλενίου-1-εξενίου για δυναμική (π.χ., έναρξη αντιδραστήρα, αλλαγή
πολιτικής τροφοδοσίας) και μόνιμη κατάσταση λειτουργίας. Σε όλες τις περιπτώσεις, οι
προβλέψεις του μοντέλου (διατηρώντας σε κάθε περίπτωση τις ίδιες κινητικές και
θερμοδυναμικές παραμέτρους) ήταν ποιοτικά και ποσοτικά ακριβείς. Η πρωτοτυπία της
παρούσας διατριβής εστιάζεται στην ολοκληρωμένη προσομοίωση της βιομηχανικής
διεργασίας παραγωγής συμπολυμερούς πολυαιθυλενίου-1-εξενίου σε δύο αντιδραστήρες
τύπου βρόχου υγρής φάσης σε σειρά αλλά και στην προσομοίωση του κάθε αντιδραστήρα
τύπου βρόχου με ένα συνεχή αντιδραστήρα πλήρους ανάμιξης συνδεδεμένο σε σειρά με
μία συσκευή ασυνεχούς απομάκρυνσης προϊόντος.

Οι παρακάτω δημοσιεύσεις σε επιστημονικά περιοδικά και συμμετοχές σε διεθνή


συνέδρια επετεύχθησαν κατά τη διάρκεια της διδακτορικής διατριβής:

Δημοσιεύσεις σε επιστημονικά περιοδικά (5):

Touloupides V., Kanellopoulos V., Krallis A., Pladis P. and Kiparissides C., 2010,
‘Modeling and Simulation of Particle Size Distribution in Slurry-Phase Olefin Catalytic
Polymerization Industrial Loop Reactors’, Computer Aided Chemical Engineering, 28, pp.
43-48.

Krallis A., Pladis P., Kanellopoulos V., Saliakas V., Touloupides V. and
Kiparissides C., 2010, ‘Design, Simulation and Optimization of Polymerization Processes
ii

Using Advanced Open Architecture Software Tools’, Computer Aided Chemical


Engineering, 28, pp. 955-960.

Touloupides V., Kanellopoulos V., Pladis P., Kiparissides C., Mignon D. and Van-
Grambezen P., 2010, ‘Modeling and Simulation of an Industrial Slurry-phase Catalytic
Olefin Polymerization Reactor Series’, Chem. Eng. Sci., 65, pp. 3208-3222.

Kanellopoulos V., Tsiliopoulou E., Dompazis G., Touloupides V. and Kiparissides


C., 2007, ‘Evaluation of Morphology of Polymer Particles Produced in Catalytic Gas-
phase Olefin Polymerization Reactors’, Ind. Eng. & Chem. Res., 46, pp. 1928-1932.

Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C., 2007,


‘Development of a Multi-scale, Multi-phase Multi-zone Dynamic Model for the Prediction
of Particle Segregation in Catalytic Olefin Polymerization FBRs’, Chem. Eng. Sci., 63, pp.
4735-4753.

Συμμετοχές σε συνέδρια (15):

Touloupides V., Kanellopoulos V. and Kiparissides C., ‘Prediction of Dynamic


Particle Size Distribution in Industrial Slurry-Phase Olefin Catalytic Polymerization Loop
Reactors’, AIChE 10/ Annual Meeting, 7-12 November, 2010, Salt Lake City, UT.

Touloupides V., Kanellopoulos V., Pladis P. and Kiparissides C., ‘Modeling and
Simulation of Particle Size Distribution in Slurry-Phase Olefin Catalytic Polymerization
Industrial Loop Reactors’, 20th European Symposium on Computer Aided Process
Engineering (ESCAPE 20), 6-9 June, 2010, Ischia, Naples, Italy.

Krallis A., Pladis P., Kanellopoulos V., Saliakas V., Touloupides V. and
Kiparissides C., ‘Design, Simulation and Optimization of Polymerization Processes Using
Advanced Open Architecture Software Tools’, 20th European Symposium on Computer
Aided Process Engineering (ESCAPE 20), 6-9 June, 2010, Ischia, Naples, Italy.

Touloupides V., Kanellopoulos V., Pladis P., Krallis A. and Kiparissides C.,
‘Modeling and Simulation of Slurry-Phase Olefin Catalytic Polymerization Industrial Loop
Reactors. Prediction of Polymer Particle Growth, Molecular, Rheological and
Morphological Polymer Properties’, AIChE 09/ Annual Meeting, November 8-13, 2009,
Nashville, USA.
iii

Touloupides V., Kanellopoulos V., Pladis P., Krallis A. and Kiparissides C., ‘Real-
time Simulation of an Industrial Scale Slurry Loop Reactor’, 8th World Congress of
Chemical Engineering (WCCE8), 23-27 August, 2009, Montreal, Canada.

Τουλουπίδης Β., Κανελλόπουλος Β., Πλαδής Π., Κράλλης Α. και Κυπαρισσίδης Κ.,
‘Δυναμική Προσομοίωση Βιομηχανικού Καταλυτικού Αντιδραστήρα Ανακύκλωσης
Πολυμερισμού’, 7ο Πανελλήνιο Επιστημονικό Συνέδριο Χημικής Μηχανικής, 3-5 Ιουνίου,
2009, Πάτρα, Ελλάδα.

Touloupides V., Kanellopoulos V., Pladis P., Krallis A. and Kiparissides C.,
‘Modeling of Industrial Catalytic Olefin Polymerization Slurry Reactors’ 7th Panhellenic
Polymer Conference, 28 Sep.-1 Oct., 2008, Ιωάννινα, Ελλάδα.

Touloupides V., Kanellopoulos V., Pladis P. and Kiparissides C., ‘Mathematical


Modelling of an Industrial Catalytic Olegin Copolymerzation Slurry-Loop Reactor’, 1st
Intrnational Conference on the Reaction Engineering of Polyolefins (INCOREP), June 22-
27, 2008, Montreal, Canada.

Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C., ‘Development


of a Dynamic Multi-compartment Model for the Prediction of Particle Size Distribution
and Molecular Properties in a Catalytic Olefin Polymerization FBR’, 3rd International
Conference on Population Balance Modeling, September 19-21, 2007, Quebec City,
Canada.

Touloupides V., Kanellopoulos V., Dompazis G. and Kiparissides C., ‘Development


of a Multi-scale Dynamic Model for the Prediction of Polymer Distributed Properties in
Catalytic Olefin Polymerization Slurry Loop Reactors’, European Congress of Chemical
Engineering (ECCE-6), 16-20 September, 2007, Copenhagen, Denmark.

Τουλουπίδης Β., Δομπάζης Γ., Κανελλόπουλος Β. και Κυπαρισσίδης Κ., ‘Δυναμική


Προσομοίωση του Καταλυτικού Πολυμερισμού Ολεφινών σε Αντιδραστήρες Συνεχούς
Λειτουργίας Αέριας και Υγρής Φάσης’, 6ο Πανελλήνιο Επιστημονικό Συνέδριο Χημικής
Μηχανικής, 31 Μαίου-2 Ιουνίου, 2007, Αθήνα.

Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C., ‘Multi-scale,


Multi-phase, Multi-zone Dynamic Model for the Prediction of Particle Segregation in
Catalytic Olefin Polymerization FBRs’, Hangzhou International Polymer Forum,
Advanced Materials and Reaction Engineering, May 13-17, 2007, Hangzhou, China.
iv

Kanellopoulos V., Gustafsson B., Touloupides V. and Kiparissides C., ‘Gas-phase


Olefin Polymerization in the Presence of Supported and Self-supported Ziegler-Natta
Catalysts’, Hangzhou International Polymer Forum, Advanced Materials and Reaction
Engineering, May 13-17, 2007, Hangzhou, China.

Dompazis G., Kanellopoulos V., Touloupides V., and Kiparissides C., ‘Development
of a Multi-Scale, Multi-Phase, Multi-Zone Dynamic Model for the Prediction of Particle
Segregation in Catalytic Olefin Polymerization Fbrs’, AIChE 06/ Annual Meeting, Nov.
12-17, 2006, San Francisco.

Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C., ‘Development


of a Multi-scale, Multi-phase, Multi-zone Dynamic Model for the Prediction of Polymer
Distributed Properties in Catalytic Olefin Polymerization FBRs’, 6ο Πανελλήνιο Συνέδριο
Χημικής Μηχανικής, 3-5 Νοεμβρίου 2006, Πάτρα, Ελλάδα.
v

Περίληψη

Στην παρούσα διδακτορικη διατριβή αναπτύσσεται ένα ολοκληρωμένο μαθηματικό


μοντέλο πολλαπλών κλιμάκων, πολλαπλών φάσεων, για την περιγραφή της δυναμικής
λειτουργίας της βιομηχανικής διεργασίας δύο αντιδραστήρων καταλυτικού
συμπολυμερισμού αιθυλενίου-1-εξενίου τύπου βρόχου υγρής φάσης σε σειρά. Πιο
συγκεκριμένα, μελετήθηκε η επίδραση των λειτουργικών συνθηκών του αντιδραστήρα στη
δυναμική συμπεριφορά και λειτουργία του (π.χ., θερμοκρασία και πίεση, τροφοδοσία
καταλύτη, μονομερών και διαλύτη, κ.α.) όπως επίσης και στις μοριακές, ρεολογικές και
μορφολογικές ιδιότητες του παραγόμενου πολυμερούς (π.χ., μέσο κατά αριθμό και κατά
βάρος μοριακό βάρος (Mn και Mw), κατανομή μοριακού βάρους (ΚΜΒ), ιξώδες, κατανομή
μεγέθους σωματιδίων (ΚΜΣ)).

Σύμφωνα με το παρόν μαθηματικό μοντέλο, κάθε ένας από τους αντιδραστήρες


τύπου βρόχου (μαζί με τα δοχεία κατακάθισης), προσομοιώνεται με έναν συνεχή
αντιδραστήρα πλήρους ανάμιξης συνδεδεμένο με μία συσκευή ασυνεχούς απομάκρυνσης
προϊόντος. Για την πρόβλεψη της δυναμικής εξέλιξης των συγκεντρώσεων όλων των
μοριακών ειδών σε όλες τις φάσεις αλλά και της θερμοκρασίας του αντιδραστήρα και του
ψυκτικού υγρού, επιλύονται δυναμικά ισοζύγια μάζας και ενέργειας για τον κάθε
αντιδραστήρα της σειράς. Ο ρυθμός της ασυνεχούς απομάκρυνσης προϊόντος, όπως επίσης
και οι συγκεντρώσεις των μοριακών ειδών του ρεύματος εξόδου υπολογίζονται μέσω ενός
μοντέλου προσομοίωσης για τα δοχεία κατακάθισης. Ο υπολογισμός της δυναμικής
εξέλιξης των κατανεμημένων μοριακών ιδιοτήτων (Mn, Mw, ΚΜΒ) επιτυγχάνεται μέσω
ενός καταλυτικού κινητικού μηχανισμού Ziegler-Natta πολλαπλών ενεργών καταλυτικών
κέντρων και της στατιστικής μεθόδου των ροπών. Για τον καθορισμό του ελάχιστου
αριθμού διαφορετικών ειδών καταλυτικών κέντρων όπως επίσης και των τιμών των
κινητικών σταθερών εφαρμόστηκε μία μεθοδολογία αποσυνέλιξης της πειραματικής
ΚΜΒ. Για τον υπολογισμό των συγκεντρώσεων θερμοδυναμικής ισορροπίας όλων των
μοριακών ειδών (π.χ., μονομερή, διαλύτης, υδρογόνο) στις δύο φάσεις (στερεή και υγρή)
εφαρμόστηκε η καταστατική εξίσωση Sanchez-Lacombe. Στο μοντέλο ενσωματώθηκαν
όλοι οι βρόχοι ελέγχου που χρησιμοποιούνται σε βιομηχανικές διεργασίες αντιδραστήρων
τύπου βρόχου σε υγρή φάση. Επιπλέον, ο υπολογισμός των ιξωδοελαστικών ιδιοτήτων του
παραγόμενου πολυμερούς (ιξώδες και δείκτης τήγματος) επιτυγχάνεται μέσω ενός
ρεολογικού μοντέλου. Τέλος, για τον υπολογισμό των μορφολογικών ιδιοτήτων του
vi

παραγόμενου πολυμερούς (ΚΜΣ), επιλύεται το γενικευμένο δυναμικό πληθυσμιακό


ισοζύγιο.

Ένας μεγάλος αριθμός προσομοιώσεων πραγματοποιήθηκε προκειμένου να


παρουσιαστεί η επίδραση των συνθηκών λειτουργίας στη δυναμική συμπεριφορά της
διεργασίας όπως και στις μοριακές, ρεολογικές και μορφολογικές ιδιότητες του
παραγόμενου πολυμερούς. Οι προβλέψεις του μοντέλου είναι ποιοτικά συνεπείς με την
αναμενόμενη συμπεριφορά των μεταβλητών της διεργασίας και τις ιδιότητες του
πολυμερούς. Επιπλέον, είναι και ποσοτικά ακριβείς, συγκρινόμενες με δυναμικές
μετρήσεις βιομηχανικής μονάδας παραγωγής πολυαιθυλενίου διεργασίας αντιδραστήρων
τύπου βρόχου σε υγρή φάση σε σειρά, υπό διαφορετικές λειτουργικές συνθήκες. Το παρόν
μοντέλο μπορεί να βρει εφαρμογή στο σχεδιασμό, τη βελτιστοποίηση και τον έλεγχο μιάς
βιομηχανικής μονάδας παραγωγής πολυαιθυλενίου μέσω της διεργασίας αντιδραστήρων
τύπου βρόχου σε υγρή φάση σε σειρά.
vii

Abstract

In the present research study, a comprehensive multi-scale, multi-phase,


mathematical model has been developed to describe the dynamic operation of an industrial
slurry-phase ethylene-1-hexene co-polymerization loop-reactor series. More specifically,
the effects of various operating conditions on the dynamic reactor behaviour (i.e., reaction
temperature and pressure, inflow rates of catalyst, monomers and diluent, etc.) as well as
on the molecular, rheological and morphological polyolefin properties (i.e., Mn, Mw,
MWD, complex viscosity, particle size distribution, etc.) were fully assessed.

According to the proposed modelling approach, each loop reactor (i.e., consisting of
the loop reactor and the settling legs) is modelled as an ideal CSTR in series with a semi-
continuous product removal unit. Dynamic macroscopic mass and energy balances are
derived for each loop reactor in the series to predict the time variation of the concentrations
of the various molecular species as well as the reactor and jacket temperatures in the two
loop reactors. The non-continuous product withdrawal rate as well as the outflow species
concentrations from each reactor of the configuration is properly calculated via the solution
of a settling leg model. A multi-site Z-N kinetic mechanism is employed to calculate the
dynamic evolution of the average and distributed molecular properties (i.e., Mn, Mw,
MWD) of the polyolefin in each reactor of the series, in conjunction with the well-known
method of moments. In order to determine the minimum number of different catalyst active
sites for the reconstruction of the experimentally measured MWDs, as well as the kinetic
parameters values, a MWD deconvolution analysis was also employed. The Sanchez-
Lacombe equation of state is utilized to calculate the thermodynamic equilibrium
concentrations of the various molecular species (i.e., monomer(s) diluent, hydrogen, etc) in
the two phases (i.e., solids and liquid) present in the reaction mixture. All different control
loops, typically employed by an industrial slurry-phase loop reactor, are incorporated in the
reactor model. Moreover, a rheological model is employed for the prediction of the
viscoelastic behaviour of polyolefins produced in slurry-loop reactors (i.e., complex
viscosity and melt flow index). Finally, a generalized dynamic population balance model is
employed to investigate the particle-size distribution developments.

Extensive numerical simulations were carried out to assess the effects of process
operating conditions on key process variables as well as on the molecular, rheological and
morphological properties of polyolefins. The results are qualitatively consistent with the
viii

expected behavior for the process variables and polymer properties. Moreover, it has been
found that model predictions are in excellent agreement (quantitatively) with dynamic
measurements under different plant operating policies (i.e., startup, grade transition, etc.)
obtained from an industrial slurry-phase catalytic olefin polymerization loop-reactor series.
The proposed model should find wide application in the design, optimization and control
of industrial polyolefin slurry-phase loop-reactor processes.
ix

Contents

1. Introduction ................................................................................................................................................ 1

1.1. Polymers ............................................................................................................................................. 1


1.1.1. Polymer History ........................................................................................................................... 2
1.1.2. Polymer Uses ............................................................................................................................... 3
1.2. Polyolefins .......................................................................................................................................... 4
1.3. Polyethylene........................................................................................................................................ 6
1.3.1. History of Polyethylene ............................................................................................................... 7
1.3.2. Polyethylene Structures................................................................................................................ 8
1.3.3. Polyethylene Uses...................................................................................................................... 11
1.4. Polypropylene ................................................................................................................................... 11
1.4.1. History of Polypropylene ........................................................................................................... 12
1.4.2. Polypropylene Structure............................................................................................................. 12
1.4.3. Polypropylene Uses ................................................................................................................... 14
1.5. Polymer Market................................................................................................................................. 14
1.6. Polyolefin Market ............................................................................................................................. 17
1.7. Research Goals.................................................................................................................................. 22
1.8. References......................................................................................................................................... 23

2. Industrial Polymerization Processes ....................................................................................................... 25

2.1. Introduction....................................................................................................................................... 25
2.2. Polyolefin Production Processes ....................................................................................................... 25
2.3. Processes Overview .......................................................................................................................... 33
2.3.1. High Pressure Processes............................................................................................................. 34
2.3.1.1. Autoclave Process............................................................................................................... 34
2.3.1.2. Tubular Process .................................................................................................................. 35
2.3.2. Solution Processes ..................................................................................................................... 36
2.3.2.1. Dowlex (Propietary: Dow).................................................................................................. 37
2.3.2.2. Sclairtech/ AST (Licensor: Nova)....................................................................................... 38
2.3.2.3. Compact (Licensor: Stamicarbon-SABIC) ......................................................................... 39
2.3.3. Slurry Processes ......................................................................................................................... 40
2.3.3.1. Phillips Slurry Loop (Licensor: Chevron Phillips) ............................................................. 41
2.3.3.2. Solvay Loop (Licensor: Solvey) ......................................................................................... 41
2.3.3.3. Hostalen (Licensor: LyondellBasell) .................................................................................. 43
2.3.3.4. Nissan (Licensor: Nissan) ................................................................................................... 43
2.3.3.5. CX Process (Licensor: Mitsui)............................................................................................ 44
2.3.4. Gas-phase Processes .................................................................................................................. 45
x

2.3.4.1. Unipol (Licensor: Univation).............................................................................................. 47


2.3.4.2. BASF Novolen Process ...................................................................................................... 48
2.3.4.3. Amoco – Chisso Horizontal Stirred-Bed Process ............................................................... 50
2.3.4.4. Innovene (Licensor: BP Amoco) ........................................................................................ 52
2.3.4.5. Spherilene (Licensor: Basell).............................................................................................. 53
2.4. References......................................................................................................................................... 55

3. Catalytic Polymerization Kinetics ........................................................................................................... 57

3.1. Introduction....................................................................................................................................... 57
3.2. Catalytic Systems .............................................................................................................................. 57
3.2.1. Polyolefin Catalytic Systems Historical Survey ........................................................................ 58
3.2.2. Ziegler-Natta Catalysts .............................................................................................................. 59
3.2.2.1. Stereochemical Control by Z-N Catalysts .......................................................................... 63
3.2.2.2. Mechanism of Z-N Catalyzed Polymerization.................................................................... 63
3.2.2.3. Evolution of Z-N catalysts .................................................................................................. 64
3.2.3. Metallocene Catalysts ................................................................................................................ 67
3.2.3.1. Stereoregulation .................................................................................................................. 68
3.2.3.2. Mechanism of Metallocene Catalyzed Polymerization....................................................... 69
3.2.4. Cromium Catalysts..................................................................................................................... 72
3.2.4.1. Mechanism of Chromium Catalyzed Polymerization ......................................................... 73
3.3. Generalized Polymerization Kinetic Scheme.................................................................................... 75
3.3.1. Activation of Potential Sites ...................................................................................................... 76
3.3.2. Chain Initiation .......................................................................................................................... 76
3.3.3. Chain Propagation...................................................................................................................... 77
3.3.4. Chain Transfer ........................................................................................................................... 77
3.3.5. Site Transformation ................................................................................................................... 77
3.3.6. Catalyst Deactivation ................................................................................................................. 78
3.4. Method of Moments.......................................................................................................................... 80
3.5. Calculation of Molecular Properties.................................................................................................. 84
3.6. Molecular Weight Distribution Reconstruction ................................................................................ 86
3.7. Slurry Loop Reactor Kinetic Scheme................................................................................................ 88
3.7.1. Kinetic Modelling Literature Review ........................................................................................ 89
3.7.2. Kinetic Rate Constants Sensitivity Analysis.............................................................................. 93
3.7.2.1. Definition............................................................................................................................ 93
3.7.2.2. Sensitivity Analysis Procedure ........................................................................................... 94
3.7.3. Simplified Slurry Loop Reactor Kinetic Scheme..................................................................... 101
3.7.3.1. Simplified Production/ Consumption Rates...................................................................... 101
3.7.4. Deconvolution of Molecular Weight Distribution ................................................................... 104
3.7.5. Kinetic Rate Constants Estimation Procedure ......................................................................... 110
xi

3.8. Notation........................................................................................................................................... 113


3.9. References....................................................................................................................................... 116

4. Slurry-loop Reactor Modelling.............................................................................................................. 121

4.1. Introduction..................................................................................................................................... 121


4.2. Slurry-loop Polymerization Process................................................................................................ 121
4.2.1. The Slurry-phase Olefin Polymerization Cascade-loop Reactor Process ................................ 124
4.2.2. Settling Legs Operation ........................................................................................................... 128
4.3. Process Models ............................................................................................................................... 131
4.3.1. Multi-scale Modelling Approach ............................................................................................. 132
4.3.1.1. Microscale Kinetic Modelling .......................................................................................... 133
4.3.1.2. Mesoscale Physical Modelling ......................................................................................... 133
4.3.1.3. Macroscale Reactor Modelling ......................................................................................... 134
4.3.2. Slurry-loop Reactor Modelling Literature Review .................................................................. 134
4.4. Model Development........................................................................................................................ 136
4.4.1. Model Development Overview ................................................................................................ 136
4.4.2. Solution Procedure................................................................................................................... 139
4.4.3. CSTR Modelling Approach ..................................................................................................... 140
4.4.4. Kinetic Modelling .................................................................................................................... 141
4.4.5. Dynamic Mass Balances .......................................................................................................... 142
4.4.6. Settling Leg Modelling ............................................................................................................ 146
4.4.7. Dynamic Energy Balances ....................................................................................................... 149
4.4.7.1. Energy Balance for Reactor Mixture ................................................................................ 149
4.4.7.2. Energy Balance for Coolant Jacket................................................................................... 151
4.4.7.3. Overall Heat Transfer Coefficient..................................................................................... 152
4.4.8. Reactor Control System ........................................................................................................... 154
4.4.8.1. PID Controller .................................................................................................................. 156
4.4.8.2. PID Controller Tuning ...................................................................................................... 157
4.4.9. Thermodynamic Considerations .............................................................................................. 158
4.4.9.1. Equilibrium Sorbed Monomer Concentration................................................................... 159
4.4.9.2. Sanchez-Lacombe Equation of State ................................................................................ 159
4.4.9.3. S-L EOS Parameter Estimation ........................................................................................ 162
4.4.9.4. Crystallinity Calculation ................................................................................................... 166
4.4.9.5. Polymer density ................................................................................................................ 168
4.5. Calculation of Physical Properties .................................................................................................. 170
4.5.1. Slurry Viscosity ....................................................................................................................... 170
4.5.2. Thermal Conductivity .............................................................................................................. 172
4.5.2.1. Thermal Conductivity of Gases ........................................................................................ 172
4.5.2.2. Thermal Conductivity of Liquids...................................................................................... 173
xii

4.5.2.3. Thermal Conductivity of Polyethylene ............................................................................. 174


4.5.2.4. Mixture Thermal Conductivity ......................................................................................... 174
4.5.3. Constant Pressure Heat Capacity ............................................................................................. 174
4.5.3.1. Constant Pressure Heat Capacity of Ethylene................................................................... 174
4.5.3.2. Constant Pressure Heat Capacity of 1-Hexene ................................................................. 176
4.5.3.3. Constant Pressure Heat Capacity of i-butane.................................................................... 178
4.5.3.4. Constant Pressure Heat Capacity of Polyethylene ............................................................ 180
4.5.3.5. Mixture Constant Pressure Heat Capacity ........................................................................ 182
4.5.4. Heat of Polymerization ............................................................................................................ 182
4.5.5. Physical Properties of Coolant Water ...................................................................................... 183
4.6. Rheological Properties .................................................................................................................... 187
4.7. Particle Size Distribution Calculation ............................................................................................. 191
4.7.1. Introduction.............................................................................................................................. 191
4.7.2. Calculation of PSD .................................................................................................................. 193
4.7.3. The Orthogonal Collocation on Finite Elements Method ........................................................ 197
4.8. Notation........................................................................................................................................... 200
4.9. References....................................................................................................................................... 205

5. Results and Discussion ........................................................................................................................... 215

5.1. Introduction..................................................................................................................................... 215


5.2. Simulation Results .......................................................................................................................... 216
5.2.1. Slurry-loop Reactor Modelling ................................................................................................ 216
5.2.1.1. Process Startup.................................................................................................................. 216
5.2.1.2. Effect of the Total Number of Settling Legs in Operation................................................ 236
5.2.1.3. Effect of the Co-monomer Inflow Rate ............................................................................ 238
5.2.1.4. Effect of the Hydrogen Inflow Rate.................................................................................. 240
5.2.2. Simulation of an Industrial Reactor ......................................................................................... 242
5.2.2.1. Molecular Properties......................................................................................................... 242
5.2.2.2. Reactor Startup ................................................................................................................. 244
5.2.2.3. 28 Days Operation ............................................................................................................ 263
5.3. References....................................................................................................................................... 281

6. Concluding Remarks and Future Work ............................................................................................... 283

6.1. Concluding Remarks....................................................................................................................... 283


6.2. Suggestions for Future Work .......................................................................................................... 284 
xiii

List of Tables

Table 2.1: High density polyethylene processes.............................................................................................. 28


Table 2.2: Linear low density polyethylene processes. ................................................................................... 29
Table 2.3: Polypropylene processes. ............................................................................................................... 30
Table 2.4: The technical applicability of processes for polyethylene manufacture. ........................................ 31
Table 3.1: Comparison of different types of polyethylene. ............................................................................. 62
Table 3.2: Generalized polymerization kinetic scheme................................................................................... 79
Table 3.3: Numerical values for kinetic rate constants.................................................................................... 90
Table 3.4: Proposed kinetic mechanisms adopted by previous workers on the field. ..................................... 93
Table 3.5: The simplified kinetic scheme...................................................................................................... 101
Table 3.6: Deconvolution of various TPRF grades. ...................................................................................... 110
Table 3.7: Numerical values of kinetic rate constants for a four-site Ziegler-Natta ethylene1-hexene
co-polymerization mechanism. ............................................................................................................ 112
Table 4.1: Values of S-L EOS parameters..................................................................................................... 162
Table 4.2: S-L EOS binary interaction parameters........................................................................................ 163
Table 4.3: Ethylene heat capacity.................................................................................................................. 174
Table 4.4: Heat capacity of saturated ethylene. ............................................................................................. 175
Table 4.5: Liquid 1-hexene pressure heat capacity........................................................................................ 177
Table 4.6: Heat capacity of liquid 1-hexene (Kalinowska and Woycicki, 1985). ......................................... 177
Table 4.7: Liquid 1-hexene heat capacity...................................................................................................... 179
Table 4.8: Heat capacity of liquid i-butane (Perry and Green, 1984, 3-199)................................................. 179
Table 4.9: Polyethylene heat capacity. .......................................................................................................... 180
Table 4.10: Heat of polymerization for polyethylene.................................................................................... 182
Table 4.11: Physical properties of water (deMan, J. M., 1993)..................................................................... 184
Table 4.12: Numerical values of the rheological model parameters for polyethylene. ................................. 190
xiv
xv

List of Figures

Figure 1.1: Polymer demand 2008: breakdown by types. ................................................................................. 4


Figure 1.2: Total consumption of energy and raw materials in kilogram equivalents of oil, to obtain 1 kg
of material indicated. ............................................................................................................................... 6
Figure 1.3: Schematic molecular structure of polyethylene. ........................................................................... 10
Figure 1.4: Commercial uses of polyethylene. ................................................................................................ 11
Figure 1.5: Polypropylene chain structures. .................................................................................................... 13
Figure 1.6: Commercial uses of polypropylene............................................................................................... 14
Figure 1.7: Polymer production....................................................................................................................... 16
Figure 1.8: World polymer production 2008: By country and region. ............................................................ 16
Figure 1.9: Polymer demand 2008: Breakdown by types................................................................................ 17
Figure 1.10: PE product map........................................................................................................................... 18
Figure 1.11: PE and PP growth. ...................................................................................................................... 19
Figure 1.12: PE and PP consumption. ............................................................................................................. 19
Figure 1.13: Polyolefins versus other plastics. ................................................................................................ 20
Figure 1.14: Polyolefin demand 2008: Breakdown by types........................................................................... 20
Figure 1.15: European polyolefin producers. .................................................................................................. 21
Figure 1.16: Global polyolefin producers........................................................................................................ 21
Figure 2.1: Classification of polyolefin processes........................................................................................... 26
Figure 2.2: Evolution of polyolefin improvements. ........................................................................................ 27
Figure 2.3: Capabilities of different processes in terms of density and melt index. ........................................ 32
Figure 2.4: Effect of polymer density and MFI on end-use properties (Foster, 1991). ................................... 32
Figure 2.5: Generic olefin polymerization process.......................................................................................... 34
Figure 2.6: Autoclave process flowsheet......................................................................................................... 35
Figure 2.7: Tubular process flowsheet. ........................................................................................................... 36
Figure 2.8: Dowlex process flowsheet. ........................................................................................................... 38
Figure 2.9: Sclairtech process flowsheet. ........................................................................................................ 39
Figure 2.10: Compact process flowsheet......................................................................................................... 40
Figure 2.11: Phillips slurry loop flowsheet...................................................................................................... 41
Figure 2.12: Solvay loop flowsheet................................................................................................................. 42
Figure 2.13: Hostalen process flowsheet. ........................................................................................................ 43
Figure 2.14: Nissan process flowsheet. ........................................................................................................... 44
Figure 2.15: CX process flowsheet.................................................................................................................. 45
Figure 2.16: Unipol process flowsheet. ........................................................................................................... 48
Figure 2.17: Novolen Vertical bed process flowsheet..................................................................................... 50
Figure 2.18: Horizontal bed process flowsheet. .............................................................................................. 52
Figure 2.19: Innovene process flowsheet. ....................................................................................................... 53
Figure 2.20: Sperilene process flowsheet. ....................................................................................................... 54
xvi

Figure 3.1: Ziegler-Natta Catalytic System. .................................................................................................... 60


Figure 3.2: The Cossee mechanism. ................................................................................................................ 64
Figure 3.3: The relative yield improvement. ................................................................................................... 65
Figure 3.4: Bis-cyclopentadienyl-titanium chloride. ....................................................................................... 67
Figure 3.5: Mechanism for metallocene polymerization according to Kaminsky’s model. ............................ 70
Figure 3.6: Mechanism for metallocene polymerization accordind to Corradini’s model. ............................. 71
Figure 3.7: Supported chromium catalysts for ethylene polymerization. ........................................................ 73
Figure 3.8: Polymerization steps for ethylene polymerization over a chromium oxide based catalyst. .......... 74
Figure 3.9: Possible co-catalyst reactions for a chromium oxide based catalyst. ............................................ 74
Figure 3.10: Overall MWD given by the weighted sum of polymer fractions produced over the different
catalyst active sites................................................................................................................................. 88
Figure 3.11: Effect of Ka on MWD. ................................................................................................................ 95
Figure 3.12: Effect of K01 on MWD................................................................................................................ 95
Figure 3.13: Effect of K02 on MWD................................................................................................................ 96
Figure 3.14: Effect of Kp11 on MWD............................................................................................................... 96
Figure 3.15: Effect of Kp12 on MWD............................................................................................................... 97
Figure 3.16: Effect of Kp21 on MWD............................................................................................................... 97
Figure 3.17: Effect of Kp22 on MWD............................................................................................................... 98
Figure 3.18: Effect of KtrSP on MWD. ............................................................................................................. 98
Figure 3.19: Effect of KtrH2 on MWD.............................................................................................................. 99
Figure 3.20: Effect of Kd on MWD. ................................................................................................................ 99
Figure 3.21: Effect of kinetic rate constants on number average molecular weight, Mn. .............................. 100
Figure 3.22: Effect of kinetic rate constants on polymerization rate. ............................................................ 100
Figure 3.23: Effect of number of different catalyst active site types on % deviation of the predicted
MWD to the actual MWD.................................................................................................................... 108
Figure 3.24: Effect of number of different catalyst active site types on the derivative of the deviation of
the predicted MWD to the actual MWD. ............................................................................................. 108
Figure 3.25:GPC deconvolution results for a representative ethylene/1-hexene co-polymer grade from
TPRF industrial plant........................................................................................................................... 109
Figure 4.1: Schematic representation of an industrial slurry-phase olefin catalytic polymerization
cascade-loop reactor series. ................................................................................................................. 123
Figure 4.2: Settling Leg Operation Cycle...................................................................................................... 129
Figure 4.3: Schematic representation of loop reactor modelling approach. .................................................. 138
Figure 4.4: Process modelling logic diagram. ............................................................................................... 139
Figure 4.5: The recycle reactor...................................................................................................................... 141
Figure 4.6: Schematic representation of the settling legs employed for product withdrawal. ....................... 146
Figure 4.7: Face-centered cubic packing arrangement. ................................................................................. 148
Figure 4.8: Energy Balance Calculations. ..................................................................................................... 149
xvii

Figure 4.9: Slurry-loop reactor control system.............................................................................................. 155


Figure 4.10: Solubility of ethylene in LLDPE-1-Hexene at different temperatures...................................... 164
Figure 4.11: Solubility of 1-hexene in LLDPE-1-Hexene at different temperatures..................................... 165
Figure 4.12: Solubility of 1-hexene in LLDPE-1-Hexene at different temperatures..................................... 165
Figure 4.13: Semi-crystalline polymer structure. .......................................................................................... 166
Figure 4.14: Literature review on the effect of 1-hexene incorporation on the crystallinity of the
produced polyethylene. ........................................................................................................................ 167
Figure 4.15: Effect of temperature on polymer specific volume. .................................................................. 169
Figure 4.16: Effect of temperature on polyethylene density.......................................................................... 169
Figure 4.17: Effect of solids volume fraction on reactor-phase viscosity. .................................................... 171
Figure 4.18: Ethylene heat capacity. ............................................................................................................. 175
Figure 4.19: 1-Hexene heat capacity. ............................................................................................................ 178
Figure 4.20: i-butane heat capacity................................................................................................................ 180
Figure 4.21: Polyethylene heat capacity........................................................................................................ 181
Figure 4.22: Effect of temperature on water density. .................................................................................... 185
Figure 4.23: Effect of temperature on water specific heat............................................................................. 185
Figure 4.24: Effect of temperature on water viscosity................................................................................... 186
Figure 4.25: Effect of temperature on water thermal conductivity................................................................ 186
Figure 5.1: Dynamic evolution of the ethylene inflow rate and number of settling legs in the first (a.)
and second (b.) loop reactor of the series during process startup......................................................... 218
Figure 5.2: Dynamic evolution of the operating pressure in the first reactor. ............................................... 219
Figure 5.3: Dynamic evolution of the operating pressure in the second reactor............................................ 220
Figure 5.4: Dynamic evolution of the ethylene mass fraction and catalyst inflow rate in the first reactor.... 221
Figure 5.5: Dynamic evolution of the ethylene mass fraction and catalyst inflow rate in the second
reactor. ................................................................................................................................................. 221
Figure 5.6: Dynamic evolution of the 1-hexene and hydrogen off-gas composition in the first reactor. ...... 223
Figure 5.7: Dynamic evolution of the hydrogen mass fraction in the first reactor. ....................................... 223
Figure 5.8: Dynamic evolution of solids concentration and iso-butane inflow rate in the first reactor......... 224
Figure 5.9: Dynamic evolution of the solids concentration and iso-butane inflow rate in the second
reactor. ................................................................................................................................................. 224
Figure 5.10: Dynamic evolution of density in the first reactor...................................................................... 225
Figure 5.11: Dynamic evolution of density in the second reactor. ................................................................ 225
Figure 5.12: Dynamic evolution of the solids enhancement factor and valve opening cycling period in
the first reactor. .................................................................................................................................... 226
Figure 5.13: Dynamic evolution of the solids enhancement factor and valve opening cycling period in
the second reactor. ............................................................................................................................... 227
Figure 5.14: Dynamic evolution of temperature in the first and second reactor............................................ 228
xviii

Figure 5.15: Dynamic evolution of the coolant’s inlet temperature and the temperature difference
between the coolant outlet and inlet values in the first reactor. ........................................................... 228
Figure 5.16: Dynamic evolution of the coolant’s inlet temperature and the temperature difference
between the coolant outlet and inlet values in the second reactor........................................................ 229
Figure 5.17: Dynamic evolution of polymer production rate at the exit of the first and second reactor of
the series. ............................................................................................................................................. 230
Figure 5.18: Dynamic evolution of product density at the exit of the first and second reactor of the
series. ................................................................................................................................................... 230
Figure 5.19: Dynamic evolution of average molecular weight at the exit of the first and second reactor
of the series. ......................................................................................................................................... 231
Figure 5.20: Dynamic evolution of polydispersity at the exit of the first and second reactor of the series. .. 232
Figure 5.21: Dynamic evolution of the MWD of the polymer produced in the first (a.) and in the second
reactor (b.) and overall MWD at the exit of the second reactor of the series (c.) (startup case). ......... 233
Figure 5.22: Dynamic evolution of the complex viscosity of the polymer produced in the first reactor
(startup case). ....................................................................................................................................... 235
Figure 5.23: Dynamic evolution of the complex viscosity of the polymer at the exit of the second reactor
(startup case). ....................................................................................................................................... 235
Figure 5.24: Dynamic evolution of the solids enhancement factor and valve opening cycling period in
the second reactor. ............................................................................................................................... 236
Figure 5.25: Dynamic evolution of the solids concentration and iso-butane inflow rate in the second
reactor. ................................................................................................................................................. 237
Figure 5.26: Dynamic evolution of the number average molecular weight and polydispersity index in the
second reactor. ..................................................................................................................................... 238
Figure 5.27: Dynamic evolution of the co-polymer composition in the first and second reactor of the
series. ................................................................................................................................................... 239
Figure 5.28: Dynamic evolution of the product density in the first and second reactor of the series............ 240
Figure 5.29: Dynamic evolution of the MWD and complex viscosity in the first reactor (effect of
hydrogen concentration). ..................................................................................................................... 241
Figure 5.30: Dynamic evolution of the MWD and complex viscosity in the second reactor (effect of
hydrogen concentration). ..................................................................................................................... 241
Figure 5.31: Molecular weight distribution of polymer at the exit of the first and second reactor of the
series, regarding a low-co-monomer content polymer grade. .............................................................. 243
Figure 5.32: Molecular weight distribution of polymer at the exit of the first and second reactor of the
series, regarding a medium-co-monomer content polymer grade. ....................................................... 243
Figure 5.33: Molecular weight distribution of polymer at the exit of the first and second reactor of the
series, regarding a high-co-monomer content polymer grade.............................................................. 244
Figure 5.34: Dynamic evolution of ethylene feed in the first reactor (Startup case). .................................... 245
Figure 5.35: Dynamic evolution of hexane feed in the first reactor (Startup case). ...................................... 246
xix

Figure 5.36: Dynamic evolution of hydrogen feed in the first reactor (Startup case).................................... 246
Figure 5.37: Dynamic evolution of iso-butane feed in the first reactor (Startup case). ................................. 247
Figure 5.38: Dynamic evolution of catalyst feed in the first reactor (Startup case). ..................................... 247
Figure 5.39: Dynamic evolution of ethylene off-gas composition in the first reactor (Startup case). ........... 248
Figure 5.40: Dynamic evolution of hexane off –gas composition in the first reactor (Startup case)............. 248
Figure 5.41: Dynamic evolution of hydrogen off-gas composition in the first reactor (Startup case). ......... 249
Figure 5.42: Dynamic evolution of solids concentration in the first reactor (Startup case). ......................... 249
Figure 5.43: Dynamic evolution of reactor phase density in the first reactor (Startup case)......................... 250
Figure 5.44: Dynamic evolution of temperature in the first reactor (Startup case). ...................................... 250
Figure 5.45: Dynamic evolution of coolant’s inlet temperature in the first reactor (Startup case)................ 251
Figure 5.46: Dynamic evolution of and the temperature difference between the coolant outlet and inlet
values in the first reactor (Startup case)............................................................................................... 251
Figure 5.47: Dynamic evolution of the valve opening cycling period in the first reactor (Startup case). ..... 252
Figure 5.48: Dynamic evolution of polymer production rate in the first reactor (Startup case). ................... 252
Figure 5.49: Dynamic evolution of product density in the first reactor (Startup case).................................. 253
Figure 5.50: Dynamic evolution of ethylene feed in the second reactor (Startup case). ............................... 253
Figure 5.51: Dynamic evolution of hydrogen feed in the second reactor (Startup case)............................... 254
Figure 5.52: Dynamic evolution of iso-butane feed in the second reactor (Startup case). ............................ 254
Figure 5.53: Dynamic evolution of ethylene off-gas composition in the first reactor (Startup case). ........... 255
Figure 5.54: Dynamic evolution of hexane off –gas composition in the second reactor (Startup case). ....... 255
Figure 5.55: Dynamic evolution of hydrogen off-gas composition in the second reactor (Startup case)...... 256
Figure 5.56: Dynamic evolution of solids concentration in the second reactor (Startup case)...................... 256
Figure 5.57: Dynamic evolution of reactor phase density in the second reactor (Startup case). ................... 257
Figure 5.58: Dynamic evolution of temperature in the second reactor (Startup case)................................... 257
Figure 5.59: Dynamic evolution of coolant’s inlet temperature in the second reactor (Startup case). .......... 258
Figure 5.60: Dynamic evolution of and the temperature difference between the coolant outlet and inlet
values in the second reactor (Startup case). ......................................................................................... 258
Figure 5.61: Dynamic evolution of the valve opening cycling period in the second reactor (Startup case).. 259
Figure 5.62: Dynamic evolution of polymer production rate in the second reactor (Startup case). .............. 259
Figure 5.63: Dynamic evolution of product density in the second reactor (Startup case). ............................ 260
Figure 5.64: Dynamic evolution of number average molecular weight in the first and second reactor
(Startup case). ...................................................................................................................................... 260
Figure 5.65: Dynamic evolution of polydispersity in the first and second reactor (Startup case). ................ 261
Figure 5.66: Dynamic evolution of the PSD in the first reactor (startup case).............................................. 262
Figure 5.67: Dynamic evolution of the PSD in the second reactor (startup case). ........................................ 262
Figure 5.68: Initial catalyst size distribution (Startup case). ......................................................................... 263
Figure 5.69: Dynamic evolution of ethylene feed in the first reactor (28 days case). ................................... 265
Figure 5.70: Dynamic evolution of hexane feed in the first reactor (28 days case). ..................................... 265
xx

Figure 5.71: Dynamic evolution of hydrogen feed in the first reactor (28 days case)................................... 266
Figure 5.72: Dynamic evolution of iso-butane feed in the first reactor (28 days case). ................................ 266
Figure 5.73: Dynamic evolution of catalyst feed in the first reactor (28 days case)...................................... 267
Figure 5.74: Dynamic evolution of ethylene off-gas composition in the first reactor (28 days case). .......... 267
Figure 5.75: Dynamic evolution of hexane off –gas composition in the first reactor (28 days case)............ 268
Figure 5.76: Dynamic evolution of hydrogen off-gas composition in the first reactor (28 days case).......... 268
Figure 5.77: Dynamic evolution of solids concentration in the first reactor (28 days case).......................... 269
Figure 5.78: Dynamic evolution of reactor phase density in the first reactor (28 days case). ....................... 269
Figure 5.79: Dynamic evolution of temperature in the first reactor (28 days case)....................................... 270
Figure 5.80: Dynamic evolution of coolant’s inlet temperature in the first reactor (28 days case). .............. 270
Figure 5.81: Dynamic evolution of and the temperature difference between the coolant outlet and inlet
values in the first reactor (28 days case). ............................................................................................. 271
Figure 5.82: Dynamic evolution of the valve opening cycling period in the first reactor (28 days case)...... 271
Figure 5.83: Dynamic evolution of polymer production rate in the first reactor (28 days case). .................. 272
Figure 5.84: Dynamic evolution of product density in the first reactor (28 days case). ................................ 272
Figure 5.85: Dynamic evolution of ethylene feed in the second reactor (28 days case)................................ 273
Figure 5.86: Dynamic evolution of hydrogen feed in the second reactor (28 days case). ............................. 273
Figure 5.87: Dynamic evolution of iso-butane feed in the second reactor (28 days case)............................. 274
Figure 5.88: Dynamic evolution of ethylene off-gas composition in the first reactor (28 days case). .......... 274
Figure 5.89: Dynamic evolution of hexane off –gas composition in the second reactor (28 days case). ...... 275
Figure 5.90: Dynamic evolution of hydrogen off-gas composition in the second reactor (28 days case). .... 275
Figure 5.91: Dynamic evolution of solids concentration in the second reactor (28 days case). .................... 276
Figure 5.92: Dynamic evolution of reactor phase density in the second reactor (28 days case). .................. 276
Figure 5.93: Dynamic evolution of temperature in the second reactor (28 days case). ................................. 277
Figure 5.94: Dynamic evolution of coolant’s inlet temperature in the second reactor (28 days case). ......... 277
Figure 5.95: Dynamic evolution of and the temperature difference between the coolant outlet and inlet
values in the second reactor (28 days case). ........................................................................................ 278
Figure 5.96: Dynamic evolution of the valve opening cycling period in the second reactor (28 days case). 278
Figure 5.97: Dynamic evolution of polymer production rate in the second reactor (28 days case)............... 279
Figure 5.98: Dynamic evolution of product density in the second reactor (28 days case). ........................... 279
Figure 5.99: Dynamic evolution of number average molecular weight in the first and second reactor
(28 days case)....................................................................................................................................... 280
Figure 5.100: Dynamic evolution of polydispersity in the first and second reactor (28 days case). ............. 280
Chapter 1: Introduction 1

Chapter 1

1. Introduction

1.1. Polymers

Historians frequently classify the early ages of man according to the materials that he
used for making his implements and other basic necessities. The most well known of these
periods are the Stone Age, the Iron Age and the Bronze Age. Such a system of
classification cannot be used to describe subsequent periods for with the passage of time
man learnt to use more materials. By the time of the ancient civilisations of Egypt,
Babylonia and Greece, man was employing a range of metals, stones, woods, ceramics,
glasses, skins, horns and fibres. Until the 19th century, man’s inanimate possessions, his
home, his tools, his furniture, were made from varieties of these eight classes of material.

During the last century and a half, a new class of material has been introduced which
has not only challenged the older materials for their well-established uses but has also
made possible new products which have helped to extend the range of activities of
mankind. Synthetic polymers have revolutionized our life. Today, without this class of
materials, it is difficult to conceive how such everyday features of modern life could ever
have been developed (Brydson, 1999). Plastics, fibers, elastomers, adhesives, rubber,
protein, cellulose-these all common terms in our modern vocabulary, and all a part of the
fascinating world of polymer chemistry. Innumerable examples of synthetic polymers may
be cited. Synthetic polymer materials such as nylon, polyethylene, teflon, and silicone have
formed the basis for a burgeoning polymer industry. Synthetic polymers today find
2 Chapter 1: Introduction

application in nearly every industry and area of life. Polymers are widely used as adhesives
and lubricants, as well as structural components for products ranging from children's toys
to aircraft. Moreover, they have been employed in a variety of biomedical applications
ranging from implantable devices to controlled drug delivery. Conclusively, it would not
be an exaggeration to say that we live in the Polymer age.

Polymers are macromolecules built up by the linking together of large numbers of


much smaller molecules. The small molecules that combine with each other to form
polymer molecules are termed monomers, and the reactions by which they combine are
termed polymerizations. There may be hundreds, thousands, tens of thousands, or more
monomer molecules linked together in a polymer molecule. When one speaks of polymers,
he is concerned with materials whose molecular weights may reach into the hundreds of
thousands or millions (Odian, 2004).

The word polymer is derived from the Greek words πολυ- (-poly-) meaning ‘many’;
and μέρος (-meros-) meaning ‘part’ (polymeric comes directly from the Greek ‘πολυμερές’
which means ‘having many parts’). The term was coined in 1833 by Joens Jakob Berzelius,
although his definition of a polymer was different from the modern definition. Initially,
according to Berzelius, the term polymer described organic compounds which shared
identical empirical formulas but differed in overall molecular weight. The larger of the
compounds was described as ‘polymers’ of the smallest. Thus, according to this -now
obsolete- definition, glucose, C6H12O6, would be a polymer of formaldehyde, CH2O.

1.1.1. Polymer History

Starting in 1811, Henri Braconnot did pioneering work in derivative cellulose


compounds, perhaps the earliest important work in polymer science. The development of
vulcanization later in the nineteenth century improved the durability of the natural polymer
rubber, signifying the first popularized semi-synthetic polymer.

The first truly synthetic polymer to be used on industrial-scale, was a phenol-


formaldehyde resin produced at precisely controlled temperature and pressure. Developed
in 1907 by the belgian-born chemist Leo Baekeland (who had already earned considerable
success by his invention of light-sensitive photographic paper), it was known commercially
as Bakelite (publicly introduced in 1909). By the decade of 1920, Bakelite had found its
Chapter 1: Introduction 3

way into a wide spectrum of consumer products, and its inventor had achieved the ultimate
visibility, featuring on a cover of Time. Other polymers, notably alkyd (polyester) paints
and polybutadiene rubber, were introduced about the same time. Yet despite such
commercial successes, most scientists had no clear concept of polymer structure. The
prevailing theory was that polymers were aggregates of small molecules, much like
colloids, but were held together by some ‘mysterious’ secondary force.

This aggregation or association theory eventually gave way, with no small amount of
resistance, to the theories of the german chemist Herman Staudinger, who attributed to the
remarkable properties of polymers to ordinary intermolecular forces between molecules of
very high molecular weight. Staudinger was also the first to introduce the term
‘macromolecül’ (i.e., macromolecule). In recognition of his contributions, Staudinger was
awarded the Nobel Prize in Chemistry in 1953. In the 1930’s, the brilliant work of the
American chemist Wallace Hume Carrothers placed the theories of Staudinger on a firm
experimental basis and led to the commercial development of neoprene rubber and
polyamide (nylon) fibers.

World War II led to significant advances in polymer chemistry, particularly with the
development of synthetic rubber when the natural rubber-growing regions of the Far East
became inaccessible to the Allies. Among the more significant developments of the
postwar years, was the discovery by Karl-Ziegler in Germany of new coordination
catalysts for initiating polymerization reactions and the application by Giulio Natta in Italy
of these new systems to development of polymers having controlled stereochemistry. Their
work has revolutionized the polymer industry, for these so-called stereoregular polymers
have mechanical properties superior in most instances to those of non-stereoregular
polymers. The importance of their discoveries was recognized by the award of the Nobel
Prize in Chemistry jointly to Ziegler and Natta in 1963. Equally significant was the work
of Paul Flory (Nobel Prize 1974), who established a quantitative basis for polymer
behaviour, whether it be physical properties of macromolecules in solution or in bulk or
such chemical phenomena as crosslinking and chain transfer (Stevens, 1999).

1.1.2. Polymer Uses

Nowadays, it is difficult to find an aspect of our lives that is not affected by


polymers. It should be noted that, just 50 years ago, materials we now take for granted
4 Chapter 1: Introduction

were non-existent. Polymers exhibit a very wide range of properties which allow for their
extensive use in society. Packaging remains the biggest end-use for polymers (38%)
followed by building and construction (21%), automotive (7%) and electrical and
electronic uses (6%). Other applications, which include medical and leisure, use 28%.
(Figure 1.1/ Source: PlasticsEurope Market Research Group (PEMRG)).

Packaging
38%
Others
28%

E&E
6% Building and
Construction
Automotive
7% 21%

Figure 1.1: Polymer demand 2008: breakdown by types.

1.2. Polyolefins

Among the broad variety of polymeric materials, polyolefins have received a special
attention during recent years. Polyolefins are among the most important commodity
plastics today due to their low production costs, reduced environmental impact, and a very
wide range of applications. They are found in products as diverse as prosthetic implants,
gas pipelines, automobile parts and accessories, synthetic fabrics, films, containers, and
toys. It may seem surprising that polymers composed only of carbon and hydrogen atoms
can be so flexible, but the versatility of polyolefins can be easily explained by the way the
monomer molecules ethylene, propylene, and higher a-olefins are connected to form the
polymer chains. It is true for any polymer, but dramatically so for polyolefins, that chain
microstructure is the key to understanding their physical properties (Meyer and Keurentjes,
2005).
Chapter 1: Introduction 5

Polyolefin is a polymer produced from a simple alkene (also called an olefin, with
the general formula CnH2n) as a monomer. This type of polymers includes large volume
materials such as homopolymers of ethylene (polyethylene, PE), and propylene
(polypropylene, PP), co-polymers of ethylene, propylene, higher olefins dienes and cyclic
olefins, as well as specialty materials such as ethylene-propylene elastomers (EPR and
EP(D)M) and polybutene-1 (PB-1).

We may state, even considering the universality of materials, that polyolefins are
unique in their ability to perfectly match the basic prerequisites for the successful growth
of a material (Galli, 1994):

 cost/ performance balance, and,


 economic production with consideration given to environmental factors.

Polyolefins (i.e., polyethylene and polypropylene) are the world’s mostly produced
and fastest growing polymer family. The basis of their dynamic development and still
tremendous potential lies in:

1) Their versatility with respect to physical and mechanical properties and


applications. Polyolefins are available in many varieties. They range from rigid
materials, which are used for car parts, to soft materials such as flexible fibres.
Some are as clear as glass; others are completely opaque. Some, such as
microwave food containers, have high heat resistance while others melt easily.

2) Their nontoxicity and bioacceptability.

3) The energy savings during their production and use, in comparison with other
materials.

4) Their low cost and the easily available raw materials.

5) Their economic, versatile, and non-polluting production. In Figure 1.2 (Galli,


1994), the total consumption of energy and raw materials in kilogram
equivalents of oil, to obtain 1 kg of different materials is presented. It is clear
that, for the production of 1 kg of PE or PP is consumed at least 7 and 3 times
less energy comparing to aluminium and steel production, respectively.
6 Chapter 1: Introduction

Aluminium

Copper

Steel

PS

HDPE

PVC

LDPE

PP

0 2 4 6 8 10 12 14 16
Mass Fraction of Oil Equivalents to Raw Material

Figure 1.2: Total consumption of energy and raw materials in kilogram equivalents of oil,
to obtain 1 kg of material indicated.

1.3. Polyethylene

Polyethylene is the largest synthetic commodity polymer in terms of annual


production and is widely used throughout the world due to its versatile physical and
chemical properties (Xie et al., 1994). Polyethylene is a generic name for a family of
semicrystalline polymers made from the polymerization of ethylene gas. Despite ethylene's
simple structure, the field of polyethylene is a complex one with a very wide range of types
and many different manufacturing processes. It is produced either in radical polymerization
reactions or in catalytic polymerization reactions. From a comparatively late start,
polyethylene production has increased rapidly to make polyethylene the major tonnage
plastics material worldwide (exceeds 70 Mtonnes capacity in 2007/ Source: CMAI report).

The main attractive features of polyethylene are (Brydson, 1999):

 low price,
 excellent electrical insulation properties over a wide range of frequencies,
 very good chemical resistance,
 good processability, toughness,
 flexibility, and,
 transparency (in thin films of certain grades).
Chapter 1: Introduction 7

1.3.1. History of Polyethylene

In the 1920’s, research into the polymerization of unsaturated compounds such as


vinyl chloride, vinyl acetate, and styrene led to industrial processes being introduced in the
1930’s, but the use of the same techniques with ethylene did not lead to high polymers.
The chance observation in 1933 by an ICI research team that traces of a waxy polymer
were formed when ethylene and benzaldehyde were subjected to a temperature of 170°C
and a pressure of 190 MPa, led to the first patent in 1936 and small-scale production in
1939. The polymers made in this way, by using free radical initiators, were partially
crystalline, and measurement of the density of the product was quickly established as a
means of determining the crystallinity. Due to the side reactions occurring at the high
temperatures employed, the polymer chains were branched, and densities of 0.915–0.925
g/cm3 were typically obtained. The densities of completely amorphous and completely
crystalline polyethylene would be 0.880 g/cm3 and 0100 g/cm3, respectively.

During the 1950’s three research groups working independently discovered three
different catalysts which allowed the production of essentially linear polyethylene at low
pressure and temperature. These polymers had densities in the region of 0.960 g/cm3, and
became known as high-density polyethylenes (HDPE), in contrast to the polymers
produced by the extensively commercialized high-pressure process, which were named
low-density polyethylenes (LDPE). These discoveries laid the basis for the coordination
catalysis of ethylene polymerization, which has continued to diversify. Of the three
discoveries at Standard Oil (Indiana), Phillips Petroleum, and by Karl Ziegler at the Max-
Planck-Institut für Kohlenforschung, the latter two have been extensively commercialized.
More recently the observation that traces of water can dramatically increase the
polymerization rate of certain Ziegler catalysts has led to major developments in soluble
coordination catalysts and later their supported variants.

The coordination catalysts allowed for the first time the co-polymerizaton of
ethylene with other olefins such as butene, which by introducing side branches reduces the
crystallinity and allows a low-density polyethylene to be produced at comparatively low
pressures. Although Du Pont of Canada introduced such a process in 1960, worldwide, the
products remained a small-volume specialty until 1978 when Union Carbide announced
their Unipol process and coined the name linear low-density polyethylene (LLDPE). In
addition to developing a cheaper production process, Union Carbide introduced the
8 Chapter 1: Introduction

concept of exploiting the different molecular structure of the linear product to make
tougher film. Following this lead, LLDPE processes have been introduced by many other
manufacturers. Present market needs, combined with the broad range of polyolefin
applications, have forced the polyolefin industry to operate under frequent grade transition
policies, producing more than one type of product (swing plants).

Additionally, co-polymers are made by both types of process (i.e., free radical and
catalytic polymerization). The free-radical process is used to produce co-polymers of vinyl
acetate, acrylates, methacrylates, and the corresponding acids, but chain transfer prevents
the use of higher olefins because of the drastic reduction in molecular mass of the polymer.
The coordination catalysts are able to co-polymerize olefins, but are deactivated by more
polar materials. Because of the complex interplay of the capabilities of modern plants, it is
convenient to treat separately the products, the catalysts, and the processes (Ullmann’s
Encyclopedia of Industrial Chemistry).

1.3.2. Polyethylene Structures

Most polyethylene molecules contain branches in their chains which are formed
spontaneously, in case of radical polymerization, or deliberately by copolymerization of
ethylene with α−olefins, in case of catalytic polymerization. In general, the density of PE
decreases with an increase in branch numbers; the more branching, the lower the density
(Shirayama et al., 1972, Sinclair, 1983). The crystallinity of PE decreases significantly
with an increase in branch frequency and size. Polyethylene resins are classified according
to their density which partly depends on the type of branching. The American Society for
Testing and Materials (ASTM) has classified PE into four groups:

 I (low density) at 0.910-0.925 g/cm3,


 II (also low density) at 0.926-0.940 g/cm3,
 III (high-density copolymers) at 0.941-0.959 g/cm3, and,
 IV (high-density homopolymer) at 0.960 g/cm3 and above (Redman, 1991).

In the literature, however, PE is normally classified as:


Chapter 1: Introduction 9

 low-density polyethylene (LDPE) at 0.915-0.935 g/cm3, and,


 high-density polyethylene (HDPE) at 0.935- 0.975 g/cm3.

It is noted that low-density polyethylene is further classified as low-density


polyethylene (LDPE) and linear low-density polyethylene (LLDPE) based on polymer
chain microstructure (i.e., molecular architecture) and synthesis processes. The three
different molecular architectures result in a wide variety of physical properties and
molding characteristics. They particularly differ in the degree and type of branching. Thus,
PE is a semicrystalline material, that is, a mixture of interconnected crystalline and
amorphous forms. Branching disrupts the crystallinity, forcing more of the polymer into
the amorphous form. As crystalline phase exhibits higher values of density (i.e., about 1.0
g/cm3) than the one in the amorphous phase (i.e., about 0.87 g/cm3), branching is easily
detected by the density. Figure 1.3 shows schematic structures for the three types of
polyethylene, with the main features exaggerated for emphasis (McDaniel, 2010).

High density polyethylene (HDPE) is essentially free of both long and short
branching and thus has stronger intermolecular forces (although very small amounts may
be deliberately incorporated to achieve specific product targets). Low degree of branching
leads to a white opaque solid material The MWD depends on the catalyst type but is
typically of medium width. It is produced mainly in slurry and gas-phase polymerization
processes (Ullmann’s Encyclopedia of Industrial Chemistry). Medium density
polyethylene (MDPE) is a type of polyethylene defined by a density range of 0.926-0.940
g/cm3. It is less dense than HDPE, which is more common. MDPE has a high degree of
resistance to chemicals and is very easy to keep clean. MDPE can be produced by
chromium/ silica catalysts, Ziegler-Natta catalysts or metallocene catalysts.

Low density polyethylene (LDPE) has a random long-branching structure, with


branches on branches. The short branches are not uniform in length but are mainly four or
two carbon atoms long. The ethyl branches probably occur in pairs (Mattice and Stehling,
1981), and there may be some clustering of other branches (van der Ven, 1990). The
molecular weight distribution (MWD) is moderately broad. LDPE is a translucent solid. It
is more highly branched (both short and long branches) than HDPE and is therefore lower
in crystallinity (40–60% vs. 70–90%) and density (0.91–0.93 g/cm3 vs. 0.931–0.97 g/cm3).
It is noted that LDPE is produced mainly in high-pressure poymerization processes (Odian,
2004). Linear low density polyethylene (LLDPE) is a substantially linear polymer, with
10 Chapter 1: Introduction

significant numbers of short branches, produced mainly by copolymerization of ethylene


with longer-chain olefins. LLDPE has branching of uniform length which is randomly
distributed along a given chain, but there is a spread of average concentrations between
chains, the highest concentrations of branches being generally in the shorter chains
(Mirabella and Ford, 1987). The catalysts used to minimize this effect generally also
produce fairly narrow MWDs. LLDPE is a translucent solid.

PE type LDPE HDPE LLDPE

Molecular
architecture

Branching
High Little or none High
amount

Branching
Long and short Short Short
type

Density,
0.915-0.935 0.935-0.975 0.915-0.935
g/cm3

Melt index 0.5-20 0-100 0.5-60

Main
Film Blow and injection molding Film
application

Production Free radical in molten Slurry, gas phase, solution Slurry, gas phase, solution
process polymer phase phase

Pressure,
69-276 2-5 2-5
Mpa

Figure 1.3: Schematic molecular structure of polyethylene.


Chapter 1: Introduction 11

1.3.3. Polyethylene Uses

The analysis of the end-use applications of polyethylene in the world market is


presented in Figure 1.4 (Source: CMAI report). In general, LDPE (together with LLDPE
which is sold into the same market) is used predominantly for films, not all of which is for
packaging. Because of its greater rigidity and better creep properties, HDPE is used in
more structural applications, and also has important applications in the packaging of
aggressive liquids such as bleach, detergent, and hydrocarbons.

Film
52%

Other non-
extrusion
3%
Other extrusion Injection
2% Molding
12%
Sheet
2% Pipe and
Conduit
Rotational
molding Extrusion 7%
1% Wire and cable Blow Molding Coating
4%
3% 14%

Figure 1.4: Commercial uses of polyethylene.

1.4. Polypropylene

Polypropylene is a synthetic, high molecular mass linear addition polymer of


propene (also called propylene). Commercial interest lies primarily in highly crystalline
PP, together with its further modifications through co-polymerization. Its superior strength
to weight ratio, combined with resistance to chemicals, and thermal stability has rendered
the plastic as most popularly used after polyethylene and polyvinyl chloride.
12 Chapter 1: Introduction

1.4.1. History of Polypropylene

In the 1950’s Karl Ziegler discovered new catalysts which were to revolutionize the
plastics industry. Before then, propene could not be polymerized to high molecular mass
products with the catalysts then available, namely, free-radical, anionic and cationic
systems. Even the most favorable of these yielded only liquids consisting of many isomers
and greaselike materials unsuitable for making hard plastic products. Industrial-scale
production of polypropylene started at Ferrara, Italy, in 1957 with a 6×103 tonne/ year
plant. Nowadays, global capacity exceeds 70×103 tonne/ year (see also 1.6. Polyolefin
Market). There are several reasons for this phenomenal growth. At first propene was
readily available, almost at byproduct prices, from petrochemical cracker plants making
ethylene. The polymer itself was suitable for a wide range of existing and new applications
such as films, various fibers, large and small moldings from boat hulls to instrument parts.
In addition, new manufacturing technologies based on MgCl2 supported catalyst systems
yielded both cost savings and improved products.

1.4.2. Polypropylene Structure

Polypropylene can be produced in different forms regarding its tacticity. The


differences between the different forms of polypropylene produced (i.e., isotactic,
syndiotactic, atactic), assuming equal molecular weight distribution, equal percentage of
branching, etc., are remarkable. In Figure 1.5, the different types of polypropylene
structures are depicted. In perfectly isotactic PP, each monomer unit in the chain is
arranged in a regular head-to-tail assembly without any branching or 1,3 additions (Figure
1.5a.). This is the result of template-type constraints by the heterogeneous stereospecific
catalyst. In practice such perfection is hard to achieve. An occasional error (Figure 1.5b.),
averaging about 0.3–1.5 per hundred chain links, occurs in some, but not equally in all
chains. The multisite nature of solid catalysts is thought to be responsible for this
behaviour. Typical melting point and crystallinity values range from 165-171oC, and, 55-
65%, respectively (Sinn and Kaminsky, 1980).

Metallocene-based single-site catalysts (i.e., SSCs) also produce stereoregularity


errors, but these seem to be distributed more uniformly along the chains, regardless of their
length (see also Figure 1.5e.). This contrasts with multisite catalysts (i.e., MSCs), for
Chapter 1: Introduction 13

which misplacements vary in frequency and in number. The tendency here is for the high
molecular mass fraction to contain fewer faults, while the short chains suffer from
excessive disruptions, which lead to reduced crystallinity.

Until recently, syndiotactic PP (Figure 1.5c.) was not commercially important


because of unacceptable polymerization difficulties at low temperatures. Some metallocene
catalysts overcome this problem to produce a regular sequence of racemic (r) propene
placements yielding a crystalline polymer (50-75% crystalinity) melting at 130°C. Mitsui
Toatsu and Fina have entered into a joint venture to assess costs and customer reaction to
syndiotactic PP. Complete loss of steric regulation generates atactic polymer chains
(Figure 1.5d.), incapable of crystallizing. Hitherto this low molecular mass byproduct had
only a few commercial outlets for adhesives. It should be noted that those stereoisomers of
polypropylene can also be found in other polymers regarding their tacticity.

Polymerization of
polypropylene

Sawhorse projection Modified Fischer Projection

a. Isotactic

b. Chain imperfection

c. Syndiotactic

d. Atactic

e. Effect of catalyst type

Figure 1.5: Polypropylene chain structures.


14 Chapter 1: Introduction

1.4.3. Polypropylene Uses

Polypropylene has a wide range of application in packaging, home products,


consumer goods, automotive products, industrial products, textile yarns, fibres and fabrics.
Polypropylene is readily processed in conventional equipment used for other
thermoplastics. Injection molding, commonly using screw-type reciprocating machines,
accounts for 40% of all applications. Extrusion processes account for the remainder with
domination by fiber and film (Figure 1.6) (Source: PlasticsEurope Market Research
Group).

Film and sheet


22%
Raffia
14%

Fiber
15%
Injection
Molding
Blow molding
37%
1%
Other Pipe and
8% profile
3%

Figure 1.6: Commercial uses of polypropylene.

1.5. Polymer Market

The production of polymers has become a major item in the economy. Polymers
contribute to sustainable development and bring quality of life to citizens. It is sufficient to
say that the polymer industry within the European Union provide employment to 1.6
million people (two thirds of the number of employees in the automotive sector) and
indirectly to many times more in industries which are enabled or depend on plastics for
their products.

The diversity of applications, and the fact that present technology allows the
tailoring of different products for different applications make synthetic polymers very
Chapter 1: Introduction 15

attractive (Chiovetta, 1983). As an example, plastics make many goods in our daily life
more affordable and reduce the wastage of many valuable resources. The consumption of
polymers has increased considerably in the last tricennial and, currently, the requirement
for new products has risen. In Figure 1.7, both world and Europe production of polymers is
presented (Source: PlasticsEurope Market Research Group (PEMRG)). Nowadays, the
overall annual world polymer production may exceed 250 Mt. The data reflect a constant
increase in polymer production in spite of criticism from environmentalists. Nevertheless,
nowadays, polymer waste has become an urgent topic for industry, providing new and
challenging areas of research and development on recycling, reuse, and degradation.
Today, Europe produces approximately 60 million tonnes of plastics, representing 25% of
the global plastics production. Plastic production facilities are well placed across Europe.
Germany is the major producer, accounting for 7.5% of global production followed by
Benelux (4.5%), France (3%), Italy (2%) and the UK and Spain (1.5%) (Figure 1.8)
(Source: PlasticsEurope Market Research Group).

There are around 20 distinct groups of polymers, each with numerous grades
available to help make the best choice for each specific application. There are five high-
volume plastics families; polyethylene (including low density (LDPE), linear low density
(LLDPE) and high density (HDPE)), polypropylene (PP), polyvinylchloride (PVC),
polystyrene (solid PS and expandable EPS) and polyethylene terephthalate (PET).
Together the big five account for over 90% of all polymer demand (see Figure 1.9). It
should also be noted that polyolefins (i.e., HDPE, LDPE, LLDPE and PP) cover 60% of
the total polymer production (source: CMAI report).
16 Chapter 1: Introduction

300
World production 2008: 245
250
Annual Production, Mt Europe production
2002: 200
200

150
1989:100
100
2008: 60
1976: 50
50
1950: 1.5
0
1950 1960 1970 1980 1990 2000 2010
Year

Figure 1.7: Polymer production.

Latin America Japan


4% 5% China
NAFTA
23% 15%

Middle East, Rest of Asia


Africa 16%
8%
CIS
3% Germany
Other EU 7%
Spain
5% Benelux
2%
5% UK
France Italy
3% 2% 2%

Figure 1.8: World polymer production 2008: By country and region.


Chapter 1: Introduction 17

PP
ABS 24%
4%

PVC
19% HDPE
17%

PET
7%

PC LDPE
2% 11%
PS
6% LLDPE
10%

Figure 1.9: Polymer demand 2008: Breakdown by types.

1.6. Polyolefin Market

Polyolefins are the most widely used plastics today due to their low production cost,
reduced environmental impact, and wide range of applications (e.g., packaging, building
and construction, transportation, electrics and electronics, furniture, etc.) (see Figure 1.4
and Figure 1.6). It is believed that the degree of technological and scientific sophistication
in relation to the polyolefin manufacturing has no equal among other synthetic polymer
production processes.

The following main driving forces for commercial expansion of polyolefins are:

1) The further strengthening of the already excellent cost/ property balance and the
good environmental image of polyolefins with respect to all other materials.

2) The continuous polymer quality improvement and property expansion of


polyolefins, leading to new application fields and market sectors.

3) Production cost reduction. In the PE area, this means continuous market growth,
mostly in the areas of the most specialized materials, such as high-density
polyethylene (HDPE) and linear-low-density polyethylene (LLDPE).
18 Chapter 1: Introduction

This progress has been made possible by continuous polymer property developments
perfectly matching the needs of many sophisticated areas (see Figure 1.10) (Galli and
Vecellio, 2004). The history of polyolefin growth speaks for itself. The explosion in the
dynamic development of PE and PP, which is still in progress (Figure 1.11), started mainly
in the 1970’s (Figure 1.12). It appears even more impressive if it is compared with the
growth of all plastic materials. The polyolefin world market share was around 20% of the
total plastics market in the sixties, while reached almost 65% in the year 2000 (see Figure
1.13) (Galli and Vecellio, 2001, CMAI report).

Presently, the total annual world polyolefins capacity exceeds 120 Mt (Source:
CMAI report). Polyethylene and polypropylene cover 60% and 40% of the total
polyolefins production, respectively (Figure 1.14) (source: CMAI report). The annual
world-wide polyolefins market growth in the coming years is foreseen to be 4-6%, making
polyolefin manufacturing a very active research area (Mulhaupt, 2003, Galli and Vecellio,
2004, CMAI report). The major polyolefin manufacturers for 2008 are presented in Figure
1.15 and Figure 1.16 (source: Borealis Group, CMAI, LyondellBasell).

Figure 1.10: PE product map.


Chapter 1: Introduction 19

120
Total polyolefin
production

World Annual Production, Mt


100 PE

PP
80

60

40

20

0
1945 1955 1965 1975 1985 1995 2005
Year

Figure 1.11: PE and PP growth.

140
PP
120
Annual Consumption, Mt

HDPE

100 LLDPE
LDPE
80

60

40

20

0
1970 1980 1990 2000 2005 2010
Year

Figure 1.12: PE and PP consumption.


20 Chapter 1: Introduction

70

60

Polyolefins Volume % 50 PP

of Total Plastics
40

30
PE
20

10

0
1951
1955
1959
1963
1967
1971
1975
1979
1983
1987
1991
1995
1999
2003
Year

Figure 1.13: Polyolefins versus other plastics.

PP
40%

HDPE
27%
LDPE
16%

LLDPE
17%

Figure 1.14: Polyolefin demand 2008: Breakdown by types.


Chapter 1: Introduction 21

Lyondel Bassel

Borealis
Ineos Borealis
Ineos 13% 14%
Dow
Dow 9%
Lyondel Bassel
SABIC 19%
SABIC
9%
Total

Polimeri Europa
Total
Exxon Mobil Other
8%
8%
Repsol Polimeri Europa TVK
6% Exxon Mobil 4%
TVK Repsol
5% 5%
0 1 2 3 4 5 6
Annual Production, Mt

Figure 1.15: European polyolefin producers.

Lyondel Bassel

Exxon Mobil
Other CNPC
Dow 45% 3% Formosa Grp
4%
SINOPEC Borealis
4%
SABIC Total
4%
Ineos
Ineos
5%
Total
SABIC
Borealis 5%

Formosa Grp Lyondel Bassel SINOPEC


9% Exxon Mobil 6%
CNPC 8% Dow
8%
0 2 4 6 8 10 12
Annual Production Mt

Figure 1.16: Global polyolefin producers.


22 Chapter 1: Introduction

1.7. Research Goals

This research project is concerned with the development of a comprehensive


mathematical simulation model of industrial slurry-loop reactor series for the
polymerization of olefins using Ziegler-Natta catalysts. More specifically, this research
project aims at determining the effects of various operating conditions on the dynamic
reactor behaviour (i.e., reaction temperature and pressure, inflow rates of catalyst,
monomers and diluent, etc.), on the molecular rheological and morphological polyolefin
properties (i.e., Mn, Mw, MWD, complex viscosity, particle size distribution, etc.).

Mathematical models for an industrial reactor are a unique tool to the study of
different research areas:

 Reactor design and simulation,


 Process control and optimization,
 Study of reactor dynamic behaviour and grade transitions,
 Calculation of molecular, rheological and morphological properties of the
polymer produced.

The thesis is organized according to the following structure:

Chapter 1 Introduction

Chapter 2 Industrial Polymerization Processes

Chapter 3 Catalytic Polymerization Kinetics

Chapter 4 Slurry-loop Reactor Modelling

Chapter 5 Results and Discussion

Chapter 6 Conclusions and Suggestions for Future Research


Chapter 1: Introduction 23

1.8. References

Brydson J.A., 1999, ‘Plastics Materials, Seventh Edition’, Butterworth-Heinemann,


Oxford.

Chiovettta M.G., 1983, ‘Heat and Mass Transfer During the Polymerization of Alpha-
olefins from the Gas Phase’, Phd Thesis.

Galli P., 1994, ‘The Breakthrough in Catalysis and Processes for Olefin Polymerization:
Innovative Structures and a Strategy in the Materials Area for the Twenty-first
Caentury’, Prog.Polym.Sci., 19, pp.959-974.

Galli P. and Vecellio G., 2001, ‘Technology: Driving Force Behind Innovation and Growth
of Polyolefins’, Prog. Polym. Sci., 26, 8, pp. 1287-1336.

Galli P. and Vecellio G., 2004, ‘Polyolefins: The Most Promising Large-Volume Materials
for the 21st Century’, Journal of Polymer Science, Part A: Polymer Chemistry, Vol.
42, pp. 396-415.

Mattice W.L. and Stehling F.C., 1981, ‘Branch Formation in Low-density Polyethylene’,
Macromolecules, 14, pp. 1479-1484.

Meyer T.and Keurentjes J., 2005, ‘Handbook of Polymer Reaction Engineering’, Wiley-
VCH, Weinheim.

Mirabella F.M., Ford E. A.,1987, ‘Characterization of Linear Low-density Polyethylene:


Cross-fractionation According to Copolymer Composition and Molecular Weight’,
J. Polym. Sci. Polym. Phys. Ed, 25, pp. 777-790.

Mulhaupt R., 2003, ‘Catalytic Polymerization and Post Polymerization Catalysis Fifty
Years After the Discovery of Ziegler’s Catalysts’, Macromol. Chem. Phys., 204, 2,
pp. 289-327.

Odian G., 2004, ‘Principles of Polymerization, Fourth Edition’, Wiley Interscience, New
Jersey.

Redman J., 1991, ‘Polyethylene’, Chem. Eng., 504, pp. 26-33.

Shirayama K., Kita S. and Watabe H., 1972, ‘Effects of Branching on Some Properties of
Ethylene/a-Olefin Copolymers’, Makromol. Chem., 151, pp. 97-120.
24 Chapter 1: Introduction

Sinclair K.B., 1983, ‘Characteristics of Linear LDPE and Description of UCC Gas Phase
Process’, Process Economics Report, SRI International, Menlo Park, CA.

Sinn H. and Kaminsky W., 1980, ‘Ziegler-Natta Catalysis’, Advances in Organometallic


Chemistry, 18, pp. 99-149.

Stevens M.P., 1999, ‘Polymer Chemistry, an Introduction, 3rd Edition’, Oxford University
Press, New York.

Ullmann’s Encyclopedia of Industrial Chemistry, 6th edition, 2002, John Wiley & Sons.

van der Ven S., 1990, ‘Polypropylene and other Polyolefins’, Elsevier, Amsterdam.

Xie T., McAuley K.B., Hsu J.C.C. and Bacon D.W., 1994, ‘Gas Phase Ethylene
Polymerization: Production Processes, Polymer Properties, and Reactor Modeling’,
Ind. Eng. Chem. Res., 33, pp. 449-479.

Borealis Group, ‘Key Figures and Market Position’, URL:


http://www.borealisgroup.com/about-borealis/key-figures-and-market-position/,
Date retrieved: May 10, 2010.

PlasticsEurope Market Research Group (PEMRG), ‘The Compelling Facts About Plastics
2009’, URL: http://www.plasticseurope.org/Documents/Document/20100225141
556- Brochure_UK_FactsFigures_2009_22sept_6_Final-20090930-001-EN-v1.pdf/,
Date retrieved: May 10, 2010.
Chapter 2: Commercial Polymerization Processes 25

Chapter 2

2. Industrial Polymerization Processes

2.1. Introduction

The purpose of this chapter is to present an overview of the processes employed for
the industrial production of polyolefins. The major industrial polymerization processes are
discussed in detail. The basic flowsheet structure of the processes is presented along with
their typical operating conditions.

2.2. Polyolefin Production Processes

The diversity of available processes for olefin polymerization is a result of the


improvements in catalytic systems, increases cost of energy and sophisticated needs of the
market in terms of new product properties and applications (Zacca, 1995). The continuous
product property improvements boosted by endless technology development still remains
in progress. Each different process produces unique combinations of polymer
characteristics.

Current industrial polyolefin processes must take advantage of the catalyst


capabilities and produce the desired range of products with consistent high quality and at a
competitive cost, taking into account environmental considerations. To satisfy the diverse
product quality specifications required by the broad range of polyolefin applications,
polymerization plants are forced to operate under frequent grade transition policies. Thus,
26 Chapter 2: Commercial Polymerization Processes

large scale production commodity polyolefins require simplified processes with efficient
transitions between product grades.

Industrial polyolefin production processes can be broadly classified in high and low
pressure processes, as shown in Figure 2.1 (Ullmann’s Encyclopedia of Industrial
Chemistry). High and low pressure processes differ in both the type of reactor used and in
the physical state of the reaction media. Moreover, in high pressure processes
polymerization takes place through free-radical kinetic mechanism, while, in low pressure
processes, catalytic systems are the usually employed. High pressure processes can be
subdivided to autoclave and tubular technologies. On the other hand, low pressure
technologies include solution, slurry-phase (suspension) and gas-phase, regarding the
reaction medium. In a solution process both the catalyst and the resulting polymer remain
dissolved in a solvent that must subsequently be removed to isolate the polymer. A slurry
process is conducted in an inert diluent in which the catalyst is affixed to an inert support,
and the polymer formed during production remains suspended in the liquid medium, never
dissolving. Slurry processes can be sub-divided into conventional slurry loop processes and
stirred tank reactor processes. Gas-phase polymerization processes are divided in stirred-
bed or the fluidized-bed process. It is noted that supported catalysts are the rule in gas-
phase processes as well. The evolution of industrial processes for polyolefin production is
briefly presented in Figure 2.2.

Figure 2.1: Classification of polyolefin processes.


Chapter 2: Commercial Polymerization Processes 27

Figure 2.2: Evolution of polyolefin improvements.

In general, high-pressure processes can produce LLDPE in addition to the normal


range of LDPEs and ester co-polymers. As well as HDPE some low-pressure plants can
also produce LLDPE and VLDPE, and in many cases these compete for the same market
as LDPE and the ester co-polymers. The essential features of industrial processes for high
density polyethylene, HDPE (polymer density>0.931 g/cm3), linear low density
polyethylene, LLDPE (0.926 g/cm3< polymer density< 0.94 g/cm3) and polypropylene, PP,
may be seen in Table 2.1, Table 2.2 and Table 2.3 (Chandrasekhar et al., 1988).

Each type of process has limitations on the breadth of resins it can produce in terms
of polymer density and melt flow index (see Table 2.4) which directly affect the end-use
properties of the product (Figure 2.4). Production capabilities of all different process types,
regarding PE, are shown schematically in Figure 2.3. Gas-phase processes exhibit the
widest operating window, in terms of polymer density and melt flow index of the polymer
produced. However, gas-phase processes encounter difficulties in producing resins with
very low densities due to polymer stickiness and particle agglomeration. Slurry-phase
processes produce PE, covering a broad range of melt index. However, the range of
densities attainable is limited (i.e., density values greater than 0.937 g/cm3). As density
decreases, resin solubility in the dilute increases. At a density of about 0.930 g/cm3,
28 Chapter 2: Commercial Polymerization Processes

sufficient dissolution occurs to foul the reactor. Hence, slurry processes are not suitable for
LLDPE production. On the other hand, solution processes can produce PE over a broad
range of densities, but only a limited range of molecular weights. As molecular weight
increases, solution viscosity increases also. At some point, the increased viscosity limits
reactor operability and productivity. Thus, solution processes are not suitable for making
high molecular weight PE.

Table 2.1: High density polyethylene processes.

Solution Slurry-phase Gas-phase

Liquid- Multi-zone
Process type Loop reactor Autoclave Fluidized bed
phase autoclave

Temperature, oC 180-270 150-300 85-105 70-90 85-100

Pressure, g/cm2 50-200 1500 7-35 7-10 20

Residence time 2 min 30-40 sec 1.5 hr 2 hr 3-5 hr

Cyclo- Hyper-critical Iso-butane, Iso-butane,


Solvent None
hexane ethylene iso-pentane n-hexane

Propylene, Propylene,
Co-monomer 1-butene-1 1-butene 1-hexene
1-butene 1-butene

Bis(triphenyl
silyl)
chromate/
1-3% chromic
Catalyst Ziegler-Natta Ziegler-Natta Ziegler-Natta silica-alumina
oxide/ silica
dialkyl-
aluminium
alkoxide

Ethylene feed,
Adiabatic Adiabatic
Heat removal Jacket cooling Jacket cooling adiabatic
reactor reactor
reactor

Yield, kg/kg cat. 4.4 5-6 5-50 10 4


Chapter 2: Commercial Polymerization Processes 29

Table 2.2: Linear low density polyethylene processes.

Solution Gas-phase

Process type Adiabatic Liquid-phase Liquid-phase Fluidized bed


autoclave CSTR CSTR

Temperature, oC 150-300 130-180 180-270 85-100

Pressure, kg/cm2 1100 25-30 100-150 20

Residence time,
<1 5-20 2 180-300
min

Conversion per
17-20 95 95 5
pass, %

Diluent Hyper-critical
Hydrocarbon Cyclo-hexane None
ethylene

Co-monomer 1-butene, 1-octene,


1-butene, 4-
4-methyl-1- 1-butene, 1-octene 1-butene, 1-hexene
methyl-1-pentene
pentene

Catalyst Low activity High activity high activity Z-N


Ziegler-Natta
Ziegler-Natta Ziegler-Natta or Cr-Ti on silica

Resin type Pellet Pellet Pellet Granular powder

Density, g/cm3 0.918-0.940 0.917-0.935 0.918-0.96 0.910-0.935


30 Chapter 2: Commercial Polymerization Processes

Table 2.3: Polypropylene processes.

Bulk Slurry-phase Gas-phase

Process type Autoclave or loop Autoclave Fluidized bed Agitated

powder bed

Temperature, oC 50-60 70 50-88 80

Pressure, kg/cm2 25-30 8-12 16-40 30

Residence time,
45 97 - 90
hr

Conversion per
2 2 - 1-2
pass, %

Diluent Liquid propylene n-hexane None None

Catalyst Ziegler-Natta Ziegler-Natta Ziegler-Natta Ziegler-Natta

Yield 300 kg/g Ti 300 kg/g Ti 20 kg/g cat. 20 kg/g cat.

Space-time yield,
94 37.5 - 100
kg/m3/hr

Isotacticity, % 94-98 96-98 95-98 95


Chapter 2: Commercial Polymerization Processes 31

Table 2.4: The technical applicability of processes for polyethylene manufacture.

High High Gas Phase Slurry Phase


Solution
Pressure Pressure
Fluidized Autoclave/
Autoclave
Autoclave Tubular Bed Loop

LDPE + + - - -

EVA co-polymers + + - - -

Acrylate co-
+ + - - -
polymers

HDPE * - + + +

HMW HDPE - - + + -

UHMPE - - * + -

LLDPE + * + * +

VLDPE + * + - +

+ = suitable

* = technically feasible with limitations

- = unsuitable or not possible


32 Chapter 2: Commercial Polymerization Processes

0.97
Homopolymer

0.96 Cascade
Hostalen Slurry
Mitsui Philips
Nissan Solvay
0.95 Equistar USI
Density (kg/lt)

Maruzen
Nisseki
Borstar
0.94

0.93
Solution
Gas
Dowlex
Unipol
Compact
Innovene
0.92 Sclairtech
Spheripol
Evolue
Lupotech G

0.91

0.01 0.1 1.0 10 100

Melt Index (g/10min)

Figure 2.3: Capabilities of different processes in terms of density and melt index.
Property

Property

Figure 2.4: Effect of polymer density and MFI on end-use properties (Foster, 1991).
Chapter 2: Commercial Polymerization Processes 33

2.3. Processes Overview

It is almost impossible to discuss reactors for polyolefin production without


describing the entire polymerization process. Modern processes for polyolefin production
are extremely efficient in producing large quantities of polymer. What has changed from
the basic 1960s process design is that now there is higher efficiency and higher throughput,
and requires less capital investment. Moreover, advances in catalyst technology allowed
the design of more efficient reactors and the elimination of some of the intermediate steps,
such as catalyst deactivation and deashing.

One of the main concerns during the design and control of polymerization reactors is
how to remove the heat of reaction efficiently, since polymerizations are very exothermic
reactions. In polymerization reactors, it is also imperative to be able to produce polymer
that can be easily separated from unreacted monomer, catalyst residues and other
byproducts, while assuring that polymer properties remain on target for a given
polymergrade. These stringent requirements led to the design of several reactor types, each
with their own advantages and disadvantages (Asua, 2007).

However, all processes follow a generic olefin polymerization process scheme as


shown in Figure 2.5 (Siemens, 2010).

 Feedstock materials and additives must be purified and catalyst material must be
prepared. In case of a high pressure process, the gas must be compressed in
several stages.
 Polymerization takes place either in the gas phase (fluidized bed or stirred
reactor), the liquid phase (slurry or solution), or in a high pressure environment.
Polymerization is the heart of the processes. On any one unit, only one of the
three processes is used. More details will be explained on the next pages.
 Polymer particles are then separated from still existing monomers and diluents,
pelletized, dried and dispatched.
 Monomers and diluents are recovered and fed again to the process.
34 Chapter 2: Commercial Polymerization Processes

Feedstock Feed Purification High-pressure


(Options)
Ethylene Suspension Separation Drying
Propylene Gas Compression
Co-monomer
Solvent Slurry-phase Recovery Pelletizing
Nitrogen
Oxygen Catalyst Preparation Gas-phase

Figure 2.5: Generic olefin polymerization process.

The major processes for polyolefin production are summarized below (Source:
Chemical Market Resources). Polymerization processes are discussed in detail. Typical
operating conditions are reported along with their basic flowsheet structure. Extensive
analysis on the commercial production of polyolefin can be found in the work of Choi and
Ray, 1985a, Choi and Ray, 1985b, Beach and Kissin, 1986 and Malpass, 2010.

2.3.1. High Pressure Processes

In high pressure processes, the reactor may take one of two forms: a high pressure
autoclave or a jacketed tube. The overall processes are similar except for the design of the
reactor itself. Fresh ethylene enters the reactor and is mixed with the low pressure recycle.
After further compression the mixture enters the reactor for polymerization. Oxygen or
peroxide may be used as initiators. The reaction pressure ranges from 150-350 MPa and
temperature varies from 150-300oC. Such high pressures call for very specialized
technology and many key features have remained proprietary information. The design of
thick-walled cylinders requires a different type of analysis from that for lower pressure
vessels, and fatigue is a major design consideration for pumps and compressors.

2.3.1.1. Autoclave Process

A typical autoclave design is shown in Figure 2.6. The reaction pressure ranges from
150–200 MPa. Typical overall residence time ranges from 30–60 sec, with corresponding
reactor volumes in larger plants larger than 1 m3. The autoclave functions as an adiabatic
continuous stirred-tank reactor (CSTR), with the heat of reaction being removed by the
fresh ethylene entering the reactor. The conversion of monomer to polymer is thus related
Chapter 2: Commercial Polymerization Processes 35

to the difference in temperature between the feed gas and the final reaction temperature.
Most modern reactors have two or more zones with increasing temperatures. The reaction
temperatures are maintained constant by controlling the speeds of the pumps feeding
initiators into the respective zones. The temperature is typically set equal to 180°C and
290°C for the first zone and the final zone, respectively.

Figure 2.6: Autoclave process flowsheet.

2.3.1.2. Tubular Process

A tubular reactor typically consists of several hundred meters (i.e., 1000-2000 m) of


jacketed high-pressure tubing arranged as a series of straight sections connected by 180°
bends (Figure 2.7). The reaction pressure ranges from 240–310 MPa, while reaction
temperature ranges from 150-300oC. Inner diameters of 25–50 mm are typical of modern
tubular reactors (Malpass, 2010). A ratio of outer to inner diameters of about 2.5 is used to
provide the necessary strength for the high pressures involved. Unlike the autoclave
process, no after-cooler is required for the secondary compressor. The first section of the
tubular reactor functions as a preheater to raise the ethylene to a sufficiently high
temperature for the reaction to start. This temperature depends on the initiator employed,
36 Chapter 2: Commercial Polymerization Processes

ranging from 190°C for oxygen to 140°C for a peroxydicarbonate. The latter part of the
reactor functions as a product cooler similar to that of the autoclave process.

A tubular reactor works in the plug flow regime with heat transfer to the jacket. Plug
flow is achieved by the correct choice of pipe diameter relative to the flow rate so as to
give sufficient turbulence and good axial mixing. Although heat is transferred through the
reactor wall, it is not generally possible to maintain isothermal conditions, and temperature
peaks occur. Injection of initiator at various positions along the tube produces temperature
peaks, increasing the overall conversion. Thus, higher conversions than in the autoclave
reactor can be achieved, but at a higher cost in compression energy. Although conversions
of up to 35% (compared with 20% for the autoclave) have been claimed, the maximum
useful conversion depends on the product quality required, since quality deteriorates
markedly with increasing conversion.

Figure 2.7: Tubular process flowsheet.

2.3.2. Solution Processes

The first commercially available technology for the production of high density,
linear polyethylene grades was the solution process. According to this technology,
polymerization is carried out in a liquid which is a solvent for the resulting polymer, which
is then flashed to remove unreacted monomer. The solvent is recovered and recycled back
in the process. The polymer is then filtered, dried and extruded in a pellet form. Solution
processes have been developed by various companies including Du Pont, Dow, DSM, and
Mitsui for the manufacture of LLDPE or HDPE/ LLDPE on a swing basis. The advantages
Chapter 2: Commercial Polymerization Processes 37

are that they readily handle a wide range of co-monomer types and product densities and,
depending on catalyst type, tend to produce narrow molecular weight distribution products
more readily. Like the high-pressure process (which is also a solution process) they are
unable to handle high-viscosity products. Polymerization occurs in a hydrocarbon solvent
(e.g., cyclohexane) or in a bulk monomer at a temperature above melting point of the
polymer. This type of process is employed for the production of LDPE (e.g., Atochem,
Sumimoto), HDPE and LLDPE (e.g., Dow, Novacor, DSM, etc.), PP (e.g., Eastman
Kodak) and polybutene-1 (e.g., Shell).

Solution processes have some unique advantages over slurry processes: MWD can
be better controlled, and, the process variables are also more easily controlled because the
polymerization occurs in a homogeneous phase. Since polymer properties such as
molecular weight distribution can be controlled by temperature, a very broad spectrum of
product grades can be obtained. The high polymerization temperature (130- 15OoC) also
leads to high reaction rates and high polymer throughputs from the reactor and low
residence times (on the order of minutes). This results in remarkable volume productivities
(up to 7500 kg/h/m3) and small reactors which are specially suited to perform grade
transitions.

Disadvantages of the process include poor monomer conversion ratio (as monomer
conversion is limited by polymer solubility; typical polymer content values in the reactor
vary from 20-35% by weight), high catalyst residuals in the polymer which must be
deashed, and high utilities consumption due to solvent recovery, purification and recycle
steps.

Solution processes are more efficient for production of low molecular weight resins
due to high operating temperatures. However, very high molecular weights cannot be
produced easily at high temperatures as a result of high viscosity in the reactor, and since
the solids content is relatively low compared with the slurry process, greater diluent
recovery may be required (Choi and Ray, 1985a).

2.3.2.1. Dowlex (Propietary: Dow)

Dow developed the Dowlex process to produce linear polyethylenes comprising


octene-1 co-monomer. As shown in a schematic representation in Figure 2.8, the principle
hardware consists of 2 stirred tank reactors in series filled with an isoparaffin solvent (a
38 Chapter 2: Commercial Polymerization Processes

mixture of C8 and C9 isoparaffins). Ethylene, octene-1, and catalyst are introduced into the
reactors, where they remain in solution. The reactors are run at approximately 160°C and
27 bar, and total residence time in the reactors is 30 minutes. When the solution exits the
second reactor it is flashed to remove ethylene monomer. Solvent and residual octene-1 are
removed in a 2-stage flash unit, and the remaining polymer melt is compounded with
additives, extruded, and pelletized.

Figure 2.8: Dowlex process flowsheet.

2.3.2.2. Sclairtech/ AST (Licensor: Nova)

NOVA acquired the rights to Sclairtech technology, when it acquired DuPont


Canada's polyethylene business in 1994. Sclairtech operates at temperatures up to 300°C in
cyclohexane solution. Polymerization is very rapid, requiring less than 2 minutes, so small
reactors can turn out relatively large volumes of polymer. The extrusion and stripping
sections of a Sclairtech plant are also designed to facilitate very quick transitions between
products, on the order of 30 minutes. The Sclairtech process produces polymers from the
LLDPE to the HDPE range using butene-1 or octene-1 co-monomers.
Chapter 2: Commercial Polymerization Processes 39

Advanced Sclairtech technology, or AST, is a significantly updated version, based


on a number of modifications and centered on a set of 2 sequential autoclave reactors, each
fitted with independent monomer, co-monomer, and catalyst feed systems plus high-
intensity mixing technology (Figure 2.9). Discrete polymers can be produced in each
reactor under a different set of tightly controlled conditions, enabling production of novel
polyethylenes with bimodal MWDs having specific end-use properties. Densities ranging
from 0.905-0.967 g/cm3 with melt index ranging from 0.2-150 dg/10 min can be produced
in the AST process.

Figure 2.9: Sclairtech process flowsheet.

2.3.2.3. Compact (Licensor: Stamicarbon-SABIC)

The Compact solution process was originally developed by DSM in the 1970’s to
produce HDPE. It was adapted in the 1980’s to produce linear medium- and low- density
polyethylenes based on butene-1 or octene-1 co-monomers using proprietary Ziegler
catalysts. The Compact solution process is capable of producing polyethylenes across a
density range of 0.900-0.967 g/cm3 and a melt-index range of 0.8-100 using propylene,
butene-1, or octene-1 co-monomers. In the Compact solution process, depicted in Figure
2.10, ethylene and co-monomer, usually octene-1, are dissolved in hexane, cooled, and fed
to a stirred, liquid-filled reactor. The reactor is operated adiabatically, so it is necessary to
precool the reactor feed to control temperature. The enthalpy of polymerization raises the
reactor to its normal operating temperature, 200°C. Molecular weight is controlled by
40 Chapter 2: Commercial Polymerization Processes

hydrogen. Ethylene conversion of at least 95% is achieved during a short reactor residence
time. The hot solution exits the reactor into vessels where volatiles plus the major portion
of the hexane are flashed off, and catalyst residues are deactivated. The polymer
concentrate is fed to a degassing extruder where the residual hexane and octene-1 are
flashed, the appropriate additives are compounded, and the polyethylene is pelletized and
transported to silos for pack out.

Figure 2.10: Compact process flowsheet.

2.3.3. Slurry Processes

The first commercially available technology for the production of high density,
linear polyethylene grades was the solution process. It was soon discovered that a more
efficient way to produce various PE grades was to carry out the polymerization in a slurry-
phase (Hottovy et. al., 2004). In slurry processes, catalyst and polymer particles are
suspended in an inert solvent, typically a light or heavy hydrocarbon. Slurry processes run
in loop reactors with the solvent circulating, stirred tank reactors with a high boiling
solvent or a ‘liquid pool’ in which polymerization takes place in a boiling light solvent. A
variety of catalysts can be used in these processes. Advanced processes combine a loop
reactor with one or two gas-phase reactors, placed in series, where the second stage of the
reaction takes place in the gas-phase reactors. For bimodal polymers, lower molecular
weights are formed in the loop reactor, while high molecular weights are formed in the
gas-phase reactor. It is noted that slurry loop reactor is further discussed and analyzed in
Chapter 4: Slurry-Loop Reactor Modelling.
Chapter 2: Commercial Polymerization Processes 41

2.3.3.1. Phillips Slurry Loop (Licensor: Chevron Phillips)

In 1961, Phillips Petroleum commercialized a slurry ethylene polymerization process


utilizing loop reactors (Figure 2.11). Supported catalyst, ethylene, and co-monomer,
usually hexene-1, are injected into a vertical, loop-shaped reactor filled with isobutane
diluent. Reactor pressure is approximately 40 bar, and reactor temperature is in the range
of 85-100°C. The mixture of catalyst particles, growing polymer particles, co-monomers,
and diluent is pumped around the loop, and polymer particles are harvested by directing a
portion of the slurry to the settling legs, where the polymer particles settle toward the
bottom, then venting the concentrated slurry to a medium-pressure separator to remove
solvent and unreacted monomers then conveying the particles to an extruder where they
are melted, compounded with additives, and pelletized.

Ethylene
To ethylene
Comonomer recovery
make-up Compressor
Fractionator

Loop reactor

To diluent
recovery

Catalyst
Purge
N2 Column
Fresh
Diluent To extruder

Figure 2.11: Phillips slurry loop flowsheet.

2.3.3.2. Solvay Loop (Licensor: Solvey)

Catalyst, ethylene, co-monomer (if desired), and hydrogen are injected into a
vertical, loop-shaped reactor filled with isobutene diluent, as shown in Figure 2.12. Reactor
pressure is approximately 30 bar, and reactor temperature is in the range of 75-80ºC. The
42 Chapter 2: Commercial Polymerization Processes

mixture of catalyst particles, growing polymer particles, co-monomer, hydrogen, and


diluent is pumped around the loop. Polymer particles are harvested via the settling legs by
venting the concentrated slurry to a medium-pressure separating vessel. The solvent and
unreacted monomer are flashed, treated, and recycled back to the loop reactor. The
polymer particles are further stripped, dried, and conveyed to an extruder where they are
melted, compounded with additives, and pelletized.

If bimodal resin is desired, the polymer particles from the first loop reactor can be
transferred to another loop reactor by directing a portion of the slurry to settling legs,
where the polymer particles settle towards the bottom, then re-directed to another loop
reactor. Hydrogen is flashed off prior to introduction in the second loop, and additional
monomer, co-monomer and hydrogen (if desired) are added, and the mixture recirculated
in the second loop, before being harvested.

Figure 2.12: Solvay loop flowsheet.


Chapter 2: Commercial Polymerization Processes 43

2.3.3.3. Hostalen (Licensor: LyondellBasell)

Hostalen process is a low-pressure slurry process technology for the production of


high-performance multimodal HDPE grades. At the heart of the new Hostalen ‘Advanced
Cascade Process’ (Hostalen ACP) technology are three reactors in cascade, enabling the
production of multimodal HDPE resins. Extension of the well-proven and highly regarded
Hostalen bimodal process technology to multimodal capability marks another significant
milestone in LyondellBasell’s rich history of innovation.

A mixture of monomer, co-monomer, hydrogen and hexane (diluent) is continuously


fed to the reactors operating at temperatures ranging from 75-85ºC and pressures of 5-10
bar, as depicted in Figure 2.13. The reactor conditions in each reactor may be varied
independently. The polymer suspension flows from the reactors into a common post
reactor where conversion rate approaches 98%. The suspension is centrifugally separated,
the polymer dried, then it is sent to the extruder for pelletization. The CSTR reactors may
be operated in parallel or series to produce unimodal or bimodal grades, respectively.

Figure 2.13: Hostalen process flowsheet.

2.3.3.4. Nissan (Licensor: Nissan)

The Nissan slurry process, used to produce HDPE, is based on a stirred tank reactor
as shown in Figure 2.14 . Catalyst, ethylene, co-monomer, and hydrogen are fed to the
44 Chapter 2: Commercial Polymerization Processes

reactor which uses hexane diluent. The reactor is operated in a temperature range of 65-
85°C at a pressure of 10-14 bar. Polymer slurry exiting the reactor is sent to a flash vessel
to remove volatiles then fed to a centrifuge where the diluent is removed. The dried
powder is compounded, extruded, and pelletized. The process has been refined through the
joint efforts of Nissan, Maruzen, and Equistar.

Figure 2.14: Nissan process flowsheet.

2.3.3.5. CX Process (Licensor: Mitsui)

The Mitsui process comprises two continuous, stirred-tank reactors of identical size
that are operated in series to produce film resins with bimodal MWDs (Figure 2.15). The
diluent is hexane, and the usual co-monomer is butene-1. Typical conditions are an
operating pressure of 7.8 bar, an operating temperature of 85°C, and a 45-minute residence
time per reactor. Polymer is isolated and pelletized in straightforward fashion in the Mitsui
process.
Chapter 2: Commercial Polymerization Processes 45

Figure 2.15: CX process flowsheet.

2.3.4. Gas-phase Processes

Gas-phase processes are by far the most common process in modern ethylene
production plants. The distinguishing characteristic of gas phase polymerization is that the
system does not involve any liquid phase in the polymerization zone. Polymerization does
occur at the interface between the solid catalyst and the polymer matrix, which is swollen
with monomers during polymerization. The gas phase plays a role in the supply of
monomers, mixing of polymer particles, and removal of reaction heat.

The reactor is operated at a temperature below the melting point of the polymer
particles. For HDPE production, the operating temperature is 90-110oC. An operating
temperature of about 90oC or lower is preferred for production of LLDPE, which contains
about 15 mol% of one or more of the C3 to C6 α-olefins (Levine and Karol, 1977).
Operating pressure ranges from 350-550 psi.

Commercial gas phase processes involve either a fluidized bed reactor, vertical
stirred reactor or a horizontal stirred reactor. In the fluidized-bed process, the monomer
flows through a perforated distribution plate at the reactor bottom and rapid gas circulation
ensures fluidization and heat removal. Unreacted polymer is separated from the polymer
particles at the top of the reactor and recycled. Fluidized-bed plants are able to produce
46 Chapter 2: Commercial Polymerization Processes

either LLDPE or HDPE and are free of constraints from viscosity (solution process) or
solubility (slurry process). A modification uses a second reactor connected in series to
perform copolymerization. The stirred-bed process uses a horizontal or vertical reactor
with compartments, in which the bed of polymer particles is agitated by mixing blades.

Gas phase processes provide an attractive environment for olefin polymerization.


They possess good versatility to produce a broad range of molecular weights and densities
for polyethylenes (see also Figure 2.3). Moreover, they do not involve any liquid phase in
the reaction zone and thus they are not constrained by solubility and viscosity. Other major
advantages of gas phase processes over liquid phase processes, include production of high
co-monomer content polyethylenes, such as high impact strength ethylene propylene
copolymers, ethylene-propylene-diene rubbers. Co-monomer content is limited in the
liquid phase processes because of problems with viscosity effects from dissolved
copolymer and co-monomer (Brockmeier, 1991, Rhee et al., 1991, Burdett. 1992).
Furthermore, gas phase processes can reduce construction costs up to 30% and operating
costs up to 35% over conventional liquid phase processes (Burdett, 1992, Krieger, 1992).

Compared with slurry and solution processes, however, gas phase processes do have
the following disadvantages:

1) Reactor operating temperature is limited to the resin softening point. Thus, the
productivity of catalyst is also limited.

2) The poor heat-transfer efficiency of the gas phase is a disadvantage. Hence,


additional inert heat transfer agent is required to maintain stable reactor
operating conditions at high production rates.

3) There is a possibility of sintering and agglomeration of the polymer particles


due to formation of local hot spots when a high-activity catalyst is used.

4) As polymerization progresses, fine particles will deposit on heat-transfer


surfaces, compressor blades, and sloped walls of the reactor.

5) Sensitivity to variation in operating conditions, such as flow rates of catalyst and


monomer feed and polymer discharge, is a disadvantage.
Chapter 2: Commercial Polymerization Processes 47

2.3.4.1. Unipol (Licensor: Univation)

In 1968 the Union Carbide Corporation (UCC) installed the first commercial gas
phase reactor for the production of HDPE. Later, in 1977, the process was also extended
for the production of LLDPE. Currently, it represents the largest single type of technology
for LLDPE production. In 1983, a joint venture between UCC and Shell Oil combined the
fluidized bed technology with Shell high activity supported Ti catalyst (SHACK catalyst)
to extend the process to PP production (Unipol process). Figure 2.16 depicts a schematic of
the Unipol process. Supported catalyst is fed into a reactor where it is fluidized in a stream
of ethylene, co-monomer, and hydrogen. A gas distributor plate, located near the bottom of
the reactor, supplies a uniform distribution of monomer for fluidization. Polyethylene
forms on the fluidized catalyst particles. Unreacted gases are continually withdrawn from
the reactor, compressed, cooled, and recycled. Average residence times of the particles in
the bed range from 3-5 hours during which time the particles grow to average sizes of
about 500μm in diameter. Temperatures in the range 50-70oC and pressures in the range
20-40 bar are typical. Polymer particles are removed from the reactor through a differential
valve, conveyed to an extruder, compounded with the appropriate additive formulations,
and pelletized.

Unipol II, which comprises not one but two reactors, appeared on the scene in 1992
with the intent to provide Union Carbide (and potential licensees) with the ability to
produce polyethylenes having a bimodal MWD. There are only 2 Unipol II lines in the
world, a Union Carbide plant in Taft, Louisiana, and one belonging to EQUATE
Petrochemical Company K.S.C., a Union Carbide joint venture with Petrochemical
Industries Company and Boubyan Petrochemical Company in Shuaiba, Kuwait. The
production rate of a Unipol reactor is determined by the amount of heat that can be
removed. To increase the heat-removal capacity of the fluidized bed Union Carbide
developed what it called condensed-mode operation. The recycle stream is partially
condensed, and some liquid is re-injected into the system. The enthalpy of vaporization
absorbs extra heat, enabling higher production rates. In the mid 1990’s Exxon developed
and patented what it refers to as Super Condensed Mode Technology that enables even
higher production rates. SCMT is one of the technical cornerstones of Univation
Technologies.
48 Chapter 2: Commercial Polymerization Processes

Cycle
Reactor compressor

Product
chamber

Catalyst Product blow


tank

Ethylene
Comonomer
Degassing
Hydrogen Recycle
stream
To extrusion
and pack out

Figure 2.16: Unipol process flowsheet.

2.3.4.2. BASF Novolen Process

In 1967, BASF was the first company to develop a gas phase process for
polypropylene (Novolen process). During the eighties, the process was extended to
propylene-ethylene random and high-impact co-polymers. Figure 2.17 shows the BASF
continuous process. All the monomers, solvents and nitrogen fed into the process are first
passed through purification columns to remove compounds which act as poisons to the
catalyst. Co-catalyst blend and catalyst slurry are fed from special tanks to the reactor.
Polymerization temperatures are 50-150oC at corresponding pressures of 10-40 bar,
ensuring that the monomer phase is gaseous in the reactor. Low concentrations of
hydrogen are used to control molecular mass over wide ranges. In general, reactor
pressures higher than other gas phase processes are used in order to increase productivity.
The temperature is controlled by removing gaseous propene from the reactor head space,
condensing it with cooling water, and then recirculating it back into the reactor, where its
evaporation provides the required cooling, as well as further aeration of the stirred powder
bed. Each tonne of polymer made requires 6 tonnes of liquid propene to be evaporated as
coolant.
Chapter 2: Commercial Polymerization Processes 49

Typical reactor sizes are 25, 50, or 75 m3, equipped with proprietary helical
agitators, which give excellent agitation. Homopolymerization needs only the primary
reactor, into which the catalyst components are fed. These must be very well dispersed in
the powder bed to avoid build-up.

Powder and associated gas discharge continuously from the primary reactor dip tube
directly into a low-pressure cyclone. Propene carrier gas from this cyclone is recycled to
the reactor after compression, liquifaction, and, sometimes, distillation. The powder then
passes to a purge vessel where a deactivator quenches all residual catalyst activity, and
nitrogen strips out traces of propene from the hot powder. From here powder is conveyed
into silos for stabilization and extrusion into granules. BASF also offers a post-granulation
steam-stripping package to remove any oligomers and oxidized residues from the granules
for demanding applications.

BASF pioneered their gas-phase process with commercial production in 1969. The
products made, Novolen 1300 series, were based on high molecular mass total work up
polymers (i.e., containing atactic PP and catalyst residues) having reduced stereoregularity.
Today, such grades still find niche markets, although they are vulnerable to competition
from random co-polymers. Production is to be phased out shortly. BASF also uses their
process with a cheaper second-generation catalyst TiCl3/AlEt2Cl, which then requires an
additional dry-powder dechlorination stage.
50 Chapter 2: Commercial Polymerization Processes

To recycle

Cyclone

Condenser
Product
Monomer

Hydrogen
T
Catalyst C

1st Reactor To recycle

Polymer
Cyclone

Condenser
Product
Monomer

Hydrogen
T
Catalyst C

2nd Reactor

Figure 2.17: Novolen Vertical bed process flowsheet.

2.3.4.3. Amoco – Chisso Horizontal Stirred-Bed Process

This process was originally developed by Amoco in 1979 for the production of
isotactic polypropylene and then extended to polypropylene co-polymers by Chisso in
1987. The reactor consists essentially of a horizontal cylindrical pressure vessel, stirred by
paddles or an axial shaft (Figure 2.18). Its lower part is divided by weirs into four
compartments. Stirring is used as way of homogenization to avoid hot spots. Monomer, a
recycle gas stream and hydrogen for molecular weight control are fed into the bottom of
each compartment at a low enough velocity to avoid fluidizing the bed. Catalyst
components and a quench stream are sprayed into the bed from inlets in the top of the
reactor. The quench stream, added to provide evaporative cooling, consists of a light
Chapter 2: Commercial Polymerization Processes 51

hydrocarbon such as butane or isopentane and a light co-monomer such as propylene or 1-


butene is fed from the top. The flow of liquid is controlled to achieve a desired temperature
profile. The evaporated monomer leaving the reactor is partially condensed and
recirculated. Fresh make-up monomer is added to this stream. The uncondensed portion is
returned under pressure to the reactor.

Polymer flows from one compartment to another propelled by the axial stirrer till it
reaches the exit section. The powder residence time in a single vessel is equivalent to 3-5
well mixed vessels. Product is discharged from the reactor through a series of sequenced
locks. Vapors from the reactor are removed and separated into a recycle stream. The
catalytic system is based on Ti and Mg modified by a combination of esters and
chlorosulphonic acids. Catalyst poisons such as moisture, oxygen, oxygenated compounds
or sulphur must not be present at levels above one part per million.

For homopolymer production, propylene is the only monomer fed to the system. For
random co-polymer production, ethylene is additionally added. In the case of high impact
polypropylene, a second reactor is required. A gas lock system transfers powder from the
homopolymer reactor to the co-polymer reactor. After the polymerization reactors, the
powder is released periodically to a gas/ powder separation system that includes a purge
column. A countercurrent stream composed of nitrogen and steam deactivates the catalyst.
Product is sent to addition of chemical additives, extrusion and pelletization.

The main advantages of this process are its narrow solids residence time distribution,
resulting in small grade transition times, and the ability of separately controlling reaction
conditions (e.g., temperature, feed compositions etc.) in each of the compartments, leading
to products with a broad range of properties. However, achieving controlled bed uniformity
represents one of the most important operational challenges in this kind of technology.
52 Chapter 2: Commercial Polymerization Processes

Figure 2.18: Horizontal bed process flowsheet.

2.3.4.4. Innovene (Licensor: BP Amoco)

The other major fluidized-bed production technology for polyethylene production is


BP Amoco’s Innovene process. A block diagram of the Innovene process is shown in
Figure 2.19. Though Innovene is very similar to Unipol, there are a few differences. In
Innovene the recycle stream is passed through a cyclone to capture solids in the gas stream
in order to eliminate fouling of the compressor and the heat exchangers and minimize cross
contamination during product transitions. The Innovene design features 2 heat exchangers
in the recycle loop whereas the Unipol design has only one heat exchanger. In 1995 BP
introduced its version of condensed-mode operation that it called ‘High Productivity’
technology. Univation and BP ended up in court over the rights to condensed-mode
operation, and Univation prevailed.
Chapter 2: Commercial Polymerization Processes 53

Ethylene
Comonomer
Hydrogen

Cyclone
Reactor Gas Recycle

Degasser

Catalyst
Degassing
Compressor column

To extrusion
and pack out

Figure 2.19: Innovene process flowsheet.

2.3.4.5. Spherilene (Licensor: Basell)

Spherilene (Figure 2.20) is an advanced swing-gas-phase process for the production


of LLDPE, MDPE and monomodal and bimodal HDPE. Long the major player in
polypropylene production technology, Basell has a polyethylene production process as
well. In 1990 Himont embarked on an effort to design a polyethylene process based on the
extensive experience it had garnered during development of the very popular Spheripol
polypropylene process, and it announced it had perfected the process in 1993, calling it
Spherilene. Taking another cue from its polypropylene technology, Himont designed a
catalyst for the process based on its Reactor Granule Technology that gives the final
product a spherical morphology. A schematic of the Spherilene process is shown in Figure
2.20. Catalyst is fed initially to a prepolymerization reactor, actually a small slurry loop,
where it is precoated with polyethylene, then injected into the first reactor.
Prepolymerization gives advantages in polymer particle size control and control of the
catalyst activity in the fluidized bed reactor (Xie et al., 1994). Spherilene is a multireactor,
fluidized-bed process. Residence time in the process is 2.5 hours. The advantages of
Spherilene over competitive gas-phase processes include:(i) the reactor can be started up
empty, (ii) fouling is minimal, (iii) grade transitions are rapid, and, (vi) steam stripping of
the powder prior to compounding minimized hydrocarbon contaminants.
54 Chapter 2: Commercial Polymerization Processes

Catalyst Prepolymerization Gas phase


reactor Cycle reactor Cycle
compressor compressor

Gas phase
reactor
Product
chamber
Flash
Diluent Product blow
tank
Recycle Recycle
stream stream
Degassing
Ethylene
Comonomer
Hydrogen
To extrusion
and pack out

Figure 2.20: Sperilene process flowsheet.


Chapter 2: Commercial Polymerization Processes 55

2.4. References

Asua J.M., 2007, ‘Polymer Reaction Engineering’, Blackwell Publishing Ltd, UK.

Beach D.L. and Kissin Y.V., 1986, Encylcopedia of Polymer Science and Engineering;
Kroschwitz J.I. Ed., John Wiley & Sons, New York, Vol. 6, p 454.

Brockmeier N.F., 1987, Encyclopedia of Polymer Science and Engineering, Kroschwitz


J.I. Ed., John Wiley & Sons, New York, Vol. 7, p 480.

Burdett I.D., 1992, ‘A Continuing Success: The Unipol Process’, Chemtech, Oct, pp. 616-
623.

Chandrasekhar V., Srinivasan P.R. and Sivaram S., 1988, ‘Recent Developments in
Ziegler-Natta Catalysts for Olefin Polymerization and Their Processes’, Indian
Journal of Technology, 26, pp. 53-82.

Choi K.Y. and Ray W.H., 1985a, ‘Recent Developments in Transition Metal Catalyzed
Olefin Polymerization-A Survey. I. Ethylene Polymerization’, Macromol. Chem.
Phys., C25(1), pp. 1-55.

Choi K.Y. and Ray W.H., 1985b, ‘Recent Developments in Transition Metal Catalyzed
Olefin Polymerization-A Survey. II. Popylene Polymerization’, Macromol. Chem.
Phys., C25( l), pp. 57-97.

Foster G., 1991, Polymer Reaction Engineering Course, McMaster University, Hamilton,
ON.

Krieger J., 1992, ‘AMOCO to Launch New Technology for Ethylene/Propylene


Polymerization’, Chem. Eng. News, March 30, 17.

Levine I.J. and Karol F.J., 1977, ‘Preparation of Low and Medium Density Ethylene
Polymer in Fluid Bed Reactor’, US. Patent No. 4,011,382.

Malpass D.B., 2010, ‘Introduction to Industrial Polyethylene, Properties, Catalysts,


Processes’, John Wiley & Sons.

Rhee S.J.; Baker E.C., Edwards D.N., Lee K.H., Moorhouse J.H., Scarola L.S. and Karol
F.J., 1991, ‘Process for Producing Sticky Polymers’. US. Patent No. 4,994,534.

Ullmann’s Encyclopedia of Industrial Chemistry, 6th edition, 2002, John Wiley & Sons.
56 Chapter 2: Commercial Polymerization Processes

Xie T., McAuley K.B., Hsu J.C.C. and Bacon D.W., 1994, ‘Gas Phase Ethylene
Polymerization: Production Processes, Polymer Properties, and Reactor Modeling’,
Ind. Eng. Chem. Res., 33, pp. 449-479.

Zacca J.J., 1995, ‘Distributed Parameter Modelling of the Polymerization of Olefins in


Chamical Reactors’, PhD thesis, University of Wisconsin-Madison.

LyondellBasel,_‘Licenced_Technologies’,_URL:_http://www.lyondellbasell.com/Technol
ogy/LicensedTechnologies/, Date retrieved: May 10, 2010.

Siemens, ‘Process Analytics in Polypropylene (PP) Plants’, URL:


http://www.sea.siemens.com/us/internet-dms/ia/AppliedAutomation/AppliedAutoma
tion/docs/CS_Process_Analytics_in_PP_Plants.pdf, Date retrieved: May 10, 2010.
Chapter 3: Catalytic Polymerization Kinetics 57

Chapter 3

3. Catalytic Polymerization Kinetics

3.1. Introduction

The purpose of this chapter is to describe the catalytic systems employed for
polymerization. Specifically, the catalytic systems presented are Ziegler-Natta,
metallocenes and chromium-based catalysts. The method of moments, which is a useful
mathematical treatment for developing rate expressions for reactor modelling and
extracting molecular developments, is described. For the definition of a simplified kinetic
scheme capable of describing olefin polymerization in slurry—phase reactors, a literature
review is presented. Moreover, a sensitivity analysis procedure is followed in order to
define the dominant reactions in terms of productivity and molecular weight. Finally, based
on the extracted simplified kinetic scheme, a deconvolution procedure, based on industrial
data, is followed for the calculation of both number of different catalyst active sites and
kinetic rate constants.

3.2. Catalytic Systems

Catalysts are the key to making high performance polymeric materials. Except for
the high temperature and pressure radical initiated ethylene polymerization, all other
processes require a catalyst system for monomer activation. Typical catalytic systems
58 Chapter 3: Catalytic Polymerization Kinetics

employed by commercial olefin polymerization processes are Ziegler-Natta (i.e., Z-N)


catalysts, Phillips-chromium catalysts and metallocenes. Since this research project is
concerned with the simulation of ethylene-1-hexene co-polymerization process via Ziegler-
Natta catalysts, this will be the catalytic system more thoroughly studied.

3.2.1. Polyolefin Catalytic Systems Historical Survey

The high pressure and high temperature processes, developed in the 1930’s, produce
branched, low density polyethylene, LDPE, with unique optical and processing yet less
impressive mechanical properties for more demanding applications. Additionally, harsh
polymerization conditions (few thousands of atmospheres of pressure and temperatures
ranging from 200–300oC) combined with the high hydrocarbon emission nature of the
older reactors still in use are considered as environmental disadvantages.

Alternative processes have been developed for ethylene polymerization and are in
use since the 1950’s. Two classes of co-ordination catalysts, the Phillips catalysts based on
silica supported CrO3 (Chromium catalysts), and Ziegler–Natta catalysts based on MgCl2
supported titanium chloride, have been employed since to produce polyethylene homo- and
co-polymer more efficiently and safely. The medium and high density polyethylene,
M:HDPE, produced with CrO3 based Phillips catalysts have both mechanical and
processing advantages but the optical properties of corresponding articles are poor and
because of relatively low catalyst activity the products contain catalyst residuals. The
linear low density polyethylene (i.e., LLDPE) produced with MgCl2 supported TiCl3 based
catalyst in a fluidized bed gas phase reactor was developed by Union Carbide researchers
in the late 1970’s. It was initially designed to replace the LDPE, it has improved
mechanical properties and its production is less environmentally stressful yet its processing
is more difficult and requires severe extrusion pressure and temperatures. Thus, the so
called LLDPE has become available as an alternative but could not replace the LDPE
resins. With respect to polypropylene until recently, only processes that employed Z-N
type catalysts existed. The catalyst system developed by Natta, initially for producing
crystalline polypropylene, had poor activity and the resulting isotactic propylene product
required several steps of washing and de-ashing with huge quantities of solvent and waste
involved. Since then, at least four new generations of catalysts have been developed each
more active and stereo-selective, eliminating more and more washing and de-ashing steps
Chapter 3: Catalytic Polymerization Kinetics 59

and creating less and less wastes and atactic fractions. However, the production of isotactic
polypropylene with the highly selective Z-N catalysts requires the addition of organic
compounds the so called internal and external electron donors, as modifiers. Additionally,
the MgCl2 supported Z-N catalysts contain substantial amounts of chlorine. The search for
versatile multipurpose clean catalysts finally led to the discovery of the metallocene
catalysts, which after years of gestation have been emerging slowly but pervasively in the
last few years. These catalysts have reached now their full industrial maturity and offer
themselves as a new class of commercially viable and environmentally clean alternative for
the production of a wide range of varieties of polyolefin homo- and co-polymers (Razavi,
2000).

3.2.2. Ziegler-Natta Catalysts

When Z-N catalysis was first discovered, it was fascinating to see ethylene, a
monomer normally difficult to activate, being polymerized under atmospheric pressure at
room temperature in a Weck-glass vessel containing gasoline as the solvent, with small
amounts of TiCl4 and Al2Et3Cl3, the latter components forming the colloidal suspended
catalyst. However, in practice, it was difficult to fabricate the resulting polymer for
practical use. It was necessary to deactivate the catalyst after polymerization, remove the
diluent and the residues of catalyst with HCl and alcohols. Washing of the polymer with
water and drying with steam was then necessary. Other complications involved purification
of the recovered diluent, feedback of the monomer after a repurification step, and, finally,
the necessity of adjusting the molecular weight by oxidative or thermal processes. The cost
of these steps diminished the advantage of low pressure processes (Sinn and Kaminsky,
1980).

The large-scale production of polyolefins has been enhanced by the development of


very high activity MgCl2-supported Ziegler-Natta catalysts as higher activities mean lower
production costs (Dusseault and Hsu, 1993). The Z-N catalyst is a complex formed by
reaction of a transition metal compound (halide, alkoxide, alkyl or aryl derivative) of group
IV-VII transition metals with a metal alkyl halide of group I-III base metals (Boor, 1979)
(Figure 3.1). The former component is usually called catalyst and the latter co-catalyst. In
practice, only a few group I-III metal alkyls are effective. Aluminum alkyls (such as AlEt3,
Al-i-Bu3, AlEt2Cl, AlEtC12 and AlEt2OR) have been overwhelmingly preferred. Also,
60 Chapter 3: Catalytic Polymerization Kinetics

transition metal compounds containing titanium (Ti), vanadium (V), chromium (Cr) and, in
special cases, molybdenum (MO), cobalt (Co), rhodium (Rh) and nickel (Ni) are primarily
used (Huang and Rempel, 1995). The most common catalyst system is titanium
tetrachloride, TiCl4, with triethyl aluminum, Al(C2H5)3, and it is used for polyethylene
production. However, various catalyst-co-catalyst combinations yield behaviour that can
be markedly different. They can vary in activity, stability, ability to polymerize specific
monomers and incorporate co-monomers, ability to control tacticity, and their effects on
the molecular weight distribution of the polymer, among other properties.

Figure 3.1: Ziegler-Natta Catalytic System.

Electron donor compounds, such as amines, ethers and esters, have the potential of
complexing and reacting with the components of the catalyst or the active centers. They
have been used in controlled amounts in many Ziegler-Natta catalytic systems as a third
component to increase catalyst activity and/ or stereoselectivity.

Catalysts of this type operate at low pressures (up to 30 atm) and at temperatures up
to 120oC. The polyethylene produced is characterized by the absence of large amounts of
long- or short-chain branching, which caused variability in density, crystallinity and
melting point in the earlier discovered high-pressure, low-density polyethylene product
formed by radical-included polymerization (Table 3.1).

All Ziegler–Natta commercial polymers have broad molecular mass distributions,


with large polydispertions (i.e., weight average molecular weight, Mw, to number average
molecular weight, Mn,), Mw/Mn= 5–10, for ex-reactor material. Any reduction is achieved
by peroxide-induced chain clipping in an extruder.

There are two existing theories that explain the ability for Ziegler-Natta catalysts to
produce polydisperse polymers. The first is diffusive and the second is kinetic. The former
Chapter 3: Catalytic Polymerization Kinetics 61

emphasizes the effect of monomer concentration on molecular weight development. The


latter is more concerned with the influence of active centers and associated kinetic
parameters. The diffusive theory involves the assumption that mass-transfer limitations
affect the ability of the monomer molecules to reach some of the active sites on the catalyst
particle, due to the presence of pores or other structural effects that hinder transport to the
sites. As a result, different sites polymerize monomer at different rates, leading to a
polydisperse polymer product (Xie et al., 1994).

The kinetic theory involves the assumption that there exist different catalyst site
types, each with its own relative reactivity, due to variations in the local chemical
composition of each site type. An example is the presence of different ligands on the active
transition metal. Experimentation has provided compelling evidence that this second
theory is a more accurate description of the physical reality (Xie et al., 1994, de Carvalho
et al., 1989, Peacock, 2000). These site types tend to produce chains that differ in both
length and co-monomer composition. More specifically, the sites that produce longer
chains tend to incorporate less co-monomer. It should be noted that each different catalyst
site type can produce polymer having polydispersity number equal to 2. However, polymer
exhibiting broad molecular weight distribution can also be produced if more than one
catalyst site types are taking part in the polymerization. In this work, this theory is
followed for the extraction of the kinetic equations, rather than the mass-transfer theory.

Ziegler-Natta catalysts are used in soluble, colloidal, and heterogeneous forms. Most
Z-N catalysts systems are heterogeneous, but some homogeneous systems are also known.
It is not clear that stereoregulation and stereoselectivity are the results of heterogeneity.
Each different catalyst form has its own advantages, depending on the application.
Supported versions are used in gas-phase and slurry processes, while soluble catalysts are
used in solution processes. Although the supported types tend to have higher activity than
the non-supported ones, likely due to an increase in the number of catalyst sites that are
active, the catalytic mechanism is the same (Peacock, 2000).

Several articles have been published that give detailed information on Z-N catalysis.
These include books by Boor, 1979, Kissin, 1985 and van der Ven, 1990 as well as Choi
and Ray, 1985a,b.
62 Chapter 3: Catalytic Polymerization Kinetics

Table 3.1: Comparison of different types of polyethylene.

Polymethylene

Polymer type HP-HD Phillips

High Pressure Ziegler

Density, g/cm3 0.91 0.93 0.95 0.97

Crystallinity, % at 20oC 65 75 85 95

Optical melting point, oC 105 120 125 135

Flexural stress at a given strain,


≈85 ≈100-140 ≈300-370 ≈325-350
kP/cm2

Tensile strength, kP/cm2 140 200-250 200-450 280-430

Notched impact strength, kP.cm/cm2 10 5 4 3

Shore hardness 40-45 45-60 ≈65 ≈68

CH3 groups per 1000 C atoms 20 1-3 1.5


Chapter 3: Catalytic Polymerization Kinetics 63

3.2.2.1. Stereochemical Control by Z-N Catalysts

Ziegler-Natta catalysis provided for the first time stereochemical control of the
polymerization process. By carefully selecting the combination of catalyst and co-catalyst,
one is able to produce polymers with the desired steric structure. The polyethylene
produced in Z-N polymerization is linear, which is characterized by the absence of long or
short chain branching. For a-olefin polymerization, polyolefins of isotactic or syndiotactic
structure can be obtained by using special Ziegler-Natta catalysts. There are even more
choices in steric structures for polydienes; polydienes of the 1,4-cis-, the 1,4-trans-, and the
1,2-structures, as well as the 3,4-structure in the case of substituted dienes, can be
produced with proper Z-N catalysts. Thus, under conditions easy to realize, isotactic and
syndiotactic structures are preferred to statistical atactic structure (see also Figure 1.5).

3.2.2.2. Mechanism of Z-N Catalyzed Polymerization

The mechanism proposed by Cossee (Cossee, 1964) seems to be the best


representation of what is happening at the active center (Figure 3.2). The titanium atom is
in an octahedral coordination environment with one site vacant and an adjacent
coordination site bonded to an alkyl group (polymer chain). The two-step mechanism for
propagation involves π-coordination of an incoming monomer by the titanium atom at its
vacant coordination site, followed by insertion, via a four-center transition state, into the
titanium-alkyl (polymer) bond. It should be noted that for stereospecific polymerization,
the growing polymer chain must migrate back to its original position after each insertion in
order to maintain sterically identical propagation steps. The weakness of this mechanism is
that active Ti3+ has only one outer shell electron available for the formation of π back-
bonding, which is a requirement for metal-olefin n-bond formation (Dusseault and Hsu,
1993).
64 Chapter 3: Catalytic Polymerization Kinetics

Polymer Polymer
H R
X X
C
X M CH2 CHR X M
C
X X
H H
X X

Initial center with one vacancy π-complex with olefin

Polymer
CHR
X X

X M CH2 CHR Polymer X M CH2

X X

X X

The center after olefin insertion Transition state

Figure 3.2: The Cossee mechanism.

3.2.2.3. Evolution of Z-N catalysts

In the sixty years of the first commercialization of propylene polymerization, four


distinct ‘generations’ of catalysts can be identified. Starting from the Natta’s discovery, it
is possible to see how the studies on the catalyst made a continuous and very good
progress, in terms of yields achievable, the fundamental stages of which are reported in
Figure 3.3. But the most important fact in the future of scientific, technological and
commercial development was not just the increase in the catalyst activity itself. It has been
the capability to reduce in parallel the process constraints to zero, maximizing the
technology versatility so as to have a significant expansion of the product property
envelope (Galli and Vecellio, 2001).
Chapter 3: Catalytic Polymerization Kinetics 65

120

1994
100

Yield, Kg PP/g cat


80

1984
60

40

20 1975

1964 1968
1958
0
st nd rd rd th th
1 gen 2 gen 3 gen 3 gen 4 gen 4 gen
super super

Figure 3.3: The relative yield improvement.

First Generation: Conventional Catalysts

Conventional catalysts were the first systems studied by K. Ziegler and G. Natta,
which were able to polymerize olefins in a stereospecific way. These catalysts consist
basically of solid TiCl3 co-crystallized with aluminium halides as a result of the reduction
of TiCl4 by organo-aluminium compounds, under controlled temperature and mixing
conditions. Commonly mentioned in the literature is the formula TiCln.xAlCl3, where
2.5<n<3.0 and 0<x<0.5. Only a small percentage of Ti atoms (1-4%) seem to be
responsible for the formation of active sites. Their activity and selectivity appears to be
related to their position in the structure of TiCl3 crystals. Those crystals aggregate,
generating porous aggregates. During polymerization, these aggregates tend to break into
smaller particles exposing new sites to the reaction medium. According to the ‘replica’
phenomenon, the final polymer particle morphology closely replicates that of the original
catalyst particle.

Second Generation: Modified Catalysts

This type of catalysts is obtained through the modification of conventional catalysts


by means of:
66 Chapter 3: Catalytic Polymerization Kinetics

 Chemical treatment (contact with electron donors, e.g., diisoamilic ether, di-n-
butyl ether, etc.)
 Deposition on inert supports (SiO2, Al2O3. CaO, etc.) (Soga and Ikeda, 1980).

The enhancement in activity, 6-10 times of the conventional catalysts, is attributed to


an increase in the number of active sites as a consequence of the better accessibility of Ti
atoms and the large micro-porosity obtained when chemical treatment is applied.

Third Generation: High Yield and Stereospecificity Catalysts

This class of catalysts appeared during the seventies as a result of research in the
area of commercial production of polyethylene and polypropylene. Its main characteristics
are the use of MgCl2 based supports as well as the introduction of Lewis bases (e.g., esters,
silanes, amines, etc.) as effective catalytic components in the preparation of the support
(‘internal electron-donor’) and during the polymerization (‘external donor’) to obtain high
isotacticity values (i.e., greater than 96%).

As in the case of TiCl3, MgCl2 particles tend to agglomerate, decreasing the support
active surface. However, electron-donor compounds adsorb the cleavage surfaces,
stabilizing the crystallites and partially regulating the stereospecific control, during the
preparation of the support. In the production of polyethylene and polypropylene, catalytic
activity values can reach 1000 tonnes/kg Ti and 1000 tonnes/kg Ti, respectively.

Fourth Generation: Catalysts with Spherical Support

These catalytic systems are prepared by deposition of TiCl4 on spherical particles


MgCl2 in order to produce polymer particles with similar morphological characteristics
(i.e., replica phenomenon). The main improvements with respect to the previous catalytic
systems are:

 Morphologic control of catalyst and polymer particles,


 higher activity, and,
 controlled stereospecificity (e.g., 90-98%) which is independent of polymer
molecular weight.
Chapter 3: Catalytic Polymerization Kinetics 67

The appearance of fourth generation catalysts allowed polymer structure, yield,


selectivity and morphology to be controlled in the reaction stage and important
simplifications in the production process.

3.2.3. Metallocene Catalysts

During the last two decades the development of industrial catalysts based on
metallocenes was a significant innovation process in polyolefin chemistry (Zechlin et al.,
2000). Metallocene compounds such as bis-cyclopentadienyl-titanium (Figure 3.4) or
zirconium complexes have been known to be mildly active for the polymerization of
olefins. In the mid 70’s Kaminski and co-workers (Anderson et al., 1976) discovered that a
significant increase in ethylene polymerization activity could be obtained by contacting
bis-cyclopentadienyl-titanium with methyaluminoxane (MAO). Methylaluminoxane is
synthesized by the addition of water in stoichiometric amounts to trimethylaluminumto
produce a cyclic oligometric structure. Apparently, the role of MAO is to form an ion pair
with an active cationic form of the metallocene. Several articles have been published that
give detailed information on metallocene catalysis (Huang and Rempel, 1995, Brintzinger
et al., 1995, Soares and Hamielec, 1995).

Figure 3.4: Bis-cyclopentadienyl-titanium chloride.

An important aspect in which homogeneous olefin polymerization by metallocene


catalysts differs from heterogeneous Ziegler-Natta catalysis is the narrow molecular weight
distribution of homogeneously produced polymers. While polyolefins obtained with
heterogeneous catalysts have broad molecular weight distributions with large
polydispersities (i.e., Mw/Mn=5-10), homogeneous catalysts (metallocene) polymers
exhibit narrower distributions (Mw/Mn= 2-4) as a consequence of their single-site
68 Chapter 3: Catalytic Polymerization Kinetics

mechanism. This avoids the need for peroxide additives, which produce objectionable odor
and volatile by-products. A polydispersity of 2 is predicted by Schulz-Flory distribution for
polymers arising from identical catalyst centers with fixed rates of chain propagation and
chain termination. A polydispersity of 2 is thus regarded as evidence that only a single
catalyst species contributes to polymer growth in a homogeneous catalyst system
(Brintzinger at al., 1995).

3.2.3.1. Stereoregulation

The breakthrough for propylene polymerization happened with the discovery that
bridged or ansa-zirconocene catalysts could produce stereospecific configurations of
polypropylene (Huang and Rempel, 1995). The addition of bridging substituent made by
zirconocene stereorigid, resulting in different kinds of symmetry. This technique allowed
the production of polypropylene with specific chain configuration structures such as
atactic, isotactic, hemi-isotactic and syndiotactic (see also Figure 1.5).

Atactic Polymers

The metallocenes, such as Cp2ZrC12 and meso-Et(Ind)2TiCl2, have the same Cp (or
Cp’) ligands and an achiral structure. These catalysts are non-stereospecific and furnish
atactic polypropylene and other poly(α-olefins).

Isotactic Polymers

The stereorigid chiral ansa-metallocene catalysts, such as rac-Et(Ind)2ZrC12 and rac-


Me2Si(H41nd)2ZrC12, catalyze the isotactic polymerization of propylene. Before these
catalysts were developed, isotactic polypropylene could only be made using heterogeneous
Ziegler-Natta catalyst systems.

Syndioitactic Polymers

The syndiospecific catalyst, isopropylidene-(cyclopentadienylfluorenyl)zirconium


dichloride, iPr(CpFlu)ZrCl2, produces syndiotactic polypropylene . The catalyst is not
chiral, but has an internal plane of symmetry. The two Cp ligands are of very different
sizes.
Chapter 3: Catalytic Polymerization Kinetics 69

Hemi-isotactic Polypropylene

Introduction of a methyl group into the Cp ring of iPr(CpFlu)ZrCl2 led to an amazing


result: the new catalyst, iPr(3-MeCp-Flu)ZrCl2 activated with MAO produces hemi-
isotactic polypropylene (Hemi- from the Greek, half.). It has become possible for the first
time to generate hemi-isotactic polypropylene through the direct polymerization of
propylene. Before this catalyst was discovered, hemi-isotactic polypropylene, the only
polymer of hemi-tactic structure obtained at that time, was prepared by polymerization of
2-methylpentadiene. Since then, syntheses of hemi-isotactic poly(1-butene) and poly(l-
hexene) with iPr(3-MeCpFlu)ZrC12MA0 catalyst have also been reported. Hemi-tactic
polymers represent a very interesting case in the field of macromolecular stereochemistry.
Their typical feature is the co-existence of order and disorder. More precisely, order and
disorder alternate along the chain in a well-defined way. In a hemi-tactic vinyl polymer,
there are two alternating series of chemically equivalent stereogenic carbons: in the odd
series the arrangement of the substituents follows a well-defined rule, while in the even
series their arrangement is random. A polymer in which the positions of substituents at
odd-numbered carbons are the same is defined as hemi-isotactic, and a polymer in which
substituents are related by a mirror glide plane is defined as hemi-syndiotactic.

3.2.3.2. Mechanism of Metallocene Catalyzed Polymerization

The role of the catalytically active species in the polymerization of a-olefins by


homogeneous Ziegler-Natta catalysis is now well established. The cationic metallocene
alkyls exhibit a strong tendency to coordinate with the weak Lewis base olefin molecules.
A study based on molecular orbital theory indicated that once an olefin coordinated to the
metallocene alkyl, the insertion of the olefin into the alkylmetal bond would proceed
rapidly. The driving force for the insertion is the energy gain on transforming M-R and M-
(C=C) bonds into an M-C and C-R bond (M = Ti, Zr or Hf; R = alkyl; C=C represent
olefin). The insertion leads to a new d0 alkyl complex that can then coordinate another
olefin which would also insert, and so forth, leading eventually to a polymer.

Kaminsky’s Model

Figure 3.5 shows an insertion mechanism model proposed by Kaminsky and Steiger
for ethylene polymerization with Cp2ZrC12/MA0 catalysts. Through a Zr-O-Al bond,
electron density is withdrawn from the Zr atom. If there is no ethylene present, the electron
70 Chapter 3: Catalytic Polymerization Kinetics

deficient Zr atom interacts with P-H through an agostic hydrogen bond, which can lead to a
chain transfer by β-H elimination. If ethylene is present, it can become bonded to Zr
forming a π-complex, which may be followed by ethylene insertion. The model was
explained by the existence of electron deficient compounds in penta-coordinated bimetal
complexes.

Figure 3.5: Mechanism for metallocene polymerization according to Kaminsky’s model.

Corradini’s Model

Figure 3.6 shows a possible mechanism for polymerization as proposed by Corradini


and Guerra. The basic assumptions about this mechanism are:

1) the mechanism is monometallic and the active center is a transition metal-carbon


bond,

2) the mechanism consists of two stages: the coordination of the olefin to the active
site, followed by insertion into the metal-carbon bond through a cis opening.

In Corradini’s model the active species is the Cp2MR+ complex. The cationic
character of the active species has been substantiated by the recent discovery of base-free
cationic metallocene catalysts. According to a quantum mechanical analyses for d0
complexes of the type Cp2ML (L=ligand), the best coordination of L may not be the one
(most symmetrical and presumably sterically most favorable) along the symmetry axis
which relates the two bent Cp rings. Energy minimum situations for a do complex and for
Chapter 3: Catalytic Polymerization Kinetics 71

L=H would correspond to angular deviations Δα≈65o from the symmetry axis. On the other
hand, an additional stabilization of a geometry α ≠ 0ο at the strongly unsaturated metal
center could occur because of an agostic interaction of a hydrogen in the γ-position.
Corradini predicted that α will differ from 0o whenever a monomer or solvent molecule
becomes coordinated to the equatorial belt of the bent metallocene cation, together with the
polymer chain. (Note that a geometry with α ≠ 0ο makes the metal atom chiral in cation
Cp’Cp”MR+ if the two cyclopentadienyl ligands Cp’ and Cp” are different).

As seen in Figure 3.6, at the end of each polymerization step the chain occupies the
coordination site previously occupied by the alkene monomer. Hence, most consecutive
polymerization steps correspond to a model obtained by exchanging the relative positions
of the growing chain and of the monomer. This is extremely relevant for the stereospecific
behaviour of the catalysts.

Figure 3.6: Mechanism for metallocene polymerization accordind to Corradini’s model.


72 Chapter 3: Catalytic Polymerization Kinetics

3.2.4. Cromium Catalysts

In 1953, Hogan and Banks, discovered that chromium oxide supported on silicate
and other carriers, would polymerize ethylene to high polymers (Hogan and Banks, 1958,
McDaniel, 2010). Since then, this catalyst has evolved to become the basis of the Phillips-
slurry loop process. There are three main classes of supported chromium catalysts. The
first one uses organochromium compounds, the second one is based on chromium oxide,
and the third one is a hybrid catalyst.

The first type of catalyst is prepared by calcining a high surface area carrier with
further deposition of an organochromium compound onto it anhydrously. Apparently, the
role of the calcining treatment is to partially dehydroxylate the carrier surface, leaving
mostly isolated single hydroxyls, which can then react with the organochromium
compound through a single oxide bond (Figure 3.7). One or more ligands may remain
attached to the chromium center.

In the second class of catalysts, chromium oxide binds to a fully hydrated surface.
Then, at a temperature around 200oC, a chromate species is formed through an
esterification reaction between CrO3 and two surface hydroxyls. The catalyst does not
become fully activated until more calcinations further dehydrates the surface, which
suggest that neighboring hydroxyls may interfere with polymerization if not removed. The
activity of the catalyst improves with calcining temperature up to the sintering temperature
of the support.

A variety of organo-Cr compounds are suitable for the production of active catalysts.
Examples include π-bonded compounds (e.g., bis(arene)-chromium(0)) and the
chromocene derivatives (e.g., allyl-chromium(II) and –chromium(III), alkyls of Cr(II) and
Cr(IV)). Adequate support for all three types of catalysts include silica, alumina,
aluminophosphates and silica-titanica.
Chapter 3: Catalytic Polymerization Kinetics 73

Figure 3.7: Supported chromium catalysts for ethylene polymerization.

3.2.4.1. Mechanism of Chromium Catalyzed Polymerization

A sequence of steps is involved in the polymerization mechanism (Figure 3.8).


When chromium catalyst contacts ethylene, it is reduced from the oxidation state Cr(VI) to
a lower valent active species, Cr(II), Cr(III) or Cr(V), producing formaldehyde. The second
step consists of an alkylation reaction in which the first polymer chain starts growing.
There are two main steps occurring continuously and simultaneously during
polymerization: (i) propagation, in which ethylene is incorporated into the polymer chain
by insertion into a Cr-alkyl bond, and (ii) β-elimination which ends the living polymer
chain, beginning a new one at the same time. Typical chain life times are of the order of a
second (Zacca, 1995).

Metal alkyls are commonly employed as co-catalysts to promote the activity of


Cr/SiO2 catalysts. These compounds can be strong reducing agents without forming
inhibiting by-products. They may also serve as scavengers to remove redox by-products.
Metal alkyls may alkylate the chromium site as happens in the case of Z-N polymerization
catalysts (Figure 3.9).
74 Chapter 3: Catalytic Polymerization Kinetics

Figure 3.8: Polymerization steps for ethylene polymerization over a chromium oxide
based catalyst.

Figure 3.9: Possible co-catalyst reactions for a chromium oxide based catalyst.
Chapter 3: Catalytic Polymerization Kinetics 75

3.3. Generalized Polymerization Kinetic Scheme

In this section, a comprehensive kinetic scheme for the catalyzed polymerization of


olefins in slurry-phase is presented. Olefin polymerization processes in both slurry and gas
phase reactors require the use of solid-supported catalysts (McKenna and Soares, 2001).
Ziegler-Natta catalysts are used to illustrate the basic principles but chromium-based
catalysts or metallocene catalysts can also be modelled by the same methodology. The set
of equations is generic enough to describe the different types of industrial catalysts. The
kinetics of heterogeneous Ziegler-Natta polymerization has been the subject of numerous
publications (Xie et al., 1994, Dusseault and Hsu, 1993, de Carvalho et al., 1989,
Hutchinson and Ray, 1991, Zacca and Ray, 1993, Lorenzini et al., 1991).

The relative chemical complexity involved in the process of polymerization of


olefins with Ziegler-Natta catalysts led to the postulation of kinetic mechanisms composed
of several reaction steps (de Carvalho et al., 1989, de Carvalho et al., 1990). Following the
ideas of Chen, 1993, Hutchinson, 1990 and Yiannoulakis et al., 2001, a comprehensive
multi-site kinetic scheme is employed. A general kinetic scheme for polymerizations using
Ziegler-Natta catalysts comprises a series of elementary reactions including the following
elementary reactions (McAuley et al., 1990, Xie et al., 1994):

1) activation of potential sites,

2) chain initiation,

3) chain propagation,

4) chain transfer,

5) chain transformation, and,

6) site deactivation.

The model allows up to N s different active site types and up to N m different


monomers, and detailed site kinetics with site dependent live and bulk polymer moments.
The scheme considers chain length, n , and active end group, i , distribution of the co-
polymer species at each type of catalyst site, k . P0k and D kn denote the concentrations of

the activated vacant catalyst sites of type ‘ k ’ and ‘dead’ co-polymer chains of length ‘ n ’
produced at the ‘ k ’ catalyst active site, respectively. The symbol Pnk,i denotes the
76 Chapter 3: Catalytic Polymerization Kinetics

concentration of ‘live’ co-polymer chains of total length ‘ n ’ ending in an ‘ i ’ monomer


unit, produced at the ‘ k ’ catalyst active site.

It should be noted that not all reactions may be applicable in every process. For
example, chain transfer to hydrogen occurs only when hydrogen is added as a chain-
transfer agent. All the different types of reactions are discussed below and are presented in
Table 3.2.

3.3.1. Activation of Potential Sites

Activation is the reaction through which a potential site is converted into a reactive
vacant site. Several mechanisms are proposed in the literature (de Carvalho et al., 1989,
Kissin Y.V., 1985, Nirisen and Rytter, 1986). The present work considers the following
four ways for the catalyst to become activated (Xie et al., 1994):

 by hydrogen,
 by co-catalyst,
 by monomer, and,
 spontaneous.

If the modelled process uses a co-catalyst, and no specific information about the
susceptibility of the catalyst to activate upon reaction with either monomer or hydrogen
exists, a first approach should utilize only the activation by co-catalyst, since this is the
primary activation path. It is always important to minimize the model complexity unless
sufficient information exists to do otherwise.

3.3.2. Chain Initiation

k
During chain initiation, a new polymer chain is started. A vacant site of type k , P0

reacts with a monomer molecule, M i , to produce an occupied reactive site of length 1,

P1,ki . The reaction is assumed to be first order with respect to each component.
Chapter 3: Catalytic Polymerization Kinetics 77

3.3.3. Chain Propagation

Chain propagation is the mechanism step in which the polymer chain grows. A new

monomer molecule, M j , is inserted to the live polymer chain, Pnk,i , that is attached to the

catalyst site, increasing its chain length by one unit, Pnk1, j .

3.3.4. Chain Transfer

Chain transfer is a type of reaction that terminates a live chain, producing dead
polymer and a vacant site. Generally, a chain transfer agent remains attached to the active
site, cutting off the polymer chain. Many chain transfer agents and mechanisms have been
proposed. A generalized kinetic scheme considers chain transfer:

 to hydrogen,
 to co-catalyst,
 to solvent,
 to monomer i, and,
 spontaneous chain transfer.

It should be noted that most of the processes use a transfer agent for molecular-
weight control. Hydrogen is the most common species chosen (Vandenberg, 1962),
reacting with the catalyst site to disengage the polymer chain.

3.3.5. Site Transformation

In order to explain the experimental kinetic behaviour present by different catalysts,


some investigators, have proposed the possibility of the site transformation reaction, in
which, any site type can be converted to another one by means of specific reactions. A
generalized kinetic scheme considers site transformation:

 to hydrogen,
 to co-catalyst,
 to solvent,
 to poison, and,
78 Chapter 3: Catalytic Polymerization Kinetics

 spontaneous.

This produces an empty live site, unlike the previous chain-transfer reactions that
produce an initiated site, and a dead polymer chain.

3.3.6. Catalyst Deactivation

Site deactivation is the reaction step generally accepted as the explanation for the
loss in activity experienced in some catalytic systems during the polymerization (Floyd,
1986, Nirisen and Rytter, 1986, Ray, 1972). Both occupied and vacant sites are assumed to
deactivate according to deactivation:

 by hydrogen,
 by co-catalyst,
 by by-product,
 by poison,
 by monomer j, and,
 spontaneous deactivation.
Chapter 3: Catalytic Polymerization Kinetics 79

Table 3.2: Generalized polymerization kinetic scheme.

K kl
Site activation to solvent: Pnk,i  S 
tS
 P0l  Dnk
k
By hydrogen: S p  H 2 K
  P0k K kl
P0k  S   P0l
aH
tS

k
by co-catalyst: S p  A K
 P0k  B to poison: K kl
Pnk,i  X   P0l  Dnk
aA
tX

Kk K kl
by monomer i: S p  M i 
aM 1
P0k  M i P0k  X 
tX
 P0l
k
S p K
  P0k K kl
spontaneous: spontaneous: Pnk,i   P0l  Dnk
aSP
tSP

Chain initiation K kl
P0k 
tSP
 P0l

by monomer i: Kk Site deactivation


P0k  M i 
0i
 P1k,i
k
Chain propagation by hydrogen: Pnk,i  H 2 K

dH
CDk  Dnk
Kk k
by monomer 1: Pnk,1  M 1 

p 11
 Pnk1,1 P0k  H 2 K
dA
CDk
Kk k
Pnk, 2  M 1 

p 21
 Pnk1,1 by co-catalyst: Pnk,i  A K
dA
CDk  Dnk
Kk k
by monomer 2: Pnk,1  M 2 

p 12
 Pnk1, 2 P0k  A K
dA
C Dk
Kk k
Pnk, 2  M 2 

p 22
 Pnk1, 2 by by-product: Pnk,i  B K
dB
 CDk  Dnk
k
Chain transfer P0k  B K
dB
 CDk
k
to hydrogen: Pnk,i  H 2 
K trH
 P0k  Dnk by poison: Kk
Pnk,i  X 
dX k
 CD  Dnk
k
to co-catalyst: Pnk,i  A K
 P0k  Dnk Kk
P0k  X  k
trA
dX
 CD
Kk
to solvent: Kk by monomer j: Pnk,i  M j 
 C Dk  Dnk
Pnk,i  S   P0k  Dnk
dMij
trS

Kk Kk
to monomer i: Pnk,i  M i 
trMi
P1k,i  Dnk P0k  M j 

dMj
 C Dk
k k
spontaneous: KtrSpi
Pnk,i   P0k  Dnk
spontaneous: Pnk,i K

dSP
 CDk  Dnk
k
Site transformation P0k K

dSP
 CDk

to hydrogen: K kl
Pnk,i  H 2 
tH
 P0l  Dnk

K kl
P0k  H 2 
tH
 P0l

K kl
to co-catalyst: Pnk,i  A 
tA
 P0l  Dnk

K kl
P0k  A 
tA
 P0l
80 Chapter 3: Catalytic Polymerization Kinetics

3.4. Method of Moments

The moments are averages of the concentrations of polymer molecules that are
weighted by their chain lengths. The method of moments, which is a useful mathematical
treatment for developing rate expressions for reactor modelling is employed for the
calculation of molecular properties. This treatment is a statistical technique that enables us
to track various polymer chain properties without the need to include the very large
number of equations and unknowns required to account for chains of every possible length
(Hatzantonis et al., 2000, Khare, 2003, Dompazis et al., 2008).

The following ‘live’, k , ‘dead’, k , and ‘bulk’,  k , moments of the corresponding

number chain length distributions (NCLDs) are defined.

Live moments:
Nm Nm 
λνk  k
 λν,i    nν  Pn,i
k 
 ;   0, 1, 2, ... (3.1)
i 1

i 1 n 1

Dead moments:

k   n  Dnk  ;   0, 1, 2, ... (3.2)
n 2
 

Bulk moments:
 νk  kν  k ;   0, 1, 2, ... (3.3)

λν,ik and k denote the ‘ ’ moment of ‘live’ co-polymer chains ending in an ‘ i ’

(  1, 2, ..., M m ) monomer unit and the ‘ ’ moment of ‘dead’ co-polymer chains, produced

at the ‘ k ’ catalyst site, respectively.  νk is the respective ‘ ’ ‘bulk’ moment. The three
leading moments, namely, the zeroth, first, and second, are sufficient for describing the
molecular weight distribution of most commercial polymers.

According to the kinetic mechanism, one can derive the net production/ consumption
rates of the various molecular species for a multi-site Z-N polymerization.
Chapter 3: Catalytic Polymerization Kinetics 81

Polymer Moments

A live polymer chain balance, applied over the kinetic scheme, gives the following
equation:

d  Pnk,i  N
RP k       k   P k   s  kl   P k    k   P k 
d  n,i   st  n,i  ti  n,i 
n ,i dt l 1 k

 Nm 
k k k  k k 
   (n)  ktMij  0, j   ( n )  k0i  0  0   M i 
 P P (3.4)
  ; for n=1, ∞
 j 1 

Nm Nm
 1   (n)   k kpij   Pnk1, j    M i    k kpij   Pnk,i    M j 
j 1 j 1

1 for n  1
where  (n)  0 for n  1

A dead polymer chain balance, applied over the kinetic scheme, gives:

 Ns Nm  Nm
RD k     stk   tik   dk     Pnk,i  (3.5)
n  l 1 k i 1
 i 1  
 

Polymer moments rates can de derived, multiplying all live and dead polymer chains

rates to nν and summing up for n  2,  . The final moments’ rates can be calculated:

Live moments:
 Ns Nm  Nm  v v 
R k  vk,i   dk    stkl  tik   k kpij   M i     k kpij   M i        qk,i 
v ,i  l 1 k j 1  j 1  q 0  q  
   
(3.6)
 Nm 
k k k  k 
   ktMij  0, i  k0i  0    M i 
 P
 j 1 
 

v v!
where  q   q !(v  q)!
 
82 Chapter 3: Catalytic Polymerization Kinetics

Bulk moments:
Nm Nm Nm Nm  v v 
k l k k k
R k    ktMij  0, i   M i     k pij   M i        q,i  v,i 
v
i 1 j 1 i 1 j 1  q 0  q  
 
(3.7)
Nm
  k0ki   P0k    M i 
i 1
 

Liquid Phase Components

Monomer i:

s N mN sN m s N N
RM i    k0ki   M i    P0k     k Pji k k
  M i   0, j    ktrMi k k
  M i   0, j
k 1
  j 1 k 1 j 1 k 1
(3.8)
Nm Ns Ns
 k
 kdMji k k  k
  M i   0, j   kdMi   M i    P0 
j 1 k 1 k 1

Solvent:

Ns  Nm  Ns Ns Nm Ns Ns
k
RS     ktrS   S    0,k 
i
kl
   ( ktS   S    0,k
i )    ktS kl
  S    P0k  (3.9)

k 1  i 1
  
 k 1 l 1 i 1 k 1 l 1

Hydrogen:

sN Ns  Mm  Ns Ns  Mm 
k
RH 2    kaH   H 2   S kp  k
  ktrH k 
  H 2    0, kl k
i      ktH   H 2    0,i 
k 1 k 1  i 1  k 1 l 1  i1 
(3.10)
Ns Ns Ns Mm Ns
kl
  ktH   H 2   P0k   ( kdH
k k
  H 2    0, k
i )   kdH   H 2   P0
k
k 1 l 1 k 1 i 1 k 1

Co-catalyst by-product:

Ns Ns  Mm  Ns
RB  k
 kaA   A   S kp     kdB
k k 
  B    0, k  k
i    kdB   B    P0  (3.11)
k 1
  k 1  i 1  k 1
Chapter 3: Catalytic Polymerization Kinetics 83

Poison:

Ns Ns  Mm  Ns Ns Ns  Mm 
kl k  kl k k k
RX      ktX   X    0, i     ktX   X   P0    kdX   X    0,i 
k 1 l 1  i 1 k 1 l 1
 k 1 i 1  
(3.12)
Ns
k
  kdX   X    P0k 
k 1
 

Co-catalyst:

Ns Ns  Mm  Ns Ns  Mm 
k
R A    kaA   A   S kp     ktrA
k
  A   0,k 
i     ktA kl k 
  A    0, i
k 1
  k 1  i 1
 k 1 l 1  
   i1 
(3.13)
Ns Ns Ns Mm Ns
 kl
 ktA   A   P0k    (kdA
k k
  A   0, k  k
i )   kdA   A   P0 
k 1 l 1
  k 1 i 1 k 1

Polymer Related Quantities

Potential active sites:

RSp k   ak   Sp k  (3.14)
 

Vacant catalyst sites:

Nm
RP k   ak   S kp   k01
k
  M1    P0k   k02
k
  M 2    P0k   ktrH
k
 [ H 2 ]   0,k
i
0       i 1

Nm Nm Nm
k k k k k k
 ktrA   A   0, i  ktrS   S    0,i   ktrSPi  0, i
i 1 i 1 i 1 (3.15)
Ns Nm Ns Ns
  tlk   0,l i   tlk  P0l   tkl  P0k   dk   P0k 
l 1 i 1 l 1 k 1

k
kdM k k  k
1   M1   P0  kdM 2   M 2    P0 
84 Chapter 3: Catalytic Polymerization Kinetics

where:

Nm
 ak  kaH
k k
  H 2   kaA k
  A   kaMi k
  M1   kaSP (3.16)
i 1

tkl  ktH
kl kl
  H 2   ktA kl
  A  ktS kl
  S   ktX kl
  X   ktS P (3.17)

Nm
k k k k k k
tri  ktrH   H 2   ktrA   A  ktrS   S    ktrMi   M i   ktrSPi (3.18)
i 1

Nm
 dk  k dH
k k
  H 2   k dA k
  A   ktB k
  B   k dX k
  X    k dMi k
  M i   kdSP (3.19)
i 1

It is noted that, according to the work of Hutchinson, 1990, the calculation of


reaction rates should involve the true (local species concentration) at the active catalyst
sites. When polymer particles are present, as in the case of slurry or gas phase
polymerizations, solid-phase species concentrations are calculated by the sorption
correlations corresponding to the thermodynamic equilibrium between fluid and
amorphous polymer phases. In order to account for these solubility effects, the Sanchez-
Lacombe Equation of State is employed. Thermodynamic calculations are further
discussed in 4.4.9. Thermodynamic Considerations.

3.5. Calculation of Molecular Properties

The average molecular properties of interest (e.g., number- and weight- average
molecular weights) and the polydispersity index can be calculated in terms of the overall
‘bulk’ moments,   , of ‘live’ and ‘dead’ polymer chains produced over the N s catalyst
active sites.
Chapter 3: Catalytic Polymerization Kinetics 85

Co-polymer Composition

The molar polymer composition can be calculated as a ration of the polymer first
moments.

Average instant co-polymer composition:


Nm
  R /  RMi
i
k k
Mi
k
(3.20)
i 1

Average cummulative co-polymer composition:

 ik  1,ik / 1k (3.21)

Cumulative co-polymer composition:


Ns Ns
 i   1,ik /  1k (3.22)
k 1 k 1

Degree of Polymerization

Catalyst site type, k and overall number average degree of polymerization:

DPnk  1k 0k ; DPn  1 0 (3.23)

Catalyst site type, k and overall weight average degree of polymerization:

DPwk   2k 1k ; DPw   2 1 (3.24)

Average Molecular Weight

Number-average molecular weight:

 Ns Ns

M n    1k  0
k
 MW  1 0  MW  DPn MW (3.25)
 k 1 k 1 
86 Chapter 3: Catalytic Polymerization Kinetics

Weight-average molecular weight:

 Ns Ns

M w     2k  k
1  MW   2 1  MW  DPw MW (3.26)
 k 1 k 1 

where MW is the average molecular weight of the repeating structural unit in the co-
polymer chains, that is:
Nm
MW    i MWi (3.27)
i 1

where MWi is the molecular weight of monomer i , M i .

Polydispersity

Polydispersity is defined as the ratio between the weight- to number average


molecular weight.

Mw
PD  (3.28)
Mn

It should be noted that the calculation of molecular properties described by Eq. 3.20-
Eq. 3.28 refer to instantaneous values. At any time, the cumulative expression of the
calculated property can also be determined by employing a mass balance in terms of the
fresh polymer produced, described by the instantaneous property value, and the total
polymer mass in the reactor described by the cumulative property value.

3.6. Molecular Weight Distribution Reconstruction

The molecular weight distribution of a polymer is a record of the kinetic history of


the reactions which occurred during its formation (Clay and Gilbert, 1995, Canu and Ray,
1991). Therefore, it is an important source of information about the kinetic processes that
have taken place in polymerization systems. When one speaks of the molecular weight of a
Chapter 3: Catalytic Polymerization Kinetics 87

polymer, one means something quite different from that which applies to small-sized
compounds. Polymers differ from the small-sized compounds in that they are polydisperse
or heterogeneous in molecular weight. The control of molecular weight and molecular
distribution is often used to improve certain desired mechanical and physical properties of
a polymer product (Kiparissides, 1996).

The reconstruction of the weight chain length distribution (i.e., WCLD) of the
polyolefin, produced at the ‘ k ’ catalyst site, can be achieved by using the two-parameter
Schultz-Flory distribution:

W k
x    
y k xy k
zk
e  xy
k

; k  1,2,... N s (3.29)
e ln  z 1
k

where W k  x  denotes the mass fraction of polymer chains with a degree of polymerization
x , produced at the ‘ k ’ catalyst site. Note that the Schultz-Flory parameters can be
expressed in terms of the ‘bulk’ moments (  0k , 1k ,  k2 ):

1k  0k zk 1
zk  ; yk  (3.30)
 2k 1k  1k  0k ξ 2k ξ1k

Thus, for a multi-site catalyst, the overall MWD will be given by the weighted sum
of all polymer fractions produced over the different catalyst active sites (Figure 3.10).

 
Ns
Wt  x   W k  x  1k 1 (3.31)
k 1

Ns
where 1 (   1k ) is the total polymer mass produced over the N s catalyst active sites.
k 1
88 Chapter 3: Catalytic Polymerization Kinetics

Overall
Catalyst site type 1
Catalyst site type 2
Catalyst site type 3
Catalyst site type 4
dW/dlog(MW)

log(MW)

Figure 3.10: Overall MWD given by the weighted sum of polymer fractions produced
over the different catalyst active sites.

3.7. Slurry Loop Reactor Kinetic Scheme

This research project is concerned with the simulation of ethylene-1-hexene co-


polymerization process (via Ziegler-Natta catalysts) in slurry-loop reactor series. The more
general a kinetic scheme is the more kinetic parameters demands for the description of
molecular developments. Thus, a simplified kinetic scheme, able to describe the system
studied, is preferably chosen.

It should be mentioned that the number of kinetic model equations solved at any
time is a function of:

1) the number of elementary kinetic reactions, and,

2) the number of the employed catalyst active sites.

In order to reduce the number of equations solved, a simplified kinetic scheme


should be defined. In the following paragraphs, the procedure followed in order to reduce
kinetic equations is explained.
Chapter 3: Catalytic Polymerization Kinetics 89

3.7.1. Kinetic Modelling Literature Review

Many investigators have worked on the slurry-phase ethylene polymerization. A


literature review on their work would make possible the elimination of some of the
elementary reactions included in the generalized kinetic scheme. Zacca and Ray, 1993,
suggest a kinetic scheme comprising of site activation by co-catalyst, propagation, chain
transfer to hydrogen, to monomers, site transformation by co-catalyst and spontaneous
deactivation. Wells et al., 2001 in their work also propose a rather generalized kinetic
scheme comprising of initiation by hexene and ethylene, activation by co-catalyst,
propagation, transfer to co-catalyst, to ethylene, to hexene, to hydrogen, spontaneous
transfer and site deactivation. On the other hand, Choi et al., 1997 and Fontes and Mendes,
2005 in their work, use the kinetic scheme proposed by Zacca and Ray, 1993, without the
site transformation reaction. Bhagwat et al., 1994 and Wu et al., 2005 simplified even more
the kinetic scheme eliminating site transfer reactions, while Ha et al., 2000, and Neto et al.,
2005 eliminated chain transfer reactions. Table 3.3 and Table 3.4. summarize the
information gathered from literature review.

According to the work cited on slurry-phase ethylene polymerization, site


transformation reactions do not seem to play an important role, while chain transfer and
deactivation reactions are employed by the majority of the workers on the field. More
specifically, chain transfer to hydrogen, to monomer and spontaneous chain transfer are the
transfer reactions more often selected, while spontaneous deactivation is also usually
employed.
90 Chapter 3: Catalytic Polymerization Kinetics

Table 3.3: Numerical values for kinetic rate constants.

1. Zacca and Ray, 1993

Kinetic rate constant (l/mol/s)

Activation by co-catalyst 7.04x102 (Ea = 12 kcal/mol)

Propagation 4.84x108 (Ep = 12 kcal/mol)

Transfer spontaneous 3.85x102

Transfer to hydrogen 4.4x106 (Et = 12 kcal/mol)

Transfer to ethylene 6.16x103

Site transformation to co-catalyst 7.04x102

Deactivation 7.92x103 l/s (Ed = 12 kcal/mol)

2. Bhagwat et al., 1994

Kinetic rate constant (m3/mol/s)

Initiation 1x107

Propagation 1x107

Deactivation 1 s-1

3. Choi et al., 1997

Kinetic rate constant (l/mol/min)

Initiation 0.263 (Ei = 12 kcal/mol)

Propagation 6.8x102 (Ep = 8 kcal/mol)

Transfer to co-catalyst 0.994 (Etr = 18 kcal/mol)

Transfer to monomer 0.019

Transfer to hydrogen 0.036

Deactivation 2x10-2 l/min


Chapter 3: Catalytic Polymerization Kinetics 91

4. Ha et al., 2000

Kinetic rate constant (m3/mol/s)

Activation by aluminium alkyl

Initiation 10

Propagation 1

Transfer to hydrogen

5. Wells et al., 2001

Kinetic rate constant (l/mol/s)

Activation by co-catalyst 7x107 (Ea = 9 kcal/mol)

Initiation by ethylene 2.7x106 (Ei = 9 kcal/mol)

Initiation by hexene 3.8x105 (Ei = 9 kcal/mol)

Propagation ethylene-ethylene 7x108 (Ep = 9 kcal/mol)

Propagation ethylene-hexene 6x108 (Ep = 9 kcal/mol)

Propagation hexene- ethylene 3.5x106 (Ep = 9 kcal/mol)

Propagation ethylene-hexene 2.4x106 (Ep = 9 kcal/mol)

Transfer to co-catalyst (ethylene end group) 6x105-6x106 (Et = 14 kcal/mol)

Transfer to co-catalyst (hexene end group) 6x105-1.3x106 (Et = 14 kcal/mol)

Transfer to ethylene (ethylene end group) 6x105 (Et = 14 kcal/mol)

Transfer to hexene (ethylene end group) 1.6x105-3.2x106 (Et = 14 kcal/mol)

Transfer to ethylene (hexene end group) 6x105 (Et = 14 kcal/mol)

Transfer to hexene (hexene end group) 1x107-1.6x108 (Et = 14 kcal/mol)

Transfer to hydrogen (ethylene end group) 3x107-3x108 lt0.5/mol0.5s(Et = 14 kcal/mol)

Transfer to hydrogen (hexene end group) 3x107-3.2x108 (Et = 14 kcal/mol)


92 Chapter 3: Catalytic Polymerization Kinetics

Transfer spontaneous (ethylene end group) 2.8x102 (Et = 14 kcal/mol)

Transfer spontaneous (hexene end group) 2.8x102 (Et = 14 kcal/mol)

Deactivation 5x103 (Ed = 12 kcal/mol)

6. Fontes and Mendes, 2005

Kinetic rate constant (l/mol/min)

Propagation 1.01x108

Transfer to co-catalyst 3.876x107

Transfer to ethylene 3.392x105-3.392x106

Transfer to hydrogen 1.421x107-1.421x108

Transfer to solvent 1.615x105

Deactivation 3x10-1

7. Wu et al., 2005

Kinetic rate constant

Initiation ~10-2 s-1 (Ei = 69.9 kJ/mol)

Propagation

Deactivation ~10-4 s-1 (Ed = 100.3 kJ/mol)

8. Neto et al., 2005

Kinetic rate constant (l/mol/min)

Initiation 8x104

Propagation 8x104 (Ep = 41 kJ/mol)

Transfer to ethylene 9.6-80 (tuning parameter) (Et = 83 kJ/mol)


Chapter 3: Catalytic Polymerization Kinetics 93

Table 3.4: Proposed kinetic mechanisms adopted by previous workers on the field.

transformation
Transfer to co-
By co-catalyst

(spontaneous)

Deactivation
Propagation

Transfer to

Transfer to

Transfer to
Reference*

Activation

monomer

hydrogen
Initiation

Transfer
catalyst

solvent

Site
1       

2    

3      

4    

5       

6      

7   

8   

*Reference corresponds to Table 3.3 References.

3.7.2. Kinetic Rate Constants Sensitivity Analysis

3.7.2.1. Definition

A possible definition of sensitivity analysis is the following: The study of how


uncertainty in the output of a model (numerical or otherwise) can be apportioned to
different sources of uncertainty in the model input (Saltelli et al., 2004, Saltelli et al.,
2008). Sensitivity analysis is aimed to determine what factor most needs better
determination, and to identify the weak links of the assessment chain (those that propagate
most variance in the output).

A sensitivity analysis procedure can be applied:

 To simplify models.
94 Chapter 3: Catalytic Polymerization Kinetics

 To investigate the robustness of the model predictions.


 To play what-if analysis exploring the impact of varying input assumptions and
scenarios.
 As an element of quality assurance (unexpected factors sensitivities may be
associated to coding errors or misspecifications).

A general schematic description of the steps to be followed to perform sensitivity


analysis on a model, independently of the method being used, is given below:

 Establish the goal of your analysis.


 Decide which input factors you want to include in your analysis.
 Choose a distribution function for each of the input factors. This can be taken
from the literature, derived from data by fitting an empirical distribution
function or based on an expert’s opinions.
 Choose a sensitivity analysis method on the basis of the following.
 Analyse the model outputs and draw your conclusions.

3.7.2.2. Sensitivity Analysis Procedure

A sensitivity analysis has been conducted concerning the estimation of the kinetic
parameters, so that the minimum number of elementary kinetic reactions, able to simulate
the actual process kinetics, could be defined. The effect of kinetic rate constants was
studied in terms of variability produced regarding molecular weight distribution and
productivity. After developing a single-site kinetic model, the effect of kinetic rate
constants in the output variables of interest was calculated. The kinetic scheme includes
site activation by catalyst, chain initiation and propagation reactions, chain transfer to
hydrogen and spontaneous chain transfer and spontaneous deactivation. The sensitivity
coefficient is defined as the ratio of the output variable change to the forced input variable
change. Figure 3.11-Figure 3.20 depict the effect of kinetic rate constants on the molecular
weight distribution. The results obtained from the sensitivity analysis are summarized in
Figure 3.21 and Figure 3.22. As can be seen, the elementary reactions that affect the
polymer molecular properties (e.g., average molecular weight) of the produced polymer are
site activation (Ka), chain propagation (Kp11) and chain transfer by hydrogen (KtrH) (Figure
3.21). On the other hand, the elementary reactions that affect the polymerization rate are
site activation (Ka) and chain propagation (Kp11) (Figure 3.22).
Chapter 3: Catalytic Polymerization Kinetics 95

1.5
1
Ka = 2 10 lt/mol/s
2
Ka = 2 10 lt/mol/s
2
Ka = 5 10 lt/mol/s
3
1.0 Ka = 1 10 lt/mol/s
dW/dlog(MW)

3
Ka = 2 10 lt/mol/s

0.5

Ka

0.0
2 3 4 5 6 7
log(MW)

Figure 3.11: Effect of Ka on MWD.

1.5
6
K01 = 5 10 lt/mol/s
8
K01 = 5 10 lt/mol/s
10
K01 = 5 10 lt/mol/s

1.0
dW/dlog(MW)

0.5
K01

0.0
2 3 4 5 6 7
log(MW)

Figure 3.12: Effect of K01 on MWD.


96 Chapter 3: Catalytic Polymerization Kinetics

1.5
6
K02 = 4 10 lt/mol/s
8
K02 = 4 10 lt/mol/s
10
K02 = 4 10 lt/mol/s

1.0 There is no effect of K02 on the MWD


dW/dlog(MW)

0.5

0.0
2 3 4 5 6 7
log(MW)

Figure 3.13: Effect of K02 on MWD.

1.5
8
Kp11 = 3 10 lt/mol/s
9
Kp11 = 3 10 lt/mol/s
10
Kp11 = 3 10 lt/mol/s
11
1.0 Kp11 = 3 10 lt/mol/s
dW/dlog(MW)

0.5
Kp11

0.0
2 3 4 5 6 7
log(MW)

Figure 3.14: Effect of Kp11 on MWD.


Chapter 3: Catalytic Polymerization Kinetics 97

1.5
9
Kp12 = 9 10 lt/mol/s
10
Kp12 = 9 10 lt/mol/s
11
Kp12 = 9 10 lt/mol/s

1.0
There is no effect of Kp12 on MWD
dW/dlog(MW)

0.5

0.0
2 3 4 5 6 7
log(MW)
Figure 3.15: Effect of Kp12 on MWD.

1.5
9
Kp21 = 3 10 lt/mol/s
10
Kp21 = 3 10 lt/mol/s
11
Kp21 = 3 10 lt/mol/s

1.0
dW/dlog(MW)

There is no effect of Kp21 on MWD

0.5

0.0
2 3 4 5 6 7
log(MW)

Figure 3.16: Effect of Kp21 on MWD.


98 Chapter 3: Catalytic Polymerization Kinetics

1.5
9
Kp22 = 9 10 lt/mol/s
10
Kp22 = 9 10 lt/mol/s
11
Kp22 = 9 10 lt/mol/s

1.0
dW/dlog(MW)

There is no effect of Kp22 on MWD

0.5

0.0
2 3 4 5 6 7
log(MW)
Figure 3.17: Effect of Kp22 on MWD.

1.5
6
KtrSP = 1 10 1/s
8
KtrSP = 1 10 1/s
9
KtrSP = 1 10 1/s
10
1.0 KtrSP = 1 10 1/s
dW/dlog(MW)

0.5
KtrSP

0.0
2 3 4 5 6 7 8
log(MW)

Figure 3.18: Effect of KtrSP on MWD.


Chapter 3: Catalytic Polymerization Kinetics 99

1.5
9
KtrH2 = 5 10 lt/mol/s
10
KtrH2 = 5 10 lt/mol/s
11
KtrH2 = 5 10 lt/mol/s

1.0
dW/dlog(MW)

0.5

KtrH2

0.0
2 3 4 5 6 7
log(MW)
Figure 3.19: Effect of KtrH2 on MWD.

1.5
1
Kd = 6 10 1/s
2
Kd = 6 10 1/s
3
Kd = 6 10 1/s

1.0
dW/dlog(MW)

0.5

Kd

0.0
2 3 4 5 6 7
log(MW)

Figure 3.20: Effect of Kd on MWD.


100 Chapter 3: Catalytic Polymerization Kinetics

0.13

Sensitivity Coefficient, (DC/c)/(DK/k)

0.12

0.00

Ka K01 K02 Kp11 Kp12 Kp21 Kp22 KtrH KtrSP Kd

-0.05

Figure 3.21: Effect of kinetic rate constants on number average molecular weight, Mn.

0.11
Sensitivity Coefficient, (DC/c)/(DK/k)

0.10

0.09

0.005

0.000
Ka K01 K02 Kp11 Kp12 Kp21 Kp22 KtrH KtrSP Kd
-0.005

Figure 3.22: Effect of kinetic rate constants on polymerization rate.


Chapter 3: Catalytic Polymerization Kinetics 101

3.7.3. Simplified Slurry Loop Reactor Kinetic Scheme

Literature review and sensitivity analysis results information leaded to the reduction
of the elementary reactions comprising the employed kinetic scheme. Thus, the number of
kinetic parameters that need to be estimated is reduced, while these reactions are able to
reproduce the actual molecular properties appearing in the industrial process studied. The
simplified kinetic scheme comprises of elementary reactions including site activation by
co-catalyst, chain initiation and propagation, chain transfer to hydrogen, to monomer and
spontaneous chain transfer as well as spontaneous deactivation (Table 3.5).

Table 3.5: The simplified kinetic scheme.

Site activation Chain transfer

k
Kk
by co-catalyst: S p  A K

aA
P0k  B by hydrogen: Pnk,i  H 2 
trHi
P0k  Dnk

Kk
Chain initiation by monomer j: Pnk,i  M j 
trMj
P1k, j  Dnk

Kk Kk
by monomer i: P0k  M i 
0i
 P1,ki spontaneous: Pnk,i 
trSpi
 P0k  Dnk

Chain propagation Site deactivation

Kk k
by monomer 1: Pnk,1  M 1 

p 11
 Pnk1,1 spontaneous: Pnk,i K

dSP
 CDk  Dnk

Pnk, 2  M 1 

Kk
 Pnk1,1 Kk
P0k  k
p 21
dSP
CD

Kk
by monomer 2: Pnk,1  M 2 

p 12
 Pnk1, 2
Kk
Pnk, 2  M 2 

p 22
 Pnk1, 2

3.7.3.1. Simplified Production/ Consumption Rates

According to the simplified kinetic mechanism, the net production/ consumption


rates have also been simplified. All kinetic rate constants regarding reactions that do not
occur according to the simplified kinetic scheme are set equal to zero. Specifically, the
rates describing the simplified scheme become:
102 Chapter 3: Catalytic Polymerization Kinetics

Polymer Moments

Zero live moment for monomer 1:

k
R k  k01   M1    P0k   k kp 21   M1   0,2
k
 k kp12   M 2   0,1
k
0,1  
(3.32)
k
 ktrH k
  H 2   ktrSP k
 0,1  k
 kdSP k
 0,1

Zero live moment for monomer 2:

k
R k  k02   M 2    P0k   k kp12   M 2   0,1
k
 k kp 21   M1   0,2
k
0,2  
(3.33)
k k k k k
(ktrH   H 2   ktrSP )  0,2  kdSP  0,2

First live moment for monomer 1:

k
R k  k01   M1    P0k   k kp11   M1   0,1
k
 k kp 21   M1   1,2
k
 k kp 21   M1   0,2
k
 k kp12   M 2   1,1
k
1,1  
(3.34)
k k k k k
(ktrH   H 2   ktrSP )  1,1  kdSP  1,1

First live moment for monomer 2:

k
R k  k02   M 2    P0k   k kp12   M 2   1,1
k
 k kp12   M 2   0,1
k
 k kp 22   M 2   0,2
k
1,2  
(3.35)
 k kp 21   M1   1,2
k k
 (ktrH k
  H 2   ktrSP k
)  1,2 k
 kdSP k
 1,2

Second live moment for monomer 1:

k
R k  k01
2,1
  M1    P0k   2  k kp11   M1   1,1
 
k
 k kp11   M1   0,1
k

 k kp 21   M1   2,2
k k
 2  1,2 k
 0,2 
(3.36)
k kp12   M 2   2,1
k k
 (ktrH k
  H 2   ktrSP k
)  2,1 k
 kdSP k
 2,1
Chapter 3: Catalytic Polymerization Kinetics 103

Second live moment for monomer 2:

k
R k  k02
2,2
  M 2    P0k   k kp12   M 2   (2,2
 
k k
 2  1,2 k
 0,2 
)  k kp 22   M 2   2  1,2
k k
 0,2 
(3.37)
 k kp 21   M1   2,2
k k
 (ktrH k
  H 2   ktrSP k
)  2,2 k
 kdSP k
 2,2

Zero dead moment:

k
R k  (ktrH
0
k
  H 2   ktrSP k
)  0,1 k
 (ktrH k
  H 2   ktrSP k
)  0,2 k
 kdSP k
 0,1 k
 0,2  (3.38)

First dead moment:

k
R k  (ktrH
1
k
  H 2   ktrSP k
)  1,1 k
 (ktrH k
  H 2   ktrSP k
)  1,2 k
 kdSP k
 1,1 k
 1,2  (3.39)

Second dead moment:

k
R k  (ktrH
2
k
  H 2   ktrSP k
)  2,1 k
 (ktrH k
  H 2   ktrSP k
)  2,2 k
 kdSP k
 2,1 k
 2,2   (3.40)

Liquid-phase Components

Monomer 1 (ethylene):

s N s N s N
k
RM1    k01   M1    P0k    k Pk 11   M1   0,1
k
  k Pk 21   M1   0,2
k
(3.41)
k 1
  k 1 k 1

Monomer 2 (1-hexene):

s N Ns Ns
k
RM 2    k02   M 2    P0k    k Pk12   M 2   0,1
k
  k Pk 22   M 2   0,2
k
(3.42)
k 1
  k 1 k 1

Hydrogen:

 
sN
k k k
RH 2    ktrH   H 2   0,1  0,2 (3.43)
k 1
104 Chapter 3: Catalytic Polymerization Kinetics

Co-catalyst:

sN
k
R A    kaA   A   S kp  (3.44)
k 1
 

Polymer Related Quantities

Potential active sites:

k
RSp k  kaA   A   Sp k  (3.45)
 

Vacant catalyst sites:

k
RP k  kaA
0
  A   S kp   k01
 
k k
  M1   k02   M 2    P0k   ktrH
 
k
 k
  H 2   0,1 k
 0,2  (3.46)

k k k k
 ktrSP1  0,1  ktrSP 2  0,2

3.7.4. Deconvolution of Molecular Weight Distribution

The deconvolution of a polymer MWD can determine the minimum number of


different catalyst active site types that is able to give an accurate representation of the
molecular weight distributions generated by Ziegler-Natta catalysts, as well as the weight
fraction and number average molecular weight of polymer produced by each catalyst site
type. Moreover, the deconvolution procedure can give us an indication on the
determination of the kinetic rate constant values. A detailed description of MWD
deconvolution can be found in literature (McAuley et al., 1990, Soares and Hamielec,
1995, Soares and Hamielec, 1996, de Carvalho et. al., 1989, de Carvalho et. al., 1990 and
Hakim and Moballegh, 2006).

Representation of WCLD as a Weighted Sum of Most Probable WCLD’s

The most probable weight chain length distribution (i.e., WCLD) of linear homo-
and binary co-polymers may be expressed by the equation:
Chapter 3: Catalytic Polymerization Kinetics 105

w  r, j    2 ( j )  r  exp   ( j )  r  (3.47)

where w  r, j  is the most probable WCLD (weight fraction of polymer of chain length n

produced on site type i instantaneously),  2 ( j ) is the ratio of rate of production of dead


polymer chain to rate of propagation (represents the inverse of the M n of polymer

produced), r is the polymer chain length and j is the active site type.

The instantaneous WCLD of the whole polymer is obtained by averaging the


distributions of each individual site type:
n
W  r    m( j )  w  r, j  (3.48)
j 1

where W  r  is the instantaneous weight fraction of polymer of chain length r , m( j ) is

the weight or mass fraction of polymer made by site type j and n is the number of different
catalyst active site types.

Eq. 3.48 is modified to include the constraint:


n

 m( j )  1
j 1
(3.49)

And is finally expressed as:

 n 1
W  r   w( r,n )   m( j )  w( r, j )  w( r,n ) (3.50)
j 1


We want to minimize the difference between W  r  computed by Eq. 3.50 and the

measured distribution W  r  in order to determine the adjustable parameters

 ( 1 ),  ( 2 ), ...,  ( n ) and m( 1 ), m( 2 ), ..., m( n ) .

Eq. 3.50 is valid for describing an instantaneous WCLD formed by the superposition
of several individual most probable WCLDs. However, in practice, one has information
106 Chapter 3: Catalytic Polymerization Kinetics

only about WCLDs for polymer accumulated over a finite polymerization time. Therefore,
to apply the proposed model to experimental WCLDs, we must assume that the WCLD for
the accumulated polymer is essentially the same as the instantaneous WCLD. This
assumption is valid when:

1) The polymerization reactor is operated at steady state conditions and the WCLD
is spatially independent.

2) The ratio of transfer to propagation rates of all active sites does not change
during the polymerization.

3) The relative amounts of polymer made by each site type do not change during
the polymerization.

4) Mass- and heat-transfer effects are negligible, since these effects could give an
instantaneous WCLD that is spatially dependent.

Finally, it should be mentioned that the values of mi and ti were estimated using a
non-linear parameter estimator and available experimental MWDs for different polymer
grades. The numerical methods used for the MWD deconvolution was the generalized
regression estimator (i.e., GREG).

According to the described analysis the deconvolution of the GPC data for a polymer
can be derived. Thus, the minimum number of catalyst active site types that accurately
describes the MWD, the weight fraction of polymer and the corresponding M n produced

at each catalyst site type and the predicted M n and M w for the entire distribution can be
determined.

In Figure 3.23, the effect of number of different catalyst active site types on %
square error deviation of the predicted MWD to the actual MWD is presented. It is evident
that as the number of catalyst active sites increases, the percentage square error deviation
decreases to a minimum value (i.e., 0.01).

In Figure 3.24, the effect of number of different catalyst active site types on the
derivative of the square error deviation of the predicted MWD to the actual MWD, is
depicted. As can be seen, the critical number of different catalyst active site types is 4, as
further increase of the number of site types would not significantly increase the derivative
of the deviation. With 4 different catalyst active site types, the expected deviation of the
predicted MWD to the actual MWD is less than 0.05%. Thus, the simplified multi-site
Chapter 3: Catalytic Polymerization Kinetics 107

catalytic kinetic scheme (see Table 3.5) can be deducted to a four-site Z-N polymerization
kinetic scheme.

In Figure 3.25, the experimentally measured MWD (marked with discrete points) is
compared with the reconstructed MWD using a two- to five-catalyst active sites model
(regarding Grade 1 data). The results indicate that at least a four-site model should be
considered to predict the shape of MWD.

Table 3.6 shows a representative set of deconvolution results for MWD provided by
TPRF plant data (regarding 6 different grades data). Table 3.6 depicts the fraction of each
site type for all different grades, based on the extracted results, for a four-site kinetic
scheme. The fraction values for each grade give us an indication on catalyst type
distribution. The average value of these calculated site type fractions for all grades can
give a ‘statistically’ correct catalyst site type distribution. Based on similar calculations, on
different grades, the site type fractions were determined equal to 0.1:0.2:0.3:0.4 for the 4
different catalyst active site types. It should be mentioned that all catalyst sites of the same
type have exactly the same kinetic rate constants. The values of number average molecular
weight produced by each catalyst site type are also presented. This information can guide
us for the determination of the kinetic rate constants through kinetic model tuning process.
108 Chapter 3: Catalytic Polymerization Kinetics

100

10
% Deviation

0.1

0.01

1E-3
1 2 3 4 5 6
Number of distributions

Figure 3.23: Effect of number of different catalyst active site types on % deviation of the
predicted MWD to the actual MWD.

0
Derivative of deviation

-4

-8

-12

-16

-20
1 2 3 4 5 6

Number of distributions

Figure 3.24: Effect of number of different catalyst active site types on the derivative of the
deviation of the predicted MWD to the actual MWD.
Chapter 3: Catalytic Polymerization Kinetics 109

a. 1.0 b. 1.0
Plant data 3 different catalyst active site types Plant data 3 different catalyst active site types
Predicted distribution Predicted distribution
Site type 1 Site type 1
0.8 0.8
Site type 2 Site type 2
Site type 3
dW/dlog(MW)

0.6 0.6

dW/dlog(MW)
0.4 0.4

0.2 0.2

0.0 0.0
2 3 4 5 6 7 8 2 4 6 8
log(MW) log(MW)

c. 1.0 d. 1.0
4 different catalyst active site types 5 different catalyst active site types
Plant data Plant data
Predicted distribution Predicted distribution
Site type 1 0.8 Site type 1
0.8
Site type 2 Site type 2
Site type 3 Site type 3
Site type 4 dW/dlog(MW) Site type 4
Site type 5
dW/dlog(MW)

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
2 4 6 8 2 4 6 8
log(MW)
log(MW)

Figure 3.25:GPC deconvolution results for a representative ethylene/1-hexene co-polymer


grade from TPRF industrial plant.
110 Chapter 3: Catalytic Polymerization Kinetics

Table 3.6: Deconvolution of various TPRF grades.

Grade Grade 1 Grade 2 Grade 3 Grade 4 Grade 5 Grade 6

Fraction 1st site 0.125 0.142 0.132 0.110 0.110 0.070

2nd site 0.170 0.218 0.138 0.230 0.210 0.150

3rd site 0.300 0.310 0.330 0.290 0.290 0.370

4th site 0.405 0.330 0.400 0.370 0.390 0.410

logMn 1st site 5.82 6.02 4.52 6.00 6.00 3.90

2nd site 4.43 4.62 6.00 4.60 4.60 5.62

3rd site 5.29 5.47 5.42 5.47 5.47 5.05

4th site 4.82 5.02 4.91 5.00 5.00 5.60

3.7.5. Kinetic Rate Constants Estimation Procedure

The exact estimation of kinetic parameters is a fundamental step in process


modelling. From the deconvolution of the first reactor we can estimate  i and mi for each

active site. As the  i is equal to the summation of all rates of deactivation and transfer
divided by the rate of polymerization, the rate constants for each site should be adjusted
such that the calculated  i for each site becomes equal to the  i obtained from
deconvolution. In the slurry-loop process, in the first reactor, the spontaneous chain
transfer constant is more important than the other transfer constants (i.e., transfer to
hydrogen. Hydrogen is mainly added in the second reactor of the process.). Therefore, by
changing the spontaneous chain transfer and propagation rate constants,  i was adjusted so
that it equalled to the one obtained from deconvolution of the GPC graph related to the first
reactor of the plant. Moreover, the production rate, molecular weight, and MWD of the
product from the first reactor could be predicted. Other kinetic rate constants, such as chain
transfer to hydrogen, propagation rate constants related to the co-monomer were adjusted
in a way to achieve the desired molecular weight, production rate, MWD, and the shape of
Chapter 3: Catalytic Polymerization Kinetics 111

MWD of the produced polymer in the second reactor. The GPC graph related to the second
reactor was used just for checking the shape of the resulted MWD. If in a trial the desired
targets were not satisfied, in the next trial, the kinetic parameters of the first reactor would
be adjusted again. Considering the polymerization kinetic parameters should be the same
in the two reactors, the procedure would be repeated until the targets were achieved in each
reactor. The numerical values of all the kinetic rate constants are reported in Table 3.7
(Touloupides et al., 2010).

It should be noted that the kinetic rate constants tuning (regarding molecular
developments) is a procedure that was employed after the accomplishment of the whole
reactor model, as concentrations appearing in kinetic rate equations correspond to the local
species concentration at the active catalyst sites. Thus, only after the total reactor model
development (including kinetic, thermodynamic and reactor modelling) it was made
possible to account for the solubility effects in order to calculate the local species
concentration.
112 Chapter 3: Catalytic Polymerization Kinetics

Table 3.7: Numerical values of kinetic rate constants for a four-site Ziegler-Natta
ethylene1-hexene co-polymerization mechanism.

Pre-exponential factor (L/mol/s) Activation energy (kcal/mol)

Site 1 Site 2 Site 3 Site 4 Site 1 Site 2 Site 3 Site 4

Activation

k
K aA 3.5 104 3.5 104 3.5 104 3.5 104 12 12 12 12

Initiation

K 0i
k 4.0 109 4.0 109 4.0 109 4.0 109 12 12 12 12

Propagation

k
K p11 5.0 1010 5.0 1010 5.0 1010 5.0 1010 12 12 12 12

k
K p21 5.0 1010 5.0 1010 5.0 1010 5.0 1010 12 12 12 12

k
K p12 1.0 1010 1.0 1010 1.0 1010 1.0 1010 12 12 12 12

k
K p22 1.0 108 1.0 108 1.0 108 1.0 108 12 12 12 12

Chain
transfer

k
K trH 2.6 1012 5.6 1011 1.8 101 6.5 1010 14 14 14 14

k
K trSpi 2.0 109 8.3 108 2.4 108 5.5 107 14 14 14 14

Deactivation

k
K dSP 6.0 103 6.0 103 6.0 103 6.0 103 12 12 12 12
Chapter 3: Catalytic Polymerization Kinetics 113

3.8. Notation

Symbols

Cd   : concentration of deactivated catalyst sites, mol/l

Dn : concentration of ‘dead’ co-polymer chains of length ‘ n ’, mol/l

DPn   : Number average degree of polymerization

DPw   : Weight average degree of polymerization

k0 : kinetic rate constant of initiation reaction, l/mol/s

ka : kinetic rate constant of activation reaction, l/mol/s

kd : kinetic rate constant of deactivation reaction, l/mol/s

kp : kinetic constant of propagation reaction, l/mol/s

kinetic constant of site transformation reaction at ‘ k ’ catalyst site to


ktkl :
‘ l ’catalyst site, l/mol/s

ktr : kinetic constant of chain transfer reaction, l/mol/s

Mn : number average molecular weight, g/mol

Mw : weight average molecular weight, g/mol

MW : average molecular weight of monomer unit, g/mol

MWi : molecular weight of monomer ‘ i ’, g/mol


114 Chapter 3: Catalytic Polymerization Kinetics

n : polymer chain length

Nm : total number of monomers

Ns : total number of catalyst active site types

PD : Polydispersity

t : Time, sec

 A : co-catalyst concentration, mol/l

 B  : co-catalyst by-product concentration, mol/l

 M i  : monomer ‘ i ’ concentration, mol/l

Pk  : Vacant catalyst sites of type ‘ k ’ concentration, mol/l


 o 

Ri : production/consumpiton rate of molecular species ‘ i ’, mol/l

 S  : diluent concentration, mol/l

 Sp k  : concentration of potential active sites of type ‘ k ’, mol/l


 

 X  : Poison concentration, mol/l

Subscripts and Superscripts

A : co-catalyst

B : co-catalyst by-product

Dn   : ‘dead’ polymer chains of length ‘ n ’

H2 : Hydrogen
Chapter 3: Catalytic Polymerization Kinetics 115

i : Referring to molar species ‘ i ’

k : catalyst type active site

Mi : monomer ‘ i ’

P0k : vacant catalyst sites

‘Live’ polymer chain of length ‘n’, ending in ‘i’ monomer, produced in


Pnk,i :
catalyst site type ‘ k ’

S : diluent

SP : spontaneous

Sp k : Potential active sites of type ‘ k ’

X : Poison

Greek letters

νth moment of the total number chain length distribution of ‘live’ co-
v :
polymer chains

νth moment of the total number chain length distribution of ‘live’ co-
v,i :
polymer chains, with terminal monomer ‘ i ’

 : viscosity, kg/m/s

νth moment of the total number chain length distribution of ‘dead’ co-
v :
polymer chains

νth moment of the total number chain length distribution of ‘bulk’ co-
 :
polymer chains

i : Average instant co-polymer composition

 ik : Average cummulative co-polymer composition


116 Chapter 3: Catalytic Polymerization Kinetics

3.9. References

Anderson A., Cordes H.G., Herwig J., Kaminski A., Mark A., Mottweils R., Sinn J.H. and
Vollmes H.J., 1976, Anfew. Chem. Int. Ed. Engl., 15, 630.

Bhagwat M.S., Bhabwat, S.S. and Sharma M.M., 1994, ‘Mathematical Modeling of the
Slurry Polymerization of Ethylene: Gas-Liquid Mass Transfer Limitations’ Ind, Eng.
Chem. Res., 33,pp.2322-2330.

Boor J., 1979, ‘Ziegler-Natta Catalysts and Polymerizations’, Academic Press, New York.

Brintzinger H.H., Fischer D., Mulhaupt R., Rieger B. and Waymouth R.M., 1995,
‘Stereospecific Olefin Polymerization with Chiral Metallocene Catalysts’, Angew.
Chem. Int. Ed. Engl., 34, pp. 1143-1170.

Canu P. and Ray W.H., 1991, ‘Discrete Weighted Residual Methods Applied to
Polymerization Reactions’, Comp. Chem. Eng., 15, 549-564.

Choi K.Y. and Ray W.H., 1985a, ‘Recent Developments in Transition Metal Catalyzed
Olefin polymerization-A Survey. I. Ethylene Polymerization’, Polymer Reviews,
25:1, pp. 1-55.

Choi K.Y. and Ray W.H., 1985b, ‘Recent Developments in Transition Metal Catalyzed
Olefin polymerization-A Survey. II. Propylene Polymerization’, Polymer Reviews,
25:1, pp. 57-97.

Choi H.K., Kim J.H., Ko Y.S. and Woo S.I., 1997, ‘Prediction of the Molecular Weight of
Polyethylene Produced in a Semibatch Slurry Reactor by Computer Simulation’, Ind.
Eng. Chem. Res., 36, pp. 1337-1342.

Clay, P.A. and Gilbert, R.G., 1995, ‘Molecular Weight Distributions in Free-Radical
Polymerizations. 1. Model Development and Implications for Data Interpretation’,
Macromolecules, 28, 552-569.

Cossee P., 1964, ‘Ziegler-Natta Catalysis I. Mechanism of Polymerization of a-Olefins


with Ziegler-Natta Catalysts’, Journal of Catalysis, 3, pp. 80-88.

de Carvalho A.B., Gloor P.E., Hamielec A.E., 1989, ‘A Kinetic Mathematical Model for
Heterogeneous Ziegler-Natta Copolymerization’, Polymer, 30, pp. 281-296.
Chapter 3: Catalytic Polymerization Kinetics 117

de Carvalho A.B., Gloor P.E., Hamielec A.E., 1990, ‘A Kinetic Model for Heterogeneous
Ziegler-Natta Copolymerization. Part 2: Stereochemical Sequence Length
Distributions’, Polymer, 31, pp. 1294-1311.

Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C., 2008, ‘Development
of a Multi-scale, Multi-phase, Multi-zone Dynamic Model for the Prediction of
Particle Segregation in Catalytic Olefin Polymerization FBRs’, Chem. Eng. Sci., 63,
pp. 4735-4753.

Dusseault J.J.A. and Hsu C.C., 1993, ‘MgCl2-Supported Ziegler-Natta Catalysts for Olefin
Polymerization: Basic Structure, Mechanism, and Kinetic Behavior’, Rev.
Macromol. Chem. Phys., C32(2), pp. 103-145.

Floyd S., 1986, ‘Heat and Mass Transfer Resistances in Polymerization of Olefins over
Solid Catalysts’, PhD thesis, University of Wisconsin-Madison.

Fontes C.H. and Mendes M.J., 2005, ‘Analysis of an Industrial Continuous Slurry Reactor
for Ethylene–Butane Copolymerization’, Polymer, 46, pp. 2922-2932.

Galli P. and Vecellio G., 2001, ‘Technology: Driving Force Behind Innovation and Growth
of Polyolefins’, Prog. Polym. Sci., 26, 8, pp. 1287-1336.

Ha K-S., Yoo K-Y. and Rhee H-K., 2001, ‘Modeling and Analysis of a Slurry Reactor
System for Heterogeneous Olefin Polymerization: The Effects of Hydrogen
Concentration and Initial Catalyst Size’, Journal of Applied Polymer Science, Vol.
79, pp. 2480-2493.

Hakim S. and Moballegh L., 2006, ‘Simulation of a Series of Industrial Slurry Reactors for
HDPE Polymerization Process Using Deconvolution of the GPC Graph of Only the
First Reactor’, Iranian Polymer Journal, 15, pp. 655-666.

Hatzantonis H., Yiannoulakis H., Yiagopoulos A. and Kiparissides, C., 2000, ‘Recent
Developments in Modeling Gas-Phase Catalyzed olefin Polymerization Fluidized-
Bed Reactors: The Effect of Bubble Size Variation on the Reactor’s Performance’,
Chem. Eng. Sci., 55, pp. 3237-3259.

Hogan J.P. and Banks R.L., 1958, ‘Polymers and Production Thereof’, U.S. Patent No.
2,825,721.
118 Chapter 3: Catalytic Polymerization Kinetics

Huang J. and Rempel G.L., 1995, ‘Ziegler-Natta Catalysts for Olefin Polymerization:
Mechanistic Insights from Metallocene Systems’, Prog. Polym. Sci., 20, pp. 459-
526.

Hutchinson R., 1990, ‘Modeling of Particle Growth in Heterogeneous Catalyzed Olefin


Polymerizations’, PhD thesis, University of Wisconsin-Madison.

Hutchinson, R. and Ray, W.H., 1991, ‘Polymerization of Olefins Through Heterogeneous


Catalysis. The Effect on Condensation Cooling on Particle Ingition’, d. Appl. Polym.
Sci., 43, pp. 1387-1390.

Khare N.P., 2003, ‘Predictive Modeling of Metal-Catalyzed Polyolefin Processes’, PhD


thesis, Virginia Polytechnic Institute and State University.

Kiparissides C., 1996, ‘Polymerization Reactor Modeling: A Review of Recent


Developments and Future Directions’, Chem. Eng. Sci., 51, pp. 1637-1659.

Kissin Y.V., 1985, ‘Isospecific Polymerization of Olefins with Heterogeneous Ziegler-


Natta Catalysts’, Springer-Verlag, New York.

Lorenzini P. and Bertrand P. and Vilermaux J., 1991, ‘Modeling Ethylene and a-Olefin
Copolymerization Using Ziegler-Natta Catalyst’, Can. J. Chem. Eng., 69, 682.

McAuley K.B., MacGregor J.F. and Hamielec A.E., 1990, ‘A Kinetic Model for Industrial
Gas-Phase Ethylene Polymerization’, AIChE Journal, 36, pp. 837-850.

McDaniel M.P., 2010, ‘A Review of the Phillips Supported Chromium Catalyst and Its
Commercial Use for Ethylene Polymerization’, Advances in Catalysis, 53, pp. 123-
606.

McKenna T.F. and Soares J.B.P., 2001, ‘Single Particle Modelling for Olefin
Polymerization on Supported Catalysts: A Review and Proposals for Future
Developments’, Chem Eng Science, 56, pp. 3931–3949.

Neto A.G.M., Freitas M.F., Nele M. and Pinto J. C., 2005, ‘Modeling Ethylene/1-Butene
Copolymerizations in Industrial Slurry Reactors’, Ind. Eng. Chem. Res., 44, pp.
2697-2715.

Natta G., 1955, J. Polym. Sci., 16, 143.

Natta G., 1956, Angew. Chem., 68, 393.


Chapter 3: Catalytic Polymerization Kinetics 119

Nirisen O. and Rytter E., 1986, ‘First Order Activation Second Order Deactivation
Kinetics Applied to the Polymerization with Supported Ziegler-Natta Catalysts’,
Macromol. Chem., 7, pp. 103-108.

Peacock A.J., 2000, ‘Handbook of Polyethylene’, Marcel Dekker, New York, pp. 1-10, 43-
66.

Ray W.H., 1972, ‘On the Mathematical Modeling of Polymerization Reactors’,


J.Macromol.Sci., 8, pp. 1-56.

Razavi A., 2000, ‘Metallocene Catalysts Technology and Enviroment’, C. R. Acad. Sci.
Paris, Serie Iic, Chimie/Chemistry, 3, pp. 615-625.

Saltelli A., Tarantola S., Campolongo F. and M. Ratto, 2004, ‘Sensitivity Analysis in
Practice: A Guide to Assessing Scientific Models’, John Wiley & Sons, Ltd.

Saltelli A., Ratto M., Andres T., Campolongo F., Cariboni J., Gatelli D., Saisana, M., and
Tarantola S., 2008, ‘Global Sensitivity Analysis. The Primer’, John Wiley & Sons.

Sinn H. and Kaminsky W., 1980, ‚ ‘Ziegler-Natta Catalysis’, Advances in Organometallic


Chemistry, 18, pp. 99-149.

Soares J.B.P. and Hamielec A. E., 1995, ‘Deconvolution of Chain-length Distributions of


Linear Polymers Made By Multiple-site Catalysts’, Polymer, 11, pp. 2257-2263.

Soares J.B.P. and Hamielec A.E., 1995, ‘Metallocene/aluminoxane catalysts for olefin
polymerization. A review’, Polymer Reaction Engineering, 3, pp. 131-200.

Soares J.B.P. and Hamielec A.E., 1996, ‘Copolymerization of Olefins in a Series of


Continuous Stirred-tank Slurry-reactor Using Heterogeneous Ziegler-Natta and
Metallocene Catalysis. 1. General Dynamic Mathematical Model’, Polym. Reac.
Eng., 4, pp. 153-191.

Soga K.T. and Ikeda S., 1980, ‘Polymerization of Propylene over Metal Oxides-Supported
TiCl3 Catalysts’, Polymer Bulletin, 1, pp. 849-856.

Touloupides V., Kanellopoulos V., Pladis P., Kiparissides C., Mignon D. and Van-
Grambezen P., 2010, ‘Modeling and Simulation of an Industrial Slurry-phase
Catalytic Olefin Polymerization Reactor Series’, Chem. Eng. Sci., 65, pp. 3208-
3222.
120 Chapter 3: Catalytic Polymerization Kinetics

Vandenberg E.J., 1962, ‘Process for Polymerizing Olefins Wherein Hydrogen Is Utilized
as a Molecular Weight Control Agent’, U.S. Patent No. 3,051,690.

van der Ven S., 1990, ‘Polypropylene and Other Polyolefins: Polymerization and
Characterization’, Elsevier Science Publishers B. V., Netherlands.

Wells G.J., Ray W.H. and Kosek J., 2001, ‘Effects on Catalyst Activity Profiles on
Polyethylene Reactor Dynamics’, AIChE Journal, Vol. 47, No. 12, pp. 2768-2780.

Wu L., Bu N. and Wanke E., 2005, ‘Kinetic Behavior of Ethylene_1-Hexene


Copolymerization in Slurry and Solution Reactors’, Journal of Polymer Science,
Vol. 43, pp. 2248-2257.

Xie T., McAuley K.B., Hsu J.C.C., Bacon, D.W., 1994, ‘Gas Phase Ethylene
Polymerization: Production Processes, Polymer Properties, and Reactor Modeling’,
Industrial and Engineering Chemistry Research, 33, pp. 449-479.

Yiannoulakis H., Yiagopoulos and A., Kiparissides C., 2001, ‘Recent developments in the
particle size distribution modeling of fluidized-bed olefin polymerization reactors’,
Chem. Eng. Sci., 56, pp. 917-925.

Zacca J.J. and Ray H.W., 1993, ‘Modelling of the liquid phase polymerization of olefins in
loop reactors’, Chem. Eng. Sci., Vol. 48, No. 22, pp. 3743-3765.

Zacca J.J., 1995, ‘Distributed Parameter Modelling of the Polymerization of Olefins in


Chamical Reactors’, PhD thesis, University of Wisconsin-Madison.

Zechlin J., Hauschild K. and Fink G., 2000, ‘Silica Supported Metallocene/MAO-Systems:
Comparison of the Polypropylene Growth During Bulk Phase polymerization with
Slurry phase experiments’, Macromol. Chem. Phys., 201, pp. 597-603.
Chapter 4: Slurry-loop Reactor Modelling 121

Chapter 4

4. Slurry-loop Reactor Modelling

4.1. Introduction

In the present chapter, the modelling approach followed for the dynamic simulation
of a typical industrial-scale slurry-loop reactor series is presented. All the steps followed
for the simulation of the industrial slurry-loop reactor, including kinetic modelling, mass
and energy balances, thermodynamic considerations, physical and transport properties
calculation, reactor control system modelling, rheological properties calculation, particle
size distribution calculation, etc., are thoroughly analyzed.

4.2. Slurry-loop Polymerization Process

The first commercially available technology for the production of high density,
linear polyethylene grades was the solution process. It was soon discovered that a more
efficient way to produce various PE grades was to carry out the polymerization under
slurry conditions (Franklin et al., 2004, Hottovy et. al., 2004). In the sixties, Phillips
Petroleum Co. developed the first loop reactor technology for the production of polyolefins
in a slurry-phase. According to this technology, the polymerization is carried out in one or
more continuous loop reactors operating in series. The product is withdrawn from the
bottom of the loop reactor via the semi-continuous operation of a product withdrawal
system (e.g., settling legs). The slurry-phase loop reactor technology has gained
122 Chapter 4: Slurry-loop Reactor Modelling

considerable acceptance and large quantities of ethylene co-polymers are annually


produced via this process (Hottovy et. al., 2004). While technology for preparing polymers
has changed with respect to catalyst and reactants, the same general loop reactor
technology, employed in the 1960’s is still in use (Noll, 2005). Nowadays, continuous
slurry-phase polymerization, in the presence of a heterogeneous Z–N catalyst, is one of the
most commonly employed processes in the production of polyolefins, including high-
density polyethylene (HDPE), isotactic polypropylene (IPP) as well as their co-polymers
with higher olefins (Fontes and Mendes, 2005, Kufeld et al., 2006). Presently, 35% of the
total polypropylene and 57% of the total polyethylene volume are produced via the slurry-
phase technology (Reginato et al., 2003).

The main reasons for the wide use of slurry-phase loop reactors are:

 their simple design and operation,


 their well-defined mixing conditions,
 their excellent heat transfer capabilities,
 the low power requirements, and,
 high conversion rates.
Catalyst
Co-Catalyst

Figure 4.1: Schematic representation of an industrial slurry-phase olefin catalytic polymerization cascade-loop reactor series.
124 Chapter 4: Slurry-loop Reactor Modelling

4.2.1. The Slurry-phase Olefin Polymerization Cascade-loop Reactor Process

In Figure 4.1, a schematic representation of an industrial slurry-phase olefin


polymerization cascade-loop reactor series is illustrated. The process consists of two
jacketed loop reactors. The reaction mixture (i.e., consisting of monomer(s), diluent,
catalyst, hydrogen and polymer) flows in the loop reactor by means of an axial centrifugal
pump placed at the bottom of the reactor. The reactor’s cross-sectional area is usually
uniform and the reactor operates free of any obstruction (in order to reduce fouling) that
can interfere with the circulation of the reaction mixture. An ‘O’ shape or any similar
arrangements (e.g., a vertical double loop) are the most commonly employed reactor
designs. The length of the loop reaction zone to the cylindrical wall usually defines a
length/ diameter ratio greater than 250 (Kufeld et al., 2006). The reactor is cooled by two-
pipe heat exchangers formed by the pipes and jackets, while, more heat exchangers may be
provided, if desire or useful, in the horizontal segments of the reactor (Kufeld et al., 2006).

The first reactor of the series is continuously fed with monomer, co-monomer,
hydrogen, diluent and catalyst. The olefin feed used for the polymerization is at least one
olefin selected from 1-olefins having a maximum of 8 carbon atoms per molecule and no
branching nearer the double bond than the 4-position. Examples of olefins which can be
polymerized in a slurry-loop reactor include ethylene, propylene, 1-butene, 1-pentene and
1,1-butadiene. Particle form polymer can be prepared from ethylene and mixtures of
ethylene with minor amounts of other unsaturated hydrocarbons as co-monomers, such as
propylene, 1-butene, 1-pentene, 1-hexene (Scoggin, 1966). It should be noted that the
monomers are introduced into the reactor as a gas. The monomers dissolve in the liquid
phase for the reaction to take place on the solid catalyst. Typical ethylene content in
reactors’ feed varies from 70-250 wt. % based on the weight of the diluents (Marechal,
2006). Suitable diluents include parafins such as those having from 3-12 (preferably 3-8)
carbon atoms per molecule (e.g., n-butane, n-pentane, iso-pentane, n-hexane, n-decane,
etc.) and saturated cyclic hydrocarbons (e.g., cyclohexane, cyclopentane and
methylcyclopentane, methylcyclohexane, etc.) (Scoggin, 1966).

Catalyst is introduced via a catalyst introduction port, before the centrifugal pump
(Kufeld et al., 2006). Typical catalyst concentration ranges from 0.001-5% by weight
based on the diluent (Scoggin, 1966). It should be noted that the catalyst type greatly
affects the polymer particles morphology that, in turn, can affect the particle circulation
Chapter 4: Slurry-loop Reactor Modelling 125

behaviour and particle agglomeration that can cause the appearance of temperature hot
spots in the loop reactor (Marechal, 2006). Moreover, the physical properties of a
polyethylene product vary depending on the catalytic system employed for the polymer
production. This is because different catalytic systems tend to yield different MWDs. A
number of catalytic systems have been employed for the production of polyethylene in a
slurry-loop reactor (Lonfils at al., 2000). The slurry-loop process originally employed
chromium-based catalysts (i.e., Phillips catalyst). Chromium-based catalytic systems for
the loop reactor are discussed by the Patent of Lonfils at al., 2000. Today, industrial scale
olefin polymerization reactors are commonly operated in the presence of third and fourth-
generation Z-N catalysts (Marechal, 2006, Reginato et al., 2003). The system employed for
this project work is ethylene-1-hexene Z-N polymerization with iso-butane as a diluent.

During polymerization, the polymer solids are gradually collected in the settling legs
placed at the lower part of the loop reactor. The settling legs periodically open to remove
the highly concentrated slurry (i.e., consisting of polymer solids and a fraction of the liquid
phase). The product stream leaving the first loop reactor is fed to the second reactor of the
series together with fresh monomer(s), diluent and hydrogen. The highly concentrated
slurry product leaving the second reactor is then flashed from the operating pressure (e.g.,
41 bar) to 2 bar (Salmon, 2003). The vaporized unreacted monomer(s) and diluent exit the
flash chamber via a conduit for further processing which includes condensation by simple
heat exchange using a recycle condenser, and return to the system without the necessity for
compression via a recycle diluents line. Thus, the diluent is completely recovered due to
the high monomer(s) conversion (i.e., 95%-98%) while there is no need for monomer(s)
recovery. Polymer particles are withdrawn from the high pressure flash chamber for further
processing. They are passed to the low pressure flash chamber and they are recovered as
the polymer product (Kufeld et al., 2006), which finally is dried and pelletized.

The reaction slurry is circulated by an impeller driven by a motor (Kufeld et al.,


2006). It is also possible (and preferred) that the reactor may have more than one impeller/
motor combination in series around loop. The series impellers may be close together,
distributed about the loop, or otherwise arranged in any appropriate manner. The existing
pump technology can support a reactor of roughly 170 m3, depending on the diameter and
configuration, assuming a flow velocity equal to 9.75 m/sec. In general, the fluid
circulation pump is designed to provide high flow velocities of the reaction mixture (i.e., 3-
10 m/s, Ayres, 1986) resulting in the establishment of very intensive and well-defined
126 Chapter 4: Slurry-loop Reactor Modelling

mixing conditions (Zacca and Ray, 1993). The velocity should be maintained high enough
to avoid saltation or deposition of solids from the slurry. The saltation velocity is defined
as the minimum flow velocity needed to avoid saltation and varies with process conditions.
The value of the saltation velocity is proportional to the reactor diameter, as well as, to the
percentage of solids in the slurry. Since high solids are desirable, one way to maintain flow
above the saltation velocity at high solids content is to use relatively small-diameter reactor
(Kufeld et al., 2006). Moreover, the highly turbulent flowrate ensures a high heat transfer
rate from the reaction mixture to the water coolant, flowing inside the tube’s jacket and, at
the same time, prevents reactor fouling caused by particle deposition on the internal
surface of the tube’s wall. The high flow circulation rates make possible the high-solids
operation of the loop reactor (e.g., larger than 30 % w/w). This represents a clear
advantage of the loop reactors over the operation of conventional stirred tanks where
serious mixing and heat transfer limitations can arise at high-solids concentrations (Drusco
and Rinaldi, 1984, Ferrero and Chiovetta, 1990).

The reactor is run ‘liquid full’ (Kufeld et al., 2006). In general, in a slurry-phase loop
reactor two phases, namely, a liquid phase (i.e., consisting of diluent, monomer, co-
monomer and hydrogen) and a polymer phase (i.e., consisting of polymer and sorbed
quantities of diluent, monomer, co-monomer and hydrogen) are present. Typical reactor
mass fractions for polymer, diluent and monomer would be 30%, 68% and 2%,
respectively, although these fraction values may vary considerably (Ayres, 1986).

Various polyolefin grades with broad or/ and bimodal MWDs can be produced using
two slurry-loop reactors in series. It is well known that in order to achieve good
processability of polyethylene resins, it is desired that the flow properties and the shear
response of the polyethylene are improved by broadening the MWD. Moreover, in order to
improve the mechanical properties of the resin, it is desired to separate as far as possible
the polymerization reactions occurring in the two reactors so that the difference in density
and molecular weight between the two resin fractions is enhanced for any given target
density of the resin.

One of the reactors of the series produces a high-molecular weight fraction, and the
other of the reactors produces a low- molecular weight fraction. The resultant bimodal
resin, comprising a chemical blend of the two fractions has particular applications (e.g., for
use as a pipe resin which exhibits good mechanical properties such as environmental stress
crack resistance and slow crack growth resistance) (Marechal, 2006).
Chapter 4: Slurry-loop Reactor Modelling 127

A typical proportion between the high and low molecular weight fractions is 50/ 50
wt. % blend between the two fractions. For a 50/ 50 wt. % blend between the two fractions,
it is clear to achieve a target density of the combined blend forming the pipe resin, which
target density is required commercially, for any decrease in density of high-density, low
molecular weight fraction there must be a corresponding increase in the density of the low
density high molecular weight fraction. However, the proportion between the high and low
molecular weight fractions may vary depending on the final product properties desired.
Typically, for pipe resins, the blend comprises from 40-57 wt. % high molecular weight
fraction. For film resins, the range is broader depending on the key resin properties that are
required (e.g., 30-75 wt. %). Reducing the high molecular weight fraction, decreases the
density of the high molecular weight resin at constant final resin density. Increased reactor
independence enables the achievement of enlarged density differences or enlarges
differences in molecular weight between the two fractions, leading to improved mechanical
properties (Marechal, 2006).

Typically, polyolefins of high molecular weight and low density are produced in the
first reactor of the series, commonly operated at low hydrogen concentrations and high co-
monomer concentrations (e.g., 2-15 wt. % based on the weight of the diluent). On the other
hand, in the second reactor of the series, high hydrogen (e.g., 0-5mole % based on the
weight of the diluents in the second reactor) and low co-monomer concentrations result in
the production of low molecular weight and high density polyethylene (Marechal, 2006).

The polymerization temperature in the first loop reactor typically ranges from 70-
100oC (preferably about 80oC) when the high molecular weight fraction is produced in the
first reactor and for the second reactor from 80-120oC (preferably around 95oC) when the
low molecular weight fraction is produced in the second reactor. It should be noted that the
temperature affects the solids content as an increase in temperature tends to lower the
viscosity of the diluent (Marechal, 2006). On the other hand, there is a practical limit on
increasing temperature because the point is quickly reached where the polymer goes into
solution and fouls the reactor. In general, reactor fouling occurs when polymer particles
dissolve into reactors’ liquid phase. Specifically, the polymer particles when dissolving
increase the volume that they occupy by several orders of magnitude. This increase in
volume causes the viscosity to increase in the reactor. If the polymer particles continue to
dissolve, thereby increasing the viscosity, eventually a gel-like substance is formed which
128 Chapter 4: Slurry-loop Reactor Modelling

plugs the reactor. As it is well known, unplugging a fouled reactor is a time-intensive and
costly undertaking (Benham et al., 1991).

Solids content is at least 30 wt. % based on the weight of the diluent, for a Z-N
catalyst (Marechal, 2006), while solids concentrations values equal to 45 wt. % are typical.
The solids concentration can be increased by using a circulating pump having a diameter
that is larger than the diameter of the reactor tube. This is achieved by providing a
localized enlargement in the reactor tube at the location of the propeller of the pump.
Furthermore, it is known that the replacement of the conventional settling legs in a slurry-
loop reactor, which are provided for periodic and sequential take-off of the polyethylene
fluff by a so called ‘continuous product take-off’ can also lead to higher solids
concentration in the reactor (Marechal, 2006).

Finally, the polymerization pressure ranges from 30-90, preferably about 41 bars
(Marechal, 2006). For the case of loop reactors in series, second reactor’s pressure is kept
in lower values in order the transport of first reactor’s outflow to the second reactor to be
feasible by means of pressure difference.

For stable as well as dynamic operation of an industrial slurry-phase cascade-loop


reactor process, several feedback controllers are typically required. For example, the
reaction temperature is controlled by manipulating the coolant inflow rate into the reactor
jacket; the ethylene mass fraction in the reactor-phase is controlled by manipulating the
catalyst inflow rate; the solids concentration in the reactor is controlled by the iso-butane
inflow rate; the reactor pressure is controlled by manipulating the product removal rate
(i.e., valve opening cycling period of settling legs).

4.2.2. Settling Legs Operation

A major consideration in the efficient operation of a continuous path loop reactor is


the removal of the product from the reactor. Prior methods of removal of the product
require the simultaneous removal of large quantities of diluent and reactants which then
need to be separated from the product and processed for return to the reactor (Scoggin,
1966). Settling legs are an apparatus for continuously removing polymer from the
pressurized loop olefin polymerization reactor containing a slurry of polymer particles and
fluids (Salmon., 2003). As the name implies, settling occurs in the legs to increase the
Chapter 4: Slurry-loop Reactor Modelling 129

solids concentration of the slurry finally recovered as product slurry. They are attached to
the lower horizontal section of the loop reactor to gravitationally draw off a concentrated
fraction of the solid polymer (Marwill, 1966).

Polymer slurry is periodically removed in small batches through lines and valves
which serve to isolate the severe conditions in the reactor from the downstream equipment.
Two valves are incorporated within the receiving zone permitting the entrapment of a
portion of the settled polymer product. Periodically, the bottom valve closes and the upper
valve opens to permit the introduction of polymer product into the lock. Then, the top
valve closes and the bottom opens to prevent the escape of the polymer with a small
amount of accompanying diluent and reactant (Scoggin, 1966) (see Figure 4.2).

Figure 4.2: Settling Leg Operation Cycle.

The non-continuous product withdrawal results in a periodic variation of the reactor


pressure. In typical control schemes, according to the semi-continuous operation of the
settling legs, each time that the operating pressure in the loop reactor exceeds an upper
limit the discharge valve of a settling leg opens so its polymer-solids content is emptied
from the leg (settling leg ‘firing’). This results in a decrease of the pressure in the loop
reactor that is followed by the closing of the discharge valve of the settling leg.
Subsequently, as fresh monomer(s) and diluent are continuously fed into the reactor and
new polymer is formed, the reactor pressure gradually reaches again its maximum
allowable value. Thus, the discharge valve of the next settling leg opens. This on-off valve
opening of the sequence of the settling legs results in a periodic oscillations (i.e., ‘zig-zag’)
130 Chapter 4: Slurry-loop Reactor Modelling

of the pressure in the loop reactor, affecting reactant concentrations and product quality
and consistency (Salmon, 2003).

Note that, during a given reactor period (e.g., startup, steady-state, etc.), a certain
number of settling legs can be in operation (i.e., 1-6). Thus, each settling leg will have its
next firing after a full firing-cycle (i.e., firing period) of all other settling legs in operation
is completed. This permits polymer settling in the leg, enhancing solids concentration.
During steady state operation of polymerization process, firing period values range from
20-50 sec (depending on reactor pressure and number of settling legs in operation)
(Marwil, 1966). The use of settling legs in a loop reactor for manufacturing an ethylene
butane co-polymer is well described by the patent of Marwil, 1966.

Settling legs do present two problems. First, they represent the imposition of a
‘batch’ technique onto a basic continuous process. Each time a settling leg reaches the
stage where it ‘fires’, accumulated polymer it causes an interference with the flow of slurry
in the loop reactor upstream and the recovery system downstream. Also, the valve
mechanism essential to periodically seal off the settling legs from the reactor upstream and
the recovery system downstream requires frequent maintenance due to the difficulty in
maintaining a tight seal with the large diameter valves needed for sealing the legs.
Secondly, as reactors have gotten larger, logistic problems are present by the settling legs.
If a pipe diameter is doubled, the volume of the reactor goes up four-fold. However,
because of the valve mechanism involved, the size of the settling legs cannot easily be
increased further. Hence, the number of legs required begins to exceed the physical space
available.

Despite the problems that may arise from the semi-continuous operation of the
settling legs, they are still employed in slurry-phase olefin polymerization processes
(Franklin et al., 2004), particularly for the production of high density polyethylene (Noll,
2005). This is because, unlike bulk slurry polymerizations (i.e., where the polymer is the
diluent) where solids concentrations are higher than 60%, olefin polymer slurries in a
diluent are generally limited to no more than 40% weight. Hence, settling legs have been
believed to be necessary to give a final slurry product at the exit of the settling legs of
greater than 40% (Franklin et al., 2004, Hottovy at al., 2004). The enhancement factor (i.e.,
the ration of solids concentration at the exit of settling legs to solids concentration in the
reactor) in steady-state operation for settling legs operation is around 1.25. It must be
Chapter 4: Slurry-loop Reactor Modelling 131

emphasized that in a commercial operation as little as one percent point increase in solids
concentration is of major significance (Franklin et al., 2004).

4.3. Process Models

A model is a mathematical abstraction of a real process. It requires comprehensive


understanding of polymerization processes, physical phenomena, and chemical reaction
mechanisms. As can be easily understood, the set of equations that comprise the model are
at best an approximation to the true process. The level of physical and chemical detail in
the model, requiring corresponding amount of time and effort, is directly related to the
expected benefits to be derived from its use.

The importance and benefits of reactor modelling have been widely recognized by
both industrial and academic researchers. Process models play a significant role in
improving product quality, plant efficiency, and safety. A mathematical model that can
reliably predict the behaviour of a specific unit or process, becomes a valuable tool that can
be applied to all tasks of process operation. The use of these models in process design,
simulation, optimization, and control, promises to have a profound commercial impact on
the chemical industry is highly desirable. In the past 30 years, the development of
computer simulation tools for the chemical industry has dramatically advanced.
Mathematical models can be helpful in process analysis and control in the following ways:

 To improve understanding of the process. Process models can be used in a


computer simulation of the process to investigate process behaviour without the
expense or unexpected hazards of operating the real process.
 To train plant operating personnel.
 To design the control strategy of a new process.
 To select controller settings.
 To design the control law. Modern control techniques incorporate a process
model into the control law (i.e., model-predictive or model-based control).
 To optimize process operating conditions.
132 Chapter 4: Slurry-loop Reactor Modelling

In recent years, interest has focussed on the prediction of the end-use polymer
properties (i.e., number- and weight average molecular weight, molecular weight
distribution, co-polymer composition, polymer density, complex viscosity, particle size
distribution, etc.). To develop a predictive model, account must be taken of the chemistry
and physics of all of the relevant microscopic processes which occur in the polymerization
process. Detailed physical property and thermodynamic data on the partitioning of species
among phases is required to quantitatively calculate the concentrations of reactants at the
loci of polymerization. Valid kinetic rate constants (frequency factors and activation
energies) are also required (Ulmann’s Encyclopedia of Industrial Chemistry). The fact that
each polymerization process involves a number of unique physical and chemical
phenomena (e.g., reaction kinetics, physical and transport phenomena, thermodynamics,
reactor configuration, etc.) increases the scope for the development of custom-made
computer aided design software tools for specific polymerization processes (Krallis et al.,
2010).

4.3.1. Multi-scale Modelling Approach

A major objective of polymerization reaction engineering is to understand how the


reaction mechanism, the physical transport phenomena (e.g., mass and heat transfer,
mixing), reactor type and reactor operating conditions affect the ‘polymer quality’ of the
final product. The term ‘polymer quality’ includes all the molecular structure properties
(e.g., MWD, stereoregularity, etc.) and the macroscopic morphological properties of the
polymer product (e.g., particle size distribution, porosity, bulk density, etc.). Complex
systems are characterized by hierarchical multi-scale nature with respect not only to space
but also to time, showing dissipative structures induced by inherent nonlinear and non-
equilibrium interactions and stabilized by exchanging energy, matter and information with
their surroundings. It should be noted that the phenomena occurring over time and length
scales may differ by several orders of magnitude. Bridging these scales with integrated,
multi-scale models, correlating the molecular behaviour with the processes occurring in
reactors, is a challenging problem for control and optimization (Li et al., 2004). According
to Ray, 1986 and Ray, 1991, a modelling hierarchy can be defined. The various chemical
and physical phenomena occurring in a polymer reactor can be classified into the following
three levels of modelling:
Chapter 4: Slurry-loop Reactor Modelling 133

1) Microscale chemical kinetic modelling

2) Mesoscale physical/ transport modelling

3) Macroscale dynamic reactor modelling

4.3.1.1. Microscale Kinetic Modelling

Polymer chain reactions (e.g. chain growth, chain branching, chain termination, etc.)
related with the kinetic mechanism occur at the microscale. If the elementary reaction steps
of a polymerization mechanism are known, several mathematical techniques are available
for calculating the molecular property distributions, in terms of the kinetic rate constants
and the concentration of the reactants. The most powerful approach for modelling
polymerization kinetics is the detailed species balance method. Based on the classical
physical law of conservation of mass, one can derive an infinite set of differential or
algebraic difference equations, depending on the reactor type, for the different molecular
species (e.g., monomer(s), growing polymer chains, dead polymer chains) present in the
reaction mixture. The resulting system of species balance equations must be solved to
obtain information on the desired molecular property distributions.

4.3.1.2. Mesoscale Physical Modelling

At the mesoscale, both interphase (e.g. thermodynamic equilibrium, kinetics) and


intraphase (e.g. heat and mass transfer between different phases) phenomena as well as
micromixing play an important role and can influence the polymer molecular and
morphological properties. Significant contributions have been made by Ray and his co-
workers toward a better understanding of interphase and intraphase transport phenomena in
heterogeneous Ziegler-Natta olefin polymerizations (Floyd et al., 1987, 1988); sorption
phenomena, and particle morphology development (Hutchinson and Ray, 1990, 1991). It
should be noted that modelling of these mesoscale physical phenomena is strongly
dependent on the specific process design and operation.
134 Chapter 4: Slurry-loop Reactor Modelling

4.3.1.3. Macroscale Reactor Modelling

Finally, at the macroscale, one has to deal with the development of models
describing the macromixing phenomena in the reactor, the overall mass and energy
balances, particle population balances, the heat and mass transfer from the reactor as well
as the reactor dynamics and control.

The whole process model integrates information coming from the solution of
different sub-models at all different space/ time scales (e.g., kinetic, particle, PSD and
reactor). It is apparent that a comprehensive mathematical model capable of predicting the
molecular and morphological developments in a polymer reactor in terms of the process
operating conditions should include appropriate model representations of all chemical and
physical phenomena occurring at the three scales. Guidelines for the development of a
mathematical model are thoroughly analysed in the work of Kiparissides, 1996.

4.3.2. Slurry-loop Reactor Modelling Literature Review

The slurry-phase catalytic olefin polymerization has been extensively studied in the
last two decades. Ha et al., 2001 investigated the operation of a semi-batch slurry ethylene
polymerization reactor by employing a multi-scale, multi-phase model. Their model,
describing the gas phase, the liquid phase (including gas bubbles), and the solid polymer
particles inside the reactor was developed to assess the effect of the hydrogen
concentration and the initial size of the catalyst particles on the polymerization reaction. It
was reported that the average molecular weight of the polymer material can be influenced
by the initial size of the catalyst particles because of the existence of mass transfer
limitations, an effect that may become very important when catalyst activities are high.
Khare et al., 2002 developed a generalized polymerization process simulator to describe
the steady-state operation of a slurry-phase industrial plant. They studied the effects of
operating conditions on the polyolefin production rate and molecular properties (i.e.,
average molecular weights and polydispersity index). The model was capable of simulating
the transient operation of the polymerization plant, and, thus, it was possible to calculate
the optimal operating conditions to increase the polyolefin production rate. Fontes and
Mendes, 2005 developed a model for a continuous slurry ethylene-butene co-
polymerization reactor that was able to predict the production rates and the number and
Chapter 4: Slurry-loop Reactor Modelling 135

weight molecular weight averages of the final polymer. The kinetic scheme employed
comprises of initiation, propagation, first order deactivation, hydrogen transfer, ethylene
transfer, transfer to co-catalyst and β-hydride elimination, assuming two different catalyst
site types taking part in polymerization. The method of moments is employed for the
calculation of molecular developments. A non-uniform solid phase is considered. The
multi-grain model (i.e., single particle model) is employed for the calculation of the radial
distribution of chemical species such as the living and dead polymer chains in o a single
polymer particle. Neto et al., 2005 developed a mathematical model to simulate the
dynamic behaviour of ethylene/1-butene industrial slurry-phase polymerization reactor.
Their model is able to describe the dynamic evolution of the molecular weight averages,
co-monomer content, particle size averages, melt index, and density of the final polymer
resin.

Despite the profound importance of the modelling of olefin polymerization loop


reactors, there is only limited experimental and theoretical information in the open
literature. Uvarov and Tsevetkova, 1974 developed a mathematical model for the
polymerization in a loop reactor. The model assumes that the reactor behaves as a CSTR
and the main objective is to model and control the polymer yield. A simplified kinetic and
thermodynamic description is employed and the concentrations of species of interest and
temperature are calculated as functions of time. Lepski and Inkov, 1977 modelled the loop
reactor as being composed of two main perfectly mixed regions. The main body of the
reactor is interconnected to a smaller volume reactor, corresponding to settling legs. Both
regions are flooded all the time. The purpose of the settling leg is to increase the polymer
concentration in the outlet mixture of the reactor. Differential equations describing the
mass and energy balances, corresponding to the two regions, are solved to calculate the
influence of operational variables such as reactor temperature, catalyst activity and
monomer conversion. Ferrero and Chiovetta, 1990 proposed a preliminary way of
designing a loop reactor for bulk propylene polymerization. Reactor dimension estimations
are based on polymerization data obtained from catalyst performance tests. Reactor length
and diameter are then specified according to the required heat production rate and the
desired productivity levels. The influence of operational parameters such as solid volume
fraction, polymer yield and reactor production was also studied.

All previous workers (i.e., Uvarov and Tsevetkova, 1974, Lepski and Inkov, 1977,
Ferrero and Chiovetta, 1990) have employed the continuous stirred-tank reactor (CSTR)
136 Chapter 4: Slurry-loop Reactor Modelling

model to describe the loop reactor. On the other hand, Weimin et al., 1991 described a
steady state mathematical model for polypropylene production in a loop reactor, employing
plug flow tubular equations. The influence of different operational parameters was
analyzed through simulation. Since only high ratios of recycle ratios were considered, no
significant concentration and temperature gradients along the length of the reactor were
found. Zacca and Ray, 1993 assumed that the flow behaviour in the loop reactor can be
approximated by a series of tubular sections interconnected by perfectly mixed inlet and
outlet zones, allowing the investigation of recycle rate, axial dispersion and heat transfer.
Reginato et al., 2003 simulated the operation of a loop reactor by assuming a non-ideal
CSTR behaviour. In their work, they introduced a constant discharge parameter accounting
for the difference in the concentrations of polymer solids in the reactor and the product
withdrawal stream.

4.4. Model Development

4.4.1. Model Development Overview

In the present study, a comprehensive dynamic mathematical model is developed for


the simulation of the dynamic operation of an industrial slurry-loop olefin catalytic
polymerization process, consisting of two loop reactors in series. According to the
proposed modelling approach, each loop (i.e., consisting of the reactor tube and the settling
legs) is modelled as an ideal CSTR (loop reactor model) in series with a semi-continuous
process (settling legs operation modelling). In particular, the polymer product is removed
from the reactor in a non-continuous mode through the dynamic operation of the reactor’s
settling legs. In fact, the product withdrawal rate from a loop reactor is controlled by the
pressure feedback controller that keeps the polymerization pressure in the loop reactor at a
specified value. To account for the non-continuous product removal rate and calculate the
actual concentration of solids in the product stream, a detailed model for the dynamic
operation of the settling legs is developed. Dynamic macroscopic mass and energy
balances are derived for each loop reactor in the series to predict the time variation of the
concentrations of the various molecular species as well as the reactor and jacket
temperatures in the two loop reactors. It should be noted that for energy balances,
regarding coolant jacket, plug flow reactor (i.e., PFR) approach was followed in order to
Chapter 4: Slurry-loop Reactor Modelling 137

calculate the temperature along the tube. A schematic representation of the proposed model
is depicted in Figure 4.3. The polymer molecular properties (i.e., number- and weight-
average molecular weights and molecular weight distribution) are determined by
employing a generalized multi-site, Ziegler-Natta (i.e., Z-N) kinetic scheme in conjunction
with the well-known method of moments. The Sanchez-Lacombe Equation of State (i.e.,
S-L EOS) is applied to calculate the thermodynamic equilibrium concentrations of the
various molecular species (i.e., monomer(s) diluent, hydrogen, etc) in the various phases
(i.e., solids, liquid and gas) present in the reaction mixture. Thermal effect related to
monomers adsorption in the diluents as well as Heat transfer resistances between the
phases are assumed negligible. Finally, a rheological model based on the reptation and
Rouse relaxation theories is employed to calculate the rheological behaviour (i.e., melt
viscosity versus frequency) of polyolefins produced in the industrial slurry-phase loop
reactor series in terms of the calculated MWDs. Numerical simulations are carried out to
investigate the effects of the process operating conditions (i.e., reactor temperature and
pressure, inflow rates of catalyst and monomers, feed composition, etc.) on the dynamic
behaviour of the cascade-loop reactor series (i.e., startup and grade transition) in terms of
plant operability and molecular and rheological properties of polyolefins. It is shown that
the proposed comprehensive model is capable of simulating the dynamic operation of an
industrial slurry-phase cascade-loop reactor series under different plant operating policies
(i.e., startup, grade transition, etc.).
Figure 4.3: Schematic representation of loop reactor modelling approach.
Chapter 4: Slurry-loop Reactor Modelling 139

4.4.2. Solution Procedure

The modelling approach followed is depicted in a logic diagram form in Figure 4.4.
The simulation model incorporates all different models presented in the precious sections
under FORTRAN environment. It should be noted that the differential equations solved at
every time, corresponding to the ones coming from mass balances for species of interest
and from energy balances, regarding reactor temperature, are solved using the ordinary
differential equation solver routine, divpag, from the numerical library of IMSL.

Figure 4.4: Process modelling logic diagram.


140 Chapter 4: Slurry-loop Reactor Modelling

4.4.3. CSTR Modelling Approach

The actual flow sheet of the process modelled is shown in Figure 4.3. According to
our modelling approach, each loop reactor of the whole process is simulated as a CSTR. A
loop reactor can be modelled as a recycle reactor (Figure 4.5a.). The recycle ratio,
Rrecycle , of a loop reactor, defined as the ration of volumetric flow rates between the

recycle stream, Qrec , and the outlet stream, Qout :

Qrec
Rrecycle 
Qout

can be made to vary from zero to infinity. Reflection suggests that as the recycle ratio is
raised, the behaviour shifts from plug flow ( Rrecycle  0 , Plug Flow Reactor, PFR) to

mixed flow ( Rrecycle   , Continuous Stirred Tank Reactor, CSTR) (Figure 4.5b.). Thus,
recycling provides a means for obtaining various degrees of backmixing with a PFR
(Levenspiel, 1972). A loop reactor must converge to the stirred tank reactor (CSTR) when
the recycle ration tends to extremely large values. In practice, there is no general way to
determine this threshold recycle ratio where the loop reactor begins to behave as a CSTR.
The recycle ratio of a loop reactor can be defined as the ratio of volumetric flow rates
between the recycle stream and the outlet stream:

Volume of fluid returned to the reactor entrance Qrecycle


Rrecycle  
volume leaving the system Qout

According to Zacca and Ray, 1993, at very high recycle ratios (i.e., Rrecycle  150 )
the reactor is fairly uniform presenting almost no concentration and temperature gradients
along the axial distance. They propose, as a ‘rule of thumb’ that for recycle ratios above
30, the loop reactor behaves as a CSTR.

For industrial loop reactor plants operating conditions, a typical recycle ratio can be
calculated:
Chapter 4: Slurry-loop Reactor Modelling 141

Qrecycle Average circulation velocity  cross  sec tion 2.5m3 / s


Rrecycle     500  30
Qout output volumetric flow rates 0.005m3 / s

As can be easily understood, the assumption that industrial loop reactors behave as
CSTRs holds true, since the recycle ratio remains at large values. Thus, the whole process
(2 loop reactors in series) can be safely simulated as a two CSTR train.

Figure 4.5: The recycle reactor.

4.4.4. Kinetic Modelling

To calculate the monomers consumption rates and molecular weight developments


of polyolefins produced in the slurry-phase, cascade-loop reactor series, a simplified four-
site kinetic model for the ethylene-1-hexene co-polymerization is employed. The
simplified kinetic scheme comprises of elementary reactions including site activation by
co-catalyst, chain initiation and propagation, chain transfer to hydrogen, to monomer and
142 Chapter 4: Slurry-loop Reactor Modelling

spontaneous chain transfer as well as spontaneous deactivation (see Table 3.5). The
methodology followed in order to derive a simplified kinetic scheme able to simulate the
molecular developments in a slurry-loop reactor is thoroughly analyzed in section 3.7
Slurry-loop reactor kinetic scheme.

4.4.5. Dynamic Mass Balances

According to the proposed modelling approach, each loop (i.e., consisting of the
reactor tube and the settling legs) is modelled as an ideal CSTR (in terms of perfect mixing
conditions) in series with a semi-continuous process. In other words, the whole process
consists of a semi-continuous (in terms of reactor outflow) CSTR, where reactor outflow
may have different concentration (enhanced in solids) compared to the one in the reactor.

th
According to the modelling approach, j reactor of volume VR , j , in series may

have a continuous inflow, regarding all species of interest, i , Fin,i , j . Moreover, species of

interest may also have a consumption/ production rate term, Ri, j . This term corresponding
to the consumption/ production rates, related to the kinetic scheme employed, are
calculated in 3.5 Slurry-loop reactor kinetic scheme section (see Eq. 3.32- Eq. 3.46). It
should be noted that, the molecular species of the reactor’s bulk phase that are taken into
account are ethylene, 1-hexene, iso-butane, hydrogen, co-catalyst, vacant catalyst sites,
potential catalyst sites, zero, first and second ‘live’ and ‘dead’ moments for each catalyst

site and for each monomer. For reactor pressure values, PR, j , not exceeding the maximum

allowable value (i.e., PR, j ,max ):

dCi , j Fin,i, j
  Ri , j PR, j  PR, j ,max (4.1)
dt VR, j ,

As can be seen in Eq. 4.1, no outflow term is taking part in the mass balances
equation. The outflow depends on the pressure of each reactor and the maximum allowable
pressure value. After accomplishing the appropriate thermodynamic calculations, the
outflow stream of the two reactors, as well as the corresponding inflow from first to second
Chapter 4: Slurry-loop Reactor Modelling 143

reactor was directly linked with reactor pressure. Thus, every time in which reactor

pressure, PR, j , exceeds its maximum allowable pressure, PR, j ,max , a product volume
proportional to the settling legs volume is automatically discharged from the loop reactor
(i.e., settling leg firing). The reduced values of compositions of all species of interest (in
terms of mol/ m3), as reactor volume remains the same, also reduces reactor pressure value.
Thus, the model follows a ‘zig zag’ pattern regarding the overall reactor pressure, as fresh
monomers and diluent fed into the reactor causing a new pressure increase. The
composition of the outflow volume can be either the same as in the reactor either enhanced
in solids, due to the settling legs operation. Moreover, first reactor’s outflow is fed to the
second reactor, increasing second reactor’s species of interest concentrations. Reactor
concentrations are then corrected as can be seen in Eqs. 4.2, 4.2 and 4.4.

First Reactor’s Firing Corrected Concentrations

For both ‘liquid’ and ‘solid’ species

Ci',1 
 
Ci,1  VR,1  VL,1  Cout ,i,1  VL,1
(4.2)
, PR,1  PR,1,max
VR,1  VL,1

and,

Ci',2 
 
Ci,2  VR,2  VL,2  Cout ,i,1  VL,1
(4.3)
VR,2  VL,2

Second Reactor’s Firing Corrected Concentrations

For both ‘liquid’ and ‘solid’ species

Ci',2 
 
Ci,2  VR,2  VL,2  Cout ,i,2  VL,2
, PR,2  PR,2,max (4.4)
VR,2  VL,2
144 Chapter 4: Slurry-loop Reactor Modelling

VL, j represents the settling leg volume (i.e., the bulk volume that is discharged every time

th
there is a leg firing). Cout ,i, j represents the outflow concentration of the i species of
th
interest from the j reactor. In order to calculate this term, all species of interest,
regarding this part of calculations, are separated into ‘liquid’ or ‘solid’ phase species. In
this way, a partial, reduced concentration, regarding liquid or solid phase can be calculated
for all species, assuming that the two phases are perfectly separated. Thus, monomers,
diluent, hydrogen and co-catalyst are treated as ‘liquid’ species, while catalyst active sites
and polymer moments are treated as ‘solid’ species. Thus, the reduced concentrations of
both ‘liquid’ and ‘solid’ species can be calculated in terms of solids volume fraction in the

reactor,  s, j .

‘Liquid’ species


Cred , j  Ci, j  1   s, j  (4.5)

‘Solid’ species

Ci, j
Cred , j  (4.6)
 s, j

According to all previous calculations, assuming that solids volume fraction in the

settling leg (i.e., in the reactor outflow), out , j , can be calculated, outflow concentration

th th
of the i species of interest from the j reactor, Cout ,i , j , can be therefore calculated.
The final equation forms for both reactors and for both ‘liquid’ and ‘solid’ species, are
presented in Eq. 4.7-4.12.

It is noted that the solids volume fraction in the settling leg, out , j , is calculated
through settling leg modelling.
Chapter 4: Slurry-loop Reactor Modelling 145

First Reactor’s Firing Corrected Concentrations

For ‘liquid’ species

Ci',1 
  
Ci,1  VR,1  VL,1  Cred ,i,1  VL,1  1  out ,1 
, PR,1  PR,1,max (4.7)
VR,1  VL,1

and,

Ci',2 
  
Ci,2  VR,2  VL,2  Cred ,i,1  VL,1  1  out ,1  (4.8)
VR,2  VL,2

For ‘solid’ species

Ci',1 
 
Ci,1  VR,1  VL,1  Cred ,i,1  VL,1  out ,1
, PR,1  PR,1,max (4.9)
VR,1  VL,1

and,

Ci',2 
 
Ci,2  VR,2  VL,2  Cred ,i,1  VL,1  out ,1
(4.10)
VR,2  VL,2

Second Reactor’s Firing Corrected Concentrations

For ‘liquid’ species

Ci',2 
  
Ci,2  VR,2  VL,2  Cred ,i,2  VL,2  1  out ,2 
, PR,2  PR,2,max (4.11)
VR,2  VL,2

For ‘solid’ species

Ci',2 
 
Ci,2  VR,2  VL,2  Cred ,i,2  VL,2  out ,2
, PR,1  PR,1,max (4.12)
VR,2  VL,2
146 Chapter 4: Slurry-loop Reactor Modelling

4.4.6. Settling Leg Modelling

Product withdrawal is achieved through the settling legs operation. Settling legs play
an important role in the whole loop reactor process. Their operation is crucial as they affect
reactor concentration, which in turn, affect reactor pressure, which again affects settling
leg operation. The quantification of the difference between the concentration inside the
reactor and the one at the reactor output is achieved through the settling leg modelling. To
calculate the solids concentration in the settling legs of a loop reactor as well as their
periodic operation (i.e., discrete times of valve openings), a comprehensive model,
describing the dynamic operation of the settling legs, was developed (Touloupides et al.,
2010a). According to the proposed approach, each settling leg in a loop reactor is assumed
to comprise two zones (see Figure 4.6), namely:

1) A well packed polymer zone at the bottom of the settling leg. The void volume
fraction in the packed polymer zone is assumed to be filled with a liquid phase
of the same composition as that in the loop reactor).

2) An upper zone at the top of the settling legs containing a slurry-phase of similar
composition with that in the loop reactor.
Operating Status

Figure 4.6: Schematic representation of the settling legs employed for product withdrawal.
Chapter 4: Slurry-loop Reactor Modelling 147

Thus, when the pressure in the loop reactor reaches a specified set-point value (i.e.,
maximum allowable operating pressure), the discharge valve of a specific settling leg
opens automatically (i.e., settling leg firing) so that the product content of the leg is
discharged. This also results in a decrease of the reactor pressure. Note that, during a given
reactor period (e.g., startup, steady-state, etc.), a certain number of settling legs can be in
operation (e.g., 1-6).

The time period between the valve openings of two consecutive legs is called firing
period, t f , and corresponds to two consecutive maxima in the reactor pressure.

Assuming that N legs are in operation, the total time available for polymer solids to be
settled in a specific leg, will be equal to the summation of the last N firing periods. The
valve opening cycling period, ts , is defined as the time period between two consecutive
openings of the same valve:

ts   t f (4.13)
N

In this research study, it is assumed that the volume fraction of solids in the packed
zone depend on the solids concentration in the reactor-phase as well as on the valve
opening cycling period,. Accordingly, the solids volume fraction, f p , in the packed

polymer zone of a settling leg will be given by:

f p  Sr  s ts (4.14)

where Sr is a packing coefficient related with the solids settling rate in a leg and  s is the

solids volume fraction in the reactor slurry-phase. Packing coefficient, Sr , depends on the
viscocity and the geometric characteristics of the reactor studied. The value of this
parameter was defined based on industrial plant data, provided by our industrial partner. It
should be noted that the value of this parameter practically remains stable as recycling
velocity also remains stable. Thus, for every loop reactor operating settling legs, a different
but stable value can be defined. It should be noted that the number of settling legs in
operation strongly affects the solids concentration in the product withdrawal stream. As the
number of settling legs in operation increases, the valve opening cycling period increases,
148 Chapter 4: Slurry-loop Reactor Modelling

giving more time for polymer settling and, leading to higher values of solids volume
fraction in the product outflow stream (or higher values of enhancement factor). This in
turn results in an increase of solids concentration in the product outflow stream (as more
time available for solids’ settling is given).

Similarly, the solids volume fraction in the product outflow stream,  out , will be
given by:

 out  f p  p  1  f p   s (4.15)

where  p is the volume fraction of the packed polymer particles in a settling leg. For this
research study, a face-centered cubic packing arrangement of spherical polymer particles
of equal size was assumed (see Figure 4.7). Thus, after simple geometric calculations one
can define that 0.7405 of the space is occupied by spheres and 0.2595 is unoccupied space.

Thus,  p is equal to 0.7405.

Figure 4.7: Face-centered cubic packing arrangement.

To account for the difference in the solids concentrations in the reactor and the
product stream, a solids enhancement factor, ef , was introduced.

e f   out  s (4.16)

Contrary to previous studies (Reginato et al., 2003), it was found that the value of
the enhancement factor did not remain constant during reactor startup, exhibiting an
Chapter 4: Slurry-loop Reactor Modelling 149

asymptotic decrease to its final steady-state value. Comparative results, regarding settling
legs’ operation are given in Chapter 5: Results and Discussion.

4.4.7. Dynamic Energy Balances

During the operation of the slurry-phase olefin catalytic polymerization cascade-loop


reactor series, the generated polymerization heat is transferred from the polymer solids to
the liquid phase, and, then through the reactor’s wall to the coolant flowing inside the
reactor’s external jacket. Thus, the following dynamic energy balances can be derived for
the reaction mixture in the two loop reactors.

4.4.7.1. Energy Balance for Reactor Mixture

Figure 4.8: Energy Balance Calculations.

Heat removal from the reactor is achieved via heat exchangers. Because of the good
mixing conditions in the loop reactor, it can be assumed that reactor temperature is, at any
time, the same in the reactor (but not constant). In other words, we simulate the loop
reactor as a CSTR.

All possible terms are taken into account for the calculation of energy balance,
regarding reactor mixture, including heat input, heat outpout, heat from reaction, heat
transferred to reactor jacket, heat loss to environment and heat provided by the pump
power (Figure 4.8).
150 Chapter 4: Slurry-loop Reactor Modelling

Heat accumulation = Heat input


+Heat from reaction
-Heat output
(4.17)
-Heat transfered to reactor jacket
-Heat loss to environment
+Pump power

dTr , j N
VR , j  R , j C p , R , j    qin ,i , j i , j C p.i , j Tin ,i , j  Tref 
dt i 1

  H r , j  R p , j

 qout , j  R , j C p , R , j TR , j  Tref  (4.18)


 qc , j c , j C p ,c , j Tc ,out , j  Tc ,in , j 

 U e , j Ae , j TR , j  Te 

 Pp

q in,i, j , q out, j and q c, j represent the inlet and outlet volumetric flowrates in the ‘ j ’ reactor

of the series and the coolant flowrate in the ‘ j ’ reactor jacket, respectively. Tref and

Tc,out, j denote the reference temperature and the jacket outlet temperature, respectively.

Finally, the term U e, jA e, j Tr, j  Te   accounts for the heat losses from the free reactor

sections (i.e., non-jacketed parts) to the environment.

It should be noted that for the establishment of the energy balances a similar to the
mass balance approach has been followed. Thus, heat input and heat output terms take part
in the equations only when volume exits the reactor (i.e., only when pressure reaches
maximum pressure limit ant settling leg volume exits on/ and enters reactor, as explained
for mass balances).
Chapter 4: Slurry-loop Reactor Modelling 151

4.4.7.2. Energy Balance for Coolant Jacket

Similarly, one can derive the following pseudo-steady-state energy balances for the
coolant flowing in the reactor jackets. The temperature in the coolant cannot be assumed
constant through the coolant tube length. Thus, contrary to reactor mixture energy balance
calculation, the coolant jacket is simulated as a PFR.

dTc, j
dz
 
 d r, jU j / q c, jc, jCpc, j Tr, j  Tc, j 
(4.19)
0  z  Lc, j , Tc, j  Tc,in, j at z0

d r, j and U j are the inside tube diameter and the overall heat transfer coefficient from the

reaction mixture to the coolant in the ‘ j ’ reactor, respectively. By integrating Eq. 4.19, one

can easily show that the axial variation of the coolant temperature in the jacket will be

given by the following equation (Kiparissides et al., 2005):

dTc   dr U r dT   dr U r

dVc ' mc  Cpc  Ac '

 Tr  Tc   c 
dz ' m c  Cpc
Tr  Tc  

dTc  T T 
 a  dz '   ln  r c   az 
Tr  Tc   Tr ,0  Tc,0


 (4.20)
Tr  Tc
Tr ,0  Tc,0
 
 e  a z  Tr  Tc  Tr ,0  Tc,0 e  a  z 

 
Tc  Tr  Tr ,0  Tc,0 e  a  z

where:

  dr U r
a (4.21)
m c  Cpc
152 Chapter 4: Slurry-loop Reactor Modelling

Tr ,0  Tr at z=0 (4.22)

Tc,0  Tc at z=0 (4.23)

where Tc ,in , j (t , 0) and Tc, j (t , z ) denote the jacket inlet temperature and the jacket

temperature at a distance ‘ z ’ from the jacket inlet at time, respectively.

4.4.7.3. Overall Heat Transfer Coefficient

The overall heat transfer coefficient is calculated by the following equation (Brodkey
and Hersey, 1988).

1
U  Rf (4.24)
RR  Rc  Rw

R f denotes heat transfer resistances because of reactor fouling. The resistances accounting

for the reaction mixture, or the coolant jacket and the reactor wall are calculated as:

dc
RR  (4.25)
d R  hR

1
Rc  (4.26)
hc

d 
d w  log  w 
Rw   dR  (4.27)
2  w

Thus, Eq. 4.24 becomes:


Chapter 4: Slurry-loop Reactor Modelling 153

d 
d w log  w 
1 d
 c   dR   1  R (4.28)
f
U d c hR 2k w hc

where hR and hc denote the inside and outside heat transfer coefficients (i.e., reactor and
coolant jacket heat transfer coefficients), respectively. For high Reynolds Number values
(i.e., turbulent flow and Re > 10,000), the inside and outside partial heat transfer
coefficient is calculated using the equations based on the work of Sieder and Tate (Coulson
and Richarson, 1999, Geankoplis, 1993):

0.14
hd  
Nu   0.027  Re0.8  Pr1/ 3   b  (4.29)
k  w 

where b and  w are the reaction mixture viscosities calculated at the bulk reaction (or
coolant) temperature and near the reactor wall temperature, respectively.

It should be mentioned that Eq. 4.29 holds for Re  10000 , 0.7  Pr  16000 and
length to diameter ratio, Leq / Deq  60 . It should also be noted that Eq. 4.29 provides very

good results when the term  Re Pr Deq  / Leq takes values higher than 10. If this number

take numbers lower than 10, the outside partial heat transfer coefficient is underestimated.

For low values of Reynolds number (i.e., laminar flow and Re  2100 ) the outside
partial heat transfer coefficient is calculated using a different equation (Bird et. al., 2002):

0.33 0.14
hd  d  b 
Nu   1.86  Re Pr    (4.30)
  l  w 

The dimensionless Reynolds and Prandlt numbers appearing in Eq. 4.29 and Eq.
4.30 are defined as follows:

d   V
Re  (4.31)

Cp  
Pr  (4.32)

154 Chapter 4: Slurry-loop Reactor Modelling

4.4.8. Reactor Control System

For the stable as well as dynamic operation of the simulated industrial slurry-phase
cascade-loop reactor series, the simulation of the various PID feedback controllers, shown
in Figure 4.9, was included. In particular, the dynamic process model included the
simulation of the two temperature controllers, one ethylene mass fraction controller, two
polymer solids concentration controllers and two reactor pressure controllers.

More specifically, five different control actions are taken:

 The catalyst inflow is manipulated through a feedback controller using ethylene


off gas composition measurements (see Figure 4.9, [1]).
 Iso-butane inflow is manipulated by controlling the solids concentration in the
reactor ([2]).
 Settling leg operation (reactor outflow rate) is manipulated by controlling the
reactor pressure ([3]). Settling leg operation is well deployed in 4.4.6. Settling
Leg Modelling section.
 During reactor startup the reactor is heated until a desired operating temperature
is reached. For this purpose a steam heat exchanger is typically used. After the
operating temperature is reached, a cascade split-range control system is
utilized. The control system consists of a master PID and a slave PID controller
which define the ratio of the recycling cooling water flows that will bypass the
heat exchanger ([4]). Master controller compares reactor temperature to a
setpoint temperature value and produces a signal. Slave controller compares
master controller’s signal to coolant inflow temperature. Slave controller’s
signal defines the ratio of the recycling cooling water flows that will bypass the
heat exchanger, while coolant inflow rate remains stable.
 The cooling water inflow coming from the cooling tower is controlled by
measuring the outflow stream temperature of the heat exchanger ([5]).
4 4
Tsp Tsp

TC TC From cooling TC TC From cooling


tower tower

5 5

TC TC

CC

CC CC

2 2

Iso-butane Iso-butane
Ethylene Ethylene
1-Hexene
Hydrogen
Hydrogen
Settling legs Settling legs

PC PC

3 3
Catalyst
Co-Catalyst

Diluent
Polyethylene

Figure 4.9: Slurry-loop reactor control system.


156 Chapter 4: Slurry-loop Reactor Modelling

4.4.8.1. PID Controller

The three basic feedback control modes that are employed are proportional (P),
integral (I) and derivative control (D). The controller compares the measured value to the
setpoint and takes the appropriate corrective action by sending an output signal to the
control valve. The ideal continuous (analog) proportional-integral-derivative (i.e., PID)
controller operation can be described by the following equation (Seborg, 1989):

 1 t de(t ) 
p (t )  p  K c  e(t )   e(t ') dt '   D  (4.33)
 I 0 dt 

To convert the ideal continuous (analog) PID controller expression (Eq. 4.33) to its
digital equivalent, the following finite difference approximations can be followed.

Finite Difference Approximation

According to this approximation, integral and differential terms become:

t n
 e(t ')dt '   ek t (4.34)
0 k 1

de(t ) en  en 1
 (4.35)
dt t

The digital PID controller equation can be written in two ways, the position form and
velocity form. By substituting Eqs. 4.34 and Eq. 4.35 into Eq. 4.33, the position form of
the digital PID control algorithm can be obtained:

 t n  
pn  p  K c en   ek  D (en  en 1)  (4.36)
  I k 1 t 
Chapter 4: Slurry-loop Reactor Modelling 157

or,

 t n  
pn 1  p  K c  en 1   ek  D (en 1  en  2 )  (4.37)
  I k 1 t 

Subtracting Eq. 4.37 from Eq. 4.36 we obtain the velocity form of the PID
controller:

 t  
Pn  pn  pn 1  K c  (en  en 1 )  en  D (en  2en 1  en  2 )  (4.38)
 I t 

Trapezoidal approximation

Another variation of the digital PID controller is based on the more accurate
trapezoidal approximation for the integral:

t n  e e
k k 1  t
 e(t ')dt '    2
 (4.39)
0 k 1  

After this expression is substituted into Eq. 4.38 the velocity form of the control law
takes the form employed in this research project.

 t  ek  ek 1   D 
Pn  pn  pn 1  K c  (en  en 1 )    (en  2en 1  en  2 )  (4.40)
 I  2  t 

4.4.8.2. PID Controller Tuning

After the control system is installed, the controller settlings must be adjusted until
the control system performance is considered to be satisfactory. This activity is referred to
as controller tuning or field tuning of the controller. Controller tuning is usually done by a
trial and error procedure. A typical approach can be summarized as follows:
158 Chapter 4: Slurry-loop Reactor Modelling

1) Eliminate integral and derivative action by setting  D at its minimum value and
 I at its maximum value.

2) Set K c at a low value.

3) Increase the controller gain K c by small increments until continuous cycling


occurs after a small setpoint or load change. The term ‘continuous cycling’
refers to a sustained oscillation with constant amplitude.

4) Reduce K c by a factor of two.

5) Decrease  I in small increments until continuous cycling occurs again. Set  I


equal to three times this value.

6) Increase  D until continuous cycling occurs. Set  D equal to one-third of this


value.

4.4.9. Thermodynamic Considerations

In general, in reactor modelling, the termodynamic properties of the system have to


be accurately calculated as they directly affect kinetic calculations, which in turn affect
polymer molecular properties. Moreover, for the case of the reactor process studied,
thermodynamic calculations also play a significant role regarding pressure calculation as
reactors’ outflow depends on the value of reactor pressure. Thus, the termodynamic
properties of the system must be readily available.

In the reactor, two phases co-exist, namely a liquid phase (including isobutene,
ethylene, 1-hexene and hydrogen) and a solid-polymer phase (including polymer and
sorbed quantities of iso-butane, ethylene, 1-hexene and hydrogen). To calculate the species
consumption/ production rates, the concentration of the sorbed species in the amorphous
polymer phase must be known. Moreover, reactor pressure must be accurately calculated
because this is what defines when settling leg volume exits the reactor, decreasing the
pressure. For the thermodynamic calculations (i.e., both solubility and pressure
calculations) the Sanchez-Lacombe Equation of State is employed.
Chapter 4: Slurry-loop Reactor Modelling 159

4.4.9.1. Equilibrium Sorbed Monomer Concentration

In the reactor, two phases co-exist, namely a liquid phase (including isobutene,
ethylene, 1-hexene and hydrogen) and a solid-polymer phase (including polymer and
sorbed quantities of iso-butane, ethylene, 1-hexene and hydrogen). It should be noted that
the gas phase may be neglected due to its low fraction (tends to be equal to zero) in the
reactor phase. Moreover, the gas phase formation in the reactor is undesired due to
operating problems that the gas phase causes (i.e, trapped gas phase in the reactor top,
cavitation problems, etc.). For the calculation of the monomer(s) consumption rate(s) and
molecular weight properties of polyolefins in the slurry-phase loop reactor, the
concentrations of the sorbed monomers and other species in the amorphous polymer phase
should be known. In the heterogeneous catalytic olefin polymerization, the initially formed
polymer chains are amorphous. However, as the ‘dead’ polymer chains move away from
the polymerization loci (i.e., active catalyst sites) due to chain propagation and termination
reactions, crystallization of the amorphous polymer chains takes place. Thus, it can be
assumed that the produced polyolefin will consist of two phases, namely, an amorphous
polymer phase swollen with the sorbed reaction species and a crystalline one that acts as a
barrier to the reaction species transfer to the active catalyst sites (Hutchinson et al., 1992,
Kanellopoulos et al., 2004). Based on the mass balances, bulk concentrations in the reactor
are calculated. However, only the species which are sorbed in the amorphous polymer
phase take part in the polymerization reactions. Thus, their ‘active’ concentration is not the
same as in the bulk phase. Thus, sorbed concentrations have to be calculated.

4.4.9.2. Sanchez-Lacombe Equation of State

Hutchinson et al., 1992, by analyzing experimental solubility data on the sorption of


different low molecular weight penetrants in semi-crystalline PE, concluded that a Henry's
law correlation can be employed to calculate the monomer solubility in amorphous PE in
terms of the temperature. According to Henry’s law, at a constant temperature, the amount
of a given gas dissolved in a given type and volume of liquid is directly proportional to the
partial pressure of that gas in equilibrium with that liquid.

pi  k H  C (4.41)
160 Chapter 4: Slurry-loop Reactor Modelling

pi is the partial pressure of the solute in the gas above the solution, C is the

concentration of the solute and k H is a constant with the dimensions of pressure divided
by concentration. The constant, known as the Henry's law constant, depends on the solute,
the solvent and the temperature.

Based on solubility experimental measurements, Hutchinson, 1990, suggested a

correlation for Henry’s law constant, kH (mol/atm/l amorphous polymer), for


hydrocarbons, in terms of critical temperature:

2
log  k H   2.38  1.08  c 
T
(4.42)
 T

However, Henry's law will not be applicable to polymer–vapor systems involving


heavier penetrant vapors, high pressures or/ and low temperatures (Chen, 1993). Moreover,
in the presence of iso-butane, significant deviations from Henry’s law can be obtained in
the predicted ethylene and 1-hexene solubility values (see also 4.4.9.3. Sanchez-Lacombe
Parameter Estimation section). It is noted that, for this research project, Henry’s Law
approach is employed only for the case of hydrogen.

The Sanchez-Lacombe equation of state (i.e., S-L EOS) is employed to calculate the
solubilities of the various reaction species (i.e., ethylene, 1-hexene, isobutene) in semi-
crystalline polyolefins over a wide range of pressures and temperatures (Sanchez and
Lacombe, 1978, Orbey et al, 1998, Rodgers, 1993).

Sanchez and Lacombe derived a dimensionless equation of state for classical fluids
based on the Ising (lattice) fluid model (see Eq. 4.43). According to the lattice theory,
fluids are mixtures of molecules and holes, confined to sites in a lattice. The S-L EOS can
be applicable for polymer species as well as conventional components. It is noted that S-L
EOS can accurately predict molar volume, fugacity coefficients, heat capacities, and
departures for enthalpy, entropy, and Gibbs free energy.

 2  P  T[ln(1   )  (1  1/ f )  ]  0 (4.43)
Chapter 4: Slurry-loop Reactor Modelling 161

P ,  , T (i.e., the dimensionless pressure, mass density and temperature of a pure


substance, respectively), are defined as:

P  P / P* (4.44)

   /  * (4.45)

T  T / T * (4.46)

* * *
where P ,  and  are characteristic parameters for pure substances.

The number of sites (mers) occupied in the lattice, f , can be related to the
molecular weight of the pure component, MW, according to the following equation:

f  P*  MW /( R  T *   * ) (4.47)

Since f remains explicit in the reduced equation of state, a simple corresponding-


states principle is not, in general, satisfied. However, for a high molecular weight polymer
the value of f can be considered to be infinite.

Ethylene, 1-hexene, iso-butane and polyethylene are taken into consideration for the
thermodynamic calculations in the reactor. It should be noted that, the reactor bulk phase
almost consists of those four species. S-L EOS can calculate the existence of two different
phases, namely, the liquid and the solid phase, their weight fraction, their density as well as
the weight fraction of each component in each phase.

By employing S-L EOS we can calculate the mass fraction of monomers in the solid
phase. The concentration in the solid phase is calculated as:

frsolid ,i   solid
Csorbed , solid ,i  (4.48)
MWi
162 Chapter 4: Slurry-loop Reactor Modelling

4.4.9.3. S-L EOS Parameter Estimation

It is important to point out that in all theoretical calculations, the values of the
*
characteristic parameters of pure components P* , T * and  , appearing in the S-L EOS
were directly obtained from molecular dynamics simulations (Kanellopoulos et al, 2006).
*
The values P* , T * and  parameters used for ethylene, 1-hexene, iso-butane and LDPE
are depicted in Table 4.1.

In all cases, the binary interaction parameters, kij and ij , accounting for the

interactions between the sorbed molecules and the polymer chains were introduced to fit
the EOS-based predictions to available solubility measurements (Moore and Wanke 2001,
Novak et al., 2006, Parrish, 1981) for solubility and pressure calculations. The binary

interaction parameters, kij and ij , were estimated by fitting the S-L EOS to binary
equilibrium data and they are reported in Table 4.2. The binary interaction parameters for
the systems ethylene-iso-butane and hexene-iso-butane for which there is not experimental
data, are set equal to zero.

Table 4.1: Values of S-L EOS parameters.

species P* , bar T* , K  * , g/cm3

Ethylene 3395 283 0.680

1-hexene 3252 450 0.814

iso-butane 2877 398 0.720

HDPE 3576 673 0.887


Chapter 4: Slurry-loop Reactor Modelling 163

Table 4.2: S-L EOS binary interaction parameters.

Ethylene 1-hexene iso-butane PE

kij  1.91 kij  0 kij  0.0221


Ethylene
ij  0.22 ij  0 ij  0.15

kij  1.91 kij  0 kij  0.0161


1-hexene
ij  0.22 ij  0 ij  0

kij  0 kij  0 kij  0.03


iso-butane
ij  0 ij  0 ij  0

kij  0.0221 kij  0.0161 kij  0.03


PE
ij  0.15 ij  0 ij  0

In Figure 4.10-Figure 4.12, the experimental measured solubilities, at different


pressures and temperatures are compared with model predictions obtained by the S-L EOS.
It should be pointed out that the theoretically calculated solubilities were found to be in
excellent agreement with the corresponding experimentally measured values,
demonstrating the capability of the S-L EOS to predict the solubility of the various reaction
species of interest in semi-crystalline (i.e., ethylene-1-hexene) co-polymers.

In Figure 4.10, the experimentally measured ethylene solubilities in LLDPE (i.e.,


ethylene-1-hexene co-polymer) at different pressures and temperatures are compared with
model predictions obtained by the S-L EOS. It should be pointed out that the numerical
value of the interaction parameter, kij , depends on temperature due to the presence of the

1-hexene co-monomer in the co-polymer chains. The variation of the interaction


parameter, k ij , with respect to temperature is depicted in the inserted scheme of

Figure 4.10. As can been seen, the binary interaction parameter exhibits an exponential
decrease with temperature. In fact, at higher temperatures (e.g., T > 65 oC), the value of k ij

asymptotically approaches zero (i.e., the interactions between the polymer segments and
the sorbed molecules become negligible).
164 Chapter 4: Slurry-loop Reactor Modelling

In Figure 4.11 and Figure 4.12, experimental and predicted solubilities of 1-hexene
and iso-butane in LLDPE (ethylene-1-hexene) co-polymer grades are depicted. Contrary to
the ethylene-LLDPE binary system, the solubilities of 1-hexene and iso-butane in LLDPE
exhibit a strongly nonlinear behaviour with pressure. Clearly, the theoretical predictions
obtained by the S-L EOS are in excellent agreement with the experimentally measured
solubilities. It is important to point out that the numerical value of the binary interaction
parameter, k ij , for the (iso-butane-LLDPE) binary system is independent of temperature.

On the other hand, the binary interaction parameter for the (1-hexene-LLDPE) binary
system follows a linear decrease with temperature (see inserted scheme in Figure 4.11).
This means that, at higher polymerization temperatures, the molecular interactions between
the unlike molecules (i.e., polymer segments and 1-hexene) decrease.

-2
6.0x10
0.05
Predicted by S-L EOS
Binary Interaction Parameter, kij

Polynomial Approximation
Solubility, greth/gram LLDPE-1-Hexene

0.04

0.03

0.02

-2 0.01
4.0x10
0.00
-5 2
kij = 0.075 - 0.019*T + 1.125*10 T
-0.01
25 30 35 40 45 50 55 60 65 70
o
Temperature, C

-2
2.0x10 o
Moore and Wanke, 2001, T = 27.7 C
S-L EOS kij = 0.030
o
Moore and Wanke, 2001, T = 47.7 C
S-L EOS kij = 0.010
o
Moore and Wanke, 2001, T = 67.7 C
S-L EOS kij = 0.0

0.0
0 10 20 30 40
Pressure, bar

Figure 4.10: Solubility of ethylene in LLDPE-1-Hexene at different temperatures.


Chapter 4: Slurry-loop Reactor Modelling 165

1.0

Solubility, gr1-hexene/gram LLDPE-1-Hexene


o
Novak et al., 2006, T=70 C
S-L EOS kij=0.025
0.8 o
Novak et al., 2006, T=90 C
S-L EOS kij=0.016
o
Novak et al., 2006, T=150 C
0.6 S-L EOS kij=-0.033

0.4
Predicted by S-L EOS)
0.4

Binary interaction Parameter, kij


Polynomial Approximation
0.3

0.2

0.1
0.2
0.0
-3
kij = 0.48 - 3.5*10 T
-0.1
60 80 100 120 140 160
Temperature, K
0.0
0 2 4 6 8 10 12
Pressure, bar

Figure 4.11: Solubility of 1-hexene in LLDPE-1-Hexene at different temperatures.

0.3
Solubility, grisobutane/gram LLDPE-1-hexene

o
Parrish, 1981, T=74 C
S-L EOS kij=0.02
o
Parrish, 1981, T=82 C
S-L EOS kij=0.02
0.2
o
Parrish, 1981, T=93 C
S-L EOS kij=0.02

0.1

0.0
0 5 10 15
Pressure, bar

Figure 4.12: Solubility of 1-hexene in LLDPE-1-Hexene at different temperatures.


166 Chapter 4: Slurry-loop Reactor Modelling

4.4.9.4. Crystallinity Calculation

In semi-crystalline polymers, sorption takes place only in the amorphous polymer


phase and not in the crystalline phase (Figure 4.13). Thus, the concentration of the sorbed
species in the solid phase must be corrected in order to correspond to the concentration in
the amorphous phase of the polymer:

Csorbed , solid ,i
Csorbed , am,i  (4.49)
1  Wcrys

The degree of crystallinity, Wcrys , is defined as:

Volume of crystalline polymer


Wcrys  (4.50)
Total polymer volume

Figure 4.13: Semi-crystalline polymer structure.

As it is well known, the crystallinity decreases with the percentage of the


incorporated co-monomer. As the 1-hexene incorporation in the polymer increases,
crystallinity decreases. Based on the experimental work of Quijada et al., 1997, a linear
correlation between hexene incorporation and crystallinity can be extracted:

%Wcrys  4.1679  (% Hexene mol fraction in PE )  55.6376 (4.51)


Chapter 4: Slurry-loop Reactor Modelling 167

In Figure 4.14 the linear fit as well as the corresponding experimental data is shown.
Moreover, the effect of hexene incorporation on the crystallinity of the produced
polyethylene according to other workers on the field is depicted. It is clear that crystallinity
data exhibit a degree of dispersion. According to plant crystallinity measurements
(provided by our industrial partner), it was observed that number average molecular weight
of the polymer produced may also affect crystallinity. Based on the plant data, the
following correlation can be extracted, as a function of hexene incorporation as well as
number average molecular weight, M n .

M n
%Wcrys      (% Hexene mol fraction in PE )    exp( ) (4.52)

It should be noted that, Eq. 4.52 can be employed for the calculation of both
instantaneous (i.e., for fresh polymer produced) or cumulative crystallinity value (i.e, the
weighted sum of polymer crystallinity regarding all polymer mass in the reactor),
depended on the %Hexene mol fraction in PE which may refer to the instantaneous or
cumulative value of the co-monomer composition, respectively.

80
Quijada et al., 1997, DSC measurements
Quijada et al.,1997, Density Measurements
Ko et al., 1998, DSC measurements
quijada et al., 1995, DSC measurements
60 Quijada et al., 1995, DSC measurements
Villar et al., 2001, DSC measurements
Linear fit
%Crystallinity

40

20

0 5 10 15
Hexene fraction (% mol)

Figure 4.14: Literature review on the effect of 1-hexene incorporation on the crystallinity
of the produced polyethylene.
168 Chapter 4: Slurry-loop Reactor Modelling

4.4.9.5. Polymer density

The density of the produced polymer can be calculated as:

%Wcrys %Wcrys
 polymer    polymer ,cryst  (1  )   polymer , am (4.53)
100 100

For polyethylene, the corresponding values for  polymer ,cryst and  polymer , am at
25oC, are 0.997 kg/m3 and 0.854 kg/m3, respectively (Yiagopoulos et al., 2001).

As it is well known, polymer density is affected by temperature (Zoller and Walsh,


1995). In Figure 4.15, the effect of temperature on linear polyethylene specific volume is
presented. Density is the inverse of specific heat. For the temperature range of interest (i.e.,
50-100oC) it is clear that there is a linear correlation connecting temperature and density
(see Figure 4.16). Assuming that the same slope remains for the co-polymer ethylene-1-
hexene (at low co-polymer composition), the following correlation can be extracted for the
corrected value of polymer density,  polymer ,T at temperature T (oC):

 polymer ,T   polymer ,25o C  1  4.193  T  25   (4.54)

where  polymer ,25o C is the polymer density at 25oC as calculated from Eq. 4.53.
Chapter 4: Slurry-loop Reactor Modelling 169

1.45
0 MPa
1.40 20 MPa

1.35
Specific Volume, cm /g
3

1.30

1.25

1.20

1.15

1.10

1.05

1.00
0 50 100 150 200 250 300
o
Temperature, C

Figure 4.15: Effect of temperature on polymer specific volume.

0.96
Experimental values
Linear fit

0.95
3
Density, g/cm

0.94

0.93

0.92

0.91
20 30 40 50 60 70 80 90 100 110
o
Temperature, C

Figure 4.16: Effect of temperature on polyethylene density.


170 Chapter 4: Slurry-loop Reactor Modelling

4.5. Calculation of Physical Properties

4.5.1. Slurry Viscosity

The apparent viscosity of slurry is higher than the one of the diluent alone. The solid
particles behave as volumes where no velocity gradient occurs, increasing the velocity
gradient and the shear rate to the remaining distance to the reactor wall, and thus the
apparent viscosity. Different equations have been proposed to estimate to slurry viscosity
as a function of the solids content.

The viscosity of the reaction mixture can be calculated by the following correlations.

Thomas Viscosity Correlation

Thomas, 1965 suggested the following expression for the calculation of viscosity of
suspensions of uniform spherical particles, mix , in terms of the viscosity of the diluent,

liquid , and the solids volume fraction,  s (Eq. 4.55).

mix  liquid 1  2.5 s  10 2s  2.73103 exp 16.6  s   (4.55)

Barnea and Mizrahi Viscosity Correlation

Barnea and Mizrahi, 1976, have also suggested a correlation for the calculation of
viscosity of suspensions of spherical solid particles (Eq. 4.56).

 5 3 s 
mix  liquid exp  (4.56)
 1 s 

Gay, Nelson, Armstrong Viscosity Correlation

Gay, Nelson, Armstrong, 1969, have introduced the extra term,  max , describing the

maximum obtainable volume fraction solids for the system.  max is a tuning parameter,

depending on the suspension system studied (Eq. 4.57). Practically,  max expresses the

limit-value of solids volume fraction,  s , where viscosity tends to infinity values. For the

system studied (i.e., polyethylene particles suspension in i-butane), the value of  max may
range between 0.45-0.70.
Chapter 4: Slurry-loop Reactor Modelling 171

  s   s 
0.48

mix  
 liquid  exp 2.5      (4.57)
 
 max   s   
 max 

For the viscosity of the liquid phase it is assumed that equals to i-butane viscosity.
According to Noll, 2005, i-butane viscosity can be calculated in terms of temperature, T .

 7.3891  2582.6 
liquid   exp   (4.58)
 172.23  T 

In Figure 4.17, all the results of all the proposed correlations are compared, for a
typical reactor temperature, equal to 90oC. As can be seen, all the correlations are in good
agreement between them for the values of volume fraction from 0-0.3. For values of
volume fraction higher then 0.3, Gay, Nelson and Armstrong and Thomas correlation,
which are in agreement between them, predict higher values of viscosity comparing to the
Barnea and Mizrahi correlation. In the present research project, Thomas correlation was
employed for the results presented in Chapter 5: Results and Discussion.

3.0
Thomas, 1965
Gay, Nelson and Armstrong, 1969
2.5 (max. volume fraction=0.65 )
Barnea and Mizrahi, 1973

o
2.0 Temperature=90 C
Viscosity, cP

1.5

1.0

0.5

0.0
0.0 0.2 0.4 0.6
Volume fraction

Figure 4.17: Effect of solids volume fraction on reactor-phase viscosity.


172 Chapter 4: Slurry-loop Reactor Modelling

4.5.2. Thermal Conductivity

4.5.2.1. Thermal Conductivity of Gases

Roy and Thodos estimation technique is employed for the calculation of thermal
conductivity of gases (i.e., ethylene and 1-hexene). According to the methodology
employed, a reduced thermal conductivity may be expressed as:

r     (4.59)

1
 T MW 3  6
  210  c  (4.60)
 P4 
 c 

where  is the reduced, inverse thermal conductivity (W/(m.K)-1), while MW is the


molecular weight (g/mol). Tc (K) and Pc (bar) denote the critical temperature and critical

pressure, respectively. The reduced thermal conductivity, r , can be separated into two
parts. The first, attributed only to translational energy, was obtained from a curve fit of the
data of rare gases (Roy, 1967), varies only with the reduced temperature, Tr . The second
term corresponds to the contribution of rotational and vibrational interchange, relating to
the reduced temperature and a specific constant estimated from group contributions. The
equations may be written as:

r    ()tr  ()int (4.61)

()tr  8.757 exp(0.0464  Tr )  exp( 0.2412  Tr )  (4.62)

()int  C  f (Tr ) (4.63)

T
Tr  (4.64)
Tc
Chapter 4: Slurry-loop Reactor Modelling 173

For olefins, f (Tr ) is calculated as:

f (Tr )  0.255  Tr  1.065  Tr 2  0.19 Tr 3 (4.65)

where C coefficient is a function of the number of carbons and their positions in the
carbon chain. For ethylene C  2.73 and for 1-hexene C  10.62 .

Due to the low mass fraction of ethylene and 1-hexene in the reaction mixture, it is
assumed that pressure has no effect on the value of thermal conductivity for those gases.
However, if necessary, Stiel and Thodos modification may estimate the effect of pressure
on thermal conductivity of gases. Roy and Thodos method for the calculation of thermal
conductivity of gases as well as Stiel and Thodos modification are presented in detail by
Ried et al., 1986.

4.5.2.2. Thermal Conductivity of Liquids

Latini et al. method is employed for the calculation of thermal conductivity of pure
liquids, L (i.e., i-butane).

0.38
A 1  Tr 
L  (4.66)
Tr1/ 6

A*  Tb a g g
A for 50 mol  MW  250 mol (4.67)
MW   Tc

where L is the thermal conductivity of the liquid in W/(m.K) and Tb is the normal boiling

point temperature, K. The terms A* ,  and  depend on the compound studied. For

saturated hydrocarbons, A*  0.0035 ,   1.2 ,   0.5 and   0.167 .

For simple organic liquids, the thermal conductivities values are 10 to 100 times
higher than the corresponding thermal conductivities of the low pressure gases at the same
temperature. There is little effect of pressure, and by increasing the temperature usually the
174 Chapter 4: Slurry-loop Reactor Modelling

thermal conductivity values decrease. These characteristics are similar to those noted for
liquid viscosities, although the temperature dependence of the latter is nearly exponential.

4.5.2.3. Thermal Conductivity of Polyethylene

According to Anderson, 1966 thermal conductivity of polyethylene may be assumed


constant, PE  3.84035W / m / K .

4.5.2.4. Mixture Thermal Conductivity

Mixture thermal conductivity is calculated via the following mixing rule:

mix   i  mi (4.68)

where mi denotes the components’ mass fraction.

4.5.3. Constant Pressure Heat Capacity

4.5.3.1. Constant Pressure Heat Capacity of Ethylene

Table 4.3 reports values of ethylene heat capacity as can be found in literature.. In
Table 4.4 the values of saturated ethylene for the operating range of temperature are also
reported.

Table 4.3: Ethylene heat capacity

Cp, J/kg/K Temperature range Reference

2402 16-169 K Chao et al., 1983

2396 15-170 K Egan and Kemp, 1937

1870 Neto et al, 2005


Chapter 4: Slurry-loop Reactor Modelling 175

Table 4.4: Heat capacity of saturated ethylene.

Cp, J/mol/K,
Temperature, K P, bar
(Perry and Green, 1984)

230 13.18 2706

240 17.71 2915

250 23.28 3260

260 30.03 3775

270 38.11 4990

275 50.97 -

6000
Saturated ethylene, Perry's Handbook
Chao, et al., 1983
5500 Egan and Kemp, 1937
Neto et al, 2005

5000
Heat Capacity, J/Kg/K

4500

4000

3500

3000

2500

2000

160 180 200 220 240 260 280 300 320 340 360
Temperature, K

Figure 4.18: Ethylene heat capacity.


176 Chapter 4: Slurry-loop Reactor Modelling

4.5.3.2. Constant Pressure Heat Capacity of 1-Hexene

Table 4.5 reports the values of 1-hexene heat capacity, as can be found in literature.
Table 4.6 reports the values heat capacity of liquid 1-hexene for the operating range of
temperature. The following correlation is proposed by McCullough, et al., 1957 for
hydrocarbons with 6 atoms of carbon (temperature range: T=150-310K):

C p  4,185.8(47.384  0.15437  T  0.6502 103T 2  0.579  106 T 3 ) / 84.163 , J/kg/K (4.69)

In Figure 4.19, the measured values of 1-hexene heat capacity are compared to the
calculated values (according to McCullough, et al., 1957).
Chapter 4: Slurry-loop Reactor Modelling 177

Table 4.5: Liquid 1-hexene pressure heat capacity.

Cp, J/kg/K Temperature range Reference

2171 180-300 K Kalinowska and Woycicki, 1985

2177 11-360 K McCullough et al., 1957

Table 4.6: Heat capacity of liquid 1-hexene (Kalinowska and Woycicki, 1985).

T, K Cp, J/kg/K T, K Cp, J/kg/K

186.581 1876 248.079 2010

189.434 1882 252.102 2022

192.401 1889 255.427 2031

195.643 1895 259.555 2043

199.039 1903 263.577 2055

202.451 1910 267.667 2068

205.899 1917 270.222 2076

208.802 1923 274.352 2089

212.865 1931 278.448 2102

217.150 1939 282.559 2115

221.24 1948 287.157 2130

229.449 1965 289.907 2140

233.557 1974 294.339 2155

235.928 1980 298.555 2172

239.938 1989 301.060 2182

243.956 1999
178 Chapter 4: Slurry-loop Reactor Modelling

2600

Specific Heat at Constant Pressure, J/(Kg K)


McCullough et al., 1957
Kalinowska and Woycicki, 1985
2500
McCullough, et al., 1957 correlation

2400

2300

2200

2100

2000

1900

1800
160 180 200 220 240 260 280 300 320 340 360 380
Temperature, K

Figure 4.19: 1-Hexene heat capacity.

4.5.3.3. Constant Pressure Heat Capacity of i-butane

Table 4.7 reports constant values of i-butane heat capacity, as suggested in the
literature. Table 4.8 reports the values of saturated i-butane for the operating range of
temperature.
Chapter 4: Slurry-loop Reactor Modelling 179

Table 4.7: Liquid 1-hexene heat capacity.

Cp, J/mol/K Temperature range Reference

2231.4 20-260 K Aston et al., 1940

2209.0 79-261 K Park et al., 1937

Table 4.8: Heat capacity of liquid i-butane (Perry and Green, 1984, 3-199).

Temperature, K P, bar Cp, J/kg/K

300 3.7365 2530

310 4.934 2610

320 6.392 2700

330 8.14 2816

340 10.21 2920

350 12.64 3040

360 15.46 3170

370 18.72 3310

380 22.48 3451

390 26.82 3620

400 31.86 3849


180 Chapter 4: Slurry-loop Reactor Modelling

4000
Aston et al, 1940

Specific Heat at Constant Pressure, J/(Kg K)


Parks et al, 1937
3500
Perry's Handbook

3000

2500

2000

1500

1000

500

0
0 50 100 150 200 250 300 350 400
Temperature, K

Figure 4.20: i-butane heat capacity.

4.5.3.4. Constant Pressure Heat Capacity of Polyethylene

Table 4.9 reports constant values of solid polyethylene constant pressure heat
capacity as can be found in literature.

Table 4.9: Polyethylene heat capacity.

Cp, J/kg/K Temperature range Reference

890.0 5-380 K. Value per monomer unit. Chang, 1976

883.6 or 1,147.8 5-350 K. Value per monomer unit. Chang, Westrum, et al., 1975

874.3 1-420 K. 100% crystalline phase Wunderlich, 1962

1,170.9 1-420 K. 100% amorphous phase. Wunderlich, 1962

872.2 or 1039.7 90-415 K. Passaglia and Kevorkian, 1963

2249.9 Value per monomer unit. Neto et al, 2005


Chapter 4: Slurry-loop Reactor Modelling 181

Heat capacity of 100% crystalline and 100% amorphous polyethylene can be


calculated via the correlations (Wunderlich, 1962, a, Wunderlich, 1962, b):

C p,cryst  4,185.8 0.1205  1.810  103T  , J/kg/K, 310K to melting, Crystalline PE (4.70)


C p , am  4,185.8 0.3985  0.538  103 T , J/kg/K, 250-470K, Amorphous PE  (4.71)

The heat capacity of partially crystalline PE can be calculated as:

C p  C p ,cryst  Wcrys  C p, am  (1  Wcrys ) , J/kg/K (4.72)

where Wcrys is the crystallinity coefficient. In Figure 4.21 the PE heat capacity range
between fully amorphous and fully crystalline morphology is displayed.

2600
Specific Heat at Constant Pressure, J/(Kg K)

2400

2200

2000
Wunderlich, B., 1962
100% Amorphous polymer
100% Crystalline polymer
1800
320 340 360 380 400
Temperature, K

Figure 4.21: Polyethylene heat capacity.


182 Chapter 4: Slurry-loop Reactor Modelling

4.5.3.5. Mixture Constant Pressure Heat Capacity

Mixture heat capacity is calculated via the following mixing rule:

C p, mix   C p,i  mi (4.73)

where mi is components mass fraction.

4.5.4. Heat of Polymerization

Table 4.10 reports the heat of polymerization for polyethylene as can be found n
literature. As can be seen, the value of heat of polymerization may vary from 83-136
kJ/mol depending on the characteristics of the process where polymerization takes place. It
is noted that the heat of polymerization value selected for this research project was the one
proposed by Zacca and Ray, 1993.

Table 4.10: Heat of polymerization for polyethylene.

ΔH, kJ/mol Comments Reference

83.716 Liquid-phase polymerization in loop reactors. Zacca and Ray, 1993

107.5 Gaseous monomers to partially crystalline polymer. Brandrup, J. and


Immergut, 1989, II/ 297
Method of determination: reaction calorimetry.

108.5 Gaseous monomers to partially crystalline polymer. Brandrup, J. and


Immergut, 1989, II/ 297
Method of determination: Combustion of monomer or
polymer or both.

89.576 High pressure tubular PE reactors. Chen et al, 1976

136.671 Gas phase polymerization Kanellopoulos et al, 2004


Chapter 4: Slurry-loop Reactor Modelling 183

4.5.5. Physical Properties of Coolant Water

The physical properties of coolant water are given by the following equations.

Density (Xie et. al., 1991)

c , j  1011.0  0.4484 Tc , j  273.15  (4.74)

Heat Capacity (Daupert and Danner, 1985):


Cpc , j  4.02 exp 1.99 104 Tc , j  (4.75)

Viscosity (Daupert and Danner, 1985):

c , j  4.8exp  1.5366 102 Tc , j  / 60 (4.76)

Thermal Conductivity (Daupert and Danner, 1985):

kc , j  (8.212 103 ln Tc , j   1.0661102 ) / 60 (4.77)

Table 4.11 shows a set of experimental measurements regarding water’s physical


properties. In Figure 4.21-Figure 4.25 are shown both experimental and calculated values
of water’s physical properties.
Chapter 4: Slurry-Loop Reactor Modeling 184

Table 4.11: Physical properties of water (deMan, J. M., 1993).

Temperature, K Fit

273.15 293.15 313.15 333.15 353.15 373.15

c, j  1000.21  0.0603 Tc, j  273.15  0.00362 Tc, j  273.15


2
Density, kg/m3 999.8 998.2 992.2 983.2 971.8 958.3

Specific heat, Cpc , j  4.215  2.31 103 Tc , j  273.15   3.851 105 Tc , j  273.15   1.555 107 Tc , j  273.15 
2 3

4.216 4.180 4.177 4.183 4.195 4.215


J/g/K

Viscosity, Pa s 17.9 10-4 10 10-4 6.53 10-4 4.66 10-4 3.55 10-4 2.82 10-4 c , j  0.00153  exp  (Tc , j  273.15) / 29.187   0.00025

Thermal
kc , j  5.645 10 4  1.887 10 6  (Tc , j  273.15)  7.267 10 9  (Tc , j  273.15) 2
conductivity, 5.65 10-4 5.98 10-4 6.27 10-4 6.52 10-4 6.69 10-4 6.80 10-4
kJ/m2/s/K
Chapter 4: Slurry-loop Reactor Modelling 185

1025
Experimental measurement (deMan, 1999)
Fit on experimental data
Correlation of Xie et. al., 1991

1000
3
Density, Kg/m

975

950
0 20 40 60 80 100
o
Temperature, C

Figure 4.22: Effect of temperature on water density.

4.35
Experimental data (deMan, 1999)
Fit of experimental data
Correlation of Daupert and Danner, 1985
Specific heat capacity, KJ/Kg/K

4.30

4.25

4.20

4.15
0 20 40 60 80 100
o
Temperature, C

Figure 4.23: Effect of temperature on water specific heat.


186 Chapter 4: Slurry-loop Reactor Modelling

0.002
Experimental data (deMan, 1999)
Fit of experimental data
Correlation of Daupert and Danner, 1985
Viscosity, Pas

0.001

0.000
0 20 40 60 80 100
o
Temperature, C

Figure 4.24: Effect of temperature on water viscosity.

7.0
Experimental data (deMan, 1999)
Fit of experimental data
6.8 Correlation of Daupert and Danner, 1985
Thermal conductivity, 10 KJ/(msK)

6.6
-4

6.4

6.2

6.0

5.8

5.6

5.4
0 20 40 60 80 100
o
Temperature, C

Figure 4.25: Effect of temperature on water thermal conductivity.


Chapter 4: Slurry-loop Reactor Modelling 187

4.6. Rheological Properties

Viscoelastic behaviour, directly affected by the distributed molecular properties of


polyolefins (i.e., including the molecular weight distribution (i.e., MWD), the co-polymer
composition distribution (i.e., CCD), etc.), to a large extent control the processing
behaviour as well as the end-use properties of polyolefins. For example, high molecular
weight polyolefins exhibit improved mechanical properties (e.g., toughness, strength,
impact resistance and environmental stress cracking resistance, etc.). However, high
molecular weight resins have higher melt viscosities and, therefore, it is more difficult to
process them. On the other hand, polyolefins having a broad or bimodal MWD exhibit
improved processability and, particularly, extrudability. In this way, rheological
parameters act as a link between molecular structure and final properties of a polymer.
Thus, the prediction of the viscoelastic behaviour of polymers is of great importance.

In the present research project, Pladis et al., 2008 rheological model is employed for
the prediction of the viscoelastic behaviour of polyolefins produced in slurry-loop reactors
(i.e., complex viscosity and melt flow index). In what follows, an advanced rheological
model for the calculation of the viscoelastic properties of linear polyolefins in terms of
their respective molecular weight properties is presented. The methodology for the
calculation of melt index (i.e., MI) (i.e., referring to the number of grams of polymer melt
per ten minutes pushed out of a die of prescribed dimensions under the action of a
specified load) is presented in detail in the work of Pladis et al., 2008.

The rheological model employed is based on the reptation and Rouse relaxation
theories, to calculate the rheological curve (i.e., melt viscosity versus frequency) of linear
polyolefins in terms of their respective MWDs. The dependence of the longest relaxation
time,  , the polymer chain diffusion coefficient, D , and the polymer melt viscosity,  ,
on the polymer chain molecular weight, MW , are given by the following correlations:

  MW 2 ; D  MW 1 ; η  MW for MW  MWe

  MW 3 ; D  MW 2 ; η  MW 3 for MW  MWe

where MWe is the chain entanglement molecular weight.


188 Chapter 4: Slurry-loop Reactor Modelling

According to Tsenoglou, 1991 and des Cloizeaux, 1990 the dynamic relaxation

modulus, Grept (t ) , of a polymer melt can be related to the molecular weight distribution

(i.e., w( MW ) ) of the polymer, using the following mixing rule:


  
0  1/   (4.78)
Grept (t )  G N  Fm ( t , MW ) w ( MW ) d log( MW )
 
 log( MWe ) 

0
where G N is the plateau modulus and Fm (t , MW ) is a monodisperse relaxation function
for polymer chains of molecular weight, MW , at time t . The parameter  is usually set
equal to two. However, slightly higher values for b are also often used. In this study, the
value of b was set equal to 2.25 (Van Ruymbeke et al., 2002a, Van Ruymbeke et al.,
2002b).

The following expression for Fm (t , MW ) was employed:

8 exp( i 2U (t , MW ))
Fm (t , MW )1/    (4.79)
 2 iodd i2

The function U (t , MW ) relates the polymer molar mass to the relaxation time:

 
t MW *  Mt 
U (t , MW )   g (4.80)
 rept ( MW ) MW  M * rept ( MW ) 
 

where MW * is some reference molecular weight.

The reptation time,  rept ( MW ) , exhibits a power law dependence on the polymer

molar mass, MW :

 rept ( MW )  K rept MW 3 (4.81)


Chapter 4: Slurry-loop Reactor Modelling 189

Finally, the function g , appearing in Eq. 4.80 has the following functional
dependence on x :

2
 1  e i x
 
0.5
g ( x)     x  x  ( x )0.5   (4.82)
i 1 i2

To account for the contribution of the polymer chains below the entanglement
molecular weight to the dynamic relaxation modulus, the following equation was
employed (Pattamaprom and Larson, 2001, Van Ruymbeke et al.,2002a):


0
G R (t )  G N  FR (t , MW ) w( MW ) d log( MW ) (4.83)
log( MWe )

where FR (t , MW ) is the Rouse relaxation function, given by:

 i 2 t i 2 t 

1   ( MW ) 1 N  R ( MW ) 
FR (t , MW )    e R  e  (4.84)
N  iN 3 i 1 
 
 

 R ( MW ) is the Rouse relaxation time, exhibiting a squared dependence on the molecular

weight, MW :

 R ( MW )  K R MW 2 (4.85)

Finally, N  MW / MWe is a normalized molecular weight with respect to the chain


entanglement molecular weight. The first term on the right hand site of Eq. 4.84 represents
the Rouse relaxation process for the whole polymer chain, restricted to subchains smaller

than MWe , while the second term accounts for the slower relaxation modes, but only in the
longitudinal dimension. Following the original developments of Van Ruymbeke et al.,
2002a the overall dynamic relaxation modulus will be given by the sum of the reptation
and Rouse contributions (see Eq. 4.78 and Eq.4.83):
190 Chapter 4: Slurry-loop Reactor Modelling

G (t )  Grept (t )  GR (t ) (4.86)

The numerical values of the model parameters were taken from Van Ruymbeke et
al., 2002a and Doelder, 2006 and are reported in Table 4.12. It should be pointed out that
the numerical values of the rheological model parameters can be obtained either from the
literature or from combined experimental measurements of the MWD and dynamic storage
and loss moduli of selected industrial PE grades. The estimated values of the model
parameters can then be employed to predict the rheological behaviour of other polymer
grades independently of the MWD type (e.g., unimodal, bimodal, monodisperse, broad).

Table 4.12: Numerical values of the rheological model parameters for polyethylene.

MWe (1) 0
GN (1)
MW * (1) K rept (1) KR (2)

g/mol Pa g/mol s  mol 3 / g 3 s  mol 2 / g 2

1.5x103 2.6x106 7.0x104 2.0x10-17 2.0x10-12

(1) Van Ruymbeke et al., 2002a

(2) Doelder, 2006

The well known Schwarzl approximations are employed for the calculation of the
' ''
storage and loss muduli, G and G (Schwarzl, 1971):

G ' ( )  G (t )  0.142  G (4t )  G (8t )   0.717  G (2t )  G (4t ) 


 1/ t
+0.046  G (t )  G (2t )   0.099  G (t / 2)  G (t ) 
(4.87)
+0.103  G (t / 4)  G (t / 2)   0.001 G (t / 8)  G (t / 4) 
+0.00716  G (t /16)  G (t / 8)   0.000451 G (t / 64)  G (t / 32) 
Chapter 4: Slurry-loop Reactor Modelling 191

G '' ( )  0.470  G (2t )  G (4t )   1.674  G (t )  G (2t ) 


 1/ t
+0.198  G (t / 2)  G (2)   0.620  G (t / 4)  G (2) 
(4.88)
+0.012  G (t / 8)  G (t / 4)   0.172  G (t /16)  G (t / 8) 
+0.043  G (t / 64)  G (t / 32)   0.0108  G (t / 256)  G (t /128) 

' ''
It should be noted that the above approximations for G ( ) and G ( ) provide
excellent results in the frequency range of commonly-available experimental
*
measurements. From Eq. 4.87 and Eq. 4.88, the complex viscosity, |  ( ) | , can easily be
obtained, using the following expression:

 
1 1/ 2
|  * ( ) | (G ' ( )) 2  (G '' ( )) 2 (4.89)

Accordingly, the loss factor is given by:

G ' ( )
tan   (4.90)
G '' ( )

Moreover, the zero-shear viscosity, 0 , and the steady state recoverable compliance,

J e0 , will be given by the following integrals:

 
0  0 G (t ) dt ; J e002  0 tG (t ) dt (4.91)

4.7. Particle Size Distribution Calculation

4.7.1. Introduction

An important property of many particulate processes is the particle size distribution


(i.e., PSD), which controls key aspects of the process and affects the end-use properties of
the product. Particulate processes are generally characterized by PSDs that can vary in time
with respect to the mean particle size as well as to the PSD form (i.e., broadness and/ or
192 Chapter 4: Slurry-loop Reactor Modelling

skewness of the distribution, unimodal and/ or bimodal character, etc.). For reactive
particulate processes, the quantitative calculation of the evolution of the PSD presupposes
a good knowledge of the particle nucleation, growth, breakage, and aggregation
mechanisms. These mechanisms are usually coupled to the reaction kinetics,
thermodynamics (e.g., solubility of a reactant in the particulate phase), and other
microscale phenomena including mass and heat transfer between the different phases
present in the system (Kiparissides et al., 2004).

Particle nucleation often results in the formation of a large number of small particles
within a short period of time. Particle growth due to chemical reaction results in an
increase of the mean particle size and can affect the form of the PSD, particularly in size-
dependent particle growth processes. Finally, particle aggregation and breakage can result
in significant changes in the form of the PSD. The time evolution of the PSD is commonly
obtained from the solution of the general population balance equation (i.e., PBE)
governing the dynamic behaviour of a particulate process (Hulburt and Katz, 1964,
Ramkrishna, 2000).

In general, the numerical solution of the dynamic PBE for a particulate process,
especially for a reactive one, is a notably difficult problem because of both numerical
complexities and model uncertainties regarding the particle nucleation, growth,
aggregation, and breakage mechanisms that are often poorly understood. Usually, the
numerical solution of the PBE requires the discretization of the particle volume domain
into a number of discrete elements that results in a system of stiff, nonlinear differential or
algebraic/ differential equations (i.e., DAEs) that are solved numerically.

In the open literature, several numerical methods have been developed for solving
the steady-state or dynamic PBE. These include the full discrete method (Hidy, 1965) the
method of classes (Marchal et al., 1988, Chatzi and Kiparissides, 1992) the discretized
PBE (i.e., DPBE) (Batterham et al., 1981, Hounslow et al., 1988) the fixed and moving
pivot DPBE methods (Kumar and Ramkrishna, 1996a, Kumar and Ramkrishna, 1996b),
the high-order DPBE methods (Sastry and Gaschignard, 1981, Landgrebe and Pratsinis,
1990) the orthogonal collocation on finite elements (i.e., OCFE) (Gelbard and Seinfeld,
1979) and the Galerkin method (Nicmanis and Hounslow, 1998). Kiparissides et al., 2004
presents a detailed description of all the numerical methods employed for the solution of
the system of nonlinear DAEs.
Chapter 4: Slurry-loop Reactor Modelling 193

Zacca et al., 1996 developed a population balance model using the catalyst residence
time as the main coordinate, to model particle-size developments in multistage olefin
polymerization reactors. A two-site copolymerization kinetic scheme was employed to
account for the polymerization kinetics and some functional, obtained by fitting simulated
data from a detailed particle growth model, was used to imitate the effect of mass transfer
resistances (effectiveness factor) on particle growth. Choi et al., 1994 followed the
population balance approach of Kunii and Levenspiel, 1991, to investigate the effect of
feed catalyst size distribution on the average molecular properties and the particle size
distribution of the polyolefin produced in a fluidized-bed reactor for both non-deactivated
and deactivated catalysts. A two-site copolymerization kinetic scheme and a multigrain
solid core model were used to describe the molecular weight developments and monomer
diffusion in the growing polymer particle. Soares and Hamielec, 1995 developed a
mathematical model to analyse the effects of ideal and non-ideal reactor residence time
distribution on the polymer particle-size distribution. For reaction rates relevant to typical
Ziegler-Natta and metallocene catalysts, the effects of non-ideal flow and catalyst
deactivation were analysed. In Touloupides et al., 2010b, a comprehensive multi-scale
mathematical model was developed to predict the dynamic evolution of morphological
(i.e., particle size distribution (PSD)) and molecular (i.e., molecular weight distribution
(MWD)) distributed polymer properties in catalytic Ziegler-Natta (Z-N) gas and slurry
phase olefin polymerization reactors. According to the multi-scale description of the
continuous olefin polymerization reactors, models at three different levels, namely, a
kinetic model, a single particle model and a population balance model were employed.
Moreover, the effect of reactor media (e.g., slurry versus gas phase) on the MWD and PSD
of the produced polymer particles was thoroughly analyzed.

In the present research study, a generalized dynamic population balance model is


employed to investigate the particle-size distribution developments in slurry-phase loop
reactors. The model accounts for the combined effects of particle growth, attrition and
agglomeration for distributed catalyst feed.

4.7.2. Calculation of PSD

To calculate the dynamic evolution of particle size distribution, a population balance


model needs to be solved (Kanellopoulos, et al., 2004). The operation of loop reactors can
194 Chapter 4: Slurry-loop Reactor Modelling

be approximated by a perfectly backmixed, continuous flow reactor. In the present research


study, it is assumed that particle attrition and elutriation is negligible. Moreover, for the
shake of simplicity, it is assumed that particles are spherical and of constant density.per
mass of polymer.

According to the developments of Randolph and Larson, 1971, Litster et al., 1995,
Dompazis, et al., 2008 and Touloupides et al., 2010b, the following dynamic population
balance equation, accounting for both particle growth and particle agglomeration can be
derived to describe the evolution of PSD in a continuous stirred tank reactor:

n p  D,t   G  D  n p  D,t   1
   Bt  D,t   Dt  D,t    Fc nc  D,t   Fp n p  D,t   (4.92)
t D W

where:

Bt  D,t  
D 2
(D3  Dmin
3
)1 / 3
K ag   D  D 
3 3 13
  D  D  ,t  n  D,t  dD
,D n p 3 3 13
p

 (4.93)
2 Dmin
 D  D
3

3 23

and,

Dmax
Dt  D,t   n p  D,t   K ag  D,D  n p  D,t  dD (4.94)
Dmin

Fc and Fp (g/s) denote the catalyst particles feed into the reactor and the product

withdrawal rate respectively. W  t  (g) is the mass of the solids in the reactor that can vary

with time. n p  D,t  and nc  D,t  , expressed in (#/g/cm), are the corresponding number

density function of the particles in the reactor and in the feed stream, respectively. The
term n p  D,t  dD denotes the number of particles in the size range  D, D  dD  .

According to Hatzantonis et al., 1998, the particle growth rate, G  D  , can be

expressed in terms of the overall particle polymerization rate, R p , as follows.

G  D   2 R p  D  ρ p πD 2 (4.95)
Chapter 4: Slurry-loop Reactor Modelling 195

The product withdrawal rate, Fp (g/s), can be easily calculated via a mass balance

equation:

Dmax
 ρ p πD 3 
Fp  Fc  W  G  D  n p  D,t  d 
 6  (4.96)
Dmin  

Notice that the second term on the right-hand side of Eq.4.96 accounts for the total
polymer production rate in the reactor. From Eq. 4.96, one can calculate the product
withdrawal rate, Fp , provided that the solids weight, W  t  , the particle growth rate,

GD  D  , and the number density function, n p  D,t  , are known.

For the special case of a uniform size catalyst feed, the total number of catalyst
particles in the feed stream per unit mass will be given by:

N c  Dc   6  p Dc3 (4.97)

K ag  D,D  is a temperature and particle size dependent kernel, governing the

agglomeration rate of particles of sizes D and D (Yiannoulakis et al., 2001).

Despite the large number of publications on particle agglomeration (e.g., aerosols,


granulation, crystallization, etc.), there is little experimental and theoretical evidence
regarding particle agglomeration in olefin polymerization reactors. Following the
developments of Kapur, 1972, Hartel and Randolph, 1986 and Arastoopour et al., 1988 and
Dompazis et al., 2006 the agglomeration kernel can be written as a product of two
functions:

K ag  g Tav ,Tsf   f  Di ,D j   Tav   miTs ,i  m jTs , j  /  mi  m j  (4.98)

where Tav denotes the average temperature of the two colliding particles of sizes Di and

D j , and Tsf is the average polymer softening temperature of the two particles, calculated

in terms of the copolymer compositions and densities of the particles. ( Ts ,i , Ts , j ) and ( mi ,


196 Chapter 4: Slurry-loop Reactor Modelling

m j ) are the respective temperatures and masses of the two colliding particles.

Yiannoulakis et al., 2001 proposed the following exponential function for g :

g Tav ,Tsf   K0  exp  Eg Tav / Tsf  (4.99)

K0 and Eg are proportionality constants that can be estimated from experimental data on

particle agglomeration.

In general, the particle collision function, f  Di ,D j  , is assumed to depend on the

hydrodynamic behaviour of the polymer particles in the reactor (e.g., the mechanism of
dual particle collision, the diameters of the colliding particles Di and D j , etc.) and,

consequently, will be influenced by the reactor geometric characteristics as well as by the


reactor operating conditions (e.g., circulation velocity, solids concentration, etc.).

In this study, a simple agglomeration kernel1 that favours the collisions of small
with small particles was employed.

K ag  K0  f  Di ,D j   K0   Di2  D 2j    Di  D j 
4
(4.100)

Finally, one can define in terms of n p  D,t  , the probability density function

Pp  D,t  dD , which represents the mass fraction of particles in the size range  D, D  dD  :

 p D 3
Pp  D,t  dD  n p  D,t  dD (4.101)
6

As a result, Pp  D,t  satisfies the following normalization condition:

 Pp  D,t  dD  1 (4.101)
Dmin
Chapter 4: Slurry-loop Reactor Modelling 197

4.7.3. The Orthogonal Collocation on Finite Elements Method

The continuous form of the PBE (Eq. 4.92) can be solved using the orthogonal
collocation on finite elements (i.e., OCFE) method. Gelbard and Seinfeld, 1978 first
employed the OCFE method to solve the dynamic PBE. The method was successfully
applied to the solution of a class of pure aggregation problems and to the general PBE (Eq.
4.92) for short aggregation times. Recently, Nicmanis and Hounslow, 1998 employed a
collocation method to determine the steady state PSD in a continuous particulate process
undergoing combined particle growth and aggregation.

In the OCFE method, the particle volume domain is first divided into ‘ ne ’ elements

based on an appropriately selected volume discretization rule. Then, ‘ nc ’ internal


collocation points are specified in each element. Accordingly, the unknown number
density function is approximated at the internal and boundary collocation points of each
element, ‘ e ’, in terms of Lagrange basis functions,  j :

nc 1
nV V ,t    nej  t    j  t  (4.102)
j 0

where nej denotes the value of nV V ,t  at the ‘ j ’ internal or boundary collocation point.

The above discretization approach, results in a total number of  ne  nc  1  2  unknown

values of the number density function, nej .

Following the general developments of Finlayson, 1980, Eq. 4.92 is recast into a
system of ( ne  nc ) residual equations corresponding to all the internal points of the ‘ ne ’
volume elements.
198 Chapter 4: Slurry-loop Reactor Modelling

nc 1
 d j  e  dGV  e
nie  t       n j  t  J GV Vi     ni  t 
e

j 0  d  i  dV i
ne nc
 nie  t   wkG J  Vi e ,Vk f  nkf  t 
f

f 1 k 1
g 1 nc (4.103)
   wGj J  Vi e  V j f ,V j f  nie nV Vi e  V j f  
f

f  1 j 1
 
Vie / 2

   V  V ,V  nV V nV Vi e  V  dV
e
i
V1g

According to the standard finite element formulation, the global volume domain of
each element ‘ e ’ is linearly transformed to a local domain [−1, 1]. The index ‘ g ’ denotes
e f
the element containing Vi / 2 . J is the Jacobian of the volume transformation and wkG

are the integration weights of the Gauss–Legendre quadrature rule. At the boundary points
between the various elements, the number density function and its first derivative are
forced to be continuous. Thus, the following  ne  1 continuity conditions between all the

adjacent pair of elements e and e  1 are written:

 
ne nc
n01  t   n01  t   w  j  J  V01 ,V j f  nkf  t 
f
(4.104)
f 1 k 1

Since the total number of unknown nodal values of nej (i.e.,  ne  nc  1  1 ), is less

than the total number of residual equations and continuity conditions (i.e.,

n n
e c  1  1 ), two additional equations are needed to produce a closed system of DAEs.

These equations correspond to the values of nej at the minimum Vmin ( e  1, j  0 ) and

maximum Vmax ( e  ne , j  nc  1 ) value of the volume integration domain. At V  Vmax , a

residual equation similar to Eq. 4.103 can be written. On the other hand, at V  Vmin , the
residual equation for the free boundary condition takes the following form:
Chapter 4: Slurry-loop Reactor Modelling 199

nc
 d j  nc
 d j  e 1
 j     n j t  J
e e 1
  n e
t J   (4.105)
j 1  d   j 1  d  
nc1 0

In the present research study, the orthogonal collocation on finite elements method
(i.e., OCFE) was employed for solving the dynamic PBE (i.e., Eq. 4.92) (Alexopoulos et
al., 2004). A third order Lagrange polynomial is chosen to approximate the actual solution
within the integration domain. Note that the selected polynomial approximation satisfies
the numerical solution criteria on convergence and computational efficiency. The resulting
system of stiff, non-linear algebraic differential Eqs. 4.103–4.105 was solved using the
double precision Petzold-Gear BDF method (IMSL, routine DASPG).
200 Chapter 4: Slurry-loop Reactor Modelling

4.8. Notation

Symbols

A : Heat transfer surface area, m2

Ci , j : Reactor phase concentration of molecular species ‘ i ’ in ‘ j th ’ reactor, mol/l

: Corrected reactor concentration, mol/l


Ci', j

Cout ,i, j : Outlow concentration of molecular species ‘ i ’ in ‘ j th ’ reactor, mol/l

Cp : Heat capacity, kJ/(kg K)

: Reduced reactor concentration, mol/l


Cred , j

Csorbed , am,i : Concentration of the sorbed molecular species ‘ i ’ in the amorphous part of
the solid polymer phase, mol/l

Csorbed , solid ,i : Concentration of the sorbed molecular species ‘ i ’ in the solid polymer
phase, mol/l

d : Diameter, m

e : Error signal

ef : Enhancement factor

en : Most recent value of e

en 1 : Next most recent value of e

Fin,i , j : Inflow of molecular species ‘ i ’ in ‘ j th ’ reactor, mol/s

fp : Volume fraction of the leg volume that consists of packed/ settled polymer
Chapter 4: Slurry-loop Reactor Modelling 201

frsolid ,i : Mass fraction of the molecular species ‘ i ’ in the solid polymer phase

hc : Outside (coolant) heat transfer coefficient, W/m2/k

hR : Inside (reactor) heat transfer coefficient, W/m2/k

l : Length, m

m : Mass flowrate, kg/s

MW : Molecular weight, g/mol

Nu : Nusselt number

P : Pressure, bar

p (t ) : Controller output

: Nominal value for constant error


p

: Dimensionless pressure
P

pi : Partial pressure, bar

Pp : Pump power, kJ/s

Pr : Prandtl number

: Pressure in ‘ j th ’ reactor, bar


PR , j

PR , j ,max : Maximum allowable pressure in ‘ j th ’ reactor, bar

qin ,i , j : Volumetric flowrate of molecular species ‘ i ’ entering the ‘ j th ’ reactor


reactor, l/s

qout , j : Volumetric flowrate of molecular species ‘ i ’ stream leaving the reactor, l/s
202 Chapter 4: Slurry-loop Reactor Modelling

R : Heat transfer resistance, m2K/W

Re : Reynolds number

Rf : Reactor fouling factor, W/m2/k

Ri, j : production/ consumpiton rate of molecular species ‘ i ’ in ‘ j th ’ reactor,


mol/l

Rp , j : Rate of polymerization, mol/s

Sr : Polymer particles settling rate, s-1

t : Time, sec

T : Temperature, K

: Dimensionless temperature
T

Tc ,in , j : Coolant inlet temperature in the ‘ j th ’ reactor reactor, K

Tc ,out , j : Coolant outlet temperature in the ‘ j th ’ reactor reactor, K

Tin ,i , j : Temperature of molecular species ‘ i ’ entering the ‘ j th ’ reactor reactor, K

Tref : Reference temperature, K

U : Heat transfer coefficient, W/m2/k

V : Volume, l

Wcrys : Degree of crystallinity

Subscripts and Superscripts

am : Amorphous polymer phase

b : bulk
Chapter 4: Slurry-loop Reactor Modelling 203

c : Coolant

cryst : Crystalline polymer phase

e : Environment

i : Referring to molar species ‘ i ’

j : Referring to ‘ j th ’ reactor

L : Settling leg

mix : Mixture

polymer : Polymer product property

R : Reactor

solid : Solid (polymer) phase

w : Wall

Greek letters

H r : Heat of reaction, J/mol

t f : Firing period, s

ts : Valve opening cycling period, s

k : Thermal conductivity, J/m/s/K

Kc : Controller gain

kH : Henry’s coefficient, bar/(mol/l)

 : Viscosity, Pa s

 : Density, kg/l
204 Chapter 4: Slurry-loop Reactor Modelling

 : Dimensionless density

D : Derivative time

I : Integral time

p : Volume fraction of the packed polymer particles in a settling leg

 s, j : Solids volume fraction in the reactor

out , j : Solids volume fraction in the settling leg


Chapter 4: Slurry-loop Reactor Modelling 205

4.9. References

Alexopoulos A. H., Roussos A. I. and Kiparissides C., 2004, ‘Part I: Dynamic evolution of
the particle size distribution in particulate processes undergoing combined particle
growth and aggregation’, Chem. Eng. Sci., 24, 5751-5769.

Anderson D.R., 1966, ‘Thermal conductivity of polymers’, Chem. Rev, 66, pp. 677–690.

Arastoopour H., Huang C.H. and Weil S.A., 1988, ‘Fluidization Behaviour of Particles
Under Agglomerating Conditions’, Chem. Eng. Sci., 43, pp. 3063-3075.

Aston J.G., Kennedy R.M. and Schumann, S.C., 1940, ‘The heat capacity and entropy,
heats of fusion and vaporization and the vapor pressure of isobutane’, J. Am. Chem.
Soc., 62, pp. 2059-2063.

Ayres C.A., Scott Jr. J.N. and Shrek F.T., 1986, Sep. 23, ‘Loop Reactor Settling Leg
System for Separation of Solid Polymers and Liquid Diluent’, US Patent No.
4,613,484.

Barnea E. and Mizrahi J., 1976, ‘On the ‘Effective’ Viscosity of Liquid-Liquid
Dispersions’, Ind. Eng. Chem., 15, No. 2, pp. 120-125.

Batterham R.J., Hall J.S. and Barton G., 1981, ‘Pelletizing Kinetics and Simulation for
Full-Scale Balling Circuits. In Proceedings of the 3rd International Symposium on
Aggregation’, Nurnberg, West Germany, p A136.

Benham E.A., McDaniel M.P., McElvain R. and Schneider R.O., 1991, Dec. 10, ‘High-
temperature Slurry Polymerization of Ethylene’, US Patent No. 5,071,927.

Bird R.B., Stewart W.E. and Lightfoot E.N., 2002, ‘Transport Phenomena’, Second
Edition, John Wiley and Sons.

Brandrup J. and Immergut E.H., 1989, ‘Polymer Handbook’, third edition, John Wiley and
Sons, Inc.

Brodkey R.S. and Hersey H.C., 1988, ‘Transport Phenomena, a Unified Approach’,
McGraw Hill International Editions, Chemical Engineering Science.

Chang S-S., 1976, ‘Heat capacities of polyethylene IV. High molecular weight linear
polyethylene’, J. Res., NBS A80, pp. 51-57.
206 Chapter 4: Slurry-loop Reactor Modelling

Chang S-S., Westrum E.F., Jr. and Carlson, H.G., 1975, ‘Heat capacities of polyethylene
III. One linear and one branched sample from 5 to 350 K’, J. Res., NBS 79A, pp.
437-441.

Chao J., Hall K.R. and Yao J.M., 1983, ‘Thermodynamic properties of simple alkenes’,
Thermochim. Acta, 64(3), pp. 285-303.

Chatzi E.G. and Kiparissides C., 1992, ‘Dynamic Simulations of Bimodal Drop Size
Distributions in Low-coalescence Batch Dispersion Systems’, Chem. Eng. Sci., 47
(2), 445-456.

Chen C.H., Vermeychuk J.G., Howell J.A. and Ehrlich, P., 1976, ‘Computer Model for
Tubular High-Pressure Polyethylene Reactors’, AIChE Journal, 22,3, pp. 463-471.

Chen C.M., 1993, ‘Gas Phase Olefin Polymerization with Ziegler-Natta Catalysts’, PhD
Thesis, University of Wisconsin.

Choi K.Y., Zhao X. and Tang S., 1994, ‘Population Balance Modelling for a Continuous
Gas Phase Olefin Polymerization Reactor’, J. Appl. Polym., Sci., 53, pp. 1589-1594.

Coulson J.M. and Richardson J.F., 1999, ‘Chemical Engineering’, The Bath Press, Great
Britain.

Daupert T.E. and Danner R.P., 1985, ‘Data Compilation Tables of Properties of Pure
Compounds’, Design Institute for Physical Properties Data, AIChE: New York.

deMan, J.M., 1999, ‘Principles of Food Chemistry’, 3rd Edition, Springer, Maryland.

des Cloizeaux J., 1990, ‘Relaxation of entangled polymers in melts’, Macromolecules,


23(17), pp. 3992-4006.

Doelder J.D., 2006, ‘Viscosity and Compliance from Molar Mass Distributions Using
Double Reptation Models’, Rheol. Acta, 46, 195.

Dompazis G., Kanellopoulos V. and Kiparissides C., 2006, ‘Assessment of Particle


Agglomeration in Catalytic Olefin Polymerization Reactors Using Rheological
Measurements’, Ind. Eng. Chem. Res., 45, pp. 3800-3809.

Dompazis G., Kanellopoulos V.,Touloupides V. and Kiparissides C., 2008, ‘Development


of a Multi-scale, Multi-phase, Multi-zone Dynamic Model for the Prediction of
Particle Segregation in Catalytic Olefin Polymerization FBRs’, Chem. Eng. Sci., 63,
pp. 4735-4753.
Chapter 4: Slurry-loop Reactor Modelling 207

Drusco G., Rinaldi R., 1984, ‘Polypropylene-Process Selection Criteria’, Hydrocarbon


Processing, pp.113-117.

Egan C.J. and Kemp J.D., 1937, ‘Ethylene. The heat capacity from 15°K to the boiling
point. The heats of fusion and vaporization. The vapor pressure of the liquid. The
entropy from thermal measurements compared with the entropy from spectroscopic
data’,J. Am. Chem. Soc., 59, pp. 1264-1268.

Ferrero M.A. and Chiovetta, M.G., 1990, ‘Preliminary Design of a Loop Reactor for Bulk
Propylene Polymerization’, Polym. Plast. Technol. Eng., 29 (3), pp. 263-287.

Finlayson B.A., 1992, ‘Numerical Methods for Problems with Moving Fronts’, Ravenna
Park Publ. Inc, Seattle.

Floyd S., Heiskanen T., Taylor T.W., Mann G.E. and Ray W.H., 1987, ‘Polymerization of
Olefins Through Heterogeneous Catalysis VI. Effect of Particle Heat and Mass
Transfer on Polymerization Behavior and Polymer Properties’, J. Appl. Polym. Sci.,
33, pp. 1021-1065.

Floyd S., Heiskanen T. and Ray W.H., 1988, ‘Solid Catalyzed Olefin Polymerization’,
Chem. Eng. Prog., 84, 11, pp. 56-62.

Fontes C.H. and Mendes M.J., 2005. ‘Analysis of an Industrial Continuous Slurry Reactor
for Ethylene–Butane Co-polymerization’, Polymer, 46, pp. 2922-2932.

Franklin ΙΙΙ, R.K., Hottovy J.D., Hensley H.D., Przelomski D.J., Cymbaluk T.H. and Perez
E.P., 2004, Jun. 1, ‘High Polymer Solids Slurry Polymerization Employing 1-Olefin
Comonomer’, US Patent No. 6,743,869 B2.

Gay E.C., Nelson P.A. and Armstrong, W.P., 1969, ‘Flow Properties of Suspensions with
High Solids Concentration’, AIChe Journal, 15, 6, pp. 815-822.

Geankoplis C.J., 1993, ‘Transport Processes and Unit Operations’, 3rd edition, Prentice
Hall.

Gelbard F. and Seinfeld J.H., 1979, ‘Exact Solution of the General Dynamic Equation for
Aerosol Growth by Condensation’, J. Colloid Interface Sci., 68 (1), pp. 173-183.

Ha K., Yoo K. and Rhee H., 2001, ‘Modeling and Analysis of a Slurry Reactor System for
Heterogeneous Olefin Polymerization: The Effects of Hydrogen Concentration and
Initial Catalyst Size’, J. Appl. Polym. Sci., 79, pp. 2480-2493.
208 Chapter 4: Slurry-loop Reactor Modelling

Hartel R.W. and Randolph A.D., 1986, ‘Mechanisms and Kinetic Modeling of Calcium
Oxalate Crystal Aggregation in a Urinelike Liquor. Part I: Mechanisms’, AIChE J.,
32, pp. 1176-1185.

Hatzantonis H., Goulas A. and Kiparissides C., 1998, ‘A comprehensive model for the
prediction of particle-size distribution in catalyzed olefin polymerization fluidized-
bed reactors’, Chem. Eng. Sci., 53, 18, pp. 3251-3267.

Hidy G.M., 1965, ‘On the Theory of the Coagulation of Noninteracting Particles in
Brownian Motion’, J. Colloid Sci., 20, 123-144.

Hottovy J.D., Hensley H.D., Przelomski D.J., Cymbaluk T.H., Franklin III R.K. and Perez
E.P., 2004, Oct. 19, ‘High Solids Slurry Polymerization Using Heat Exchange to
Condense the Flashed Diluent’, US Patent No. 6,806,324 B2.

Hounslow M.J., Ryall R.L. and Marshall V.R., 1988, ‘A Discretized Population Balance
for Nucleation, Growth, and Aggregation’, AIChE J., 34 (11), 1821-1832.

Hulburt H.M. and Katz S., 1964, ‘Some Problems in Particle technology. A Statistical
Mechanical Formulation.’, Chem. Eng. Sci., 19, pp. 555-574.

Hutchinson R.A., 1990, ‘Modeling of Particle Growth in Heterogeneous Catalyzed Olefin


Polymerizations’, PhD thesis, University of Wisconsin-Madison.

Hutchinson R.A., Chen C.M. and Ray W.H., 1992, ‘Polymerization of Olefins through
Heterogeneous Catalysis X: Modelling of Particle Growth and Morphology’, J. Appl.
Polym. Sci., 44, 1389-1414.

Kalinowska B. and Woycicki W., 1985, ‘Heat capacities and excess heat capacities of (an
alcohol + an unsaturated hydrocarbon). 1. (Propan-1-ol + n-hex-1-ene)’, J. Chem.
Thermodynam., 17, pp. 829-834.

Kanellopoulos V., Dompazis G., Gustafsson B. and Kiparissides C., 2004, ‘Compregensive
Analysis of Single-Particle Growth in Heterogeneous Olefin Polymerization: The
Random-Pore Polymeric Flow Model’, Ind. Eng. Chem. Res., 43, pp. 5166-5180.

Kanellopoulos V., Mouratides D., Pladis P. and Kiparissides C., 2006, ‘Prediction of
Solubility of α-olefins in Polyolefins Using a Combined Equation of State-Molecular
Dynamics Approach’, Ind. Eng. Chem. Res., 45 (17), pp. 5870–5878.
Chapter 4: Slurry-loop Reactor Modelling 209

Kapur P.C., 1972, ‘Kinetics of Granulation by Non-random Coalescence Mechanism’,


Chem. Eng. Sci., 27, pp. 1863-1869.

Khare N.P., Seavey K.C., Liu Y.A., Ramanathan S. and Lingard, S., Chen, C., 2002,
‘Steady-State and Dynamic Mdeling of Commercial Slurry High-Density
Polyethylene (HDPE) Process’, Ind. Eng. Chem. Res., 41, pp. 5601-5618.

Kiparissides C., 1996, ‘Polymerization Reactor Modeling: A Review of Recent


Developments and Future Directions’, Chem. Eng. Sci., 51, pp. 1637-1659.

Kiparissides C., Alexopoulos A., Roussos A., Dompazis G. and Kotoulas C., 2004,
‘Population Balance Modeling of Particulate Polymerization Processes’, Ind. Eng.
Chem. Res., 43, pp. 7290-7302.

Kiparissides C., Baltsas A., Papadopoulos S., Congalidis J.P., Richads J.R., Kelly M.B.
and Ye Y., 2005, ‘Mathematical Modeling of Free-Radical Ethylene
Copolymerization in High Pressure Tubular Reactors’, Ind. Eng. Chem. Sci., 44, pp.
2592-2605.

Ko Y.S., Han T.K., Sadatoshi H. and Woo S.I., 1998, ‘Analysis of Microstructure of
Ethylene–1-Hexene Copolymer Prepared over Thermally Pretreated
MgCl2/THF/TiCl4 Bimetallic Catalyst’, Inc. J. Polym Sci A: Polym Chem, 36, pp.
291-300.

Krallis A., Pladis P., Kanellopoulos V. and Kiparissides C., 2010, ‘Development of
Advanced Software Tools for Computer-Aided Design, Simulation, and
Optimization of Polymerization Processes’, Macromolecular Reaction Engineering,
4, pp. 303-318.

Kufeld S.E, Reid T.A., Burns D.H., Verser D.W., Hensley H.D., Przelomski D.J.,
Cymbaluk T.H., Franklin R.K.,Perez E.P. and Hottovy J.D., 2006, Apr. 25, ‘Slurry
Polymerization Reactor Having Large Length/ Diameter Ratio’, US Patent No.
7,033,545 B2.

Kumar S. and Ramkrishna D., 1996a, ‘On the Solution of Population Balance Equations
by Discretizations I. A Fixed Pivot Technique’, Chem. Eng. Sci., 51 (8), 1311-1332.

Kumar S. and Ramkrishna D., 1996b, ‘On the Solution of Population Balance Equations
by Discretizations II. A Moving Pivot Technique’, Chem. Eng. Sci., 51 (8), 1333-
1342.
210 Chapter 4: Slurry-loop Reactor Modelling

Kunii D. and Levenspiel O., 1991, ‘Fluidization Engineering’, Butterworth-Heinemann,


Boston MA., 2nd Ed.

Landgrebe J.D., Pratsinis S.E., 1990, ‘A Discrete-Sectional Model for Powder Production
by Gas-Phase Chemical reaction and Aerosol Coagulation in the Free-Molecular
Regime’, J. Colloid Interface Sci., 139 (1), pp. 63-86.

Lepski D.M. and Inkov A.M., 1977, ‘Mathematical Modelling of Polymerization of


Propylene in Loop Reactors’, Sb. Tr. Vses. Ob’edin. Neftekhlin, 13, 34.

Levenspiel O., 1972, ‘Chemical Reaction engineering’, Wiley, New York.

Li J., Zhang J., Ge W. and Liu X., 2004, ‘Multi-scale methodology for complex systems’,
Chem. Eng. Sci., 59, pp. 1687-1700.

Litster J.D., Smit D.J. and Hounslow M.J.,1995, ‘Adjustable Discretized Population
Balance for Growth and Aggregation’, AIChE J., 41, pp. 591-603.

Lonfils N., Bodart P. and Debras G., 2000, Aug. 1, ‘Catalysts for Polyethylene Production
and Use Thereof’, US Patent No. 6,096,679.

Marchal P., David R., Klein J.P. and Villermaux J., 1988, ‘Crystallization and Precipitation
Engineerings-I. An Efficient Method for Solving Population Balance in
Crystallization with Agglomeration’, Chem. Eng. Sci., 43 (1), 59-67.

Marechal P., 2006, Apr. 25, ‘Process for Producing Bimodal Polyethylene Resins’, US
Patent No. 7,034,092 B2.

Marwil S.J., 1966, Dec. 20,‘Withdrawal of Solids from a Flowing Stream Comprising a
Slurry of Same’, US Patent No. 3,293,000.

McCullough J.P., Finke H.L., Gross M.E., Messerly J.F. and Waddington G., 1957, ‘Low
temperature calorimetric studies of seven 1-olefins: effect of orientational disorder in
the solid state’, J. Phys. Chem., 61, pp. 289-301.

Moore S.J. and Wanke S.E., 2001, ‘Solubility of Ethylene, 1-butene and 1-hexene in
Polyethylenes’, Chem. Eng. Sci., 56, pp. 4121-4129.

Neto A.G. M., Freitas M.F., Nele M. and Pinto J.C., 2005, ‘Modeling Ethylene/1-Butene
Copolymerizations in Industrial Slurry Reactors’, Ind. Eng. Chem. Res., 44, pp.
2697-2715.
Chapter 4: Slurry-loop Reactor Modelling 211

Nicmanis M. and Hounslow M.J., 1998, ‘Finite-Element Methods for Steady-State


Population Balance Equations’, AIChE J., 44, pp. 2258-2272.

Noll P., 2005, Jan. 27, ‘Process for Preparing Polyethylene’, US Patent No 2005/ 0020784
A1.

Novak A., Bobak M., KoseK J., Banaszak B.J., Lo D.,Widya T., Ray, W.H and de Pablo
J.J., 2006, ‘Ethylene and 1-Hexene Sorption in LLDPE under Typical Gas-Phase
Reactor Conditions’, Journal of Applied Polymer Science, 100, pp. 1124-1136.

Orbey H., Bokis C.P. and Chen, C.C., 1998, ‘Equation of State Modeling of Phase
Equilibrium in the Low Density Polyethylene Process: The Sanchez-Lacombe,
Statistical Associating Fluid Theory and Polymer-Soave-Redlich-Kwong Equations
of State’, Ind. Eng. Chem. Res., 37, 4481-4491.

Park G.S., Shomate C.H., Kennedy W.D. and Crawford B.L. Jr., 1937, ‘The entropies of n-
butane and isobutane, with some heat capacity data for isobutane’, J. Chem. Phys., 5,
pp. 359-363.

Parrish WM.R., 1981, ‘Solubility of Isobutane in Two High-Density Polyethylene Polymer


Fluffs’, Journal of Applied Polymer Science, 26, pp. 2279-2291.

Passaglia E. and Kevorkian, H.K., 1963, ‘The heat capacity of linear and branched
polyethylene’, J. Appl. Poly. Sci., 7, pp. 119-132.

Pattamaprom C., Larson R.G., 2001, ‘Predicting the Linear Viscoelastic Properties of
Monodisperse and Polydisperse Polystyrenes and Polyethylenes’, Rheol. Acta, 40,
pp. 516-532.

Perry R. H. and Green, D., 1984, ‘Perry’s Chemical Engineer’s Handbook’, sixth edition,
McGraw-Hill International Editions, Chemical Engineering Series.

Pladis P., Kanellopoulos V., Chatzidoukas C. and Kiparissides C., 2008, ‘Effect of
Reaction Conditions and Catalyst Design on the Rheological Properties of
Polyolefins Produced in Gas-Phase Olefin Polymerization Reactors’, Macromol.
Theory Simul., 17, pp. 478-487.

Quijada R., Scipioni R.B., Mauler R.S., Galland G.B. and Miranda M.S.L., 1995,
‘Synthesis and Characterization of Ethylene-1-hexene Copolymers Using
Homogeneous Ziegler-Natta Catalysts’, Polymer Bulletin, 35, pp. 299-306.
212 Chapter 4: Slurry-loop Reactor Modelling

Quijada R., Rojas R., Mauler R.S. Galland G.B. and Scipioni R.B., 1997, ‘Study of the
Effect of the Monomer Pressure on the Copolymerization of Ethylene with 1-
Hexene’, Inc. J Appl Polym Sci, 64, pp. 2567-2574.

Ramkrishna D., 2000, ‘Population Balances: Theory and Applications to Particulate


Systems in Engineering’, Academic Press, San Diego, CA.

Randolph A. D., Larson M. A., 1971, ‘Theory of Particulate Processes: Analysis and
Techniques of Continuous Crystallization’, Academic Press, N.Y.

Ray W.H., 1986, ‘Modelling of polymerization phenomena’, Ber. Bunsengesellshaft Phy.


Chem., 90, pp. 947-955.

Ray W.H., 1991, ‘Modeling of Addition Polymerization Processes-Free Radical, Ionic,


Group Transfer, and Ziegler-Natta Kinetics’, Can. J. Chem. Eng., 69, pp. 626-629.

Reginato A.S., Zacca J.J. and Secchi A.R., 2003, ‘Modeling and Simulation of Propylene
Polymerization in Nonideal Loop Reactors’, AIChe Journal, 49, No. 10, pp. 2642-
2654.

Reid R.C., Prausnitz J.M. and Poling B.E., 1986, ‘The Properties of Gases and Liquids’,
Fourth edition, McGraw-Hill international Editions.

Rodgers P., 1993, ‘Pressure-Volume-Temperature Relationships for Polymeric Liquids: A


Review of Equations of State and Their Characteristic Parameters for 56 Polymers’,
J. Appl. Chem. Sci., 48, pp. 1061-1080.

Roy D., 1967, Ms. Thesis, Northwestern University, Evanston, III.

Salmon E.J., 2003, May 20, ‘Continuous Recovery of Polymer from a Slurry Loop
Reactor’, US Patent No. 6,566,460 B1.

Sanchez I.C. and Lacombe R.H., 1978, ‘Statistical Thermodynamics of Polymer


Solutions’, Macromolecules, 11, 6, pp. 1145-1156.

Sastry K.V.S. and Gaschignard P., 1981, ‘Discretization Procedure for the Coalescence
Equation of Particulate Processes’, Ind. Eng. Chem. Fundam., 20, pp. 355-361.

Scoggin J.S., 1966, Mar. 22, ‘Method and Apparatus for the Recovery of Solid Olefin
Polymer From a continuous Path Reaction Zone’, US Patent No. 3,242,150.
Chapter 4: Slurry-loop Reactor Modelling 213

Seborg D.E., Edgar T.F. and Mellichamp D.A., 1989, ‘Process Dynamics and Control’,
Wiley Series in Chemical Engineering, John Wiley and Sons, USA, Chapter 8, pp.
183-198, Chapter 13, pp. 294-309, Chapter 26, pp. 614-616.

Schwarzl F.R., 1971, ‘Numerical Calculation of Storage and Loss Modulus from Stress
Relaxation Data for Linear Viscoelastic Materials’, Rheol. Acta, 10, pp. 165-173.

Soares J.B.P. and Hamielec A.E., 1995 ‘Effect of Residence Time Distribution on the Size
Distribution of Polymer Particles Made With Heterogeneous Ziegler-Natta and
Supported Metallocene Catalysts. A Generic Mathematical Model’, Macromol.
Theory Simul., 4, pp. 1085-1104.

Thomas, D.G., 1965, ‘Transport characteristics of suspension: VIII. A Note on the


Viscosity of Newtonian Suspensions of Uniform Spherical Particles’, Journal of
Colloid Science, 20, pp. 267-277.

Touloupides V., Kanellopoulos V., Pladis P., Kiparissides C., Mignon D. and Van-
Grambezen P., 2010a, ‘Modeling and Simulation of an Industrial Slurry-phase
Catalytic Olefin Polymerization Reactor Series’, Chem. Eng. Sci., 65, pp. 3208-3222.

Touloupides V., Kanellopoulos V., Pladis P. and Kiparissides C., 2010b, ‘Modeling and
Simulation of Particle Size Distribution in Slurry-Phase Olefin Catalytic
Polymerization Industrial Loop Reactors’, Computer Aided Chemical Engineering,
28, pp. 43-48.

Tsenoglou C., 1991, ‘Molecular weight polydispersity effects on the viscoelasticity of


entangled linear polymers’, Macromolecules, 24, 1762-1767.

Ullmann’s Encyclopedia of Industrial Chemistry, 6th edition, 2002, John Wiley & Sons.

Uvarov, B.A. and Tsevetkova V.I., 1974, ‘Development of a Mathematical Model for
Controlling the Yield of Propylene Polymerization in Loop Reactors’, Polim.
Protsessy Appar. 165.

Van Ruymbeke E., Keunings R., Bailly C., 2002a, ‘Determination of the Molecular
Weight of Entagled Linear Polymers from Linear Viscoelasticity Data’, J. Non-
Newtonian Fluid Mech., 105, pp. 153-175.

Van Ruymbeke E., Keunings R., Stephenne V., Hagenaars A. and Bailly C., 2002b,
‘Evaluation of Reptation Models for Predicting the Linear Viscoelastic Properties of
Entangled Linear Polymers’, Macromolecules, 35, pp. 2689-2699.
214 Chapter 4: Slurry-loop Reactor Modelling

Villar M.A., Failla M.D., Quijada R., Mauler R.S., Valles E.M., Galland G.B. and
Quinzani L.M., 2001, ‘Rheological Characterization of Molten Ethylene-a-olefin
copolymers synthesized with Et[Ind]2ZrCl2/MAO catalyst’, Polymer, 42, pp. 9269-
9279.

Wang Z., Tham M.T. and Morris A.J., 1992, ‘Multilayer feedforward neural networks:
Canonical form approximation of nonlinearity’, Int. J. Control, 56, pp. 655-672.

Weimin Z., Weiduan S. and Hongsi W., 1991, ‘Modeling and Analyis of Propylene Loop
Reactor-Steady State Behavior and Operable Region’, Chemical Reaction
Engineering and Technology, 7, pp. 68-76.

Wunderlich B., 1962,a, ‘Motion in polyethylene. I. Temperature and crystallinity


dependence of the specific heat’, J. Chem. Phys., 37, pp. 1203-1207.

Wunderlich B., 1962,b, ‘Motion in polyethylene. III. The amorphous Polymer’, J. Chem.
Phys., 37, pp. 2429-2432.

Xie T.Y., Hamielec A.E., Wood P.E. and Woods, D.R., 1991, ‘Experimental Investigation
of Vinyl Chloride Polymerization at High Conversion: Mechanism, Kinetics and
Modeling’, Polymer, 32(3), pp. 537-557.

Yiannoulakis H., Yiagopoulos A. and Kiparissides C., 2001, ‘Recent developments in the
particle size distribution modeling of fluidized-bed olefin polymerization reactors’,
Chem. Eng. Sci., 56, pp. 917-925.

Yiagopoulos A., Yiannoulakis H., Dimos V. and Kiparissides C., 2001, ‘Heat and Mass
Transfer Phenomena During the Early Growth of a Catalyst Particle in Gas-phase
Olefin Polymerization: the Effect of Prepolymerization Temperature and Time’,
Chem. Eng. Sci., 56, pp. 3979-3995.

Zacca, J.J. and Ray, H. W., 1993, ‘Modelling of the liquid phase polymerization of olefins
in loop reactors’, Chem. Eng. Sci., Vol. 48, No. 22, pp. 3743-3765.

Zacca J.J., Debling J.A. and Ray W.H., 1996, ‘Reactor Residence Time Distribution
Effects on the Multistage Polymerization of Olefins-I. Basic Principles and
Illustrative Examples, Polypropylene’, Chem. Engng Sci., 51, pp. 4859-4886.

Zoller P. and Walsh D., 1995, ‘Standard pressure-volume-temperature data for polymers',
Technomic publishing, Lancaster, Basel, pp. 36-37.
Chapter 5: Results and Discussion 215

Chapter 5

5. Results and Discussion

5.1. Introduction

The comprehensive mathematical model described in Chapter 4: Slurry-loop


Reactor Modelling, was numerically solved to assess the effects of various process
operating conditions (i.e., ethylene, 1-hexene, iso-butane, hydrogen and catalyst inflow
rates, reactor temperature, feed composition, etc.) on key process variables (i.e., pressure,
liquid and polymer phase densities, molecular species outflow concentrations, etc.) as well
as on the polyolefin production rate and the associated molecular, rheological and
morphological polymer properties (i.e., Mn, Mw, MWD, complex viscosity, particle size
distribution, etc.). Moreover, the predictive capability of the model was validated by
accurately simulating the dynamic operation of an industrial slurry-loop reactor series. The
scaling of plant data provided by our industrial partner is deliberately concealed due to
confidentiality reasons.

Model tuning has been conducted in terms of plant data, provided by our industrial
partner. Model input variables comprises ethylene, 1-hexene and hydrogen inflow rates
policies, reactor temperature, pressure, solids concentration and ethylene off-gas
composition (regarding first reactor operation) setpoint policies as well as the policy on the
number of settling legs that will be in operation. It should be pointed out that in all
different simulation runs, the same set of kinetic, transport and other model parameters
were utilized.
216 Chapter 5: Results and Discussion

5.2. Simulation Results

5.2.1. Slurry-loop Reactor Modelling

The following simulation results correspond to the modelling of a typical slurry-loop


reactor, utilizing settling legs for product withdrawal. It should be noted that reactor
geometry corresponds to a typical industrial-scale reactor dimensions but not to the
dimensions of the actual slurry-loop reactor series of our industrial partner, presented in
section 5.2.2. Simulation of an Industrial Reactor. Moreover, monomers, diluents and
hydrogen inflow rates correspond to typical inflow rate values but not to the actual grade
recipes (as presented in the scaled-figures of the next section regarding the industrial loop
reactor.).

The volume of each reactor in the cascade-loop reactor configuration was assumed
to be equal to 45 m3 (i.e., tube diameter and tube length were assumed equal to 0.48 m and
248 m, respectively). The volume of each settling leg was set equal to 0.07 m3. Finally, for
the stable as well as dynamic operation of the simulated industrial slurry-phase cascade-
loop reactor series, the simulation of the various PID feedback controllers, shown in Figure
4.9, was included. In particular, the dynamic process model included the simulation of the
two temperature controllers, one ethylene mass fraction controller, two polymer solids
concentration controllers and two reactor pressure controllers.

5.2.1.1. Process Startup

To simulate a typical startup policy for the cascade-loop reactor series, the ethylene
inflow rates as well as the number of settling legs in operation in the first and second
reactor were assumed to follow the discrete dynamic profiles shown in Figure 5.1. That is,
at time t  0 , the two loop reactors were assumed to be filled with iso-butane and known
concentrations of ethylene and hydrogen. In fact, the initial mass fractions of ethylene and
hydrogen in the first and second loop reactor were assumed to be equal to 0.6 and 0.14 and
0.0 and 0.0, respectively. The initial pressure in each reactor was assumed to be equal the
respective set-point value of the two pressure controllers. The mass flowrate ratios of (1-
hexene to ethylene) and (hydrogen to ethylene) in the feed stream were set equal to 0.07
and 1.15 10-5 in the first reactor and 0.0 and 1.5 10-3 in the second reactor, respectively.
Chapter 5: Results and Discussion 217

It should be pointed out that during the process startup, the setpoint values of the
temperature controllers in the first and the second reactor of the series were set equal to
T  85o C and T  95o C , respectively. In addition, the setpoint values of the ethylene mass
fraction, solids concentration and reactor pressure controllers were kept constant and equal
to typical values employed in industrial slurry-phase cascade-loop reactors (Marechal,
2006, Noll, 2005).
218 Chapter 5: Results and Discussion

a.
12 tn/hr
10.5 tn/hr
Ethylene Inflow Rate

1.5 tn/hr

6
5 Settling Legs in Operation

4
Startup
3
First reactor
2

b.
12 tn/hr
10.5 tn/hr
Etylene Inflow Rate

0 tn/hr

6
Settling Legs in Operation

4 Startup

Second reactor
2

0 10 20 30 40
Time, hr

Figure 5.1: Dynamic evolution of the ethylene inflow rate and number of settling legs in
the first (a.) and second (b.) loop reactor of the series during process startup.
Chapter 5: Results and Discussion 219

In Figure 5.2 and Figure 5.3, the dynamic evolution of the pressure in the first and
second reactor is depicted. According to the semi-continuous operation of the settling legs,
each time that the operating pressure in the loop reactor exceeds an upper limit (i.e., see
dashed line in Figure 5.2 and Figure 5.3), the discharge valve of a settling leg opens so its
polymer-solids content is emptied from the leg. This results in a decrease of the pressure in
the loop reactor (i.e., lower limit in Figure 5.2 and Figure 5.3) that is followed by the
closing of the discharge valve of the settling leg. Subsequently, as fresh monomer(s) and
diluent are continuously fed into the reactor and new polymer is formed, the reactor
pressure gradually reaches again its maximum allowable value. Thus, the discharge valve
of the next settling leg opens. This on-off valve opening of the sequence of the settling legs
results in a periodic variation (i.e., ‘zig-zag’) of the pressure in the loop reactor (see Figure
5.2 and Figure 5.3).

45
45
First reactor
44
Reactor Pressure, bar

43

44 Upper limit
42
Set point
41
Lower limit
40
Reactor Pressure, bar

43 39
46900 46920 46940 46960 46980 47000
Time, sec

42

41

40

39
0 10 20 30 40
Time, hr

Figure 5.2: Dynamic evolution of the operating pressure in the first reactor.
220 Chapter 5: Results and Discussion

44 44

Second reactor 43

Reactor Pressure, bar


42

43 Upper limit 41

Set point
Dimensionless Reactor Pressure
40

Lower limit 39

42 38
46800 46840 46880 46920 46960 47000
Time, sec

41

40

39

38
0 10 20 30 40
Time, hr

Figure 5.3: Dynamic evolution of the operating pressure in the second reactor.

Figure 5.4 illustrates the dynamic evolution of the ethylene off-gas composition (i.e.,
% w/w ethylene mass fraction in the reactor-phase, except for the polymer) and the catalyst
inflow rate into the first reactor of the series. It should be pointed out that the ethylene
mass fraction in the reactor-phase in the first reactor is controlled by manipulating the
catalyst inflow rate. As can be seen, after approximately 30 hours of dynamic operation,
the reactor reaches its final steady-state value with respect to the ethylene concentration.
Note that the total ethylene concentration in the reactor-phase (i.e., liquid phase and
swollen polymer particles) is a very important process variable since high ethylene
concentrations can lead to the formation of a separate gas-phase in the reactor with
undesired effects regarding the reactor operability (i.e., appearance of a stagnant ethylene
gas phase, insufficient recirculation rate of the reaction mixture, appearance of cavitation
phenomena, etc.). On the other hand, there is no controller for the ethylene off-gas
composition in the second reactor (as there is no fresh catalyst feed in the second reactor).
Figure 5.5 illustrates the dynamic evolution of the ethylene off-gas composition in the
second reactor. As can be seen, ethylene off-gas composition reaches higher values
compared to the ones in the first reactor. Moreover, a delay in the polymerization rate,
leading to an increase of the ethylene off-gas composition is observed due to the fact that
while the reactor is fed with fresh ethylene and un-reacted ethylene, coming from the first
reactor, there is not enough catalyst in the reactor to propagate the polymerization.
Chapter 5: Results and Discussion 221

1.5 150
First reactor

Ethylene Off-gas Composition, % w/w


Controlled Variable

Catalyst Inflow Rate, kg/hr


1.0 100
Set point

0.5 50

Manipulated Variable

0.0 0
0 10 20 30 40
Time, hr

Figure 5.4: Dynamic evolution of the ethylene mass fraction and catalyst inflow rate in the
first reactor.

4
Second reactor
Ethylene Off-gas Composition, % w/w

0
0 10 20 30 40
Time, hr

Figure 5.5: Dynamic evolution of the ethylene mass fraction and catalyst inflow rate in the
second reactor.
222 Chapter 5: Results and Discussion

The dynamic evolution of the 1-hexene and hydrogen off-gas composition in the first
and second reactor of the series are illustrated in Figure 5.6 and Figure 5.7, respectively. It
should be pointed out that the 1-hexene and hydrogen mass fractions reach their final
steady-state values after approximately 15 hours of operation.

The dynamic evolution of the solids concentration and iso-butane inflow rate (i.e.,
manipulated variable) in the first and second reactor are depicted in Figure 5.8 and Figure
5.9, respectively. As can be seen, the solids concentration controller is inactive during the
first 6 hours of the startup because of the very low concentration of solids in the reactor. It
should be pointed out that, based on the selected reactor startup policy, after about 30
hours of operation the solids concentration in the first reactor reaches its final steady-state
value.

As has already been mentioned, in the reactor, two phases co-exist, namely a liquid
phase (including isobutene, ethylene, 1-hexene and hydrogen) and a solid-polymer phase
(including polymer and sorbed quantities of iso-butane, ethylene, 1-hexene and hydrogen).
The time evolution of the reactor, polymer and liquid phase density in the first and second
reactor are depicted in Figure 5.10 and Figure 5.11, respectively. As can be seen, regarding
the first reactor, the reactor-phase density at t  0 is equal to the density of the pure liquid
phase (i.e., liquid  440 kg / m3 ). However, as the solids concentration increases, the

density of the reaction mixture (i.e., liquid plus solids) in the first reactor reaches its final
value (i.e., b  640 kg / m3 ).
Chapter 5: Results and Discussion 223

5 0.04
First reactor

Hydrogen Off-gas Composition, % w/w


1-Hexene off-gas Composition, % w/w
4
0.03

0.02

0.01
1

0 0.00
0 10 20 30 40
Time, hr

Figure 5.6: Dynamic evolution of the 1-hexene and hydrogen off-gas composition in the
first reactor.

4.0 3
Second reactor

Hydrogen Off-gas Composition, % w/w


1-Hexene Off-gas Composition, % w/w

3.5

3.0

2
2.5

2.0

1.5
1

1.0

0.5

0.0 0
0 10 20 30 40
Time, hr
Figure 5.7: Dynamic evolution of the hydrogen mass fraction in the first reactor.
224 Chapter 5: Results and Discussion

60 20
First reactor

Iso-butane Inflow Rate, kg/hr x 10


50
Set-point
Solids Concentration, % w/w 15

40
Controlled Variable

30 10

20
Manipulated Variable
5

10

-3
0 0
0 10 20 30 40
Time, hr

Figure 5.8: Dynamic evolution of solids concentration and iso-butane inflow rate in the
first reactor.

60 15
Second reactor

50

Iso-butane Inflow Rate, kg/hr x 10


Set-point
Solids Concentration, % w/w

Controlled Variable
40 10

30

20 5
Manipulated Variable

10
-3

0 0
0 10 20 30 40
Time, hr

Figure 5.9: Dynamic evolution of the solids concentration and iso-butane inflow rate in
the second reactor.
Chapter 5: Results and Discussion 225

900
First reactor

800

Reactor phase
3
Density, kg/m

700 Polymer phase


Liquid phase

600

500

400
0 10 20 30 40
Time, hr

Figure 5.10: Dynamic evolution of density in the first reactor.

900

800
Second reactor

Slury phase
3
Density, kg/m

700 Polymer phase


Liquid phase

600

500

400
0 10 20 30 40
Time, hr

Figure 5.11: Dynamic evolution of density in the second reactor.


226 Chapter 5: Results and Discussion

Figure 5.12 depicts the dynamic evolution of the solids enhancement factor as well
as the valve opening cycling period of the settling legs in the first reactor during process
startup. It is apparent that the value of the enhancement factor (i.e., ef   out  s )

decreases exponentially to its final steady-state value (i.e., e f  1.25 ). This means that the

solids volume fraction in the product stream is about 25% higher than that in the reactor-
phase. Notice that during the process startup (i.e., the startup period) the number of settling
legs in operation changes from 2 to 6 (see Figure 5.1). At the same time, the valve opening
cycling period changes from about 100 seconds at the beginning of the startup to about 48
seconds at the final reactor steady-state. In Figure 5.13 the dynamic evolution of the solids
enhancement factor as well as the valve opening cycling period of the settling legs in the
second reactor during process startup is depicted. As can be seen, at t  10hours , there is
an abrupt increase in both enhancement factor and valve opening cycling period due to the
corresponding increase f the number of settling legs in operation (i.e., from 4 to 6, see
Figure 5.1).

3.0 100
First reactor

Valve Opening Cycling Period, sec


80
2.5
Enhancement Factor

60

2.0

40

1.5
20

1.0 0
0 10 20 30 40
Time, hr

Figure 5.12: Dynamic evolution of the solids enhancement factor and valve opening
cycling period in the first reactor.
Chapter 5: Results and Discussion 227

6.0 120
Second reactor
5.5

Valve Opening Cycling Period, sec


100
5.0

4.5
Enhancement Factor

80
4.0

3.5 60

3.0
40
2.5

2.0
20
1.5

1.0 0
0 10 20 30 40
Time, hr

Figure 5.13: Dynamic evolution of the solids enhancement factor and valve opening
cycling period in the second reactor.

In Figure 5.14-Figure 5.16 the dynamic evolution of temperature, coolant’s inlet


temperature into the reactor’s jacket and the time variation of the temperature difference
between the coolant outlet and inlet values, in the first and second reactor is depicted. It
should be pointed out that the reactor temperature is kept constant during its startup
operation by manipulating the coolant’s inlet temperature. It is apparent that as the
ethylene inflow rate increases, the polymerization rate increases and so does the heat
generation rate. As a result, in order to maintain an isothermal operation in the first loop
reactor the coolant’s inlet temperature decreases while the temperature difference between
the coolant outlet and inlet values increases.
228 Chapter 5: Results and Discussion

100

Second reactor's set-point


95
Coolant Inlet Temperature, C
o

90 Controlled Variable First reactor


Second reactor

First reactor's set-point


85

80

75
0 10 20 30 40
Time, hr

Figure 5.14: Dynamic evolution of temperature in the first and second reactor.

80 8
First reactor
Coolant Inlet Temperature, C
o

70
ΔT Coolant, C

4
o

60
Manipulated Variable 2

50 0
0 10 20 30 40
Time, hr

Figure 5.15: Dynamic evolution of the coolant’s inlet temperature and the temperature
difference between the coolant outlet and inlet values in the first reactor.
Chapter 5: Results and Discussion 229

100 8
Second reactor

o
Coolant Inlet Temperature, C 90
6

DT Coolant, C
80

70

o
Manipulated Variable 2
60

50 0
0 10 20 30 40
Time, hr

Figure 5.16: Dynamic evolution of the coolant’s inlet temperature and the temperature
difference between the coolant outlet and inlet values in the second reactor.

In Figure 5.17 the dynamic evolution of polymer production rate at the exit of the
first and second reactor of the series is depicted. As can be seen, polymer production rate
follows ethylene feed rate, exhibiting high conversion values (i.e., higher than 95%).

In Figure 5.18, the variation of the polyolefin density at the exit of the first and
second reactor (i.e., final product density, free of any sorbed solvent) is shown. As
expected, polymer produced in the first reactor exhibits lower values of density due to the
higher co-monomer content in the reactor’s feed. In the inserted figure, the polymer
crystallinity is presented.
230 Chapter 5: Results and Discussion

30
First reactor
Second reactor
25
Polymer Production Rate, tn/hr
20

15

10

0
0 10 20 30 40
Time, hr

Figure 5.17: Dynamic evolution of polymer production rate at the exit of the first and
second reactor of the series.

0.96
First reactor
Second reactor
Product Density, kg/lt

64
62
60
58
56
54
Crystallinity

0.94
52
First reactor
50
Second reactor
48
46
44
42
40
38
36

0 10 20 30 40
Time, hr

0.92
0 10 20 30 40
Time, hr

Figure 5.18: Dynamic evolution of product density at the exit of the first and second
reactor of the series.
Chapter 5: Results and Discussion 231

In Figure 5.19 and Figure 5.20 the dynamic evolution of the average molecular
weight and polydispersity at the exit of the first and second reactor is presented. Moreover,
in Figure 5.21 the dynamic evolution of the MWD of the polymer produced in the first
(Figure 5.21a.) and in the second reactor (Figure 5.21b.) and overall MWD at the exit of
the second reactor of the series (Figure 5.21c.) is depicted. As can be seen, the MWDs
produced in the first and the second reactor are relatively narrow. On the other hand, the
overall MWD at the exit of the second reactor is broader reflecting the formation of a low
molecular weight polymer fraction due to the high hydrogen concentration in the second
reactor. The MWD of the polymer produced by all different catalyst site types in the first
and the second reactor, regarding steady-state (i.e., at t  14hours ), is presented, in the
inserted figures of Figure 5.21a. and Figure 5.21b., respectively.

1.0
-5

First reactor
Number Average Molecular Weight, Mn, x10

Second reactor
0.8

0.6

0.4

0.2

0.0
0 10 20 30 40
Time, hr

Figure 5.19: Dynamic evolution of average molecular weight at the exit of the first and
second reactor of the series.
232 Chapter 5: Results and Discussion

35

30

25
Polydispersity

20
First reactor
Second reactor
15

10

0 10 20 30 40
Time, hr

Figure 5.20: Dynamic evolution of polydispersity at the exit of the first and second reactor
of the series.
Chapter 5: Results and Discussion 233

a. 0.8 0.8

2 hours Overall
Catalyst site type 1
Catalyst site type 2

6 hours 0.6 Catalyst site type 3


Catalyst site type 4

dW/dlog(MW)
14 hours 0.4

0.6 0.2

Polymer produced
in the first reactor 0.0

dW/dlog(MW)
2 3 4 5 6 7 8
log(MW)

0.4

Time

0.2

0.0
b. 0.8 0.8
Overall
2 hours Catalyst site type 1
Catalyst site type 2

6 hours 0.6 Catalyst site type 3


Catalyst site type 4

dW/dlog(MW)
14 hours 0.4

0.6
0.2
dW/dlog(MW)

0.0
2 3 4 5 6 7 8
log(MW)

0.4
Time

0.2

Polymer produced
in the second reactor
0.0
c. 0.8
2 hours Polymer at the exit
8 hours of the second reactor of the series
14 hours
0.6
dW/dlog(MW)

0.4 Time

0.2

0.0
2 3 4 5 6 7 8
log(MW)

Figure 5.21: Dynamic evolution of the MWD of the polymer produced in the first (a.) and
in the second reactor (b.) and overall MWD at the exit of the second reactor of the series
(c.) (startup case).
234 Chapter 5: Results and Discussion

In the present study, the complex viscosity of polyolefins produced in the slurry-
phase cascade-loop reactor series was calculated by employing a rheological model based
on the reptation and Rouse relaxation theories (Pladis et. al., 2008). The application of this
model enables the prediction of the rheological behaviour of polymer melts (i.e., complex
viscosity versus frequency) for linear polymer chains in terms of their respective MWD.
Figure 5.22 and Figure 5.23 depict the calculated rheological curves of the polymer
produced in the first reactor and polymer at the exit of the second reactor of the series,
respectively. It is evident that, the complex viscosity in the first reactor is higher than that
in the second reactor reflecting the higher molecular weight values of the polymer
produced in the first reactor. Moreover, the transition from the Newtonian plateau to the
shear thinning behaviour is more abrupt for narrow MWDs (i.e., see discrete lines in the
inserted figure in Figure 5.22).

In the following sections of the research project, the effect of the total number of
settling legs in operation, the effect of the co-monomer inflow rate and the effect of the
hydrogen inflow rate on the slurry-loop reactor series process is studied. It should be noted
that the simulations regarding all those effects investigation, begin at t  0hours ,
corresponding to the end of steady-state simulation, studied at 5.2.1. Process startup.
Chapter 5: Results and Discussion 235

6
10
2 hours
6 hours
14 hours
Complex Viscosity, η, Pa.s 5
10

4
10

3
10

2
10
1E-3 0.01 0.1 1 10 100 1000
Frequency, 1/sec

Figure 5.22: Dynamic evolution of the complex viscosity of the polymer produced in the
first reactor (startup case).

6
10
2 hours
8 hours
14 hours
5
Complex Viscosity, η, Pa.s

10

4
10

3
10

2
10
1E-3 0.01 0.1 1 10 100 1000
Frequency, 1/sec

Figure 5.23: Dynamic evolution of the complex viscosity of the polymer at the exit of the
second reactor (startup case).
236 Chapter 5: Results and Discussion

5.2.1.2. Effect of the Total Number of Settling Legs in Operation

It is well known that the selection of the number of settling legs in operation is
directly related with the overall process performance and product quality. For example, a
decrease in the number of settling legs in operation may occur due to either problematic
operation of settling legs (e.g., settling leg fouling due to extensive particle agglomeration,
etc.) or settling legs maintenance. In such a case, the non-operating leg is directly set out of
order and, thus, the reactor operates with a reduced number of settling legs.

In Figure 5.24, the dynamic evolution of the enhancement factor as well as of the
reactor valve opening cycling period (i.e., time period between two consecutive valve
openings of the same leg) are depicted. In this simulation, the number of settling legs in
operation in the second loop reactor of the series changes from 6 to 5 at time t  10hours
and finally to 4 at t  20hours . It can be seen that as the number of settling legs in
operation decreases the valve opening cycling period decreases. This results in a decrease
of the solids enhancement factor since the time period available for solids settling is
reduced. This in turn leads to a significant increase in the iso-butane inflow rate (see Figure
5.25) caused by the higher iso-butane removal rate (i.e., low values of e f ).

1.4 100

Valve Opening Cycling Period, sec


80
1.3
Enhancement Factor

60

1.2

40

1.1
6
20
5
4
Settling Legs in Operation
1.0 0
0 10 20 30
Time, hr

Figure 5.24: Dynamic evolution of the solids enhancement factor and valve opening
cycling period in the second reactor.
Chapter 5: Results and Discussion 237

48 12

Iso-butane Inflow Rate, kg/hr x 10


46
Controlled Variable
Solids Concentration, % w/w Set-point
10
44

42
8

Manipulated Variable
40

-3
38 6
0 10 20 30
Time, hr

Figure 5.25: Dynamic evolution of the solids concentration and iso-butane inflow rate in
the second reactor.

In Figure 5.26, the dynamic evolution of the number average molecular weight and
polydispersity index of the polymer produced in the second reactor of the series are
depicted. It is apparent that as the number of settling legs in operation decreases, the
number average molecular weight increases while the polydispersity index decreases due
to the increase of the ethylene concentration in the reactor (i.e., in both liquid and polymer
phase, see inserted figure in Figure 5.26). This can be attributed to the decrease of that the
ethylene conversion as a result of the decrease of the reaction mixture residence time (due
to the increase of the iso-butane inflow rate, see Figure 5.25).

In summary, the number of settling legs in operation affects both the reactor
performance (i.e., the iso-butane inflow rate in the second reactor changes from 7.5 x 103
kg/hr to 10.0 x 103 kg/hr when the number of settling legs in operation changes from 6 to
4, see Figure 5.25) as well as the molecular properties of the polyolefin (see Figure 5.26).
238 Chapter 5: Results and Discussion

1.0 32

-4
1.05

Number Average Molecular Weight, Mn x 10

Dimensionless Ethylene Mass Fraction


1.00

0.95

0.90 Reactor phase


Polymer phase
0.85

0.80 30
0.75

0.70

Polydispersity
0 10 Time, hr 20 30

0.9 28

26

0.8 24
0 10 20 30
Time, hr

Figure 5.26: Dynamic evolution of the number average molecular weight and
polydispersity index in the second reactor.

5.2.1.3. Effect of the Co-monomer Inflow Rate

In order to change the density of the final polymer product, a step change in the 1-
hexene inflow rate in the first reactor of the series was introduced. Figure 5.27 presents the
dynamic evolution of the co-polymer composition (i.e., in 1-hexene) in the first and second
reactor of the series. In this simulation, a transition from a polymer grade having a co-
polymer content equal to 0.5 % mol, at the exit of the second reactor of the series, to a
polymer grade with a co-polymer content equal to 0.95 % mol is depicted. Moreover, the
dynamic evolution of the co-polymer composition at the exit of the first reactor of the
series is illustrated (see dotted line in Figure 5.27). In the inserted figure, the time variation
of the mass fractions of 1-hexene in the liquid and polymer phase in the first reactor are
depicted, during the above grade transition.
Chapter 5: Results and Discussion 239

2
First reactor
Second reactor

1-Hexene Composition, % mol/mol

12
First reactor

Hexene Mass Fraction, % w/w


Liquid phase
Polymer phase
8

0
0 10 20 30
Time, hr
0
0 10 20 30
Time, hr

Figure 5.27: Dynamic evolution of the co-polymer composition in the first and second
reactor of the series.

In Figure 5.28, the variation of the polyolefin density in the first and second reactor
(i.e., final product density, free of any sorbed solvent) as well as the applied step-change in
the 1-hexene inflow rate (i.e., from 800 kg/hr to 1600 kg/hr) are depicted with respect to
the polymerization time. It is apparent that the product density at the exit of the second
reactor of the series is higher than that in the first reactor due to the lower co-monomer
incorporation rate in the second reactor. It should also be noted that the product density
reaches its new steady-state value in about 5 hours after the introduction of the step change
in the 1-hexene inflow rate. Note that, in the first reactor of the present slurry-phase
cascade-loop reactor series, low crystalline polymer (e.g., 35-40%) is produced due to the
higher co-monomer concentration, while, in the second reactor of the series, high
crystalline polymer (e.g., 55-60%) is produced due to the lower co-monomer incorporation
rate.
240 Chapter 5: Results and Discussion

0.96 5000

4000

1-Hexene Inflow Rate, kg/hr


First reactor
Product Density, kg/L

0.94 Second reactor

3000

0.92 2000

1000

0.90
0 10 20 30
Time, hr

Figure 5.28: Dynamic evolution of the product density in the first and second reactor of
the series.

5.2.1.4. Effect of the Hydrogen Inflow Rate

In this case, a step change in the hydrogen inflow rate was introduced to change the
final MWD of the polyolefin. More specifically, the hydrogen to ethylene mass flowrate
ratios in the first and second reactor of the series changed from 1.0 10-5 to 1.0 10-4 and
from 6.0 10-4 to 4.0 10-3, respectively. It is well known that, in Z-N catalytic olefin
polymerization, the hydrogen concentration is employed to control the molecular weight of
the polyolefin. In Figure 5.29 and Figure 5.30, the effect of the hydrogen concentration on
the dynamic evolution of the MWDs of polyolefin produced in the first and the second
reactor of the series are depicted. It is evident that as the hydrogen to ethylene mass
flowrate ratio increases (i.e., the hydrogen concentration increases), the MWD shifts to
lower molecular weights. The inserted figures in Figure 5.29 and Figure 5.30 show the
corresponding rheological curves during the imposed transition in the hydrogen inflow
rate. It is apparent that higher molecular weight polyolefins exhibit higher complex
viscosities and vice versa.
Chapter 5: Results and Discussion 241

0.8 10
6

Initial
Initial 2 hours

Complex Viscocity, η, Pa.s


10
5 4 hours
2 hours
4 hours 10
4

dW/dlog(MW) 0.6 10
3

2
10
1E-3 0.01 0.1 1 10 100 1000
Frequncy, 1/sec

0.4 Time

0.2

0.0
2 3 4 5 6 7 8
log(MW)

Figure 5.29: Dynamic evolution of the MWD and complex viscosity in the first reactor
(effect of hydrogen concentration).

0.8 6
10

Initial Initial
2 hours
Complex Viscocity, η, Pa.s

4 hours
2 hours
5
10

4 hours 10
4

0.6 10
3

2
10
dW/dlog(MW)

1E-3 0.01 0.1 1 10 100 1000


Frequncy, 1/sec

0.4
Time

0.2

0.0
2 3 4 5 6 7 8
log(MW)

Figure 5.30: Dynamic evolution of the MWD and complex viscosity in the second reactor
(effect of hydrogen concentration).
242 Chapter 5: Results and Discussion

5.2.2. Simulation of an Industrial Reactor

The model developed under this research project was employed for the modelling of
an industrial slurry-loop reactor. The predicting capability of the model has been tested by
simulating the actual reactor operation and comparing model’s predictions to plant
measurements. Model predictions were compared with actual plant data as provided by our
industrial partner. The following representative results correspond to a typical reactor start
up and a whole month operation which includes a grade transition. It should be noted that
the model has been tested over different operating conditions for the production of
different polymer grades, exhibiting great accuracy in all cases. Finally, it should be
mentioned that all results are presented scaled due to confidentiality issues.

5.2.2.1. Molecular Properties

The reactor simulation model’s predicting capability in terms of molecular properties


has been thoroughly tested over different MWD polymer grades, provided by our industrial
partner. It has been found that the multi-site Z-N kinetic scheme employed (using the
kinetic rate constants coming from the MWD deconvolution procedure), together with the
reactor model can describe the molecular developments occurring in the actual reactor. In
the following figures, representative results are presented, regarding 3 different polymer
grades in terms of co-monomer distribution and consequently different polymer density (of
the final polymer product at the exit of the second reactor of the series, outside the reactor).
Figure 5.31 corresponds to polymer produced having low 1-hexene content (polymer
density>0.96 kg/m3), Figure 5.32 corresponds to polymer produced having medium 1-
hexene content and Figure 5.33 corresponds to polymer produced having high 1-hexene
content (polymer density<0.95 Kg/m3). In all cases, it is clear that the simulation model
can reproduce the polymer MWD.
Chapter 5: Results and Discussion 243

0.8
First reactor, model prediction Grade 1
First reactor, plant data
Second reactor
Second reactor, plant data
dW/dlog(MW) 0.6

0.4

0.2

0.0
1 2 3 4 5 6 7 8
log(MW)

Figure 5.31: Molecular weight distribution of polymer at the exit of the first and second
reactor of the series, regarding a low-co-monomer content polymer grade.

0.8
First reactor, model prediction Grade 2
First reactor, plant data
Second reactor
Second reactor, plant data
0.6
dW/dlog(MW)

0.4

0.2

0.0
1 2 3 4 5 6 7 8
log(MW)

Figure 5.32: Molecular weight distribution of polymer at the exit of the first and second
reactor of the series, regarding a medium-co-monomer content polymer grade.
244 Chapter 5: Results and Discussion

0.8
First reactor, model prediction Grade 3
First reactor, plant data
Second reactor
Second reactor, plant data
0.6
dW/dlog(MW)

0.4

0.2

0.0
1 2 3 4 5 6 7 8
log(MW)

Figure 5.33: Molecular weight distribution of polymer at the exit of the first and second
reactor of the series, regarding a high-co-monomer content polymer grade.

5.2.2.2. Reactor Startup

Model’s capability to predict the dynamic response of the two loop-reactor series
process is tested over a typical startup of the reactor series. In Figure 5.34-Figure 5.36 the
reactor first reactor’s feed policies regarding ethylene, hexane and hydrogen are presented.
As can be seen, reactor feed is gradually increased to reach steady state operation (in terms
of polymer production rate) after 40 hours of operation. In Figure 5.37 the iso-butane feed
is depicted. It is noticed that iso-butane feed is manipulated in order to control reactor
solids concentration (see Figure 5.42). Manipulated catalyst feed is presented in Figure
5.38. As can be seen in Figure 5.39-Figure 5.42, the model is capable of accurately
predicting reactor composition, regarding ethylene, hexane, hydrogen and solids
concentration, leading to an accurate calculation of reactor phase density (Figure 5.43).
The dynamic evolution of temperature, coolant’s inlet temperature into the reactor’s jacket
and the time variation of the temperature difference between the coolant outlet and inlet
values can be predicted by the model (see Figure 5.44-Figure 5.46). The settling leg model
Chapter 5: Results and Discussion 245

presented in this research project is able to calculate the dynamic evolution of the valve
opening cycling period presented in Figure 5.47. Finally, the dynamic evolution of
polymer production rate and polymer density (i.e., final product density, free of any sorbed
solvent) is presented in Figure 5.48 and Figure 5.49, respectively. In Figure 5.50-Figure
5.63, simulation results regarding the second reactor of the series are presented.

The dynamic evolution of polymer density for the first and second reactor is
presented in Figure 5.49 and Figure 5.63, respectively. Moreover, the dynamic evolution of
number average molecular weight and polydispersity for the first and second reactor is
presented in Figure 5.64 and Figure 5.65, respectively. As can be seen in Figure 5.64, for
the first 4 hours of reactor series operation, the polymer produced in both reactors exhibit
the same number average molecular weight. This is attributed to the fact that the second
reactor is not fed with fresh catalyst but with catalyst coming from the first reactor outflow
and ‘live’ polymer which continues to polymerize. It is noted that the transition time for
polymer production with the defined end-use properties in terms of number average
molecular weight, polydispersity and density does not exceed 15 hours.

1.2
Startup

1.0
Scaled Ethylene Feed, Kg/hr

0.8

0.6

0.4

First reactor
0.2
Plant Data
Model Input
0.0
0 10 20 30 40 50
Time, hr

Figure 5.34: Dynamic evolution of ethylene feed in the first reactor (Startup case).
246 Chapter 5: Results and Discussion

1.2
Startup

1.0
Scaled Hexene Feed, Kg/hr
0.8

0.6

0.4

First reactor
0.2
Plant Data
Model Input
0.0
0 10 20 30 40 50
Time, hr

Figure 5.35: Dynamic evolution of hexane feed in the first reactor (Startup case).

1.2
Startup
Scaled Hydrogen Feed, 10 x Kg/hr

1.0
-3

0.8

0.6

0.4

First reactor
0.2
Plant Data
Model Input
0.0
0 10 20 30 40 50
Time, hr

Figure 5.36: Dynamic evolution of hydrogen feed in the first reactor (Startup case).
Chapter 5: Results and Discussion 247

1.4
Startup
1.2

Scaled Iso-butane Feed, Kg/hr


1.0

0.8

Manipulated Variable
0.6

0.4
First reactor
0.2
Plant Data
Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.37: Dynamic evolution of iso-butane feed in the first reactor (Startup case).

1.6
Startup
1.4
Scaled Catalyst Feed, Kg/hr

1.2

Manipulated variable
1.0

0.8

0.6

0.4 First reactor

0.2 Plant Data


Model Input
0.0
0 10 20 30 40 50
Time, hr

Figure 5.38: Dynamic evolution of catalyst feed in the first reactor (Startup case).
248 Chapter 5: Results and Discussion

Scaled Ethylene Off-gas Composition, % w/w


Startup First reactor

Plant Data
Model Prediction

Controlled Variable
Setpoint
1

0
0 10 20 30 40 50
Time, hr

Figure 5.39: Dynamic evolution of ethylene off-gas composition in the first reactor
(Startup case).

1.4
Startup
Scaled Hexene Off-gas Composition, % w/w

1.2

1.0

0.8

0.6

0.4
First reactor

0.2 Plant Data


Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.40: Dynamic evolution of hexane off –gas composition in the first reactor
(Startup case).
Chapter 5: Results and Discussion 249

16

Scaled Hydrogen Off-gas Composition, % w/w


Startup First reactor

Plant Data
Model Prediction
12

0
0 10 20 30 40 50
Time, hr
Figure 5.41: Dynamic evolution of hydrogen off-gas composition in the first reactor
(Startup case).

1.4
Startup
Scaled Solids Concentration, % w/w

1.2

Setpoint
1.0

Controlled Variable
0.8

0.6

0.4
First reactor

0.2 Plant Data


Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.42: Dynamic evolution of solids concentration in the first reactor (Startup case).
250 Chapter 5: Results and Discussion

1.2
Startup First reactor

3
Scaled Reactor Phase Density, Kg/m
1.1

1.0

0.9

0.8
Plant Data
Model Prediction
0.7
0 10 20 30 40 50
Time, hr

Figure 5.43: Dynamic evolution of reactor phase density in the first reactor (Startup case).

1.1
Startup
Scaled Temperature, C
o

Setpoint
1.0

Controlled Variable

First reactor

Plant Data
Model Prediction
0.9
0 10 20 30 40 50
Time, hr

Figure 5.44: Dynamic evolution of temperature in the first reactor (Startup case).
Chapter 5: Results and Discussion 251

1.6
Startup First reactor

Scaled Coolant Inlet Temperature, C


Plant Data
o Model Prediction
1.4

1.2

1.0
Manipulated Variable

0.8
0 10 20 30 40 50
Time, hr

Figure 5.45: Dynamic evolution of coolant’s inlet temperature in the first reactor (Startup
case).

1.4
Startup
1.2

1.0
Scaled ΔT coolant, C
o

0.8

0.6

0.4
First reactor

0.2 Plant Data


Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.46: Dynamic evolution of and the temperature difference between the coolant
outlet and inlet values in the first reactor (Startup case).
252 Chapter 5: Results and Discussion

2.0
Startup

Scaled Valve Opening Cycling Period, sec 1.5

1.0

0.5 First reactor

Plant Data
Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.47: Dynamic evolution of the valve opening cycling period in the first reactor
(Startup case).

1.2
Startup
Scaled Polymer Production Rate, Kg/hr

0.8

0.4

First reactor

Plant Data
Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.48: Dynamic evolution of polymer production rate in the first reactor (Startup
case).
Chapter 5: Results and Discussion 253

1.10
Startup
1.6

Scaled Crystallinity, %
1.4
First reactor
Scaled Product Density, kg/lt 1.2

1.0
1.05
0.8
0 10 20 30 40 50
Time, hr

1.00

Plant Data
Model Prediction
0.95
0 10 20 30 40 50
Time, hr

Figure 5.49: Dynamic evolution of product density in the first reactor (Startup case).

1.2
Startup
Scaled Ethylene Feed, kg/hr

0.8

0.4

Second reactor

Plant Data
Model Input
0.0
0 10 20 30 40 50
Time, hr

Figure 5.50: Dynamic evolution of ethylene feed in the second reactor (Startup case).
254 Chapter 5: Results and Discussion

0.18
Startup
0.16

0.14
Scaled Hydrogen Feed, Kg/hr

0.12

0.10

0.08

0.06

0.04 Second reactor

0.02 Plant Data


Model Input
0.00
0 10 20 30 40 50
Time, hr
Figure 5.51: Dynamic evolution of hydrogen feed in the second reactor (Startup case).

2.4
Startup Second reactor

2.0
Iso-butane Feed, kg/hr

1.6

1.2

Manipulated Variable
0.8

0.4
Plant Data
Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.52: Dynamic evolution of iso-butane feed in the second reactor (Startup case).
Chapter 5: Results and Discussion 255

Scaled Ethylene Off-gas Composition, % w/w


Startup Second reactor

Plant Data
Model Prediction
0
0 10 20 30 40 50
Time, hr

Figure 5.53: Dynamic evolution of ethylene off-gas composition in the first reactor
(Startup case).

0.8
Startup Second reactor
Hexene Off-gas Composition, % w/w

0.6

0.4

0.2

Plant Data
Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.54: Dynamic evolution of hexane off –gas composition in the second reactor
(Startup case).
256 Chapter 5: Results and Discussion

300

Scaled Hydrogen Off-gas Composition, % w/w


Startup Second reactor

250

200

150

100

50
Plant Data
Model Prediction
0
0 10 20 30 40 50
Time, hr

Figure 5.55: Dynamic evolution of hydrogen off-gas composition in the second reactor
(Startup case).

1.4
Startup Second reactor
1.2

Setpoint
%wt Solids Concentration

1.0

0.8
Controlled Variable

0.6

0.4

0.2 Plant Data


Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.56: Dynamic evolution of solids concentration in the second reactor (Startup
case).
Chapter 5: Results and Discussion 257

1.1
Startup Second reactor

3
Scaled Reactor Phase Density, Kg/m
1.0

0.9

0.8

0.7
Plant Data
Model Prediction
0.6
0 10 20 30 40 50
Time, hr
Figure 5.57: Dynamic evolution of reactor phase density in the second reactor (Startup
case).

1.20
Startup Second reactor

1.15
Scaled Temperature, C
o

Setpoint

1.10
Controlled Variable

1.05

Plant Data
Model Prediction
1.00
0 10 20 30 40 50
Time, hr

Figure 5.58: Dynamic evolution of temperature in the second reactor (Startup case).
258 Chapter 5: Results and Discussion

1.8
Startup Second reactor

Scaled Inlet Coolant Temperature, C


o 1.6

1.4

1.2

Manipulated Variable
1.0
Plant Data
Model Prediction
0.8
0 10 20 30 40 50
Time, hr

Figure 5.59: Dynamic evolution of coolant’s inlet temperature in the second reactor
(Startup case).

1.4
Startup Second reactor

1.2
Scaled ΔT Coolant, C

1.0
o

0.8

0.6

0.4

0.2 Plant Data


Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.60: Dynamic evolution of and the temperature difference between the coolant
outlet and inlet values in the second reactor (Startup case).
Chapter 5: Results and Discussion 259

2.4
Startup

Scaled Valve Opening Cycling Period, sec


2.0

1.6

1.2

0.8

Second reactor
0.4
Plant Data
Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.61: Dynamic evolution of the valve opening cycling period in the second reactor
(Startup case).

1.2
Startup
Scaled Polymer Production Rate, Kg/hr

0.8

0.4

Second reactor

Plant Data
Model Prediction
0.0
0 10 20 30 40 50
Time, hr

Figure 5.62: Dynamic evolution of polymer production rate in the second reactor (Startup
case).
260 Chapter 5: Results and Discussion

1.08
1.7
Startup
1.6

Scaled Crystallinity, %
1.5
Scaled Product Density, kg/lt 1.06
1.4

1.3

1.04 1.2
0 10 20 30 40 50
Time, hr

1.02

Second reactor
1.00
Plant Data
Model Prediction
0.98
0 10 20 30 40 50
Time, hr

Figure 5.63: Dynamic evolution of product density in the second reactor (Startup case).

1.4
Scaled Nunber Average Molecular Weight, Mn

Startup
1.2

1.0

0.8
First reactor
Second reactor
0.6

0.4

0.2

0.0
0 10 20 30 40 50
Time, hr

Figure 5.64: Dynamic evolution of number average molecular weight in the first and
second reactor (Startup case).
Chapter 5: Results and Discussion 261

5.0
Startup
4.5

Scaled Polydispersity 4.0

3.5

3.0 First reactor


Second reactor
2.5

2.0

1.5

1.0

0.5
0 10 20 30 40 50
Time, hr

Figure 5.65: Dynamic evolution of polydispersity in the first and second reactor (Startup
case).

The dynamic evolution of the particle size distribution in the first and second reactor
of the series, is illustrated in Figure 5.66 and Figure 5.67, respectively. It is important to
point out that during the process startup, the PSD moves to higher diameters as the system
approaches its final steady state operation. It can be seen that a steady state PSD is
achieved after approximately 14 hours of operation. As has already been mentioned,
polymer particles produced in the first reactor, are fed into the second reactor of the series
where polymerization continues without fresh catalyst feed. Thus, the PSD of the polymer
particles at the exit of the second reactor shifts to higher diameters, compared to the one in
the first reactor, as the same polymer particles, coming from the first reactor continue
growing. In Figure 5.68, the initial catalyst size distribution is depicted. It should be
mentioned that the initial catalyst size distribution is important for the PSD of the polymer
produced in the reactors.
262 Chapter 5: Results and Discussion

60
Startup First reactor

50 4 hours
6 hours
10 hours
40 14 hours
Pp, 1/cm

30

Time
20

10

0
200 400 600 800 1000
Diameter, μm

Figure 5.66: Dynamic evolution of the PSD in the first reactor (startup case).

70
Startup Second reactor
60 6 hours
10 hours
50
12 hours
14 hours
Pp, 1/cm

40

30
Time
20

10

0
200 400 600 800 1000
Diameter, μm

Figure 5.67: Dynamic evolution of the PSD in the second reactor (startup case).
Chapter 5: Results and Discussion 263

0.15
Catalyst Size Distribution

0.10
Pp, 1/cm

0.05

0.00
10 20 30 40 50
Diameter, μm

Figure 5.68: Initial catalyst size distribution (Startup case).

5.2.2.3. 28 Days Operation

In this simulation run, the reactor model simulates a whole month reactor series
operation, containing a grade transition. More specifically, the 1-hexene feed rate in the
first reactor is increases by over 6 times, for the production of less crystalline polymer
grade. At the same time, hydrogen feed in the first reactor is decreased in order to shift
MWD of polymer produced in the first reactor to higher molecular weights.

In Figure 5.69-Figure 5.71 the reactor first reactor’s feed policies regarding ethylene,
hexane and hydrogen are presented. As can be seen at t=11 days, 1-hexene feed rate shifts
to a higher value while hydrogen feed rate is lowered. In Figure 5.72, the dynamic
evolution of iso-butane feed rate is depicted. Iso-butane feed rate changes in order to
control the solids concentration to follow the desired set point policy (see Figure 5.77). The
dynamic evolution of reactor composition, regarding ethylene, 1-hexene and hydrogen is
presented in Figure 5.74-Figure 5.76. As expected, 1-hexene composition increases and
hydrogen composition decreases after the corresponding changes in the reactor’s feed rates
increase of 1-hexene feed in the first reactor at t=11days. The dynamic evolution of reactor
phase density is presented in Figure 5.78. The dynamic evolution of temperature, coolant
264 Chapter 5: Results and Discussion

inlet temperature as well as the temperature difference between the coolant outlet and inlet
values are depicted in Figure 5.79, Figure 5.80 and Figure 5.81, respectively. The dynamic
evolution of the valve opening cycling period, regarding the first reactor is shown in Figure
5.82. It is clear that the model is able to follow the variability exhibited by the valve
opening cycling period.

In Figure 5.83 the dynamic evolution of polymer production rate is depicted. The
dynamic evolution of polymer density for polymer produced in the first reactor is
presented in Figure 5.84. As can be seen, at t=11 days, the polymer density increases due
to the corresponding increase of the 1-hexene inflow rate. It should be noticed that the
polymer density reaches its new steady state 6 hours after the 1-hexene inflow rate step
change. It should be noticed that the density of the polymer produced in the first reactor
also affects the polymer at the exit of the second reactor (see Figure 5.98).

In Figure 5.85-Figure 5.98, simulation results regarding the second reactor of the
series are presented. The dynamic evolution of number average molecular weight and
polydispersity for the first and second reactor is presented in Figure 5.99 and Figure 5.100,
respectively. As can be seen, the step change in hydrogen inflow rate in the first clearly
affects the number average molecular weight of the polymer produced in the first reactor,
shifting it to higher values.
Chapter 5: Results and Discussion 265

2.0
28 days First reactor

1.6
Scaled Ethylene Feed, kg/hr

1.2

0.8

0.4
Plant Data
Model Input
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.69: Dynamic evolution of ethylene feed in the first reactor (28 days case).

2.4
28 days

2.0
Scaled Hexene Feed, kg/hr

1.6

1.2

0.8

First reactor
0.4
Plant Data
Model Input
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.70: Dynamic evolution of hexane feed in the first reactor (28 days case).
266 Chapter 5: Results and Discussion

4.0
28 days First reactor
3.5

Scaled Hydrogen Feed, 10 x Kg/hr


Plant Data
Model Prediction
3.0
-3

2.5

2.0

1.5

1.0

0.5

0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.71: Dynamic evolution of hydrogen feed in the first reactor (28 days case).

2.4
28 days First reactor

2.0
Scaled Iso-butane Feed, kg/hr

1.6

1.2
Manipulated Variable
0.8

0.4
Plant Data
Model Prediction
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.72: Dynamic evolution of iso-butane feed in the first reactor (28 days case).
Chapter 5: Results and Discussion 267

2.4
28 days First reactor

2.0
Scaled Catalyst Feed, kg/hr

1.6

1.2

0.8 Controlled Variable

0.4
Plant Data
Model Prediction
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.73: Dynamic evolution of catalyst feed in the first reactor (28 days case).

2.0
Sclaed Ethylene Off-gas Composition, % w/w

28 days First reactor

1.5

1.0

Controlled Variable
0.5

Plant Data
Model Prediction
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.74: Dynamic evolution of ethylene off-gas composition in the first reactor (28
days case).
268 Chapter 5: Results and Discussion

2.0
28 days

Sclaed Hexene Off-gas Composition, % w/w


1.6

1.2

0.8

First reactor
0.4
Plant Data
Model Prediction
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.75: Dynamic evolution of hexane off –gas composition in the first reactor (28
days case).

6
Scaled Hydrogen Off-gas Composition, % w/w

28 days First reactor

Plant Data
Model Prediction

0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.76: Dynamic evolution of hydrogen off-gas composition in the first reactor (28
days case).
Chapter 5: Results and Discussion 269

1.1
28 days First reactor

Scaled Solids Concentration, % w/w


Plant Data
Model Prediction

1.0

0.9
Controlled Variable

0.8
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.77: Dynamic evolution of solids concentration in the first reactor (28 days case).

1.1
28 days First reactor
3
Scaled Reactor Phase Density, Kg/m

1.0

Plant Data
Model Prediction
0.9
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.78: Dynamic evolution of reactor phase density in the first reactor (28 days case).
270 Chapter 5: Results and Discussion

1.05
28 days First reactor

Scaled Temperature, C
o

Controlled Variable

1.00

Plant Data
Model Prediction
0.95
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.79: Dynamic evolution of temperature in the first reactor (28 days case).

1.4
28 days First reactor
Scaled Coolant Inlet Temperature, C
o

1.3
Plant Data
Model Prediction
1.2

1.1

1.0

0.9

0.8

0.7

0.6
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.80: Dynamic evolution of coolant’s inlet temperature in the first reactor (28 days
case).
Chapter 5: Results and Discussion 271

2.0
28 days First reactor

Scaled ΔT Coolant, C 1.6


o

1.2

0.8

Plant Data
Model Prediction
0.4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.81: Dynamic evolution of and the temperature difference between the coolant
outlet and inlet values in the first reactor (28 days case).

2.0
28 days First reactor
Scaled Valve Openinfg Cycling Period, sec

1.6

1.2

0.8

0.4
Plant Data
Model Prediction
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.82: Dynamic evolution of the valve opening cycling period in the first reactor (28
days case).
272 Chapter 5: Results and Discussion

2.0
28 days First reactor

Scaled Polymer Production Rate, Kg/hr 1.6

1.2

0.8

Plant Data
Model Prediction
0.4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.83: Dynamic evolution of polymer production rate in the first reactor (28 days
case).

1.05
1.4

28 days
Scaled Crystallinity, %

1.2
Scaled Product Density, kg/lt

1.0

0.8
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

1.00

First reactor

Plant Data
Model Prediction
0.95
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.84: Dynamic evolution of product density in the first reactor (28 days case).
Chapter 5: Results and Discussion 273

2.0
28 days

Scaled Ethylene Feed, kg/hr 1.6

1.2

0.8 Second reactor

Plant Data
Model Input
0.4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.85: Dynamic evolution of ethylene feed in the second reactor (28 days case).

0.25
28 days

0.20
Scaled Hydrogen Feed, kg/hr

0.15

0.10

Second reactor
0.05
Plant Data
Model Input
0.00
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.86: Dynamic evolution of hydrogen feed in the second reactor (28 days case).
274 Chapter 5: Results and Discussion

2.0
28 days Second reactor

Scaled Iso-butane Feed, kg/hr 1.5


Manipulated Variable

1.0

0.5

Plant Data
Model Prediction
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.87: Dynamic evolution of iso-butane feed in the second reactor (28 days case).

5
28 days Second reactor
Scaled Ethylene Off-gas Compositon, % w/w

1
Plant Data
Model Prediction
0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.88: Dynamic evolution of ethylene off-gas composition in the first reactor (28
days case).
Chapter 5: Results and Discussion 275

1.0

Scaled Hexene Off-gas Composition, % w/w


28 days

0.8

0.6

0.4

Second reactor
0.2
Plant Data
Model Prediction
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.89: Dynamic evolution of hexane off –gas composition in the second reactor (28
days case).

300
Scaled Hydrogen Off-gas Composition, % w/w

28 days Second reactor

250

200

150

100

50
Plant Data
Model Prediction
0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.90: Dynamic evolution of hydrogen off-gas composition in the second reactor
(28 days case).
276 Chapter 5: Results and Discussion

1.05
28 days Second reactor

Scaled Solids Concentration, % w/w

1.00

0.95

Plant Data
Model Prediction
0.90
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.91: Dynamic evolution of solids concentration in the second reactor (28 days
case).

1.10
28 days Second reactor
3
Scaled Reactor Phase Density, kg/m

1.05

1.00

0.95

0.90

0.85
Plant Data
Model Prediction
0.80
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.92: Dynamic evolution of reactor phase density in the second reactor (28 days
case).
Chapter 5: Results and Discussion 277

1.15
28 days Second reactor

Scaled Temperature, C 1.14


o

1.13
Controlled Variable

1.12

1.11
Plant Data
Model Prediction
1.10
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, days

Figure 5.93: Dynamic evolution of temperature in the second reactor (28 days case).

1.4
28 days Second reactor
Scaled Coolant Inlet Temperature, C
o

1.2

Manipulated Variable
1.0

0.8

Plant Data
Model Prediction
0.6
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.94: Dynamic evolution of coolant’s inlet temperature in the second reactor (28
days case).
278 Chapter 5: Results and Discussion

2.0
28 days

1.5
Scaled ΔΤ Coolant, C
o

1.0 Manipulated Variable

0.5 Second reactor

Plant Data
Model Prediction
0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.95: Dynamic evolution of and the temperature difference between the coolant
outlet and inlet values in the second reactor (28 days case).

2.0
28 days Second reactor
Scaled Valve Opening Cycling Period, sec

1.6

1.2

0.8

Plant Data
Model Prediction
0.4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.96: Dynamic evolution of the valve opening cycling period in the second reactor
(28 days case).
Chapter 5: Results and Discussion 279

2.0
28 days Second reactor

Scaled Polymer Production Rate, kg/hr 1.6

1.2

0.8

Plant Data
Model Prediction
0.4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.97: Dynamic evolution of polymer production rate in the second reactor (28 days
case).

1.05
2.0
28 days
1.8
Scaled Crystallinity, %

1.04
Scaled Product Density, kg/lt

1.6

1.4

1.03 1.2
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

1.02

Second reactor
1.01
Plant Data
Model Prediction
1.00
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.98: Dynamic evolution of product density in the second reactor (28 days case).
280 Chapter 5: Results and Discussion

1.4

Scaled Number Average Molecular Weight, Mn


28 days
1.2

1.0

0.8

First reactor
0.6
Second reactor

0.4

0.2

0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.99: Dynamic evolution of number average molecular weight in the first and
second reactor (28 days case).

6
28 days

5
Scaled Polydisperity

3 First reactor
Second reactor

0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Time, Days

Figure 5.100: Dynamic evolution of polydispersity in the first and second reactor (28 days
case).
Chapter 5: Results and Discussion 281

5.3. References

Marechal P., 2006, Apr. 25, ‘Process for Producing Bimodal Polyethylene Resins’, US
Patent No. 7,034,092 B2.

Noll P., 2005, Jan. 27, ‘Process for Preparing Polyethylene’, US Patent No 2005/ 0020784
A1.

Pladis P., Kanellopoulos V., Chatzidoukas C. and Kiparissides C., 2008, ‘Effect of
Reaction Conditions and Catalyst Design on the Rheological Properties of
Polyolefins Produced in Gas-Phase Olefin Polymerization Reactors’, Macromol.
Theory Simul., 17, pp. 478-487.
282
Chapter 6: Concluding Remarks and Future Work 283

Chapter 6

6. Concluding Remarks and Future Work

6.1. Concluding Remarks

In the present research study, a comprehensive multi-scale multi-phase dynamic


mathematical model has been developed to describe the dynamic operation of an industrial
slurry-phase ethylene-1-hexene co-polymerization loop-reactor series. More specifically,
the effects of various operating conditions on the dynamic reactor behaviour (i.e., reaction
temperature and pressure, inflow rates of catalyst, monomers and diluent, etc.) as well as
the on the molecular (i.e., Mn, Mw, MWD, etc.), rheological (i.e., complex viscosity, etc.)
and morphological (i.e., PSD) polyolefin properties were fully assessed.

According to the proposed modelling approach, each loop reactor (i.e., consisting of
the loop reactor and the settling legs) is modelled as an ideal CSTR in series with a semi-
continuous product removal unit. Dynamic macroscopic mass and energy balances are
derived for each loop reactor in the series to predict the time variation of the concentrations
of the various molecular species as well as the reactor and jacket temperatures in the two
loop reactors. The non-continuous product withdrawal rate as well as the outflow species
concentrations from each reactor of the configuration is properly calculated via the solution
of a settling leg model.
284 Chapter 6: Concluding Remarks and Future Work

Moreover, a multi-site Z-N kinetic mechanism is employed to calculate the dynamic


evolution of the average and distributed molecular properties (i.e., M n , M w , MWD) of
the polyolefin in each reactor of the series, in conjunction with the well-known method of
moments. In order to determine the minimum number of different catalyst active sites for
the reconstruction of the experimentally measured MWDs, a MWD deconvolution analysis
was also employed.

The Sanchez-Lacombe EOS is utilized to calculate the thermodynamic equilibrium


concentrations of the various molecular species (i.e., monomer(s) diluent, hydrogen, etc) in
the two phases (i.e., solids and liquid) present in the reaction mixture. Finally, all different
control loops, typically employed by an industrial slurry-phase loop reactor, are
incorporated in the reactor model.

Extensive numerical simulations were carried out to assess the effects of process
operating conditions on key process variables as well as on the molecular, rheological and
morphological properties of polyolefins. The results were qualitatively consistent with the
expected behavior for the process variables and polymer properties. Moreover, it was
found that model predictions were in excellent agreement (quantitatively) with dynamic
measurements under different plant operating policies (i.e., startup, grade transition, etc.)
obtained from an industrial slurry-phase catalytic olefin polymerization loop-reactor series.
It was demonstrated that the present model can simulate bimodal MWDs produced in
multistage reactor configurations. The proposed model should find wide application in the
design, optimization and control of industrial polyolefin slurry-phase loop-reactor
processes.

6.2. Suggestions for Future Work

New research topics can be suggested as future work in the field.

 A recommended area of future work is the extension of the current model,


assuming perfect mixing conditions in the reactor tube to a multi-zone reactor
model, in order to be able to calculate the variation of the various species
concentrations, temperature, molecular and other properties of interest with
respect to the reactor length (i.e., axial reactor direction). This will help to
elucidate the effect of various reactor(s) operating conditions, startup policies
Chapter 6: Concluding Remarks and Future Work 285

(e.g., catalyst feed rate, etc.) and mixing conditions on the molecular and
morphological properties as well as on the molecular species dispersion
characteristics, energy requirements (i.e., recirculation pump power
consumption, pressure losses along the reactor length, etc). Moreover, the
employed multi-zone modelling approach will be capable of evaluating various
reactor instabilities and problematic operations in the slurry loop reactor
process.
 Particle agglomeration phenomena as well as the operability of the slurry loop
reactors largely depend on the comprehension of the growth of single catalyst/
polymer particles. Particle agglomeration and/ or attrition phenomena in the
slurry loop reactors can be further studied. To achieve this, a single particle
model should be employed to investigate if possible hot spots formation (i.e.,
particle agglomerates) in the reactor zones may take place (i.e., axial reactor
locations) taking into account the local reactor operating and mixing conditions
along with the particle size distribution model.
 Application of model predictive control (i.e., MPC) on the developed reactor
model to define the optimum reactor policy (regarding reactor feed rates,
controllers’ setpoints, etc.) in order to minimize grade transition times or reach a
preferred steady state under the most advantageous reactor policy route.
CV of V. Touloupides i

PERSONAL INFORMATION
Name Vassileios Touloupides
Address Auxendiou Str. 3, Triandria 55337, Thessaloniki, Greece
Phone Number +30-2310-922469 (Home) +30-6972-268369 (Mobile)
Email v.touloupides@gmail.com, vast@cperi.certh.gr
Nationality Greek
Date of Birth 16.05.1982
Place of Birth Thessaloniki/ Greece
Marital Status Single
EDUCATION
2009- Aristotle University of Thessaloniki (Α.U.TH.)
Medical School
2005-10 Ph.D in Chemical Engineering
Aristotle University of Thessaloniki (Α.U.TH.)
Department of Chemical Engineering, Faculty of Engineering.
Ph.D Thesis: ‘Mathematical Modelling and Simulation of an
Industrial α-Olefins Catalytic Slurry Phase Loop-reactor Series’
Corresponding Professor: Costas Κiparissides, Professor of Chemical
Engineering, Aristotle University of Thessaloniki.
2000-05 Diploma in Chemical Engineering
Aristotle University of Thessaloniki (Α.U.TH.)
Department of Chemical Engineering, Faculty of Engineering.
Graduation Degree: 8.5/10.0 Ranking: First within 110 Graduated
Students.
Diploma Thesis: ‘MATLAB Applications in Chemical
Engineering’
Corresponding Professor: Costas Κiparissides, Professor of Chemical
Engineering, Aristotle University of Thessaloniki.
EXPERIENCE
2004-05 Technical & Economical Study on Ethylene-oxide Production Plant,
Department of Chemical Engineering, Faculty of Engineering
(A.U.TH.).
July-August 2004 Training on process dynamics, Centre for Research and Technology
Hellas.
AWARDS & GRANTS
2011 ‘Borealis Student Innovation Award’ for the best doctoral graduate
thesis.
2005-10 Scholarship from Centre for Research and Technology Hellas.
2005 Award for Excellent Academic Achievements from the Technical
Chamber of Greece.
2005 Award for Excellent Academic Achievements from the Pan-Hellenic
Association of Chemical Engineers.
2004 Award for Excellent Academic Achievements from the Department
of Chemical Engineering, Aristotle University of Thessaloniki.
2004-05
2003-04 Scholarship from the National Institute of Scholarships.
2000-01
ii CV of V. Touloupides

INDUSTRIAL RESEARCH PROJECTS


2009-10 ‘Multi-Scale Multi-Zone Mathematical Modeling of Industrial
Polyethylene Slurry Loop Reactor(s) for Ziegler-Natta Catalysts’-
TOTAL PETROCHEMICALS
2007-09 ‘Mathematical Modeling of Industrial Polyethylene Slurry Loop
Reactor(s) for Ziegler-Natta Catalysts’-TOTAL PETROCHEMICALS.
PUBLICATIONS
 Publications: 6
Touloupides V., Kanellopoulos V., Orfanidou K. and Kiparissides C., 2010,
‘Mathematical Modeling and Simulation of Gas- and Slurry-Phase Catalytic
Olefin Polymerization Reactors’, to be submitted to Molecules.
Touloupides V., Kanellopoulos V., Krallis A., Pladis P. and Kiparissides C., 2010,
‘Modeling and Simulation of Particle Size Distribution in Slurry-Phase
Olefin Catalytic Polymerization Industrial Loop Reactors’, Computer Aided
Chemical Engineering, 28, pp. 43-48.
Krallis A., Pladis P., Kanellopoulos V., Saliakas V., Touloupides V. and
Kiparissides C., 2010, ‘Design, Simulation and Optimization of
Polymerization Processes Using Advanced Open Architecture Software
Tools’, Computer Aided Chemical Engineering, 28, pp. 955-960.
Touloupides V., Kanellopoulos V., Pladis P., Kiparissides C., Mignon D. and
Van-Grambezen P., 2010, ‘Modeling and Simulation of an Industrial Slurry-
phase Catalytic Olefin Polymerization Reactor Series’, Chem. Eng. Sci., 65,
pp. 3208-3222.
Kanellopoulos V., Tsiliopoulou E., Dompazis G., Touloupides V. and
Kiparissides C., 2007, ‘Evaluation of Morphology of Polymer Particles
Produced in Catalytic Gas-phase Olefin Polymerization Reactors’, Ind. Eng.
& Chem. Res., 46, pp. 1928-1932.
Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C., 2007,
‘Development of a Multi-scale, Multi-phase Multi-zone Dynamic Model for
the Prediction of Particle Segregation in Catalytic Olefin Polymerization
FBRs’, Chem. Eng. Sci., 63, pp. 4735-4753.
 Conference Proceedings: 15 (Oral & Poster Presentations)
Touloupides V., Kanellopoulos V. and Kiparissides C., ‘Prediction of Dynamic
Particle Size Distribution in Industrial Slurry-Phase Olefin Catalytic
Polymerization Loop Reactors’, AIChE 10/ Annual Meeting, 7-12
November, 2010, Salt Lake City, UT.
Touloupides V., Kanellopoulos V., Pladis P. and Kiparissides C., ‘Modeling and
Simulation of Particle Size Distribution in Slurry-Phase Olefin Catalytic
Polymerization Industrial Loop Reactors’, 20th European Symposium on
Computer Aided Process Engineering (ESCAPE 20), 6-9 June, 2010, Ischia,
Naples, Italy.
Krallis A., Pladis P., Kanellopoulos V., Saliakas V., Touloupides V. and
Kiparissides C., ‘Design, Simulation and Optimization of Polymerization
Processes Using Advanced Open Architecture Software Tools’, 20th
European Symposium on Computer Aided Process Engineering (ESCAPE
20), 6-9 June, 2010, Ischia, Naples, Italy.
Touloupides V., Kanellopoulos V., Pladis P., Krallis A. and Kiparissides C.,
‘Modeling and Simulation of Slurry-Phase Olefin Catalytic Polymerization
CV of V. Touloupides iii

Industrial Loop Reactors. Prediction of Polymer Particle Growth,


Molecular, Rheological and Morphological Polymer Properties’, AIChE 09/
Annual Meeting, November 8-13, 2009, Nashville, USA.
Touloupides V., Kanellopoulos V., Pladis P., Krallis A. and Kiparissides C.,
‘Real-time Simulation of an Industrial Scale Slurry Loop Reactor’, 8th
World Congress of Chemical Engineering (WCCE8), 23-27 August, 2009,
Montreal, Canada.
Touloupides V., Kanellopoulos V., Pladis P., Krallis A. and Kiparissides C.,
‘Dynamic Simulation of an Industrial Catalytic Slurry-loop Polymerization
Reactor’ 7th Panhellenic Chemical Engineering Conference, 3-5 June, 2009,
Patras, Greece.
Touloupides V., Kanellopoulos V., Pladis P., Krallis A. and Kiparissides C.,
‘Modeling of Industrial Catalytic Olefin Polymerization Slurry Reactors’,
7th Panhellenic Polymer Conference, 28 Sep.-1 Oct., 2008, Ioannina,
Greece.
Touloupides V., Kanellopoulos V., Pladis P. and Kiparissides C., ‘Mathematical
Modelling of an Industrial Catalytic Olegin Copolymerzation Slurry-Loop
Reactor’, 1st Intrnational Conference on the Reaction Engineering of
Polyolefins (INCOREP), June 22-27, 2008, Montreal, Canada.
Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C.,
‘Development of a Dynamic Multi-compartment Model for the Prediction of
Particle Size Distribution and Molecular Properties in a Catalytic Olefin
Polymerization FBR’, 3rd International Conference on Population Balance
Modeling, September 19-21, 2007, Quebec City, Canada.
Touloupides V., Kanellopoulos V., Dompazis G. and Kiparissides C.,
‘Development of a Multi-scale Dynamic Model for the Prediction of
Polymer Distributed Properties in Catalytic Olefin Polymerization Slurry
Loop Reactors’, European Congress of Chemical Engineering (ECCE-6),
16-20 September, 2007, Copenhagen, Denmark.
Touloupides V., Dompazis G., Kanellopoulos V. and Kiparissides C., ‘Dynamic
Simulation of Olefin Polymerization in Gas- and Slurry-phase reactors’, 6th
Panhellenic Chemical Engineering Conference, 31 May-2 June, 2007,
Athens, Greece.
Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C., ‘Multi-
scale, Multi-phase, Multi-zone Dynamic Model for the Prediction of Particle
Segregation in Catalytic Olefin Polymerization FBRs’, Hangzhou
International Polymer Forum, Advanced Materials and Reaction
Engineering, May 13-17, 2007, Hangzhou, China.
Kanellopoulos V., Gustafsson B., Touloupides V. and Kiparissides C., ‘Gas-phase
Olefin Polymerization in the Presence of Supported and Self-supported
Ziegler-Natta Catalysts’, Hangzhou International Polymer Forum,
Advanced Materials and Reaction Engineering, May 13-17, 2007,
Hangzhou, China.
Dompazis G., Kanellopoulos V., Touloupides V., and Kiparissides C.,
‘Development of a Multi-Scale, Multi-Phase, Multi-Zone Dynamic Model
for the Prediction of Particle Segregation in Catalytic Olefin Polymerization
Fbrs’, AIChE 06/ Annual Meeting, Nov. 12-17, 2006, San Francisco.
Dompazis G., Kanellopoulos V., Touloupides V. and Kiparissides C.,
‘Development of a Multi-scale, Multi-phase, Multi-zone Dynamic Model
iv CV of V. Touloupides

for the Prediction of Polymer Distributed Properties in Catalytic Olefin


Polymerization FBRs’, 6th Panhellenic Polymer Conference, 3-5 November,
2006, Patras, Greece.
LANGUAGE PROFICIENCY
 Greek (Native).
 English (Fluent, Level: Advanced, Certificated).
 Spanish (Fluent, Level: Advanced, Certificated).
 French (Basic).

COMPUTER PROGRAMS
Excellent Knowledge of:
 WINDOWS 2000/NT/XP/VISTA/7 Operation Systems and Internet
Applications,
 MICROSOFT OFFICE,
 FORTRAN 77/90, MICROSOFT DEVELOPER STUDIO,
MATLAB,
 ORIGIN, PHOTOSHOP and VISIO Software Packages.

FURTHER INFORMATION
Member of the Technical Chamber of Greece (2005).
Member of the Panhellenic Association of Chemical Engineers
(2007).
Driving License.
Personal Interests:
Playing Guitar, Music, Medicine, Gardening, Reading, Travel.

You might also like