Sei sulla pagina 1di 469

Emulsion polymerization

@i; PROYECTOCONACYT

Contributors
A. Berge
D. C. Blackley
A. S. Dunn
T. Ellingsen
Carlton G. Force
Robert G. Gilbert
A. E. Hamielec
F. K. Hansen
A. A. Khan
Gottfried Lichti
J. F. MacGregor
P. C. M0rk
Donald H. Napper
Mamoru Nomura
R. H. Ottewill
Gary W. Poehlein
Vivian T. Stannett
J. Ugelst~d
V. 1. Yeliseyeva

'.

EMULSION

POLYMERIZATION

Editad by

IRJA PIIRMA
Institute of Polymer Science
The Unversity of Akron
Akron, Oha

ACADEMIC

PRESS

A Subsidiary of Harcourt Brace Jovanovich, Publishers

New York London -Toronto Sydney San Francisco 1982

----

COPYRIGHT @ 1982, BY ACADEMIC PRESS, INC.


ALL RIGHTS RESERVED.
NO PART OF THIS PUBLlCATION MAY BE REPRODUCED OR
TRANSMITTED IN ANY FORM OR BY ANY MEANS, ELECTRONIC
OR MECHANICAL, INCLUDING PHOTOCOPY, RECORDING, OR ANY
INFORMATION STORAGEANO RETRlEVAL SYSTEM, WITHOUT
PERMISSION IN WRITING FROM THE PUBLlSHER.

ACADEMIC PRESS, INC.

111 Fifth Avenue, New York, New York 10003

United Kingdom Edition published by


ACADEMIC PRESS, INC. (LONDON)

24/28 Oval Road, London NWI

7DX

Ubrary of Congress cataloging


Main entry under ti tle:

LTD.

in Publication
.

Data

Emulsion polymerization.
Includes bibliographies
and index.
l. Emulsion polymerization.
l. Piirma,
QD)B2.E48E48
660.2'8448
ISBN 0-12-556420-1

81-17626
AACR2

PRINTED IN THE UNITED STATES OF AMERICA


82 83 84 85

9 8 7 6 S 4 3 2 J

Irja,

Contents
Contributors
Preface

ix
xi

1 The Stability and Instability of Polymer Latices


R. H. Ottewi/l
1. Introduction
11. The Nature of Polymer Latex Particles
111. The Effect of Electrolytes on a Latex
IV. The Theory of the Stability of Lyophobic Colloids
V. Coagulation as a Kinetic Process
VI. An Alternative Approach to the Critical Coagulation Concentration
VII. The Determination of ccc Values
VIII. The Effect of lons That Interact with Water
IX. Secondary Minimum Effects
X. The Effects of Organic lons: Added Surfactants
XI. lonic Head Group with a Charge of the Same Sign as the Particle
XII. lonic Head Group with a Charge of Opposite Sign to the Particle
XIII. Nonionic Surfactants
XIV. Mixed Electrolyte Systems
xv.
Heterocoagulation
XVI. Surface Coagulation
XVII. Peptization
XVIII. The Effects of Adsorbed or Grafted Macromolecules
XIX. Particle Stability in Emulsion Polymerization
XX. Summary
References

1
2
6
8
14
16
17
19
22
26
27
28
31
35
36
39
40
42
45
47
47

2 Particle Formation Mechanisms


F. K. Hansen and J. Ugelstad

1.

Introduction

11. Micellar Nucleation: The Smith-Ewart


111. Radical Absorption Mechanisms
IV. Micellar Nucleation: Newer Models
V.
Homogeneous Nucleation
VI.
VII.

Theory

Particle Coagulation during the Formation Period


Nucleation in Monomer Droplets
References

51
54
56
63
73
82
86'
91

Contents

vi

3 Theoretical Predictions of the Particle Size and Molecular Weight


Distributions in Emulsion Polymerizations
Gottfried Lichti, Robert G. Gilbert, and Donald H. Napper
1.
11.

Prediction of the PSD


Molecular Weight Distributions
111. Separability of MWD and PSD
IV. Conclusions
References

4 Theory ofKinetics ofCompartmentalized


Reactions

94
115
141
142
143

Free-Radical Polymerization

D. C. Blackley
1. Introduction
11. Reaction Model Assumed
111. The Time-Dependent Smith-Ewan Differential Difference Equations:
Methods Available for Their Solution
IV. Solution for the Steady State
V. Solutions for the Nonsteady State
VI. Predictions for Molecular Weight Distribution and Locus-Size
Distribution
VII. Theory for Generation of Radicals in Pairs within Loci
List of Symbols
References

146
149
156
164
167
183
185
187
189

Desorption and Reabsorption of Free Radicals


in Emulsion Polymerization
Mamoru Nomura

1. Introduction
11. Polymerization Rate Equations Involving Free-Radical Desorption
111. Derivation of Rate Coefficient for Radical Desorption from Panicles
IV. Effect of Free-Radical
Desorption on the Kinetics of Emulsion
Polymerization
List of Symbols
References

191
192
199
210
217
219

6 Effects of the Choice of Emulsifier in Emulsion Polymerization


A. S. Dunn
1. Introduction
Monomer Emulsification
11.
111. Emulsion Polymerization with Nonionic Emulsifiers
IV. Emulsion Polymerization with lonic Emulsifiers
V. Latex Agglomeration
Other Effects of Emulsifiers
JVI.
References

221
224
229
230
236
237
243

Contents

vii

7 Polyrnerization

of Polar Monorners

V. l. Yeliseyeva
1. Introduction
11. Interface Characteristics of Polymeric Dispersions
111. Relationship between Emulsifier Adsorption and the Difference in the
Boundary-Phase Polarity
IV. Mechanism of Particle Generation
V. Colloidal Behavior of Polymerization Systems
VI. Kinetics of Emulsifier Adsorption
VII. Mechanism of Formation and Structure of Particles
VIII. Polymerization Kinetics
IX. Relationship
between
Polymerization
Kinetics and Adsorption
Characteristic of the Interface
Nomenclature
References

247
249
250
257
261
268
27q
278
283
286
287

8 Recent Developrnents and Trends in the Industrial Use of Latex


Car/ton G. Force
1.
11.
111.
IV.
V.
VI.

289
291
300
312
313
314
316

Introduction
Factors in Adhesion
Bonding Applications
Construction Applications
Rubber Goods
Properties of Various Latexes
References

9 Latex Reactor Principies:

Design, Oneration,

and Control

A. E. Hamie/ec
and J. F. MacGregor
1. Introduction
11. Batch Reactors
111. Continuous Stirred-Tank Reactors: Steady-State Dperation
IV. ContinuQus Stirred-Tank Reactors: Dynamic Behavior
V. Dn-Une Control of Continuous Latex Reactors
VI. Summary
Nomenclature
References

10

319
320
333
339
345
351
351
353

Ernulsion Polyrnerization in Continuous Reactors.


Gary W. Poeh/ein

1. Introduction

. 11.

Smith-Ewart Case 2 Model for a CSTR


111. Deviations from Smith- Ewart Case 2
IV. Transient Behavior bf CSTR Systems
V. Strategies for Process and Product Development
VI. Summary
References

357
361
367
375
378
381
381

Contents

viii

11

Effect of Additives on the Formation


and Polymer Dispersions

of Monomer

Emulsions

J. Uglestad, P. C. MrjJrk, A. Berge, T. Ellingsen, and A. A.Khan


1. Introduction
11. ThermodynamicTreatment of Swelling and Phase Distributions
111. Rate of Interphase Transport
IV. Preparation of Polymer Dispersions
V. Effectof Additionof Water-InsolubleCompounds to the Monomer Phase
VI. Emulsificationwith Mixed EmulsifierSystems
Listof Symb91s
References

383
384
392
396
401
408
411
412

12 Radiation-Induced Emulsion Polymerization


Vivan T. Stannett
1. Introduction
11. Laboratory Results with Different Monomers
111. Copolymerizations
IV. Radiation-Induced Emulsion Polymerization Using Electron Accelerators
V. Pilot Plant and Related Studies
References

lndex

415
418
433
436
437
447

451

Contributors
Numbers in parentheses indicate the pages on which the authors' contributions begin.

A. Berge (383), SINTEF, Applied Chemistry Division, 7034-NTH,


Trondheim, Norway
D. C. Blackley (145), National College of Rubber Technology, The Polytechnic of North London, Holloway, London N7 8DB, England
A. S. Dunn (221), Chemistry Department, University of Manchester
Institute of Science and Technology, Manchester M60 lQD, England
T. Ellingsen (383), SINTEF, Applied Chemistry Division, 7034-NTH,
Trondheim, Norway
Carlton G. Force (289), Westvaco Corporation, Research Center, North
Charleston, South Carolina 29406
Robert G. Gilbert (93), Departments of Physical and Theoretical Chemistry,
University of Sydney, New South Wales 2006, Australia
A. E. Hamielec (319), Department.of Chemical Engineering, McMaster
University, Hamilton, Ontario L8S 4L7, Canada
F. K. Hansen (51), DYNO Industrier, Lillestr~m Fabrikker, N-2001
Lillestrs6m, Norway
A. A. Khan (383), E. 1. du Pont de Nemours & Co., Polymer Products
Department, Experimental Station, Wilmington, Delaware 19898
Gottfried Lichti (93), Australian Institute of Nuclear Science and Engineering, New South Wales, Australia
J. F. MacGregor (319), Department of Chemical Engineering, McMaster
University, Hamilton, Ontario L8S 4L7, Canada
P. C. Mr)rk (383), Laboratory of Industrial Chemistry, The University of
Trondheim, N-7034 Trondheim, Norway
Donald H. Napper (93), Department ofPhysical and Theoretical Chemistry,
University of Sydney, New South Wales 2006, Australia
Mamoru Nomura (191), Department oflndustrial Chemistry, Fukui University, Fukui, Japan
R. H. Ottewil/ (1), School of Chemistry, University of Bristol, Bristol
BS8 1TS, England
Gary W. Poehlein (357), School of Chemical Engineering, Georgia Institute
ofTechnology, Atlanta, Georgia 30332
Vivian T. Stannett (415), Department of Chemical Engineering, North
Carolina State University, Raleigh, North Carolina 27650
ix

Contributors

J. Uge/stad (51,383), Laboratory of Industrial Chemistry, The Norwegian


Institute of Technology, The University of Trondheim, N-7034
Trondheim, Norway
V. l. Ye/iseyeva (247), Institute ofPhysical Chemistry, Academy ofSciences
USSR, Moscow, Union of Soviet Socialist Republics

Contributors

J. Ugelstad (51, 383), Laboratory of Industrial Chemistry, The Norwegian


Institute of Techno1ogy, The University of Trondheim, N-7034
Trondheim, Norway
V. l. Yeliseyeva (247), Institute ofPhysical Chemistry, Academy ofSciences
USSR, Moscow, Union of Soviet Socialist Republics

1m

Preface
Emulsion polymerization has been a very successful industrial
process for four decades. In recent years it has undergone revitalization:
while some old factories are closing, new and much more sophisticated
ones using emulsion polymerization are sprouting.
Historically, in an attempt to produce synthetic rubber, the polymerizations involving the use of an aqueous emulsifier solution resulted
in a product that in physical appearance resembled nature's latex. It also
gave a greatly improved synthetic rubber over that produced previously by
other processes, e.g., the sodium-butadiene process. This industrial success
was subsequently followed by a more theoretical approach to the problem
during the 1940s that resulted in the publication of the first scientific papers
in this field. Even then, since emulsion polymerization had been established
first as an industrial process, the theories proposed by the scientists were
primarily concerned with trying to put the industrial observations into a
general and workable scientific framework. Unfortunately, the emulsion
polymerization of styrene and the copolymerization of styrene-butadiene
fit beautifully into a simple kinetic scheme proposed by Smith and Ewart,
and during the two decades that followe9 almost all research efforts in the
field attempted to make the other monomers fit into this framework.
We now know that emulsion polymerization is not just another polymer
synthesis method and that the complexity of the interactions, whether
chemical or physical, must be considered before any control is possible
over the outcome of the reaction. The creation and nucleation of particles,
for example, is not necessarily and simply explained by the presence or
or absence of micelles, but needs the understanding of interactions of all the
ingredients present. Variables such as hydrophilic and hydrophobic associations or repulsions, polarity of the monomers, chemical structure of the
surfactants, have to be taken into account.
Research in the field is flourishing all over the world, and although numerous papers have been published and collections of papers have appeared
recently, they have the disadvantage of presenting fragments of the subject
and never the total picture. This book presents a collection of chapters, each
written by scientists in their fields of expertise. These experts come from all
over the world and thus sometimes represent different viewpoints of the same
subject. It is the hope ofthe editor and the contributors that we have been successful in presentirig a total picture of the current understanding ofthe subject.
xi

'"

1
xii

Preface

1 would like to thank Silvia Dolson for providing her time and talents to
assist in the book's cover designo

1
!1
S
~
-

1
The Stability and Instabiliy
01 Polymer Latices
R. H. Ottewill

1. Introduction .
11. The Nature of Polymer Latex Particles
111.
IV.
V.
VI.
VII.
VIII.
IX.
X.
XI.
XII.
XIII.
XIV.
xv.
XVI.
XVII.
XVIII.
XIX.
XX.

l.

The Effect of Electrolytes on a Latex .


The Theory of the Stability of Lyophobic Colloids
Coagulation as a Kinetic Process .
An Alternative Approach
to the Critical
Coagulation Concentration
The Determination of ccc Values .
The Effects of lons that Interact with Water
Secondary Minimum Effects .
The Effects of Organic lons: Added Surfactants
lonic Head Group with a Charge ofthe Same Sign
as the Particle
lonic Head Group with a Charge of Opposite Sign
to the Particle
Nonionic Surfactants
Mixed Electrolyte Systems .
Heterocoagulation
.
Surface Coagulation
Peptization
The Effects of Adsorbed or Grafted Macromolecules.
Particle Stability in Emulsion Polymerization
Summary.
References

1
2
6
8
14
16
17
19
22
26
27
28
31
35
36
39
40
42
45
47
47

Introduction

Over the last two decades work on the formation and properties
of polymer latices has developed extensively, and a very substantial amount
of work has been devoted to the study of the processes of formation of
polymer particles in a latex and to the characterization of the particles once
formed (see for example, Fitch 1980). It is now generalIy recognized that in
1
EMULSION POLYMERIZATlON
Copyright @ 1982 by Academic Press. Inc.
AIl rights of reproduction in any form reserved.
ISBN 0-12.556420-1

R. H. Ottewill

the majority of latices, the particles are within the size range 1 nm to 1 pm,
which designates them as colloidal, and the name "polymer colloids" is
becoming commonly used to describe this type of system. Consequently, the
entire field of polymer latex manufacture and utilization is very dependent
on an understanding of the basic principIes of colloid science in the widest
sense whether it is to stabilize or coagulate the particles or to control the
rheological properties of latex formation.
In using the word stability in the colloidal sense we understand that the
particles in the dispersion will remain for long periods of time, often years,
dispersed as single entities in Brownian motion. The size range 1 nm to
1 pm is chosen to exclude at the lower end of the range small single ion s and
molecules and at the higher end to exclude particles that settle under the
influence of gravity and do not remain dispersed by Brownian motion.
These are definitions of convenience rather than rigor and in considering
polymer latices we shall frequently need to consider particles with a size.
greater than 1 pm; in this case the colloidal arguments can be maintained
but, in addition, the effects of gravity have to be included. When the la,tex
loses its stability because the particles aggregate under the influence of
chemical additives or mechanical action, the terms coagulation and flocculation are applied (see later). By the term rheology we describe the flow
and deformation properties of the system, Le., whether it is viscous (either
Newtonian or non-Newtonian) or elastic or whether it possesses both these
properties and is viscoelastic.

ll.

The Nature o Polymer Latex Particles

A typical polymer latex particle will be composed of a large number of


polymer chains, with the individual chains having molecular weights in the
range of about 105 to 107. According to the arrangement of the polymer
chains within the particle, the latter can be amorphous, crystalline, rubbery,
or glassy. Moreover, in many cases monomer is retained by the particle and
hence the particles can also be, in cases where the polymer is soluble in the
monomer, either extensively or minutely swollen. The physical state of the
particle can be important in close-range interactions and in drying. For
example, if the particles are soft, coalescence of the particles can occur to
give continuous film formation, whereas with hard particles their individuality is retained in the dry state.
In determining the colloidal behavior of a latex, the surface properties
playa very important role, and these are directly related to the preparative
. method employed. They frequently depend on (i) groupings arising from the
initiator used; (ii) adsorbed or grafted surfactants; and (iii) adsorbed or
grafted polymers, particularly, those soluble in the dispersion medium.

I
j
j

I
j

j!

1. The Stability and Instability of Polymer Latices

In aqueous-based emulsion polymerizations using water-soluble initiators, the surface groupings formed are frequently determined by the
nature of the initiator used (Ottewill et al., 1967; van den HuI et al., 1970;
Goodwin et al., 1973) and the following have been reported:
from hydrogen peroxide, persulfate,
bisazocyanopentanoic
acid

Weak acid

Strong acid

-0-S03

or

from persulfate

CH3 NH2
Base

Nonionic

,f'

-C-C~,
I
NH2
CH2
}
-OH

from azobisisobutyramidine

from hydrogen peroxide or hydrolysis


sulfate groups

of

In addition, latex particles with mixed anionic and cationic groups on the
surface can be prepared (Bolt, 1978).
In an ionizing medium of high relative permittivity (e.g., water) the acidic
and basic groupings exist in the ionized form, depending on their pKa and
pKb values and the pH, and consequently the surface of the particle
becomes electrically charged. In addition the adsorption of other ionic
species, such as surface-active ions, can also contribute to the surface
charge. In physical terms the water is a good solvent for the ions and poor
for the latex particle; that is, most of the polymers used for latex preparation are totally insoluble in water. A schematic illustration of this situation
is shown in Fig. 1 where the particle surface is shown to be that of a smooth
sphere with the charges evenly distributed over the spherical surface. The
condition of electroneutrality is maintained by balancing the charge on the
latex surface by the charges on small ions of opposite sign in the solution
phase (counterions). This forms the so-called electrical double layer in
which an equilibrium is set up between electrostatic forces and diffusion
forces. As a consequence of its surface charge, the lat.ex particle surface has
an electrostatic surface potential rjlswhich can be either positive or negative,
depending on the nature of the surface groupings, relative to 'earth. This
potential falls off exponentially with distance from the surface of the sphere
according to the equation
rjlr= rjls(a/r)exp[

K(r - a)]

(1)

where rjlris the surface potential at a distance (r - a) from the surface of the'
sphere where a is the radius of the sphere. K is defined by
(2)
K2 = (8ne2NAl1000ekT)I

R. H. Ottewill

- -+ - -

,,/

o
+

/
I
I

"-

+ "-

+ \

+
\
\
I
I

++

\
\

+
"-

"-

-- +

.,//

Fig. 1. SchematiciIlustration of a negatively charged spherical polymer latex particle


with an electricaldouble layer. --- representsthe rangeof influenceof electrostaticforces.

and is dependent on the ionic .strength 1 of the solution phase, e the


fundamental electronic charge, NA Avogadro's number, and ; the dielectric
constant of the medium. It is K that determines how rapidly the electrostatic
potential falls off with distance from the particle surface and consequently
the range of electrostatiC interaction forces. The dashed line in Fig. 1
indicates that the range of electrostatic forces can extend well beyond the
physical size of the particle.
When latices are prepared in nonaqueous media such as hydrocarbons
then charged-surface groups no longer pro vide a practical means of stabilizing the particles formed. Under these conditions polymer chains, soluble in
the dispersion medium, can be grafted to the core polymer particle which
remains insoluble in the dispersion medium. A typical example is the use of
poly(12-hydroxystearic acid) chains to stabilize particles of poly(methyl
methacrylate) in dodecane. This leads to what might be termed a "hairy
particle" with the noncharged "hairs" extending into the solvent medium as
shown in Fig. 2. Again the range to which the chains extend is important as
it determines the distance at which one particle of this sort can start to
interact sterically with another, giving the so-called sterically stabilized
systems.
Returning again to ionizing media, a combination of these two effects
can be employed by grafting to a polymer core particle polyelectrolyte
chains. This is illustrated schematically in Fig. 3. It pro vides a combinatorial effect of electrostatic and steric interactions. These will have
different interactive ranges as illustrated in Fig. 3 by the dotted line for the
electrostatic range and the dashed line for steric effects.

1.

The Stability and Instability of Polymer Latices

"-

\
.\
\
I
,

/
Fig. 2.

SchematiciIIustrationof a nonchargedpolymerlatex particlewith adsorbed or

grafted nonionic

polymer chains. ---

represents

the range of inftuence of steric forces.

The chains shown in Fig. 3 are those of long poly(ions) deliberately


added. However, there is sorne evidence to indicate that even in conventional ernulsion polyrnerizations the particles forrned rnay not be as
srnooth as those shown in Fig. 1, and the charged groups rnay be floating a
short distance in the rnediurn as "rnicrohairs." In practice one should not be
rnisled by the convenience of the srnooth sphere rnodel for theoretical
rnodeling of colloidal phenornena.

. . . . . .'".. .
+

'

.,

- .

. +

.+

+
+

. +

+
. . .. . . .

Fig. 3.
Schematic iIIustration of a polymer latex core with grafted polyelectrolyte
chains attached to the core surface.
represents the range of electrostatic forces, and --represents the range of steric forces.

R. H. Ottewill

From this qualitative description of latex particIes we can immediately


recognize the origins of three basic forces that ha ve to be considered in
understanding the behavior of dispersions in both aqueous and nonaqueous
media. These can be summarized as
1. Electrostatic effects: usualIy repulsive but opposite charges on particIes can lead to attraction.
2. Steric effects: arising from the geometry and conformation of adsorbed or grafted mo.1ecules.
3. Solvation effects: arising from the organization of solvent molecules
near an interface or between the chains of adsorbed macromolecules.
It will be noticed that these effects do not take into account to any great
extent the bulk phase of the particIe. However, this also has to be taken into
account and both the polarizability and the density are of importance in
determining the attractive forces between particIes. Thus, we can add to the
list

4. Attractive effects: which have their origin in molecular dispersive


interactions (oftentermed van del Waals interactions).

m.

The Effect of Electrolytes00 a Latex

In an aqueous latex that has been cIeaned to remove various materials


such as surface active agents one is dealing in many cases with a dispersion
of charged spherical particIes. FrequentIy, the distribution of particIe sizes is
very narrow and the term monodisperse is used to describe latices of this
type. In an attempt to obtain a concise overview of the properties of
aqueous latices in electrolyte solutions some of the essential features are
summarized schematicalIy in Fig. 4.
At intermediate electrolyte concentrations (-10-3 mol dm-3) and at low
volume fractions of the dispersed phase, the charged particIes occupy
random positions in the system and undergo continuous Brownian motion
with transient repulsive contacts when the particIes approach each other.
The range of the electrostatic repulsive forces is represented by the dashed
circIe in Fig. 1, which implies that when a similar circIe on another particIe
overlaps with it on a collision trajectory, a transient electrostatic repulsion
occurs and the particIes move out of range. With most latices the particIes
.have a real refractive index and their visual appearance is "milky" white.
Ir, however, either the electrolyte concentration is reduced to about

10- 5 mol dm- 3, which increases the range of the eIectrostatic repulsive
forces, or the concentration of the particIes is increased, a situation is

1. The Stability and Instability of Polymer latices


Order
Regular lattice
Strong repulsion

Electrolyte 10- M
5

Remove electrolyte
Imixed-bed ion exchang~

Dlsperslon
Electrolyte

- sol
10-

resin)

Disorder
Random arrangement
of particles
Brownian motion
Repulsive contacts

Add electrolyte

Stable
Unstable
Electrolyte 0.2 M

Schulze

Hardy
Flocculatlon
Disorder
Weak attraction

Fig. 4.
partides

Coagulation
Disorder
Strong attraction

The effectof electrolyteand partide size on the properties of polyrner latex

in an aqueous

rnediurn. (Reproduced

with perrnission

of Chem. nd. (London).

reached where the partic1es must maintain repulsive contacts over a long
period of time. Con sequen tIy, an ordered arrangement.of the partic1es is set
up so that the partic1es sit in lattice positions but remain well separated, Le.,
a "liquid crystal" arrangement is formed (see Fig. 4). When the interpartic1e
spacing is of the order of the wavelength of light, Bragg diffraction effects
become superimposed on the scattering from the partic1es and bright
iridescent colors can be seen visually. This is well demonstrated with
mono disperse latices which, in the right size and concentration range, show
brilliant colors (Hiltner and Krieger, 1969; Hachisu, 1973; Goodwin et al.,
1980).

R. H. Ottewill

In the lower part of Fig. 4 a schematic illustration is given of the


transition from a stable dispersion, where all the particles exist essentially as
single entities, to an unstable state where aggregation of the particles occur.
This change from stability to instability has received a great deal of
attention from several generations of colloid scientists. It is usually associated, when inorganic electrolytes are used, with the names of Schulze
(1882, 1893) and Hardy (1900) who investigated this phenomenon at about
the turn of the century. Today, our explanations of colloidal particle
aggregation processes are largely based on the theories of interactions
between particles that were first enunciated in quantitative form in the
1940s by Derjaguin and Landau (1941) and Verwey and Overbeek (1948),
frequently now termed the DL VO theory in honor of the four authors.

IV.

The Theory of the Stability of Lyophobic CoUoids

The quintessence of the ideas put forward by DL VO was that the


potential energy of electrostatic repulsion VRbetween the particles and the
potential energy of the van der Waals attraction VAcould be added together
to obtain the total potential energy of interaction VT. When the very
shortrange Born repulsive energy was included, a potential energy against
distance of surface separation curve of the form shown in Fig. 5 was
obtained. This type of curve exhibits a number of characteristic features. At
short distances, a deep potential energy minimum occurs, the position of
which decides the distance of cIosest approach ho and is hence termed the.
primary minimum. At intermediate distances, the electrostatic repulsion
makes the largest contribution and hence a maximum occurs in potential
energy of magnitud e Vm;this position is termed the primary maximum. At
greater distances, the exponential decay of the electrical double layer term
causes it to fall off more rapidly than the power law of the attractive term
and another min!mum occurs in the curve, of depth VSM, termed the
secondary minimum.
In more quantitative terms VRfor weak interactions can be written in the
form
VR =

3.469 X 1019 B(kT)2ay2 exp( -

Kh)/V2

(3)

where Bis the dielectric constant of the dispersion medium, k is Bolzmann's


constant, T is"the absolute temperature, v the magnitud e of the valency of
the counterion, and y = [exp(veljs/2kT)- 1J/[exp(veljsl2kT)+ 1J with e as
the fundamental electron charge.
This expression given by Reerink and Overbeek (1954) is for the electrostatic interaction between two spheres of equal radius a of constant

1.

The Stability and Instability of Polymer Latices


10

Primary
Maximum

. ...

l....

o
Secondary
Minim~m

Pri';;a~y- ,
Mnimum

h
Fig. 5.

SchernaticiIIustration of a potential energy against distance of surface sepa-

ration curve to illustrate the rnain features used in discussing colloid stability. ~ Vr= energy
barrier to coagulation;

~Vb

energy barrier to peptization;

i'd = dispersion

free energy of the

polyrner-water interface; Vm= height of the prirnary rnaxirnurn; VSM= depth of the secondary

rninirnurn.

surface potential and material 1 separated by a distance h in a medium 2


(Fig. 6). It is valid for the Ka range 3 to 10. For Ka < 3 and Ka > 10 useful
expressions for small potentials are, respectively
VR = [ea21jJ;/(h
VR = tealjJ;

+ 2a)] exp( -'Kh)

ln[l + exp(- Kh)]

(4)
(5)

The question of interaction at constant surface charge has also been


discussed by several authors (Frens, 1968; Wiese and Healy, 1970; Gregory,
1975).

R. H. Ottewill

10

Fig. 6. Interaction betweentwo sphericalparticlesof material 1 in a liquid mediumof


material2. a = radius of sphericalparticle;h = distanceof surfaceseparation.

For the attractive part of the interaction it was shown by Hamaker


(1937) that VAcan be given for two spheres of the same radius and the same
material by the expression
V A-

12{ X2 + 2x+ X2+ 2x+ 1+

2l

X2

+ 2x

n(X2+ 2x+ 1)}

6)

where x = hl2a and A = the composite Hamaker constant for the particles
in the mediumas givenby
(7)
TABLEI
Values of Reported Hamaker Constants for Various Polymers

Material
Poly(vinyl acetate)
Poly(vinyl chloride)
Poly(methyl methacrylate)

Styrene-butadiene
Polytetrafluoroethylene

Polystyrene

AII/J
(x 10-2)
8.84
12.4
10.0
6.3
5.6
3.7
7.6

a Lifschitz, small separation distances.


b Lifschitz,
long separation distances.

AjJ
(xlO-21)
5.4
5.5
7.2
5.5
3.0-3.8
4.0
2.9
3.6
3.5
9.0
3.2
6.5
7.0

Reference
Dunn and Chong (1970)
Evans and Napper (1973)
Visser (1972)
Visser (1972)
Friends and Hunter (1971)
Evans and Napper (1973)
Force and Matijevi (1968a)
Fowkes (1967)
Rance (1976)
Gingell and Parsegiana (1973)
Visser (1972)
Krupp et al. (1972)
Gingell and Parsegiana (1973)
Gingell and Parsegianb (1973)
Evans and Napper (1973)
Lichtenbelt et al. (1974)

11

1. The Stability and Instability of Polymer Latices

= the Hamaker constant of the particIes and A22 that of the


medium. The Hamaker constant is directly related to the nature of the
material by the expression

with All

(8)
where Vj is the dispersion frequency of the material (c/A.o,with A.o= the
dispersion wavelength), Ctjis the static polarizability and qj is the number of
atoms/molecules per unit volume of the material. Some typical values of

A 11 for polymericmaterials are listed in Table I.


For conditions such that x ~ 1 a useful approximation is obtained from
Eq. (6),in the form
(9)

VA = - Aa/12h

As the distance of separation between the particIe surfaces increases,


particularly beyond 100 nm, a weakening of the attraction starts to occur,

150
11

.,
100 t+--

\
\

\
\

\.,

:i.

11.

kT

50':'

'\
"-

\
\

...

.\

"-

"". -- .;-..,

.. .........

"........

,.,

- - =.= .'::..~,

10

...-.--.-20

h (nm)
Fig. 7. Potential energy diagrams for the interaction between two spherical latex
particles of radius 0.1 Jlm at a constant potential of 1/1,= 50 mV, with A = 7.0 X 10-21 J and
T=298.2K
in a 1:1 electrolyte. -,
1O-3moldm-3;
---,
1O-2moldm-3;
-'-',
0.075 mol dm-3;
0.5 mol dm-3.

R. H. Ottewill

12

100
VT
kT

,......,

I!

r:

.,

"

'.

.'.

e.-- ..

'.

......-.

- .-.

k~..".-

O . ..-"

-8-

::
V..a.~.- . _8-

--

..- -
'8.~ I...:e-=."\:

f
h (nm)

Fig. 8.

Potential energy diagrams for the interaction between two spberical latex

partic1es of radius 0.1 Jlm at a constant

1:1 eIectrolyte

concentration

of 10-3 mol dm-3

A = 7.0 X 10-21 J. T = 298.2K. and various values of "', as follows:

37.5mV;-u-o 25mV;

12.5mV;

5 mV.

-.

witb

50 mV'; -'-,

the so-called retardation effect, and this has to be allowed for in the
calculation of VT. Equation (9) can then be rewritten in order to allow for
retardation as (Kitchener and Schenkel, 1960)
VA = -(Aa/h)(2.45/p

- 2.17/180p2 + 0.59/420p3)

(10)

with p equal to 21th/A.o.Usually, A.ocan be taken as -100 nm and Eq. (10) is


a reasonable approximation for a ;p h and 0.5 < p < oo.
An alternative treatment of attractive interactions between particles is
that given by Lifshitz and co-workers (1961); it is beyond the scope of this
article but is discussed by Ninham and Parsegian (1970, 1971) and
Richmond (1975).
From the various equations given, under the appropriate conditions of
Ka, h. etc. the total potential energy of interaction between the particles can
be calculated from the basic DL VO assumption that for lyophobic
dispersions,
(11)

"=""

13

1. The Stability and Instability of Polymer latices

"
\

200H-'I
I
I

I
I

\
\
\
\
\
\

100

O""'O\

'o

r!
,1

~,f
o

!II-:! 8_'-..1 -=

~ -= .":--'::.
20

-- ~ _8-:--=30

(n m)

Fig. 9. Potential energydiagramsfor the interactionbetweensphericallatexparticlesin


a 1:1 electrolyte at a concentration of 1O-2moldm-3 with A=7.0xl0-2IJ
and
T = 298.2Kover a range of particle radii: -,
a = 2 Jlm;---, a = 1Jlm;-'-, a = 0.5Jlm;
, a = 0.1Jlm,t/J, = 25 mV.
.

-.
-

A number" of curves of VT against h are given in Figs. 7 to 9 to illustrate


various points. For example, Fig. 7 shows the effect of electrolyte for
spherical particles of radius 0.1 J1.mat a constant surface potential t/ls.For
the same particle radius Fig. 8 shows the effect of changing t/ls at a constant
electrolyte concentration of 10- 2 mol dm - 3. The effect of particle size for
particles of the same t/ls at the same electrolyte concentration is demonstrated in Fig. 9.
It is apparent from these curves that when the magnitude of Vmis greater
than about 10kT conditions are favorable for the formation of a stable
dispersion. Moreover, the form of the potential energy curve obtained by
this approach indicates that the dispersion exists in a higher free energy
state than that of the solid polymer with the depth of the primary rpinimum
equal to twice the dispersion free energy of the polymer (Fowkes, 1964). The
stability of the dispersion is kinetic in origin as a consequence of the large
activation energy ~Jrwhich renders the forward transition into the primary
minimum improbable. It also becomes apparent that because the activation

R. H. Ottewill

14

energy for the backward reaction /!:;.


Vbis much larger than /!:;.
Vr(see Fig. 5), and
a decrease of free energy in the system has occurred once the primary
minimum has been accessed, then spontaneous redispersion is unlikely to
occur and mechanical work is needed to redisperse the aggregated particles.
Mreover, time has to be considered in this context since prolonged contact
of particles in the primary minimum can lead to welding of the particles, as
interparticle diffusion of polymer chains occurs.

V. Coagulation as a Kinetic Process


In a fundamental paper von Smoluchowski (1917) presented a theoretical
model for the kinetics of the coagulation process. He showed that in the
initial stages of coagulation the rate of disappearance of the primary
particles, i.e., those present as single particles in the original dispersion,
could be written as
-dN/dt

= kN~

(12)

where No is number of primary particles per unit volume present initially


and k is arate constant. For rapid coagulation, i.e., coagulation in the
absence of an energy barrier, the process is diffusion controlled and
k

= ko = 8nDR

where D is the diffusion coefficient of a single particle and R

the collision radius of the particle.


In subsequent analyses it was shown that if diffusion in the presence of
an energy barrier is considered (Fuchs, 1934; Overbeek, 1952) then the
initial rate of disappearance of the primary particles could be written as
-dN/dt

= koNUW

(13)

where W, termed the stability ratio, is related to VTby


W

= 28 r'1) exp(VT/kT)
Jo

dh

(14)

(h + 2a)2

In the absence of an energy barrier, i.e., setting VT = O,we find W = 1 and


the equation for rapid coagulation is obtained. In the presence of an energy
barrier VTbecomes positive and W becomes greater than unity and, clearly
for these conditions, the rate of coagulation is slowed down; hence slow
coagulation occurs. In practice at intermediate electrolyte concentrations

and medium potentials (ljIs'" 50 mV) W can attain values of the order of
107 so that coagulation is imperceptible on a reasonable time scale. The
approach emphasizes the kinetic nature of the stability of lyophobic
colloids.

15

1. The Stability and Instability of Polymer Latices

Measurements of the rate of coagulation of latex particles can be carried


out by a number of techniques, two of those most commonly used being
particle counting and light scattering (Ottewill and Shaw, 1966). These give
values for the rate constant k and since a well-marked transition usually
occurs between the slow and rapid coagulation regions the assumption is
usually made that for rapid coagulation W = 1, whence for this region
k = ko, and thus in the slow coagulation regon, W can be obtained from
the ratio ko/k. A typical example of the type of experimental data obtained
is shown in Fig. 10 plotted in the form log Wagainst log C., where C. is the
concentration of electrolyte in the dispersion after mixing has occurred at
zero time. As can be seen, the transition between slow and rapid coagulation is clearly marked and a well-defined electrolyte concentration can be
obtained from the graph at this point. This electrolyte concentration is
termed the critcal coagulaton concentraton and will be abbreviated as cee.
As a further development, it was shown by Reerink and Overbeek (1954)

::
C
o
...J

0.5

-2.5

-2.0
Log

--~

-1.5

ce

Fig. 10.
Log W against log C. eurve for a polystyrene latex (a = 0.21 Jlm) in barium
nitrate solutions (Ottewill and Shaw, 1966).
indieates results obtained from light-seattering

measurements; O indieates resuIts obtained using a partic1e eounting teehnique;


cee value.

indieates

R. H. Ottewill

16

that the gradient of the curve just before the cee was given as a first
approximation, based on Eqs. (3) and (9), by
(15)
From this equation it is clear that the gradient should be directly proportional to the particle radius a and inversely proportional to the square
of the valence of the counterion used. A number of experimental studies
have been made tha~ do not seem to confirm these predictions (Ottewill and
Shaw, 1966) and even more refined treatments of the kinetic process have
not removed the discrepancy (Derjaguin and Muller, 1967; Honig et al.,
1971). There is little doubt that as a kinetic process coagulation is rather
complicated; further discussion on this point will be given later.

VI.

An Alternative Approach to the Critical


Coagulation Concentration

From the previous discussion.it is apparent that provided the magnitude


of Vmis substantial the probability of transition into the primary minimum
is small. However, as shown above when VT = Vm = Othe transition is facile
(Fig. 11) and the particles coagulate. Therefore, we can define conditions for
the onset of instability as (Verwey et al., 1948)
and

(16)

whence using Eqs. (3) and (9) we find


Kcrit= 2.04 X 1O-Sy2/Av2

(17)

Distance
>~

IJ)
c:
W

...

...

...

...

Fig. 11.
Sehematie potential energy eurve to iIIustrate eonditions leading to a theoretical definition ofthe cee. -,
VT;---, VAonly.

17

1. The Stability and Instability of Polymer Latices

and since for a symmetrical electrolyte 1( can be related directly to the


concentration of the electrolyte C expressed in mol dm - 3 by
1(2

= 81tv2e2NAC/ekTx103

(18)

with NA = Avogadro'snumber, wethen find using Eqs. (17)and (18)that,


CCril=

3.86 x 1O-25y4
-3
mol dm
A 2v
~

(19)

where Ccritis the ccc and we note the inverse dependence on the sixth power
of the valence of the counterion.
For low surface po tentials, t/ls< 25 mV, a further simplification can be
made to give
mol dm-3

(20)

which gives an inverse dependence on the square of the valence, so that for
univalent, divalent, and trivalent counterions, we obtain

The DL VO theory thus enabled theoretical significance to be given to the


valence sequence in coagulation experiments that had been observed many
years earlier by Schulze (1882) and Hardy (1900). Although expressions of
this type are useful in a qualitative predictive sense, the implication that
there is a simple rule applicable to alf systems must be treated with
considerable caution. It must be borne in mind that coagulation is a
complicated phenomenon involving quite a range of kinetic and specific ion
effects.

VD.

Tbe Determination of eee Values

From an experimental viewpoint, however, there is no doubt that the ccc


is a very important quantity to know for a polymer latex, since it represents
the electrolyte concentration at which complete loss of stability occurs.
Experimentally, the ccc can be obtained by a variety of methods and the use
of light scattering and partic1e counting ha ve already been mentioned.
Possibly the simplest method of all is visual observation in test tubes
containing the same concentration of latex and different concentrations of
electrolyte. A slightly more elaborate version of this method is to use a
simple spectrophotometer to measure the optical density of the dispersion
at specific time intervals after addition of electrolyte; the most convenient

R. H. Ottewill

18

time period having been determined from some preliminary experiments


(OttewilI and Rance, 1977). An example is given in Fig. 12 of a plot of
optical density against pH which shows the boundary between slow and
rapid coagulation regions. Extrapolation of the rapidly descending portion
of the curve to the abscissa yields a value for the cec. As can be seen the
value obtained wilI be time dependent so that for comparative purposes,
e.g., to evaluate the effectiveness of different electrolytes, standardization of
the time period of the experiment is essential.
The time dependence is not unexpected since each time corresponds to a
different stage in the overall kinetic process. There is no doubt that the most
precisely defined value is that obtained from the log Wagainst log Ce curve
using initial rates, since this is always defined as t -+ O. We can anticipate
that there will be some variation in the actual values of the ccc obtained by
different methods and also between different workers. The general pattern
of the results observed, however, should not change with either of the latter
factors.

0.2

>-

VI
c:
CD

c.
O
0.1

2.4

2.0

1.6
pH

Fig. 12.
Optieal density against pH al various times after addition of hydroehlorie aeid.
-0-,1
hr; -8-,
2 hr; -/;:,.-,26 hr; ---, cee 1 hr; -'-', cee 2 hr;
, cee 26 hr.

1.

19

The Stability and Instability ot Polymer latiees


TABlE

11

cee Values tor Various Polymer latiees


Latex
Polystyrene
(Carboxylsurface)
Polystyrene
(Amidinesurface)
Divinylstyrene
Styrene-butadiene

Poly(vinyl chloride)

Counterion

cccjmmol dm - 3

H+
Na+
BaH
La3+ (pH 4.6)
CIBr1Na+
Na+
K+
MgH
BaH
La3+ (pH 3)
Na+
MgH

1.3
160
14.3
0.3
150
90
43
160-560
200
320
6
6
0.5
50-200
2-10

Reference
Ottewill and Walker (1968)
Storer (1968)
Ottewill and Shaw (1966)
Ottewill and Walker (1968)
Pelton (1976)

Neimann and Lyashenko (1962)

Force and Matijevi (1968a)

Bibeau and Matijevi (1973)

The cee values for a number of polymer latiees have been determined
and some typical values are reported in Table II. The trends observed are
qualitatively in agreement with those expected fram the theoretieal approach for particles with smooth surfaces, with 1/1.everywhere the same,
using simple eleetrolytes, i.e., those which do not interact chemically with
water to form new ionie species. These yalues should only be used for
qualitative guidance since, in addition to the factors already mentioned,
there can be variations of the cec with particle size, type and density of
surface groupings, and the presence or absence of stabilizing materials such
as surfactants. In practice it is advisable to determine the actual value for a
particular latex system.

VIII.

Tbe Effects of IODSTbat Interact with Water

So far, the assumption has been made that the ions used in the
coagulation experiments do not interact with water. In a number of cases,
however, the ions do react with water under certain pH conditions to form
hydrolyzed species. For example, in the case of aluminum, the AI3+ ion
exists at pH values below about 3.3 as the hexaaquo ion, with six water
molecules in the octahedral coordinate positions. As the pH is slowly
increased, reaction occurs with water to form a sequence of species. The
chemistry involved in these reactions is somewhat complex and has not

1-

l
20

R. H. Ottewill

been fulIy resolved but a plausible reaction scheme can be proposed, for the
present purpose, as
~20

20
H20

",

--.

'~'~I~+
""

H;O

H20:
.'.

Ji20

:
'. ..

H20:
".
--.

''':~I~+
'..
'......

..........

lI20

~20
OH

'. H20

H;O

OH

/i.....
HO

'H20

:
:
"AI+

H20

"H20

H20

/
1+

H'~ r~~",r~H
"~I

/'~H/'"
[ HO

..

~!

((

lI20

..

"'}{20

1120

1
Alx[OHy]"+

where Alx[OHy]n+ represents an inorganic polymer soluble in water.


Polymeric species of this type can adsorb strongly onto negatively charged
particles and reduce the effective surface potential on the particle to zero. As
anticipated from Eq. (3) this situation leads immediately to coagulation. At
higher concentrations of the aluminum species, superequivalent adsorption
can take place, thus conferring a positive charge on the particle and leading
to restabilization of the dispersion as one containing positively charged
rather than the original negatively charged particles. In addition, it is also
possible for the positive polymeric species to "bridge" two negatively
charged particles.
The exact nature of the polymeric species in solutions of aluminium salts
at the pH conditions for charge reversal is not known with certainty. It is
possible that several species coexist, depending on their stability constants,
and that these also change with time with the ultimate product of hy-'
drolysis being aluminum hydroxide particles. A number of species ha ve
been pro posed in the literature and Matijevi et al. (1964) ha ve provided
evidence for the existence of Als(OH)1t from coagulation studies. The
higher valence of this type of species again reduces the concentration of ions

1.

21

The Stability and Instability of Polymer Latices

required to produce coagulation. The combined effects of high valence and


reduction of the surface charge to zero makes aluminum salts very effective
coagulants in the pH range of about 4 to 5.5. Coagulation can be achieved
at very low salt concentrations and since most of the aluminum is adsorbed
by the particles there is little salt left in the filtrate after remo val of the
coagula. This factor is exploited in the use of aluminum salts for the
treatment of potable water.
The basic pattern of the coagulation of polymer latices with aluminum
salts has been clearly demonstrated by the work of Matijevi and his
collaborators (1968) using styrene-butadiene, poly(vinyl chloride) (1977),
and PTFE (1976) latices. The results obtained by Matijevi and Force
(1968b) for the coagulation of styrene-butadiene latices using alumirium nitrate are shown in Fig. 13. From these it can be seen that up to a
pH of ",3.4 the ccc remains constant at 5 x 10-4 mol dm-3 and then decreases between pH 3.4 and 4.8 to reach a constant value of

'"

'";;
O
z

-2

c:

.2 -3
-;;
...
:

Q)
()
c:
o

RESTABILlZATlON

... -4
al
o
~
ti'
o
...J

STABlE

-5

REGION

..

pH
Fig.13.
Log[(AIN03h/mol dm-3J against pH showing the positions of the coagulation domains for a styrene-butadiene latex. Curves constructed from the data of Matijevi
and Force (1968); reproduced with permission of Ko/loid Z.u.Z.fur Polymere.

.. ..

R. H. Ottewill

22
3.0

- 3.4

"O

<5
E
"-

,..,
u
u
u

':::! 3.8

'"

4.2

pH

Fig. 14.

Log[ccc] against pH for a PTFE latex using aluminium

nitrate

as the

coagulating electrolyte.

'" 2.5 x 10- 6 mol dm - 3

between pH 4.8 and 6.0. The region of restabili-

zation as positively charged particles can also be seen on this "domain"


plot.
In the case of PTFE latices the behavior seems to be strongly dependent
on the amount of stabilizer present in the system. Kratohvil and Matijevi
(1976) found evidence for both charge reversal and restabilization. On the
other hand, Ottewill and Rance (1979) using well-dialyzed PTFE latices
obtained the results shown in Fig. 14. The change in cee with pH is clearly
defined but the cee above pH 5 is at a fairly high concentration of
electrolyte and insufficient positive charge is built up to stabilize the
dispersion. It is also possible with this system that the very hydrophobic
polytetrafluoroethylene parts of the surface do not adsorb the hydrolyzed
species, and in the well-dialyzed system adsorption can only occur on the
sparsely charged sites.

IX.

Secondary MinimumEtIects

One of the pronounced features of the curve of VT against h shown in


Fig. 5 is the secondary minimum, and as can be seen from Fig. 9, although
this is a feature that is not very pronounced for small-diameter particles, it
becomes more distinctive as the particle size increases. With an increase in

1.

23

The Stability and Instability of Polymer Latices

salt concentration, for a constant particle radius, the depth of the secondary
minimum increases; and although the magnitud e of Vrnis reduced, it
nevertheless remains distinct and positive. These trends can be clearly seen
in Fig. 15. The form of the potential energy curves indicates the possibility
that over this range of electrolyte concentration, once a particle enters a
secondary minimum, it will have a long residence time there and remain
separated from the second spherical particle by distances of the order of 6
to 10 nm. However, there remains a substantial primary maximum in
po tential energy to be overcome before the particles can come into contact
or enter the primary minimum. Association in the latter state clearly
corresponds to a condition where the particles come into close contact,
providing the possibility with subsequent thermal diffusion that they will
20 o

,"'
, '.

100

l'

i/\

... .' \
\ \

:!:L
kT

'

-',-'.- \
\

'.

".
.

:,

-100

:.,I

'.

------

;',.y

-----

--_o
.....;;-;:.-:~._- - - - -

00'

,,"

:,

h (nm)
Fig.

15.

VT/kT against

h for spherical

partic1es

of radius

= 1.62

J.lm at various

con-

centrations of al: 1 eIectrolyte. -,


0.05 mol dm-3; -'-, 0.1 mol dm-3; -"-, 0.15 mol dm-3;
,0.3 mol dm-3; ---,0.4 mol dm-3; 1jI,= 25 mV; and A = 7 X \0-21 J at a temperature of
298.2K.

i;

R. H. Ottewill

24

fuse together. Under these conditions, therefore, one would expect the units
formed to be hard, compact, and essentially nonreversible, and there is
compelling logic to term this state coagulation. On the other hand, when
association occurs in a secondary minimum, the particles remain separated
by a liquid film, which renders thermal diffusion of polymer chains between
the solid particles unlikely, and leaves the possibility that by decreasing the
salt concentration the particles can redisperse. The logical term for this
state is flocculation and there are strong reasons for distinguishing between
the coagulated and flocculated states when simple electrolytes are used to
produce them. It is clear from Fig. 15 that with continued addition of
electrolyte, a transition from the flocculated to the coagulated state should
also occur.
The spherical nature of the particles and the high degree of monodispersity in polystyrene latices makes them ideal systems for testing such a
hypothesis and exploring its practical implications. In some preliminary
experiments by Mardle (1980) the ccc values were determined for a series of
latices of different particle sizes; and then over a range of salt concentrations at and above the ccc he examined the effect on the associated
state of using dialysis to remove the salto The results obtained are summarized in Table III.
TABLE

111

Reversibility of Aggregates Formed on Addition of Electrolyte


Particle
diameter
(Jlm)

eee/mol dm - 3
value for NaCI

0.21

0.208

0.208

Large aggregates
Nonreversible

0.40

0.155

Some redispersion
Some redispersion
Nonreversible

0.58

0.150

1.34

0.108

3.24

0.158

0.155
0.25
0.40
0.150
0.60
0.70
1.00
0.108
0.400
1.500
0.158
0.400
1.0

Eleetrolyte
NaCI eone
(mol

dm - 3)

Behavior
on dialysis

Some redispersion
Some redispersion
Some redispersion
Nonreversible
Redispersion
Redispersion
Redispersion
Redispersion
Redispersion
Redispersion

1.

25

The Stability and Instability of Polymer Latices

On a kinetic basis the presence of a pronounced secondary minimum


should lead to a steady-state condition in which the rate of particles
entering the secondary minimum to form associated units should be
balanced by their rate of return to single particles. Direct evidence for this
situation has been obtained recently by Cornell et al. (1979), using optical
microscope observations on particles of 2-Jlm diameter. The results are
given in Fig. 16 as plots of the percentage of the total number of particles
remaining as single particles as a function of time at several difIerent
electrolyte concentrations. A marked dependence of the rate of disappearance of single particles on electrolyte concentration is apparent. In the
most dilute electrolyte, 10- 5 mol dm - 3 sodium chloride, no change was
observed in the number of single particles present over a period of 6.3 hr.
At salt concentrations of 3 x 10-3 and 10-2 mol dm-3 the percentage of
single particles initially decreased and then subsequently became constant
over an extended period of time, indicating a steady-state condition in the

C/')
:: 60
UI
41

:u 40
Q.
~

20

200

100
Time

300

(min)

Fig. 16.
Experimental determination of percentage of particles remaining at various
times after addition of electrolyte using a polystyrene latex with a particle radius of 1 Jlm.
Sodium chloride concentrations. -0-,
10-5 mol dm-3; -e-,
3 x 10-3 mol dm-3;
-f:::,-, 10-2 mol dm-3; --,
6 x 10-2 mol dm-3.

R. H. Ottewill

26

aggregation process. Assuming that at this stage doublets form the predominant associated units the rate of disappearance of single particIes
becomes
(21)
where k1 is the rate constant for entry into the secondary minimum, k2 the
rate constant for exit from the secondary minimum, N1 the number of single
particIes, and N2 the number of doublets. For a steady-state condition we
have
.
-dN/dt

=O

(22)

and consequentiy the number of doublets is given by


N2

= k1NI/k2

(23)

From these arguments it would be anticipated that once a steady-state


condition was achieved the percentage of single particIes would become
constant. With a further increase in electrolyte concentration, however, a
deepening of the secondary minimum would occur and therefore more
particIes would reside in secondary minima. The experimental observations
shown in Fig. 16 appear to be in accord with this.
With this size of latex particIe it becomes possible to make direct
observations on particIes over a period of time and record them with a
high-speed camera. Using this technique Comell et al. (1979) discovered
that particIes in an associated unit could be quite mobile. It was observed
that as well as so~e particles leaving the aggregated unit as single particIes
and retuming to the disperse phase there was a continued rearrangement of
the particles. This was also observed with floccules at salt concentrations
well above the cee. These observations clearly support the contention that
association can occur in a secondary minimum and that in this situation a
liquid film is maintained between the particles.

x.

The Effects of Organic Ions: Added Surfactants

In general, the types of surfactant added to a latex will be either anionic,


cationic, or nonionic as classified by the nature of the head group. If it is
assumed that the latex particle has a surface free of adsorbed materials and
is negatively charged, then the various possibilities of adsorption of the
surfactant can be envisaged by the schematic diagram given in Fig. 17. The
discussion of the various phenomena observed can then be based on these
models.

27

1. The Stability and Instability of Polymer Latices

ANIONIC
ENHANCEO

/:~

STAS ItlTY

COAGU LA TI ON

CATIONIC

RESTAS ItI ZATION

STASltlTY

--

..

Fig. 17.
Schematic iIlustration of the adsorption of anionic, cationic, and nonionic
surfactants on anionic polystyrene latex particles.

XI.

Ionie Head Group with a Charge of the Same Sigo as the Particle

In this situation the hydrophobic chains of the surfactant adsorb onto


the hydrophobic areas of the latex particle and hence leave the anionic
group exposed to the solution phase, thus increasing the overall surface
charge on the particle.
The phenomena can be illustrated by data obtained by Ottewill and
Rance (1977) using polytetrafluoroethylene (PTFE) latices. With welldialyzed latices the cee values were those recorded in the first column of
Table IV. After it had been established that the perfluorooctanoate ion can
adsorb on PTFE particles to give a monolayer, with an area per adsorbed
perfluorooctanoate ion of 54 A2, at ~20% of the critical micelle con- .
centration (2.4 x 10- 2 mol dm - 3) the cee values were redetermined in
2.5 x 10-2 mol dm-3 ammonium perfluorooctanoate. The results are given
in Table IV.
In the presence of ammonium perfluorooctanoate, it can be seen that
with sodium chloride and aluminum nitrate the stability of the PTFE latex

R. H. Ottewill

28

TABLEIV
cee Values for PTFE latiees

Eleetrolyte
NaCI
Ba(N03h
AI(N03h (pH 3.0)

eee/mol dm - 3
dialyzed latex

eee/mol dm 3
in the presenee of
2.5 x 1O-2moldm-3
ammonium perfluoroaetanoate

4.7 X 10-2
7.4 X 10-3
1.6 X 10-4

2.2 X 10-1
7.4 X 10-3
5.4 X 10-3

increased and that nearly an order of magnitude increase in salt concentration was needed to produce coagulation. On the other hand with
barium nitrate the ccc value did not change. The explanation was that
barium perfluorooctanoate is essentially insoluble in water and the addition
of Ba2+ ions to the solution phase stripped the adsorbed perfluorooctanoate ions from the surface.
A further important point arises in the context of adsorption. It was
found that despite the fact that both fluorocarbons and hydrocarbons are
hydrophobic dodecanoate ions are only very weakly adsorbed on a PTFE
latex surface and do not form a monolayer, whereas perfluorooctanoate
forms a monolayer at relatively low concentrations (Rance, 1976).
Compatibility of the polymer surface and the surfactant is therefore an
important factor to consider if enhanced stability is required from addition
of the surfactant.
On basically hydrocarbon-hydrophobic substrates such as polystyrene,
it is well established that even on the negatively charged partic1es there is
adsorption of surfactant anions via the hydrocarbon chains. This is demonstrated in the work of Kayes (1976), who found a substantial increase in
the electrophoretic mobility of polystyrene latices with increase in the
concentration of dodecyl sulfate in the system, and in the work of Cebula et
al. (1978) on the adsorption of dodecanoate ions on polystyrene latex
partic1es.

XII.

Ionic Head Group with a Charge o Opposite Sigo


to the Particle

As illustrated in a very simple fashion in Fig. 17 in the case of a


negatively charged partic1e, the first stage of adsorption of a cationic
surfactant is via the positive head group to neutralize the charge on the
r

:
L

IJ!!II

1.

29

The Stability and Instability of Polymer Latices

particle, whence IjIs = O and consequently VR= O also. In this region on a


hydrocarbon surface it is probable that the hydrocarbon tails lie flat on the
surface. Some data obtained for the coagulation of polystyrene latices by a
series of alkyl trimethylammonium halides are shown in Fig. 18 in the form
of curves of log Wagainst log Ce; for comparison results are also included
for a simple 1: 1 electrolyte, potassium bromide (Connor, 1968; Ottewill,
1980). It can be seen from these data that the range of concentrations over
which coagulation occurs is very narrow and that the cee is strongly
dependent on the chain length of the hydrocarbon tail of the surfactant
molecule. Studies of the electrophoretic mobility of the particles confirm
that at the cee the particle mobility becomes zero (Connor and Ottewill,
1971).
Once the negative charges on the particle surface have been neutralized
further adsorption of the surfactant occurs via the tail onto the hydrophobic patches of the surface and also by association of the hydrocarbon chains. Detailed adsorption studies on polystyrene latices have
been reported (Connor and Ottewill, 1971). The additional adsorption
pro vides a positive charge to the particles and restabilization occurs. The
sharpness of this phenomenon is clearly illustrated by the experimental data
2

;: 1
c>
o
..J

-1

Log [Cone, S. A.A./ moI dni3]


Fig. 18. Log W against the log of the concentration of alkyl trimethyl ammonium
bromidesof variouschain lengths: ., C16; ., C12;., CIO;O, Cs; O, C4. For comparison,
data for potassium bromide is included. f:::,..Radius of polystyrene latex particles
Reproduced

with permission

of American

Chemical

Society.

= 48 nm.

R. H. Ottewill

30

given in Fig. 18, and it has also been examined theoretically (Ottewill et al.,
1960). At much higher additions of surfactant the electrolyte concentration
is appreciably increased and compression of the electrical double layer
occu.rs, leading to a second coagulation region. Under these conditions, i.e.,
with the ionic groups of the surfactant exposed to the solution phase, the
particles are well-wetted and the coagula usually sink if the density of the
particles is greater than that of the media. The conditions of the first
coagulat"ion region, however (i.e., zero surface potentiaI with hydrocarbon
chains orientated' toward the solution phase) leads to particles that are
easily dewetted and flotation is frequently observed (Connor, 1968).
In Fig. 19 values of the logarithm of the ccc obtained with a series of
alkyl trimethylammonium bromides are plotted against the chain length of
the surfactant for the coagulation of polystyrene latices. The curve is
essentially linear for chain lengths between C4 and Cl2 but deviates from
linearity for the short- and long-chain materials. A similar trend was
reported by Tamaki (1960) for the coagulation of silver iodide soIs by
alkylamine hydrochlorides. It appears that for chain lengths shorter than
C4 the ions behave in a manner similar to hydrated inorganic ions and that
the short hydiocarbon chain is'hydrated by the solvation sheath of the head

-1

-2
r--r

'"
le -3
"C
'O

-4
L

'" -5
o
...J

-6

Carbon

Atoms

10
in

12

14

Chain

Fig. 19.
Log cee against ehain length for the eoagulation of a polystyrene latex
(radius = 48 nm) by alkyl trimethyl ammonium bromides.

31

1. The Stability and Instability of Polymer Latices


TABLE V
Coagulation of Polystyrene and PTFE Latices by Cationic Surfactants
eee/mol

Surfaetant
CaH17 NMe; BrC1oH21 NMe; BrC12H2SNMe; Br-

polystyrene

dm - 3

eee/mol dm -

latex

PTFE latex

1.59 X 10-4
1.99 X 10-5
2.93 X 10-6

4.5 X 10-5
3.1 X 10-5
2.9 X 10-5

i
g

I
i
I

group. In the case of the hexadecyl trimethylammonium bromide there .is


some evidence for dimerization in solution, which would help to screen the
hydrophobic nature of the hydrocarbon chain.
In the coagulation of PTFE latices by cationic hydrocarbon surfactants,
however, a different behavior is observed in that only very small differences
are observed in the magnitude of the ccc with variation of chain length
(Richardson, 1979). This effect is illustrated by the data in Table V. Again, it
appears to demonstrate the lack of affinity of hydrocarbon chains for
fluorocarbon surfaces.

XIII.

Nonionic Surfactants

The most extensively studied type of nonionic surfactant is that with a


head group of ethylene oxide units. It was shown by Ottewill and Walker
(1968) that these materials can adsorb on polystyrene latex particles, below
the cloud paint of the surfactant, to give a monolayer on the surface
with the alkyl chains adsorbed on the particle surface and the ethylene
oxide groups extending into the solution phase, as illustrated in Fig. 17. By
means of ultracentrifugation studies they measured the thickness of the
adsorbed layer O and for dodecylhexaoxyethylene glycol monoether
[C12H2S(CH2CH20)60H(C12E6)] found that the adsorbed layer contained
as much as 70% water. Such a homogeneous hydrated layer pro vides a
steric barrier to the approach of the particles and ts effect needs to be
considered in some detail.
In order to account for the steric effects, an additional potential energy
term V., which pro vides a measure of the steric interaction, needs to be
introduced. On the assumption that the adsorbed layer is homogeneous
and using Flory-Krigbaum (1950) statistics to describe the mixing of the
two layers, the expression found by Ottewill and Walker (1968) for V. was
V.

= (41tC;kT/3V1P~)(t/Jl -

Xl)(O - h/2)2(3a + 20 + h/2)

(24)

R. H. Ottewill

32

where Ca the concentration of the surfactant in the absorbed layer, VI the


molecular volume of the solvent molecules, P2 the density of the adsorbed
material, rfl an entropy parameter which for ideal mixing can be taken as
0.5, and Xl a parameter characterizing the interaction of the surfactant with

the solvent. It is immediately clear that if rfl is taken as 0.5, then for
Xl < 0.5, V. is negative and the term becomes attractive, whereas for
Xl < 0.5, V. becomes positive and the term becomes repulsive. Moreover,
for this type of interaction no effect would be anticipated until h = 2<5,when
the adsorbed layers touch. An extension of this model which allows for a
redistribution of the adsorbed material during interaction has been proposed by Dorozklowski and Lambourne (1971), and application of similar
models to adsorbed polymer layers has been extensively examined by
Napper (1977) and his collaborators. The various models may be too simple.
but it does appear that for well-solvated materials XI is less than 0.5 and in
general, the interaction energy rises fairly steeply once the adsorbed layers
touch. The net effect of v., when it can be used in these circumstances, is to
impart considerable stability to the dispersion since it effectively presents
the particles from entering the primary minimum.
The form of V. against h is 'shown schematically in Fig. 20. It will,
however, seldom if ever act on its own, and usually the van der Waals

'. \ 1

..

'.\ 1
.\

'

.\
>-

O)

..
Q)

c:
W

.~

-Co
Q)

o
Q.

'.\
..
'.\
'. \
..

..

..

.. ........

h
Fig. 20.

Schematic potential energy curves. -.-',

VT= Vs + VR+ VA,

Vs only;

, vT = Vs+ vA; -,

33

1. The Stability and Instability of Polymer Latices


attraction will act in combination with

v.,so that

VT = Vs+

(25)

VA

a situation that can give rise to a shallow energy minimum, as shown in Fig.
20, and to a flocculated state in a dispersion that is easily reversed by
mechanical energy (Long et al., 1973).
In addition in an aqueous dispersion there will be some charge on the
layer, and hence for this situation
VT = Vs

+ VR + VA

(26)

and a potential energy curve of the form shown in Fig. 20 is obtained. Such
a system will be very well stabilized at low electrolyte concentrations and
will remain stabilized even at very high electrolyte concentrations since Vs
still provides the repulsive interaction energy.
Data which illustrate the latter point were obtained by Ottewill and
Walker (1968, 1974) using polystyrene latices of various sizes and various
electrolytes in th~ presence of nonionic surfactants. Figure 21 shows the
2

C)

...J

-4

-3
Log [La (N03h/mol

-2
dr3J

Fig. 21.
Log W against log concentration of lanthanum nitrate at pH 4.6 for a
polystyrene latex (radius = 52 nm) in the presence of various concentrations of a nonionic
surface active agent, C12E6. -0-,
CI2E6 absent; -e-,
lO-s mol dm-3 C12E6; -D.-,
1.5 x lO-s mol dm-3 C12E6; -0-,
2 x lO-s moldm-3.

R. H. Ottewill

34

curves of log Wagainst log C. obtained using polystyrene particles of 103nm diameter at various concentrations of C12E6 using lanthanum nitrate as
the coagulating electrolyte. It is clear that the ccc moves to higher lanthanum nitrate values as the concentration of C12E6 is increased. At
10-5 mol dm-3 C12E6, well below the critical micelle concentration, the
increase in ccc is over an order of magnitude. Close to the critical micelle
concentration with the system containing 5 x 10- 5 mol dm - 3 C12E6 the
latex was stable even at 0.3 mol dm - 3 lanthanum nitrate. The effectiveness
of C12E6 as a stabiiizing agent to the addition of electrolyte is particlesize dependent its effect being greatest with the smallest particles. This is
demonstrated by the data given in Table VI.
It should be noted, however, in connection with the use of nonionic
surfactants that tlocculation is usually observed at a temperature just below
that of the cloud point of the surfactant. The exact temperature at which
this occurs, however, can depend on the type of salt used and the presence
of other surfactants.
An interesting feature occurs in the use of nonionic surfactants with
PTFE latices. Experiments indicate that PTFE particles adsorb ethylene
TABlE VI
Coagulation of Polystyrene latices by Lanthanium Nitrate at pH 4.6 in
the Presence of C12E6
Number average
particle diameter
(nm)
60:t 10

Total C12E6
eoneentration

Equilibrium C12E6
eoneentration
(mol dm-3)

zero
X 10-6
x 10- 5
x 10- 5
x 10- 5
zero
1 X 10-5
1.5 X 10-5
2 x 10- 5
5 x 10- 5
zero
1 x 10- 5
2 x 10- 5
5 x 10- 5
10-4

zero
2 X 10-6
3 X 10-6
7 X 10-6
-

2.8
6.9
1.7
4.2

zero
6 X 10-6
1 x 10- 5
1.4 X 10-5
-

5.6
3.3
3.2
5.2

x 10-4
X 10-4
X 10-3
X 10-3
sa
x 10-4
X 10-3
X 10-3
X 10-3

zero
9.5 X 10-6
1.9 X 10-5
5.0 X 10-5
10-4

5.2
2.0
2.0
1.0

sa
x 10-4
X 10-3
X 10-3
X 10-2

5
1
2
5

103 :t 8

368 :t 17

S indieates

b cee eould

eee/mol dm - 3
La(N03h

(mol dm - 3)

that eoagulation
not be determined

did not occur in 0.3 mol dm - 3 lanthanium


but very

slow eoagulation

oeeurred.

_b
nitrate.

1.

35

The Stability and Instability of Polymer latices

oxide-type surfactants (Bee, 1978). Since as indicated earlier hydrocarbon


chains are reluctant to adsorb on polytetrafluoroethylene surfaces the
circumstantial evidence is strong that the initial adsorption occurs via
interaction of the ethylene oxide group with the ionie groups on the PTFE
latex partic1es. Indeed experimentally a small drop is noted in the cee value
at low concentrations of C12E6, which supports this idea. Subsequent
adsorption of the nonionic surfactant can then nuc1eate around the hydrocarbon chain of the head-group "down" molecules to give eventually
monolayer coverage and enhanced stability.

XIV.

Mixed Electrolyte Systems

In many cases mixed electrolyte systems can be added to a latex either


adventitiously or deliberately. Their effects on the stability of dispersions
can best be summar~zed by Fig. 22. The axes of this figure are plotted as a
percentage, so that the abscissa is the cee value of salt 1 expressed as a
pereentage of its cee value in the absenee of the second salt; salt 2 is
expressed on a similar basis. Thus, the simplest possible case is additivity in

\
80

C\I
Cii

60

CI)

.. .

\ '.
\
\
\
\

'\.
40

"

"

20
100
Yo

-,

Salt

Fig. 22.
Various effects obtained using mixed electrolyte systems as coagulating agents.
antagonism; ---, superadditivity;
, additivity; -'-', syngergism.

R. H. Ottewill

36

,,

,,

20

,,

,,

,,

,,

,,

,,

,,

...o.100

.
Fig. 23.
Results obtained for the coagulation of a polystyrene latex (radius = 0.109 pm)
using mixtures of magnesium sulfate and sodium nitrate at pH 8.5 and 298.2K. Percentages
are expressed in terms of the molar concentrations of the salts; --- indicates additivity.

which case the plot joining the two values is a straight line. Superadditivity
gives a curved line which is convex to both the abscissa and the ordinate.
Antagonism gives an even more pronounced convexity and the gradient at
the point of intersection on the ordinate is distinctly positive. The fourth
case is that of synergism, Le., when the curve is concave toward the axes.
Some results obtained by Storer (1968) using well-dialyzed polystyrene
latices at pH 8.5 and mixtures of magnesium sulfate and sodium nitrate as
the coagulating electrolytes are shown in Fig. 23. Distinct synergism was
observed over the entire concentration range. The use of activities of the
ions in the mixed systems, rather than concentrations, gave a reasonable
explanation for the form of the data, but it is also of interest that the
discrete-ion treatment of Levine and Bell (1965) also predicts synergism for
certain cases.

XV. Heterocoagulation
In the previous section the strong affinity of cationic surfactants for
negatively charged polystyrene latices was noted. This concept of a single
po sitive ion interacting with a negative charge on a surface can be extended

37

1. The Stability and Instability of Polymer latices

to the interaction between a positively charged particIe and a negatively


charged particIe. Moreover, the particIes of different charges can have the
same chemical composition or be composed of different materials. Thus,
mixing of latices contaning particIes of opposing charge can lead to
coagulation

and this phenomenon

is usually. termed heterocoagulation.

Measurements of heterocoagulation can be made in much the same way


as were measurements for the addition of an electrolyte to a latex. For
example, with an appropriate mixing device (Cheung, 1979) an anionc latex
can be added to a cationic one and the progress followed by turbidity
measurements to obtain a stability ratio. Results of this type are shown in

Fig. 24 in the form of log Wagainst N-/(N + + N-) where N+ and N_are
the number concentrations of the cationic and anionic lattices respectively.
The anionic latex used was composed of polystyrene particIes (diameter

0.1 6

O>
o

...1

0.0 8

0.1

Fig. 24.

0.2

0.3

004

Log W against N_/(N- + N+) for mixtures of an anionic polystyrene latex

(particleradius = 26 nm) and a cationic polystyrenelatex (particleradius = 22 nm). N+ and


N-

= the

number concentration of the cationic and anionic lattices, respectively. Sodium

chloride concentration

= 10-3

mol dm-3.

3~

R. H. Ottewill

52.7 nm) with sulfate. groups on the surface; the cationic latex was also
polystyrene (diameter 43.4 nm) with surface amidine groups. It can be seen
from Fig. 24 that the system becomes completely unstable w~en the ratio
N_/(N+ + N_) reaches -0.25.
When the particles are of different sizes complete coverage of the bigger
particles by the smalIer can occur (Goodwin et al., 1978). This is demonstrated by the scanning electro n micrograph shown in Fig. 25.
An "alternative method of studying heterocoagulation is to use the
spinning disk technique, a method that provides well-defined hydrodynamic
conditions for examining the deposition of spherical particles on planar
surfaces. The method was used by Clint et al. (1973) to examine the
heterocoagulation of polystyrene latex particles (diameter 418 nm) onto
planar polystyrene surfaces in the presence of barium nitrate. The highest
rates .of deposition of the latex particles began to occur at about
0.02 mol dm - 3 barium nitrate. This corresponded to IjIs values of about
10 mV on the particles and about 6 mV on the planar surface. The
experiments were satisfactorily explained using the Levich theory (1962) of
diffusion to a rotating disk, with modification to include a potential energy
of interaction between the plate and a sphere.

Fig. 25.
Scanning electron micrograph showing the heterocoagulation of cationic latex
particles (radius = 22 nm) onto a negativeIy charged particle (radius = 1.07 11m).

1.

39

The Stability and Instability of Polymer Latices

XVI.

Suace Coagulation

There are a number of cases with polymer colloid systems where the
coagulation process can occur at the liquid-air interface under conditions of
electrolyte concentration that are far removed from those required to
produce coagulation in the bulk solution. The processes that occur in this
type of coagulation are illustrated schematically in Fig. 26. The effect was
originally called mechanical coagulation, by Freundlich, but waf! more
appropriately termed surface coagulation by Heller and Peters (1970) who,
in conjunction with several co-workers, have carried out an extensive
investigation of this phenomenon. A theory of surface coagulation was
developed by these authors based on the following assumptions:
1. That the coagulation process proceeded exclusively at the liquid-air
interface and was a biparticle association.
2. That the contribution of aggregates to the rate could be neglected.
Moreover, since the aggregates returned to the bulk phase they had an
insignificant effect on the interfacial area available for occupation by
unreacted primary particles.
3. That sufficient convection occurred to exclude the diffusion of particles to and from the interface as a rate-determining factor.
4. That a steady state occurred for the distribution of primary particles
between the bulk dispersion and the surface, and this distribution could be
described by a Langmuir adsorption isotherm.
5. That the rate of adsorption was la,rge enough relative to the rate of
formation of fresh surface for the adsorption equilibrium to be unaffected.
Thus, putting the Langmuir adsorption isotherm in the form

c/r

= K1

+ K2c

(27)

where c is the concentration of colloidal particles in the bulk phase, r the

surface concentration, and K I and K 2 are constants, and taking the


biparticular surface reaction rate as
(28)
they obtained
(29)
Air
-----..

Water

Dispersion
Fig.26.

Schematic

(5

Adsorption
illustration

co
Surface
Coagulation

of the process of surface coagulation.

Redispersion

40

R. H. Ottewill

When S is the surface area at a constant rate of surface renewal and Vis the
volume of solution, the rate at which the bulk concentration was changed
was given by
-dcjdt = (KOSjV)C2j(K2

+ K2c)2

(30)

From a seriesof experimentsHelleret al. (1970b,1971a,b)found that

1. Colloidal dispersions that required a relatively large amount of


electrolyte in order .to obtain conventional coagulation in the bulk phase
were not susceptible to surface coagulation. .
2. Coagulation at the surface required a low dispersion stability although the latter could be adequate to avoid coagulation occurring in the
absence of a renewable liquid-air interface.
3. The rate data indicated a lack of participation of the secondary and
tertiary aggregates in the surface reaction.
In some cases the aggregates formed by surface coagulation were found
to differ in form from those formed in bulk coagulation processes. They
appeared to be laminar aggregates, as indicated by the "silkiness" exhibited
in mildly agitated dispersions.
The mechanism appears to be connected with dewetting of the particle at
the water-air interface either as a consequence of desorption of stabilizing
surfactant or the fact that the particle surface is not homogeneous. PTFE
latices are particularly prone to surface coagulation and this may be partly
due to their nonspherical shape. The latter is a consequence of crystallinity
and polymer chain folding which may mean that the ionic surface groups at
the chain ends are concentrated on some surfaces, whereas the other
surfaces are devoid of stabilizing entities. Consequently,the latter have a
high contact angle against water. One method of preventing surface coagulation is to store the latex in containers without a water-air interface.

XVll.

Peptization

The reverse of the coagulation process is peptization. It is well known


from analytical chemistry that fresh precipitates are easier to disperse than
old ones, which indicates qualitatively that an aggregate of colloidal
particles is not in equillibrium and that irreversible, time-dependent processes occur in coagulation. As pointed out by Frens and Overbeek (1971)
the interpretation of peptization phenomena with aggregated systems is not
possible unless the data are obtained in experiments with a shorter time
scale than the aging time of the aggregate. They demonstrated that it was
possible to follow the kinetics of peptization by suddenly diluting the sol

t
41

1. The Stability and Instability of Polymer Latices

containing electrolyte after a short period of coagulation. From their


experiments they concluded that peptization was a rapid, spontaneous
process and that electrical double-Iayer repulsion probably provided the
driving force.
The process of peptization can be explained in terms of potential energy
diagrams following the suggestions of Overbeek (1977). The relevant diagrams are shown in Fig. 27 for systems that do not have pronounced
secondary minima. In Fig. 27a, for the reason explained earlier, peptization
is unlikely to occur since going from L to R involves an increase in free
energy and a substantial energy of activation. In Fig. 27b the transition
from L to R involves a free energy decrease, which is favorable for
peptization; but stilI a high activation energy barrier has to be surmounted
and again peptization is unlikely to occur. The most favorable conditions
for peptization are those ilIustrated in Fig. 27c and d; these involve a decrease in free energy and only a small activation energy barrier of the order
of 1 kT. Thus, the question arises as to how can this be achieved in practice.

>-

Q)

c:

L-- ( a)

(b)

Distance

-------

Fig.27.
repeptization.

AVb-kT

AVb-kT

(e)

(d)

Potential energy curves to iIIustrate the differences between coagulation and

42

R. H. Ottewill

One possibility is to provide the particles with a thin steric barrier such as
the "micro hairs" which could occur on the surface of a latex by solvation
of the polymer chain beyond the ionic end group. Indeed some evidence for
this occurs with certain latices. For example, Smithan et al. (1973) have
reported evidence of steric stabilization with polystyrene latices with a high
content of carboxyl groups on the surface prepared by an essentially
conventional emulsion polymerization method. Microsteric stabilization
with latices could be an important factor and this is undoubtedly an area
that needs more extensive investigation.

XVffi.

The Effects of Adsorbedor Grafted Macromolecules

Space prohibits a detailed discussion of this topic but a few general


points can be made following the comments made in the previous section.
With hairy particles of the type shown in Fig. 3, polyelectrolyte molecules
can be chemically linked to the surface or adsorption can occur by several
mechanisms, including ionic bonding-particularly
via charges of opposite
sign-hydrogen bonding, couplig with multivalent inorganic ions, and by
hydrophobic bonding of the hydrophobic regions of the marcomolecule
to the surface.
The net result is shown schematically in Fig. 28. Instead of the array of
surface charges leading to a well-defined surface charge density and surface
potential there is now a distribution of charges in space which contri bu te to
the electrical double layer surrounding the particle. At low electrolyte
concentrations the latter will extend into the space beyond the polyelectrolyte layer so that VR will be significant. However, with increase in
electrolyte concentration and compression of the electrical double layer, VR
can become small or zero. However, under these conditions the particle will
still be coated with an extensively hydrated layer of polymeric molecules
which provide a steric barrier V.. Hence, this type of system provides a twotier mechanism of stabilization against electrolyte additions and the classical protective agents for colloidal particles such as gelatin, gum arabic, etc.
almost certainly act in this way.
A number of polymer latices falling into this category have been
described in the literature but they are still relatively novel and have not
received the extensive attention given to the more conven.tional latices.
Probably the systems of this type most extensively characterized are those
described by Hoy (1979) and Bassett and Hoy (1980) which were prepared
by copolymerizing methyl methacrylate, butyl acrylate, and ethyl acrylate
with an unsaturated acid such as itaconic, acrylic, or methylacrylic. The
particles obtained appeared to consist of a spherical core particle sur-

t~

43

1. The Stability and Instability of Polymer latices

I
I
/
POLYION
/

LAYER
/

/'

Fig. 28. Schematic ilIustration of a latex partic1e with an adsorbed or grafted layer of
polyelectrolyte.
represents the extension in space of the adsorbed layer, and --- represents
the extension in space of the electrical double layer. 'Reproduced with permission. of American
Chemical Society.

rounded by an acid-bearing polyion shell. The latter expanded at high pH


values as the acid groups were neutralized to give a structure similar to that
shown schematically in Fig. 28. These authors have carried out extensive
ultracentrifugation studies on this type of system as a function of pH in
order to determine values for the expanded shel1thickness (see c5in Fig. 17).
Latices prepared by grafting polyacrylate chains onto a polystyrene core
have also been described recently by Buscall and Comer (1980). These
authors also examined the stability behavior of their latices as a function of
the degree of neutralization of the polyacrylate and as a function of
temperature at different electrolyte concentrations. As can be seen from the
results given in Fig. 29 the systems were stable over a certain range of
temperature and degree of neutralization but flocculated both on heating
and on cooling. The behavior at the upper temperature appears to be
similar to that observed with nonionic surfactants as stabilizing molecules
at or near the c10ud point of the surfactant and basically arises as a

R. H. Ottewill

44

20

Fig. 29.
Data iIIustrating the temperature behavior of a latex with a polystyrene core
stabilized by grafted poly(acrylic acid) in 1.10 mol dm - 3 sodium chloride solution; IX= degree
of neutralization of the latex.

consequence of desolvation of the chains; in the polymer terms the cloud


point would be close to the () temperature. The authors explain the lower
temperature behavior as a consequence of the dissimilarity between the free
volume of the polymer and the solvent.
The behavior of this type of polyelectrolyte system is of fundamental
interest and it is hoped that in the near future more detailed stability studies
will be reported. An interesting feature from the point of view of polymer
morphology is whether all the chains are on the surface or some are buried in
the particle.
The behavior of aqueous latex dispersions, in which the polymer core
particles were stabilized by block copolymers of poly(ethylene oxide) and a
vinyl or acrylic monomer, has been investigated in some detail by Napper
(1969). The particles in these latices were shown by electrophoresis to be
noncharged. Flocculation occurred when the solvency of the dispersion
medium for the polymer chains was decreased as, for example, by raising
the temperature when a critical flocculation temperature was observed that
was found to be insensitive, over a limited range, to the molecular weight of
the poly(ethylene glycol) chains. The influence of a number of other factors
on the stability of the latices was also investigated (Napper, t970a) including the nature of the anchoring groups of the stabilizing polymers, the

1. The Stability and Instability of Polymer Latices

45

nature of the disperse phase, the particle size, the surface coverage and the
molecular weight of the stabilizing polymer. It was found that to obtain
colloid stability, it was necessary to use a dispersion medium that was better
than a () solvent for the stabilizing chains.
The implication of this observation was that the second virial coefficient
of the stabilizer needed to be po sitive so that the segmental excluded
volume was also positive. Under these conditions once overlap of the
stabilizing polymer layers occurred, the configurational entropy of the
molecules in the overlap region would become less than that of the
molecules in the dispersion medium and an excess osmotic pressure would
occur in the overlap volume. As a consequence of this, molecules of the
dispersion medium would difIuse into the overlap region, forcing the
stabilizing layers apart. This is essentially the basis upon which Eq. (24) was
formulated. It was found that with this type of system that peptization
could be achieved spontaneously after centrifugation or flocculation,' as
would be anticipated from Fig. 27d.
Napper (1970b) also investigated the flocculation of poly(vinyl acetate)
particles stabilized by poly(ethylene glycol) chains with a series of electrolytes. The order for the cations was

The controlling factor appeared to be the capability of the ion to convert


water into a ()solvent for the stabilizing chains.

XIX. Particle Stability in Emulsion Polymerization


Any consideration of the stability of polymer latices would be incomplete without some discussion of the stability of the colloidal polymer
particles formed during the course of an emulsion polymerization. As
pointed out by Dunn and Chong (1970) the adsorption of the emulsifier
plays a major role in determining the surface charge density of the particle
and hence in determining the final particle size.
The case in which there is an absence of added emulsifier has been
considered by Goodwin et al. (1978) on the basis that" the particulate units
initially formed contain only a small number of chains; they therefore have
a low surface charge and are colloidally unstable. Hence, coagulation
occurs until the particles formed reach values of surface charge density and
radius sufIciently large to render them stable colloidal particles. The
arguments can be developed in terms of the stability ratio starting with Eq.
(13), which gives the rate of coagulation, and recapitulating that as W
becomes greater than unity the rate of coagulation is reduced. Some

R. H. Ottewill

46

fundamental questions therefore arise, namely, At what size does the


particle become a colloidally stable entity and at this, point what is its
surface charge density and how many polymer chains does it contain? A
further point of importance is to understand how this size varies with the
ionic strength of the aqueous phase.
In order to obtain a qualitative understanding of these points we can
proceed by making some simple assumptions. These are (i) the particles
formed are spherical, (ii) each polymer chain has the same molecular weight
Me, (iii) each chain has two end groups, and (iv) all the end groups are
anchored on the surface of the particle. Hence, if the latex particle has a
molecular weight M L and a density of PL, then the number of polymer
chains per particle is given by
Ne

= 4na3PLNAi3Me = MdMe

(31)

The number of charged end groups per particle is therefore given by


Ne

= 2Ne = 8na3 PLN Ai3Me

(32)

and the surface charge density by


(33)
Thus, as is directly proportional to a. For spherical particles the surface
potential1/ls is given by
1/Is

= 4naa./[B(1 +

Ka)]

(34)

an equation which holds reasonably well up to 1/Isvalues of 50 mV. From


this we find that, taking Me = 150,000: for a = 5 nm, 1/Is= 8 mV; for
. a = 10 nm, 1/1.~ 20 mV; and for a = 22 nm, 1/1.~ 50 mV. Using a combination of Eqs. (4) and (6) to calculate VT as a function of h it is then
possible to calculate W by numerical integration of Eq. (14). Since 1/Isis
known as a function of particle size, then W can also be obtained as a
function of r at an appropriate ionic strength. The results obtained are
shown in Fig. 30 in the form of curves of log W against a. It is clear from
these curves that the size of the first stable colloidal particle formed is
controlled to a large extent by the ionic strength of the dispersion medium,
i.e., at 4 x 10-4 mol dm - 3 log W = 2 is achieved with a = 3.7 nm, where to
achieve the same Wat 4 x 10-3 mol dm-3 a has to grow to 11.3 nm. Since
the size of the initial stable particles controls the nu~ber concentration of
the latex during the diffusional growth period, then for the same initial
monomer concentration and for the same percentage conversion of monomer, the final particle diameter in the medium of higher ionic strength
will be the larger. This conforms to the clear trend found in the preparation

47

1. The Stability and Instability of Polymer Latices


12

10

:;:
O>

+
2

.'
5

10
Radius

15

20

(nm)

Fig. 30. Log Wagainst latex partic1eradius as a function of the concentration of 1: 1


electrolytein the system; ,4 x 10-4 mol dm-3; -,
4 x 10-3 mol dm-3.

of polystyrene latices in the absence of added emulsifier (Goodwin et al.,


1976).
XX.

Summary

In this chapter I have attempted to show in a broad sense how the


application of the basic principIes of colloid science can be applied to
develop our understanding of the various mechanisms involved in the
stabilization of polymer latices. In the space available, it was not possible
to go into very specific details of the many systems that have been
investigated nor to deal with nonaqueous polymer latices. The latter,
however, have been discussed in the recent comprehensive book by Barrett
(1975). The literature on polymer colloids appears to be growing exponentially and to the authors of the many excellent papei-s which I have not
quoted, I offer my sincere apologies.
References
Barrett, K. E. J. (1975). "Dispersion Polymerization in Organic Media." Wiley, New York.
Bassett, D. R., and Hoy, K. L. (1980). In "Polymer Colloids 11" (R. M. Flitch, ed.), pp. 1-25.
Plenum Press, New York.

R. H. Ottewill

48

Bee, H. (1978). B.Sc. thesis, Univ. of Bristol.


Bibeau, A. A., and Matijevi, E. (1973). J. Colloid Interface Sci. 43,330.
Bolt, P. (1978). B.Sc. thesis, Univ. of Bristol.
Buscall, R., and Comer, T. (1980). Org. Coat. Plast. Chem. 43, 203.
Cebula, D. J., Thomas, R. K., Harris, N. M., Tabony, J., and White, J. W. (1978). Faraday
Discuss. Chem. Soco 65, 76.
Cheung, W. K. (1979). Ph.D. thesis, Univ. of Bristol.
Clint, G. E., Clint, J. H., Corkill, J. M., and Walker, T. (1973). J. Colloid Interface Sci. 44,121.
Connor, P. (1968). Ph.D. thesis, Univ. of Bristol.
Connor, P., and Ottewill, R. H. (1971). J. Colloid Interface Sci. 37, 642.
Comell, R. M., Goodwin, J. W., and Ottewill, R. H. (1979). J. Co/loid Interface Sci. 71, 254.
Derjaguin, B. V., and Landau, L. (1941). Acta Physicochim. URSS 14, 633.
Derjaguin, B. V., and Muller, V. M. (1967). Dokl. Phys. Chem. 176, 738.
Doroszkowski, A., and Lamboume, R. (1971). J. Polym. Sci. Part C 34,253.
Dunn, A. S., and Chong, L. C-H. (1970). Br. Polym. J.2. 49.
Dzyaloshinskii, 1. E., Lifshitz, E. M., and Pitaevskii, L. P. (1961). Adv. Phys. 10, 165.
Evans, R., and Napper, D. H. (1973). J. Co/loid Interface Sci. 45, 138.
Fitch, R. M. (1980). "Polymer Colloids 11." Plenum Press, New York.
Flory, P. J., and Krigbaum, W. R. (1950). J. Chem. Phys. 18, 1086.
Force, C. G., and Matijevi, E. (1968a). Kolloid Z. Z. Polym. 224, 51.
Force, C. G., and Matijevi, E. (l968b). Ko/loid Z. Z. Polym. 225,33.
Fowkes, F. M. (1964). Ind. Eng. Chem. 56, 40.
Fowkes, F. M. (1967). In "Surfaces and Interfaces" (J. J. Burke, ed.), Vol. 1, p. 199. Syracuse
Univ. Press, New York.
Frens, G. (1968). Doctoral thesis, Univ. of Utrecht.
Frens, G., and Overbeek, J. Th. G. (1971). J. Co/loid Interface Sci. 36, 286.
Friends, J. P., and Hunter, R. J. (1971). J. Co/loid Interface Sci. 37, 548.
Fuchs, N. (1934). Z. Phys. 89, 736.
Gingell, D., and Parsegian, V. A. (1973). J. Co/loid Interface Sci. 44, 456.
Goodwin, J. W., and Ottewill, R. H. (1978). Faraday Discuss. Chem. Soco 65, 338.
Goodwin, J. W., Heam, J., Ho, C. c., and Ottewill, R. H. (1976). Co/loid Polym. Sci.60, 173.
Goodwin, J. W., Ottewill, R. H., Pelton, R., Vianello, G., and Yates, D. E. (1978). Br. Polym. J.

10,173.

Goodwin, J. W., Ottewill, R. H., and Parentich, A. (1980). J. Phys. Chem. 84, 1580.
Gregory, J. (1975). J. Co/loid Interface Sci. SI, 44.
Hachisu, S., Kobayashi, Y., and Kose, A. (1973). J. Co/loid Interface Sci. 42, 342.
Hamaker, H. C. (1937). Physica 4, 1058.
Hardy, W. B. (1900). Proc. R. Soco London Ser. A 66, 110; Z. Phys. Chemie. 33.385.
Heller, W., and Peters, J. (l970a). J. Co/loid Interface Sci. 32, 592.
Heller, W., and Peters, J. (1970b). J. Co/loid Interface Sci. 33, 578.
Heller, W. and de Lauder, W. B. (l97Ia). J. Co/loid Interface Sci. 35, 60.
Heller, W. and de Lauder, W. B. (197Ib). J. Co/loid Interface Sci. 35, 308.
Hiltner, P. A., and Krieger, 1. M. (1969). J. Phys. Chem. 73, 2386.
Honig, E. P., Roeberson, G. J., and Wiersema, P. H. (1971). J. Colloid Interface Sci. 36, 97.
Hoy, K. L. (1979). J. Coat. Technol. SI, 27.
Kayes, J. B. (1976). J. Co/loid Interface Sci. 56,426.
Kitchener, J. A., and Schenkel, J. H. (1960). Trans. Faraday Soco 56, 161.
Kratohvil, S., and Matijevi, E. (1976). J. Co/loid Interface Sci. 57, 104.
Krupp, H., Schnabel, W., and Walter, G. (1972). J. Co/loid Interface Sci. 39,421.
Levich, V. G. (1962). "Physico-chemical Hydrodynamics." Prentice Hall, Englewood ClilTs,
New Jersey.

,,

~
r

1. The Stability and Instability of Polymer Latices

49

Levine, S., and Bell, G. M. (1965). J. Colloid Sci. 20, 695.


Lichtenbelt, J. W. Th., Pathmamanoharan, C., and Wiersema, P. H. (1974). J. Colloid Interface
Sci. 49, 281.
Long, J., Osmond, D. W. J., and Vincent, B. (1973). J. Colloid Interface Sci. 42, 545.
Mardle, R. (1980). B.Sc. thesis, Univ. of Bristol.
Matijevi, E. (1977). J. ColJoid Interface Sci. 58, 374.
Matijevi, E., and Force, C. G. (1968). KolJoid Z. Z. Po/y. 225,33.
Matijevi, E., Janauer, G. E., and Kerker, M. (1964). J. ColJoid Interface Sci. 19, 333.
Napper, D. H. (1969). J. ColJoidInterface Sci. 29, 168.
Napper, D. H. (1970a). J. ColJoid Interface Sci. 32, 106.
Napper, D. H. (1970b). J. ColJoidInterface Sci. 33, 384.
Napper, D. H. (1977). J. ColJoid Interface Sci. 58, 390.
Neiman, R. E., and Lyashenko, O. A. (1962). ColJoidJ. USSR (English Trans.) 24, 433.
Ninham, B. W., and Parsegian, V. A. (1970). J. Chem. Phys. 52, 4578.
Ottewill, R. H. (1980). Chem. Ind. 377.
Ottewill, R. H., and Rance, D. G. (1977). Croatica Chem. Acta SO,65.
Ottewill, R. H., and Rance, D. G. (1979). Cr,oatica Chem. Acta 52, 1.
Ottewill, R. H., and Shaw, J. N. (1966). Discuss. Faraday Soco 42, 154.
Ottewill, R. H., and Shaw, J. N. (1967). KolJoid Z. Z. Po/y. 218, 34.
Ottewill, R. H., and Walker, T. (1968). KolJoid Z. Z. Po/y. 227, 108.
Ottewill, R. H., and Walker, T. (1974). J. Chem. Soco Faraday 170, 917.
OttewilI, R. H., Rastogi, M. C., and Watanabe, A. (1960). Trans. Faraday Soco 56, 854.
Overbeek, J. Th. G. (1952). In "ColIoid Science" (H. Kruyt, ed.), Vol. 1. Elsevier, Amsterdam.
Overbeek, J. Th. G. (1977). J. ColJoid Interface Sci. 58, 408.
Parsegian, V. A., and Ninham, B. W. (1971). J. Col/oid Interface Sci. 37, 332.
Pelton, R. (1976). Ph.D. thesis, Univ. of Bristol.
Rance, D. G. (1976). Ph.D. thesis, Univ. of Bristol.
Reerink, H., and Overbeek, J. Th. G. (1954). Discuss. Faraday Soco 18, 74.
Richardson, R. (1979). B.Sc. thesis, Univ. of Bristol.
Richmond, P. (1975). In "ColIoid Science" (D. H. Everett, ed.), Vol. 2, p. 130. Chemical Society,
London.
Schulze, H. (1882). J. Prakt. Chem. 25, 431.
Schulze, H. (1883). J. Prakt. Chem. 27, 320.
Smitham, J. B., Gibson, D. V., and Napper, D. H. (1973). J. ColJoid Interface Sei. 45, 211.
Storer, C. S. (1968). Ph.D. thesis, Univ. of Bristol.
Tamaki, K. (1960). KolJoid Z. 170, 113.
van den HuI, H. J., and Vanderhoff, J. W. (1970). Br. Po/y. J. 2, 121.
Verwey, E. J. W., and Overbeek, J. Th. G. (1948). "Theory of the Stability of Lyophobic
ColIoids." Elsevier, Amsterdam.
Visser, J. (1972). Adv. ColJoid Interface Sei. 3, 331.
von Smoluchowski, M. (1917). Z. Phys. Chem. 92,129.
Wiese, G., and Healy, T. W. (1970). Trans. Faraday Soco 66, 490.

~
\

2
Particle Formation Mechanisms
F. K. Hansen and John Ugelstad

1.
11.
111.
IV.
V.
VI.
VII.

Introduction.
Micellar Nucleation: The<Smith-Ewart Theory .
Radical Absorption Mechanisms.
".
Micellar Nucleation: Newer Models .
Homogeneous Nucleation.
Partide Coagulation during the Formation Period.
Nucleation in Monomer Droplets
References .

51
54
56
63,
73
82
86
91

l. Introduction
The nuc1eation stage constitutes the so-called Interval I in an
emulsion polymerization, the initial period in which the partic1e number is
changing. In Intervals n and nI the paftic1e number is believed to be
essentially constant. Nuc1eation of new partic1es may in some cases also
take place during Intervals n and In. This phenomenon is often referred to
as secondary huc1eation and may be encountered in systems with poor
stability (coagulation) or with changing composition (continuous and semicontinuous polymerizations). The present chapter will attempt to treat all
mechanisms that may lead to formation of polymer partic1es, in whatever
stage of the polymerization they take place.
All discussions of partic1e nuc1eation start with the Smith-Ewart theory
in which Smith and Ewart (1948) in a quantitative tre!ltment of Harkins'
micellar theory (Harkins, 1947, 1950) managed to obtain an equation for
the partic1e number as a function of emulsifier concentration and initiation
and polymerization rates. This equation was developed mainly for systems
of monomers with low water solubility (e.g., styrene), partly solubilized in
micelles of an emulsifier with low critical micelle concentration (CMC) and
p'?"Ts{lIledto work well for such systems (Gerrens, 1963). Other authors have,
,

however, argued against the Smith-Ewart theory on the grounds that (i) .
partic1es are formed even if no micelles are present, (ii) the equation for the
51
EMULSION POLYMERIZATlON

Copyright; 1982 by Academic Press, Inc.


Al! rights of reproduction in any form reservcd.ISBN 0-12-556420-1

52

F. K. Hansen and J. Ugelstad

partic1e number gives an estimate that is a factor of 2 higher than that


found experimentally even for styrene, (iii) more water-soluble monomers
do not fit the theory, and (iv) a maximum in polymerization rate at the end
of the nuc1eation period is predicted, but has rarely been observed. On this
basis other theories for the nuc1eation have been put forward, based on the
idea of self-nuc1eation of oligomer radicals produced in the aqueous phase.
These mechanisms of partic1e formation were first treated quantitatively by
Fitch and Tsai (1971). These ideas, which have been further elaborated by
other authors, seem to have solved the first problem of the theory of partic1e
nuc1eation but leave open the question of whether the micelles, when
present, play any role at all in partic1e formation. More recent work seems
not only to confirm the importance of the micelles but to stress the necessity
of inc1uding more detailed features of absorption and reaction in micelles .
and particles in order to explain the other discrepancies of the Smith-Ewart
theory.
Usually, monomer droplets are believed not to play any role in'emulsion
polymerization other than as a source of monomer. Ugelstad and associates
ha ve shown, however, that in cases with very small monomer droplets, these
may become an important, or even the sole, loci for partic1e nuc1eation. The
system may then be regarded as a microsuspension polymerization with
water-soluble initiators. It has therefore been pointed out (Hansen and
Ugelstad, 1979c) that partic1e nuc1eation models should inc1ude a~l three
initiation mechanisms-micellar, homogenous, and droplet-since all these
mechani~ms may compete and coexist in the same system, even if one of
theni usually dominates.
Figure 1 illustrates the main components and phases in an emulsion
polymerization system. The arrows indicate the possible distribution ofthe
components between the phases.
Monomer (i) will usually exist as monomer droplets; (ii) some monomer,
depending on water solubility, will be dissolved in the continuous phase;
and (iii) some monomer will be solubilized in micelles.
Emulsifier (i) will be partly dissolved in the continuous phase; (ii) if
concentration is above the CMC, the excess will form emulsifier inicelles;
and (iii) some emulsifier will be adsorbed on monomer droplets, and may
even be dissolved into the droplets.
lnitiator (i) will mostly be dissolved in the continuous phase as watersoluble initiators are usually applied. For special applications partly or
completely oil soluble initiators may be used. These will be distributed
similarily to the monomer.
A so-called ordinary emulsion polymerization, Le., similar to the case
treated by Smith and Ewart, is characterized by large monomer droplets

'2. Particle Formation Mechanisms

53

Honamer
H

Hice//es

Drap/ets

nitiatar

Emu/sifier

,
Aqueaus
sa/utian
Fig. 1.

Schematic iIIustrationsof the components and phases usually present in an

emulsion polymerization
among phases.

system. The arrows indicate

the possible distribution

of components

with consequently smaIl total surface, miceIle-forming emulsifiers with


reIatively low CMC, and ionic initiators that decompose either by a thermal
andjor by a redox mechanism.
In the Smith-Ewart theory as weIl as in other treatments, the primary
radicals formed by decomposition of the initiator in the continuous phase
were assumed to enter emulsifier miceIles and polymer partieles. It has been
pointed out by several authors (Alexander and Napper; 1971, Nomura et
al., 1975; Barrett, 1975) that these usuaIly ionic, very water-soluble radicals
are rarelyabsorbed directIy into a miceIle or partiele but must add some
monomer units in the aqueous phase to become sufficientIy oil soluble to be
absorbed. It now seems to be generaIly accepted that formation of these
oligomers in the aqueous phase is the first step in the nUeleation (and
polymerization) process. The presen'ce and molecular weight of oligomers in
some systems have been analyzed by means of GPC and spectrophotometric methods (Fitch and Tsai, 1971; GoodaIl et al., 1975; Chen and Piirma,
1980). Degrees of polymerization have been found to lie between 1 and 6070. The water-soluble oligomers may be destroyed or may nUeleate partieles. The different possibilities are listed in Table I.
It should be added that oligomers that are surface active may also be
adsorbed onto the surface of partielesjdropletsjmiceIles rather than being
absorbed into the interior. Being on the end of the oil-soluble chain, the
active site of the radical wilI no doubt' be able to propagate into the interior.

11

54

F. K. Hansen and J. Ugelstad


TABLEI
Reaction Possibilities of an Oligomer Radical in the Aqueous Phase and
the Probable Result
Process

A
B
e
D
E
F
G

Result
Micellar initiation

Absorption into a micelle


Absorption into a monomer droplet
Absorption into an earlier formed
particle
Propagation in the aquous phase
Termination in the aquous phase

Droplet initiation.
Radical disappearance

(particle growth)

Higher oligomers
"Dead" oligomers (may or may not lead
to nucleation)
HoIt1ogeneous nucleation

Precipitation in the aquous phase


(self-nucleation)
Mixed-micelle formation
(with or without emulsifiers)

Nucleation

(homogeneous

or micellar)

This may lead to a core and shell morphology. Process E in Table 1,


termination in the aqueous phase, may also lead to homogeneous nucleation if the "dead" molecules are sufficiently water insoluble. When
partic1es have been formed, transfer reactions to monomer (or chain transfer
agent if present) willlead to monomer radicals, which may be desorbed into
the aqueous phase. This is also indicated in Fig. 1. The monomer radicals
may act in a way similar to the initiator radicals. ~elow is given a detailed
description of the different processes listed in Table 1 and of the theories
that have been advanced for these processes.

n.

MiceUar Nucleation: The Smitb-Ewart Theory

Smith and Ewart (1948) proceed as follows: radicals are absorbed into
monomer-swollen emulsifier micelles which then are transformed into polymer partic1es; the rate of radical absorption Is equal to the rate of initiator
decomposition p, which means that
dN/dt

= p

(1)

where N is the partic1e number. The rate of growth of a partic1e is assumed


to be constant and is expressed as
dv/dt

J1.

(2)

where v is the volume of the partic1e. The number of micelles will decrease
as the partic1es grow, giving an increasing surface which will adsorb

2.

55

Particle Formation Mechanisms

emulsifier. Particle formation stops when a11emulsifier is adsorbed on the


particles, which means that the particle surface area Ap is equal to the total
surface area of emulsifier asS where as is the specific surface area per unit of
emulsifier and S the amount of emulsifier. To be precise S should be the
amount in excess of the CMC. The difference will not be of any importance
if the CMC is much lower than S. As dv/dt is assumed to be constant, Ap
may be expressed by J1.and t by integration over a11formation times from O
to t
(3)
where
(4)
From Eq. (3) one obtains the time tl when Ap = asS,which inserted in Eq.
= Pitl' The result is

(1)givesthe particle number N

(5)

This is the so-ca11edupper limit equation. The lower limit is derived in a


similar way. In this case it is also assumed that polymer particles may
absorb radicals leading to a decrease in the rate of nucleation. For
computational purposes the rate of radical absorption is set as proportional
to the particle surface area, Ap =

api'

where

apj is the surface

area of one

particle. Equation (1) is then transforme9 to


(6)
The total area Ap is expressed by an integral equation in a way similar to
that for the upper limito Inserting for Ap in Eq. (6) gives by a somewhat
complicated integration
(7)
which is identical to Eq. (5) except for the constant. From diffusion laws
(Fick's first law) it is expected that the radical absorption rate is proportional to the particle surface area divided by the radius (i.e., 4nrN) so
that sma11 particl~s absorb more radicals per unit area than do large
particles. This fact was realized by Smith and Ewart who stated that the
true value of k should be between 0.37 and 0.53 and accordingly the particle
number somewhere between those values predicted by Eqs. (5) and (7). It
may be shown (Section IV) that this procedure is not quite correct, the
constants and ~will also be slightly altered (the former decreased and the
latter increased). In addition the constancy of dv/dt = is somewhat doubtfuI, both because the average radical number per particle in the lower case wilI

J1.

56

F. K. Hansen and J. Ugelstad

not be constant (willdecrease)and because the monomer concentration in


the particles is expected to increase with increasing size. These two factors
will counteract each other and are not expected to have a great influence on
the value of N.
The Smith-Ewart theory has been successful in describing the experimental results with some systems,especiallythe predicted orders of N with
respect to initiator and emulsifier as obtained from double logarithmic plots
of N against the two variables (Gerrens, 1963). Other authors have,
however,found a wide range of exponents(citedby Fitch, 1973).AIso,other
discrepancies exist, as mentioned in the introduction. The Smith-Ewart
theory has been modified and recalculated by several worker~ (Parts et al., 1965; Gardon, 1968a-f, 1971; Harada et al., 1972). Parts et al. applied a
numericaf integration of the nucleation equations and reached the same
conclusions as Smith and Ewart. They found ihat the average number of
radicals per particle is approximately 1 through the entire nucleation
period because the rate of radical adsorption in micelles is so much higher
than that in new particles (very large number of micelles). They propose
that the absorption efficiency of micelleshas to be lower in order to explain
experimental findings. Gardo~ has recalculated the lower limit of the
Smith-Ewart theory by a seminumerical method and finds that decreases
from 1 to 0.67 during the formation periodo This does not, however,
significantly influence the exponents 0.4 and 0.6. The particle numbers (or
more correctly, particle sizes) calculated by Gardon were found to describe
some experimental results for styrene and methyl metacrylate fairly well,
whereas other data on particle numbers were 2-3 times lower than
predicted.
Another feature of the Smith-Ewart theory is that the reaction rate at
the end of the nucleation period is expected to be higher than in the steady
state because is higher than the steady-state value of 0.5 (Smith-Ewart
Case 2 kinetics). There is little experimental evidence for such a maximum
in rate (Ugelstad and Hansen, 1976), and this discrepancy may be explained
by more details about the radical absorption rates in micelles and particles.
Before any further discussion of particle-formation mechanisms, it therefore
seems logical-to review the mechanisms responsible for radical absorption,

111. Radical Absorption Mechanisms

Gardon based his calculations on a geometric derivation of the radical


absorption rate which gave the result that the rate should be proportional
to the particle surface area. This derivation, which also was adopted by
Fitch and Tsai (1971), has been criticized for not taking the concentration

2. Partiele Formation Meehanisms

57

gradient of radicals around a particle into account (as in Fick's laws)


(Barrett, 1975; Ugelstad and Hansen, 1976; Hansen and Ugelstad, 1978).
This was also realized by Fitch and Shih (1975) who, using seeded experiments, found the expected proportionality of the absorption rate to Nr.
Similar conclusions may also be drawn from the seed experiments of Gatta
et al. (1969), the kinetic results of Ugelstad et al. (1969) with vinyl chloride,
and the recent findings of the authors (Hansen and Ugelstad, 1979a) using a
polystyrene seed. Fick's first law for the diffusion of a component A in a
stationary medium B around a spherical particle may be written (Byron
Bird et al.; 1960)
(8)
where 1)ABis the diffusion constant for A in B, e is the total molar,
concentration of A and B, and XA is the molar fraction of A. For the
geometric configuration see Fig. 2.
'\

\
\
\

\
\

\
1
I

Halar fraction,
Concentration

r.e

XAW'CW

---

------

R Distance

Fig. 2.
Geometric and concentration conditions around an absorbing spherical partic1e. e is concentration and X is molar fraction of dilfusing species.

F. K. Hansen and J. Ugelstad

58

Assuming a stationary diffusion layer outside the particle of thickness b


one has for dXAldR.
dXAldR

= (r/R2)[(r + b)/b](l -

Inserting into Eq. (8) for R

XA) In[(1

XAa)j(l - XAw)]

(9)

=r
(10)

Usually the molar fraction of the diffusing species (A) is much smaller
than that of the' stationary component (B) so that 1 - XA ~ 1 and
In(l - XA) ~ XA, which gives
(11)
where
(12a)
and
(12b)
Ir b ~ r, Eq. (11) simplifies to.
A = -Dw4nr(Cw- Ca)

(13)

Ir one has an electrostatic repulsion between charged oligomers and


equally charged particles the diffusion rate is given by (Hansen and
Ugelstad, 1978)
(14)
A = -4nDwr(Cw - CaeZ)/W'
where z = et/lo/kTand W' is given by
,

W=r-

r+~

r+ b
b f.r

et/lodR

exp--

( kT ) R2

(15)

where R is the distance from the oligomer to the center of the particle, e is
the electronic charge, and t/lois the (effective)surface potential.
The value of W' has been calculated by numerical integr~tion for the
case where b ~ r and has been given as a function of et/lo/kT and Kr where K.
is the inverse double-Iayer thickness (Hansen and Ugelstad, 1978). The term
eZin Eq. (14) may be considered as an activation energy (Boltzmann factor)
and accounts for the fact that in order to become absorbed the charged
radical s have tq surmount an energy barrier at the surface. The'term W' is the'
integral of this activation energy factor giving the retardation of a unit charge
upon diffusion to the surface. In the case where Cw~ Caez Eq. (14) reduces
to'
(16)

\.

..

2.

59

Particle Formation Mechanisms

Equation (16)describesthe situation with irreversiblediffusionand seems


to apply to cases of relatively large particles. The folIowing simplifications
are inherent in Eq. (16):

1. The concentration of the diffusing component has to be low (lower


than 1 M). This condition is always fulfilIed for free radicals.
2. The stationary layer "thickness" (<5)has to be much larger than the
particle size. UsualIy, <5is believed to be on the order of 10 .,tm,decreasing
with increasing stirring intensity. For particles in the nucleation range (220 nm) this condition is very welI fulfilIed; for seeded experiments with
larger particles the assumption may be doubtful.
3. The .concentration of particles has to be low, so that the "stationary
layers" do not overlap. This condition is equivalent to saying that <5will be
dependent on particle size, with the same consequences as above.
4. The concentration at the particle surface must be much smalIer than
in the bulk, meaning that the particles must act as a radical "sink."
This last condition is the most doubtful, especialIy in the case of smalI
particles. A thorough treatment of diffusion rates, based on Eq. (13), has
recentIy been presented (Hansen and Ugelstad, 1978). In this derivation the
concentration profile inside the particles, as illustrated in Fig. 2, is also
taken into account. The derivation is based on the equations for diffusion
with chemical reaction (Danckwerts, 1951), expressing the steady-state rate
of diffusion at the inside particle surface by
A

= -4nrDpC*(X ooth X - 1)

(17)

where Dp is the diffusion constant inside the particles and

= r(k/Dp)1/2

(18)

= kpMp + nktp/vp

(19)

with

Here kp is the propagation constant, Mp the monomer concentration in the


particles, n the number of radicals in the particle, ktp the bimolecular
termination constant in particles (in molecular units), and vp the particle
volume.
.
The total rate of absorption is obtained by assuming a rapidly established equilibrium distribution for the radical between the inner and the
outer surface of the particle. Assuming no radical accumulation at the
interface the value of A in Eq. (14) is equal to that in Eq. (17). The
equilibriu.m distribution between the inner and outer surface of the particles
is expressed by
C*

= aCa

(20)

60

F. K. Hansen and J. Ugelstad

leading to the following expression for the rate of absorbtion per particle

nA= 4nrDwCwF

(21) ,

where the "efficiency factor" for absorption F is given by the expression


1/F = (Dw/aDp)[1/(X coth X - l)]eZ + W'

(22)

Note that in Eq. (21) nA has a positive sign since it stands here for rate of
absorption.
It will appear from Eqs. (21) and (22) that if
W' ~ (Dw/aDp)[1/(X coth X

1)]eZ

the value of nAwill in any case be given by


nA = 4nrDwCw/W'

(23)

which corresponds to Eq. (16). Equation (23) describes the situation when
the transport of the oligomer radical to the outer surface is the ratedetermining step in the absorption process.
In the case where W' 1
::::::

(24)
Figure 3 gives 1/F as a function of r for some probable values of the
parameters involved and with W' obtained by numerical integration of Eq.
(15). The more important features of Eq. (22) are evident from the figure.
1. When particles are large (r > 1000 nm) and a h.sa high value (oil-

soluble oligomericradicals)then F::::::1 and the absorption folIowsEq. (24)


(irreversible absorption) with W' equal to unity.
2. When particles are smalIer (r < 1000 nm), the absorption rate is
lower than that given by Eq. (24), and may be very low if the particles are
small and/or the radicals are relevatively water soluble. Radicals with a very
high water solubility (a ~ 1) will not be absorbed.
3. The electrostatic repulsion may playa role when F < 1.
The electrostatic effect may be diminished by the so-calIed tunneling

effect(Fitch and Shih, 1975)in which the uncharged end pulls the oligomer
through the surface or reacts while the charged end still is in the aqueous
phase. This may be important especialIy for higher oligomers and may be
calculated if the van der Waals attraction is included in the activation
energy termo
Equation (22) may in some limiting cases be simplified.
Condition A:
X ~ 1

i.e.

X coth X

r(k/Dp)1/2~ 1

(25)

1 '" X2/3

(26)

2.

61

Particle Formation Mechanisms

10

10
r{nm}
Fig. 3.

Inverse absorption

efficiency factor 17F as a function

of particle radius r for

distribution constants a = 1 and 104. Elfective surface potential i/to = O (1), 50 mV (11),and
100mV
(III).
Other
parameters:
Dw = 5 X 10-8 dm2jsec,
Dp = 1 X 10-8 dm2jsec,
kp = 300 dm3jmol sec, Mp = 5 moljdm3, T= 298 K.

which

gives
(27)

Case A usually implies that the term W' in Eg. (22) may be neglected. In this
case Eg. (27) describes the curves with a slope of - 2 iti Fig. (3). The rate of
absorption is then
(28)
In this case the concentration gradients both on the agueous and the
particle side may be neglected. The rate of absorption is proportional to the
~

rate of reaction of the radical inside the particle and thus to the particle
volume.

62

F. K. Harsen and J. Ugelstad

Condition B:

x ~ 1 i.e. r(k/Dp)1/2~ 1
X cothX - 1 "" X

(29)
(30)

which gives
(31)
If condition B is operating it is more like1ythat W' is the dominant term
in the equation for "l/F, leading to the conclusion that the irreversible
absorption Eq. (23) will be operating. With low values of a and Dp and/or
high values of Z one may also find that the term W' may be neglected in this
case. Equation (31) then describes curves with a slope of -1, as found in
part of the most righthand curve in Fig. 3. The absorption rate will if the
term W' is neglected be
lA= 41tr2a(kDp)1/2/e'= apa(kDp)1/2/e'

(32)

where ap is the particle surface area.


AIso in this case the concentration gradient on the aqueous side may be
neglected. The rate-determining step is diffusion with reaction inside the
particle, and the absorption rate is proportional to the particle surface. This
rate will also be low compared to the case for irreversible absorption,
although not lowered to the same degree as when Eq. (28) is operating.
Condition B, including also the negligible influence of W', gives a situation
wherein Gardon's absorption equation is applicable.
Both Eqs. (28) and (32) (with F ~ 1) describe cases where the electrostatic
repulsion should be determined by the Boltzmann factor e'. When the
particle size and/or the value of a are large, the value of F is determined
chiefly by the value of W'. The rate of absorption is then given by Eq. (23).
The electrostatic repulsion will be relatively small. Usually the value of W'
willlie between 1 and 10.
Contrary to the above treatment it may be argued that it would be more
correct to express the equilibrium distribution between the inner and
outside particle surface by
(33)
where ao is a constant for the case where z = O.This means that the value of
a applied above is given by
(34)
The term eZ in the expression for l/F is said to take into account the fact
that the charged radical at the outside surface of the particle is brought up
to a higher energy level, expressed by the Boltzmann factor e'. However,

2., Particle Formation

63

Mechanisms

this fact should also come into play in that it leads to an increase in the
value of a with the same factor.
Application of a value of a given by Eq. (34) transforms Eq. (22) to the
form

!
F

Dw

aoDp X coth X - 1

+ w'

(22a)

Similarly,for Eqs. (28)and (32),respectively


(28a)

A = vpaokCw

A = apao(kDp)

1/2

(32a)

Equation (23) for irreversible absorption will remain the same, the only
necessary condition for Eq. (23) should be to replace eZlawith 1lao.

IV. Micellar Nucleation: Newer Models


It has been suggested that the charged radicals are not at all able to enter
the micelles and that the micelles only act as a reservoir for emulsifier. Roe
(1968) showed that the Smith-Ewart expression for N may be derived from
a homogeneous nucleation mechanism, substituting the critical micellar
concentration with a critical stabilization concentration somewhat lower
than the CMC. This result is not surprising, since he applied exactly the
same assumptions for the nucleation rate,'radical absorption, and constant
volume growth as did Smith-Ewart. The experimental arguments of Roe
were among others based upon the fact that a fairly constant number of
micelles in a mixture of nonionic and anionic emulsifier system gives widely
varying particle numbers. However, on the basis of the absorption equations in the preceding section it may well be that micelles give differing
particle numbers if the radical absorption rate varies significantly because of
variations in micellar size, monomer content, internal diffusion constant and
oligomer solubility. Piirma and Wang (1976) have shown that particle
numbers vary with micellar size in a mixed-emulsifier system. The results of
Dunn and AI-Shahib (1980) show that styrene polyrnerization rates, and
hence particle numbers, are constant when the molar concentration of
micelles is constant, irrespective of CMC, in the case of potassium alkyl
soaps (Cs-Cd and sodium alkyl sulfates (CS-C1S)'
One main objection to the micellar theory has been that even if the
particle number determined by polymerization of sparingly water-soluble
monomers like styrene or octyl acrylate changes steeply at or directly below
the CMC, this is not the case for more water-soluble monomers (Stterlin

~I

64

F. K. Hansen

9 seed

J. Ugelstad

Nsrs.1(

ea

and

O
11..9

21..21.

M
'e

1016

'Z
/'
./
;'

.,/

C=O

CMC

1
101'
001

0.1

1.0

e (free 50S)
Fig. 4.

Emulsion polymerization of styrene.Number

the freeconcentration of SDS


60C,

initiator

K2S20S

10.0

(g dm-3)
of new partic1esN as a function of

with and without seed. Seed partic1e radius

(0.6 gjdm3).

(Reproduced

by permission

80 nm, temperature

of J. Polym. Sci.)

et al., 1976, 1980; Ugelstad et al., 1969). The often undramatic change in N
at the CMC may well be explained by homogeneous nucleation. The
question still remains as to whether the micelles present above the CMC
play any important role in the nucleation process. One method for investigating the competition between homogeneous and micellar nucleation
uses seed experiments. Results from such runs are shown in Figs. 4 and ~
for styrene and methyl metacrylate, respectively. The log(new particle
number) is plotted as a function of log(free emulsifier concentration) both in .
seeded and unseeded runs. Free emulsifier denotes the concentration of
emulsifier not adsorbed on seed particles.
In the case of styrene the curve for the unseeded runs shows a sharp
increase at CMC. Above the CMC, the increase in particle number with
increasing emulsifier concentration becomes smaller but is still significant.

2.

Particle Formation Mechanisms

65

MMA

u
~
.u
0.1

1.0

10

Free SDS(gdm-3
Fig. 5. Emulsion polymerizatjon of methyl metacrylate.Number of new particles, N as
a fimction of the freeconcentration of SDS witb and witbout seed"

With seed present the particle number is drastically decreased at emulsifier


concentrations below the CMC. Above the CMC the curves without and
with seed almost coincide. The results indicate that above the CMC the far
more numerous micelles are the loei for particle nuclea"tion both with and
without seed.
For the case of methyl methacrylate the result without seed is strikingly
different from that with styrene in that there is no noticeable change in the
slope of the curve as one passes the CMC. This fact has previously been
used as an argument for asserting that with methyl methacrylate the
micelles do not play any role.in particle formation. It would appear from
Fig. 5 that the application of seed leads to a drastic decrease in the number

66

F. K. Hansen and J. Ugelstad

of particles below the CMC when compared to the case without seed.
Slightly above the CMC, however, the curves with and without seed
coincide completely. This would seem to indicate that with methyl methacrylate also the particle formation takes place in micelles when operating
above the CMC. It seems as if the main difference between styrene and
methyl methacrylate is that the chance for self-nucleation is better for the
latter. In conclusion it may be said that the seed experiments pro vide strong
arguments for the importance of micellar nucleation even for more watersoluble monomers. .
Further derivations based on the micellar theory have been advanced to
create a more detailed picture than that of Smith-Ewart. With these
theories the equations become so complex that a simple analytical solution,
like that of Smith-Ewart, cannot be given. Instead, numerical methods
must be applied to the integrations. Apart froIl1the calculations of Parts et
al. (1965) and Gardon (1968) already mentioned, such work has been
performed by Harada et al. (1972) (styrene), Nomura et al. (1976) (viny)
acetate), and Hansen and Ugelstad (1979b, styrene and 1979d, different
monomers). For most of these calculations nonsteady-state equations are
used to calculate the rate of particle nucleation and rate of volume growth
given by the average number of radicals per particle (). Generally, the rate
of micellar nucleation may be expressed by
(35)

wherePAis the total rate of radical generationin the aqueous phase and Pm
is the probability of absorption of radical s (oligomers) in the micelles. Often
(as in the Smith-Ewart theory) PA is set equal t 'Pi, the rate of initiator
splitting. Generally, radicals desorbed from the particles should also be
included in PA (Nomura et al., 1976; Ugelstad and Hansen, 1976). The
expression for PAwill then be
PA

= Pi +

kdN

(36)

where kd is the rate constant for desorption. .


In Eq. (36) no aqueous-phase termination is taken into account. Because
the particles are small, they will hardly be able to contain more than 2 (or
even 1) radicals, and N may be expressed by
(37) .
where Ni and N2 are the numbers of particles containing 1 and 2 radicals,
respectively. Neglecting the N2 particles the rate of formation of Ni
particles is
(38)

2.

Particle Formation Mechanisms

67

where Po and Pt are the probabilities for radical absorption in particles with
O and 1 radical, respectively. The rate of particle growth is given by

(39)
where kp is the propagation constant, N A Avogadro's number, <Pmpthe volume
. fraction of monomer in particles, and dm and dp the densities of monomer
and polymer, respectively. The last term includes the contribution to
volume growth by the micellar volume and is usually small. The volume,
radius, and surface area of each particle may be calculated from Vp and N
(this procedure involves an error in the calculated surface area because the
size distribution of the particles is not included, but it has been shown that
this error is small, usually about 5%).
Ir the above equations are integrated numerically with kd = O, constant
<Pmp,Pm = 1, and Po and Pt = O, one arrives at the Smith-Ewart upper

limit case. By setting Pm = 1 - Ap/asS, Po = Apo/asS, and Pt = Ap/asS


one arrives at the Smith-Ewart lower limit case. The important question is
therefore the magnitude of the parameters involvedin the above equations.
Expressions for kd may vary depending upon whether the monomer
radicals are transported back and forth between particles without addition
of monomer in the aqueous phase or if they add monomer in the aqueous
phase before reentering the particles (i.e., reversible or irreversible absorption, Ugelstad and M0rk, 1970). The two possibilities may be combined in
the equation (Hansen and Ugelstad, 1979d)
(40)

where kf is the transfer constant to monomer, k~ the propagation constant


for a monomer radical (to distinguish it from a polymer radical), Mp the
molar concentration of monomer in particles,and q = (reversibleabsorption) or 1 (irreversible absorption). The parameter kdm, the diffusional
desorption constant, is
(41)
where
(42)
where a is the distribution constant for the monomer radical between the
particle and water.
Equations (40)(with q = ), (41), and (42)were also derived by Harada et
al. (1971)and Nomura et al. (1971)by a somewhat differentapproach. The
probability for radical absorption may be calculated from the ratio of

68

F, K, Hansen and J. Ugeistad

absorption rate nto the particle in question to the sum of all absorption
rates. For instance, the probability for absorption into micelles is given by
(43)
where Nm is. the number of micelles presento The appropriate rates of
absorption to be inserted into Eq. (43) is obtained from Eq. (21) with the
value of F given by Eq. (22).
Because oligomers of different degrees of polymerization may be absorbed, the expression for P may become much more complex than Eq. (43).
AIso, because the efficiency constant F for initiator radicals may differ from
monomer radicals, one should in fact operate with different probabilities for
those two radicals, which leads to further complications. These problems
will be dealt with in the next section. Satisfactory results have been
obtained with the relatively simple model outlined above. Nomura et al.
(1978) expressed Pmby
(44)
where
(45)
Here Mm is the agregation number for the micelles. It is seen that Eq. (44)
is identical to Eq. (43) if the efficiency factor in particles containing O and 1
radicals is the same. By numerical integration, using constants for vinyl
acetate, Nomura et al. found
(46)
where So and lo are the initial emulsifier and initiator concentrations,
respectively. Equation (46) explained well their experimental results with
vinyl acetate when B was held constant (1.2 x 107). A very important result
of these calculations is that the exponents in Eq. (46) are higher and lower,
respectively, than the Smith-Ewart values 0.6 and 0.4, respectively. This has
been shown

to be a common

feature

when desorptionjreabsorption

be-

comes important (Nomura, private communication; Hansen and Ugelstad,


1979b,d). The present authors used a steady-state approximation for the
average radical number , as well as a constant value for fJmp'The value of "
Pm was expressed

by

(47)
Here b is an efficiency factor for radical capture in a micelle relative to a
particle of the same size and x is the order of capture rate with respect to
radius.

2.

69

Particle Formation Mechanisms

From the discussion of absorption rates (Section III) it follows that


1 :::;;x :::;;3, and furthermore that m!p = (j(rm/rpt. The steady-state radical
number was expressed by (Harada et al., 1971; Ugelstad and Hansen, 1974)
(48)
.Some results of calculations of particle numbers for different monomers are
given in Figs. 6 and 7. In these calculations, q in Eq. (40) and the values of (j
and x are all set equal to unity. Figures 6 and 7 give particle number as a
function of concentration of micellar emulsifier at a given initiator concentration and as a function of initiator concentration at a given emulsifier
concentration, respectively. It will be observed that only with styrene, when
the value of the desorption constant is low due to a low value of kf and a low
water solubility of the monomer radical (Le., a high value of a in Eq. (42)),
- do the results approximately conform to the Smith-Ewart equation. The
results with the other monomers deviate considerably from this scheme, the

0.1

1
10
50/(9 dm-3)

100

Fig. 6. The number of partic1es N as a function of the initial emulsifier concentrations,


So, calculated by numerical integration of Eqs. (35) and (39) with Pmfrom Eq. (47) and from
Eq. (48). Parameters:
p = 1 x 1016 radicals/dm3 sec, b = 1, x = 1, q = 1. Monomers: MMA,
methyl metacrylate; BMA, n-butyl metacrylate; S, styrene, VAc, vinyl acetate; ve, vinyl
chloride. (Reproduced by permission of Makromol. Chem.)

70

F. K. Hansen and J. Ugelstad


102,

ve

1019

.... ....

...

-",,'"
--~.,,-'"
.... ..........
1_"''''''

'upper

,....-

...
I
~Iower

5-E

1017
10'5

1017

p!(dm-3

Fig. 7.

Sol)

The number of particles N as a function of the rate of initiator decomposition p;

calculated by numerical integration. So


by permission of Makromo/. Chem.)

I gJdm3, other parameters

as in Fig. 6. (Reproduced

more so the higher the value of the e[ective desorption constant. It turns
out that the particlenumber may be expressedby
(49)
where 0.6 < Z < 1.0.
Equation (49) was first obtained by Nomura et al. from a nonsteadystate treatment. The value of Z increases as the desorption and reabsorption of radicals increases. This also leads to an increase in the
particle number because desorbed radicals may also produce new particles.
On the other hand the polymerization rate per particle diminishes as the
high degree of desorption and reabsorption of radicals leads to a value of.
~1.
The above calculations explain well the qualitative results with monomers like vinyl chloride, vinyl acetate, methyl methacrylate, and styrene.
Quantitatively, the correlation between theory and experiment will depend
on the magnitude of J and x, which may also change during the formation
periodo Generally, if Z > 0.6, the particle number, as well as the value of Z,

2. Particle Formation Mechanisms

71

will decrease with increasing values of x and decreasing values of b. The


effect of x will be relatively low if Z ~ 1, because then ~ 1, which means
that the particIes will grow more slowly. It may be estimated that the value
of rpjrm is approximately 10 and 3 for styrene and vinyl acetatejvinyl
chloride, respectively. The experimental results indicate that with a value of
x = 3 the value of b is approximately unity for styrene. For vinyl chloride
nnd vinyl acetate the value of b is 1.0 x 10- 3 and 1.6 x 10-4 for x = 3 and
with q equal to in Eq. (40), which seems to be the more probable situation
for vinyl chloride and vinyl acetate. Ir q in Eq. (40) is set to unity
(irreversible absorption) the value of b increases, most significantly for vinyl
acetate. In any case, it appears that the apparent value of b is considerably
lower for vinyl chloride and vinyl acetate than for styrene and the acrylates.
This may be taken to indicate that the values of b for the radicals formed
by chain transfer are considerably lower than for the polymer radicals,
which in tum may be explained if the radicals formed by chain-transfer
reactions are less reactive than the polymer radical s (Hansen and Ugelstad,
1979d). For styrene, a nonsteady-state calculation has been carried out
without the simplifications used above (Hansen and Ugelstad, 1979a). Thus,
instead of Eq. (35), the following equation was used for the rate of particIe
formation
(50)
where PImand PMmare absorption probabilities for radicals stemming from
initiator and monomer radical s, respectively. AIso, a distinction is made
between Po and P1. In the calculation of the absorption probabilities, the
complete expressions for nobtained from Eqs. (21) and (22) were used.
In addition the monomer volume fraction in particIes (<Pmp)
was calculated as a variable, applying the Morton equation (Morton et al., 1954). The
termination constant in particIes was calculated as a function of the
monomer concentration from the equation of Friis and Hamielec (1976) and
applied to Eq. (19) to calculate the total rate of radical reaction inside the
particIes. An interesting result of these calculations is shown in Fig. 8 where
N, , and <Pmp
are plotted as a function of time for a total concentration of
7 gjdm3 SDS. It appears that N reaches its final value in 2.6 rnin (conversion 8.6 gjdm3) during which decreases from 1 to slightly less than 0.5
(due to a small amount of radical desorption). An important result is that
the total rate, as expressed by Eq. (39), will not go through any maximum
during the formation period, as predicted by Smith-Ewart theory. The
reason for this is that the rate of radical absorption will be faster in particIes
containing one radical than in micelles and particIes without a radical
because the termination rate in small partic1es containing two radicals is
very high. In larger particIes, the importance of the termination term in Eq.

72

F. K. Hansen and J. Ugelstad

t51

-I

r7.5

tO

5.0;I

g,

"
IC

0.5

)(

2.5

--

t (min)
Fig. 8.
The number of particles N mean radical concentration, , and monomer volume
fraction in particles, rPmp,as a function of time from numerical. integration of Eqs. (38-39) with
P m from

Eq. (43) and from

Eq. (21-22).

Monomer,

styrene;

emulsifier

SDS with

So

7 g/dm3,

initiator K2S20S (0.6 g/dm3), temperalure 60C, Dw = 5 X lO-s dm2/sec, Dp = I X lO-s


dm2/sec, 1/10~ 100 mV, kp = 300 dm3/mol sec, a = lOs, k~p= 7 X 107 dm3/mol seco(Reproduced by permission of J. Polym. Sci.)

(19) becomes less. However, radical absorption may still be practically


irreversible because of more space for diffusion inside the particles. From
Fig. 8 it appears that the usual assumption of a constant 4Jmpseems to be
relatively good except in the early stages of the formation periodo The
ca1culations showed good agreement between theoretical and experimental
particle numbers and reaction rates both in seeded and unseeded runs and
turned out to be little sensitive to the choice of distribution constant (a) for
the radical s between particles and water (103-105). The results were somewhat more sensitive to the surface potential of the micelles and particles,
which were set to 100 mV. It therefore seems that both electrostatic
repulsion and reversible diffusion are important factors for the absorption
efficiency of micelles and very small particles.
In all theories hitherto cited, it has been assumed that only one sort of
(oligomeric) radical stemming from the initiator is absorbed, even if this
happens reversibily. This means that oligomeric radicals with a lower
degree of polymerization are not absorbed at all, or only to a very low
degree. For some monomers with low water solubility (as styrene) this may
be justified, for other monomers the assumption is more doubtful. A more
detailed discussion of the mechanistic aspects of oligomer growth and
absorption is given below, including a treatment of the mechanisms for
homogeneous particle formation.

73

2. Particle Formation Mechanisms


V. Homogeneous Nucleation

Homogeneous nuc1eation is a result of processes shown as D-G in Table


1, in which the initiator radical propagates in the aqueous phase to form an
oligomer which will then be the nuc1eus for a new partic1e. The exact
mechanism for this nuc1eation has been discussed and several propositions
have been set forth. The best-known theory on this subject, and one of the
few quantitative treatments, has been proposed by Fitch and co-workers
(Fitch and Tsai, 1971; Fitch, 1973; Fitch and Shih, 1975). They base their
model on process F, involving the self-precipiation of the oligomer radical
when a critical degree of polymerization is reached. Nuc1eation will be
hindered if the oligomers are absorbed in earlier formed partic1es (process
C) before they reach the critical chain length. The rate of partic1e generation
is given by
(51)
dN/dt = p - PA
Aqueous-phase termination was not considered as a separate rate process
but was inc1uded as anefficiency factor for p. The rate of radical absorption, PAwas expressed by
PA= npLNr~
(52)
where L is the average diffusion distance of an oligomer before selfnuc1eation. This parameter was expressed by Einstein's relationship
(53)
where t is the time between initiation and precipitation and was expressed by
t = jcr/kpMw

(54)

Fitch and Tsai applied their theory to describe homogeneous partic1e


formation for the case of methyl methacrylate. The critical degree of

polymerizationUcr) was found by GPC analysisof the aqueous phase to be


~65. The time dependence of rp was calculated from the volume growth
which Fitch et al. (1969) found could be described by homogeneous kinetics
during the formation period, Le., polymer formation as well as termination
takes place mainly in the aqueous phase. The expression for the partic1e

volumeis then

(55)
where v is the molar volume of the repeating unit in the polymer, ktw the
aqueous-phase termination constant (in molecular units) and cjJppthe volume fraction of polymer in partic1es. The rate of partic1e formation is then
from Eq. (51)
dN/dt

= p[1 -

(nN)1/3{[3vkpMw/(4ktwcjJppNA)]

x In[cosh(pktw)1/2t]}2/3L]

(56)

..

74

F. K. Hansen and J. Ugelstad

The partic1e number was obtained by numerical integration of Eq. (56).


It was found that, in order to achieve agreement with experimental data,
one had to assume a very low initiator efficiency. This does not seem
unreasonable, taking into account the high degree of aqueous-phase termination. The more formal argument against the Fitch theory maintains, as
mentioned, that it assumes that the absorption rate is proportional to the
partic1e surface area [Eq. (52)]. This assumption is in sharp contrast to Eqs.
(23) and (24) and has been criticized by several authors. Equation (52) will
lead to a lower absorption rate and thereby to a higher partic1e number
than Eq. (24). The fact that Eq. (56) seems to give reasonable results for
methyl metacrylate may possibly result from a balance between the low
rate of collision implied in Eq. (52) and the assumption of an irreversible
absorption. As stated above, it is more likely that the rate of collision should
be given by the diffusion theory which will result in a much higher rate of
collision, and that this in small partic1es may be counteracted by a reversible
absorption.
One may look upon the homogeneous process as a chain reaction in
which every chain has its specific probability for propagation, termination,
and absorption and then choose the most probable mechanism by which
new partic1es are formed. Such a treatment has been proposed by several
authors (Barrett, 1975; Peppard, 1974; Hansen and Ugelstad, 1975, 1978;
Arai et al., 1979). These authors all use the basic assumption of Fitch and
co-workers, namely that the oligomers have to reach a certain critical
degree of polymerization at which they self-nucleate (process F in Table 1).
Arai et al. also inc1ude a second mechanism. If termination by combintion
in the aqueous phase leads to an oligomer which by this reaction reaches a
chain length that exceeds the critical chain length this oligomer may
precipitate to form a partic1e. This possibility has been neglected by other
authors because it has been assumed that such a terminated molecule with
two ionic end groups would be soluble as long as each of the terminating
oligomers were soluble. It is possible that this argument is more valid for
sparingly water-soluble monomers such as styrene, where the critical chain
length is low (4-5), than for more soluble monomers with higher critical
sizes and thereby a decreased influence of end group on oligomer solubility.
The possibility for nuc1eation by mixed-micelle formation (process G)
has also been proposed (van der Hoff, 1960), and quite recently experimental work has investigated this mechanism (Munro et al., 1979; Chen and
Piirma, 1980). Evidence has been found indicating association of the surface
active oligomers in systems both with and without emulsifiers (below the
CMC). No quantitative treatment was given by these authors. It may be
questioned whether there is any real difference between the self-nuc1eation
and the mixed-micelle models. When emulsifier is absent, there is c1early a

2.

Particle Formation Mechanisms

75

strong coagulation among the small "primary"


particles, because their
surface charge is too small to give stability. Coagulation will lead to higher
surface charge density and thereby to a greater stability (so-called limited
coagulation). This phenomenon was also observed by Fitch and Tsai (1971).
Emsulsifiers will decrease this coagulation tendency up to the point where all
new particles formed are stable. It may be argued that this coagulation and
the association into mixed micelles are in reality the same process, determined by the same physical parameters (diffusion rates, repulsion, and
attraction energies). An argument against apure mixed-micelle model stems
from the fact that, with VC and VAc, desorbed monomer radicals produce
new particles. These radicals are uncharged and therefore not surface active.
Most often the problem encountered in describing the particle formation
quantitatively
involves formulating a mathematical
model that is close
enough to the real process but not so complex as to make a solution
impossible. Common to the chain-reaction models is the fact that a simple
solution for the particle number is difficult to obtain analytically without
considerable simplification. Below are given the rate equations for the chain
model based on a previous treatment (Hansen and Ugelstad, 1978). In
addition the present model includes the desorption of monomer radical s
and the different

absorption rates, as outlined in Table 1.

The differential rate expression


aqueous phase are as follows:

for the different types of radicals in the

lnitiator radicals

dR/dt

= p -

kpRMw

Monomer radicals (desorbed)

ktwRRw

Lp kaipNp

(57)

(58)
l-mers from initiator radicals
(59)
l-mers from monomer radicals
(60)

76

F. K. Hansen and J. Ugelstad

j-mers from initiator radicals

dRr
d/ = kpRIO-1)Mw- RIj kpMw+ ktwRw + ~kaIjpNp

(61)

j-mersfrom monomerradicals
(62)
Here R and RM are the concentrations of radicals produced from the
initiator and by chain transfer to monomer, respectively, RIj and RMjare the
concentration of j-mers originating from these radical s, and Rw is the total
concentration of alI radicals in the aqueous phase
jIcr

Rw = R

L:

jMcr
RIj

+ RM +

j=l

L:

(63)

RMj

j=l

The sum of all absorption rate terms p is


L: kaIjpNp
p

kaMjpNp

== kaIjmNm

== kaMjmNm

+ kaIjONO + kaIjlNl
+ kaMjoNO +

kaMj1N1

+ kaIjdNd
+ kaMjdNd

(64)

(65)

Here index m denotes micelIes,index Oand 1 particles with Oand 1 radical,


respectively (i.e., "dead" and "living" particles), and d monomer droplets.
The absorption rate "constants" ka, are generalIy expressed by Eq. (21)
(omitting the concentration). For instance
(66)
(67)
for absorption into micelIes. Similar expressions are obtained for particles
and monomer droplets, applying the proper expressions for radius (r) and
efficiency factor (F). The rate of particle formation is given by
(68)
which means that oligomers stemming from initiator and monomer radicals
precipitate as particles when they propagate beyond their respective critical
degrees of polymerization jIcr and jMcp which may or may not be equal.

2.

77

Partiele Formation Meehanisms

Probably, jMcr will be less than jIcr because the monomer radical is uncharged. The rate of formation of living (N1) particles is

and the rate of formation of dead particles (No)is obviously


dNo/dr

= dN /dt -

dN/dt

(70)

The total rate of particle growth is expressed by Eq. (39), only the micellar
volume Vrnshould be exchanged by the volume of the precipitated oligomers. We do not differentiate between the volumes of dead and living
particles, since the particles rapidly change from active to inactive and vice
versa.
Equations (57-70) may be integrated numerically, expressing F by Eq.
(22). A simplification may be obtained by means of the steady-state
approximation, in which all rates of radical change in the aqueous phase are
set equal to O,expressing Rj by means of Rj-1
RIj/RIO-1)
RMj/RMO-1)

= kpMw/(kpMw + ktwRw + LpkaIjpNp)

= fJIj

= kpMw!(kpMw + ktwRw + Lp
kaMjpNp)= fJMj
,

(71)
(72)

Thus, fJ is the probability for propagation of a j-mer. By multiplication of


all f3Ijand fJMj,independently, and inserting into Eq. (68),one obtains
dN
jlcr
jMcr
(73)
= Pi~i n f3Ij+ kdN1f3Mn fJMj
dt
j=1
j=1
By writing out the expressions for fJ and dividing each term in the
denominator by kpMw, one obtains

(74)
where
(75)

78

F. K. Hansen and J. Ugelstad

and
PM = 1 + ktwMRw + Lp kaMpNp
kpMMw
kpMMw

(76)

The expression for Rw is obtained by expressing Eqs. (71) and (72) as


j
p
Rlj

(77)

= k p M w p T1
PII
1=1

and
j
kdN1
RMj

(78)

= k p M w PM T1
PMI
1=1

FinalIy we have

R.~ k:~A~+%Q]p,.)]
(79)

+ ::Z~PM[[+%(QPM.)]

In the first paper on this subject (Hansen and Ugelstad, 1978) desorption
of monomer radical s was neglected. Furthermore, in order to simplify Eq.
(74), Pi was set equal to 1 (Le., termination and absorption of initiator
radicals was neglected). AlI ka1jpvalues were set equal to some average
absorption constant Kc(when micelIes are absent). As an approximation, Rw
could then be expressed by
Rw

= {[(KcN)2 + 4pktw]I/2

KcN}/2ktw

(80)

and dN /dt by
dN/dt

= p(1 + ktwRw/kpMw + KcN/kpMw)-hcr

(81)

Ir I<cand Rw are constant during the nucleation period, Eq. (81) may be
integrated numericalIy. Assuming that as an approximation radical absorption may be neglected in the calculation of Rw we ha ve
Rw

= (p/ktw)I/2

(82)

and
N(t)

..

= {[k1Pjcrt +

(k2 + l)jcr]IUcr - k2

1}/k1

(83)

..

2.

79

Particle Formation Mechanisms

where
(84)
and
kz = (ktwpY/Z/kpMw

(85)

Because r will increase with time, ke will also increase, and dN/dt will fall off

more with time than is expressedby Eq. (81). By using a plausible average
value for r, Eq. (83) may give an idea of the number of particles that may be
expected to be formed homogeneously. Assuming jer = 5, as found by
Goodall et al. (1975) for styrene, and F = 1 (irreversible absorption) it turns
out that N will be of the same order of magnitude as found experimentally
in emulsifier-free systems. In systems with emulsifiers, below the CMC, the
experimental particle numbers are much higher, indicating that the value of
ke (and F) is in reality much lower. The low particle number in emulsifierfree systems must therefore be attributable to limited coagulation.
By numerical integration of Eq. (73) it was shown (Hansen and Ugelstad,
1979b) that after 1 hr the ca1culated particle number was close to the
experimental particle number for styrene with SOS as emulsifier, Le.,
immediately below the CMC. The following simplifications were made: (1)
no aqueous phase termination and (2) radical s with lower degree of

polymerizationthan jer are not absorbed. The rate of nucleation is then


dN/dt
in close resemblance

= PPIh + kdNPMh

to Eq. (50) for nucleati~n

(86)

in micelles. Here

(87)
and
PMh

= [1 +

(88)

(Lp4nDwrpNpFMp)/kpMwrl

with FIpand FMpexpressed by Eq.(22).


The total rate of radical absorption may be expressed by
jIcr

pA

jMcr

L (RIj Lp k.IjpNp) + j=O


L (RMj Lp k.MjpNp)
j=O

(89)

where R1j and R..jare given by Eqs. (77) and (78). It is obviously very
difficult even to estimate PA from Eq. (89). One method of simplification
may be the second one given above. This leads to
Lp k.1pNp

PA =p.
1 kpMw

+ ktwRw + Lp k.1pNpkpMw+ ktwRw

k N

jcr

kpMw

1 kpMw

kpMw
Lp k.MpNp
ktwRw + Lp k.MpNp kpMw + ktwRw

jMCr

(90)

F. K. Hansen and J. Ugelstad

80

Ir the absorption rate of the critical oligomers is high (spontaneous), the


absorption probability for radicals stemming from the initiator is
PAI

= (1 +

ktwRw/kpMw)-her

(91)

ktwRw/kpMw)-jMer

(92)

and for radical s from chain transfer


PAM

= (1 +

The fraction 1 - PAIwill therefore represent the probability for termination


of radicals stemming from initiator in the aqueous phase before they can be
absorbed.
It is expected that the value of PAIwill be c10se to unity for monomers
like vinyl chloride and vinyl acetate where the values of kpMw are high. In
the case of styrene, kp = 240 dm3/mol sec and Mw= 4 X 10-3 mol/dm3
(50C).With a value of k = 1.0 X 10-6 sec-I, ktw = 7 X 107 dm3/mol sec,
and jlcr = 4 the values of PAIat [I] = 10-4, 10- 3, and 10- 2 mol/dm3 are,
with Rw expressed by Eq. (82), 0.72, 0.38, and 0.082, respectively. Quite
recently, Hawkett (1980) investigated the seed polymerization of styrene at
various initiator concentrations..From exact measurements of conversion as
a function of time at low conversions, so that the approach to the steady
state is taken into account, he c1aims to have determined separate values for
the rate of absorption PAland the desorption rate constant kd for the special
latex applied (N

= 4.9

x 1016 dm

- 3, diameter

93 nm). Hawkett's

values for

PAI at the above values of [I] are approximately 0.2, 0.09, and 0.03,
respectively. Hawkett points out that ktwmay have a higher value than that
given above based upon measurements in dilute polymer solution, since the
termination reactions in water involve the reaction between small molecules. It should be pointed out that the above treatment do es not involve
competition between termination and absorption as in Eq. (90). From this
equation we should expect an influence of partic1e number and micelles on
absorption efficiency.
The assumption of such a low efficiency as found by Hawkett, where the
order of PAlwith respect to [I] approaches 0.5 with increasing [I], seems to
be in contradiction to both the absolute value of partic1es formed (both
below and above the CMC) and even more so with the observed exponent
in Eqs. (5) or (7) (i.e., 0.4). AIso, it would seem to contradict the well-known
fact that at high values of [I] (i.e., high a' = 2ki[I]Vp/Nk~) the rate of
polymerization is proportional to [I] 1/2 even in Interval II (Smith-Ewart
Case 3).
The effect of termination in aqueous-phase on partic1e formation will
depend on whether the terminated and usually doubly charged molecules
stay in solution, are absorbed on existing partic1es, or form the nuc1eus for
a new partic1e. It seems unlikely that doubly charged styrene oligomers with

2.

81

Particle Formation Mechanisms

low degrees of polymerization (1-8) would form new partic1es, whereas it


seems more probable that more water-soluble monomers, forming higher
oligomers, may react according to this mechanism. A model for homogeneous partic1e formation applying this mechanism has recently been
pro posed by Arai et al. (1979). These researchers also make the distinction
between active and inactive partic1es, and their expressions for the rate of
formation are (using our terminology)

-dN1 = kpRicrMw+ 4n(foNo dt


-dN. o = (1 - A)!ktw jcr
d
t

L DwjRj

jcr

f1N1)

(93)

j= 1

jcr

L R,

Rjcr+l-j

j=1

1=1
jcr

+ 4n(f1N1 - foNo)

L DwjRj

(94)

j=1

where Dwj is the diffusion constant for j-mers in water and A is the ratio of
termination by disproportionation to the overall termination reaction. The
first term in Eq. (93) accounts for the partic1e formation by precipitation of

oligomers that have a total degree of polymerization higher than jcr. The
rate expressions for the soluble oligomers Rj, otherwise c10sely follow Eqs.
(59-62). Arai et al. assume that Dwjfollows Wilke's equation
DWJ. = Dw/':1.0.6
(95)
They do not take into account desorption of monomer radicals or the
possibility for decreased rates of desorptioll'.
The equations for dNldt and dR/dt, as well as for dVldt [Eq. (39)J are
solved by numerical integration for the polymerization system MMAK2S20S-water, with rate constants obtained from the literature. The
initiator efficiency was set-equal to unity. Partic1e numbers between 1014
and 1015were obtained for initiator concentrations of 10-3-10-2 mol/dm3.
The calculations showed that N should be almost independent of the
chosen value of jcr for values between 5 and 70 (in strong contrast to our
calculations). The reason for this is probably that aqueous-phase termination with subsequent precipitation is the dominant partic1e-formation
mechanism in Arai's model, even more so with increasing initiator concentration. The theoretical partic1e-formation time was on the order of 2
sec, a very low value compared to the experimental results of Fitch and
Tsai. Arai et al. found that their calculated partic1e numbers were approximately in accordance with the experimental results of Yamazaki et al.
(1968) for emulsifier-free polymerizations. Arai's model does not inc1ude any
coagulation mechanisms. It will therefore have the same shortcomings as
most other models, namely that the strongly increased partic1e number in

82

F. K. Hansen and J. Ugelstad

systems with emulsifier (below CMC) cannot be explained. A modification


of the radical absorption rates could possibly lead to a higher particle
number for this model too. If the parameters of Arai et al. are inserted into
the very simplified equation Eq. (83), almost the same particle numbers are
obtained. In conclusion it may be said that there is still too little experimental evidence to determine the exact mechanism of homogeneous particle
formation, except that the production of oligomers in the aqueous phase
seems to be the essential initial mechanism.
A problem of special importance to homogeneous particle nucleation
involves coagulation processes. It has usually been assumed that all new
particles will be stable if the concentration of emulsifier is just below the
CMC. With decreasing amounts of emulsifier, an increasing amount of
coagulation sets in, up to the point where no emulsifier is added and the
particles are only stabilized by surface groups ipherent in or produced by
the reaction. Roe (1968) proposed the use of some sort of absorption
isotherm of emulsifier on particles to account for the influence of the
emulsifiers. This would then have to be connected with corresponding
stability criteria (W ratios). Also, kinetic aspects of compeition between
absorption and coagulation might be of importance. No quantitative model
explaining the influence of emulsifier below the CMC (except for Roe's
modified Smith-Ewart calculation) has yet been presented. Although this
may seem less important than the mechanism for particle formation above
the CMC, as most practical applications use supermicellar consentrations, it
still represents both a theoretical and practical problem of considerable
interest.

VI.

Particle Coagulation during the Formation Period

Because of the transient character of the reaction, analysis of particle


numbers during the formation period is not as easy as the case where the
number has reached a constant value. Sampling has usually been performed
in one of two ways, either by fast sampling from the reaction mixture with
some sort of short-stop procedure or by direct observation of a monomer
drop-free system by some light-scattering technique. The former method
was used by Fitch and Tsai (1971) in their homogeneous MMA-water
system. The short-stop was achieved by spraying onto an electron microscope grid directly from the reactant mixture. When no emulsifier was used,
a pronounced maximum in particle number was observed, and the final
value was much less than the value at the maximum. A maximum was also
observed in systems with SDS as emulsifier, but the maximum became less
pronounced with increasing SDS concentration. Furthermore, formation

2.

83

Particle Formation Mechanisms

and coagulation took place during a longer time interval when emulsifier
was used, and the total particle number both at the maximum and at the
steady-state level increased.
Fitch and Watson (1979) also used a light-scattering technique to
investigate the early stages of the reaction in a homogeneous solution of
MMA in water. They used a photoinitiation procedure by which free
radical s could be p~oduced by means of a light flash, after which Rayleigh
ratios (Rgo) were measured as a function of time. Within the experimental
time scale of '" 2 min, Rgo increased linearly with time, indicating particle
coagulation. The rate of coagulation was clearly dependent on SDS concentrations and was most pronounced in emulsifier-free runs.
The stability of a dispersion may in the most simple way be described by
Smoluchowski's coagulation equation
(96)
-dNddt = kllNf
and in the integrated form
1/N1

= l/No +

kllt

= l/No(1 +

t/1:)

(97)

Here k11 is the second-order coagulation constant and 1:the half-life time
l/Nokll' This treatment assumes that kll is constant during the coagulation, which is only approximately correct during the early stages. When no
force exist between the particles, the coagulation was regarded as a diffusion
process with a value of kll given by
kll

= 4nDllrll .

(98)

where Dll and rll are the total diffusion constant and the collision radius of
the particles, respectively. Expressing Dll by the Stokes-Einstein relationship leads to the familiar expression for kll
(99)
k11 = 4n(2D1,)(2r1) = 16n(k:T/61ttlr1)r1 = 8kT/3r
When repulsive forces exist between the particles, the rate constant is
decreased and the ratio kll(slow)/kll(fast) is usually expressed as
kll(slow)/kll(fast)
in which W(fast) 1S usually set
ratio, expressed by

= 1.

= W/W(fast)

The parameter

reo
W = 2rlo exp(J!;ot/kT)

dh

(100)

W is

Fuchs' stability
(101)

where J!;otis the total potential energy and h is the distance between the
particles. The parameters for this equation [Eq. (101)] have been extensively discussed in the literature (Derjaguin and Landau, 1941; Verwey and
Overbeek, 1948).

84

F. K. Hansen and J. Ugelstad

For our purpose, it is often convenient to describe the stability during


particle formation through the variation in W with time and experimental
conditions. In the experiments cited above, Fitch and Watson (1979) found
the linear relationship

(102)
where S is the molar concentration of SDS and is less than 9 x 10-4 M
(0.26g/dm3).Becauseof the linear relationship of Rayleigh ratios with time,
W wasindependentof time for a specificexperiment,whichindicatesthat W
is independentof particlesize. This was also found by Ottewill and Shaw
(1966) to be the case for larger particle. When the total particle surface is
small, and also at low SDS concentrations, the surface charge density will
probably be independent of particle size and increase linearly with S. This
would seem to imply that W has a linear dependence on surface charge
density in the experiments described by Fitch and Watson.
Munro et al. (1979) and Goodall et al. (1980) investigated particle
nucleation in emulsifier-freepolymerizationsof styreneby means of electron
microscopy, light scattering, and photo correlation spectroscopy, the latter
in unstirred reactions in which 11 styrene layer was kept on top of the
saturated water solution. The initiator was potassium persulfate. Particle
numbers were found to decrease with time from an initial high value (Le.,a
maximum); the maximum was more pronounced the higher the polymerization temperature. The time scale for the decrease was 10-100 min and
involved particles up to a size of about 75nm. GPC analysis of both the
aqueous phase and the particles during these early stages showed that the
predominant molecular weight of the species present in the aqueous phase,
measured as polystyrene, was about 520 and that of the polystyrene in the
particles about 1000. In addition, the particles contained some higher
molecular weight polymer (up to 106), the fraction of which increased after
the formation period was finished. The conclusions drawn were that if these
low molecular weight species are produced by termination by combination
it would appear that the water-soluble oligomers are produced from 2-3
mers and the oligomers in the particles from 4-5 mers. This is the basis
for assuming that in the case of styrene jcr = 4-5.
In the work on emulsifier-free polymerization of styrene Hansen and
Ugelstad (1979a) also concluded that coagulation during the formation
period was of great importance. By seeded polymerizations, it was found
that by including coagulation between new particles and seed particles one
could describe qualitatively the number of new particles formed. For the
homocoagulation between new particles, Eq. (96) was applied with ku
given by Eqs. (97-100).Heterocoagulation between seed (Nz) and the new
particles (NI) is described by
(103)
-dN/dt = klZNINz

2.

85

Particle Formation Mechanisms

where
k12

= 41tp12r12/W2
= 2kT/317[(1/r

= 41t(D + D2)(r+ r2)/W2


+ 1/r2)(rl

(104)

+ r2)/WI2]

The expression for W2 is similar to that given in Eq. (94), exchanging 2r


with rl + r2' In order to calculate particle coagulation during the formation
period the distribution in size of the coagulating new particles should be
included, with the coagulation rate generally expressed by Eq. (103). Such a
procedure would lead to very complex calculations. It would be necessary
to calculate W12as a function of size and surface charge density, which will
vary during the reaction. AIso, particle growth by propagation should be
included leading to higher particle sizes and lower charge densities. The
present authors (1978) have made an attempt to perform such a calculation,
without including propagation. Values for W12 were calculated from the
DL VO theory for different degrees of coagulation of primary pafticles with
one surface charge. The rate of formation of new particles was expr.essed by
Eq. (68) with Rj from Eqs. (59-65) (without monomer radical desorption).
The result of this calculation was a particle number which increased up to
t

= 2.7

sec and then decreased.

The size distribution

also went through

maximum which increased with time. It seems that this result is in


qualitative agreement with the results cited above, the calculated time for
reaching the maximum particle number is, however, far too short, and in
addition the maximum is too shallow. Both discrepancies are probably
explained by the fact that the value of the efficiency factor F was set equal
to unity at Dw = 10-8 dm2/sec, resulting iij a considerable underestimation
of the number of new particles formed. This number should probably be
increased up to two orders of magnitude which would result in both more
pronounced maxima and a larger time scale.
Hansen and Ugelstad (1978) also proposed a simplified steady-state
description of the particle formation with coagulation. In this case the
following equation was applied:
dNjdt

= pkpMw/(kpMw + kcN + kcsN2) -

where kc and kcs are the absorption

constants

kllNi

k12N1N2

(105)

in primary and seed particles,

respectively. Only oligomers at the critical degree of. polymerization are


assumed to be absorbed, and aqueous-phase termination is neglected. This
could be included by applying Eq. (90) for the absorption rateo AIso, desorption of monomer radicals could easily be included. The solution of Eq. (105)
when dNjdt

O and N2

O with kcN ~ kpMw is

N = !(p317kpMwWu/4nDwkTrF)

(106)

To arrive at more exact solutions, including the increase in r and change in


Wll and F, numerical methods have to be applied.

F. K. Hansen and J. Ugelstad

86

VII.

Nucleation in Monomer Droplets

Usually, particle formation by initiation in the monomer droplets is not


considered important. This is because of the low absorption rate of radicals
into the droplets, relative to the other particle formation rates. Only in
cases where the monomer droplets are made very small may they be an
important source for partic1e formation. In a series of papers, Ugelstad and
co-workers have made a strong case for droplet nuc1eation in systems with
especially effective preparation and stabilization methods for making and
keeping the monomer in a very fine-dispersion (Ugelstad et al., 1973; 1974;
Hansen et al., 1976; Azad et al., 1976; Hansen and Ugelstad, 1979c). The
fine monomer dispersions were produced either by spontaneous emulsification by means of a mixed emulsifier system consisting of an anionic or
cationic emulsifier in combination with a long-chain fatty alcohol or amine
or by high pressure homogenization of monomer and water, using ionic
emulsifiers and a water-insoluble compound (e.g., hexadecane) to stabilize
against degradation by Ostwald ripening. The latter experiments are very
illustrative in this context. Electron micrographs of latexes produced with a
total concentration of 2 g/dm3 of'SDS (constant) with potassium persulfate
as initiator are shown in Fig. 9. The variation in partic1e size was obtained
by variation of the homogenizing conditions.
Case A is an ordinary polymerization with no homogenization. The
degree of monomer subdivision by the homogenization increases in the
order B-E. In case A the normal emulsion polymerization with initiation in
the aqueous phase leads to formation of large number of particles. The
increasing partic1e size in going from A to B and possibly C may still be
explained by :in aqueous-phase nuc1eation. The partic1es become larger
because some emulsifier is adsorbed on the monomer droplets, leaving less
emulsifier in the aqueous phase and thereby leading to the formation of a
smaller number of partic1es. In cases E and D when practically all emulsifier
is adsorbed on the monomer droplets the initiation obviously takes place
solely in the monomer droplets. The number of partic1es in case E is even
larger than that obtained in case A.
There are two factors that influence the particle formation conditions:
the adsorption of emulsifier on the droplets, which leaves less emulsifier in
the aqueous phase to facilitate partic1e formation. there, and the rate of
radical absorption in monomer droplets. The former factor may be evaluated'
by taking into account the adsorption isotherm of emulsifier on monomer
droplets. Using a modified Langmuir adsorption isotherm in combination
with a mass balance gives an expression for the aqueous phase concentration of emulsifier C
ifO<C<CMC
C

= t[(K2 + 4CXS)1/2 -

K]

(107)

2.

87

Particle Formation Mechanisms

..

..tt-

Fig.

9.

Electron micrographs

of poly-

styrene latexes produced without (A) and with


(B-E) monomer homogenization. Total SDS,
So = 2 gfdm3, initiator K2S20S (0.6 gfdm3),
temperature 60C, monomer concentration
300 cm3fdm3 H20. Concentration
of free
SDS (gfdm3): (A) 2.0, (B) 0.95, (C) 0.27,
(D) 0.14, (E) 0.12. (Reprinted by permission of
J. Polym. Sci.)

.
88

F. K. Hansen and J. Ugelstad

where
K

[{ex

(108)

+ CMC)/CMC](Ad/as) - S + ex

ifC;?;CMC
C=S-

(109)

Ad/as

Here S is the total emulsifier concentration, exthe constant in the Langmuir


adsorption equation, Ad the total monomer droplet surface, and as the
specific surface per unit amount of emulsifier. The expression for Ad is
Ad

= 3VJrd

(110)

where J-d and rd are the total volume and surface average radius of the
monomer droplets, respectively.
The rate of radical absorption in monomer droplets is given by Eq. (89) if
index p is taken to represent monomer droplets, or simplified by an
equation similar to Eq. (90)

PAd

ka1dNd

= p.
I kpMw

+ Pd

kpMw

hcr

+ ktwRw + p ka1pNp kpMw+ ktwRw

kaMdNd

kpMw

jMcr

kpMp + ktwRw + p kaMpNp kpMw+ ktwRw

(111)

where the index p in Np as usual represents all kinds of particles, droplets


and micelles and Pd is the desorption rate from all particles containing
radicals. The terms enclosed in parenthesis may, if Rw is constant, be taken
as the initiator/monomer radical absorption efficiency and the two fraction
terms containing Nd as the efficiency of droplet initiation relative to all
other initiation mechanisms. If, for simplicity, termination in the aqueous
phase and monomer radical desorption are neglected, Pd at the start of the
reaction is given by
P d--

4nDwrdNdFd

+ 4nDwrdNdFd + 4nDwrmNmFm

kpMw

(112)

and the efficiency of homogenuous nucleation Phby


P,h--

kpMw
kpMw

+ 4nDwrdNdFd + 4nDwrmNmFm

(113)

where Nm is the number of micelles which when C > CMC is given by


Nm

= N~iC - CMC)

(114)

where N~ is the number of micelles per unit amount of emulsifier.


The absorption efficiency factors Fm and Fd are as usual given by Eq.
(22). An illustration of the effect of monomer droplet size on emuIsifier

..

--

2.

89

Particle Formation Mechanisms

adsorption and initiation efficiencies in droplets and by homogeneous


nucleation calculated by means of Eqs. (107-110) and (112-113) are shown
in Table n. The values for Fm and Fd are taken from Fig. 3 with a = 104
and IjIom= 100 mV.
In these calculations the total volume of monomer was constant
(300 cm3/dm3 water) and the specific surface area of SDS on the monomer
was set to 50

A2

per molecule.

The other parameters

are given in the table

footnote. It appears that absorption of emulsifier on the monomer droplets


surface starts to become significant at droplet radii '" 5 p.m. With
S = 2 g/dm3 H20 one passes the CMC for the emulsifier at a droplet
diameter of about 1.5 p.m, with 5 g S the corresponding value is ",0.2 p.m.
With droplets below that size one should expect that the droplets may
become the important loci for particle nucleation.
It also appears from Table n that the probability of homogeneous
nucleation decreasesmarkedly as the monomer droplet radius decreases. In
considering the number of new particles that may be formed below CMC,
the stability and the growth of the new particles must also be taken into
account. The reduction in Nh most probably is lower than would be
expected from the decrease in Ph.
Ir one operates at e > CMC the number of micelles will usually be much
higher than the number of monomer droplets and consequently the number
TABLE 11
Number of Monomer Droplets. Free Emulsifier Concentrations.
for Droplet and Homogeneous Initiation8

S
rd (pm)
0.05

0.10
0.15
0.25
0.50
1.00
2.00
5.00
10.00

Nd(dm-3)
5.73
7.16
2.12
4.58
5.73

x
x
x
x
x

1017
1016
1016
1015
1014

7.16
8.95
5.73
7.16

X 1013
X 1012
X 1011
X 1010

= 2 gJliter

e (gJliter)

Pd

0.04

-1
-1
-1
-1

0.09
0.14
0.26
0.65
1.18
1.57
1.83
1.91

-1
0.997
0.847
0.162
0.037

S
Pb

e (gJliter)

and Probabilities

= 5 gJliter
Pd

1.7 x 10-5

0.11

4.2x 10-5

0.30

8.0x 10-5

0.61

-1

1.9 x
6.8 x
2.7 x
9.0 x
1.1 x
1.0 x

1.56
3.28 .
4.14
4.57
4.83
4.91

0.997
0.781
0.379
0.116
0.019
0.005

10-4
10-4
10-3
10-3
10-2
10-2

-1

-1

Pb

1.7 x 10-5
4.2
8.0
1.9
5.3
1.0
1.2
1.3
1.3

x
x
x
x
x
x
x
x

10-5
10-5
10-5
10-4
10-3
10-3
10-3
10-3

Calculated from Eqs. (107-110) and (112-113) as a function of average droplet size rd'
Monomer
volume 300cm3/dm3
H20. Other parameters:
Dw = 5 X 10-8 dm3/sec;
Dp = 1 X 10-8 dm2/sec; kp = 300 dm3/mol sec; Mw= 4 X 10-3 mol/dm3; Mp = 5 mol/dm3;
a = 104; 1/10"; 100 mV, K = 5 X 108 m-I, CMC = 1.5 g/dm3.
Q

90

F. K. Hansen and J. Ugelstad

of particIes stemming from the micelles Nmmuch higher than Nd. It should
be pointed out, however, that this does not necessarily mean that there is no
initiation in droplets. They will, however, in most cases be difficult to detect
and even to separate from particIes initiated in the aqueous phase. The
higher total rate of consumption of monomer in the large number of
particIes initiated in the aqueous phase implies that even if monomer
droplets have become initiated they will during polymerization shrink
because of transfer of monomer to the growing particIes initiated in the
aqueous phase.
The experimental results shown in Fig. 10 illustrate the condition for
droplet initiation. In this figure the total number of particIes at the end of
polymerization is plotted as a function of the concentration of free emulsifier in the aqueous phase for three different levels of emulsifier concentrations. One moves from right to left in the figure as the degree of

../

'

0\"

~~.
8,<:,0

4,\

~o

~~

\KtA8~
....
I
E

~~~,yl

"

VP'
~L~

7. I
~

A
1015.
0.1

~~

~it.
2g
5..
10..

PPS
~
4
Ct

1.0

BP
<>
V
10.0

e (free 50S) (g dm-3)


Fig. 10.
Total number of particlesafter styrene emulsion polymerizations as a concentration of free SDS in the water phase before polymerization. Initiators K2S20S (PPS) and
dibenzoyl peroxide (BP). Other experimental conditions as in Fig. 9. Theoretical curves A-E
calculated from the absorption isotherm, Eq. (107-108). Curves A-B, ex= 4 gjdm3, curves C-E,

ex= 0.4 gjdm3.(Reprintedby permissionof J. Polym.Sci.)

2.

Particle Formation Mechanisms

91

homogenizationof the monomer-water-emulsifier with a constant amount


of total emulsifier is increased. The particle number decreases because the
particle formation in this region still takes place in the aqueous phase,
because the emulsifier concentration in the aqueous phase decreases, and
because with increasing homogenization more emulsifier becomes adsorbed
on the surface of the monomer droplets.
As the degree of homogenization is further increased the particle number
goes through a minimum and then increases steeply, with a slope of - 3,
when the monomer droplets have become the dominating loci for particle
formation. Figure 10 also includes a curve from "normal" emulsion polymerization experiments where the particle number is plotted as a function of
the amount of emulsifier added. It appears that with sufficient homogenization one may get a higher number of particles formed in monomer
droplets than will be formed by a "normal" emulsion polymerizat.ion with
the total amount of emulsifier in the aqueous phase.
Experiments with benzoyl peroxide show that with the oil-soluble initiator, initiation takes place solely in the monomer droplets under the same
conditions as in Fig. 9. No minimum in the curves is encountered as the
degree of homogenization of the monomer-water-emulsifier mixture is
reduced.
References
Alexander, A. E., and Napper, D. H. (1971). Prog. Polym. Sci.3, 145.
Arai, M., Arai, K., and Saito, S. (1979). J. Polym. Sci.. Polym. Chem. Ed. 17,3655-3665.

Azad,A. R. M., Ugelstad,J., Fitch, R. M., and Hansen,F. K. (1976).Am. Chem.SocoSymp. Ser.
24, 1-23.
.
Barrett, K. E. J., (1975). "Dispersion Polymerization in Organic Media." Wiley, New York.
Byron Bird, R., Stewart, W. E., and Lightfoot, E. N. (1960). "Transport Phenomena." Wiley,
New York.
Chen, C-Y., and Piirma, I. (1980). J. Polym. Sci. Polym. Chem. Ed. 18, 1979-1993.
Danckwerts, P. V. (1951). Trans Faraday Soco 47. 1014.
Derjaguin, B. V., and Landau, L. (1941). Acta Physicochim. USSR 14, 633.
Dunn, A. S., and Al-Shabib, W. A. (1980). In "Polymer Colloids 11" (R. M. Fitch, ed.), pp.
619-628. Plenum Press, New York.
Fitch, R. M. (1973). Dr. Polym. J. 5, 467.
Fitch, R. M., and Shih, L-B. (1975). Prog. Colloid Polym. Sci. 56, 1.
Fitch, R. M., and Tsai, C. H. (1971). In "Polymer Colloids" (R. M. Fitch, ed.), p. 73.
Fitch, R. M., and Watson, R. C. (1979).J. ColloidInterfaceSci. 68, 14-20.
Fitch, R. M., Prenosil, M. B., and Sprick, K. J. (1969). J. Polym. Sei. Part C 27,95.
Friis, N., and Hamielec, A. E. (1976). Am. Chem. Soco Symp. Ser. 24, 82-91.
Gardon, J. L. (1968a). J. Polym. Sci. Part A-l 6, 623.
Gardon, J. L. (1968b). J. Polym. Sci. Part A-l 6, 643.
Gardon, J. L. (1968c). J. Polym. Sci. Part A-l 6, 665.
Gardon, J. L. (1968d). J. Polym. Sci. Part A-l 6, 687.

92

F. K. Hansen and J. Ugelstad

Gardon, J. L. (1968e). J. Po/ym. Sci. Part A-I 6, 2853.


Gardon; J. L. (1968f). J. Polym. Sci. Part A-I 6, 2859.
Gardon, J. L. (1971). J. Po/ym. Sci. Part A-I 9, 2763.
Gatta, G., Benetta, G., Talamini, G. P., and Vianello, G. (1969). Adv. Chem. Ser. 91, 158.
Gerrens, H. (1963). Ber. Bunsenges. Phys. Chem. 67, 741.
Goodall, A. R., Wilkinson, M. C., and Heam, J. (1975). Prog. Colloid Interface Sci. 53, 327.
Hansen, F. K., and Ugelstad, J. (1975). Preprint, Nato Advanced Study Institute on Polymer
Colloids.
Hansen, F. K., and Ugelstad, J. (1978). J.Po/ym. Sci. Po/ym. Chem. Ed. 16,1953-1979.
Hansen, F. K., and Ugelstild, J. (l979a). J. Po/ym. Sci. Po/ym. Chem. Ed. 17,3033-3045.
Hansen, F. K., and Ugelstad, J. (l979b). J. Po/ym. Sci. Po/ym. Chem. Ed. 17, 3047-3067.
Hansen, F. K., and Ugelstad, J. (l979c). J. Po/ym. Sci. Po/ym. Chem. Ed. 17,3069-3082.
Hansen, F. K., and Ugelstad, J. (1979d). Makromo/. Chem. 180,2423-2434.
Hansen, F. K., Ofstad, E. B., and Ugelstad, J. (1976). In "Theory and Practice of Emulsion
Technology" (A. L. Smith, ed.), pp. 13-21. Academic Press, New York.
Harada, M., Nomura, M., Eguchi, W., and Nagata, S. (1971). J. Chem. Eng. Jpn. 4, 54.
Harada, M., Nomura, M., Kojima, H., Eguchi, W., and Nagata, S. (1972). J. App/. Po/ym. Sci.
16, 811.
Harkins, W. D. (1947). J. Am. Chem. Soco 69, 1428.
Harkins, W. D. (1950). J. Po/ym. Sci. S, 217.
Hawkett, B. S. (1980). Thesis, Dept. of Physical Chemistry, Univ. of Sydney.
Morton, M., Kaisermann, S., and Altier, M. W. (1954). J. Colloid Sci. 9, 300.
Munro, D., Goodall, A. R., Wilkinson, M. c., Randle, K., and Heam, J. (1979). J. Colloid
Interface Sci. 68, 1-13.
Nomura, M., Harada, M., Nakagawara, K., Eguchi, W., and Nagata, S. (1971). J. Chem. Eng.
Jpn.4, 160.
Nomura, M., Harada, M., Eguchi, W., and Nagata, S. (1975). Po/ym. Preprint Am. Chem. Soco
div. Po/ym. Chem. 16,217.
Nomura, M., Harada, M., Eguchi, W., and Nagata, S. (1976). Am. Chem. Soco Symp. Ser. 24,
102-121.
Ottewill, R. H., and Shaw, J. N. (1966). Discuss. Faraday Soco 42, 154.
Parts, A. G., Moore, D. E., and Watterson, J. G. (1965). Makromo/. Chem. 89, 156.
Peppard, B. D. (1974). Thesis, Iowa State Univ.
Piirma, l., and Wang, poCo(1976). Am. Chem. Soco Symp. Ser. 24, 34-61.
Roe, C. P. (1968). Ind. Eng. Chem. 60, 20.
Smith, W. V., and Ewart, R. H. (1948). J. Phys. Chem. 16,592.
Stterlin, N., Kurth, H-J., and Markert, G. (1976). Makromol. Chem. 177, 1549-1565.
Stterlin, N. (1980). In "Polymer Colloids 11" (R. M. Fitch, ed.), pp. 583-587. Plenum Press, New
York.
Ugelstad, J., and Hansen, F. K. (1976). Rubber Chem. Techno. 49, 536-609.
Ugelstad, J., and Merk, P. C. (1970). Brit. Po/ym. J. 2, 31.
Ugelstad, J., Merk, P. C., Dahl, P., and Rangnes, P. (1969). J. Po/ym. Sci. Part C 27,49-68.
Ugelstad, J., EI-Asser, M., and Vanderhoff, J. W. (1973). J. Po/ym. Sci. Po/ym. Lett. Ed. 11,505.
Ugelstad, J., Hansen, F. K., and Lange, S. (1974). Makromo/. Chem. 175, 507-521.
Ugelstad, J., Hansen, F. K., and Kaggerud, K. (1977). Faserforsk. Texti/tech. 28, 309-320.
Verwey, E. J. W., and Overbeek, J. Th. G. (1948). "Theory of the Stability of Lyophobic
Colloids." Elsevier, Amsterdam.
Yamazaki, S., Fukuda, M., and Himachima, M. (1968). Kobunshi Kagaku 25, 203.

3
Theoretical Predictions
of the Particle Size and
Molecular Weight Distributions
in Emulsion Polymerizations
Gottfried Lichti, Robert G. Gilbert, and
Donald H. Napper

1. Prediction of the PSD .


A. Simple Approximate Approaches to the PSD
B. Population Balance Models of PSD .
C. Batch Polymerizations .
D. Semicontinuous Emulsion Polymerizations.
E. Continuous Emulsion Polymerizations .
F. Experimentallnvestigations of the PSDs
G. Conclusions on the PSD '.
11. Molecular Weight Distributions
A. Introduction .
B. ElementaryConcepts .
. C. General Theory of Emulsion Polymer MWDs
D. Sample Evaluationof MWDs .
E. Alternative Formulations of the MWD .
F. ExperimentalDeterminationof the MWD .
111. Separability of MWDand PSD
IV. Conclusions .
References

94
95
96
99
105
105
109
114
115
115
116
120
124
134
139
141
142
143

The technique of emulsion polymerization is characterized by


the formation of the polymer in the form of a latex. The particle size
distribution (PSD) of the latex and the molecular weight distribution
(MWD) of the contained polymer are two important measurable parameters of the latex. Not only do they inftuence the end-use behavior of the
product, but they also reftect the growth history of the emulsion polymerization process. In what follows, we review the theories that have been
developed to describe the PSD and MWD of emulsion polymers.
93
EMULSION

POLYMERIZATION

Copyright
<!:>1982 by Academic Press, Inc.
Al! rights of reproduction
in any form reserved.
ISBN 0-12-556420-1

94

l.

Gottfried Lichti el al.

Prediction of the PSD

AlI the observable properties of an emulsion polymerization, such as the


PSD and MWD, are gover1ed by a single set of rate coefficients for the
various microscopic processes occurring therein. Although the two distribution functions are interconnected, it is possible, without loss of generality,
to discuss separately the mathematical modeling required to compute these
two observables (see Section III). It must be recognized, however, that the
various rate coefficients that can be determined empiricalIy are themselves
commonly a function ofthe PSD and, less commonly, ofthe MWD.
We now consider the problem of predicting the PSD (or the inverse
problem of deducing rate coefficients from an observed PSD) without
reference to the nature of the MWD generated. There have been two broad
strategies adopted with respect to the prediction and interpretation of
experimental PSDs. The first, exemplified by the work of Min and Ray
(1974), is a "global" strategy which recognizes the considerab1e complexity
of industrial-type emulsion polymerizations. The system is modeled on the
basis of plausible theoretical arguments for each of the many microscopic
kinetic processes included therein and incorporates rate coefficients taken
from the literature for similar or analogous types of systems.
In contrast, the second approach, which has been favored by the present
authors (Hawkett et al., 1980; Lansdowne et al., 1980; Lichti et al., 1981),
adopts a more pragmatic strategy in that simple model systems are studied
experimentally' and theories constructed to permit several different determinations to be made of each required parameter. This obviates the need
to appeal to literature values that may not be applicable to the systems in
question. The global approach has the advantage that it considers problems
of direct interest to industry. It is open to criticism in that the results may
be ambiguous: the same agreement between theory and experiment could
often have been reached with a different set of parameters and/or other
equally plausible theoretical premises. Literature values for many rate
coefficients for polymerization processes display considerable variation. For
example, the reported values of the propagation rate constant (kp) for
styrene at 50C span the range 100-400 dm3 mol-1 sec-l, yet values of the
average number of free radicals per particle () are often calculated by
selecting one specific determination of kp. The proponents of the more
pragmatic approach, on the other hand, have confined their attention to
well-characterized "model" systems. Such studies are, as yet, of limited
relevance to industrial systems, although the determination of an unambiguous set of rate parameters should in the long term permit a better
understanding of systems of technical importance.

95

3. The Particle Size and Molecular Weight Distributions

A. Simple

Approximate Approaches to the PSD

We define the PSD by n(O",


t), the relativenumber of latex particles of size
O"at time t; O"may be the volume (V),area, or diameter (D)of a swollenor
unswollen particle. The most rudimentary method of describing a PSD is to
specify the average particle size, together with the breadth of the distribution. This method was adopted by Ewart and Carr (1954), who used two
separate measures for the breadth of the PSD. These were the mean square
deviation of (i) the particle diameter and (ii) the cube of the particle
diameter. The former is defined by
(1)
with an analogous definition for the latter. In terms of the continuous
distribution function n(O",t), where O" == D, we can write

~D =

{[f

D2n(D, t) dD

I f n(D,

t) dD

]-

[f Dn(D,

t) dD

If n(D, t) dD J}

1/2

It may appear at first sight that 15 and ~D in fact are the only measures
of the PSD necessary (in a unimodal system), because of the following
argument. If one supposed that each latex particle in an emulsion polymerization were to grow at constant rate in terms of the volume added per unit
time, then the mean square volume deviation of the PSD would be constant
after Interval I (the nucleation period). The PSD expressed in terms of
volume would thus translate along the volume axis (i.e., the distributions
would be superimposable) as the experiment progressed in time. On the
other hand, if each particle were to grow ata constant rate in terms of
radius added per second, then the radius distribution would translate along
the radius axis after Interval 1. In fact, a real latex system cannot obey
either type of behavior exactly because not all particles actually contain a
growing free radical at any given instant. The foregoing argument involving
the translation of the PSD on the size axis remains valid provided only that
the fluctuations in the average polymerization times from particle to
particle be negligible. This permits the real particle tha~ polymerizes for half
the time to be replaced by an "equivalent" particle that grows continuously
at half the rate of a particle containing one free radical. Such considerations
are clearly of only limited value, and more complete measures of the PSD
are required.
Another measure of the PSD, one used by Gerrens (1959), is the mean
square volume deviation, which is directly proportional to the mean square
deviation of the diameter cubed. Gerrens was aware that such a description

Gottfried Lichti et al.

96

is only meaningful when all the PSDs to be compared have the same
Gaussian or normal distributions. His experimental evaluation of several
PSDs of polystyrene latexes showed that the distributions were almost
Gaussian when expressed in terms of the particle volume.
An additional measure of the PSD is the skewness parameter s, defined
by Gardon (1968b) as the exponent of the radius such that the PSD
expressed in the size parameter (radius)' is most nearly Gaussian.
The preceding approaches to characterizing the PSD are adequate if the
PSDs always closely approximate a particular type of distribution. If,
however, the shape of the PSD changed markedly during polymerization,
or if the PSD do es not conform to a recognizable functional form, greater
mathematical sophistication is required.

B.

Population Balance Models of PSD

A comprehensive theory of the PSD should account for the time


evolution of the PSD throughout the emulsion polymerization. The basic
concepts widely accepted as describing the growth of polymer latex particles
were originally advanced by Harkins (1945, 1946, 1947, 1950) and Smith
and Ewart (1948). Each latex particle contains an integral number i
(O,1,2, 3, ...) of polymerizing free radicals. Henceforth, i will be referred 'to
as the "state" of the latex particle, The fraction oflatex particles that are in
st'ate i at any time t is denoted by N(t).
These Ni values relate directiy to the rate of polymerization through the
average number of free radical s per particle, = L iNi (with the normalization condition L Ni = 1). The Smith-Ewart equations describing the time
evolution of N are
dN;/dt

= p(Ni-l - N}+ k([i + 1JNi+l - iNi)


+ e([i + 2J[i + lJN+2 - i[i -1JNi)

(2)

for all i ~ O. Here p is the first-order rate coefficient for the entry of free
radicals into latex particles, k the first-order rate coefficient for exit (desorption) of free radicals from the particles and 2e pseudo-first-order rate
coefficient for the annihilation of pairs of free radical s by bimolecular
termination. Note that in Eq. (2), the term with a negative state coefficient
(for i

= O) is

ignored and that bimolecular

termination

only contributes

if

i ~ 2. Note, too, that p and k give, respectively, the average number of free
radical s that enter and exit from each particle in unit time and that the
average lifetime of an isolated pair of free radicals in a latex particle is
about !e.

,;

...

3. The Particle Size and Molecular Weight Distributions

97

The quantity N(t) alone contains insufficient information to specify the


PSD of the growing latex because it refers to the fraction of particIes in
state i at time t, irrespective ofparticle size. To circumvent this problem, it is
necessary to define the number density distribution of particles of size (1in
state i at time t by nM, t).
The actual fraction of particIes with sizes in the range (a, b) can be
evaluated from the integral of n(1,
t) between these limits. Thus, n(1,
t) is
related to the previously defined N(t)through

N(t)

(3)
100

n(1, t) d(1

from which it can be seen that the dimensions of n(1,


t) are (size)-l. The
observable PSD n(1,t) at any time during the experiment may be calculated
from n(1,
t) by a summation over all states:
n(1,t) =

L n(1, t)

(4)

The n(1,
t) term is a function of two independent variables, (1and t. The
equations describing them are a family of partial differential equations. The
basic forro of these equations, as derived from population balance considerations, has been recognized by many authors (O'Toole, 1969; Sunberg
and Eliassen, 1971; Pis'men and Kuchanov, 1971; Min and Ray, 1974;
Lichti et al., 1977) and may be written in the compact but general form
.
onlot

= en - o(Kn)lo(1+ e

(5)

Here n(1,t) i~ a vector whose (i + l)th component represents the distribution n(1,
t) and e is a square matrix formed from the kinetic parameters
p, k, and c. The nature of e may be recognized by expressing Eq. (2) in
matrix form and identifying the resulting square matrix with e. e is called
here the Smith-Ewart coupling matrix because its elements in Eq. (5)
describe how latex particIes change state as a consequence of the SmithEwart mechanisms (entry, exit, and bimolecular termination).
The matrix K in Eq. (5) is a diagonal matrix ~hose (i + 1, i + l)th
element specifies the rate of growth of a latex particIe in state i. The term
-o(Kn)lo(1 in Eq. (5) accounts for the particIe growth process whereby
particles of size (1 are lost to the population n(1,
t) when they grow to
another size (1+ d(1,whereas particles are gained in the population n(1,
t)
when particIes of size (1 d(1 in state i grow to size (1(Lichti et al., 1977).
One common assumption regarding K is that

K+l,+l = iK

(6)

98

Gottfried Lichti el al.

where K = rate of growth of a partiele in state 1. This assumption, which


asserts that a partiele containing i free radicals grows i times more rapidly
than a partiele containing one free radical, is valid provided that partiele
growth is not limited by, for example, monomer diffusion. Note that
K = kpCrJNA where CMequals monomer concentration in the latex partieles and NA is Avogadro's constant.
The vector c in Eq. (5) describes the creation in, andjor remo val oflatex
partieles from, the system. The creation component may arise from in situ
partiele formation (e.g., in Interval 1) or from the flow behavior in a
continuous stirred-tank reactor system (CSTR) with an arbitrary number of
reaction vessels. Partiele remo val terms may be required if coagulation
occurs or in the context of CSTR operation.
All the parameters in Eq. (5) may be functions of (J and t. Equation (5)
may also be coupled to other mass balance equations, for example, that for
the total amount of monomer present in the system. Certain parameters,
e.g., the entry rate coeffcient p, may themselves be functions oe n(J,t). Such
would be the case if a radical that exists from one particle may enter
another (Ugelstad and Hansen\ 1976). This results in highly nonlinear
behavior.
Equation (5) reduces to the Smith-Ewart equation [Eq. (2)] if c is sef
equal to zero and if both sides of Eq. (5) are integrated between (J = O and
(J = oo. This reduction further requires the assumption that all rate coef-

ficients forming the elements of a are independent of (J. It is evident that


the population balance Eq. (5) are considerably more general in scope than
the Smith-Ewart equation because the inelusion of the size parameter
enables the formalism to model the partiele formation process, as well as
both the kinetics and the evolution of the PSD.
The Range of Applicability of the Populat~onBalance Equations
The population balance equations are very general and may be applied
to batch, semicontinuous, and continuous emulsion polymerizations.
Furthermore, both seeded and ab initio polymerizations are comprehended
by Eq. (5) in all (or part) of the three commonly considered polymerization
intervals. The following sections show how the different possibilities are
reflected in different functional forms of the elements of the matrices a and
K and of the vector c. It should be remembered, however, that certain
conceivable situations are not comprehended by Eq. (5); for example, if the
monomer molecules are not freely exchanged between the latex partieles so
that the monomer concentration inside each latex partiele is determined by
its growth history.

3. The Partiele Size and Moleeular Weight Distributions


C.

99

Batch Polymerizations
1. lnterval II

Interval 11 of an emulsion polymerization is characterized by polymerization in a constant number of latex particles in the presence of monomer
droplets (Le., nucleation is absent). This situation usually exists in an ab
initio polymerization immediately on completion of Interval 1; however, for
a seeded system, it may exist from the commencement of polymerization.
Clearly, if no new particles are formed and coagulation does not occur, the
vector e in Eq. (5) is zero. Moreover, in the presence of monomer droplets,
the concentration of monomer at each polymerizing site in the particles is
approximately constant. The growth factor K is thus known. The remaining parameters p, k, and e in Eq. (5), or rather size-averaged values for them,
can be evaluated from kinetic studies (Hawkett et al., 1980, 1981;
Lansdowne et al., 1980). All the parameters in Eq. (5) are then specified and
it is possible to predict uniquely the time evolution of the PSD. Of course,
the initial distribution n~q, t = O)must also be specified (Lichti et al., 1977).
Provided that a relatively monodisperse seed latex is used, the size-averaged
kinetic parameters can be employed without introducing significant error.
In this way, the predictions of Eq. (5) can be tested against experiment and,
e.g., the accuracy of the kinetic parameters checked by PSD evolution data.
Additional experiments with polydisperse seed latexes may permit the size
dependence of the parameters to be specified.
a. The Solution of the Population Bala'lce Equations. The solution of
Eq. (5), given the initial conditions and all the requisite parameters, can be
achieved in at least three different ways: (i) by analytic solutions, if they
exist; (ii) by finite difference numerical solutions which may always be
generated (Carnahan et al., 1969), although they may require considerable
computing time and are prone to inaccuracy; or (iii) by the method of
moments (Bamford and Tompa, 1954) which provides an efficient numerical
procedure for certain systems.
Analytic solutions for Eq. (5) provide the most direct path of the
prediction of PSD evolution. For batch polymerizations in Interval 11,
however, analytic solutions have only been achieved forothe so-called zeroone system (Lichti et al., 1981). These are systems wherein negligibly few
particles contain two or more free radicals because of the rapidity of the
bimolecular termination reaction (e.g., in styrene emulsion polymerizations
with smalllatex particles). In this case, Eq. (5) may be written as follows:
ano/at = - pno + (p + k)n
(7)
andat = pno (p + k)nl - aKndaq
(8)

Gottfried lichti et al.

100

The appearance of the term pnl in Eq. (7) reflects the fact that entry into a .
state 1 latex particIe causes a state zero latex particIe to be formed as a
consequence of rapid bimolecular termination. Ir all the paramaters p, k,
and K in Eqs. (7) and (8) can be taken to be independent of time, while
being permitted to be arbitrary functions of size, analytic solutions may be
generated for arbitrary initial conditions (Lichti et al., 1981). These solutions encompass the more restricted cIass of analytic solutions provided
by O'Toole (1969) and Watterson and Parts (1971). The use of the term
analytic is retained Qere even though, for arbitrary initial conditions, one
numerical integration is required; such integrations are very rapidly performed numericalIy. The importance of this analytic solution is that it enables
one to make a thorough comparison between experimental PSDs and
theoretical predictions, with a wide range of assumptions as to the values
and functional forms of the various contributing rate coefficients.
The analytic solution derived by Lichti et al. is applicable to any zeroone system in lnterval n, with no restrictions on the rate coefficients except
that they be independent of time (but may depend on size). Although a
zero-one description is applicable to some important systems, it is not
universally

valid. For systems where exceeds

t, general

analytic solutions

have yet to be developed, and we now examine methods for solving Eq. (5)
for systems of arbitrary .
The second method for solving the PSD evolution equations is bruteforce numerical solution using first-order finite difference. Whereas a solution can always be obtained by this technique, it suffers from numerical
instability, from the lack of any automatic check on accuracy, and from
requiring large amounts of computer time.
The third method for solving Eq. (5), the method of moments, was
exploited by Katz et al. (1969) in connection with MWDs (see Section
n,E,2) and by Sundberg and Eliassen (1971) in connection with PSDs. This
method is numerically efficient. The kth moment mf of the population
n(a,t) is defined by
mNt) = 1<0akn(a,t) da

(9)

Equations involving only a finite number of mf may be derived from Eq. (5)
under appropriate cIosure approximations. As only one variable (t) is then
involved, the moment equations may be readily solved and the populations
n(a,t) generated. This method will often fail, however, when Eq. (5) is
nonlinear (e.g., if reentry is important) or if large numbers of moments are
equired (e.g., for a polymodal PSD).
b. Theoretical Results. To illustrate the use of the analytic solutions
mentioned above and to highlight the general behavior exhibited by the

3. The Particle Size and Molecular Weight Distributions

101

PSD in Interval II batch reactions, the results of a sample calculation are


provided (see Fig. 1). In this caIculation for a zero-one system, it was
assumed that p = 1, k = O,K = 1 and e ~ p, the units being arbitrarily, but
appropriately, chosen. The initial condition selected, which is typical of a
seeded emulsion polymerization system, was that

no(O",t = O)= 10 exp[ -1t(0" - 2)2]


=0

0">0
O"~O

and
(10)
which corresponds approximately to a Gaussian distribution function. In
Fig. 1, we chose O"= V (volume) as the size variable.
Figure 1 depicts the evolving population no(O",
t) and nl(O"'t), as well as
the overall observable PSD n(O",t). Initially, alllatex particles are nongrowing. As polymerization proceeds, more particles enter the growing state.
After some time, the total number of latex particles in each state is equal
(Le., = 0.5), although the population nl(O"'t) is marginally further advanced in size than the nongrowing population no(O",
t). This must be so because
the population nl(O"'t) is growing. The overall PSD moves along the size
axis and broadens (as measured by the decreasein peak height of the
normalized curves) with increasing time. The rate of broadening is greatest
when the polydispersity is least (e.g., the PSD broadens significantly more
in the initial time interval 0-2 than in the next interval 2-4).
It is stressed that the PSD may be expressed as a function of any
arbitrary size variable. For example, distributions n(V) and n(r)expressed in
terms of the particle volumeand radius, respectively,are related by
n(V) dV = n(r) dr

(11)

n(V) = n(r)/41tr2

(12)

For spheres, this implies that

Equation (12)has the interesting,but often overlooked,.consequencethat a


series of evolving PSDs if expressed in volume may broaden with time; yet
the sameseries if expressedin terms of radius may become more monodisperse. This can be seen from Eq. (12) to arise because each element of n(V)
must be multiplied by 41tr2 to yield the distribution n(r); consequently,
elements with larger values of V (and r) are weighted more and so become
more peaked than elements with smaller V values. It is therefore mandatory
to specify the size variable in which a PSD is expressed before commenting
on the nature of the evolving PSD.

102

Gottfried

Lichti el al.

(o)
100 r
80
60
>
o
c: 40
20

(b)

100 r
80
60
>
c: 40
1=2

20f

.
1

(c)
100 r
80

60
>

40

20

Fig. 1.

Parlic1esizedistribution functionsno, nI' and n (= no + ni) for p = K = 1 and

k = O,as functions of time t; size variable is volume V.

3. The Particle Size and Molecular Weight Distributions

103

It should also be noted that the functional dependence of the growth


factor K in Eqs. (5), (7), and (8) is strongly influenced by the size variable
used to describe the PSD. Suppose, for example, K is constant when
expressed as a function of particle volume (Le., K(V) = K), which would
pertain if each free radical incremented the host latex particle by a constant
volume in unit time:

dV/dt

=K

(13)

If the same PSD is now expressed in terms of the radius, the term K(r)
corresponds to the radius increment to the host particle in unit time as a
result of growth of a single free radical. Equation (13) yields
dr/dt

= K/4nr2 = r(r)

(14)

which is the requisite form of the same growth factor expressed in terms of
the new variable r.
2.

lntervall

The prediction of the evolution of the PSD in Interval n is simpler than


that in the other intervals and it was for this reason that it was discussed
first. Even the qualitative

features of particle formation

in Interval I are in

doubt and the relative importance of homogeneous (Le., oligomeric prcipitation) versus heterogeneous (Le., micelIar) nucleation mechanisms are
not fully understood. For this reason, detailed solutions to Eq. (5) in this
Interval, when e is nonzero, appear to be premature. Moreover, in many
emulsion polymerizations, the precise detail~ of events occurring in Interval
I are masked by the subsequent particle growth in Intervals n and nI.
The global strategy endeavors to enCQIIlpass all possible nucleation and
coagulation m~chanislI\s. Min and Ray (1974) have identified four different
mechanisms that lead to a change in the number of particles, although this
list may not be exhaustive: (i) the coagulation of latex particles with
micelles, (ii) the mutual coagulation of latex particles, (iii) the entry of free
radical s into micelles, and (iv) the precipitation of an aqueous-phase
oligomeric radical that has exceeded its criticallength.
The four processes
all contribute to the vector e in Eq. (5). These terms are evaluated in a
nonrigorous fashion. For example, the rate of coagulation of particles of
volume V1 with those of volume V2 is taken by Min and Rayas proportional to the product of the respective numbers of particles, the constant
of proportionality being given by
kc

= C1 exp( -

E* /kB T)(Vl V2)-1/3

(15)

In Eq. (15), E* equals the activation energy for coagulation and Cl is a


constant.

Gottfried Lichti el al.

104

With this type of approach, two comments must be borne in mind: first,
at this level of complexity, the proliferation of adjustable parameters almost
assures that any experimental data can be fitted (Le., the model is virtuaIly
unfalsifiable experimentaIly); second, only plausibility arguments were used
to establish Eq. (15). It might be argued, for instance, that E* should be a
function of particle size for charged species and that a formalism that treats
miceIles and smaIl latex particles in an identical fashion is unrealistic.
Furthermore, all particle sizes under consideration in Interval I are relatively smaIl 10 nm typicaIly); arguments leading to Eq. (15) have only
been validated for considerably larger particles. These criticisms are meant
only to stress that the coIloid science associated with Interval I is still
poorly understood. Progress in this are a is therefore likely to be slow.
Fortunately, as mentioned previously, particle formation by whatever
mechanisms are operative usuaIly produces particles whose size is smaIl
relative to that attained at the end of Interval In. It then becomes possible
to assume with little loss of accuracy that aIl particles are formed at the
same (smaIl) size.

3. IntervalIlI
.

Interval nI is characterized by polymerizationin a constant number of


latex particles in the absence of monomer droplets. The concentration of
monomer in the latex particles CM thus declines as polymerization progresses. The principal modification required for the application of Eq. (5) in
modeling Interval nI is its coupling with a monomer balance equation and
the modification of the growth factor K to incorporate the declining
monomer concentration. A full monomer balance equation for a batch
reactor demands the consideration of the monomer consumed in both the
aqueous phase and the particles (Min and Ray, 1974). Frequently, however,
the aqueous phase consumption is relatively smaIl so that only consumption in the latex phase is significant. The latter is given by

dCM/dt

= ACM

(16)

where A is kp/NAVSand Vs the average swoIlen volume ofthe paiticles.


Equations (16) and (5) are mutuaIly coupled as foIlows: CMenters the"
evaluation of the growth parameter K [see Eq. (6)] and thus the matrix K
in Eq. (5); conversely, is required in Eq. (16) and is ca1culated by reference
to Eq. (5) through the relationship

= i~O

n(a,t) da

(17)

Solutions to these relationships have been developed by Wood et al. (1981).

...

3. The Partid e Size and Molecular Weight Distributions

105

Other materials balance equations may be coupled with Eq. (5) in an


way analogous to the monomer balance equation. This is necessary if the
availability of some species changes significantIy with time, e.g., when the
initiator has a half-life comparable to (or shorter than) the polymerization
half-life.
.Jt
D.

Sem;cont;nuous

Emuls;on Polymer;zat;ons

Semicontinuous emulsion polymerizations are characterized by the continued addition of monomer to the reaction vessel. This permits the
production of latexes with high weight percentage solids while allowing the
initial burst of nuc1eation to be achieved in substantially aqueous surroundings. The theory for semicontinuous systems is substantially that set forth
for Interval In of batch polymerizations, except that the materials balance
equations [Eq. (17)] must be modified to inc1ude the flow of new material
into the reactor. The effect of the monomer input is twofold: first, the mass
of material present in the system is increased and second, the concentration
of other reagents may be reduced.
From a practical viewpoint, a semicontinuous process may be used to
increase the duration of Intervals n or In. Thus, for example, by extending
the duration of an Interval In which displays a Trommsdorff gel effect with
its concomitant rate acceleration, a shorter reaction time may be achieved.
E.

Cont;nuous Emuls;on Polymer;zat;ons

A continuous emulsion polymerization is characterized by (i) the continual addition of monpmer, surfactant, and initiator to the reaction vessel
and (ii) the withdrawal of a steady stream of reactor fluido Several such
reaction vessels may be set up in series with one another so that the effiuent
from the first feeds the second, etc. Any species in such a system remains in
the reaction vessel for a finite time, termed the "residence time." Such
reactor vessels may be either completely or incompletely mixed and operate
under either isothermal or nonisothermal conditions. The incompletely
mixed, nonisothermal reactor is characterized by the existence of spatial
concentration and temperature gradients in the reactor.
The principal advantage of continuous reaction vessels is that they
operate (after an initial transient period) under steady-state conditions that
are conducive to the formation of a highly uniform and well-regulated
product. In this section, we shall confine the discussion to continuous
stirred-tank reactors (CSTRs). These reactors are characterized by isothermal, spatially uniform operation.

106

Gottfried lichti et al.

Equation (5) may be used to model a eSTR by modifying the vector e to


allow for the removal of latex particles in the effluent stream while
incorporating the particle creation terms of the type discussed above. The
form of the term describing latex particle removal is given simply for all i by
c =

-NJr

(18)

where t is the mean residence time. This may be readily caIculated from the
volume of the reactor and the efflux rateo
The addition of a term such as that in Eq. (18) to Eq. (5) makes a
profound difference to the evolution of the PSD in time. For example, true
steady-state solutions to Eq. (5) may be generated by setting the lefthand side equal to zero. The PSD does not evolve in time in that case but
remains static under the competing influences of particle nucleations and
remo val. This type of behavior is unique to continuous systems. Uually,
steady-state conditions prevail in eSTRs only after a period equal to
several mean residence times has elapsed.
Thompson and Stevens (1977) have numerically solved the steady-state
version of Eq. (5) for a multistate emulsion eSTR, allowing for all the
usually considered free-radical events. The problem contains only one
independent variable, the particle size (1.It may be solved with considerable
precision using sophisticated finite difference methods. The actual shape of
the steady-state PSD is quite unlike that observed in the batch process (see
Fig. 2), being a monotonically decreasing function of particle volume rather
than a peaked function. This feature of the PSDs of eSTRs arises from the
greater probability of removal of particles in the effluent as their times of
residence (Le., growth times) increases. It is emphasized that the foregoing
argument is valid only when the PSD is expressed in terms of volume; the
PSD produced in a eSTR does show a maximum when expressed in terms
ofradius [see Eq. (12)].
The PSDs produced in emulsion eSTRs tend to extend over a wider
range of particle sizes than those generated in a batch process. This arises
from the constant creation of new particles in a eSTR compared with the
short burst of nucleation in a batch system. One consequence of this
broadness is that reasonably accurate predictions of the steady-state PSD
can be achieved in such reactors without considering "stochastic" broadening factors (Le., broadening caused by the distribution of free radicals
among the particles, as illustrated in Fig. 1 for a batch system). The
formalism of Eq. (5) fully encompasses the contribution from stochastic
broadening by considering the portion of the overall population in each
state i. eSTR PSDs often display insensitivity to stochastic broadening.
This insensitivity to stochastic broadening considerably simplifies the problem of predicting the PSD in a eSTR system. De Graaf and Poehlein

3. The Particle Size and Molecular Weight Distributions

107

(a)

>-

t-

Z
t.s
e

lb)

a:
t.s
ID
~
=>
z

VOLUME

Fig. 2.

Schematic comparison of the PSDs produced by (a) batch and (b) CSTR

emulsion polymerizations.

(1971) showed that it was possible to calculate the PSD from an expression
for the distribution of residence times (and hence growth times), as well as a
growth rate parameter that depends upon the particle size [see Eq. (21)]. The
crux of their argument will now be presented,
The lifetime distribution of any species in the CSTR g(t) is given by
g(t)

Co exp(

tlr:)

(19)

if the effiux rate is uniformo Here Cois a normalizing factor. Suppose that it
is possible to assign a unique value of particle size O'(t)to a particle that
leaves the reactor after a residence time t. This assumption actually
disregards stochastic broadening by ignoring the fact that particles that
leave the reactor after residence time t will in reality possess a size
distribution. The lifetime distribution expressed by Eq. (19) can then be
transformedinto the steady-statePSD as follows[cf. Eq."(11)]:
g(O')

Co exp[

- t(O')Ir:](dtldO')

(20)

where t(O') is obtained by inverting O'(t). In order to calculate t(O'),or


equivalently O'(t),an equation describing the average growth rate of the
particle is used:
dO'ldt

= K"

(21)

Gottfried Lichti et al.

W8

where ",is the size-dependent average number of free radicals per particle
calculated from a knowledge of the size-dependent parameters p, k, and e
[see Eq. (2)].
The foregoing argument is equivalent to the approach of Stevens and
Funderburk (1972), who used a more abstract line of reasoning. Suppose
that a summation of Eq. (5) is performed over all states i (Le., all the rows
are summed). Then Eq. (5) becomes
.0

L ni/ot = -oKL inJoa- L n;/r:

(22)

Equation (22) follqws from Eq. (5) if the vector e is given the form
appropriate to CST,s [see Eq. (18)] for removal of latex particles in the
effluent stream and if the particle creation terms are considered as boundary
conditions operating at a = o. The term involvingthe Smith-Ewart coupling matrix vanishes on summation because of the conservation of particle
numbers. The value of ",is given by
(23)

",= L ini/L ni
so that Eq. (22) may be rewritten,as
on/ot

(24)

= - oK", n/Ba - n/-r:

where n represents the overall PSD (Le., n = Li n).


For steady-state conditions, Eq. (24) becomes
(25)

oK", n/oa = -n/-r:


This equation

has the general solution

n
where

= coex{

Co is the number

-(f:da/Kn",-r:)JIK

of newly formed

particles

(26)
present

at a

= O.

Equation (26) is identical to Eq. (20),ifviewed in conjunction with Eq. (21).


The utility of the steady-state CSTR methods lies in the fact that the
problem of the steady-state PSD may be couched in terms of uncoupled
equations that are considerably easier to solve than the full problem. Some
caution must be exercised in omitting the stochastic termo The results of the
comprehensive calculations of Thompson and Stevens (1977) show that in
certain instances, the steady-state PSD is sensitive to stochastic broadening.
Finally, we note that the PSD in a CSTR is strongly sensitive to the
residence time distribution, which may be varied over a wide range.
Consequently, the production of a latex with a desired PSD is usually more
readily achieved with a CSTR process than a batch or semicontinuous
process, for the latter depend in a complex manner upon many mechanisms.
The production of monodisperse latexes is an exception to this rule: these

3. The Particle Size and Molecular Weight Distributions

109

are more readily prepared by a batch or semicontinuous process since the


CSTR product has a much broader distribution.

F. Experimentallnvestigations of the PSDs


1.

Batch Polymerizations

Min and Ray (1974) in their review article detailed many experimental
studies on the PSDs of emulsion polymers. In this section, some more
recent results that compare the predictions of the population balance
equations [Eq. (5)] with experimental data will be discussed.
The actual acquisition of reliable experimental PSD data is a task of
considerable complexity. Such techniques as soap titration, turbidity, ultracentrifugation, and light scattering (Morton et al., 1954; Kerker, 1969;
Shaw, 1970) usualIy pro vide only average values for the particle diameter,
although some of these methods can provide PSD data if some functional
form for the distribution function is assumed. The Coulter Counter can
usualIy be exploited only for particles oflarger size (> 0.5 ,um), although it is
not an absolute method (Eckhoff, 1967). Fractional creaming of latexes has
also been proposed as a method for determining PSD (Schmidt and Kelsey,
1951; Schmidt and Biddison, 1960). This method is based on a relationship
between the concentration of creaming agent and the size of the creamed
particles, which is claimed to be independent of the nature of the dispersed
particles. It has the advantage of simplicity, but again calibration is
mandatory. Hydrodynamic and liquid exclusion chromatography (Nagy et
al., 1980; Singh and Hamielec, 1978) have also been proposed as methods
for determining the PSD, although difficulties still exist in relating the
detector signal to the PSD.
By far the most direct method for the measurement of PSDs is the use of
transmission electro n microscopy. The measured PSD is a distribution
expressed in terms of the unswolIen radius (or diameter) of the particles. AlI
sizes are treated with the same statistical weight (provided they are able to
be resolved); this is not the case with some of the sizing methods listed
above (e.g., light scattering), which weight the larger particles more heavily.
One major difficulty with electro n microscopy is that so me latex particles
are adversely affected by the electro n beam. Shadowing techniques and
hardening procedures (Corio et al., 1979) can sometimes be used to
overcome this problem. A further difficulty inherent in any PSD measurement by electro n microscopy resides in the need to measure a large number
of images of latex particles on a photographic print. Gerrens (1959) has
stated that at least 3000 particles must be measured in order to obtain a
statisticalIy meaningful distribution.
This involves considerable
human

Gottfried Lichti et al.

110

effort because current attempts to automate the procedure have thus far
foundered on the difficulty of resolving the extensively overlapping particles
tht inevitably occur in samples of sufficient number density to give
meaningful statistics. It seems likely that sophisticated image analysis
software andjor software that can resol ve overlapping particles will become
commercially available in the near future.
An additional complication arises with relatively mono disperse samples
in that slight changes in the focus of the electro n microscope introduce
random errors in the exact magnification of the final print. These problems
of focus adjustment can be corrected for by the use of an internal standard
(frequently another polymer latex with a significantly larger average particle
size). It is usually mandatory for the PSD of the internal standard to be
completely separate from the PSD of the sample.
Sundberg and Eliassen (1971) ha ve compared the predictions of their
population balance formalism with the experimental PSD data obtained by
Gerrens (1959) for polystyrene latexes. Their model is a zero-one system
with no exit (desorption) from the particles. Although the neglect of radical
desorption for particles of this size is questionable (Hawkett et al., 1980;
Lansdowne et al., 1980), the theoretical curves are in qualitative agreement
with the experimental PSDs (see Figs. 3 and 4). It must be stressed,
however, that the experimental PSDs were taken at the conclusion of the
6

32

1
2
3
4
5
6

28

">
o
H
"

24
20

lO' [1]0 (g cm-3)


3.61
.5.41
10.8
18.0
28.9
54.1

16
12
8
4
O

500

1000

1500

2000

1018 VOLUME (cm3)

Fig. 3.

Experimental PSDs for polystyrene latexes obtained by Gerrens (1959) for

various initiator concentrations


[1]0. Gerrens' d('r.%)/dV, the derivative of the cumulative PSD,
is proportional
to n(O',t) with O'= volume (V). (After Gerrens, 1959; adapted with permission of
Springer- Verlag.)

3. The Partide Size and Molecular Weight Distributions

111

z 0.6
52
1
Z
~

I.J...
Z
O

: 0.4
~
m
a:
1IJ)

~
0.2
...J
O
>

o
Fig. 4.

0.5

1.0
1.5
VOLUME

2.0

Theoretical PSDs for polystyrenelatexes predicted by Sundberg and Eliassen

(1971). Initiator concentration curves (mol dm-3): 1, 1.1 x 10-3; 2, 2.2 x 10-3; 3, 5.5 x 10-3;
4, 1.1 x 10-2. Volume is dimensionless. (Reproduced with permission of Plenum Publishing
Corp.)

polymerization reaction and SO incorporate information from the entire


range of kinetic processes (e.g., nuc1eation, monomer-depleted growth, etc.).
It is therefore difficult to assess whether the qualitative agreement between
theory and experiment is significant or whether it results from a fortuitous
cancellation of errors in the theory, especially as the data in Figs. 3 and 4
do not correspond to the same recipe and temperature conditions.
Min and Ray (1978) have compared the predictions of the population
balance model with the experimental results obtained by Gerrens (1959) on
poly(methyl methacrylate) latexes. The correspondence between theory and
experiment is qualitatively acceptable, and the same kinetic parameters
model the kinetic behavior ofthe polymerization process (see Fig. 5). Again,
the PSD was measured at the conc1usion of the polymerization reaction,
raising once more the problem of cancelling errors in the theory. Min and
Ray (1978) used literature values for all but two of the rate coefficients

112

Gottfried Lichti et al.


0.3

N
I

o
~

0.2

>-

1Vi
Z
UJ
e
...J

~ 0.1

Ir
o
Z

80
PARTlCLE

Fig. 5.

DIAMETER

160
Inm)

Comparison of experimer,ttaland theoretical PSDs for poly(methylmethac-

rylate) latexes prepared with various initial initiator concentrations:


[1]0 =
1.8 x 10-3 g cm-3, ... [1]0 = 3.6 x 10-4 g cm-3, predictions of Min and Ray (1978); O and
1::" corresponding experimental data of Gerrens (1959). Ordinate is "normal density,"
proportiomi.l to n(u, t) with u = D (diameter). (After Min and Ray, 1978; adapted with
permission of Journal 01 Applied Polymer Science. Also after Gerrens, 1959; adapted with
permission of Springer-Verlag.) .

required in the polymerization. As noted above, such literature values


frequently display large variations from one source to another.
Lichti et al. (1981) examined the evolution of the PSD during a seeded
styrene emulsion polymerization. The polymerization was sampled only
during Interval 11, thus removing any uncertainties associated with
monomer-depleted growth. The use of well-characterized seed particles
obviates the difficulties associated with nucleation in Interval I and provides
the initial condition for Eq. (5). Exhaustive kinetic studies on this system
showed that the zero-one approximation was obeyed so that Eqs. (7) and
(8) were applicable. The same kinetic studies yielded independent, precise
estimates for the values of the parameters p and K, the effects of desorption
being negligible; values of p and K were then chosen to optimize the
agreement between the theoretical and experimental PSDs at the later time.
Figure 6 shows the good agreement achieved between the calculated and
measured PSDs at the later time (t = 45 min). A comparison between the
values of p and K obtained by the kinetic and PSD experiments was also
presented and excellent agreement found. These experiments demonstrate
that consistent rate parameters may be determined using the kinetic and

113

3. The Particle Size and Molecular Weight Distributions


RADIUS (om)
40
5
102 ('._1
1025 K

7e
~

60

50

PSD KINETICS
2.2
2.3
2.3

2.4

m3 $-1

2:
e

I
o

~\

\
.\

,\

10

Fig. 6.
Comparison of theory and experiment for a seeded emulsion polymerization.
Continuous lines: smoothed partic1e size distribution obtained experimentally; broken line:
theoretical fit using values of p and K shown in inset (values of p and K from kinetic study also
shown) at t = 45 mino (After Lichti et al., 1981; reproduced with permission of Journal of
Polymer Science.)

PSD approaches; alternatively, they demonstrate that the PSDs may be


reliably calculated
from
o
, kinetic data alone.

2. Continuous Emulsion Polymerizations


De Graaf and Poehlein (1971) and Stevens and Funderburk (1972) have
compared the predictions of the simple residence-time theory for the CSTR
with experiment. However, the results of Stevens and Funderburk must be
treated with caution in view of their use of only 30-400 particles to
establish the PSD. Figure 7 shows the comparison of theory with experiment for the cumulative PSD obtained by De Graaf and Poehlein for a
styrene CSTR with a mean residence time of 59 mino The agreement
obtained was good provided that the Stockmayer solution for in terms of
p and e was used (k = O), rather than the Smith-Ewart Case 2 (i.e., was
greater than t). Note, however, that De Graaf and Poehlein assumed that
free-radical desorption need not be taken into account; moreover, they
assumed that the initiator capture efficiency was 100%. Both assumptions

Gottfried lichti et al.

114
1.0

z 0.8
O
i=
=>
ID

Q?
1- 0.6
VI
i5
ILI
>

~ 0.4
-1
=>

=>

00.2

O'

100.

200

300

PARTICLE DIAMETER(n m)
Fig. 7.
A comparison of theory with experiment for the normalized cumulative PSD,
proportional to Son(O',t) dO',where O'= diameter, obtained by De Graaf and Poehlein (1971) for
a polystyrene CSTR: --- = 0.5; > 0.5; O experimento (After De Graaf and Poehlein,
1971; adapted with permission of Journal of Polymer Science.)

are of dubious validity for the small particle sizes studied (Hawkett et al.,
1980).
G.

Conc/usions ofthe PSD

Because of the large number of mechanistic processes opera tive in


emulsion polymerizations, complete theories for the PSD are necessarily
complex. Nevertheless, they can be formulated by a population balance
approach. Much remains to be done, however, to clarify the basic colloid
science that underpins the nucleation process in Intervall. The experimental challenge in evaluating the predictions of the theory for PSDs resides
not only in the attainment of agreement with experiment but also in
showing that such agreement is not merely fortuitous but arises from the
. correct mechanistic scheme. Considerable experimental work will be required to establish the validity of mechanistic assumptions for any particular monomer.

3. The Particle Size and Molecular Weight Distributions

115

11. Molecular Weight Distributions

A. lntroduction
The molecular weight distribution (MWD) of a polymer generated by
emulsion polymerizations can be fundamentally different from that generated in solution or bulk. For example, in styrene emulsion polymerization
the MWD of formed polymer has a much higher average molecular weight
than may be obtained using other methods. The basic reason for this was
postulated by Smith and Ewart (1948) to be the compartmentalization of
the polymerization reaction inside the latex particles which leads to the
isolation of free radicals. This isolation reduces the probability of bimolecular terminations and hence increases the degree of polymerization.
The fundamental difficulty in constructing a theory for the MWD in
emulsion polymers is to account for the compartmentalized nature of the
system. In the commonly occurring situation where particles contain only a
few free radical s at any given time, it is obviously incorrect to consider that
each latex particle behaves like a "mini-bulk" reaction vessel, and so the
conventional methods used for bulk polymerizations are inapplicable.
Nevertheless, some assumptions which introduce only minor errors may
often be made. The most important such assumptions is that the evaluation
of the MWD may be separated from that of the PSD. In other words,
provided that the MWD being produced at any given moment is the same
as would be formed in an equivalent set of monodispersed latex particle
systems [as expressed in Eq. (27) below], then the MWD evolved in a
system that is polydispersed in size may be computed trivially. Formally,
this is expressed as follows. Let S(M, a, t) be the MWD formed in a
monodisperse' system 'Of size a at time t; here M is the molecular weight
variable. In a polydisperse system with PSD n(a, t) the overall MWD at any
experimental time te, defined as S(M, te),is postulated as
""
re
(27)
S(M, te)= O dt O da n(a, t)S(M, a, t)

ii

A detailed critique of the validity of Eq. (27) is given in Section III. In brief,
two criteria are required for this validity, both of which are well satisfied in
ordinary emulsion polymerization systems. These are (i) that the time
required for formation of a single polymer chain be much less than that
over which significant changes occur in the rate coefficients governing the
MWD and (ii) that the average number of free radicals per particle of size a
is close to its steady-state value.

116

Gottfried Lichti el al.

Equation (27) may be expanded to incorporate the contribution to the


MWD from newly formed latex particles by assuming that these enter the
system at some small volume and calculating the MWD in the new particles
separately. Since latex particles grow to many times their original volume
during the course of polymerization, the component of the MWD produced
by nucleation is often negligible (at least, when measured as a weight
average). Any aqueous-phase polymerization may also be included in the
same way.
We now show how to evaluate the MWD in a monodisperse compartmentalized system. It will be seen that the problem may be solved with
complete generality if chain-branching reactions do not occur; moreover,
analytical solutions can be obtained for the steady-state regime.
B.

Elementary Concepts

In this chapter, the term MWD refers to the number density MWD,
S(M), which gives the relative number of polymer molecules of molecular
weight M. Another commonly used measure is the weight density MWD,
W(M), giving the relative weight of polymer molecules. The two distributions are related by
.

W(M)

= MS(M)

(28)

Thus, the relative number of chains in the sample with molecular weights in
the range MI::;;;M ::;;;M2 is given by
N[MI,M2J = rM2S(M)dM

(29)

JMI

and the relative number in a small range M to M + 11M by S(M) 11M. The
normalization (and thus the dimensions) of S(M) will be specified later.
In the ensuing development, S(M) will refer to the "instantaneous"
MWD, Le., the contribution to the MWD created over a comparatively
short time periodo Ir all polymerization occurs under steady-state conditions, S(M) then gives the final MWD of formed polymer directly via Eq.
(27); otherwise, the final MWD is given by a simple time integration over
S(M), as given in Eq. (27) and amplified in Section II,C,2.
1. Review of Bulk MWD Theory
In bulk or solution polymerizations, the following free-radical mechanisms are usually considered: initiation, propagation, chain transfer (to
monomer, polymer or transfer agent), and bimolecular termination (by

3. The Particle Size and Molecular Weight Distributions

117

combination or disproportionation). Except for initiation, all these reactions are bimolecular. Initiation and transfer are chain-starting reactions,
whereas transfer and bimolecular termination are chain-stopping.
A particularly simple theory for the MWD in bulk or solution polymeri. zations

is obtained if long-chain branching mechanismsare ignored and if

the growth of chains is envisaged as a continuous rather than a discrete


process. The latter approximation is valid for almost all systems of interest,
since the average degree of polymerization is usually large. The molecular
weight of a chain is proportional to the growth time t' of this chain:
M = kpCMMot' ==al'

(30)

where kp is the propagation rate coefficient, CMthe monomer concentration,


and Mo the molecular weight of monomer. We note at this point that a may
be a function of time, as it is during Interval III when CMdecreases; this
point will be pursued in Section II,C,2.
G(t),the distribution of growing chains of growth time t', is then
G(t') = Goexp(-Jet')

(31)

where Go is a constant given by the rate of formation of new, growing


chains (involving initiation and transfer rates) and Jerelates to the stopping
of growing chains (involving bimolecular termination and transfer rates).
Equation (31) expresses the fact that, once formed, a growing chain
continues growing unless it undergoes a chain-stopping reaction. As Flory
(1953) has pointed out, these kinetic probabilities imply a chain distribution
that is of the most probable type, which to a good approximation is
exponential.
The required MWD is then given by the distribution of dead (nongrowing) chains; this is readily found as follows from G(t'). For those chainstopping reactions ih which the identity of the growing chain-stopping
remains intact (i.e., transfer to monomer or to chain-transfer agent and
disproportionation), the distribution of dead polymers produced which
have (previously) polymerized for t' is simply proportional to G(t').
Denoting this distribution by Str.iM), we thus ha ve

Str.iM) oc exp( - Jet')

(32)

where we have used Eq. (31) together with Eq. (30) to relate M and t'.
However, in the case of chain stoppage by bimolecular combination,
pairs of free radicals join together to form a chain whose "equivalent"
growth time t~q is the sum of the growth times of the two contributing
chains. Now, for bulk or solution polymerization, the product G(l')G(t)
specifies the distribution of pairs of growing chains of growth times t' and
t, since in a bulk system any growing chain is equally likely to undergo

118

Gottfried Lichti et al.

bimolecular combination with any other chain (it will be seen that this does
not hold in a compartmentalized system). The distribution of dead chains
created by bimolecular combination with equivalent growth time t~q,
denoted Sbc(M) (where M = at~q) is thus proportional to the sum of alI
possible pairs of growth times giving this equivalent growth time:

t~q

Sbc(M

= at~q)oc. o

(33)

G(t' - t~)G(t~)dt~

Thus, from Eq. (31) we have


Sbc(M)oc t~qexp( -

(34)

At~q)

The overalI MWD is then the sum of Sbcand Str,d'


In order to facilitate comparison between MWDs, it is customary to
define a dimensionless quantity P, the polydispersity of the MWD, which is
the ratio ofthe weight average to the number average tnolecular weight:
P = [1'X) M2S(M) dM LX)S(M) dM

J/[LooMS(M)

dM

(35)

If alI chains have the same molecular weight, P = 1; otherwise P > 1. For
monomodal MWDs, the larger the value of P, the broader the distribution.
For the MWD of Eqs. (32) and (33), P = 2 and 1.5, respectively.
2. Emulsion Polymer MWDs: Concepts and Limiting Cases
The concepts introduced above require extension and revision when we
turn to an emulsion polymerization. In view of the above discussion, we
may take the system to be monodisperse in volume. First, we briefly
consider the kinetics of such a system. As described in Section I,B, this is
given by the distributions N(t), which are in turn given by the first-order
rate coefficients for entry (p), desorption (k) and bimolecular termination (c).
We note here that
c

= Cbc + Cd

(36)

where Cbcand Cdare the (pseudo-first-order) rate coefficients for bimolecular


combination and disproportionation, respectively.
It is convenient to introduce at this point a further pseudo-first-order
rate coefficient, that for transfer, given the symbolf We write
f

= ktr.MCM + ktr.A[A]

(37)

where ktr.Mand ktr.Aare the rate coefficients for transfer to monomer and to
chain transfer agent A; we assume that if transfer of free radical activity to
A occurs, the species so produced polymerizes further. It is postulated that

3. The Particle Size and Molecular Weight Distributions

119

desorption, if it occurs, proceeds by transfer of free-radical activity to a


small molecular weight species that subsequently leaves the latex particle.
Thus, f as given in Eq. (37) includes desorption as well as true propagative
transfer. To resolve this ambiguity, we redefine f as the rate coefficient for
transfer events that do not result in desorption, i.e., we set

f = ktr.MCM
+ ktr.A[A]- k

(38)

We will consider the MWD in two simple cases. The first is when chain
transfer is sufficiently rapid to ensure that all other chain-stopping events
can be ignored. In such a situation, whereas the compartmentalized nature
of the reaction may affect the rate of initiation of new chains, it will not
affect the lifetime distributions of the chains once they are formed. The
MWD may then be found from the bulk formulas, provided only that , the
average number of free radicals per particle, is known. Such an approach
has been used by Friis et al. (1974) to calculate the MWD evolved in a vinyl
acetate emulsion polymerization. These authors included in addition the
mechanisms of terminal bond polymerization and of transfer to polymer
(both of which cause broadening). The formulas required for the incorporation of these mechanisms could be taken from bulk theory.
It is important to note that, even in this present limiting case of a
transfer-dominated system, the chain-stoppage mechanism can be changed
by compartmentalization. Thus, the MWD formed in the polymerization of
styrene appears to be transfer-dominated in some emulsion systems (Piirma
et al., 1975) but to be combination dominated in bulk or solution (George,
1967). This difference occurs because, in ~tyrene emulsion systems, the rate
of radical entry into a particle is slow, and most particles usually contain
either zero or one free radical. In the "state one" particles (Section I,B), the
growing free,radical ras time to undergo several transfer reactions before a
further entry causes radical annihilation.
The second simple case we consider in this section arises in a zero-one
system (Section I,C,l,a), where the average time between successive entries
of free radicals into a latex particle is short compared with the time taken
for a growing chain to undergo a transfer reaction. In this case, most chains
cease growth by bimolecular termination. Ir this happens by combination
with a low molecular weight entrant species, the following approach is
possible. When entry occurs in a state zero particle, a free radical is
formed and grows uninterrupted until the next entry event. Since the rate
coefficient for entry is essentially constant, Eq. (31) gives the distribution of
growing chains. Since the terminating free radical is of negligible molecular
weight, the MWD will then be given by Eq. (32). The average chain growth
time A.-1 is the average time between successive entries into the same
particle (viz., p -1; i.e., p = },).

120

Gottfried Lichti el al.

In concluding this section, we note that except for the limiting cases
given above, the phenomenon of long-chain branching has not thus far been
incorporated into the models for emulsion polymer MWDs. In the general
MWD formulation given below, it is assumed that these branching events
occur with negligible frequency.
C.

General Theory 01 Emulsion Polymer MWDs


1. "Singly Distinguished" Particles

The development here follows that of Lichti et al. (1980). By analogy


with the MWD calculation for bulk and solution polymerizations presented
earlier, the MWD formalism for monodisperse emulsion systems requires
the evaluation of certain types of free-radical growth time distributions.
Because of the variable nature of the reaction loci (depending on the state i),
a separate growth time distribution is required for the population of
particles in each state i. It is therefore convenient to define the distribution
of singly distinguished latex pa;ticles in state i, denoted N;(t, t'), as the
relative number of latex particles inside which a certain free radical began
growth at time t, and continued growing in an uninterrupted manner for a
further growth time t'. The state of the latex particle at time t + t' equals i.
The free radical with growth time t' is called the distinguishing free radical.
When the distinguishing free radical ceases growth, the latex particle is no
longer said to be distinguished. These N; are analogous to (but quantitatively different from) the G for bulk systems (Section U.B.1). Note that for
a distribution of singly distinguished free radicals N;(t, t') evaluated at
experimental time te, we have

te = t + t'

(39)

from the above definition. It will be seen that this feature of the formalism
has the important consequence that the evolution equations governing
N;(t, t') are simply differential equations containing terms in only one of the
independent variables, viz., t'.
For N;, the minimum value of i is unity, not zero, since the definition
requires the presence of at least one growing chain. If N;(t, t') is integrated
over all values of the growth time t', the resulting integral counts each of the
free radicals inside the particle. Thus, from Eq. (36), we have the following
normalization:
rte

iNi(te)

= lo

N;(te - t, t') dt'

Since Ni is dimensionless (by the normalization condition


dimensions of reciprocal time.

(40)

L Ni = 1),N; has

3.

The Particle Size and Molecular Weight Distributions

121

Because N(t, t') is a function of two independent variables, it would at


first appear that the concomitant family of evolution equations would be,
for example, coupled partial differential equations, and thus difficult to
solve. Indeed, some formalisms for emulsion polymer MWDs (Katz et al.,
1969) are based on distribution functions that suffer from this difficulty.
However, N(t, t') is defined such that this problem is avoided, since the
evolution equation separates into equations in t' alone with t being merely
parametric. In fact, we have
oN/ot' = pN;-l - [p + ik + f + i(i - l)cJN
+ ikN+l + i(i + l)cN+2

(41)

This has been derived in detail by Lichti et al. (1980). In brief, this equation
states that singly distinguished latex particles in state i form (i) by entry into
an N-l type particle, (ii) by desorption of nondistinguishing radicals from
an N+l particle, (iii) by termination among any of the (i + 1) nondistinguishing radicals in an N+2 particle. Similar statements hold for the
loss term in Eq. (41).
Since Eq. (41) involves variation in t' only, it may be solved given the
initial conditions, viz., N(t, t' = O). These last quantities are the distributions of ordinary latex particles in state i inside which a new chain begins
growth at time t:
N(t, t' = O)= pN-l(t) + ifN{t)

(42)

Equation (42) asserts that entry and transfer (involving any of the i chains
present) are the only chain starting reactions.
Note that since N(t, t') usually decay monotonically in t', no steady-state
approximati<:>ncan ~e made to simplify Eq. (41). Nevertheless, there are
only a set of differential equations in one variable (as noted above) which
can readily be solved (e.g., as in Section II.D.2).
Having specified N, we now show how to compute the required MWD,
or more specifically, the component of the MWD arising from the mechanisms considered in this section. This is the distribution of dead chains
formed as a result of transfer, exit (desorption), and termination by disproportionation. We denote the components ofthe instantaneous MWD arising
from each of these mechanisms by Str(t., M), S.it., M) and Sd(t., M), where
t. is the experimental time. We have

L fN(t., t')
S.x(t., M) = L kN;(t., t')
i~l
Stit., M) =

Sd(t., M) =

(43)

i;?;l

L 2(i -

i?t2

I)CdN(t., t')

(44)
(45)

122

Gottfried lichti et al.

Here t' and M are related by Eq. (30). The right-hand sides of Eqs. (43-45)
are the rates at which the distinguishing growing chain can cease growth by
each of the events considered.
It can be seen that Stn Sex, and Sd each has dimensions of(time)-z. The
meaning of these distribution functions is as follows: the number of chains
formed by transfer whose growth times are between t' and t' + dt', over the
range of experimental time from te to te + dte, is given by SIrdt' dte (a
dimensionless quantity), and similarly for Sex and Sd' Note that the
expression for Sd is simpler than that given by Lichti et al. (1980), although
the more complex expression for Sd given by these authors reduces to Eq.
(45) after suitable manipulation.
2. " Doubly Distinguished" Particles
The N; values are insufficient to account for the contribution to the
MWD from bimolecular termination by combination: since tfiis requires a
knowledge of the growth times of each of a pair of growing chains.
In the bulk system considered in Section II,A,l, the distribution of free-

radical pairs with growth times '1/1 and t~ was shownto be givenby the
product G(t~)G(t~).However, the equivalent relation in an emulsion system
does not hold. A simple counter example will illustrate this.
We consider a system containing only two particles in state 2. Let
particle 1 contain growing chains with growth times t~ and tB' and particle
2 contain growing chains with growth times t~ and tD' The fraction of
particles in state 2 which contain a chain of growth time t~, which we
denote N~(t~), is 0.5, and similarly for N~(tB)'N~(tC),and N~(tD)' Note that,
as in Eq. (40)
N~(t~) + N~(tB)+ N~(t~) + N~(tD) = 2
since each particle is counted twice in this expression. Now, the fraction of
particles that contain a pair of chains of growth times t~ and t~ is evidently
zero, 110t N2,(t~)N~(tC)= 0.25, as would be given by the product formula.
The reason for this difference between bulk and compartmentalized systems
is that whereas all free radicals in a bulk system are mutually accessible, in
an emulsion system only radicals inside the same latex particle are mutually
accessible for bimolecular termination.
We therefore define a new distribution N?(t, t', t") as being that of
"doubly distinguished" latex particles in state i, Le., particles in which one
free radical began growth at time t and continued growing for a time t', at
which time (t + t') another free radical began growth in the same particle,
and both are still growing at time t" later, the particle being in state i. From
this definition, the experimental time te is given by
te

= t + t' + t"

(46)

3. The Particle Size and Molecular Weight Distributions

123

The two free radical s which started at t and t + t' are termed "distinguishing," and the latex particle ceases to be doubly distinguished when either or
both of these distinguishing free radicals cease growth. Obviously, the
minimum value of i for which N7 is nonzero is 2.
We define N;' to be normalized such that the integral of N7 over all
possible values of t' and t" counts all possible pairs of growing free radicals
in particles in state i. Since the number of pairs in a single particle in state i
is ti(i

1), we have
fte ftte

Jo Jo N;'(t. - t' - t", t', t") dt' dt" = ti(i - I)N(t.)


(47)
Note that this definition corrects a minor error in the work of Lichti et al.
(1980). From Eq. (47),we see that N7 has dimensions of(time)-2.
Although N;'(t, t', t'') is a function of three independent variables, the
evolution equation governing variation with t" can be shown to contain t
and t' only parametrically, analogous to that for N;(t, t'). In fact, it may be
shown (Lichti et al., 1980) that
ON;'/ot" = pN;'-l - [p + 2f + ik + i(i - l)c]N;' + (i - l)kN;'+l
+ i(i - l)cN;'+2 i ~ 2
(48)
The usual population balance methods are used to derive this equation:
thus entry into an N;'-l particle creates an N;' particle, and so on; note in
particular that transfer occurs with arate coefficient of 2f, since f is that for
transfer of a single free radical, and we here have two distinguishing free
radicals.
Since Eq. (48) involves variation with respect to t" only, it may be solved
if the initial conditions N;'(t, t', t" = O) are known. These initial conditions
are the distributions of singly distinguished particles in state i inside which a
new chain starts growing at time t + t'. Thus
N7(t, t', t" = O)= pN;-l(t, t') + (i - l)fN;(t, t')
(49)
because only entry and transfer are chain-starting reactions. The coefficient
(i - 1) arises because the chain-starting transfer must not involve the first
distinguishing radical.
We see therefore from Eqs. (49), (48), (42), and (41) that the multiplevariable distribution functions represented by N7(t, t', n are in fact obtained
as the solutions of sets of coupled hierarchical ordinary or simple partial
differential equations: the solution of one set [for N(t)] pro vides the initial
conditions for another set of partial differential equations [for N;(t, t')]
which in turn pro vides the initial conditions for N7(t, t', t"). Obviously, this
could be extended to n-tuply distinguished distributions, but this is unnecessary since the order of polymer-producing reactions never exceeds
two.

Gottfried Lichti et al.

124

We now show how to use N' to obtain the contribution to the MWD
arising from bimolecular termination by combination. When two distinguishing chains, one of which has been growing for time t" (this being the
second chain formed)

and the other for t'

+ t"

(this being the first) undergo

combination, the resulting single dead chain has an equivalent growth time
of t~q= t' + 2t". The rate coefficient for this event is 2c/(the factor of 2
arising from the usual definition of c as being per pair of free radicals).
Thus, the relative number of chains of equivalent growth time t~qpresent at
experimental time te formed by bimolecular combination is

t,~q

Sbc(te, M) =

dt" L: 2ccN'(te
i

t~q

+ t", t~q-

2t", t")

(50)

where at~q = M as usual. Equation (50) is obtained as follows. First, the


upper limit of the integral is the maximum value of t" such that a chain of
growth time t~q can be produced by combination. Second, the integral is
performed over all pairs of chains (after summation over all doubly
distinguished particles in all states i) one of whose growth time is t" and the
other whose growth time is such as to give a chain of growth time t~q, after
combination
has occurred. The te - t~q + t" is the starting time for such
chains so that the time at which combination
occurs is te; note here that
since the lifetime of chains is normally much less than experimental times,

one has te - t~q + t"

~ te

to an excellent approximation.

Equation (50) defines a distribution as do Stn Sex, and Sd, so that the
number of chains formed of growth time in the range t~qto t~q+ dt~q over
the experimental time te to te + dte, is Sbc dt~q dte. These four contributions
to the instantaneous
MWD may thus be directiy summed to give the
overall instantaneous MWD from all sources, S(te, M):
S(te, M) = SIr + Sex + Sd + Sbc
te,

(51)

Finally, we recall that the instantaneous MWD, S(te, M), may vary with
The accumulatedoverall MWD of formedpolymer at time t* is simply
Ct'

S(M, t*) ~ Jo S(te, M) dte

(52)

If S is independent of te, one simply has S = teS; such is the case, for
example, during the Interval n steady state. In other cases (e.g., Interval In)
the variation of quantities such as CM (and hence a) with time will be known
from the kinetics; the evaluation of Eq. (52) is then a simple quadrature.

D.

Samp/e Eva/uation 01 MWDs


The formalism given in Section n,e involves a large number

of equations

whch, although posing no difficulties so far as developing solutions, can


best be understood by examining their behavior in limiting cases.

3. The Particle Size and Molecular Weight Distributions

125

1. Zero-One-Two Systems
We first consider a system in which the maximum number of free
radicals per latex particle is 2. This is termed a "zero-one-two" system. An
analytic solution to the MWD equations for such a system in the Interval II
steady state has been developedby Lichti et al. (1980).
The zero-one-two model accurately describes systems wherein (the
average number of free radical s per particle) does not exceed 0.7; it is thus
applicable to small-particle styrene (Hawkett et al., 1980),vinyl acetate
(Ugelstad and Hansen, 1976), and vinyl chloride (Friis and Hamielec, 1975)
emulsion polymerizations. In the Interval II steady state, the analytic
solutions to the various functions involved in the MWD obtained by Lichti
et al. (1980)are as follows; in these expressions,the dependence on te is
suppressed since a steady state is assumed
(53)
N'l(t') = B1 exp( - A+t') + El exp( - Aj)
N~(t') = B2 exp( where

-t{all

},+ t')

E2 exp(

Aj)

(54)

= N;(O) - E, E = [(al + },+)N'I(O) + a2N~(0)]/(},+- A_), A:t =


+ a22:t [(all + a22)2 - 4(alla22
aI2a21)]1/2}, N'(O) = pNo + fNI,

N~(O)= pNI + 2fN2, whereNo = 1- NI - N2,NI = p(p + 2k + 2c)/rx,N2=


p2/rx,with rx= p(2p + 3k + 4c) + 2k(k + c), and the aij are the elements of
the matrix

-p-f - k

k + p/2
-p-2k-f-2c

We further have
N;(t',I t") = [pN'I(t') + fN~(t')] exp( - Qt")

(55)

where Q = p + 2(f + k + c). These results give the singly and doubly
distinguished particle contributions for i = 1 and 2 (no higher value of i
being required for our zero-one-two system) in terms of the various rate
coefficients for the microscopic processes involved. The various components
of the instantaneous MWD are then found from Eqs. (43-45) and (50).
Using the above results, Eq. (50) is found to reduce to

Sbc(M)= 2cc{(fB2 + pBI)[exp(-A+t~q) - exp(-tQt~q)]/(Q- 2A+)


+ (fE2 + pE}[exp(-A_t~q)- exp(-tQt~q)]/(Q - 2A_)}

(56)

Finally, there is an additional contribution to the MWD arising from the


artificial truncation used in a zero-one-two system, Le., from the assumption that entry into a particle in state 2 causes instantaneous termination.
We denote this additional term Sit(M), and find
(57)

126

Gottfried Lichti el al.

This term is in fact a numericaIly inconsequential component of the overaIl


MWD (if < 0.7).
Eqs. (53-56), (43-45), and (30) may then be used to compute MWDs for
some cases of interest; because we are considering an lnterval II steady-state
system, the overall MWD is simply proportional to the instantaneous
MWD:
s(t., M) = t.S(t., M) = t.(Str + S.x + Sd + Sbc+ Sit).
Figure 8 shows the MWD produced in a zero-one-two system where
values of the various rate coefficients have been chosen to give a MWD
dominated by transfer and exit: p = 0.1 sec-l, k = 0.3 sec-l,f=
0.9 sec-1,
e = Osec-1. This gives a steady state of 0.34. Before discussing the form of
the curves, we pause to consider the axes employed in this figure. Because
we are considering a steady state system, we may use as abscissa either
M (molecular weight, in a.m.u.) or t' (growth time, in s). We use both

0.5

0.4

10-5M
6

( b)

(o)

0.4

N
I
U
Q)
~ 0.3
z
O
::
::>
~ 0.2
a::
ti>
5

0.3

0.2

0.1

0.1

O
O

O
2

341
GROWTH TIME (sec)

Fig. 8.
Plots of instantaneous molecular weight distribution S arising from instantaneous termination, stoppage, and transfer, as a function of growth time (Iower abcissa) and
molecular weight (for a typical styrene system; upper abcissa), for a steady-state transferdominated zero-one-two emulsion polymerization, with e = O, p = 0.1 sec-I, k = 0.3 sec-I,
f = 0.9 sec-I. (a) Curves I and 2 give conlribulions lo S from N'I and N2, respeclively. (b)
Overall MWD, S. (Mter Lichli et al., 1980; reproduced wilh permission of Journal of Polymer
Science.)

127

3. The Particle Size and Molecular Weight Distributions

for illustrative purposes, with the value of a computed for styrene


emulsion polymerization
at 50C, using kp = 258 dm3 mol-1 sec-1,
CM = 5.-8mol dm - 3 (Hawkett et al., 1980). The ordinate of the instantaneous MWD plot has dimensions (time)-2, the dimensions of S; the
conversion from S to the relative number of chains of a given molecular
weight was given in Section n.c.l.
We consider the curves given in Fig. 8. In Fig. 8A, curves 1 and 2 give
the instantaneous MWD arising from distinguished partic1es in states 1 and
2, respectively. Fig. 8B gives the overall instantaneous MWD, which is a
monotonically decreasing function approximating a single exponential. This
behavior, which for a transfer dominated system is identical with that in the
bulk, will be further discussed in Section n,D,3.
Figure 9 shows the instantaneous
MWD computed for a zero-one-two
system where the rate coefficients have been chosen to give domination by
combination: p = 0.1 sec-1, Cc = 1 sec-l, k = O, f= O, Cd = O. This gives a

steady state of 0.55. Figure 9A showsthe contributions to the MWD from


Sbc [Eq. (56)] and the artifactual Sil term [Eq. (57)]. It can be seen that Sil
indeed contributes but little to the MWD, showing that the artificial
truncation method used to reduce the problem to a zero-one-two case (so
0.004
(b)

(o)

.-

...

I
<.>
CII

'"

5.... 0.002
::>
ID

ir
\
e
TT

o
o

1.0

2.0

1.0

2.0

EQUIVALENT
GROWTH
TIME (sec)
Fig. 9.

(a) Plots

of components

of instantaneous

molecular

weight

distribution

arising from bimolecular combination (Sbe' labeled BC) and instantaneous termination (Sil'
labeled TI) for a steady-state zero-one-two system with termination by combination only,
assuming Ce= I sec - 1, P = 0.1 sec - 1, k = f = Cd= O. (b) Plot of S, the overall instantaneous
MWD, for the same system. Note = 0.55. (After Lichti et al., 1980; reproduced with
permission of Journal of Polymer Science.)

128

Gottfried Lichti et al.

as to obtain analytic solutions) does not affect physical content. It can be


seen from Fig. 9B that the overall MWD shows a distinct maximum. This
resembles the behavior in bulk or solution, but as will be seen the similarity
is only qualitative.
Figure 10 shows the MWD for a zero-one-two system with rate
coefficients chosen to give a disproportionation-dominated
process:
p = 0.1 sec-l, Cd= 0.3 sec-l, Cc= O= k = f. This gives = 0.64. Here the
MWD is seen to be monotonically decreasing, again a resemblance to
solution and bulk behavior which is only qualitative.
2. Systems of Arbitrary
We next consider the evaluation of the instantaneous MWD for systems
where may be arbitrarily large. For simplicity we confine our discussion
to a steady-state system in Interval II; the minor adjustments required to
extend the computation to Interval III have been mentioned in Section
n,e,2.
For systems of arbitrarily large , analytic solutions as given in Section
n,D,l are no longer possible. Nevertheless, the numerical evaluation of S
for such systems will be seen to require negligible computational effort. The
procedure is as follows: first, from a given set of values of p, k, and c, one
numerically solves the steady-state Smith-Ewart equations for Ni, i.e., Eq.

0.03
....
I

(\)

'"
~

0.02

z
o
~

:J
CD

g:0.01
IJ)

10
GROWTHTIME (sec)

20

Fig. 10.
Plot of instantaneous MWD S as function of growtb time in a zero-one-two
Interval 11 system witb termination by disproportionation only, assuming p = 0.1 sec-1,
Cd = 0.3 sec-1, c<=

J = k = O. (After

oJPolymerScience.)

Licbti et al., 1980; reproduced

witb permission

of Journal

3. The Particle Size and Molecular Weight Distributions

129

(2) with appropriate modification to allow for truncation at a sufficiently


large value of i (Ballard et al., 1981).Equation (41) is then solved numerically by expressing them in matrix terms:
(58)

oN'jot' = O'N'

where N' is the vector whose elements are N'l' N~, oo.,and the elements of
the matrix n' are defined in Eq. (41); the solution to Eq. (58) is
N'(t') =

exp(Ajt')Lej[RejN'(t'

= O)]

(59)

where the Aj, Lej and Rej are the eigenvalues, left (column) and right (row)
eigenvectors of n'. The initial conditions N'(t' = O)are determined from Eq.
(42) using the values of the Ni obtained as described above. Solution of Eq.
(59) thus requires a single numerical eigenvalue determination for a comparatively small matrix (e.g., a 10 x 10 matrix for = 3). Sd' SIn and S.x are
then found from Eqs. (43-45). The next step is the evaluation of the N7.
These are obtained by writing Eq. (48) in matrix form, as in Eq. (58),
carrying out the eigenvalue evaluation, and then determining Ni from the
equivalent of Eq. (59). The initial conditions N7(t', t" = O) are found from
Eq. (49) using the results obtained in N;(t') computation. Specifically, we
expand Eq. (59) to give

N;(t')=

Ij a;j exp(Ajt')

where

Im Le;m(pNm-1

a;j = Re;j

+ mfNm)

Re;j and Le;m'being tbe ith components of the jth right and mth left
eigenvectors of n' (corresponding to Aj), and Ni being the above-mentioned
steady-state solutions of the Smith-Ewart equations. Writing Eq. (48) in
matrix notation, we have
oN"jot" = n"N"
whose solution is written using the eigenvalue expansion as
N;'(t', t") =

Ij a;j exp(Ait")

(60)

the Aj being the eigenvalues of n". The a;j depend on t', and one finds that
Eq. (60) may be rewritten as
N;'(t', t") =

Lj LI bijl exp().;t'

+ Ait")

Gottfried Lichti el al.

130

where
bjl

= Rej L
Le;;'lpa.m-l
m

+ (m - l)fam]

from Eq. (49). Finally, substituting into Eq. (50),one finds


Sbc(t., M) = 2cc

L LL {bjl[exp(tAjt') -

;>2

exp(l.t')]/[Aj- 2A]}

Some typical resu"lts found using the above formulas are shown in Fig.
11, where the MDS from a combination-dominated high system are
shown (=1.72, obtained with p=0.5sec-1,
cc=O.lsec-l,
k=Cd=
f = O) and, for comparison, a combination-dominated low system
( = 0.51,using p = 0.5 sec-1, Cc= 50 sec-l, k = Cd= f = O).Note that the
MWDs have been normalized, and for cIarity of graphical presentation that
for = 0.51 has been re-scaled by a factor of t. The pronounced difference
between Sbc for

= 0.51

and 1.72 is a dramatic

illustration

of the effects of

compartmentalization, a point amplified in the following section. Note too


that the polydispersity

ratio P for:

= 0.51

is greater than for

= 1.72.

This

0.04

0.51 )

0.03

'"
I

(=1.72)

~ 0.02

<11
'"

o
~
~
0.01

o
O

5
GROWTHTIME (sec)

10

Fig. 11.
Comparison of instantaneous MWD for combination-dominated systems
(where S = Sbc) with = 0.51 and 1.72, assuming p = 0.5 sec-I, k = Cd= f = O, with
Cc= 50 sec -1 and 0.1 sec -1, respectively.

131

3. The Particle Size and Molecular Weight Distributions


occurs even though the MWD for
broader than that for

= 0.51.

= 1.72

appears from Fig. 11 to be

This highlights the fact that Pis a measure

of relative, rather than absolute, broadness.


3. Polydispersity Ratios
It will be recalled that the polydispersity ratio P, defined in Eq. (35),
gives a compact description of the MWD. It is instructive to compare
values of P computed from bulk and solution MWD theory with equivalent
values of P for a compartmentalized system, Le., from the theory developed
in Sections n,D,l and 2.
As stated in Section n,B,l, for bulk or solution polymerization, Pis 2 for
a system dominated by transfer or disproportionation but 1.5 for domination by combination. For compartmentalized systems P may be computed analytically from the expressions given in Sections n,D,l and 2.
For the transfer- and exit-dominated system shown in Fig. 8, it is found
that P = 2.0. This result shows that compartmentalization does not alfect
the MWD of polymer in transfer-dominated systems, vis-a-vis their bulk/solution counterpart. Briefly, the reason for this is that the chain-stopping
mechanism does not involve free radicals other than those on the growing
chain, and so compartmentalization has no elfect.
For the combination-dominated compartmentalized system shown in
Fig. 9, the polydispersity ratio is found to be 1.9. Here it can be seen that
compartmentalization has a significant elfect, this value for P being appreciably greater than the bulk value of 1.5. Thus, for compartmentalized
combination-dominated systems the MWD is significantly broader than in
the equivalent bulk system.
In the disproportionation-dominated system of Fig. 10, P = 2.2, showing
again a significant broadening

compared

with the bulk value of P

= 2.

It is particularly informative to compute P as a function of for


combination-dominated and disproportionation-dominated
systems (note
incidentally that 0.5 is the lower limit of in such a system). It will be seen

. that

when ;(; 2, the compartmentalized

system is indeed

a "minibulk"

system, with compartmentalization elfects becoming more pronounced as


becomes smaller. The P versus results shown in Figs. 5 and 6 were
obtained by fixing p and systematically varying either Cc(Fig. 12) or Cd(Fig.
13). The values of P were obtained from the MWD expressions given in
Sections n,D,l (valid for < 0.7) and n,D,2 (valid for any ).
Figure 12 shows P versus for a combination-dominated system,
obtained

by varying Cc and with p

= 1 sec-1.

This type of plot was first

presented by Katz et al. (1969), whose work in this area will be reviewed in
Section n,E,2. We show P computed both from the zero-one-two analytic

132

Gottfried Lichti et al.


2.0
+
+
+

"c:
~
v 1.5
'"'-

---

----

"~

~
v

1.0
0.5

1.0

1.5

2.0

Fig. 12.
Plot of polydispersity ratio P = (Mw>/(Mn>, as function of for system where
termination is by combination alone. + zero-one-two system; - complete analysis.
p = l sec - 1, k = 1 = Cd = O,CC varied.

4.0

3.0

2.0

1.0
0.5

--------------

1.0

1.5

2.0

Fig. 13.
Plot of polydispersity ratio, P = (Mw>/(Mn>, versus for system where
termination is by disproportionation alone. + zero-one-two system; - complete analysis.
p = I sec-I, k =1= '.c = O,Cdvaried.

133

3. The Particle Size and Molecular Weight Distributions

formulas and from seminumerical formulas of Section 11,0,2, so as to


show the validity range of the former (i.e., < 0.7). Figure 12 shows that in
the totally compartmentalized limit ( = 0.5), P approaches 2, as shown
above for the MWO of Fig. 9. As foreshadowed above, the transition from
P = 2 (compartmentalized limit, for small ) to P = 1.5 (bulk limit, for large
) occurs in the range 0.5 ;5 ;5 2.
Figure 13 shows the analogous P versus plot for a disproportionationdominated system, found by putting p = 1 sec-1, k = f = Cc= O,and letting
Cdvary. In the totally compartmentalized limit ( = 0.5),P = 4. This can be
understood as follows. When = 0.5, only state Oand state 1 particles are
present: the entry of a second free radical causes instantaneous termination.
This termination leaves a contribution to the MWO with P = 2, as in the
combinatioil-dominated system. However, with a disproportionation mechanism operative, the entrant free radical leaves in addition a low moleeular weight species for each longer dead chain formed. These two contributions to the MWO add, to halve the number-averaged molecular weight,
leaving the weight-averaged molecular weight essentially unaffected. Thus,
the overall P changes from 2 to 4.
The P versus curves computed in Figs. 12 and 13 were for systems
where no desorption occurred (k = O)and thus gave a lower limit of 0.5 to
. By choosing a nonzero value for k, values of anywhere between O
and 00 can be obtained. Figure 14 shows the P versus curve obtained
4.0

A
;;;

3.0

~
v
""A
~

~V

-------

2.0'~-------

1.0

0.5

i'i

1.0

1.5

Fig. 14. Plot of polydispersityratio P = <Mw>I<Mn>versus for system where chain


stoppage is by exit and disproportionation. k = 10-3 sec.l, Cd= 10-1 sec-I, f= Ce= O, O
varied.

Gottfried Lichti et al.

134

for a system showing disproportionation and exit: k = 10- 3 sec-1,


Cd= 10-1 sec-1,f = O= cc, where p is varied systematically. It is seen that
P passes through a distinct maximum at = 0.5 (for k = O,the maximum
value of P is 4, as in Fig. 13). For ~ 0.5, the system is exit-dominated
(P ~ 2), whereas for ~ 0.5, the bulk value of P is equal to 2 for a
disproportionation-dominated system.
E.

Alternative Forinulations 01 the MWD

We have reviewed (above) a complete theory of the MWD in an


emulsion polymerization (in the absence of long-chain branching). We now
relate this to alternative formulations.
1. Zero-One Systems
In these systems (Section I,C,l,a) Gerrens (1959) recognized that if
combination is the only chain-stopping event, then the number-average
molecular weight (Mn)is inversely proportional to the entry rate coefficient:
Mn IXp-l

(61)

This result may indeed be proved by taking the appropriate limits (k = O,


f = O, p ~ e) to the results discussed in Section n,D,1. Gardon (1968a)
extended this concept to make allowances for p varying as a function of the
fraction conversion of monomer to polymer in Interval n (arising from the
increase in particle size). Gardon however concluded incorrectly that, in
such a system, 1 < p < 1.33. Gardon's error lay in his assumption that the
instantaneous MWD is perfectly monodisperse (P = 1),Le., that all chains
formed at a given instant have the same molecular weight, this being given
by Eq. (61). In fact, this assumption does not hold, and P has, for example,
a value in the range 1.5 ~ P ~ 2 for domination by combination. Since P at
any instant exceeds 1.5, the overall P must also exceed 1.5.
Lin and Chiu (1979) developed a theory for the MWD in a zero-one
system that correctly predicts (where M is continuous rather than discrete)
the free-radicallifetime distribution N'1 leading to a value of 2 for P. The
formalism of Lin and Chiu involves calculating the number of chains
containing m monomer units. Their final results involve sequence summations over the range 1 ~ m ~ oo.
2.

Formulaton of Katz et al.

The first comprehensive theory of the MWD in an emulsion polymerization system was given by Katz et al. (1969). These authors considered a
monodisperse system in which the only mechanisms operating were freeradical entry, chain propagation, and bimolecular termination by combination. They defined a distribution function Pi(te,MI, M2, ..., Mi), as the

135

3. The Particle Size and Molecular Weight Distributions

relative number of particles in state i at experimental time te containing one


growing chain of molecular weight MI, another of M2, and so on, with
MI > M2. The equations that describe the evolution of the p are

fJPifJt.= -a

j<'

+ 2c

fJP,lfJMj
- [p + i(i -1)c]P, + pP'-1t5M;.o

I I

'<j~'+2

ff

p,+2(MI, oo.,M" .oo,Mj, ..., M'+2) dMj dM,

(62)

Here a is defined in Eq. (30). The Kronecker t5 term gives the source terms
for each P,. These coupled partial integrodifferential equations in the
variables t, MI, oo.,Mn were derived by population balance applied to each
microscopic process. Since bimolecular termination by combination is the
only chain-stopping mechanism considered, Sbc is the only contribution to
the instantaneous MWD; this is given by

f ~j~

Sbc(t., M) = 2ccP~

x t5(Mj + M,

P,(te, MI, oo.,M}

(63)

M) dMI, oo.,dM,

Here t5 is the Dirac t5 function and p is a normalizing factor. The inner


summation in Eq. (63) accounts for combination between all possible pairs
of free radicals. Equation (63)was derivedassumingthat the polymerization
is occurring in the steady-state regime.
Because of the complexityof these equations, Sbcmust be evaluated by
numerical means. Katz et al. used the method of moments to generate a
measure of the MWD called the cumulative distribution of the tail, defined
. as

F(M)

f: Sbc(M)dM

(64)

The development of Katz et al. properly encompasses all compartmeIltalization effects in a system where combination is the only chain-stopping
mechanism. It has been shown (Lichti et al., 1980) that the equations in Eq.
(62) are equivalent to the evolution equations for N; and N;' given in
Section n,e,

with the restriction

that k

= f = Cd = O. The

P, and N;(t., t')

and N;'(t., t', t") are related by


N;(t.

Mrla, Mrla) =

j~.i P,(t., MI, ..., Mj-I'

M.. Mj+l' oo.,M,)

dMI, oo.,dMj-1dMj+l' oo.,dM,


N;'(t. - Mrla, (Mr - Ms)la,Msla) =

fI L

i~ 1

(65)

P,(t., MI, .oo,Mj-I'

l<j~i

M.. Mj+l'
dMj+l'

oo.,M'-I,

oo.,dM'-1

M.. MI+I,
dM,+I'

oo.,M,) dMI, .oo,dMj-1

.oo,dM,

i ~ 2

(66)

136

Gottfried Lichti el al.

Plots of the type shown in Fig. 12 (P versus ) were first given by Katz et al.
Their treatment, however, suffers from a number of limitations. These are (i)
the numerical labor involved in generating the moment expansion, (ii) the
imperfect resolution involved in generating the cumulative distribution [Eq.
(64)] rather than the fuIl distribution, and (iii) the neglect of exit, transfer,
and bimolecular combination by disproportionation. Point (ii) may be
iIIustrated as foIlows. Consider a combination-dominated system in two
limits: the fuIly compartmentalized case of = 0.5, when one has
Sbcex exp(-ca'), and the bulk case, where = 00, when one has
Sbc ex t exp( - at'), where a is a parameter proportional to cc' Figure 15

e
~
~

z
o
::

=>
m
ir
1(J)

e
w
>

~
=>

=>
(,)

GROWTH TIME

Fig. 15.

Plot

combination-dominated

curve 2 = oo.

of instantaneous

MWD

system, as functions

(S) and

cumulative

of growth time (arbitrary

distribution

(F) for

units). Curve 1

= 0.5;

3. The Particle Size and Molecular Weight Distributions

137

shows plots of Sbcand F [from Eq. (64)] for these two limits, with a = 2. It
can be seen that plots of the MWD, Sbc' show a pronounced qualitative
ditTerence between the compartmentalized and bulk limits. However, the
equivalent plots of F, the cumulative distribution of the tail, are qualitatively similar, and only show smalI quantitative ditTerences.This is because
the greatest ditTerencesbetween the Sbcfor the bulk and compartmentalized
limits occurs at smalI values of M (or t'), to which F(M) is insensitive.

3.

Formulation of Min and Ray

Min and Ray (1974) have presented a theory in which the MWD and
PSD evolutions are considered together. The proposition that these quantities may be evaluated separately is considered in Section III of ~his
chapter.
The treatment of Min and Ray is applicable to a system which is
mohodisperse in volume if one omits both particle nucleation terms and
terms involving O/OV,and if the volume dependences of any rate coefficients
are ignored. For such a system, Min and Ray define a function fm(i, te), the
distribution of latex particles in state i at experimental time te that contain a
growing radical with m monomer units. This is directIy related to the singly
distinguished distribution function Ni(te, t') by

Ni(te

t', t') = fm(i, te)

(67)

where t' = mMo/a, where Mo is the molecular weight of monomer. The


evolution of fm(i, te) is given by simple coupled ditTerential equations in te,
one for each m. There would typicalIy be thousands of such equations; this
is to be compared wi'th the sma:lI number of simple coupled ditTerential
equations required to evaluate Ni.
The formulation of Min and Ray alIows for exit, in the form of
permitting desorption of growing chains of any length, rather than the exit
of smalI oligomers alone permitted in the general formulation given here.
This assumption of Min and Ray does not relate desorption to any chaintransfer evento If growing chains of any length could d~sorb, it would seem
as likely that dead chains would also desorb; this possibility is however not
encompassed within Min and Ray's formalismo Their formalism in addition
has no analog of the doubly distinguished distribution functions Ni.
Instead they approximate the contribution to the MWD by combination by

a convolution expression involving products of the fm and indeed acknowledge that this assumption is open to question. It was shown in Section
II,C,2 by use of a particular example that such a product approximation is

138

Gottfried Lichti et al.

invalid. We now give a more specific illustration. Consider the correct form
for Sbcin a zero-one-two system:

it'

Sbe(te' t' ) =

"
N"2e'
(t t' - 2t" , t" ) dt

(68)

with the equivalent convolution expression of Min and Ray, which we


denote Sbc:

sbe(t e' tI) =

tO

pio

NI 2e'
(t

tI

t" ) N' 2e'


(t

t") dt"

Here the normalization constant pis. chosen so that


Figure 16 shows Sbc and Sbc for p

~~

JSbcdt' = J

SbC

dt'.

1 sec-1, Cc = 10 sec-1, k = f = Cd = O

(giving = 0.55). It is clear that the product assumption is both quantitativeiy and qualitatively inaccurate. Indeed, the limiting
,
.value of the

NI
o
~
~

o
31
~

EQUIVALENT GROWTH TIME (sed


Fig. 16.
Comparison of predictions of the formalism of (1) Min and Ray (1974) with
that of (2) Lichti et al. (1980) for the MWD of a combination-dominated system assuming
p = 1 sec -1, C, = 10 sec -1 and k = f = Cd = O. Ordinate
cissa is growth time (sec).

is instantaneous

MWD

(sec - 2), ab-

3. The Particle Size and Molecular Weight Distributions

polydispersi"ty ratio P as approaches 0.5 is P


correct value of P = 2.
F.

= 2.5 for Sbc, rather

139
than the

Experimental Determination 01 the MWD

We now briefly consider experimental methods of determining the


MWD, in order both to test and to apply the theory. In relating theory and
experiment, it is clearly insufficient to consider a single measure of the
MWD alone, e.g., the number average, since interpretation of a single
datum is prone to ambiguities. Hence, experimental techniques that only
give such a datum (e.g., light scattering) are inadequate for the present
purpose. Even the polydispersity P, being the weight-to-number-average
ratio, is ambiguous unless it can be established unequivocally that the
MWD is monomodal, and even then different-shaped MWD curves can
give similar values of P. It is clear that a measure of the full molecular
weight variation of the MWD is necessary to pro vide useful information.
The best procedure at present for obtaining this appears to be gel permeation chromatography (GPC). As this is relatively new technique, full
MWD data are sparse.
Friis and Hamielec (1975) have used GPC to study the MWD development in vinyl acetate and vinyl chloride emulsion polymerizations. For
these monomers, the main chain-stopping mechanism is thought to be
transfer, and so the compartmentalized nature of the system is relatively
unimportant. These workers found that the MWDs produced at early
times, where branching reactions are unimportant, have a P value close to
2, as expected for transfer-dominated reactions.
A careful experimental study of the MWD produced in a highly compartmentalized system has been carried out by James and Piirma (1976) and
Piirma et al. (1975). Ah ab initio emulsion polymerization was used which
produced particles small enough ('" 75 nm unswollen radius) to ensure low
values in the Interval 11 part of the reaction. P was found to be constant
during Interval 11, with a value somewhat in excess of 2. The fact that P
(which as measured experimentally is of course from the cumulative rather
than the instantaneous MWD) remains constant implies that the P of the
instantaneous MWD is also close to 2 throughout this interval. This value
of P may be rationalized on the assumption that either transfer or combination (in the = 0.5, i.e., compartmentalized limit) are the main chainstopping mechanisms. Piirma et al. showed that in fact transfer to monomer
was dominant, since the number-average molecular weight observed corresponded closely to that predicted from transfer domination. This result is
interesting, in view of the fact that the main chain-stopping mechanism for
styrene solution or bulk polymerization is combination. This result is thus a

140

Gottfried Lichti et al.

clear example of how compartmentalization can change the dominant


chain-stopping mechanism. This is because free radicals inside the growing
latex particles polymerize in isolation for periods of time sufficiently long to
make the MWD transfer limited.
The work of James and Piirma and of Piirma et al. is important because,
inter alia, it highlights the role of several commonly used surfactants (e.g.,
Triton X) as transfer agents. This discovery complicates the interpretation
of many experimental results reported in the literature. Included in this
category is the rise in molecular weight with conversion in Interval 11,used
by Grancio and Williams (1970) as evidence for the core-shell model of
latex particle morphology.
Lin and Chiu (1979) have reported measurements of P as a function of
time for an ab initio styrene emulsion polymerization. They also found that
P was slightly in excess of 2 for a considerable part of the reaction, but their
actual molecular weight averages were much lower than those of Piirma et
al. (1975). This suggests that their surfactant may have been acting as a
chain-transfer agent.
De Graaf and Poehlein (1971)'have measured the MWD of styrene in a
continuous stirred-tank reactor. A wide range of particle sizes were present,
and thus interpretation of the data is hindered by an inadequate knowledge
of the size dependence of the various rate parameters. However, the values
reported for the average molecular weights suggest again a transferdominated process; this is also consistent with the value obtained for P
which was close to 2, for a variety of conditions.
The influence of particle size on MWD has been investigated by Morton
et al. (1954). They showed that the average molecular weight of the polymer
produced was insensitive to the latex particle size. This is consistent with
the molecular weight being dominated by chain transfer to monomer, a
conclusion that holds irrespective of any differential swelling of particles of
different sizes.
The foregoing experimental review includes the major studies in which
polydispersity ratios (as distinct from average molecular weights) were
reported. The experiments are somewhat piecemeal in nature, dealing with
systems for which reliable estimates of the rate coefficients which govern the
MWD (p, etc.) were not available. It would clearly be advantageous, in
terms of mechanistic understanding, to carry out MWD measurements on
systems that are well characterized kinetically. The simplest experiment to
interpret would be on a seeded system (thereby obviating problems as to the
MWD formed during particle nucleation, as will be shown later), employing
a volume monodisperse latex with polymerization being carried out in
lnterval 11 under steady-state conditions and without significant particle
growth during the course of polymerization. Methods have been established

3. The Particle Size and Molecular Weight Distributions

141

for obtaining the various rate coefficients klP p, k, kp, and e from kinetic
studies (e.g., Hawkett et al., 1980, 1981). The theoretical evaluation of the
MWD is then trivial. For systems with a significant polydispersity in
volume, the MWD can readily be calculated using Eq. (27), given the
volume dependence of the various rate coefficients; however, it was pointed
out above that uncertainties in these quantities make an unambiguous
relation between theory and experiment impossible in such an experiment.
Experiments on a volume-monodisperse Interval III system could again be
readily compared with theory, using known values of the time variation of
CM, etc. (obtained from the kinetics), and integrating the instantaneous
MWD over time.
The ideal seeded experiment would be where seed Iatex particles are
prepared in which the MWD of the seed is significantly different from that
formed during subsequent seeded polymerization. The instantaneous MWD
formed at early times could then be estimated directly. Seed latexes of
different sizes could be used to probe the effect of compartmentalization (we
stress again that if the system is transfer-dominated, the MWD of formed
polymer should be independent of particle size).
The experimental strategy outlined above would give a unified set of rate
parameters to describe the kinetics, MWD, and PSD. If no such consistent
set could be found, new mechanisms may pro ve to be important. For
example, aqueous-phase termination, which is not usually incorporated in
MWD theories, would give rise to extensive low molecular weight fragments, which would significantly influence the MWD if these fragments are
subsequently incorporated into the latex p~rticles.

111. Separa~iIity of 1YIWDand PSD


Most of the theory surveyed in this chapter is based on the premise that,
although the MWD and PSD of a system are governed by the same
fundamental set of rate coefficients (p, klP kp, etc.), it is nevertheless possible
to decouple the computation of these quantities. This is required, for
example, for the validity of Eq. (27). The basis of this premise 's the
separation of time scales over which the various proc~sses governing these
quantities occur. This concept was first introduced by O'Toole (1969). We
now consider its validity.
It is frequently the case that the time required for a Iatex particle to
increase its volume by a measurable amount (say by 10%) is large compared
with (i) the time taken for the growth of a polymer chain and (ii) the time
taken for the distribution of free radicals inside the latex particles to reach a
steady-state value (for a given particle volume). This separation of time

142

Gottfried Lichti el al.

scales is plainly invalid at very early reaction times, when most of the
volume of the partic1e may be that of a single chain. However, since latex
partic1es usually contain many individual chains, the time scale separability
assumption for volume versus chain growth [i.e., (i) above] is c1early
relevant to mO$tof the growth process.
We now consider the second time scale separability, (ii) above. Again,
this will obviously break down at early times, but again will be valid for
most of the polymeriz.ation. This occurs for three reasons.
1. As partic1es increase in volume, a fixed percentage increment of
growth takes longer to achieve.
2. With increasing time, the system will attain and maintain its steady
state in the distribution of free radical s in all partic1es, with the possible
exception of growth during a particularly rapid Trommsdorff gel effect.
3. The PSD becomes smoother with the passage of time (e.g., Fig. 1),so
that partic1e growth [which disturbs steady state through the o/ay term in
Eq. (8)] becomes less significant.
.
Since the separate time scale aondition is c1early valid for most of the
polymerization process, one may say that each polymer chain is formed
inside a partic1e of unchanging size, wherein all rate coefficients are constant
and the distribution of free radicals has its steady-state value, for each
volume V. Any residual effect of the PSD on the MWD would reside
presumably in the effects of the PSD on the kinetic parameters (e.g., p, e,
and to a lesser extent k). Conversely, the MWD would possibly influence
the PSD through its effects on the swelling of the partic1es by the monomer;
the effect, if it exists, is likely to be small.

IV. Conclusions
The theoretical tools required to describe the evolution of the MWD and
PSD of an emulsion polymerization, in terms of a consistent set of rate
coefficients for the various microscopic processes involved, are fairly well
developed. Whereas chain branching (transfer to polymer) has yet to be
inc1uded, it is likely that bulk theories for this process (Bamford and
Tompa, 1954) can be readily modified to the emulsion polymer case.
Although the general theory is well developed, a real understanding of
the details of the processes involved is limited by the paucity of unambiguous experimental results to enable theory to be tested, modified, and
applied.

143

3. The Particle Size and Molecular Weight Distributions


Acknowledgments

The support of the Australian Research Grants Committee, and of an Australian Institute
for Nuclear Science and Engineering postdoctoral fellowship for GL, are gratefully
acknowledged.

References
Ballard, M. J., Gilbert, R. G., and Napper, D. H. (1981). J. Polymer Sci. PoI. Phys. Edn.,
(in press).
.
Bamford,C. H., and Tompa, H. (1954).Trans. Faraday Soco SO, 1097-1115.
Camahan, B., Luther, H. A., and Wilkes, J. O. (1969). "Applied Numerical Methods." Wiley,
New York.
Corio, l., Mara, L., and Salvatore, O. (1979). Makromol. Chem. 180,2251-2252.
De Graaf, A. W., and Poehlein, G. W. (1971). J. Polym. Sci. Parl A-2 9, 1955-1976.
Eckhoff,R. K. (1967).J. Appl. Polym. Sci. 11, 1855-1861.
Ewart, R. H., and Carr, C. 1. (1954). J. Phys. Chem.58, 640-644.
Flory, P. J. (1953). "Principies ofPolymer Chemistry." Comell Univ. Press, 1thaca, New York.
Friis, N., and Hamielec,A. E. (1975).J. Appl. Polym. Sci. 19,97-113.
Friis, N., Goosney, D., Wright, J. D., and Hamielec, A. E. (1974). J. Appl. Polym. Sci. 18,
. 1247-1259.
Gardon, J. L. (l968a). J. Polym. Sci. Parl A-I 6, 665-685.
Gardon, J. L. (1968b). J. Polym. Sci. Parl A-I 6, 687-710.
Gerrens, H. (1959).Forlschr. Hochpolym.Forsch. 1, 234-328.
George, M. H. (1967). /n "Vinyl Polymerizations" (G. E. Ham, ed.), Vol. 1, part 1, p. 165.
Dekker, New York.
Grancio, M. R., and Williams, D. J. (1970). J. Polym. Sci. Parl A-I 8,2733-2745.
Harkins, W. D. (1945).J. Chem. Phys. 13, 381-382.
Harkins, W. D. (1946). J. Chem. Phys. 14,47-48. ,
Harkins, W. D. (1947). J. Am.Chem.Soco69, 1428-1444.
Harkins, W. D. (1950). J. Polym. Sci. 5, 217-251.
Hawkett, B. S., Gilbert, R. G., and Napper, D. H. (1980).J. Chem.SocoFaraday Trans. /76,

1323-1343.
.

Hawkett, B. S., Napper, D. H., and Gilbert, R. G. (1981). J. Chem. Soco Faraday Trans. /
(in press).
James, H. L., and Piirma, l. (1976)./n
eds.), pp. 197-209. American

"Emulsion

Polymerization"

(1. Piirma and J. L. Gardon,

Chemical

Society, Washington, D.C.


Katz, S., Shinnar, R., and Saidel, G. M. (\969). Adv. Chem. Ser. 91, 145-157.
Kerker, M. (1969). "The Scattering of Light." Academic Press, New York.

Lansdowne, S. W., Gilbert, R. G., Napper, D. H., and Sangster, D. F. (1980). J. Chem. Soco
Faraday Trans./76,
1344-1355.
Lichti, G., Gilbert,
1957-1971.

R. G., and Napper,

D. H. (1977). J. Polym. Sci. Polym. Chem. Ed. 15,

Lichti, G., Gilbert,


1297-1323.

R. G., and Napper,

D. H. (\980).

Lichti, G., Hawkett,

B. S., Gilbert,

R. G., Napper,

Sci. Polym. Chem. Ed. 19, 925-938.

J. Polym. Sci. Polym. Chem. Ed. 18,

D. H., and Sangster,

D. F. (1981). J. Polym.

144

Gottfried Lichti el al.

Lin, C. c., and Chiu, W. y. (1979). J. Appl. Polym. Sci. 23, 2049-2063.
Min, K. W., and Ray, W. H. (1974). J. Macromol. Sci. Rev. Macromol. Chem. C 11, 177-255.
Min, K. W., and Ray, W. H. (1978). J. Appl. Polym. Sci. 22, 89-112.
Morton, M., Kaizerman, S., and Altier, M. (1954). J. Colloid Sci. 9, 300-312.
Nagy, D. J., Silebi, C. A., and McHugh, A. J. (1980). In "Polymer Colloids II" (R. M. Fitch, ed.),
pp. 121-137. Plenum Press, New York.
O'Toole, J. T. (1969). J. Polym. Sci. Part C 27, 171-182.
Piirma, J. Kamath, V. R., and Morton, M. (1975). J. Polym. Sci. Polym. Ed. 13,2087-2102.
Pis'men, L. M., and Kuclianov, S. I. (1971). Vysokomol. Soedin. A13, 1055-1065.
Schmidt, E., and Biddison, P. H. (1960). Rubber Age 88, 484-490.
Schmidt, E., and Kelsey, R. H. (1951). Ind. Eng. Chem. 43, 406-413.
Shaw, D. J. (1970). "Introduction to Colloid and Surface Chemistry." Butterworths, London.
Singh, S., and Hamielec, A. E. (1978). J. Appl. Polym. Sci. 22, 577-584.
Smith, W. V., and Ewart, R. H. (1948). J. Chem. Phys. 16,592-599.
Stevens, J. D., and Funderburk, J. O. (1972). Ind. Eng. Chem. Process Res. Dev. 11, 360-369.
Sundberg, D. C., and Eliassen, J. D. (1971). In "Polymer Colloids" (R. M. Fitch, ed.), pp.
153-161. Plenum Press, New York.
Thompson, R. W., and Stevens, J. D. (1977). Chem. Eng. Sci. 32: 311-322:
Ugelstad, J., and Hansen, F. J. (1976). Rubber Chem. Tech. 49, 53Cr609.
Watterson, J. G., and Parts, A. G. (1971). Die Makromol Chem. 146, 11-20.
Wood, D., Lichti, G., Napper, D. H., ~nd Gilbert, R. G. (1981), to be published.

4
Theory 01 Kinetics
01 Compartmentalized
Free-Radical Polymerization
Reactions
D. C. Blackley

1. Introduction.
A. Definitionsand Introductory Concepts
B. Practical Significance of the Theory
C. Scope of Chapter .
11. Reaction Model Assumed
A. Description.
B. Constancy and Uniformity of Monomer

111.

IV.
V.

VI.
VII.
VIII.

Concentration within Reaction Loci


C. Mechanism
of Fundamental
Processes;
Influence of Locus Size ,
The Time-Dependent
Smith-Ewart Differential
Difference Equations; Methods Available for
Their Solution
A. Derivation .
B. Methods Available for Their Solution .
Solution for the Steady State
Solutions for the Nonsteady State .
A. Case in which Radical Loss 15 Predominantly
by First-Order Processes
B. Case in which Generation of New Radicals
Ceases .
C. Approximate .. Poissonian" Solution to the
General Case.
D. Other Approximations
Predictions for Molecular-Weight Distribution
and Locus-Size Distribution .
Theory for Generation of Radicals in Pairs
within Loci .
List of Symbols.
References .

146
146
147
149
149
149
151
153

156
156
160
164
167
167
172
176
177
183
185
187
189

145
EMULSJON POL YMERIZA TION
Copyright ~ 1982 by Academic Press, Inc.
AII rights 01 reproduction in any lorm reserved.
ISBN 0-12-556420-1

146

D. C. Blackley

l. Introduction
A.

Definitions and Introductory Concepts

By the term "compartmentalized free-radical polymerization


reaction" is meant a free-radical polymerization reaction that is taking
place in a large number of separa te reaction loci. These loci are dispersed in
a contiguous external phase. They are "separate" from each other in the
sense that material contained within one particular locus is presumed to be
capable of transferring to another locus only insofar as it is capable of being
lost from the first locus to the external phase, and then of being subsequentIy absorbed by the second locus from the external phase. The
number of separate reaction loci is presumed to be "large" in the sense that
it is at least of the same order of magnitude as the number of propagating
free radicals present within the reaction system as' a whole. It therefore
follows that the average number of radicals per single reaction locus is
small; it is convenient to regard 10 as the absolute upper limit, 5 as the
usual upper limit, and 0.01-2 'as the range that covers many reaction
systems of practical significance. The term "compartmentalization"
as
applied to free-radical polymerization reactions of the type considered here
seems to have first been introduced by Haward (1949).
The mechanism of the polymerization reaction is presumed to be essentially that of a homogeneous bulk or solution free-radical polymerization.
The concern is exdusively with the polymerization by double-bond opening
of carbon compounds that contain at least one carbon-carbon double
bond. The reactive species that propagates to produce the polymer chain is
a free radical formed by opening of the n-bond of the carbon-carbon double
bond. The basic steps of the polymerization reaction are initiation, propagation, termination (by various means), and various transfer reactions.
The structure of the polymer produced is determined by the balance of the
propagation, termination, and transfer reactions.
Whereas the polymerization mechanism is essentially that of homogeneous bulk or solution polym~rization, there are significant differences
attributable to compartmentalization as the term has been defined above.
The most important of these is that, unlike free-radical polymerizations
taking place in a homogeneous medium, there are physical barriers that
prevent interaction between the various propagating radicals present in the
reaction system at a given time. To a large extent, the propagating radicals
are physically isolated from each other, each very small group of radicals
being provided with its own reaction vessel. This is, in fact, the significance
of the adjective "compartmentalized." An important consequence of the
compartmentalization of the propagating radicals is that opportunities for

4.

Kinetics of Compartmentalized

Free-Radical Polymerization Reactions

147

the mutual termination of radicals are reduced relative to the case of a


similar polymerization occurring in a homogeneous medium at the same
overall concentration of propagating radicals within the reaction system as
a whole.
The model for the reaction system will be considered in detail in Section
11.However, it is convenient to note here that, in principie, the free radical s
that initiate the polymerization may be generated either within the external
phase (external initiation) or within the reaction loci themselves (internal
initiation). Whereas very brief reference will be made at the conclusion of
this chapter to reaction systems of the latter type, the concern here will be
almost exclusively with reaction systems of the former type. Insofar as the
initiating radicals are generated exclusively within the external phase (and
therefore ha ve to be by some means acquired by the loci by absorption
from the external phase), we have a further important distinction between
homogeneous and compartmentalized reactions. In the latter case, the
processes that lead to the generation of the initiating radical s are physically
isolated from the propagation, termination, and transfer reactions. One
minor consequence of this is that transfer-to-initiator reactions may be
virtually eliminated in the latter case.
The primary objective of the theory of compartmentalized free-radical
polymerization reactions is to predict from the physicochemical parameters
of the reaction system the nature of the "Iocus population distribution." By
this latter term is meant collectively the proportions of the total population
of reaction loci which at any in~tant contain O,1,2,. ", i,... propagating
radicals. The theory is concerned with the prediction of these actual
populations and also with such characteristics of the locus population
distribution as the average number of propagating radical s per reaction
locus and th~ varian~e of the distribution of locus populations.
.,

B.

Practical Significance of the Theory

Apart from intrinsic interest, the theory of compartmentalized freeradical polymerization reactions is of importance primarily because it is
believed that ITIost of the polymer which is formed in the course of an
emulsion polymerization reaction is formed via reactions of this type. The
general shape of the conversion-time curve for many emulsion polymerization reactions suggests (see Fig. 1) that the reaction occurs in three more-orless distinct stages or "intervals." The first of these, the so-called Interval 1,
is interpreted as the stage of polymerization in which the discrete reaction
loci are formed. In the second and third stages - Intervals 11 and 111- the
polymerization is believed to occur essentially by compartmentalized freeradical polymerization within the loci which were formed during Interval l.

148

D. C. Blackley

c.
o

.~
Q)

>
e
o
<>

INTERVAL m

time

Fig.1.

Conversion-time

curve for typical unseeded emulsion

polymerization

reaction.

The feature that distinguishes IntervallI from Interval 111is that monomer
droplets are present as a separate phase during the former only.
The theory also has relevance to the so-called "seeded "emulsion polymerization reactions. In these reactions, polymerization is initiated in the
presence of a "seed" latex under conditions such that new particles are
unlikely to formo The loci for the compartmentalized free-radical polymerization that occurs are therefore provided principally by the particles of the
initial seed latex. Such reactions are of interest for the preparation of latices
whose particles have, for instance, a "core-shell" structure. They are also of
great interest for investigating the fundamentals of compartmentalized freeradical polymerization processes. In this latter connection it is important to
note that, in principIe, measurements of conversion as a function of time
during nonsteady-state polymerizations in seeded systems offer the possibility of access to certain fundamental properties of reaction systems not
otherwise available. As in the case of free-radical polymerization reactions
that occur in homogeneous media, investigation of the reaction during the
nonsteady state can provide information of a fundamental nature not
available through measurements made on the same reaction system in the
steady state.

4. Kineticsof Compartmentalized Free-Radical PolymerizationReactions 149.


C.

Scope 01 Chapter

Detailed consideration will first be given to the nature of the reaction


model assumed. Such consideration is important in order to gain an appreciation of the limitations of the theory that is subsequently developed.
This will be followed by derivation of the equations that govern the behavior
of compartmentalized free-radical polymerization reactions and a discussion ofmethods that are available for their solution. This will be followed by
brief consideration of the solution of these equations for reaction systems
that have attained a steady state. This part of the subject is not recent and
will not be dealt with in detail. The main part of the chapter will not be
concerned with the solution of the equations for reaction systems in the
steady state but with those that are approaching it. This aspect of the subject
is of relatively recent development. Brief reference will then be made of the
evolution of the particle-size and molecular weight distribution during a
compartmentalized free-radical polymerization; although this subject is of
recent development, it will be dealt with in detail elsewhere in this book. The
chapter will conclude with brief reference to reaction systems in which the
initiating radicals are generated in pairs within the reaction loci.

11. Reaction Model Assumed


A.

Description

The number of reaction loci is assumed not to vary with time. No


nucleation of new reaction loci occurs as polymerization proceeds, and the
number of loci is not reduced by processes such as particle agglomeration.
The monomer is assumed to be only sparingly soluble in the external phase
(a typical example is styrene as monomer and water as the external phase),
and thus polymerization is assumed to occur exclusively withip the reaction
loci and not within the external phase. The monomer is assumed to be
present in sufficient quantity throughout the reaction to ensure that monomer droplets are present as a separa te phase, and the rate of transfer of
monomer to the reaction loci from the droplets is .assumed to be rapid
relative to the rate of consumption of monomer in the loci by polymerization. The monomer concentration within the rea<,:tionloci is then taken to
be constant throughout the reaction. This assumption is important if an
attempt is made to relate the overall rate of polymerization to the average
number of propagating radicals per reaction locus. The assumption will
therefore be examined in further detail below.

150

D. C. Blackley

In general, radicals are presumed to be generated in the external phase at


a constant rate and to enter the reaction loci at a constant rate. Thus, the
rate of acquisition of radicals by a single locus is kinetically of zero order
with respect to the concentration of radicals within the locus. The principal
concern is with reaction systems that initially contain no free radicals, and
therefore in which the rate of polymerization is zero. Then, at a certain
instant, which is taken as the zero of subsequent time, radical s begin to be
generated in the external phase at a constant rateo It is then required to
predict the behavior of the reaction system as it approaches the steady rateo
Important exceptions to these generalizations include the following:
(i) where the interest is primarily with the steady state rather than with the
approach to the steady state; (ii) where the rate of generation of new
radicals varies with time, e.g., because of initiator depletion; (iii) where the
rate of generation of new radicals in a polymerizing system is suddenly
reduced to zero and the interest is in predicting the subsequent decay of the
reaction; and (iv) where new radical s are generated in pairs within the
reaction loci.
Once a radical enters a reaction locus, it is presumed to initiate a chain
polymerization reaction which then continues at a constant rate until the
activity of the radical is lost. The processes whereby the activity of the
propagating radicals is lost from the reaction loci can be classified into two
broad types:
1. Processes that are kinetically of first order with respect to the
concentration of radicals within the reaction locus. These processes include
exit from the locus into the external phase, termination by reaction with
monomer within the locus, termination by reaction with adventitious
impurities in the locus, and spontaneous deactivation.
2. Processes that are kinetically of second order with respect to the
concentration of radicals within the reaction locus. The most important of
these processes is bimolecular mutual termination between pairs of propagating radicals within the same reaction lo~us.
In attempting to analyze the behavior of reaction systems in the nonsteady sta te, it is generally necessary to make the assumption that radicals
lost from the reaction loci by exit to the external phase are not able
subsequently to reenter the reaction loci and reinitiate polymerization. They
are not, therefore, regarded as being added to the "bank" of radicals in the
external phase available for entry into reaction loci. This is undoubtedly a
restrictive assumption that cannot be justified rationally. It is unfortunately

an assumption that often has to be made in order for the mathematics


describing the behavior of the reaction system to remain tractable.
Although some progress has been made in removing this restriction, it

4.

Kinetics of Compartmentalized Free-Radical Polymerization Reactions 151

nevertheless remains a general feature of the theory developed for nonsteady state reactions.
A further general assumption is that the physicochemical parameters
that characterize the rates at which the various processes occur remain
constant throughout the reaction. However, as will appe~r subsequently,
there is one particular instance where it has been possible to remove this
restriction. Important consequences of this restriction are that the polymerj
monomer ratio in the reaction loci should not vary significantly as the
reaction proceeds, and that the volume of the reaction loci should not
change greatIy as a consequence of polymerization. The latter requirement
will be most obviously realized in a seed emulsion polYp1erization in which
the size of the initial seed is large and the extent of polymerization is small.
B.

Constancy and Uniformity of Monomer Concentration


within Reaction Loci

There has been much discussion in recent years concerning the question
of whether or not, in a reaction system of the type considered in this
chapter, the concentration of monomer in the reaction loci is constant as
the reaction proceeds and uniform across the reaction locus. Morton et al.
(1954) made the reasonable proposal that the equilibrium swelling of
polymer partic1es by excess monomer in a system such as is envisaged here
is determined by the balance of two factors: (i) the decrease of free energy
that accompanies mixing of polymer and monomer and (ii) the increase of
interfacial free energy that accompanies the concomitant swelling of the
polymer partic1es. Theoretical considerations predict that at equilibrium
large partic1es should imbibe more monomer than small partic1es, and that
reduction of the interfacial free energy between polymer partic1e and
external phase shoulcl lead to increased swelling. Polymer molecular weight
is not a significant factor in determining equilibrium imbibition, but the
polymer-monomer solution interaction parameter is. Morton et al. (1954)
and Allen (1958) have provided experimental evidence to support these
predictions, the polymers used being polystyrene and natural rubber, respectively.Further information has been provided by Vanzo et al. (1965)
concerning the solubilities of vinyl acetate, vinyl hexanoate, styrene, and
benzene in poly(vinyl ,acetate), poly(vinyl hexanoate), and polystyrene
latices. Morton et al. (1954) also provided evidencein support ofthe view
that, as long as excess monomer is present as a separate droplet phase, the
monomer-swollen polymer partic1es remain essentially in the same state of
equilibrium imbibition during polymerization as in the absence of polymerization. This implies that the process of diffusion from droplets to polymerization loci occurs rapidly relative to the rate of polymerization, so that the

152

D. C. Blackley

consumption of monomer by polymerization has little effect on the concentration of monomer at the polymerization loci. A further implication is
that the concentration of monomer is uniform throughout the polymerization locus.
The views summarized in the preceding paragraph were generally accepted without serious challenge for reaction systems for which monomer and
polymer are miscible, until Grancio and Williams (1970) proposed that the
monomer-swollen polymer particles in a styrene emulsion polymerization
are nonuniform, and that in fact they consist of an expanding polymer-rich
"core" surrounded by a monomer-rich "shell," the latter being the major
locus of polymerization. The "core-shell" model was suggested by kinetic
evidence and supported by evidence obtained by examining ultra-thin
sections of latex particles by electro n microscopy. These ideas have been
further elaborated by Williams (1971), by Keusch and Williams (1973), and
by Keusch et al. (1973).Friis and Hamielec(1973)have challengedGrancio
and William's interpretation of their kinetic data and have suggested that
the observations can be explained by means of a homogeneous particle
model in which allowance is made for decrease in the rate coefficient for
termination as polymerization proceeds. Napper (1971) has also criticized
the core-shell morphology proposed by Grancio and Williams, and has
argued in favor of a core-shell morphology of the reverse kind, Le., a
particle that comprises a monomer-rich core surrounded by a polymer-rich
shell. Williams (1973) has challenged this view. Keusch et al. (1974) have
discussed the nature of the distribution of polymer segments in a swollen
latex particle and find that the concentration of segments should be high in
the central region of the particle and fall to zero at the particle surface, thus
supporting a core-shell morphology of the Grancio-Williams type. The
evidence for and against concentration gradients in polymerizing monomerpolymer particles in a reaction system of the type envisaged here has been
reviewed by Gardon (1973). Theconclusion
reached is that such concentration gradients as may form within such a particle are unlikely to
be large enough to affect the rate of polymerization significantly. A further
review of this subject has been published recently by Vanderhoff (1976).
Blackley and Haynes (1977) have argued that the marked reduction in the
rate of the emulsion polymerization of styrene that occurs when certain
organic diluents (e.g., ethylbenzene) are added is consistent with the view
that the equilibrium concentration of monomer in the polymerization loci is
less when polymerization occurs than when it does noto This interpretation
has recently been challenged by Azad and Fitch (1980).
The conclusions to which the present evidence seems to lead can be
summarized as follows:
1. The monomer concentration is probably essentially uniform across
the polyermization locus if the monomer is a good solvent for the polymer.

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 153


However, the equilibrium concentration during polymerization may be
somewhat less than that attained if no polymerization is occurring.
2. The monomer concentration will remain constant as polymerization
proceeds only if the following conditions are fulfilled (or if, fortuitously, the
opposing effects which they tend to have upon equilibrium swelling happen
to balance): (a) the size of the loci remains essentially unchanged as
polymerization proceeds and (b) the interfacial free energy associated with
the interface between the loci and the external phase remains unchanged as
polymerization proceeds. If the loci increase in size only to a small extent as
a consequence of polymerization, and the amount of adsorbed stabilizer in
the system remains constant, then condition (b) will probably be fulfilled.
Furthermore the extent to which polymer particles are swollen by excess
monomer is usually such that the polymer/monomer ratio at the polymerization locus is high, even from the earliest stages of the reaction. The
viscosity of the reaction medium is therefore high, and allowance must be
made for this when assigning values to the rate coefficients for processes
such as propagation and termination.

C.

Mechanism 01 Fundamental Processes; lnfluence 01 Locus Size

Although this chapter is concerned primarily with the derivation of


theoretical predictions-and
for this purpose the various fundamental
processes that occur within the reaction system are characterized by
appropriate
physicochemical parameters
regardless of underlying
mechanism-it is nevertheless desirable to consider briefly the mechanism
of some of the fundamental processes in order to appreciate the significance
of the physicochemical parameters.
1. Acquisition of Radicals by Loci
Formally, the rate of acquisition of new radicals by loci is characterized
by the average rate of entry of radical s into a single locus. Denoted by a,
this quantity is equal to p/N, where p is the overall rate of entry of radicals
into all the loci contained within unit volume of the reaction system, and
N is the number of loci of all types contained in' unit volume of the
reaction system. p is not necessarily equal to the rate of generation of new
radical s within the external phase. There may be wastage, and the extent of
wastage may depend upon, inter ala, the rate of generation of new radicals
and the value of N. Thus, Hawkett et al. (1980) and Gilbert et al. (1980)
have recently shown that, for the seeded aqueous emulsion polymerization
of styrene initiated by potassium persulfate, the capture efficiency for
radicals is only about 1% at an initiator concentration of 8.33 x 10- 2 M,
but rises to about 50% at 1.6 x 10- 5 M.

154

D. C. Blackley

Controversy has surroundec the question of whether it is mainly primary


unreacted initiator fragments (e.g., sulfate radical ions) that enter the
reaction loci, or whether some polymerization of monomer dissolved in the
external phase occurs first, and then it is the oligomers (which may well be
surface-active) formed by this process that are acquired by the reaction loci.
Present opinion tends strongly to favor the latter view, especially when the
monomer is appreciably soluble in the external phase.
A related matter. concerns the physical mechanism by which radical s
(primary or oligomeric) are acquired by the reaction loci. One possibility,
first proposed by Gardon (1968) and subsequentIy developed by Fitch and
Tsai (1971), is that capture occurs by a collision mechanism. In this case, the
rate of capture is proportional to, inter alia, the surface area of the particle.
Thus, if the size of the reaction locus in a compartmentalized free-radical
polymerization varies, then u should be proportional to r2, where r is the
radius of the locus. A second possibility (Fitch, 1973)'is that capture occurs
by a diffusion mechanism. In this case, the rate of capture is approximately
proportional to r rather than to r2. A fairly extensive literature now exists
concerning this matter (see, e~g., Ugelstad and Hansen, 1976, 1978,
1979a,b). The consensus of present opinion seems to favor the diffusion
theory rather than the collision theory. The nature of the capture mechanism is not, however, relevant to the theory discussed in this chapter. It
is merely necessary to note that both mechanisms predict that the rate of
capture will depend on the size of the reaction locus; constancy of u
therefore implies that the size of the locus does not change much as a
consequence of polymerization.
2. Exit of Radicals from Loci
Of the various possible first-order processes by which radicals can be lost
from reaction loci, the most likely is that of exit from the locus back into the
external phase. It is reasonable to suppose that the likelihood of a propagating radical being lost from a reaction locus by this process falls off
sharply as the degree of polymerization increases. It is the monomeric
radicals and very short-chain oligomers that are likely to be lost in this way.
For this reason, the favored molecular mechanism for radical exit is transfer
to monomer followed by exit of the monomeric radical or of a short-chain
oligomer formed from it. In fact, the product of the transfer-to-monomer
rate coefficient (relative to that for propagation) and the solubility of the
monomer in the external phase is sometimes taken as a rough index of the
tendency for radicals to be lost from reaction loci by exit.
Whatever may be the exact nature of the molecular processes that lead
to exit, the rate of loss of radical s from a single reaction locus by exit is
written simply as ki, where i is the number of radicals in the locus and k is a

4. Kinetics of Compartmentalized Free-Radical Polymerization Reactions 155


composite coefficient. As with (1,the parameter k almost certainly depends
on the size of the reaction locus. Thus, results published by Hawkett et al.
(1980) and also by Gilbert et al. (1980) indicate that k is inversely proportional to r-2. This is consistent with an exit mechanism that involves the
diffusion of a species (e.g., a monomeric radical) from a spherical particle.
On the other hand, Smith and Ewart (1948) assumed k to be expressible in
the forro koa/v (Le., k IX.r-l), where a is the surface area of the reaction
locus, v is its volume, and ko is a constant. ko then expresses the tendency of
radical s to diffuse across a unit area of the surface of the reaction locus
when radical s are present in the locus at unit concentration. Again, for the
purposes of this chapter, the precise functional dependence of k on r matters
little; it is merely necessary to note that there are good reasons for
supposing that k varies with the size of the reaction locus.
3.

Propagation and Termination

Little needs to be said here except to note that (i) the rate of propagation
is unlikely to be appreciably dependent on the size of the reaction locus,
whereas the rate of termination is likely to be appreciably size-dependent
and (ii) the rate of both propagation and termination will be reduced if the
viscosity of the reaction medium rises, but the rate f termination will be
reduced more than that of propagation.
The physical reason why the rate of propagation is not appreciably size
dependent is simply that a propagating radical is always surrounded by the
same concentration of monomer molecules wherever the radical is in the
reaction locus and, to a first approximation, whatever the size of the locus.
However, to the extent that the monomer/polymer ratio in the reaction
locus depends on the size of the locus (see Section II,B), then some
dependence of rate of.pro paga tion on locus size will be observed.
Since most of the monomer in a compartmentalized free-radical polymerization reaction is consumed in the propagation reaction, it is customary
to write the overall rate of polymerization as
dM/dt

= 1Nkp[M]

(1)

where 1 is the average number of propagating radical s per reaction locus, kp


is the rate coefficient for propagation, and [M] is the concentration of
monomer in the reaction locus. (For discussion of the precise meaning of
dM/dt, see Blackley, 1975.) From the above discussion, it is clear that there
may be some slight size dependence of dM/dt through [M]. The rate of
propagation in a single reaction locus containing i radicals is written
forroally as kpi[M].
.
In the case of the termination reaction, the rate is expected to be
appreciably size dependent because it involves the interaction of two radical

156

D. C. Blackley

speciesthat have to find each other. Formally, the rate of termination in a


single locus is written as k,i(i - 1)/v, where k, is the rate coefficient for
mutual termination (see Blackley, 1975 and Smith and Ewart, 1948). This
expression can be most easily justified by regarding (i - 1)/v as the concentration of free radical s with which any one of the i free radical s may
react.
An important consideration is that possibly the effect of locus size, and
almost certainly the. effect of locus viscosity, will depend on the degree of
polymerization of the propagating radicals seeking to interact. The larger
these radicals have grown, the more pronounced should be the effect of
increasing locus size and viscosity in reducing the rate of termination
relative to propagation.

111. The Time-Dependent Smith-Ewart Differential DifIerence Equations;


Methods Available CorTheir Solution

A. Derivation
The fundamental equations that govern the behavior of a compartmentalized free-radical polyrnerization reaction in which the radicals are
generated excIusively in the external phase are most readily derived by considering the rates of the various processes by which loci containing exactIy
i propagating radicals are formed and destroyed. These processesare illustrated in Fig. 2 as transitions between various "states of radical occupancy"
of the loci, each state of occupancy being defined as the number of
0+2)0 +l)ni+2X

(j + Oni+1 k

i+2

i+1
ink

i(j -I)nx

njO'

H
ni-lO'

j-2
F.ig. 2. Transitions betweenstatesof radical occupancyrequired for derivation of timedependentSmith-Ewart differential differenceequation for state i.
.

4.

Kinetics of Compartmentalized Free-Radical Polymerization Reactions 157

propagating radical s present in the locus.* Only those transitions are


shown that affect the value of ni, the number of reaction loci per unit
volume of reaction system which contain i propagating radicals. This type
of locus is referred to henceforth as a "Iocus of type i."
There are three processes by which loci of type i can be formed. These
are (i) from type i

1 loci by acquisition

of a radical, (ii) from type i + 1

loci by loss of a radical, and (iii) from type i + 2 loci by mutual termination
within the loci. Likewise there are three processes by which loci of type i
can be destroyed: (i) by acquisition of a further radical, thereby becoming a
locus of type i + 1; (ii) by loss of one radical by exit, thereby becoming a
locus of type i - 1; and (iii) by loss of two radicals by mutual termination,
thereby becoming a locus of type i - 2. The differencebetween the sum of
the rates of the first group of processes and the sum of the rates of the
second group of processes gives dn;/dt, the overall rate at which loci of type
i are formed. The result for dn/dt is readily found to be
dl1;/dt

= (11-1 -

11)O'+

{(i +

l)n+1

il1}k

+ {(i + 2)(i + 1)11i+2- i(i - l)n}(kt/v)

(2)

where i = 0,1,2, ..., and n-1 = O. It will be convenient subsequently to


denote the ratio kt/v by the single symbol X. It is c1ear that constancy of the
coefficient X implies that v is constant as well as kt (apart from the unlikely
possibility that simultaneous variations in v and kt fortuitously balance
each other). These equations comprise an infinite set of linear differential
difference equations, conveniently known as the "time-dependent SmithEwart differential difference equations" because they are a generalization to
the nonsteady-state of the infinite set of equations first derived by Smith
and Ewart (1948) fOl;steady-state reaction systems. The first term on the
right-hand side of Eq. (2) represents the net effect of radical acquisition on
dn/dt. The second and third terms likewise represent the net effects of firstorder radical loss and bimolecular mutual termination, respectively. The
equations originally given by Smith and Ewart (1948) for the steady state
follow immediately from the Eq. (2) by setting all the dn;/dt equal to zero.
The time-dependent Smith-Ewart differential difference equations can
also be derived by an alternative method first used by O'Toole (1965) for
the steady state. In this method, one considers the rates of transition of
locus populations across a notional barrier situated between two neighboring states of radical occupancy. The barrier is illustrated in Fig. 3, where the
* In faet, as Gilbert and Napper (1974) have pointed out, there is a useful analogy between
the oeeupaney of quantized energy states and the presenee in reaetion loei of free radieals
whose basie "quantum" is the free radical

158

D. C. Blackley

1+2
n..,.
I

---1---1_--1--

i+1

- -------------

i-I
1-2
.

these transltlons not relevant

Fig. 3. Transitions between states of radical occupancy required for derivation of


O'Toole equation for rate of transition of loci across notional barrier between states i and
i + 1.

barrier has been drawn between the states i and i + 1. The O'Toole
approach is simpler than the earlier approach of Smith and Ewart (1948)
(which considers the balance of transitions from a particular state of radical
occupancy, as illustrated in Fig. 2) to the extent that only four transitions

have to be considered instead of six. Of these four, three (i + 1 ~ i,


i ~ i + 1,and i + 2 ~ i)are identicalto those of the Smith-Ewart approach.
The remaining three Smith-Ewart transitions (i ~ i - 1, i - 1 ~ i, and
i ~ i - 2) are irrelevant because they do not involve transitions across the
barrier between states i and i + 1. The fourth transition

that has to be taken

into account in the O'Toole approach is i + 1 ~ i - 1; this is irrelevant to


the Smith-Ewart
approach because it does not affect the number of loci
which contain i radicals. However, although the O'Toole approach is
simpler than the Smith-Ewart approach in that fewer transitions have to be
considered, and although application to steady-state reaction systems is
simple, application to nonsteady-state
reaction systems is not as simple as
in the case of the Smith-Ewart treatment.
O'Toole presents his argument in terms of what he calls "the probability
of i-fold occupancy." This probability is mathematically equivalent to the
fractional frequency of occurrence of loci that contain i radicals, Le., to
ni/N. Such fractional frequencies will bC?denoted by Vi in this chapter.
However, the argument can equally well be presented in terms of actual

locus populations ni, Inspection of Fig. 3 shows that the overall rate of

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 159


transition of loci upward across the barrier between states i and i + 1 is
given by
R-+i+l = n(1- (i + l)n+lk - {(l + l)in+l + (i + 2)(i + 1)n+2}x (3)
To obtain the equation that must be obeyed by the locus populations in the
steady state, one merely equates R-++
1 to zero. It will be shown subsequently that the resulting equation is exactly equivalent to that obtained
from Eq. (2) by setting dn;/dt

= O. If

the reaction system is not in the steady

state, then Eq. (3) gives merely the overall rate of transition of loci upward
across the barrier. To obtain an expression for dn;/dt, it is necessary to
observe that this quantity is the rate of accumulation of loci in state i, and
that this in turn is equal to the difference between the rate of transition
upward across the barrier between states i - 1 and i and the rate of
transition upward across the barrier between states i and i + 1 (see Fig. 4).
Thus, we have
(4)
Equation (2) then follows immediately by noting that
R-l-+i

= n-l(1-

ink - {i(i - l)n + (i + l)in+}x

(5)

It is desirabIe to mention briefly the matter of the dimensions and units


of the physicochemical parameters (1, k, and X. The dimensions of the
left-hand side of Eq. (2) are [lociJ[timer 1. The dimensions of (1, k, and
X are therefore, respectively [timer 1, [radicalsr 1[timer 1, and
[radicalsr2[timerl.
To the extent that numbers of loci and numbers of
radicals are regarded as "mere numbers'! and therefore dimensionless, the
dimensions of the left-hand side of Eq. (2) become [timer 1, and those of (1,
i+2
i+1

---

- - - --- - - - - -1-1

1-2
la)

lb)

Fig. 4. Transitions between states of radical occupancy required for derivation of dnJdt
by the O'Toole approach. (a) shows transitions across barrier betwe~n states i and i + 1; (b)
shows transitions between states i - l and i.

160

~o Co Blackley

k, and X are also all [timer

lo

The dimensions of the k{= Xv) of Eqo (2) are

then [radicalsr 2[volumeJ[timer

or [volumeJ[timer 1, according to

whether the dimensionality of number of radicals is or is not recognisedo


B.

Methods Available for Their Solution


lo Matrix Methods
The objective is in general to obtain explicit expressions for either the

actual locus populations n(t) (i = 0,1,2,ooo,) as functionsof t, or the


fractional locus populations v(t) (i = 0,1,2, oo.,) as functions of to The
mathematical procedures for obtaining v(t) are identical to those for
obtaining n(t),because the form of Eqo (2) is unaltered by dividing through
by No In the special case where the reaction system is in the steady state, the
objective is to obtain the steady-state values of n(or'V).
The feasibility of the matrix approach to solving the time-dependent
Smith-Ewart differential difference equations depends on the fact that each
of these equations is linear in certain n Thus, the typical equation of the
o

set [Eq (2)Jcan be rearranged to read


o

dn/dt

= an-I - {a + ik + i(i - l)x}n


+ (i + l)kni+1 + (i + 2)(i + 1)Xn+2

(6)

The infinite set of equations can therefore be written as the single matrix
equation
d[n]/dt = [nJ[n]
(7)
where [n]T is the matrix [nonl n2 . . .J and [nJ is a matrix of coefficientso
The structure of [nJ is such that the jth row contains the following elements
in the U - l)th,jth, U + l)th, and U + 2) columns respectively, and all the
other elements of the jth row are zeros:
a

- {a + U- l)k + U- l)U - 2)x} jk U + l)jx

(8) .

To the extent that the elements of the matrix [nJ are dependent neither on
time nor on the nnor on each other, it is possible immediately to write down
in matrix form the solution ofthe differential matrix Eq. (7). It may, however,
be preferable first to transform the differential matrix equation, eogo,by
using a suitable similarity transformationo Having solved the transformed
equation, the n can then be obtained from the matrix of transformed ni by
applying the inverse transformation.
It should be noted that, beca use the set of equations contained in Eq. (2)
is theoretically infinite in number, the matrices [n] and [nJ of Eq. (7) are
infinite with respect to the number of elements they contain. What is done

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 161


in practice is to truncate the set of equations in Eq. (2) at some value of i
which is believed to be such that the fraction of reaction loci that contain
more than this number of propagating radicals is negligible. An alternate
way of regarding the matter is to place a limit on the maximum number of
propagating radicals that a locus can contain. Denoting this maximum
Qumber of radicals by s, [Q] then has the form of an (s + 1) x (s + 1)
square matrix. As a consequence of this truncation, it may be necessary to
adjust the coefficients of the final equation of the set in Eq. (2) in order to
preserve constancy of the totallocus population. This adjustment inevitably
introduces a slight distortion into the set of equations, but the distortion will
be negligible if the maximum value of i is taken sufficiently large for the
reaction system whose behavior is being represented.

2. Use of a Locus-Population

Generating Function

In this method, the infinite set of linear differential difference equations


[Eq. (2)] is also con verted into a single equation, but by a different
approach to that used in the matrix approach. A new function 'P(~, t) is
introduced, defined by the equation
00

'P(~, t) =

L v(t)~
=o

(9)

where ~ is an auxiliary variable. This function is described as a "locuspopulation generating function," because once it is known, all the fractional
locus populations at all times can in principIe be obtained from it by
making use of the fact that
1 0'P
(10)
v(t) ="1
::I}'i
lo

( )
u,>

~=o

Furthermore, the average number of propagating radical s per locus at any


'instant Y(t) can be found from

0'1'

00

Y(t) = 1=1
.L iy(t)

()
~u,>

(11)

~=l

and the variance of the distribution of locus populations at any instant can
be found from
00

V(t)

L (i -

=o

~
{ o~

Y)2y(t)

~0'P

= L i2V(t) -

o~ )} ~= 1

0'1'

/2

(o~ )~=

(12)
1

162

D. C. Blackley

The function 'P(~, t) c1early has the additional property that


CJ)

'P(I, t) = L: v(t)= 1
=o

(13)

for all values of t. Thus, the time dependency of 'P(~, t) should always
disappear when the substitution ~= 1 is made-this is a consequence of the
assumption that the number of reaction loci in the system remains constant.
The function 'P(~, t) has another interesting property which as yet hasnever
been exploited as fai:as is known. Ir the substitution ~ = - 1 is made; then
we obtain
'P(-I, t) = vo(t) - v(t) + V2(t)- v3(t) + ...
(14)
Thus, 'P(-I, t) gives at any time the difference between the fraction of the
total number of loci that contain an even number of propagating radicals
and those that contain an odd number, the fraction containing an even
number inc1uding within it those containing no propagating radicals at all.
Ir it is desired to know the fraction of the total number of loci that do not
contain any propagating radicals, then this can be obtained from the relationship
vo(t) = 'P(O,t)
-(15)
Thus, it is c1ear that the single function 'P(~, t) is capable of yielding a great
deal of information concerning the distribution of locus populations in the
reaction system.
There are at least four important advantages to using locus-population
generating functions as compared with the matrix approach:
1. It is quite unnecessary to truncate the set of equations contained in
Eq. (2); in fact, the logic of the method requires that the set be maintained
infinite in extent. There is therefore no possibility that the set of equations
actually being solved is a distorted representation of the behavior of the
reaction system being investigated. To the extent that the requisite mathematical operations ha ve been correctly performed, the resultant expressions for v(t),etc. are free from approximation errors.
2. It is possible in several cases to obtain expressions for v(t), etc. in
c10sed analytic form rather than as somewhat unwieldy algebraic
expressions.
3. The method using the locus-population generating function is more
readily applicable to systems in the steady state.
4. It is in principIe possible to make inferences concerning the nature of
the distribution of locus populations if the generating function can be
recognized as that of the frequencies of some known distribution.
It should perhaps be pointed out that in several previous publications
(e.g., Stockmayer, 1957; Birtwistle and Blackley, 1977, 1978: Birtwistle et al.,

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 163


1979) the function 'P(~, t) has been detined using the actual locus 'populations n(t) and not the fractionallocus populations v(t).The difference is
trivial. The function as detined by these previous writers is merely N times
the function as detined here, N being the total number of reaction loci in
unit volume of the reaction system. The advantages of the present detinition
over the alternative detinition are that (i) the analytic expressions for 'P(~, t)
are simplitied to the extent that they do not contain N as an arbitrary
multiplier and (ii) the expressions for i(t) and v(t) in terms of 'P(~, t) are
somewhat simpler. It should be noted, however, that if 'P(~, t) is detined
using n(t)in place of v(t),then the right-hand side of Eq. (10) gives n(t),and
not v(t).
In order to convert the set of differential difference equations in Eq. (2)
into a single differential equation with 'P as the dependent variable, each
equation for dn;/dt is multiplied by ~/N [the factor N-1 changes n(t)into
the corresponding v(t)], and then all the equations so obtained are added
together. It is then noted that

(16)

where in each case the summations cover all possible values of i. The
resultant single differential equation then readily transforms to
.8'P
-

8t

t
= O\~
- 1)'P + k(l - ~)-8'P + X(l - ~2)-82'P

8~

8~2

(17)

This is the partial differential equation for 'P(~, t) which has to be solved for
each particular case. The solution has to be subject to the initial (boundary)
conditions appropriate to that particular case.
For the special case of reaction systems that are in a steady state, n(t)
are invariant with time, and therefore 'P(~, t) is not a function of t. It is therefore denoted by 'P(~) in this case. Thus, in this case 8'P/8t = O,and the righthand side of Eq. (17) must be zero. What were hitherto partial differential
coefficients of 'P(~, t) with respect to ~ become ordinary differential coefticients. Dividing through by the factor 1 - ~ then gives
d'P
d2'P
a'P - k- X(l + J:) -

d~

..

de

=O

(18)

164

D. C. Blackley

as the ordinary differential equation to be solved for 'P in this case. This is
identical to the differential equation first given by Stockmayer (1957) for the
locus-population generating function that characterizes the behavior of
reaction systems in the steady state. Equation (18) can also be obtained
from the infinite set of equations for the steady-state as given by O'Toole
(1965). It will be recalled that this is the set of which Eq. (3) with the lefthand side set equal to zero is the typical member. Again the procedure is to
multiply each eq\lation by ~i/N, add together all the equations so obtained,
and then note the relationships summarized in Eq. (16) together with the
additional relationship
.
2'P
(i + 1)ivi+l~1 = ~ ~2

(19)

The resultant single differential equation can then readily be transformed


into Eq. (18). Thus, it becomes clear that the O'Toole (1965) formulation of
the steady-state problem is exactIy equivalent to that of Smith and Ewart
(1948), notwithstanding that the approach to the problem is somewhat
different.
3.

Approximation Methods

As has already been explained, the matrix approach involves some


degree of approximation because of the necessity of truncating the set of
equations in Eq. (2). There have, however, been other attempts to solve the
Smith-Ewart differential difference equations by approximation methods
that have not used the matrix approach. These methods have been used for
steady-state systems as well as for nonsteady-state systems. Again the
procedure is to set a limit on the maximum number of propagating radicals
which a single locus can contain. This maximum number is sometimes qujte
low, e.g., either one or two. The resulting set of equations is then solved by
conventional methods for simultaneous linear algebraic equations (in the
case of equations representing the steady state) or simultaneous linear
differential equations (in the case of equations representing the nonsteady
state).

IV. Solution for the Steady State


The general solution to the Smith-Ewart differential difference equations
for reaction systems in the steady state is most readily obtained using the
locus-population generating function approach. This was first demonstrated
by Stockmayer (1957) and subsequentIy by O'Toole (1965). It is convenient
to introduce two new parameters :;and m defined as
:;= av/kt = a/x,

m = kv/kt = k/X

(20)

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 165


It is subsequently convenient to introduce a third parameter h defined by
the equation
(21)
The parameter e is a measure of the rate of radical entry relative to the rate
of bimolecular termination within loci. Similarly, the parameter m is a
measure of the rate of radical exit relative to the rate of bimolecular
termination within loci. The ratio e/m = CT/kis a measure of the rate of
radical entry relative to radical exit. The so-called Case 1, Case 2, and Case
3 of Smith and Ewart (1948) correspond to the following circumstances:
Case 1: m large relative to e;
Case 2: k = O,kt = 00 (Le., e = m = O);
Case 3: e large relative to m.
Note that both e and m increase linearly with locus volume v if CT,k, and kt
remain constant, but the ratio e/m is independent of both v and kt.
The first step is to solve Eq. (18) for 'P(~). Introducing the parameters e
and m, this equation becomes
(1 + ~) d2'P/d~2 + m d'P/d~ - e'P = O

(22)

Two new variables x and y are introduced. These are defined by the
equations
x

y = 'P(~)/xl-m

= 2)e(1 + ~),

(23)

Equation (22) then becomes


X2 d2y/dx2 + x dy/dx

{(1 - , m)2

X2}

=O

(24)

This is a modified Bessel equation of order 1 - m. Except for the physically


improbable case wliere 1 - m is exactly an integer or zero, the general
solution to Eq. (24) is
y = AIl-m(X) + Blm-l(X)
(25)
I

where Il-m(x) and Im-l(X) are modified Bessel functions of orders 1 - m and
m - 1 respectively, and A and B are constants. Physical considerations (see
O'Toole, 1965) indicate that the A must always be zero. The value of B is
determined by the requirement that 'P(1) = 1. The final result for 'P(~) is

(\/+fl+1)

2(m-1j/2

'P(~) = Im-l(h) (1 +

~)(1-m)/2Im-l

(26)

Application of Eqs. (10) and (11) gives


h

hi2m-1-3i
and

Vi

= i! Im-l(h)

Im-l+i

(J2)

(27)

(28)

166

D. C. Blackley

Factors such as Im(h)/Im-l(h) have been convenientIy referred to by van der


Hoff (1958, 1962) as "subdivision factors." In effect, they quantify the extent
to which the overall concentration of propagating radical s in the reaction
system as a whole is enhanced by compartmentalization of the propagation
steps into a large number of small reaction loci.
An important further contribution to the analysis of steady-state reaction
systems has been made by Ugelstad et al. (1967). They have shown how
account can be taken of the likely possibility that radicals that exit from the
reaction loci contribute to the stationary concentration of free radicals in
the external phase which is available for entry into a reaction loci. For this
purpose, it is necessary to distinguish bimolecular mutual termination
between radical s that occurs in the reaction loci (i.e., within polymer/
monomer partic1es)from that which occurs in the external phase. The rate at
which the former reaction occurs is characterized by the rate coefficient ktP,
the rate of the latter reaction by ktE. The total rate of entry of radicals into
all loci within unit volume of reaction system is th~n expressed as the
sum of three contributions. The first derives from the rate of formation of
new "acquirable" radicals within the external phase; the second derives
from the rate at which acquirable radicals become present in the external
phase by the process of exit from the loci; the third (which is negative)
allows for the fact that radicals can be lost from the external phase by
bimolecular mutual termination within the external phase. The resultant
equation is
00

p = p' + k

i= 1

ini - 2ktE e2

(29)

where p is the total rate of entry of radicals into all the N loci in unit
volume of reaction system, p' is the rate of generation of new radicals within
unit volume of reaction system, and e is the stationary concentration
of radicals within the external phase. The form of this equation implies that
all the radical s that become present in the external phase (whether by
generation or by exit from the loci) are potentially available for reentry,
except insofar as they are destroyed by mutual termination. Dividing
through by N, setting a = p/N, a' = p'/N, and y = pie, and introducing
parameters

e = av/ktP'

e' = a'v/ktp,

m = kv/ktP,

Z = 2NktpktE/yZv

(30)

Eq. (29) readily transforms to


e = e' + mf - Ze2

(31)

Ugelstad et al. have used Eq. (31) in conjunction with Eq. (28) to calculate y
as a function of e' for various values of m and Z. It is valid to regard the e of

4.

Kinetics of Compartmentalized

Free-Radical Polymerization Reactions

167

Eq. (31) as equivalent to the h of Eq. (28) because the reaction system is
presumed to have reached a steady state in which, inter alia, ,/,and 1 have
attained stationary values.

V. Solutionsfor the NonsteadyState


A.

Case in which Radical Loss is Predominantly by First-Order Processes

Considerable progress has been made in recent years in obtaining


solutions to the time-dependent Smith-Ewart differential difference equations for various special types of reaction system in the nonsteady state.
Although it has so far not pro ved possible to give an entirely general
solution to these equations, it has pro ved possible to obtain a general
solution to a modified set of equations which, under certain circumstances,
approximate to the exact set of equations.
The simplest case for which an exact solution has been obtained is that
of a reaction system in which radicals are lost from reaction loci almost
exdusively by first-order processes. Initially, the reaction system is devoid
of free radicals, so that no polymerization is occurring. Then, at a time
taken as the zero of subsequent time, radicals begin to be generated in the
external phase at a constant rateo This case has been discussed in detail by
BirtwistIe and BIackley (1977, 1979) using the locus-population generating
function approach, and by Gilbert and Napper (1974) using the matrix
approach. The expressions for the locus-population generating function, for
v(t), and for l(t) haye also been given by Weiss and Dishon (1976), but
without discussion of the significance of the result.
Setting X = O,the partial differential equation to be solved for this case is

..

o\f/ot

= c(~-

1)\f + k(1 - ~)o\f/o~

(32)

The boundary conditions are


vo(O)= 1
v(O) = V2(0)

= ... = O

(33)

and, furthermore, for all t


00

;=0

v(t)

vo(O)

(34)

Equation (32) with these boundary conditions is amenable to solution by


the method of separation of variables, in which the function \f(~, t) is
assumed to be of the form 2(~)'11:t),where 2(~) is a function of ~ only, and

168

D. C. Blackley

T{t) is a function of t only. The details of the process whereby the solution
can be obtained have been given by Birtwistle and Blackley (1977). The
result for 'P(~, t) is
'P(~, t) = exp{(o-k)(~ - 1)(1 - e-k/)}

(35)

It follows immediately from this that at all times the distribution of locus
populations with respect to radical occupancy is Poissonian, and that the
parameter of the distribution at any instant is (O"/k)(1- e-k/). These conclusions follow because the expression for 'P(~, t) is recognizable as the
frequency-generating function for a Poisson distribution (see Kendal and
Stuart, 1965). It then follows that

v.(t) = .!. ~(1


,
i! k

- e-k/)

} {
eXP

-~(1

- e-k/)

(36)

and

(37)

T~t) = (O"/k)(1- e-k/)

Alternatively, the results embodied in Eqs. (36) and (37) can be derived from
Eq. (35)using Eqs. (10)and (11).The general result for v(t) then shows that
the distribution of locus populations must always be Poissonian with timedependent

parameter

(O"/k)(1- e-k/). Setting t

= 00 in

these results givcs the

following predictions for the reaction system when it has attained the steady
state:
'P(~,oo)

= e(O'/k)(~-l)

1 O" i
v(oo)=i! k e-O'/k

()

Y(oo) = O"/k

(38)

Approximate solutions for vo(t),Vl(t),and V2(t)have been given for this


case by Gilbert and Napper (1974). These solutions were obtained using the
matrix method. The expressions obtained are algebraically cumbersome
compared to the general result embodied in Eq. (36), but can, of course, be
readily handled using modern computers. Provided that O"/k is small
compared to unity, the solutions given by the two methods predict almost
identical numerical results for Vi(t);in fact the ratio of the values given by
the two methods is e-O'/k/(1- 0"/2k)2,which is almost unity for O"/k~ 1.
An example of the numerical predictions that this theory gives is shown
in Figs. 5 and 6. Figure 5 shows the fractionallocus populations vo(t),Vl(t),
and

V2(t)

as functions of t for a reaction system for which

= 1 x 10- 5 sec-l, k = 5 X 10-4 sec-l, and X is zero. It is seen that the


steady state is reached after approximately 104 sec, and that at all times
most of the reaction loci are devoid of propagating radicals. Of those loci
that contain radicals, most contain only one. Figure 6 shows the prediction
for T(t) as a function of t for this reaction system. Again this shows the
O"

4. Kinetics of Compartmentalized Free-Radical Polymerization Reactions 169


1.0
=O
i=I

;,;-0.5

i=2

10

15

20

time (sec) x 10-3

Fig. 5.

Fraetional locus populations, v(t),as functions of time t for i = 0,1,2 for

reaction system for whieh radiealloss from reaction loei is exclusively by first-order processes.
Values taken ror (] and k are 1 x lO-s sec- and 5 x 10-4 see- respeetively. The ordinates
for i = Oare vo(t); those for i = 1 are 40 v,(t); those for j = 2 are 2 x 103 v2(t).(Reproduced with
permission of J. Chem. SocoFaraday l.)

..
Q

,x

10

15

20

time (sec) x 10-3

Fig. 6.
Average number of radieals per loeus I(t) as a funetion of time t for reaetion
system to which Fig. 5 refers. (Reprodueed with permission of J. Chem. SocoFaraday l.)

170

D. C. Blackley
10

10

15

20

time (se e) x 10-3


Fig. 7.
Conversion of monomer to polyrner M(t) as a function of time t for reaction
system to which Figs. 5 and 6 refer. (Reproduced with permission of J. Chem. Soco Faraday l.)

steady state being reached after about 104 seco The average number of
radical s per locus in the steady state is small (0.02) for this case.
The amount of monomer M(t) converted in a unit volume of the reaction
system after the eIapse of time t can be obtained by substituting for T(t) in
Eq. (1) from Eq. (37), and then integrating

over the range t

= O to t = t.

The

result is
(39)
The prediction for M(t) as a function of t for the reaction system to which
Figs. 5 and 6 refer is shown in Fig. 7. Marked deviation from linearity in
the predicted curve for conversion as a function of time occurs over the
range t < 5 X 103 seco
An interesting generalization to the case X = O has been given by
Birtwistle and Blackley (1979). In this generalization, the theory is extended
to include cases where a and k are time dependent. In order to emphasize
the time dependence of a and k they are written as a(t) and k(t). What
Birtwistle and Blackley (1979) have shown is that the solution to Eq. (32)
always has the form
'I'(~, t) = e<~-1)8(t)

(40)

where O(t) is a function of time that satisfies the ordinary differential


equation
dO(t)/dt = a(t) - k(t)O(t)
(41)

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 171


and is subject to the initial condition 0(0) = Ofor reaction systems for which
the boundary conditions are as stated in Eq. (33). This result for 'I'(~, t)
shows that the distribution of locus populations with respect to radical
occupancy is at aIl times Poissonian, notwithstanding the time dependence
of a and k, and that the parameter of the distribution is always equal to
O(t).It therefore foIlows that always
7(t) = O(t)

(42)

and
-8(/)
v(t) = {O(tW
., e
lo

(43)

If a and k are not time dependent, then O(t)has the value (a/k)(1 - e-kl)
and the results are identical to those obtained previously. The foIlowing
particular cases of time-dependent a and k have been considered in detail
by BirtwistIe and BIackley (1979):
1. The case where a decays exponentiaIly with time, and k is constant.
2. The case where a decays linearIy with time, and k is constant.
3. The case where a decays as a consequence of second-order depletion
of initiator, and k is constant.
4. The case where a is constant, but k varies because the loci grow at a
constant rateo
5. The case where a varies because radicals lost to the external phase
are available for reinitiation, and k is constant.
Of these cases, the first is probably of greatest interest because the
decomposition of moSt dissociative initiators is kineticaIly of first order with
respect to initiator concentration, and therefore both the initiator concentration and the rate of decomposition of the initiator wilI decay exponentiaIly with time. Setting a(t) = aoe-al, the solution for O(t)is

(44)
except for the unlikely special case where o: = k, in whichcase the result for
O(t)is
O(t)= aote-kl

(45)

Figure 8 shows examples of the variation of O(t)[and hence of T(t)] with t


for a reaction system for which k has the constant value 5 x 10-4 sec-l and
a decays exponentiaIly with time with ao = 1 x 10- 5 sec- I and o: taking
various values between zero and 5 x 10-3 sec-I. It appears that the
exponential decay of a has little effect upon the variation of O(t) with t,
provided that o:/k< '" 10- 2. But in aIl cases the effectof ~ nonzero o:is to

D. C. Blackley

172
2

a/k

=O

a/k =0.01

..
Q
X

,11

q;

15

20

Fig. 8. Loeus-populationdistribution parameter O(t)and the average number of radieals per loeus T(t) as funetions of time t for reaetion system for whieh radiealloss from reaetion
loei is exc1usively by first-order proeesses and rate of radical entry deeays exponentially with
time. The values taken for (10and k are 1 x 10- s see- and 5 x 10-4 see-, respeetively; eurves
are given for various values of c(, indieated by the C(/k ratios
(Reprodueed with permission of J. Chem. Soco Faraday l.)

appended

to the eurves.

cause 8(t) to rise to a maximum and then eventually to decay to zero. This
ultima te decay to zero is a consequence of depletion of the source of
radicals. It carries the implication that eventually all v(t) for i > O themselves decay to zero and vo((0) = 1.

B.

Case in which Generation oi New Radicals Ceases

Birtwistle et al. (1979) have obtained an explicit analytic solution to


Eq. (17), and hence to the general time-dependent Smith-Ewart differential
difference equations, for the case where the rate of formation of new radical s
in the external phase is zero (Le., (f = O).Of course, if no radicals ever ha ve
been generated within the external phase of the reaction system, then the
problem becomes trivial and admits of an obvious and simple solution,
namely, that allloci are at all times devoid of propagating radical s and the
rate of polymerization is always zero. This solution is c1early of no interest.
The case that is of interest is that of a reaction system in which radical s
have been generated within the external phase, so that a certain rate of
polymerization has developed; then the rate of generation of new radicals is
suddenly reduced to zero, so that the reaction rate then decays. It is with
the characteristics of the reaction during this period of decay, following the
cessation of the generation of new radicals, that this aspect of the theory of

4.

Kinetics of Compartmentalized Free-Radical Polymerization Reactions 173

compartmentalized free-radical polymerization is concerned. Possible ways


in which the rate of generation of new radicals might suddenly be reduced
to zero include (a) sudden release of an efficient radical scavenger into the
external phase of the reaction system and (b) cutting off the source of
radiation in the case of a radiation-initiated
reaction. It is convenient to
regard the elapsed time as measured from the time at which the generation
of new radicals ceased.
It is necessary to consider two distinct cases,namely, the general case of
reaction systems for which m is nonzero, and the special case for which
m = O. It will be recalled that m denotes the ratio kv/k" Le., k/X [see

Eq. (20)J.Thus, a nonzero value of m implies that radicals can be lost from
reaction loci by first-order processes,whereasa zero value of m impliesthat
either radicals cannot be lost from reaction loci by first-order processes or
k, is truly infinite. The latter possibility seems physically unlikely.
Whatever the value of m, the equation to be solved is Eq. (17) with (1set
equal to zero, Le.,
(46)
subject to an appropriate

boundary

condition.

In any particular

case, the

boundary condition is determined by the particular distribution of locus


populations that prevailed at the instant when the formation of new
radicals in the external phase ceased. Equation (46) can be solved by the
method of separation of variables; full details are given in the paper by
Birtwistle et al. (1979). The function 'P(~, t) is assumed to be of the form
E(~). T(t). The function T(t) is found to h~ve the form

T(t) = Ae).x

(47)

where A.is the "separation constant" and A is a constant whose value may
depend upon that of A..It is necessary to give careful consideration to the

values of

A. that

are physically acceptable. It turns out that these are all

negative and are of the form - x(m + b)(b

+ 1) where b =

0,1,2,

....

The equation which the function E(~)must obey is found to be


(48)
By means of the substitution x = (1 + W2, this equation can be readily
transformed to
x(1

x) d2E/dx2 + m(1 - x) dE/dx - (A./X)E= O

(49)

This differential equation is of hypergeometric form, and so solutions in


terms of hypergeometric functions are readily available. It is at this point
that it is necessary to distinguish between the cases m # O and m = O. The
latter case requires separate treatment.

174

D. C. Blackley

1. Casein which m is Not Zero


For this case, the result for 'I'(~, t) is found to be
00

(50)

BpJr 1.m-l)(~)e-t/tp

'I'(~, t) =

p=o

where 7:pis defined by the equation


. 7:p= 1j[Xp(P

+m-

1)]

(51)

p = O, 1,2, ...

the Jpare Jacobi polynomials, and the Bpare coefficients whose values are
determined by the requirement that 'I'(~, t) at zero time must have a
particular form 'I'(~,O). The particular form for 'I'(~,O) in any given instance
is that which characterizes
the distribution
of locus populations
which
prevailed at the instant when the generation of new radicals ceased. The
general results obtained for v(t), T(t), and M(t) are as follows:

1
v(t) =

2.,1.

00

.'Bip
p='

(52)

- l);1~:/,m+-l)(O)e-t/tp

00

T(t) =

t Bp(p + m -

(53)

l)e-t/tp

p=l

and
00

M(t) = M(O) + tkp[M] N

Bp7:p(p+ m p=l

1)(1-

(54)

e-t/tp)

where a symbol of the type (U)Kis a Pochhammer symbol denoting the


function r(u + K)jr(U). Equation (50) shows that the nature of the decay of
the reaction following the cessation of the generation of new radicals is
characterized by, inter aUa,a set of relaxation times 7:p' The number of these
relaxation times is in general infinite, although there may be special cases
for which the number is effectively finite. The values of 7:pdo not depend on
the nature of the initial distribution of locus populations from which the
reaction system is decaying, insofar as the nature of the initial distribution
does not affect the allowable values of p. Rather, tp are a property of the
reaction system itself and not of any particular distribution of locus
populations from which it may be decaying. It should be noted that one of
the relaxation times (that corresponding to p = O)is infinite in value, and its
presence ensures that a second boundary condition which obviously applies
in this case, namely 'I'(~, (0) = 1, can be fulfilled.
Birtwistle et al. (1979) have applied the general solution embodied in
Eqs. (50-53) to the decay of the reaction from the following three types of
\

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 175


initial distribution of locus populations: (i) a distribution of the
Stockmayer-O'Toole type, for which 'P(~, O) is given by Eq. (26); (ii) a
Poisson distribution, for which 'P(~, O)is of the form ell(~-)where J.lis the
parameter of the distribution; and (iii) a ..homogeneous" or ..uniform "
distribution, Le., a distribution in which all the reaction loci contain the
same number of propagating radicals. For the case of decay from an initial
distribution of locus populations which is of the Stockmayer-O'Toole type,
it is found that the coefficients Bp are given by

= 2p +

2.

1 12p+m-(h)

p + m- 1

(55)

Im-(h)

Case in which m = O

Setting m

= O in

Eq. (48) gives


d2,:=

A.

(1 - ~2) d~-; - xE

(56)

=O

as the differential equation that must now be satisfied by the function E(~).
This equation is a special case of the Gegenbauer differential equation. The
result obtained for 'P(~, t) is
C()

(57)

L KpC~-/2)(~)e-t/tp

'P(~,t) =

p=o

where "pare defined by the equation


1

..

Xp(p

(58)

p = O,1, 2, ...

1)

q-/2)(~) is a Gegenbauer polynomial of order p and parameter -!, and Kp


are coefficients whose values are determined by the form of 'P(~, O). The
general results for v(t),I(t) and M(t) for the case m = Oare as follows:

"
C()

.
( 1)(P-/)/2

v(t)= l'l. P=I


L, . Kp -

r[(p + i - 1)/2]

e -t/tp

r(i - 1/2)r([(p - i)/2] + 1)

(59)

where the summation extends over all values of p ~ i such that p - i is even
C()

I(t)

=-

L Kpe-t/tp

(60)

p=!

and
(61)

176

D. C. Blackley

The separate term for p = 1 in Eq. (61) arises because '1 is infinite in this
case, as well as 'o.
There is a very simple physical consideration that underlies the need for
separate treatment in the case m = O,and which explains in particular the
fact that the summation for v(t) given in Eq. (59)does not contain terms
that correspond

to p

i odd. It is that, if m truly is zero, and this situation

does not arise because k, is infinite, then there is no mechanism by which


the number of radicals in a reaction locus can alter by other than - 2 (it
being recalled that in the type of reaction system under consideration here
no new radicals are being generated). There is thus an essential physical
discontinuity in passing from reaction systems for which m is small but
nonzero to those for which m is truly zero. One consequence is that loci that
initially contained an even number of propagating radicals must continue
always to contain an even number of radicals (inc1uding no radical s at all);
likewise "odd" loci must always remain "odd." It is this feature that gives
rise to the absence of terms corresponding to p - i od!=iin the expression for
v(t).
Again, this general result has been applied to the decay of the reaction
from initial locus-population distributions of the Stockmayer-O'Toole,
Poisson, and homogeneous types. For the case of decay from a
Stockmayer-O'Toole distribution, the coefficients Kp are found to be given
by
Kp

= -(p - !)[I2p-l(h)/L l(h)J

(62)

It should be noted that this result does not follow from Eq. (55) by setting
m = O. However, the 'p for the case m = O can be obtained from those for
the case m # O by putting m

C.

= O.

Approximate" Poissonian" Solution to the General Case

Although it has so far not been possible to obtain a completely general


explicit analytic solution to Eq. (17) [and therefore to the set of equations in
Eq. (2)J for the case of reaction systems initially devoid of radicals and in
the external phase of which radicals suddenly begin to be generated at a
constant rateo Birtwistle and Blackley (1981a) have shown that it is possible to obtain an explicit analytic solution to a modified form of Eq. (17)
which under certain circumstances can be expected to be a reasonable
approximation of Eq. (17). The modified equation is
o\f!/ot

a(~

1)\f!

+ k(l - ~) o\f!/o~ + X(1- ~) o2\f!/O~2

(63)

Le., in the third term of the right-hand side of Eq. (17), the factor ~ is
replaced by the factor 1 -~. Some mathematical justification for the

4.

Kinetics of Compartmentalized Free-Radical Polymerization Reactions 177

approximation can be given if it is noted that (i) the use which is made of
the function 'P(~, t) is such that it is required at the extremities of the range
O::;;~ ::;;1 only, (ii) over this range 1 ~ is a reasonably good approximation for 1 - ~2, and (iii) 1 - ~ and 1 - ~2 have the same values at the
extremities of this range. Equation (63) can be obtained from a modified set
of time-dependent Smith-Ewart differential difference equations in which
the coefficient of kt/v in the final term on the right-hand side of Eq. (2) is
replaced by {(i + 2)(i+ 1)ni+2-(i+ l)ini+}' The modified set ofequations is thus expected to be a reasonable approximation, provided that the
rate of loss of radical s from reaction loci by second-order processes is not so
great as to be in effect the dominant radical-los s mechanism.
The solution of Eq. (63) subject to the boundary condition 'P(~, O)= 1 is
found to be

'P(~, t) = e(~-l)<I>(I)

(64)

where fjJ(t)is a function of time which satisfies the ordinary differential


equation
(65)
dfjJ(t)/dt= a - kfjJ(t) - X{fjJ(t)}2
and is subject to the initial condition fjJ(O)= O. Solution of Eq. (65) gives the
following explicit form for fjJ(t):
,"

'f'(t) = 2 a

tanh(at/2)

(66)

where a = (4aX + k2)1/2.


The function 'P(~, t) is immediately iecognizable as the frequencygenerating function .of a Poisson distribution whose parameter is fjJ(t).It
therefore follows that, for reaction systems whose behavior is described by
Eq. (63),the distribution oflocus populations is Poissonian at all times; this
is the justification for describing the approximation as "Poissonian." Thus,
the conclusion reached concerning the distribution of locus populations in
reaction systems for which X = O has been shown to be of much wider
applicability in that this will be approximately the case for all reaction
systems for which Eq. (63) is an adequate representation of behavior. Since
the parameter of the distribution is always the function fjJ(t),it follows that
I(t) and v(t)are given by equations analogous to Eqs. (42) and (43) in which
fjJ(t)replaces O(t).
D.

Other Approximations

Hawkett et al. (1975) ha ve used the matrix method to obtain approximate predictions of the nonsteady-state behavior of reaction systems for

178

D. C. Blackley

which radical loss from reaction loci is predominantly by second-order


processes. Loss by first-order processes, such as radical exit, is assumed to
be negligible. Reference has been made (see Section V,A) to the paper by
Gilbert and Napper (1974) in which the matrix method was used to obtain
approximate predictions of the nonsteady-state behavior of reaction systems for which radical loss is predominantly by first-order processes, the
rate of loss by second-order processes being negligible by comparison. The
paper to which r~ference is now made pro vides approximate solutions for
what is in effect the complementary problem. In particular, solutions are
given for reaction systems for which no radicalloss by first-order processes
occurs at all (Le., k = O). To the extent that the rate of radical loss by
second-order processes is large relative to the rate of radical entry (Le.,
a/x ~ 1), then such reaction systems should conform to Case 11 of the
classification scheme of Smith and Ewart (1948). For such systems, it is
known that 1(00)= -0.5. The results which Hawkett et al. (1975) give for
vo(t),Vl(t), and v2(t) for this latter case are as follows: .
vo(t) =!<1 -

P + P2)

+ 4(1 +

P+

p2)e-2at -

p2e-(2x+a-a2/20X)t

Vl(t) =!<1 - i~2) -!<1 - ip2)e-2at

(67)

V2(t)= !<p - P2) - !<p + p2)e-2at + p2e-(2x+a-a2/20x)t


where p = a/2x.These predictions are illustrated graphically in Fig. 9 for a
reaction system for which a = 1 x 10-3 sec-1 and X = 1 X 103sec-1. It can
be seen that for such a reaction system (i) the steady state is attained after
the lapse of approximately 2 x 103 sec, (ii) in the steady state, both Voand VI
are approximately 0.5, and (iii) at all times V2is very much less than either
Voor VI' Le., the fraction of loci that contain two propagatinK radicals is
always very small. Hawkett et al. (1975) also give an expression for
monomer conversion as a function of time which, in the case where p2 is
negligible, reduces to
M(t) = A[(l + a/x)/a ](2at + e- 2at- 1)

(68)

where A is constant. With regard to functional dependence on time, this is


of the same form as Eq. (39) for reaction systems in which radical loss is
exclusively by first-order processes. [Equation (39) can also be derived by
using the matrix approach.] The conclusion is therefore reached that it is
unlikely that it will pro ve possible to discriminate between these two very
different types of reaction system on the bases of the shapes of plots of M(t)
as a function of t.
Hawkett et al. (1977) have also succeeded in obtaining approximate
predictions of the nonsteady-state behavior of completely general reaction

40 Kinetics 01 Compartmentalized

Free-Radical Polymerization Reactions

179

100

-=
;;;- 005

i=2-

30

60

lime {mio}
Fig. 9.
Fractional locus populations, v~t), as functions of time t for i = 0,1,2 for
reaction system for which radical loss from reaction loci is exclusiveiy by second-order
processes. Values taken for (f and X are l x 10- 3 sec- and 1 x 103, respectively. The ordinates
for i = O and i = 1 are vo(t) and v(t), respectively; those for i = 2 are 2 x 106V2(t).
(Reproduced

with permission

of Jo Chemo Soco Faraday lo)

systems using the matrix method. They agree that the locus-population
generating function method is more dir~ct than the matrix method, but
correctly point out that treatment of the general case by the locuspopulation geneniting function method has so far not been possible. For
the case where all the loci are initially devoid of propagating radical s and
(f ~ (k + X), the general solution is given by Hawkett et al. (1977) as
co

v(t)

= L l~r~eAp'

p=o

(69)

where l~ is the zerotb component of the pth left eigenvector of the matrix
[O], r~ is the ith component of the pth right eigenvector of [O], and Ap is
the pth eigenvalue of [O]. The expression given for M(t) is
(70)
where A is a constant. Thus, the problem of obtaining predictions for the
v(t)and for M(t) reduces to that of finding the eigenvalues and eigenvectors

180

D. C. Blackley

of En]. The method which Hawkett at al. (1977) use is to express the matrix
En] in tHeform
En]

= [no] + 0"[n1]

(71)

where

-1

O
-1

[O,]

O O oo.
O O oo.

1 -1

O ...

(72)

[
and [no] is the matrix whose elements are obtained by subtracting O"times
those of [na from the corresponding elements of En]. Since O"is assumed
to be small compared with k and X, the eigenvalues and eigenvectors of En]
can be obtained by using non-Hermitian perturbation theory and treating
0"[n1] as a perturbation on [no]. Hawkett et al. consider that the restriction O"~ (k + X) is fulfilled by most, though not all, compartmentalized
free-radical polymerization re'actions in which the external phase is
aqueous.
Of the somewhat grosser approximations, consideration will be given
first to the "two-state" model of Lichti et al. (1977). Reaction systems that
conform to this model are those in which very few loci ever contain more
than one radical, Le., the rate of entry of radical s into loci is much less than
the sum of the rates of radical loss by first- and second-order processes.
Thus, the only significant populations that have to be taken into account in
calculating l(t) are those of loci for which i = O and 1. The number of loci
that contain three or more radicals is assumed to be strictly zero. The
number of loci containing two radical s is assumed to be very small and
stationary. Although the number of such loci is assumed to be insignificant
with regard to contribution to 1(t), the assumption that dn2/dt is zero can be
used to eliminate n2 from the expressions for dnjdt and dn2/dt. The
following treatment of the model is essentially that of Birtwistle and
Blackley (1981a). Setting n3 = n4 = O in the equation for dn2/dt [Le.,
Eq. (2) with i = 2], and setting dn2/dt = O,gives
n2 = O"nj[O"
+ 2(k + X)]

(73)

from which it is evident that n2 ~ n1 implies that O"~ O"+ 2(k + X),and thus
that O"can be neglected relative to 2(k + X). Substituting for n2 from Eq. (73)
into the equations for dno/dt and dnjdt [Le., Eq. (2) with i = O and i = 1,
respectively] gives equations for dl1o/dt and dnjdt which can be expressed
in matrix form as

~
dt

no

[J
n1

O" -q

no

[- O" qJ[ n1J

(74)

4. Kineticsof Compartmentalized Free-Radical

Polymerization

Reactions

181

The general solution of this differential matrix equation is known to be of


the form
...
no(t)

[ ]

= c [ A ] e)."

nl(t)

+c

[A ]
2

(75)

e).21

where Al and A.2are the eigenvalues of the matrix of coefficients in Eq. (74),
[Al] and [A2] are the corresponding column eigenvectors, and CI and C2
are arbitrary constants. The following results for vo(t), VI(t), and f(t) are
obtained by this method:
vo(t)
f(t)

=~

O'+q

= VI(t)

1+

(76)

~e-(<1+q)/

q
O'

= -{1-

(77)

e-(dq)/}

O'+q

= k + O'X/(k+ X).
Solutions for a "three-state" model have been published by Brooks

where q

(1978,1979).In this model, it is assumed that the n{t)values are strictly zero
for i > 2, so that the only locus populations that have to be considered are

'1

:+

y.

kt
Fig. 10.
Predictions for average number of radicals per locus I(t) as a function of time t
obtained by numerical solution of time-dependent Smith-Ewart differential difference equations, showing effect of increasing /. keeping (1 and k constant. Values taken for (1and k are
both 1 x 10-3 sec-'. Values taken for /. are as follows: A, O; B, 1 X 10-5 sec-'; e, 1 x 10-4
sec-I; D, 1 X 10-3 sec-t; E, 1 X 10-2 sec-'; F, I X 10-' sec-'. (Reproduced with per'mission of J. Chem. Sal'. Faraday /.)

182

D. C. Blackley

those for which i = O, 1, and 2. No assumptions are made concerning the


variation of n2(t) with t. Brooks gives explicit expressions for vo(t),Vl(t), and
V2(t)from which T(t)can be readily obtained as
(78)
Brooks regards his approximation as valid for reaction systems in which
T((0) is smalI (up to 0.4).
The matter of the decay behavior of a seeded emulsion polymerization
reaction following the cessation of the generation of new radicals in the
external phase has recently been treated by Lansdowne et al. (1980) using
the matrix approach.
Numerical procedures have been devised for solving the set of equations
in Eq. (2) in particular instances. Birtwistle and Blackley (1981b) have
described one such procedure. Examples of the results of their calculations
are shown in Figs. 10 and 11. These calculations refer to reaction systems
0.6

0.4

,-

0.2

A
time (sec)
Fig. 11 . Predictions for average number of radicals per locus, (t), as a function of time
t obtained by numerical solution of time-dependent Smith-Ewart differential difference equations, showing effect of decreasing k and increasing X, keeping u constant. Value taken for u is
1 X 10-5 sec-I. The values taken for k and X are as follows: A: k = 1 X 10-3 sec-I,
X = 1 X 10-8 sec-I; B: k = 1 x 1O-4sec-1, X = 1 X 10-6 sec-I; C: k = 1 X 10-5 sec-I,
X=lxl0-4sec-l;
D: k=lxl0-6sec-l,
X=lxl0-2sec-l;
E: k=lxl0-7sec-.,
X = 1 x 10 sec-I. (Reproduced with permission of J. Chem. SocoFaraday l.)

4.

Kinetics of Compartmentalized

Free-Radical Polymerization Reactions

183

that are initially devoid of radical s and in which, at a certain time (taken as
t = O),radicals suddenly begin to be generated in the external phase at a
constant rateo Figure 10 illustrates the effect of increasing Xon the variation
of l(t) with t, (1,and k being held constant. As expected, an increase in the
value of X leads to a reduction in the value of "t at any instant. Figure 11
shows the effect of increasing X and decreasing k on l(t) as a function of t,
with (1being held constant. As expected, for large values of X/k, the value of
i(t) at long times approaches the Smith-Ewart Case 2 value of 0.5.
Numerical solutions have also been obtained by Brooks (1980) using his
three-state model (see above). The relevant simultaneous differential equations were solved by Euler's method. Brooks has included in his examples
the case where (1 decays exponentially with time. In all the cases investigated, he finds that allowance must be made for bimolecular mutual
termination of radicals and for the re-entry of desorbed radicals into
reaction loci; he concludes that failure to take account of these possibilities
can lead to serious errors.

VI.

Predictions for Molecular-Weight Distribution and Locus-Size


Distribution

This chapter has been concerned exclusively with predictions for the
distribution of locus populations within a compartmentalized free-radical
polymerization reaction system. Other matters of considerable interest are
the distribution of polymer molecular weights which the polymerization
reaction produces, and the distribution of sizes of the reaction loci at the end
of the reaction. A significant literature concerning both these aspects is
beginning to develop, but beca use of the complexity of the subject and
limited space, oniy a brief summary of the various contributions can be
given here.
The distribution of molecular sizes produced by a given addition polymerization reaction is determined by the balance of processes such as
propagation, termination, transfer; branching, and cross-linking. In fact, the
final polymer molecule produced is the "corpse" of the kinetic chain by
which it was produced, as modified by transfer, branching, and crosslinking reactions. Early attempts to consider the distribution of molecular
weights produced by a compartmentalized free-radical polymerization reaction were made by Katz et al. (1969) and by Saidel and Katz (1969).
These workers derived a set of coupled partial differential equations, from
the solution of which the molecular-weight distribution can, in principIe, be
predicted. However, except for a few simple cases, solutions were restricted
to predicting the lower moments of the molecular-weight distribution, and

184

D. C. Blackley

even these predictions required moderately extensive numerical calculations. Furthermore, the treatment did not take into account several
important molecular events, such as the possibility of mutual termination
with disproportionation, and the various transfer reactions. An alternative
treatment, given by Min and Ray (1968), is also restricted to the prediction
of the lower moments of the molecular-weight distribution.
More recently, Sundberg and Eliassen (1971) have attempted a prediction of molecular-weight distribution for a reaction taking place isothermally in a well-mixed batch reactor. Micelles are assumed to be the only
source of reaction loci. Both micelles and loci receive radicals at arate
proportional to surface area. All the radicals produced in the external phase
are absorbed by micelles and loci. Termination takes place immediately
when two radical s become present in a locus and combination is the
outcome. Allowance is made for the reaction of transfer to monomer. The
predicted molecular-weight distribution is of the Flory "most probable"
type with a polydispersity ratio of 2.0. As might be expected, the value of
the rate coefficient for transfer to monomer relative to that for propagation
has an important effect on both the average molecular weight and the distribution of molecular weights.
An extensive treatment of this subject has been given very recently by
Lichti et al. (1980),and a brief summary was given in an earlier paper
(Lichti et al., 1978).The model assumed for this treatment is a three-state
model in which i is O,1,or 2. An earlier paper (Lichti et al., 1977)applied a
similar treatment to a two-state model in which i is O or 1. The treatment
allows for the possibility that mutual termination may result in either
combination or disproportionation. It also allows for the possibility of
transfer to monomer. It has not, however, been possible to make allowance
for branching and cross-linking. Prediction of the full distribution of
molecular sizes, and not merely of particular moments of the distribution,
has been achieved. The conclusion has been reached that compartmentalization of the reaction leads to a broadening of the molecular-weight
distribution.
An early contribution to the cognate matter of the prediction of the
eventual distribution of locus sizes was made by Ewart and Carr (1954).
The two principal factors that affect the nature of this distribution were
considered to be (i) the distribution of locus sizes present when locus
nucleation ceases, and (ii) the manner in which the loci grow subsequently as
a consequence of polymerization. A third factor that may influence the
growth pattern is the tendency for large loci to imbibe more monomer than
do smallloci.
O'Toole (1969) has discussed the extent of stochastic (as contrasted with
deterministic) con tribu tion s to the polydispersity of locus sizes. He con-

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 185


eludes that for moderately large loci, stochastic contributions are significant
if the ratio of propagation to termination rate coefficients is greater than
about 0.1. Sundberg and Eliassen (1971) have also made predictions of
locus-size distribution using their modeI. The subject has also been discussed by Watterson and Parts (1971a,b).
BirtwistIe and Blackley (in press) ha ve recently applied the locuspopulation generating function approach to the problem of the evolution of
the locus-size distribution in compartmentalized free-radical polymerization
reactions. They have introduced a generalized locus-population generating
function 'P(~, t, v), defined as
00

'P(~, t, v) =

v(t, v)~

=o

(79)

where the function v(t,v) is such that v(t,V)DVis the fraction of the total
number of reaction loci which at time t contain i propagating radical s and
which also ha ve a volume that lies between v and v + DV.The relationship
between this locus-population generating function and the generating function 'P(~, t) which has been used previously is elearly
'P(~, t) =

Vll.

(80)
Loo

'P(~, t, v) dv

Theory for Generation of Radicals in Pairs within Loci

This problem was first treated in detail by Haward (1949). He considered


the case of a bulk' polymerization that has been compartmentalized by
subdividing the reaction system into a large number of separate droplets,
each of volume v. Radicals are generated exelusively within the droplets and
always in pairs. An example would be the polymerization of styrene in
emulsified droplets dispersed in water initiated by the thermal decomposition of an oil-soluble initiator which partitions almost exelusively
within the monomer droplets, In the model considered by Haward, radicals
are unable to exit from the droplets into the external phase. The only
radical-Ioss process is in fact bimolecular mutual termination. It therefore
follows that all the droplets must always contain an even number (ineluding
zero) of propagating radicals, and that the state of radical occupancy will
change in increments of :t2. The conelusion reached by Haward is that in
this case the effect of compartmentalization is to reduce the overall rate of
polymerization per unit volume of disperse 'phase. The physical reason for
this is that, as the volume of the droplets is reduced, so are the opportunities for a radical to escape from the others-and hence to avoid mutual

186

D. C. Blackley

termination. Compartmentalization therefore enhances the rate of termination relative to propagation, and therefore reduces the rate of polymerization; indeed, according to Haward, polymerization can be effectively
suppressed altogether if the degree of compartmentalization is sufficiently
great. The average degree of polymerization of the polymer produced is also
reduced by compartmentalization. The expression derived by Haward for
the overall rate of polymerization per unit volume of disperse phase in the
steady state is as foUows:
d[M]/dt

-kp[MJX/2k\)1/2

tanh{(X/2k\)1/2vNA}

(81)

where X is the rate of formation of new radicals containing one monomer


unit, in unit volume of monomer, k\ is the rate coefficient for bimolecular
mutual termination, and NAis the Avogadro number. Comparison with the
corresponding equation for bulk polymerization without subdivision shows
that the effect of compartmentalization
is expressed by the factor
tanh(vNAi.jX/2k,), and this is, of course, always less than unity.
A more detailed treatment of, this problem has been given by O'Toole
(1965). In this treatment, allowance is made for the possibility that radical s
can exit from the loci into the external phase. However, no allowance is
made for the possibility that radicals which do so exit may reenter the
reaction loci and reinitiate polymerization there. The transitions which
affect the numbers of loci which cross a notional boundary between states
of radical occupancy i and i + 1 are illustrated in Fig. 12. The condition for
the steady state is readily found to be
k(i + l)n+1 + x(i + l)in+1 + x(i + 2)(i + l)n+2 = !CT(n + n-I)
{j+2)(i+l)n+2X

(i+l)ni+lk

(i+ I)lnl+lx

(82)

1+2

i+1

---- -- - - -- -- - 1--- - --ncr/2


i-I
nl-lcr/2
Fig. 12.
Transitions between states of radical occupancy required for derivation of
OToole equation for rate of transition of loci across notional barrier between states i and i + 1
in the case where radicals are generated in pairs within reaction loci.

4. Kineticsof Compartmentalized Free-Radical Polymerization Reactions 187


where (j is now the rate of formation of single radicals within a single locus,
and therefore (j/2 is the rate of formation of pairs of radicals within a single
locus. The remaining symbols have their previous significances. Introducirig
the parameters 8 and m as defined previously, this equation becomes

(m + i)(i + l)n+1 + (i + 2)(i + l)n+2= t8(n + n-l)

(83)

An equation of this type must be satisfied at notional boundaries between


each pair of neighboring states of radical occupancy, if the reaction system
as a whole is to be in a steady state. By introducing the locus-population
generating function 'P(~) as defined previously, this infinite set of equations
can be combined together to give the single equation
d2'P/d~2 = Cm/O

+ ~)J d'P/d~ - (8/2)'1' = O

(84)

The solution of this equation consistent with the physical constraints of the
reaction system is
l-m

'P(~) = A(l

+ ~)-r-Im;1 { 4(1 + ~)}

(85)

where A is a normalizing constant and h is as defined previously. The result


obtained for Viis

where the combinatorial symbols have their usual meanings, and j is zero
for even values of i.and unity for odd values of i. The average number. of
propagating radical s per locus is found to be given by

-1 =-h I(m+ 1)/2(h/2)


4I(m-1)/2(h/2)

(87)

For the case where loss of radical s by first-order processes is of negligible


significance, the expression for 't'assumes the simple form
1=

ih tanh(h/4)

List of Symbols
A
[AJ,(AzJ
a
B
Bp

constant
column eigenvectors
(4<TX+ kZ)IIZ or surface area of spherical reaction locus
constant
constant associated with term containing e-ti..

(88)

188
b
q-I/21(~)
nCr
e
C,C2
h
l.
;
1

Jt..lm
j
Kp
k
ko
kp
kt
ktE
k,p
[o
p
M
[M]
m
N
NA
ni

D. C. Blackley
0,1,2, ...
of order p and parameter

stationary

of radicals in external phase

concentration

constants associated with [A ] and [A 2]


(80)1/2
modified Bessel function of order t
number of propagating radicals in single reaction locus
average number of propagating
radicals in single reaction
Jacobi polynomial of order p
row of matrix [O]

rate coefficient for propagation


rate coefficient for bimolecular

mutual

rate coefficient for bimolecular


rate coefficient for bimolecular

mutual termination
mutual termination

concentration
kvlkt

of monomer

k+~

x
y

of matrix [O]

locus

number of reaction loci in unit volume of reaction system


Avogadro number
number of loci in unit volume of reaction system which contain ; propagating
radicals
matrix of ni values

T(t)
t
V
v
X

in external phase
in reaction loci

= klx

0,1,2, oo.

rip

at.reaction

loci

termination

zeroth component of pth left eigenvector


amount of monomer

locus

constant associated with term containing e-ti'.


coefficient characterizing
rate of exit of radicals from reaction
kvla

[n;]

R-i+l

-1-

Gegenbauer polynomial
combinatorial
symbol

k+X
rate of transition of loci across national boundary between states of radical
occupancy ; and ; + 1
radius of sphericaI reaction locus
ith component of pth right eigenvector of matrix [O]
maximum number of propagating radicals in single reaction locus
part of 'I'(~, 1) which depends only on 1
time
variance of distribution of locus populations
volume of reaction locus
rate of formation of new radicals containing one monomer unit, in unit volume of
monomer
2';6(1 + ~) o/' (1 + W2
'I'(~)

xl-m

Z
ex

2Nk,pk,Jy2v
see (10

p
r

(1/2X

pIe

gamma function;

(u). is a Pochhammer

symbol which denotes r(u + K)/r(u)

4. Kinetics of Compartmentalized Free-Radical Polymerization Reactions 189


o for evenvaluesof i and 1for odd valuesof i
uv/k,= u/X

e
e'

u'v/k,

O(t)

).

AI,A2
Ap
J1
Vi

E(~)

~
p
p'
u
u'
Uo and C(

,
'p
q,(t)
X

'P
[n]

= u'/x

time-dependent
parameter of Poisson distribution
separation constant
eigenvalues associated with [A tJ, [A2]
pth eigenvalue of matrix [n]
parameter of initial Poisson distribution of locus populations
fraction of loci which contain i propagating radicals
part of 'I'(~, t) which depends only on ~
auxiliary variable of locus-population
generating function
rate of entry of radicals into all reaction loci in unit volume of reaction system
rate of formation of new acquirable radicals in unit volume of reaction system
average rate of entry of radicals into a single locus reaction locus, i.e., p/N, or rate
of formation of radicals within a single reaction locus
p'/N
are such that u = u o r"
average number of propagating radicals in single reaction locus
I/XP(P + m - 1)
time-dependent
parameter of Poisson distribution
kJv
locus-population
generating function
matrix of coefficients

[no], [nI] are such that [n]

= [no]

+ [ntJ

References
Allen, P. W. (1958). J. Col/od Sci. 13,483-487.
Azad, A. R. M., and Fitch, R. M. (1980). Paper presented

at Am. Chem. Soco Emulsion

Symp.. Las Vegas Pr~prints, pp. 537-542.


Birtwistle, D. T., and Blackley, D. C. (1977). J. Chem. Soco Faraday
Birtwistle, D. T., and Blackley, D. C. (1979). J. Chem. Soco Faraday
Birtwistle, D. T., and Blackley, D. C. (l98Ia). J. Chem. Soco Faraday
Birtwistle, D. T., and Blackley, D. C. (198Ib). J. Chem. Soco Faraday

Polym.

173, 1998-2009.
175, 2051-2064.

177, 413-426.
177,1351-1358.
Birtwistle, D. T., Blackley, D. c., and JefTers, E. F. (1979). J. Chem. Soco Faraday / 75,
2332-2358.
Blackley, D. C. (1975). "Emulsion
Polymerisation:
Theory and Practice,"
Science, London.
Blackley, D. C., and Haynes, A. C. (1977). Br. Polym. J. 9, 312-321.
Brooks,

B. W. (1978). J. Chem. Soco Faraday /74,

p. 95. Applied

3022-3056.

Brooks, B. W. (1979).J. Chem.SocoFaraday /75, 2235.


Brooks, B. W. (1980). J. Chem. Soco Faraday 176, 1599-1605.
Ewart, R. H., and Carr, C. 1. (1954). J. Phys. Chem. 58, 640-644.
Fitch, R. M. (1973). Br. Polym. J. 5, 467-483.
Fitch, R. M., and Tsai, C. H. (l97Ia).
Plenum Press, New York.

"Polymer

Colloids"

(R. M. Fitch, ed.), pp. 73-102.

Fitch, R. M., and Tsai, C. H. (I97Ib).


Plenurn Press, New York.

"Polyrner

Colloids"

(R. M. Fitch, ed.), pp. 103-116.

Friis, N., and Hamielec, A. E. (1973). J. Polym. Sci. Polym. Chem. Ed.2,3321-3325.

190

D. C. Blackley.

Gardon,
Gardon,

J. L. (1968). J. Po/ym. Sci. Part A-/6, 623-641.


J. L. (1973). J. Po/ym. Sci. Po/ym. Chem. Ed. 2, 241-251.

Gilbert,

R. G., and Napper,

D. H. (1974). J. Chem. Soco Faraday 170, 391-399.

Gilbert, R. G., Napper, D. H., Lichti, G., Ballard, M., and Sangster, D. F. (1980). Paper
presented at Am. Chem. Soco Emu/sion Po/ym. Symp., Las Vegas Preprints, pp. 527-530.
Grancio, M. R., and Williams, D. J. (1970). J. Po/ym. Sci. Part A-1S, 2617-2629.
Hansen, F. K., and Ugelstad, J. (1978). J. Po/ym. Sei. Po/ym. Chem. Ed. 16, 1953-1979.
Hanseii, F. K., and Ugelstad, J. (1979a). J. Po/ym. Sei. Po/ym. Chem. Ed. 17,3033-3046.
Hansen, F. K., and Ugelstad, J. (1979b). J. Po/ym. Sci. Po/ym. Chem. Ed. 17, 3047-3068.
Haward,

R. N. (1949).

Hawkett,

B. S., Napper,

Po/ym.

Sci.

4, 273-287.

D. H., and Gilbert, R. G. (1975). J. Chem. Soco Faraday 171,2288-2295.

Hawkett, B. S., Napper, D. H., and Gilbert, R. G. (1977). J. Chem. Soco Faraday 173,690-698.
Hawkett, B. S., Napper, D. H., and Gilbert, R. G. (1980). J. Chem. Soco Faraday 176,1323-1343.
Katz, S., Shinnar, R., and Saidel, G. M. (1969). Adv. Chem. Ser. 91, A.C.S., p. 145-157.
Kendal, M. G., and Stuart, A. (1963). In "The Advanced Theory ofStatistics,"
2nd ed., Vol. 1,
p. 30. Griffin, London.
Keusch,

P., and Williams,

Keusch, P.,
Keusch, P.,
Lansdowne,
Faraday

D. J. (1973). J. Po/ym. Sci. Po/ym. Chem. Ed. 11, 143-162.

Prince, J., and Williams, D. J. (1973). J. Macromo/. Sci.-Chem. A7(3), 623-646.


Graff, R. A., and Williams, D. J. (1974). Macromo/ecu/e~ 7,304-310.
S. W., Gilbert, R. G., Napper, D. H., and Sangster, D. F. (1980). J. Chem. Soco
176, 1344-1355.

Lichti, G., Gilbert,

R. G., and Napper,

D. H. (1977). J. Po/ym. Sci. Po/ym. Chem. Ed. 15, 1957-

1971.
Lichti, G., Gilbert, R. G., and Napper,
Latex Conf, London Paper No. 1.
Lichti, G., Gilbert,
1323.
Min, K.
Morton,
Napper,
O'Toole,
O'Toole,

R. G., and Napper,

D. H. (1978). Paper presented

at P/ast. Rubb. Inst. lnt.

D. H. (1980). J. Po/ym. Sci. Po/ym. Chem. Ed. IS, 1297-

W., and Ray, W. H. (1968). J. App/. Po/ym. Sci. 22, 89-112.


N., Kaizerman, S., and Altier, M. W. (1954). J. Col/oid Sci. 9, 300-312.
D. H. (1971). J. Po/ym. Sci. Part A-19, 2089-2091.
J. T. (1965). J. Appl. Po/ym. Sci. 9, 1291-1297.
J. T. (1969). J. Po/ym. Sci. Part C 27, 171-182.

Saidel, G. M., and Katz, S. (1969). J. Po/ym. Sci. Part C. 27, 149-169.
Smith, W. V., and Ewart, R. H. (1948). J. Chem. Phys. 16,592-599.
Stockmayer, W. H. (1957). J. Po/ym. Sci. 24, 314-317.
Sundberg, D. C., and Eliassen, J. D. (1971). In "Polyrner Colloids" (R. M. Fitch, ed.), pp.
153-161. Plenum Press, New York.
Ugelstad, J., and Hansen, F. K. (1976). Rubber Chem. Techno/. 49, 53/Hi09.
Ugelstad, J., Merk, P. c., and Aasen, J. O. (1967). J. Po/ym. Sci. Part A-I 5, 2281-2288.
van der Hoff, B. M. E. (1958). J. Po/ym. Sci. 33, 487-490.
van der Hoff, B. M. E. (1962). In "Polyrnerization
and Polycondensation
Processes," Adv. in
Chem. Ser. 34, pp. 6-31. American Chemical Society.
Vanderhoff, J. W. (1976). Proc. Water Borne Higher Solids Coating Symp. Paper No. 6.
Vanzo, E., Marchessault,
R. H., and Stannett, V. (1965). J. Col/oid Sci. 20, 62-71.
Watterson, J. G., and Parts, A. G. (1971). Makromo/. Chem. 146, 11-20.
Weiss, G. H., and Dishon, M. (1976). J. Chem. Soco Faraday 172,1342-1344.
Williams, D. J. (1971). J. E/astop/ast. 3, 187-200.
Williams,

D. J. (1973). J. Po/ym. Sci. Po/ym. Chem. Ed. 2, 301-303.

5
Mamoru Nomura

1. Introduction .
11. Polymerization Rate Equations Involving FreeRadical Desorption .
A. Reaction Locus .
B. Polymerization Rate Equations
111. Derivation of Rate Coefficient for Radical Desorption
from Particles.
A. Definition of Radical Desorption
and Reabsorption
B. Derivation of Mass- Transfer Coefficient
for Radical Desorption and Reabsorption .
C. Derivation of Rate Coefficient for Radical
Desorption
IV. Effect. of Free-Radical Desorption on the Kinetics
of Emulsion Polymerization
A. Effect on the Rate of Emulsion Polymerization.
B. Effect on Micellar Particle Formation
V. List of Symbols .
References

l.

192
192
194
199
199
202
204
210
210
214
217
219

Introduction

The kinetic behavior of ernulsion polyrnerization is greatly


atfected by radical desorption from polymer particles. This has been shown
by Ugelstad et al. (1969), Ltt et al. (1970), Harada et al. (1971), Friis and
Nyhagen (1973), and Nomura et al. (1971). It is believed that the deviation
of the kinetic behavior of the emulsion polyrnerization of water-soluble
monorners such as vinyl acetate and vinyl chloride from the Srnith and
Ewart (1949) Case 2 kinetic theory is mainly due to dominant desorption of
191
EMULSION

POLYMERIZATION

Copyright
It> 1982 by Aeademie Press. 1ne.
AIl rights of reproduetion
in any form reserved.
ISBN 0-12-556420.1

192

Mamoru Nomura

radical s from particles. Although the importance of this physical phenomenon was pointed out by Smith and Ewart in the 1940s, the quantitative
understanding was insufficient to explain the kinetic deviation of vinyl
acetate and vinyl chloride emulsion polymerizations from the Smith-Ewart
Case 2 kinetic theory.
Recently, Ugelstad et al. (1969) proposed a semiempirical rate coefficient
for radical desorption in vinyl chloride emulsion polymerization. On the
other hand, Nomur.a et al. (1971, 1976) have derived arate coefficient for
radical desorption theoretically with both stochastic and deterministic
approaches and have successfully applied it to vinyl acetate emulsion
polymerization. They also pointed out that radical desorption from the
particles and micelles played an important role in micellar particle formation. Friis et al. (1973) also derived the rate coeffieient for radical
desorption in a different way. Litt et al. (1981) discussed in more detail the
chemical reactions incorporated in the physical process of radical desorption in the emulsion polymerization of vinyl acetate..
In this chapter, the polymerization rate equations for emulsion polymerization will be reviewed briefly: Then, the rate coefficient for radical desorption from the particles will be derived theoretically, and the effect of
radical desorption on the rate of emulsion polymerization and the micellar
particle formation will be discussed.

11. Polymerization Rate Equations Involving Free-Radical Desorption


A.

Reaction Locus

Harkins (1947) and Smith and Ewart (1948) assumed that in emulsion
polymerization the polymerization loei were inside the particles because the
rate of styrene emulsion polymerization, for example, was proportional to
the number of polymer particles presento However, the polymerization loci
in the emulsion polymerization of water-soluble monomers such as vinyl
acetate and vinyl chloride have long been discussed because the rate of
emulsion polymerization of these monomers was not proportional to the
number of polymer particIes presento
Patsiga et al. (1960) and Giskehaug (1965) carried out the seeded emulsion polymerizations of vinyl acetate and of vinyl chloride, respectively,
and found that in both systems the rate of polymerization was proportional
only to the 0.15-0.20 power of the number of polymer particles. From this
finding they concIuded that the main locus of the emulsion polymerization
of water-soluble monomers must be in the water phase. This conclusion
seems reasonable because the solubilities of these monomers are about 100

5. Desorption and Reabsorption of

Free Radicals

193

times greater than that of styrene. But this inference may be excluded from
the following simple discussion.
Let us consider a seeded emulsion polymerization where no particle
formation occurs and hence the number of polymer particles is constant. At
a steady state, the overall rate of radical entry into the particles PA is
expressed by

=
PA

rate of radical production


in the water phase

[
-

J+

rate of radical desorption


from polymer particles J

rate of radical termination

in the water phase

(1)

Since the value of PA is not so different from the value of Pw, the rate of
radical production in the water phase, we get
(2)
It must be noted here that the quantities PA and Pware different from one
another.
Ir the diffusion law can be applied to the radical entry into the particles,
PA is given by

PA = ka[R~JNT~ 2ndpDw[R~JNT

(3)

where [R~J is the radical concentration in the water phase, NT is the


number of polymer particles, dp is the diameter of a particle, ka is the rate
coefficient for radical entry into the particles, and Dw is the diffusion
coefficient for the radical s in the water phase.
Using Eqs. (2) and (3), one can calculate the average residence time for a
radical to be in the water phase before entering into the particles tred, the
average time needed for a radical to add one monomer unit in the water
phase tadd, and the average time needed for a radical to be deactivated by
the mutual termination reaction in the water phase ttermfrom the following
equations:
tred= [R~J/PA ~ 1/2ndpDwNT

(4)

tadd= l/kp[MwJ

(5)

tterm= [R~J/2ktw[R~J2 ~ ndpDwNT/ktwpw

(6)

where kp is the propagation rate constant, [MwJ is the monomer concentration in the water phase, and ktwis the termination rate constant in the
water phase. The values of tred, tadd' and ttermare calculated to be -10-5,
-4 x 10-4, and - 103 sec, respectively, when the following numerical constants are used: ktw= 6 X 108 (dm3/mol sec), Pw= 1016(molecules/dm3 sec),

194

Mamoru Nomura

[Mw] = 0.5 (mol/dm3), dp = 10-7 (m), Dw= 10-9 (m2/sec), kp = 5000


(dm3/mol sec), and NT = 1017(particles/dm3 water).
Comparing the values of (red' (add, and (term,one can conclude that the
radical s in the water phase enter the particles without adding any monomer
unit or terminating with other radicals in the water phase under normal
conditions. This indicates that the pro paga tion and mutual termination
reactions in the water phase can be neglected, and hence that these
reactions occur practically in the particles under normal conditions.
B.

Polymerization Rate Equations

For emulsion polymerization systems where polymerization takes place


exclusively in the particles, the rate of emulsion polymerization can be
expressed in the same way as in other radical polymerizations

Rp

= -dM/dt = kp[Mp][R;]Mw .

(7)

where Rp is the rate of emulsio~ polymerization [kJdm3(water) sec], M is


the monomer concentration in the water phase [kJdm3 (water)], kp is the
propagation rate constant (dm3/mol sec), [Mp] is the monomer concentration in the polymer particles (mol/dm3), and Mw is the molecular weight
of monomer (kJmol). [R;] and NT can be expressed as follows in terms of
the number of polymer particles containing n radicals Nn.
<XJ

[R;] = (Ni + 2N2 + 3N3 + ... + nNn + ...)/NA


<XJ

NT=No+Ni+N2+...+Nn+"'=

Nn
n=O

n=i

nNn/NA

(8)
(9)

where NA is Avogadro's number, [R;] is given in mol/dm3 water, and Nn in


particles/dm3-water. The average number of radicals per particle is
defined by

= (~i nNn)j NT

(10)

Thus, the rate of emulsion polymerization given by Eq. (7) is also rewritten
as
Rp = kp[Mp]NTMw/NA

(11)

In order for the rate of emulsion polymerization to be predicted by Eq. (7)


or (11), Nn must be represented as a function of the properties of the
emulsion polymerization system in question.
Smith and Ewart proposed the steady-state equation for Nn in terms of
the rate of radical entry into the particles, the rate of radical desorption

195

5. Desorption and Reabsorption of Free Radicals

from the particles, and the rate of radical termination within the particles.
For the nonsteady state
dNn dt

PA

()
-( )

N T Nn-l

+ k eap

PA

n+ I

(~ )
()

N n+l

+ k tp

NT Nn - keap vp Nn - ktp [

(n + 2)(n + I) N
n+2

n(n - I)
vp

J Nn

(12)

where ke is the rate constant for radical desorption from the particles
(m/sec), ap is the surface area of a particle (m2), vp is the volume of a particle
(m3), and ktp is the rate constant for termination (m3/molecule sec). Since
the rate constant for radical desorption should depend on the radical chain
length, the rate constant ke represents an average value.
The rate coefficient for radical desorption from the particles kr is defined
as
(13)
wh~e
(14)
where kj is the rate constant

for radical desorption

from the particles for

j-

mer radical s, 'j is the ratio of the number of j-mer radicals to the total
number of radicals nNn contained in Nn particle, and j = O indicates the
initiator radical.
By solving Eq. (12) for Nn and introducing it into Eq. (10), one can
calculate the value of , and accordingly, 'the rate of emulsion polymerization by Eq. (11).. Although many investigations have been reported
concerning the solution of Eq. (12), only the principal work will be reviewed
briefly here. For more details, the readers should refer to the review article
by Ugelstad et al. (1976).
Smith and Ewart (1948) did not obtain a general solution for Eq. (12) but
rather solved it for three limiting cases, applying a steady-state hypothesis,
i.e., dNn/dt = O.
Case 1 PA/NT ~ kr
In this case, the following relation holds:
(15)
Further,
(a) When termination in the water phase is dominant (ktp/vp ~ kr,
Pw~ 2ktw[R~]2)
(16)

Mamoru Nomura

196

where md indicates the partition coefficient of radicals between the particle


and water phases defined by Eq. (38).
(b) When termination in the particles is dominant (ktP/vp~ kr,
Pw~ 2(pAiNT)N1)
= (Pw/2krNT)1/2 ~ 0.5
Case 2

(17)

kr ~ pAiNT ~ ktP/vp

= 0.5

(18)

Equation (18) is most generally known as the Smith-Ewart theory.


Case 3 pAiNT ~ ktp/vp

= (PAvp/2ktpNT)1/2 ~

0.5

(19)

Further, when the termination of radicals in the water phase or the


desorption of radicals from the particles could be neglected, Eq. (19) is
rewritten as
=.(Pwvp/2ktpNT)1/2

(20)

Stockmayer (1957) obtained a general solution for Eq. (12), applying a


steady-state hypothesis, and expressed by the modified Bessel function of
the first kind. O'Toole (1965) later extended it and gave a physically more
acceptable solution for as follows:

= -!a[Im(a)/Im-l(a)]

(21)

where
a2 = 80(

O(= PAvp/ktpNT

(22)

and
m = krvJktp

(23)

Gardon (1968) has solved Eq. (12) numerically for the case of negligible
desorption of radical s from the particles without assuming the steady state,
stating that the Stockmayer solution for is incorrect because there is no
steady state in principie and because Eq. (12) includes the time-dependent
parameter vp' However, the results of numerical calculation by Gardon
coincide almost completely with those predicted by the Stockmayer solution for no radical desorption from the particles. This also supports the
validity of applying the steady-state hypothesis to the solution for Eq. (12)
under normal conditions for emulsion polymerization.
Noting that in the treatments by Smith and Ewart, by Stockmayer, and
by O'Toole the "fa te " of the desorbed radicals was not necessarily specified,

5.

197

Desorption and Reabsorption of Free Radicals

Ugelstad et al. (1967) claimed that Eq. (12) should be solved simultaneously
with the following balance equation on the radicals in the water phase.
(24)
Since the steady-state hypothesis is applicable to Eq. (24), it is rewritten in
the same form as Eq. (1):
PA = ka[R:JNT = Pw

+ kfnNn - 2ktw[R:J2

(25)

Considering that nNn = NT, Eq. (25) is rewritten in a dimensionless form


a = a' + m - Ya2
(26)
where
a = PAVp/ktpNT
a' = Pwvp/ktpNT

(27)

m = kfVp/ktp

(23)

= 2ktwktp/k;NTvp

(28)

and
y

Ugelstad et al. (1967) solved the simultaneous equations Eqs. (21) and (26)
for and plotted the calculated value of against the value of a' at fixed
value of Y, varying the value of m as a parameter. Since the radical
termination in the water phase is negligible under normal emulsion polymerization conditions as mentioned in Section ~I,A,the condition Y = Ois most
important for usual emulsion polymerizations. Figure 1 shows a plot of
10'

y=o

IC

164

10-2

104

a'
Fig. 1.
Relationship between and IX'when termination in the water phase is neglected.
IX'= Pwvp/NTk,pand m = kfVp/k,p. (Reprinted with permission of Journal of PolyrnerScience.)

198

Mamoru Nomura

log versus logO(', varying the value of m for the case of Y = O. The
treatment by Ugelstad et al. (1967) is most general and correct, and hence,
Eqs. (21) and (26) are important in predicting the value of , that is, the rate
of emulsion polymerization.
Since it is inconvenient to use Eqs. (21) and (26) directly, however, several
empirical or approximate equations for are derived for the case where
Y=O.
1. When m = O

1/2

0('

( )

n= -+4 2

(29)

2. When instantaneous termination is dominating in the particles and


< 0.5

= !(- e + J c2 + 2e)

e = O('/m= Pw/kfNT

(30)

Equation (30) leads to

= 0.5

(31)

when e -4 00, and when e < 10-2


= (Cj2)1/2

(32)

3. When < 0.2


-

0('

[(

n= -

4.

1+-

1 1/2
m)J

(33)

When m -4 00 or ~ 0.5
= (0('/2)1/2

(34)

Equation (29) is an empirical equation presented by Ugelstad and Merk


(1970) and may be useful in styrene emulsion polymerization because the
value of mis about 10-\ or less than that under normal reaction conditions
in this system. On the other hand, Eqs. (30) and (33) are applicable to vinyl
acetate and vinyl chloride emulsion polymerizations (Ugelstad et al., 1969;
Harada et al., 1971; Friis and Nyhagen, 1973). Equation (34) explains very
well the rate of aqueous dispersion polymerization of vinyl acetate in the
absenceof emulsifier(Nomura et al., 1978).
It is clear from the discussion so far that as long as the value of the rate
coefficient for radical desorption from the particles kf cannot be estimated
quantitatively, the rate of emulsion polymerization is impossible to predict.
In the next section, therefore, the quantitative expression for kf will be
derived.

5. Desorption and Reabsorption of Free Radicals

199

111. Derivation of Rate Coefficient for Radical Desorption from Particles


Ugelstad et al. (1969) first suggested experimentally that in vinyl chloride
emulsion polymerization kf may be expressed in the following form
kf

= kv-2/3
= k'd-2
p
p

(35)

Nomura et al. (1971) also discussed this problem almost at the same time
and derived the theoretical expression for kf, which was inversely proportional to the square of particle diameter, as shown by Eq. (35). In this
section, the derivation of kf by the Nomura and Harada method is mainly
explained (Harada et al., 1971; Nomura et al., 1976; Nomura and Harada,
1981).
A.

Definition 01 Radical Desorption and Reabsorption

As shown in Fig. 1, the value of becomes independent of the value of kf


in the range of ~ 0.5. This means that the rate coefficient for radical
desorption from the particles is important in the range ( < 0.5) where the
polymer particle contains at most one radical. For this reason, we consider
an emulsion polymerization system where (i) the particles contain at most
one radical and (ii) instantaneous termination takes place when another
radical enters the particle that already contains a radical.
In this case, the rate coefficients for radical desorption and reabsorption
are defined in the following balance equation on the active particles
containing one radical
(36)
where N* is the number of active polymer particles containing one radical,
ka is the rate coefficient for radical entry into the particles, and No is the
number of dead particles containing no radicals. The first term on the righthand side of Eq. (36) shows the rate of decrease in the number of active
polymer particles due to radical desorption from the particles containing a
radical. The second term usually expresses the rate of increase in the
number of active particles containing two radicals if instantaneous termination does not occur when another radical enters the particle already
containing one radical. However, since we supposed that instantaneous
termination does occur in this case, this term represents the rate of decrease
in the number of active particles containing a radical, and accordingly,
should be negative. The third term indicates the rate of increase in the
number of active particles by radical entry into the dead particles.
Let us consider the diffusion of radicals. According to the two-film
theory developed by Lewis and Whitman (1924) for mass transfer across the

200

Mamoru
INTERFACE
I
WATER PHASE

PARTlCLE

-DISTANCE

Fig. 2.
(Reprinted

Nomura

Schematic diagram of concentration


gradients near phase boundary.
with permission of Journal of Applied Polymer Science.)

Cp

= n/vp.

interface between two phases, the concentration gradients near the phase
boundary can be assumed to' be as shown in Fig. 2. Applying Fick's
diffusion law, the rate of radical desorption from a single particle that
contains, for example, n radicals with same chain length, is generally
expressed by
JRo= -d(vpCp)/dt = -dn/dt = ksap(Cp- Cpi)= kwap(Cwi- Cw) (37)
where ks and kw, respectively, denote the film mass-transfer coefficients for
the inner and outer diffusion films adjacent to the interface between the
particle and water phases, Cp is the concentration of radical s in the particle,
Cw is the concentration of radical s in the water phase and i denotes the
interface. The concentration of radicals at the interface, which is at equilibrium and usually very low, can be expressed by the linear relationship ,
Cpi

= rnd Cwi

(38)

whererndis the partition coefficient.


Using Eqs. (37)and (38),we have
J. =
RO

dn =
Cp
dt (l/ksap)

- rndCw

+ (rnd/kwap) -

Ksap(Cp

rndCw)

(39)

(40)

here

l/ks + rnd/kw

(41)

l/Kw = l/rndks + l/kw

(42)

l/Ks

5. Desorption and Reabsorption of Free Radicals

201

where Ks and Kw are the overall mass-transfer coefficients. Equations (41)


and (42) show that the overall resistance for mas s transfer is the sum of
individual ones. From Eqs. (41) and (42) we obtain
(43)
Since we are considering an emulsion polymerization system where the
particle contains at most one radical, the rate of decrease in the number of
active particles containing a j-mer radical equals the desorption rate of jmer radicals from the particles. Thus, applying Eq. (39) we have

= Ksjap[(l/vp)

-dNlfdt

- mdJR:J]N/ + Ksjap[(O/vp)
- mdJR:J]

x (No + N,* + Nt + Ni +.oo + N/-1 + N/+1 + oo.)

(44)

where N/ is the number of active particles containing a j-mer radical, mdj is


the partition coefficient for j-mer radicals between the water and particle
phases, Nt is the number of polymer particles containing an initiator
radical, and [R:J is the concentration ofj-mer radical in the water phase.
The first term on the right-hand side of Eq. (44) represents the desorption
rate of a j-mer radical and the last term expresses the absorption rate of jmer radical s from the water phase into the particles containing no j-mer
radicals. Rewriting Eq. (44) we get
dNj/dt
here

= -Ks/ap/vp)NJ

+ KsjapmdlR:j]N* + KsjapmdlR~j]No (45)

N* = Nt + Ni + Ni + ... + N/-'l + N/ + N/+1 + ...

(46)

Summation of Eq. (45) with respect to Nj* leads to


dN*/dt

= d Nlfdt = -('1 KsjN/Hap/vp)+ (KSjmdlR:J)apN*


+ ( KSjmdJR:j])apNo

(47)

The second term on the right-hand side of Eq. (47) usually represents the
absorption rate of j-mer radical from the water phase into the active
particles that already contain a radical, and hence represents the rate of
increase in the number of particles containing two radicals, one of which is
a j-mer radical. However, this term must be changed from positive to
negative because it expresses the rate of decrease in the number of active
particles when instantaneous termination in the particles is assumed. Thus,
using Eq. (43), Eq. (47) can be rewritten as
dN* /dt = - ('1KsjN/)(ap/vp)- ( KWjmdJR:j])apN*
(48)

202

Mamoru

Nomura

Equation (48) defines the radical desorption and reabsorption in the


polymer particles and corresponds to Eqo (36)0Thus, comparing the corresponding terms in Eqso (36) and (48), we obtain the following equations

whichdefinethe coefficientskf and ka

kf

= LKsiNNN?i')(ap/vp)
= LKOJ{NNN*)

ka = (LKw~[R:j]/[R:])ap

KOj =

= Kai[R:j]/[R:])

Ksiap/vp) (49)

Kaj = Kwjap (50)

where
[R:] = [R:.J + [R:] + [R:2] + o.. + [R:j_] + [R:j] + 0.0 (51)
In this case, ke and (j defined in Eqso (13) and (14) are, respectively,
represented by
ke =

B.

KSJ{Nf/ N*)

Derivation 01 Mass-Transler
and Reabsorption

(j

= (NNN*)

Coefficient lor Radicql Desorption

In order for the rate coefficients kf and kaodefined by Eqso (49) and (50),
to be predicted it is necessary to know how KSj and Kwj are expressed in
terms of the chemical and physical properties involved in the desorption
and reabsorption processeso First of all, let us consider the mass-transfer
coefficient for a j-mer radical in the individual diffusion filmso There is a
large number of published studies concerning the mass-transfer coefficient
in an external diffusion film around a spherical particleo One of these is the
following semitheoretical equation proposed by Ranz and Marshall (1952)
Sh

=2 +

006R~/2S~/3

(52)

where Sh is the Sherwood number (kwAJDwj), Re the Reynolds number


(dpuP//l), and Sc the Schmidt number (J1/pDwj)o
Since such small spheres as
emulsion polymer particles will move with the eddies of the fluid, there will
be no relative velocity u between the surface of the particle and the fluid
(water)o Therefore, the value of Re can be regarded as zero, and hence
Sh = 2. From the value of Sh = 2 we get
kwj = 2Dwidp

(53)

On the other hand, mass-transfer inside the viscous polymer particles will
occur by molecular diffusiono According to an analytical solution to the
diffusion equation inside the particles, average mass-transfer coefficient
between time O", t,ks is given as follows (Newman, 1931):

_k 0=
SJ

- dp 1n
6
t

[~ ~ 2- (
2 L-

1tn=n

2 exp

4n21t2Dpl
dp2

)]

(54)

203

5. Desorption and Reabsorption of Free Radicals

Since this coefficient is too complicated to employ for our present purpose,
we use an approximate mass-transfer coefficient derived by the following
simple treatment. The average time t spent by a radical inside a particle
before it escapes from the particle can be ca1culated using the Einstein
diffusion equation. Nomura and Harada (1981) used the following value
for t:

(55a)
Recently, Chang et a/. (1981) considered two cases: (1) when a radical enters
at the edge of the particle t is given by
(55b)
and (2) when a radical is generated anywhere in the particle t is given by
(55c)
We assume that the rate of radical desorption from the particle can be
ca1culated approximately using the following equation obtained from
Eq. (37) by neglecting the term Cpji because the value of Cpji is usually very
low compared to that of Cpj.
(56)
Using Eq. (56), the average time t spent by a radical inside the particle
before it escapes fram the particle will be given by
t =. (CpjVp)/lR*= (Cpjvp)/ksjap Cpj = dp/6ksj
By comparing

Eqs. (55) and (57) we obtain


ksj

= /3(Dp/dp)

/3 = i

-i

(57)

(58)

Substitution of Eqs. (53)and (58)into Eqs. (41)and (42)leadsto


K . = 2Dwj 1 +
SJ

mdjdp

2Dwj
/3mdjDpj

-1

= 2Dwj

bj

(59)

mdjdp

and
KWj

= m.K
dJ

sj

- 2Dwjb.
l

(60)

dp

where
(61)

Mamoru Nomura

204

where bj is the ratio of the external film mass-transfer resistance to the


overall mass-transfer resistance for j-mer radicals (mdikwj)/(l/Ksj)'
C.

Derivation 01 Rate Coefficient lor Radical Desorption

To simplify the subsequent treatments, we make the following five


assumptions: (i) polymer particles contain at most one radical, (ii) a radical
with no longer than s monomer units can desorb from and enter into the
particles with the same rate irrespective of chain length, (iii) instantaneous
termination occurs when another radical enters the particle already containing a radical, (iv) no distinction is made between radicals with or
without an initiator fragment on its end, and (v) water-phase reactions such
as propagation, termination, and transfer can be neglected from a kinetic
point of view, as shown in Section n,A. Under these assumptions, the rate
coefficient for radical desorption from the particles is derived with both
deterministic and stochastic approaches.
1. Deterministic Approach*
According to the assumptions [(ii) and (iv)] given above, we can regard the values of KOj and Kaj defined by Eqs. (49) and (50) to be constant
and equal to Ko and Ka regardless of radical chain length, j, respeCtively.
Moreover, no radicals longer than s monomer units will be found in the
water phase according to the assumption of (ii). Thus, Eqs. (49) and (50)
can be rewritten as

kf

= KQI(~I:)+ Ko(~~) + Ko(Z!) + ... + Ko(~:)

[R~I

[R~l]

[R~2]

[R~s] "-

a = Kal [R~] + Ka [R~] + Ka [R~] +... + Ka [R~] = Ka

(62)

(63)

where KQIand Kal are the coefficients for initiator radical s and Ko and Ka
are the coefficients for radicals other than initiators. Discrimination between the initiator and other radicals is made here because their chemical
and physical properties are very different. Thus, the values of KQIand Ko
are different.
In order to get the relationship between Nj and N*, we have the balance
equations with respect to N1*and Nf which contain an initiator radical and
* See Nomura

et al., 1976 and Nomura

and Harda,

1981.

5.

205

Desorption and Reabsorption of Free Radicals

a j-mer radical, respectively. Applying a stationary-state hypothesis to these


equations we obtain
dNNdt

= KaI[R:]No -

(ka[R:] + kmr[Mp]+ kTf[Tp] + k[Mp] + KOI)N,*

=0

~~

dNT/dt = Ka[R:]No + k[Mp]N*+ (kmf[Mp] + kTf[Tp])N*

dNNdt

(ka[R:] + kmf[Mp] + kTf[1;,] + kp[Mp] + Ko)Ni

= Ka[R:j]No

= O

+ kp[Mp]Nt-

- (ka[R:] + kmf[Mp]+ kTf[Tp]+ kp[Mp]+ Ko)Nt = O


dN.*fdt = Ka[R:s]No
- (ka[R:]

~~
~~

+ kp[Mp]Ns*-1
+ kmrEMp] + kTf[Tp] + kp[Mp]

+ Ko)Ns*

~~

where k is the initiation rate constant, kTfis the chain transfer rate constant
to chain transfer agent, [Tp] is the concentration of chain transfer agent in
the polymer particles and kmf is the chain transfer rate constant to
monomer. Furthermore, by taking a balance on the water-phase radicals
and applying a steady-state hypothesis we have
(68)
d[R:]/dt

= KoNi -

Ka[R:]NT

=O

(69)

d[R;j]/dt

= KoNt -

Ka[R:j]NT

=O

(70)
(71)

Since high molecular weight polymers are usually produced in emulsion


polymerization, it can be assumed that
ka[R:], kmr[Mp],kTf[Tp] ~ kp[Mp], k[Mp],Ko" Ko

(72)

From Eqs. (64),(68), and (72) and the definition N* = NT,weobtain


N,* =

Pw(1- ) -

n~

KOI + k[Mp]

By similar treatments, we have

n~

206

Mamoru Nomura

and
j-

fVt

kp[AIp]
Kon + kp[AIp]

)
)[( )+ ( )fV* + (~ )
fVt

=(

kp[AIp]

kmr

kTf[Tp]

kp

kp[ AIp]

fVt

kp

(75)

Inserting Ego (75) ioto Ego (62) we get

kmr

fVt

kr = KOI

kTr[Tp]

kfVt

kp[AIp]

( )+ Ko[(k;)+ (kp[AIp])+ (kpfV*)] j': (Ko + kp[AIp])


fV*

(76)
Introduction of Ego (73) and fV* = fVTinto Ego (76) leads to

k - K
r-

Pw(1- )

O( (KO( + k[AIp])fVT

I(

kp[AIp]

]
) [k; +
j

kmr

+ Ko j= Ko + kp[AIp]

kTf[Tp]

PW(1- )k

kp[AI!,] + (KO(

+ k[AIp])kpfVT]
(77)

Application of kr given by Ego (77) to a real emulsion polymerization


system will be shown latero
20 Stochastic Approach
In this treatment (Harada et al., 1971) we make the same assumptions as
given aboveo A radical in the particle has the probability of undergoing the
following four events in the particle: (i) initiation and propagation reactions,
(ii) chain transfer to monomer, polymer, emulsifier, and so on, (iii) termination reaction when another radical enters the particle, and (iv) desorption
from the particle into the water phaseo
If the probability for a radical to escape from the particle within the time
interval for which the radical needs to add one monomer unit is designated
as q, the probability for the radical not to desorb within the time interval p
is given" by

p=l-q

(78)

Since a radical with no longer than s monomer units is assumed here to


escape from the particles with the same rate, irrespective of chain length, the
probabilities p and q should be constant regardless of radical-chain lengtho
Considering Figo 3, the fraction <I\t for the radicals generated by chain

207

5. Desorption and Reabsorption of Free Radicals

containing (s+l) units

Fig. 3. Scheme for radical escape from polymer partic1e. (Reprinted with permission of
Journal of Chemical Engineering, Japan.)
transfer to monomer or transfer agent to desorb from the particles before
growing to (5 + 1)-mer radical is given by
<I\t = q(1 + P + p2 + p3 + ... + p' - 2 + p' - 1) = 1 - p'

(79)

The rate of radical production by chain transfer to monomer and transfer


agent can be expressed as
(kmr[Mp]+ kTf[1;,])N*

(80)

Part of these radicals will escape from the particles before growing (5 + 1)-

mer radical s and that fraction is given by


(81)
This quantity also represents the desorption rate of radicals generated
originally by chain transfer to monomer or chain transfer agent. The
appearance of small radicals that can desbrb from the particles also occurs
by entry of initiator radicals. Since termination in the water phase can be
neglected (as shown in Section n,A), all initiator radicals generated in the
water phase enter the particles. The entry of initiator radical s into the
particles containing no radicals results in an increase in the number of
polymer particles containing an escaping radical, with the rate given by
(82)
Let <1\be the probability for an initiator radical to escape the particle before
it adds one monomer unit. Then, the desorption rate.of radicals that have
grown from the initiator radicals is expressed by
(83)
The first term represents the desorption rate of initiator radicals that enter
the particles and the last term represents the desorption rate of the radicals
that have grown from initiator radicals. At steady state, the concentration
of radical s in the water phase is constant. This means that the rate of

208

Mamoru Nomura

radical entry into the partic1es should be the same as that of radical
desorption from the partic1es. Therefore, the decrease, due to radical
desorption, in the number of partic1es containing escaping radicals is partly
recovered by radical entry into the partic1es containing no radicals. Let N
be the number of polymer partic1es containing an escaping radical. Then,
KoN represents the rate of radical desorption, and KoN(N o/NT)is the rate
of recovery in the number of partic1es containing an escaping radical by
radical entry into. the partic1es containing no radicals. Therefore, the
apparent (or net) rate of radical desorption per partic1e containing an
escapingradical is given by
(84)

[KoN - KoN(N o/NT)]/N = Ko

For initiator radical s Ko in Eq. (84) should be replaced by Ko).


Considering the rate for each event (ii) to (iv) to occur, only events (i) and
(iv) are usually important in calculating the probability, p. The probability p
is therefore expressed by
p

1 -'q

(85)

kp[Mp]

Ko + kp[Mp]

Substitution of Eq. (85) into Eq. (79) leads to

~ = Ko f.

kp[Mp]

)
j

= 1-

kp[Mp]j= 1 Ko + kp[Mp]

kp[Mp]

(Ko+ kp[Mp])

(86)

. In the same way we obtain


cI\ =

(87)

KOJ

Ko) + k[Mp]

It must be noted here that the rate coefficient for radical desorption kr is
related not to the apparent (or net) rate but to the true rate of radical
desorption by molecular diffusion. Using Eqs. (81) and (83), therefore, the
apparent (or net) rate of radical desorption from the partic1es can be
expressed by the use of the coefficient kr as
krN* -

krN*(No/NT)

= <1>M(kmr[Mp]

+ kTf[Tp])N*

+ Pw(No/NT)[<1 + ~(1

- <1)]

(88)

Introducing Eqs. (86) and (87) into Eq. (88) and rearranging leads to
Eq. (77).
kr

= KOJ

Pw(1

- )

(KOJn+ k[Mp])nNT
Pw(1 -

)k

kmr

] [
]i (

(KOJ+ k[Mp])kpNT

+Ko

kp

+ -kTf[1;']
kp[Mp]

kp[M p]

j= 1 Ko

+ kp[M p]

(77)

209

5. Desorption and Reabsorption of Free Radicals

Initiator radical s are so reactive that initiator radicals would initiate


polymerization before escaping from the particles and hence, the number of
polymer particles containing an initiator radical will be maintained at a low
value. Moreover, when transfer to monomer is dominant, that is, the value
of kmf/kp is large, it is reasonable to consider that the term N,*/N* in
Eq. (76) can be neglected. Thus, kr given by Eq. (77) is simplified as
kr = Koz
where
Z

kmr + kTf[Tp]
kp

I(

kp[Mp]

kp[Mp]

(89)

(90)

j= 1 Ko + kp[Mp]

The value of Z depends upon the values of Ko and kp[Mp] and satisfies the
following inequality:
o< Z < s
(91)
If it is supposed further that only monomer radicals can escape from a

particle(that is, s = 1) and that Ko ~ kp[Mp] Eq. (89)becomes


kp[Mp]

(Ko + kp[Mp])(

kr = KO

kmr + kTf[Tp]
kp

kp[Mp]

(92)

where Ko is the desorption rate constant for monomer radicals defined by


Eq. (49), and it is supposed that Ko is also applicable to a chain transfer
agent radical. According to Eqs. (49) and (59),therefore, Ko is given by
Ko = Ks(aJvp) = 12Dw)/rndd~
Considering that the value of {3is i
most likely value for l, we have

(93)

1, because Eq. (55c) seems to give the


(94)

Since the physical properties of a monomer radical will be almost identical


to those of monomer, the values of Dw, Dp, and rndused here are those for
monomer.
We have derived Eq. (92) under the assumption that the physical and
chemical properties of a chain transfer agent radical are approximately
equal to those of a monomer radical. However, if we take into account the
difference of physical and chemical properties between the chain transfer
agent and monomer radicals, Eq. (92) should be modified as
k

r-

kmr[Mp]

o Ko + kp[Mp]

) (
+K

kTf[1;']

OT KOT

+ kiT[Mp]

(92')

where KOTand kiT are the desorption rate constant and the reinitiation rate
constant for chain transfer agent radical s, respectively.

210

Mamoru

Nomura

In the absence of transfer agent, when the rate of radical desorption from
the partic1es and the reactivity of monomer are very high, as with vinyl
acetate and vinyl chloride, the condition kp[Mp] ~ Ko is fulfilled. For such
systems Eg. (92) is simplified further to
kf

= Ko(kmelkp) = (12DwDlrndd~)(kmelkp)

(95)

Consider here the value of D. In the beginning of the polymerization, D


seems to be unity because the partic1es are saturated with monomer, and
hence the decrease in the value of Dp, the self-diffusion constant of a
monomer radical in the partic1e, is not so remarkable. However, in a higher
conversion range the value of Dp decreases with the progress of polymerization due to an increase in the viscosity inside the partic1es. For this
reason, the value of Dwill decrease in a higher conversion range. When the
condition 2DwlprndDp~ 1 is satisfied in Eg. (94), that is, the diffusion
resistance inside the partic1es becomes dominant, the rate coefficient for
radical desorption kf given by Eg. (95) is written in the f?rm
kf --

6pDp

2
dp

kmf ~ 6Dp

kp

kmf

(96)

2
dp kp

Using Eg. (55a), Friis and Nyhagen (1973) derived the following expression for kf:
kf

= (2Dpld~Df~:[Mp])

kmf[Mp]

(97a)

When the condition kp[Mp]~ 2Dpld~is fulfilled,Eg. (97)is simplifiedto


k

= 2Dp kmf
d2p

kp

(97b )

This is the same as Eg. (96) except for the value of the numerical constant.
Ugelstad and Hansen (1976) also derived arate coefficient for radical
desorption similar to those given by Egs. (92) and (95).
IV.

A.

Effect of Free-Radical Desorption on the Kinetics


of Emulsion Polymerization
Effect on the Rate 01 Emulsion Polymerization

1. Verification of Applicability of kf

In the emulsion polymerization of vinyl acetate and vinyl chloride, for


example, the polymer partic1e contains at most one radical, under normal
conditions. For such emulsion polymerization systems, the steady-state

211

5. Desorption and Reabsorption of Free Radicals

hypothesis can be applied to the following balance equations for the


number of active particles N* and the radical concentrations [R:] in the
water phase
dN*/dt

= ka[R:]No -

kfN*

ka[R:]N*

=O

d[R:]/dt = Pw+ kfN* - ka[R:]NT = O

(98)
(99)

Substituting Eq. (99) into Eq. (98) and solving for , we reproduce Eq. (30):

= 1(- e + ~ C2 + 2C)

e = pw/kfNT

(30)

In Fig. 4, theoretical values of calculated by Eq. (30) are plotted against


the value of e = Pw/kfNT. Experimental values of obtained at the
conversion point in vinyl acetate and vinyl chloride emulsion polymerizations where monomer droplets disappear in the water phase (20 -- 40%)
are also plotted in Fig. 4 and compared with the theoretical values. The
numerical values of constants used here are shown in Table I. The rate of
radical production in the water phase was calculated from
Pw = 2kd[Io]

(100)

where kd is the rate constant for radical decomposition, f is the initiator


efficiency,and [lo] is the initiator concentration.
It has generally been accepted that the main interruption reaction of the
growing polymer chain is first order with respect to the concentration of

monomor

10-1,

50.C

Yinyl
chlaride

SO.C

kfNy

literature
4
5,7
22
1
Ihalsgd
(1
.6. Gi.........
11
L:. Peggion.
24

key
()
o

resean:he1's

Frils "
Nomura
Zollars

Id

IcJ3

te.

vinyl
acotate

[-J

Fig. 4. Comparison between theoretical and experimental values of : (a) calculated


from conversion versus time curve and the number of polyrner particles. (Reprinted with
permission of Journal of Applied Polymer Science.)

212

Mamoru Nomura

TABLEI
Numerical Values of the Constantsa
Constants

(units)

Vinyl acetate

kp
(Iiterjmolsec)
kmrjkp
md
Dw
(cm2jsec) .
kd
(Iiterjsec)
[Mp). (molj1iter)
z
b

3340
2 X 10-4
28
1.9 x lO-s b
1.5 X 10-6 e
8.9
1.0
1.0

a 500C. Reprinted with permission


Applied Polymer Science.

Vinyl chloride

10,000

1.2x 10-3
35

2.5 X lO~sb
1.5 X 10-6,
6.2
1.0
1.0
of the Journal of

b From Wilke and Chang (1955).


From Morris and Parts (1968).

growing polymer chains. This was supposed to result from a chain transfer
to monomer and the rate of the interruption reaction rf was experimentally
analyzed according to the expression
rf = k[R .][M]

(101)

where [R -] is the concentra tion of growing polymer chains, [M] is the


concentration of monomer, and k is the rate constant of the interruption
reaction. Ir the main interruption reaction of growing polymer chains is a
chain transfer to monomer, the value of k corresponds to the value of kmf,
the rate constant of chain transfer reaction to monomer. However, Ugelstad
(1980) recently pointed out that although the rate of interrupting the reaction of growing chains is represented by Eq. (101) in vinyl chloride polymerization, k is not equal to kmf but is composed of many rate constants because
a growing polymer radical, after the head-to-head addition to monomer,
undergoes several reactions and finally terminates by splitting off CI. or
H- radicals. This finding, however, does not affect the present treatment as
long as the experimental value of k, based on Eq. (101), is used as kmf in the
coefficient kf, in the case of vinyl chloride emulsion polymerization.
As can be seen in Fig. 4, a comparison of the theoretical and experimental values of , assuming the value of Z = 1, gives a good agreement.
From this and the fact that the value of Ko is about an order of magnitude
smaller than that of kp[Mp] in vinyl acetate and vinyl chloride emulsion
polymerizations, we can conclude that the radical that can escape from a
particle seems to be principally a monomer radical. Furthermore, it can be

213

5. Desorption and Reabsorption of Free Radicals

conc1uded that Eqs. (30) and (95) are applicable to vinyl acetate and vinyl
chloride emulsion polymerizations.
On the other hand, for such monomers as methyl methacrylate and
styrene which are less reactive and less soluble in water than vinyl acetate
or vinyl chloride, Eqs. (30) and (76) [or (77)] are essentially applicable,
although the value of the term (kJkp)(N*jN*) may not necessarily be
neglected comparing the value of the term kmr/kp.
The rate coefficient for radical desorption in emulsion copolymerization
was also derived in the same way as described in Section III, and it was
successfully applied to explaining the rate of emulsion copolymerization of
methyl methacrylate and styrene (Nomura et al., 1978, 1979).
2. Effect of Radical Desorption on Rate of Emulsion Polymerization
Let us consider the rate of emulsion polymerization Rp where Eqs. (30)
and (95) can be applied. From Eqs. (30-32), it is c1ear that Eq. (30) can be
written in the form
(102)
The values of a varies from ! to O with an increase in the value of C.
Combining Eqs. (95) and (102) we have
I ] av2a/3
R p oc N T oc N1-a
(103)
] ad2aoc N1-a
T [I Op
T [ Op
After the complete absorption of monomer existing as monomer droplets in
the water phase, the following relationship holds:
NTvp= Vp

(104)

where Vpis the total volume of the partic1e per unit water volume. Inserting
Eq. (104) into Eq. (103) leads to
R p oc N(3
(105)
T - saJ/3[ I ] aJl:2a/3
p
When radical desorption
Eq. (105) becomes

from the partic1es can be neglected

R p oc N1.0
T [I ] 0Jl:0
p

(Le., a = O)

(106)

This corresponds to Smith-Ewart Case II kinetics and is applicabie to


styrene emulsion polymerization under normal conditions. On the other
hand, when radical desorption from the partic1es is dominant (Le., a = !)
Eq. (105) leads to
R p oc Nl/6
T [ I ] 1/2Jl:1/3
p

(107)

This theoretical equation explains very well the rate of seeded emulsion
polymerization of vinyl acetate and vinyl chloride found by Patsiga et al.

214

Mamoru Nomura

(1960) and Giskehaug (1965). Recently, Ugelstad et al. (1969) reported that
when the number of polymer particles present was large, the rate of vinyl
chloride emulsion polymerization is proportional to the 0.15 power of the
number of polymer particles and to the 0.3 power of the total volume of
polymer

particles

~.

Nomura

et al. (1971, 1976), found

that

Eq. (107)

explained the dependence of the rate of vinyl acetate emulsion polymerization on NT, [/0], and ~. Van der Hoff (1956) reported that even in
styrene emulsion polymerization, the rate of polymerization was proportional to the 0.17 power of the number of polymer particles when the
number of polymer particles was very large. These experimental findings
support the validity of the theory developed in this chapter.

B.

Effect on Micellar Particle Formation

Smith and Ewart (1948) proposed two idealized situations for the
formation of polymer particles, assuming that (i) partile nucleation occurs
in monomer-swollen emulsifier micelles, (ii) the volumetric growth rate of a
particle is constant in the interval of particle formation, and (iii) radicals do
not desorb from a particle.
Case 1: Ir only micelles can absorb initiator radicals, then
dNT/dt = Pw

(108)

In this case the number of polymer particles formed is given by


NT = 0.53(pw//l)0.4(asSo)0.6

(109)

where /l is the volumetric growth rate of a particle, as is the surface area per
unit amount of emulsifier, and Sois the total amount of emulsifier per unit
volume of water.
Case 2: Ir both particles and micelles absorb radicals at arate
portional to the surface area, then
dNT/dt = pw[A.,J(Am + Ap)]

pro-

(110)

where Am and Ap are the total surface areas of micelles and polymer
particles, respectively. In this case
NT

0.37(pw//l)0.4(asSo)0.6

(111)

However, when radical desorption from particles takes place in the


interval of particle formation, Pwin Eqs. (108) and (111) should be changed
to PA, the overall rate of radical entry into micelles and particles, and the
volumetric growth rate of a particle can no longer be a constant.

215

5. Desorption and Reabsorption of Free Radicals

Noting this, Nomura et al. (1972, 1976; Nomura, 1975) studied the effect
of radical desorption on the formation of polymer particles from micelles
for the case Y = o. They used the following non-steady-state treatment
under the assumption that the polymer particles contain at most one
radical.
Particle formation:
dNT

k1ms[R:J
klms
+ kzNT[R:J = PA klms + kzNT

dt = PA
-

k1ms[R:J

) (

PA

1 + aNT/Sm

)
(112)

where
(113)
and
PA

= (k1ms + kzNT)[R:J = Pw+ kfN*

(114)

and where Mmis the aggregation number of a micelle.


The number of polymer particles containing a radical, N*;
d~* = (k1ms + kzNo - kzN*)[R:J - kfN*
= [1 + (aNT/Sm)(l- 2N*/NT)](pw + kfN*)/[l + (aNT/Sm)]- kfN*
(115)
Monomer conversion:
(116)
where Mw is the molecular weight of monomer, Mo is the initial monomer
concentration, and NAis Avogadro's number.
Emulsifier balance:
(117)
where
(118)
The second term on the right-hand side of Eq. (117) represents the amount
of emulsifier adsorbed on the surface of polymer particles.
These equations correspond to the Smith and Ewart Case 1 model of
particle formation when the conditions a = Oand kf = Oare employed; they
further correspond to the Smith and Ewart Case 2 model of particle

216

Mamoru Nomura

formation when the conditions k1 = nd~, k2= nd~,and kf = O are employed (where dmis the diameter of a micelle).
Nomura et al. (1972) solved Eqs. (112) to (118) on a digital computer and
analyzed the particle formation in styrene emulsion polymerization where
kf = O approximately held under normal conditions and found that the
value of 8 was 1.3 x 105. This value is about 102 times greater than that
predicted by the diffusion theory which assumes that k1 = 2nDdm and
k2 = 2nDdp,and whereD is the diffusioncoefficientof a radical in the water
phase.
On the other hand, in vinyl acetate emulsion polymerization the value of
8 was 1.2 x 107 (Nomura et al., 1976). This value is also about 104 times
greater than that predicted by the diffusion theory. The reason for this may
be that radical s have greater difficulty in entering micelles than polymer
particles, or it may be that radicals, having entered a micelle, may escape
from the micelle too rapidly to cause initiation, because the micelle has too
small a volume. Both factors will decrease the apparj::nt value of k1 and
hence increase the value of 8. Therefore, 8 can be regarded as a factor that
represents "the radical capture efficiency of a micelle relative to a particle."
Furthermore, they studied the effect of radical desorption from the
polymer particles on the particle number formed in Internal 1 of emulsion
polymerization, using above equations, and showed that radical desorption
leads to an increase in the number ofparticles. The reason for this is that: (1)
The desorbed radicals reenter the micelles and take part in particle nucleation: (2) Radical desorption decreases the particle growth rate and hence,
results in a decrease in the rate of micelle consumption. This also increases
the chance of radical entry into the micelles. They also showed that radical
desorption brought about the change in the orders of particle number with
respect to emulsifier and initiator according to the following relationship:
NT oc S~p~-a

(119)

The value of a increases from 0.6 to 1.0 with increasing radical desorption.
For vinyl acetate and vinyl chloride emulsion polymerization, the calculated

values of order with respectto Soand Pw are 1 and O,respectively.Theseare


in good agreement with experimental results (Ugelstad et al., 1969; Nomura
et al., 1976). For other monomers and details see the recent paper by
Hansen and Ugelstad (1979).
Chain transfer agents such as CCI4, CBr4, and mercaptan promote the
desorption of radicals from the particles and result in an increase in the
value of kf, as can be seen from Eq. (92) (Nomura, 1975; Whang et al., 1980;
Nomura et al., in press). Thus, increasing the amount of chain transfer
agent charged initially causes both a decrease in the rate of emulsion
polymerization and an increase in the number of polymer particles formed.

5. Desorption and Reabsorption of Free Radicals

217

Furthermore, the value of a in Eq. (119) increases from 0.6 to 1.0 with
increasing transfer agent (Nomura, 1975; Nomura et al., in press).

List of Symbols
a
ap
a,
Am
Ap
C
Cp
Cp;
Cw
Cw;
dm
dp
Dp
f
I
Im,/m-1
[lo]
j
Jpo

k.
kd
k
k
kj'

(8:x)1/2
surface area of a partic1e
partic1e surface area per unit amount
total surface area of micelles

total surface area of polymer partic1es


(J.'/m
concentration
of radicals in a particle
concentration
of radicals in particle at interface
concentration
of radicals in the water phase
concentration
of radicals in the water phase at the interface
diameter of micelle
diameter of particle
self-dilfusion constant of a monomer
initiator efficiency
subscript denoting initiator radical
Bessel functions of the first kind
initiator concentration
subscript denoting
denotes an initiator

radicals
radical

rate of radical desorption


rate coefficient for radical
rate constant for initiator
rate coefficient for radical
initiation rate constant

kTf
k.
kw
m
md
'n~
Mm
Mw
[Mp]
[Mw]

radical in a particle

with j number

of monomer

film mass transfer


interface between
rate constant for
rate constant for
rate constant for

units, j = O

from a single partic1e


entry into the partic1es
decomposition
desorption from the partic1es

(j = 1,2,3,
) rate constant for radical desorption
j-mer radicals
rate constant for transfer to monomer

propagation rate constant

k,p
k,w

of emulsifier

from the partic1es for

coefficient for the inner dilfusion film adjacent


the partic1e and water phases
termination in the partic1es
termination in the water phase
transfer to transfer agent

rate constant for radical desorption from the partic1es


film mass transfer coefficient for the outer dilfusion film adjacent
interface between the partic1e and water phases
kl'plk,p
partition coefficient for radicals at the particle-water
concentration
of micelles per unit volume of water
aggregation number of a micelle
molecular weight of monomer
concentration
of monomer in partic1es
concentration
of monomer in water phase

interface

to the

to the

218

Mamoru Nomura

Mo
n

N
N*
NA
N.,

initial

monomer

number

concentration

of radicals in a particle

average number

of radicalsper particle

number

of particlescontaining an escaping radical

number

of particlescontaining one radical

Avogadro's number
(n =

NT
P

0, 1,2, 3...)

number

of particles containing

number

of polymer particlesper unit volume of water

radicals

probability of a radical not escaping fron:a particleduring the time


in"terval
of adding one more monomer

of adding one more monomer

Re
Rp
[R:]
[R:]
5
Sc
Sh
So
Sm
t

unit

probability of a radicalescaping from a particleduring the time interval


unit

Reynold's number
overallrate of polymerization
concentration of radicalsin the particlesper unit volume of water
totalconcentration of radicalsin the water phase
maximum
number of monomer
units in a desorbed radical
Schmidt number
Sherwood

number

totalamount
amount

of emulsifierper unit volume of ~ater

of micellaremulsifierper unit volume of water

average time spent


the

by a radical inside a particlebefore itescapes from

particle

tadd

average time for a radicalto add one monomer

red

average residence time of a radical in the water phase before entering a

unit in the water phase

particle

fterro

average time for a radical to be deactivated by mutual

termination in

the water phase

[1;,]
Vp
Vp
Xm

concentration of transferagent
volume

of a particle

totalvolume of particlesper unit volume of water


fractionalmonomer
conversion

ex

parameter defined by Eq. (22)

ex'

parameter defined by Eq. (27)

numerical

constant,
!- i
ratio of the outer film mass-transfer resistance to the overall mass-

transfer resistance for j-mer

Cl
J1

ratio of number

radicals

ofj-mer radicalsto the totalnumber

of radicals

volumetric growth rate of a particle

PA
Pw

rate of radical production

<1>m

fraction

overall rate of radical entry into the particles

of

tbe

transfer agent

in the water

radicals

to desorb

generated
from

phase

by

chain

transfer

the particles before

to

growing

monomer
to (5 +

or
1)-mer

radicals
<1>

probability of initiatorradicals escaping out of the particles before

t/lc

monomer

adding one monomer

unit

weight fraction

in the particles

atsaturation

with monomer

5. Desorption and Reabsorption of Free Radicals

219

References
Chang, K., Litt, M., and Nomura, M. (1981). Macromolecules (to be published).
Friis, N., and Nyhagen, M. (1973). J. Appl. Polym. Sci. 17, 2311.
Gardon, J. L. (1968). J. Polym. Sci. Part A-l, 6, 665.
Giskehaug, K. (1965). Symp. Chem. Polym. Process, London, April.
Hansen, F. K., and Ugelstad, J. (1979). Macromol. Chem. ISO, 2423.
Harada, M., Nomura, M., Eguchi, W., and Nagata, S. (1971). J. Chem. Eng. Jpn. 4(1), 54.
Harkins, W. D. (1947). J. Am. Chem. Soco 69, 1428.
Lewis, W. K., and Whitman, W. G. (1924). Ind. Eng. Chem. 16, 1215.
Litt, M., Patsiga, R., and Stannett, J. (1970). J. Polym, Sci. Part A-l S, 3607.
Morris, C. E. M., and Parts, A. G. (1968). Macromol. Chem. 119,212.
Newman, A. B. (1931). Trans. AIChE 27, 310.
Nomura, M. (1975). Ph.D. Thesis, Kyoto Uniy., Kyoto, Japan.
Nomura, M., and Harada, M. (1981). J. Appl. Polym. Sci. 26, 17.
Nomura, M., Harada, M., Nakagawara, K., Eguchi, W., and Nagata, S. (1971). J. Chem. Eng.
Jpn.4(2), 160.
Nomura, M., Harada, M., Eguchi, W., and Nagata, S. (1972). J. Appl. Polym. Sci. 16,811.
Nomura, M., Harada, M., Eguchi, W., and Nagata, S. (1976). In "Emulsion Polymerization"
(1. Piirma and J. L. Gardon, eds.), ACS Symposium Series No. 24, pp. 102-122.
Nomura, M., Harada, M., and Eguchi, W. (1978). J. Appl. Polym. Sci. 22, 1043.
.
Nomura, M., Yamamoto, K., Fujita, K., Harada, M., and Eguchi, W. (1978). Preprints, 12th
Fall Meeting ofthe Soco ofChem. Engrs., Okayama, Japan, GS-207, pp. 439-440.
Nomura, M., Yamamoto, K., Harada, M., and Eguchi, W. (1979). Preprints, Int. Conf Surface
Col/oid Sci., 3rd, Stockholm, Sweden, pp. 148-149.
Nomura, M., Minamino, Y., Fujita, K., and Harada, M. J. Polym. Sci., (in press).
Patsiga, R., Litt, M., and Stannett, V. (1960). J. Phys. Chem. 64, 801.
Ranz, W. D., and Marshall, W. R. (1952). J. Chem. Eng. Prog. 4S, 141.
Smith, W. V., and Ewart, R. H., (1948). J. Chem. Phys. 16, 592.
Stockmayer, W. H. (1957). J. Polym. Sci. 24, 313.
O'Too1e, J. T. (1965). J, Polym. Sci.9, 1291.
Ugelstad, J. (1981). Pure Appl. Chem. 53, 323-363.
Ugelstad, J., and M0rk, P. C. (1970). Dr. Polym. J. 2, 31.
Uge1stad, J., and Hansen, F. K. (1976). Rubber Chem. Technol.49, 536.
Uge1stad, J., M0rk, P. c., and Aasen, J. O. (1967). J. Polym, Sci. Part A-l 5, 2281.
Uge1stad, J., M0rk, P. C., Dahl, P., and Rangnes, P. (1969). J. Polym. Sci. Part C 27,49.
Van der HotT, B. M. E. (1956). J. Phys. Chem. 60, 1250.
Wang, B. C. Y., Lichti, G., Gilbert, R. G., and Napper, D. H. (1980). J. Polym. Sci. Po/ym.
Letter Edi. IS, 711.
Wilke, C. R., and Chang, P. C. (1955). AIChE J. 1,264.
Zollars, R. L. (1979). J. Polym. Sci. 24, 1353.

6
Effects of the Choice
of Emulsifier in
Emulsion Polymerization
A. S. Dunn

1. Introduction .
11. Monomer Emulsification
A. Hydrophile-LipophileBalance
B. Effects of Mixtures of Anionic
and Nonionic Emulsifiers .
C. Preparation of Monodisperse Latexes
111. Emulsion Polymerizationwith Nonionic Emulsifiers
IV. Emulsion Polymerizationwith lonic Emulsifiers.
A. Effectof Emulsifieron Numberof Latex
Particles Formed
B. The Effectof Micelle Size .
V. Latex Agglomeration
VI. Other Effects of Emulsifiers
A. Monomer Solubilization
B. Polymer Solubilization .
C. Initiator Decomposition Induced by Emulsifier.
D. Catalysis of the Initiation Reaction .
E. Transferto Emulsifiers .
References

221
224
225
227
228
229
230
233
234
236
237
237
238
239
241
242
243

l. Introduction
Despite the name, neither the emulsification of monomer nor the
stabilization of polymer particles formed is an essential characteristic of
emulsion polymerization. The essentlal feature of emulsion polymerization
is that polymerization occurs in a large number of isolated particles that
normally contain no more than a single polymerizing radical: this permits
high molecular weight polymer to be formed at a high rate of polymerization, by contrast with bulk or solution polymerization in which an
221
EMULSION POLYMERIZATION
Copyright It 1982 by Aeademie Pres,. Ine.
AJI righlS of reproduetion in any fonn reserved.
ISBN 0-12-556420-1

222

A. S. Dunn

increase in the rate of polymerization generally resuIts in a decrease in the


molecular weight of the polymer produced.
Emulsion polymerization without the use of an emulsifier may be
achieved even with a monomer with water solubility as low as that of
styrene provided one uses an initiator such as potassium persulfate which
introduces ionic end groups into the polymer that can stabilize the polymer
latex particIes produced electrostatically. Emulsifier-free emulsion polymerization is advantageous when the object is to obtain a well-characterized
model colloid for use in experiments on colloidal stability, etc. Then it is
usually desirable that the surfaces of the colloidal particIes be cIean. When
an emulsifier is used in the preparation, its removal (e.g., by dialysis) is
generally so incomplete that it is simpler to avoid its use in the first place.
However, emulsifier-free latexes are necessarily dilute and consequentIy of
little interest for commercial applications.
When an emulsifier is used, its type and concentration primarily affects
the number of latex particIes formed, which in turn determines the rate of
polymerization and, depending also on the rate of initiation, the molecular
weight of the polymer formed.' AIthough the physical properties of the
polymer are primarily dependent on its molecular weight and molecular
weight distribution, the properties of the latex depend on its concentration,
averageparticIe size, particIe size distribution, and the viscosity of the
aqueous phase, which may be enhanced by addition of a thickener-a
water-soluble polymer not adsorbed by the polymer phase which does not
affect the course of the reaction.
For applications in which the stability of the latex to mechanical
agitation, exposure to low temperatures, etc. is a prime consideration, it
may be expedient to add additional amounts of surfactants after the
polymerization is complete: in general the effect of such post-additions
differs from the effect obtained when the same amounts are added prior to
polymerization. The best emulsifiers for stabilizing the mo~omer emulsion
are not necessarily those that are best for stabilizing the polymer latex.
Normally, mechanical agitation is provided to keep the monomer dispersed
during polymerization, whereas the latex is likely to be required to remain
in dispersion for a prolonged period without agitation. The fact that the
latex particIes are much smaller than the monomer droplets in the original
emulsion facilitates stabilization of the latex, but emulsifiers will usually
have to be chosen with. a view to stabilizing the latex rather than the
monomer emulsion. When it is desired to maximize the solids content of the
latex, controlled agglomeration of the latex particIes may be needed. Again,
if the uItimate application of the latex is as an adhesive or a paint vehicIe
requiring formation of a continuous film from the polymer latex particIes
after evaporation of the water, the effect of the emulsifiers on film properties

6.

Effects of the Choice of Emulsifier in Emulsion Polymerization

223

(e.g., water sensitivity) may be of the utmost importance. Hence, extensive


practical testing is likely to be required to determine the least expensive
emulsifying system that will give acceptable results in a given application.
Synthetic butadiene-styrene rubber production accounts for the greatest
volume of polymer production by emulsion polymerization. Most synthetic
rubber is not used as a latex but coagulated to isolate the polymer for
subsequent compounding. Consequently, it is convenient to continue to use
carboxylate emulsifiers, fatty acid or resin acid soaps that permit the latex
to be coagulated simply by acidification. The presence of emulsifier residues
in the polymer is not objectionable in synthetic rubber because zinc stearate
is an important ingredient of must rubber compounds. Coagulation of
latexes stabilized by sulfate, sulfonate, and nonionic emulsifiers is frequently
difficult, although these do coalesce to form more or less coherent films
above the minimum film-formation temperature (MFT) of the polymer: this
is related to the glass transition temperature of the polymer (Tg) but is
generally conspicuously lower because of plasticization of emulsion polymers by water.
The effects of emulsifiers in emulsion polymerization systems may be
enumerated as follows: (1) stabilization of the monomer in emulsion, (2)
solubilization of monomer in micelles, (3) stabilization of polymer latex
particles, (4) solubilization of polymer, (5) catalysis of the initiation reaction,
and (6) action as transfer agents or retarders which leads to chemical
binding of emulsifier residues in the polymer obtained.
Commercial emulsifiers are mostly mixtures of homologs, which are
difficult or impossible to purify or even 3;nalyze. Differences in the composition of the mixture are likely to account for differences observed in the
effects of similar 'materials from different sources (Blackley, 1975).
Consequently, comparisons between the effects of different emulsifiers are
relatively rare. The authors of a number of important publications have
prudently confined their investigation to a single batch of a particular
emulsifier, although it may have been used over a range of concentrations.
When comparisons are made between different emulsifiers, it is difficult to
ensure that they are made at corresponding concentrations. The obvious
procedure is to use equal concentrations by weight (cf. Hopff and Falka,
1965): since emulsifiers are sold by weight, this is the proper procedure to
determine the best value for money. However, the molecular weights of
emulsifiers differ so much that a more chemically satisfactory procedure
would use equal molar concentrations. This should be valid provided the
concentration chosen is greatly in excess of the critical micelle concentration (CMC) of all the emulsifiers. Micellar nuc1eation of latex partic1es
dominates in the polymerization of monomers of low water solubility (e.g.,
styrene), wherein the rate of polymerization may increase by a factor of 100

A. S. Dunn

224
4

2.5

5.0
102SJ/mol

7.5

10.0

dm-3

Fig. 1.
Dependence of the rate of polymerization
(in Interval 11) of styrene emulsions on
the concentration
of the emulsifier (sodium dodecyl sulfate) at 60C. Initiator: 0.2% K2S20S
on aqueous phase. (From AI-Shahib, 1977.)

as the concentration of the emulsifier is increased through the CMC


(Fig. 1). In such a case a comparison between equal co'ncentrations of two
emulsifiers, one above its CMC and the other below, is useless; and it is
logical to choose concentrations at which the amounts of each emulsifier
that will be present as micelles under the conditions of the experiment are
equal, despite the fact that the total concentra tion s of the emulsifiers present
may differ greatly. The conventional anionic emulsifiers (typified by sodium
dodecyl sulfate) are much less strongly adsorbed by polar polymers [e.g.,
poly(vinyl acetate)] than they are by hydrocarbon polymers (e.g., polystyrene). For monomers of moderate (Le., 1-3%) solubility in water (e.g.,
methyl methacrylate, vinyl acetate), the dominant mechanism of latex
partic1e nuc1eation is oligomeric precipitation (Fitch and Tsai, 1971), and
there is no perceptible effect of the presence of emulsifier micelles. In such
cases, corresponding concentration should probably be chosen so as to give
equal surface concentrations on the polymer (i.e", the emulsifiers would
have the same effect in stabilizing the' polymer latex partic1es), but the
paucity of adsorption isotherm data makes it difficult to arrange this at
present.

11. Monomer Emulsification


The emulsification of the monomer is not usually of primary importance
in emulsion polymerization, although it has been shown (Ugelstad et al.,
1973; Ugelstad and Hansen, 1979) that if the size of the emulsion droplets
can, by the use of special methods of emulsification, be reduced considerably below that which would normally be attained, the emulsion

6.

225

Effects of the Choice of Emulsifier in Emulsion Polymerization

droplets can compete effectively for initiator radical s and become a significant or even dominant locus of polymerization. It may occasionally be
desirable to have a monomer emulsion sufficiently stable to permit stirring
to be dispensed with (e.g., this would facilitate the use of the dilatometric
method for determining the rate of an emulsion polymerization), but the
difficulty here is that an emulsion that is stable so long as it is not
polymerized will break as polymerization proceeds because of depletion of
emulsifier from the monomer droplets by adsorption on the latex paitic1es.
This might be prevented by the use of sufficiently large concentrations of
emulsifier .

A. Hydl'ophile-LipophileBalance
Hydrophile-lipophile

balance

(HLB)

pro vides a useful guide

to the

selection of a suitable emulsifier system for a given disperse phase particulady when nonionic emulsifiers are to be used. Several studies of the effect
of HLB in emulsion polymerization

have been published.

The HLB scale was introduced by W. C. Griffin of Atlas (now LC.L


America) (1949, 1954, 1980) as an aid particularly to the use of nonionic
surfactants in the formulation of cosmetics. His scale (Table 1) is related to
the solubility of the materials in water: for nonionic surfactants on
poly(ethylene

oxide) the HLB number may be found simply by dividing the

percentage of ethylene oxide in the product by five (Griffin, 1954). Davies


(1957) has devised group constants that permit ca1culation of HLB values
for other types of surfactant (Table 11). Greth and Wilson (1961), also of
TABLEI
The HLB Scale and Water Solubility of Surfactantsa
Solubility

HLB number

in water

Not dispersible
Poorly dispersible
Unstable milky dispersion
Stable milky dispersion
Translueent solution

{!6

Clear solution

{
a After Adamson,

Applieation

8
10
12
14
16
18

1976. Reprodueed

Emulsifiers for Water in Oil


Emulsions
Wetting agents
Emulsifiers

Detergents
Solubilizers

with permission

for oil in water

Emulsions

of John Wiley & Sons, lne.

226

A. S. Dunn
TABLE

11

Group Constants for Calculations


Hydrophilic

groups (constant)

-S04Na

(38.7)

-COOK
-COONa
-S03Na
-COOH
-OH
-0-{CH2CH20)-

(21.1)
(19.1)
(1l.0)
(2.1)
(1.9)
(1.3)

of HLBa Valuesb

Lipophilic

-CH2-CH=
-CH3
-{CH2CH2CH20)-

a HLB = 7 + ~Jhydrophilic group constant)


constants).
~ After Davies, 1957. Reproduced

groups (constant)

- Dlipophilic

with permission

(0.475)

(0.15)

group

of J.. T. Davies.

Atlas, extended the application' of the HLB system to the selection of


emulsifiers for the emulsion polymerization of styrene and vinyl acetate.
Emulsifiers and emulsifier blends were rated according to the stability of the
polymer latexes obtained on storage at room temperature. The best results
were obtained when the HLB for polystyrene latexes was in the range 13-16
and for poly(vinyl acetate) 14.5-17.5. The highest rates of polymerization
were also obtained in these ranges. Benetta and Cinque (1965) found the
HLB system useful for the selection of stabilizers for the suspension
polymerization of vinyl chloride, but Testa and Vianello (1969) working in
the same laboratory found that all blends were inferior to sodium dodecyl
sulfate in the emulsion polymerization of vinyl chloride. Ono et al. (1974,
1975) found that particle size and latex stability increased as HLB was
decreased by increasing the proportion of hexaoxyethylene oleyl ether
(HLB 9.9) mixed with sodium dodecyl sulfate, both in the homopolymerization of methyl methacrylate in emulsion and in the emulsion copolymerization of acrylonitrile and methyl methacrylate in a molar ratio of 3 to 1.
They determined latex stability by calculating the Hamaker constant for the
latexes from the value of Debye and Hckel's reciprocal thickness of the
ionic atmosphere 1( at the critical concentration of added electrolyte required to coagulate the latex following the procedure of Ottewill and Shaw
(1966).
Jagodic et al. (1975, 1976) also studied the effect of the HLB of
emulsifying agents in the emulsion homopolymerization of methyl methacrylate, ethyl acrylate, and acrylonitrile and in the emulsion copolymerization of methyl methacrylate and ethyl acrylate, methyl methacrylate and

6.

Effects of the Choice of Emulsifier in Emulsion Polymerization

227

styrene, and ethyl acrylate and styrene. They found the HLB to affect
reaction rate, latex stability, viscosity, and partiele size. Optimum latex
stability was obtained in the range 12.1-13.7 for poly(methyl methacrylate),
11.8-12.4 for poly(ethyl acrylate), 13.3-13.7 for polyacrylonitrile, and 11.9513.05 for the copolymer obtained from a 50% (by weight) mixture of methyl
methacrylate and ethyl acrylate. Eliassaf (1973) found a correlation between
the HLB of the suspension stabilizer and the morphology of PVC
[poly(vinyl chloride)J partieles prepared by suspension polymerization,
which varied from glassy beads to porous partieles having a high plasticizer
absorption.
However, as pointed out by Rosen (1978) the HLB method is useful only
as a rough guide to emulsifier selection since it indicates neither the
efficiency of the emulsificr (i.e., the concentration required) or its effectiveness (i.e., the stability of the emulsion produced) and does not take into
account the effect of temperature in varying the extent of hydration of
hydrophilic groups such as polyoxyethylene, with consequent change in the
action of the surfactant.

B. Effectsof Mixtures of Anionic and Nonionic Emulsijiers


As shown by Ono et al. (1974, 1975) decrease of the HLB of a mixed
emulsifier by use of an increasing proportion of a nonionic emulsifier
increases the stability of the latex to coagulation by electrolyte addition
despite an increase in its average partiele size. This is because purely
electrostatic stabilization by adsorbed ionic emulsifier is supplemented by
steric stabilization by the adsorbed nonionic emulsifier which effectively
decreases the van der Waals attractive force between the latex partieles
(which causes them to coalesce), thereby increasing their stability.
All the systems studied by Ono et al. inelude a large proportion of a
monomer of moderate water solubility (e.g., methyl methacrylate, acrylonitrile) so that partiele nueleation by the oligomeric precipitation mechanism
(Fitch and Tsai, 1971) is likely to be dominant. In cases (e.g., styrene) for
which micellar nueleation of latex partieles is dominant above the CMC of
the emulsifier another effect is also important. This is the formation of
mixed micelles that are much larger than those formed by the anionic
surfactant alone but which have a much lower surface charge density and
which may, therefore, increase the efficiency of initiation by negatively
charged radicals (e.g., SO;) from the aqueous phase. Working with a
mixture of sodium dodecyl sulfate and a tridecyloxypoly(ethyleneoxy)ethanol (Emulphogene BC-840, GAF), Piirma and Wang (1976) found that
addition of the ionic emulsifier reduced the micellar weight from 16,800
for the nonionic

alone to 3,500 with 50 mol

% anionic

emulsifier, although

A. S. Dunn

228

the micellar weight of the anionic emulsifier alone was much larger (23,500):
thus, there is a much larger number of much smaller micelles in the
mixed system than is obtained with either component alone. This was
reflected in an increase in the number of latex particIes formed and in the
subsequent steady rate of polymerization and in a reduction of the breadth of
the particIe size distribution. The highest rate and narrowest particIe size
distribution was obtained with 17 mol % of the ionic emulsifier when the
average micellar weight was 6,000: it was found that the probability of a
micelle nucIeating a latex particIe was a maximum at this composition
aIthough the probability remained low as an absolute value (0.00038).

C.

Preparation 01 Monodisperse Latexes

In the absence of particIe coalescence (i.e., when relativeiy large concentrations of emulsifiers are present sufficient to stabilize all primary
particIe nucIei formed) the essential condition for the' formation of a latex
with a very narrow particIe size.distribution is that the nucIei of all particIes
should be formed at very neariy the same time and should subsequently
grow at equal rates. The use of mixtures of anionic and nonionic emulsifiers
pro vides the requisite conditions as first shown by Woods et al. (1968).
AIthough initiation of emulsion polymerization with a water-soluble initiator has hitherto been credited with a 100% efficiency, the efficiency of
initiation of styrene by persulfate in seeded polymerizations of styrene in
which sodium dodecyl sulfate was used as the emulsifier has recently been
shown by Hawkett et al. (1980) to be rather low. Using an -10% emulsion
with 0.42% sodium dodecyl sulfate on the water phase, they find at 50C an
initiator efficiency of 44% at low persulfate concentrations (1.6 x 10-5
mol dm - 3) falling to only 1% at high concentrations (8.33 x 10- 2
mol dm - 3). Consequently, any improvement in initiator efficiency
should increase the rate of particIe nucIeation. Evidently, the increased
surface area and reduced surface charge density of mixed micelles achieves
this. The seed latex used by Hawkett et al. was monodisperse: it was prepared at a high temperature (90C) using sodium dioctylsulfosuccinate
(Aerosol MA, Cyanamid) as emulsifier with persulfate initiation. The high
temperature would give a high rate of radical formation from the initiator
but this emulsifier, having two alkyl chains, presumably forms micelles
with a lower surface charge density than emulsifiers with single alkyl
chains.
It has often been pointed out that the approach of a charged radical to a
charged micelle requires a Coulombic repulsive energy barrier to be overcome even though counterion binding reduces the effective charge on the

6.

Effects of the Choice of Emulsifier in Emulsion Polymerization

229

micelle below that which would be inferred from the aggregation number of
the micelle. Consequently, it has been surmised that a sulfate radical might
hav.e to add several monomers in an aqueous-phase polymerization before
the van der Waals attractive force between the growing radical and the
micelle suffices to overcome the Coulombic repulsion. Some evidence in
support of this is provided by the observation (Waite, 1977) that monomers
even less soluble in water than styrene (e.g., tert-butylstyrene, dodecyl
methacrylate) cannot be polymerized in an emulsion with a water-soluble
initiator unless the monomer solubility in the aqueous phase is increased by
addition of alcohol or acetone. On the other hand the only actual estimate
of the magnitude of the energy barrier (Fitch and Shih, 1975) shows it to be
negligible in the case of latex particles stabilized by ionic initiator end
groups only. Exchange of individual emulsifier molecules between micelle
and solution is rapid but a minimum Cs alkyl chain is needed for micelles
to be formed at all by n-alkyl sulfates or carboxylates, from which one
might infer that a minimum of four monomers must be added before a
sulfate radical can enter a micelle or a latex particle.
The first published formulation for producing a mono disperse polystyrene seed latex (Bobalek and Williams, 1966) involved the use of octylphenoxypolyethylene

oxide (Triton X-100, Rohm and Haas) 1.7 %with 0.08-

0.17% sodium dodecyl sulfate on the water phase. Unfortunately, as shown


by Kamath et al. (1975),this nonionic emulsifierhas a labilehydrogen atom
and is liable to peroxidize. Consequently, polymerization rates are not
reproducible unless the emulsifier used is a fresh sample, and transfer to
emulsifier invalidates any inference from molecular weight data in which
this factor is neglected. Emulphogene BC-840 was introduced as a product
with similar propert'ies free from this difficulty; but it is possible that
equivalent results could be obtained with other mixtures of ionic and
nonionic emulsifiers, indeed Renex 30 (Atlas) (a polyoxyethylene alkyl ether)
in combination with sodium dodecylbenzene sulfonate has been found
suitable (Doherty, 1978).

ill.

Emulsion Polymerization witb Nonionic Emulsifiers

Nonionic surfactants were developed subsequent to the ionic types and


are not normally used as the sole emulsifying agent in emulsion polymerizations. Consequently, the characteristics of emulsion polymerizations using
only nonionic emulsifiers have received little attention apart from a series of
papers from Medvedev's group in the Soviet Union, although an understanding of these is a prerequisite for the interpretation of their action in
combination with ionic emulsifiers. The Bobalek-Williams recipe produces

A. S. Dunn

230

latex partic1es of eomparatively large size (195-230 nm diameter) relative to


those obtained with ionie emulsifiers (typieally about 50 nm), although
smaller than obtainable by emulsifier-free emulsion polymerization of styrene (500 nm). The meehanism of emulsion polymerization with nonionie
emulsifiers alone appears to differ fundamentally from that of emulsion
polymerization with ionie emulsifiers. Instead of all the latex partic1es being
formed during a eomparatively short initial Interval I and then remaining
eonstant in number during Interval II though inereasing in size at a steady
rate until all monomer is absorbed by the latex partic1es and Interval III
eharaeterized by a diminishing rate of polymerization eommenees, the size
of the partic1es remains eonstant throughout the reaetion but their number
inereases as eonversion proeeeds (Medvedev et al., 1971). The eharaeteristies of emulsion polymerization with nonionie emulsifiers alone would thus
appear to have mueh in eommon with emulsifier-free emulsion polymerization in whieh the final latex partic1es are formed by primary partic1es
whieh eoalesee until a stable size is reaehed. This would imply that even
though mieelles are present and presumably solubilize styrene the loeus of
polymerization is exc1usively in the aqueous phase. Adsorption of the
nonionie emulsifier would greatiy inerease the size whieh oligomerie radicals eould attain without preeipitating. However, the results of Piirma and
Chang (1980) shed a different light on these experiments. Although they also
observe a eonstant rate for the first 40% of the reaetion (an extent of
reaetion exeeeded in few of the experiments reported by the Soviet group),
they find a marked inerease in rate above 40%. Eleetron mieroseopy shows
that this is not attributable to the gel effeet but to an inerease in the number
of reaetion loei beca use of the generation of a seeond erop of latex partic1es.
They attribute this to the faet that the nonionie emulsifiers are soluble in
styrene, so that the emulsifier eoneentration in the water phase inereases as
the monomer droplets are eonsumed and all monomer is present in the
swollen latex partic1es. The Soviet workers determined partic1e size by a
nephelometrie method, and it is possible that the average partic1e size may
remain eonstant, as they find when there is a bimodal distribution with
both large and small partic1es inereasing in size but with the number of the
small partic1es inereasing while the number of large partic1es remains
eonstant.
There is c1early seope for further investigations of the behavior of
nonionie emulsifiers in emulsion polymerizations.

IV.

EmulsioD PolymerizatioD with IODicEmulsifiers

The Smith-Ewart theory was eoneeived with referenee only to ionie


emulsifiers. Ir the mode of aetion bf nonionie emulsifiers differs from that of
ionie emulsifiers, the theory may be inapplieable either when nonionie

6.

Effects of the Choice of Emulsifier in Emulsion Polymerization

231

emulsifiers are used alone or when mixtures of ionic and nonionic emulsifiers are used. The parameter characteristic of the emulsifier in SmithEwart theory is as, the area occupied by an emulsifier "in a saturated
monolayer at the polymer-water interface. The criterio n for the transition
from Interval 1, in which the latex particles are formed, to Interval 11, in
which the particles increase in size at a uniform rate while their number
remains constant, is that the total surface area of the latex particles should
ha ve increased to an extent sufficient to adsorb all the emulsifier originally
present in the aqueous phase as micelles. It is assumed that the area
occupied by an emulsifier molecule on the latex particle surface and in the
micelle is the same. Since the presence of micelles containing solubilized
monomer is postulated as essential to the nucleation of latex particles, it
follows that it should not be possible to nucleate additional latex particles
once all emulsifier has been adsorbed on the existing particles. The reason
that emulsifier should adsorb on the latex particles rather than aggregate in
micelles in solution is not explained, although it is an experimental fact that
the micelles do usually disappear from solution (as shown by an increase in
the surface tension) at approximately the same time as new latex particles
cease to formo Actually, it is probable that the van der Waals interaction
energy between an emulsifier molecule and a latex particle is greater than
that between an emulsifier molecule and a micelle simply because the latex
particle is larger, but a quantitative demonstration of this has not yet been
gIven.
Despite the fact that Smith-Ewart theory (or Gardon's development of
it) does apply well to the emulsion polymeri?:ation of styrene at least, none
of these postulates seems to be strictly true. Micelles are not essential for
the nucleation of la'tex particles since the possibility of emulsifier-free
emulsion polymerization has been amply demonstrated. When high concentrations of a weakly adsorbed emulsifier are used, micelles do not
disappear at the beginning of Interval 11 and even remain at the end of the
reaction (Bakker, 1952), it being impossible to carry out a soap titration on
the latex because the surface tension has not risen above the value
characteristic of the CMC of the emulsifier (AI-Shahib, 1977). The value of
as is not independent of the nature of the polymer as was initially assumed
[and apparently confirmed by early experiments (Sawyer and Rehfeld,
1963)] but increases with the polarity of the polymer and the temperature
(Piirma and Chen, 1980) (Table III). Indeed if as were independent of temperature the van't Hoff Isochore would imply a zero enthalpy of adsorption
in contradiction to the general observation that physical adsorption is an
exothermic process. The real criterion for the cessation of micellar particle
nucleation appears to be not the absence of micelles but the fact that the
number and size of the latex particles has increased sufficiently to ensure
c~pture of all radicals generated in the aqueous phase in competition with
any micelles that may remain. Hence, as is probably not the parameter

A. S. Dunn

232
TABlE

111

Area (a's) Occupied by a Surfactant Molecule at a PolystyreneWater Interface at the Critical Micelle Concentration: Effect of
Alkyl Chain length
(a/A2)
Potassium carboxylates

Chain length
C10
C12
C14
C16
C18

Sodium alkyl sulfates


Piirma and Chen
47C
(1980)

Maron el al. AI-Shahib


50C
60C
(1954)Q
(1977)

68
66
48

41
34
25
23

52
40
25

EffeCI of temperature
Sodium dodecyl,sulfate
TemperatureOC
a~A2
Q

(Piirma and Chen, 1980)

22
47

37
49

47
52

On butadiene-styrene rubber.

required, but if it is used it needs to be determined from measurements on


the right polymer at the relevant temperature, and values derived, for
example, from studies of adsorption at the air-water interface cannot be
used. For non polar polymers, the magnitudes of polymer-emulsifier and
emulsifier-emulsifier interactions appear to be similar so that the CMC is
reached in solution at just about the concentration that is also in equilibrium with a saturated monolayer on the polymer surface. In these circumstances the soap titration method of Maron et al. (1954), involving the
addition of a surfactant solution to a latex in which the particles are only
partially covered with adsorbed surfactant initially, can be used to determine as reasonably satisfactorily; but in the case of more polar polymers
the polymer-emulsifier interaction is evidently considerably lower than the
emulsifier-emulsifier interaction and micelles are formed in solution long
before enough emulsifier has been adsorbed to complete the coverage of the
polymer surface (Paxton, 1969). The as value obtained by soap titration
is the area occupied by a surfactant molecule on the polymer surface at the
concentration at which the CMC is reached in solution, precluding any
further increase of the concentration of surfactant monomers in the aqueous
phase (cf. Table IV). It appears also (Drban, 1980) that a saturated
monolayer may be formed at the air-water interface before micelles form in

6.

233

Effects of the Choice of Emulsifier in Emulsion Polymerization

TABLEIV
Adsorption of Soaps on PVCa.b
Soap
Potassium
Potassium
Potassium
Potassium

dodecanoate
tetradecanoate
tetradecanoate
tetradecanoate

lOs (salt)jmol dm-3

a;jA2

aslA2

2.7KCI
O
1.7 K2HP04
5.7 K2HP04

51
52.9
47
40

26.5
24.4
25
25

From Paxton, 1977.

b as was determined by soap titration,


the Langmuir adsorption isotherm.

as by extrapolation

of a linear plot of

the bulk of the solution, so that differing estimates of as are obtained


according to whether surface tension or conductivity measurements are
used to determine the CMC. Urban used both methods on the same
surfactant sample, whereas most other investigators ha ve confined themselves to a single technique and have attributed any discrepancies with
literature results obtained by alternative techniques to differences in sample
purity (which is certainly the major factor in explaining discrepancies in the
reported results).
A.

Effect 01 Emulsifier on Number 01 Latex Particles Formed

According to the theories of Smith and Ewart and of Gardon, the effect
of emulsifiers on the number of polymer latex particles formed N (and hence
on the rate of polym~rization during Interval 11)is determined by the area
as which the emulsifier molecule occupies in a saturated monolayer at the
polymer-water interface,
N oc (asS)3/5

where S is the surfactant concentration. Although Bartholome et al. (1956)


showed that the slope of a logarithmic plot of the rate of emulsion
polymerization of styrene against surfactant concentration could indeed be
fitted with a straight line with the expected gradient of 0.60, the only studies
in which a series of different surfactants were used (Ivanova and Yurzhenko,
1958; Onischenko et al., 1970) showed that the latex particle size decreased
rapidly and the rate of polymerization increased correspondingly as the
alkyl chain length of the emulsifier increased. However, at the concentrations used in these studies the lower members of the homologous series
of emulsifiers were below their CMC, and the effect is basically similar to
that observed when the concentration of a single emulsifier is increased .
from below to above its CMC.

A. S. Dunn

234

Using equal weight concentrations of alkyl carboxylates (all above the


CMC of the emulsifier), Carr el al. (1950) observed only a slight increase in
the Interval 11 rate, although there was a large difference in the time
required for a steady rate to be attained. In fact, the literature values of as
(Table III) decrease with increasing alkyl chain length in the homologous
series of alkyl carboxylates so that when equal concentrations of these
emulsifiers are used a decrease in Interval 11 rate should be expected.
Actually, when equal (molar) concentrations chosen to be above the CMC
of the lowest member of the homologous series are used a slight decrease in
particle size with a corresponding increase in the Interval II polymerization
rate is found (Dunn and AI-Shahib, 1978). However, the CMC decreases as
alkyl chain length increases so that although the total emulsifier concentration is the same, the concentration of micellar emulsifier increases
with alkyl chain length. Since, for styrene, a large increase in the number of
latex particles formed is observed as emulsifier concentration is increased
through the CMC, showing the micellar nucleation of latex particles to be
dominant, this suggests that the increase in rate may be attributable to the
increase in the concentration of micellar emulsifier and that it may be the
concentration of micellar emulsifier that should be considered. Although
the Harkins model postula~s micellar nucleation of latex particles, the
micelles are not actually considered in the Smith-Ewart theory: it is
assumed that the total concentration of emulsifier is all adsorbed on the
latex particles at the end of Interval I and that the equilibrium concentration in solution (which is likely to be of the order of the CMC) may
be neglected. This is possible for many commonly used emulsifiers that have
a low CMC (e.g., potassium stearate) but it is unsatisfactory for the more
weakly adsorbed emulsifiers (e.g., resin acid salts and even for sodium
dodecyl sulfate).
When the concentration of emulsifiers are chosen so that the micellar
concentrations are equal (although total concentration may differ greatly),
the size of the latex particles formed is the same. This means that the
number of latex particles formed is also the same and, consequently, the
Interval 11 rates of polyq.erization are also the same (Dunn and AI-Shahib,
1980). The duration ofInterval I does increase for the more weakly adsorbed
emulsifiers, however (Fig. 2).
B.

The Effect of Micelle Size

Considerable discrepancies exist between the results of different techniques for determining micelle size (Anacker el al., 1964); but it has been
shown (Kratohvil, 1980) that serious discrepancies can arise from oversimplification of the treatment of experimental results and that the results of

6.

235

Effects of the Choice of Emulsifier in Emulsion Polymerization


100

ea

e 10

'

e 12

::./
O

e 16

/!

~ 40

e 14

LL l
600

600
Time(mio)

'

600

e la

,l

600

600

60

Fig. 2. Course of the emulsion polymerization of styrene at 60C using the same
concentrations (0.012 mol dm - 3)above their critical micelle concentrations of the homologous
series of sodium alkyl sulfates. lnitiator: 0.20% KzSzOs on aqueous phase. (From AI-Shahib,
1977.)
different techniques can be reconciled when proper allowance for the effects
of counterion binding is made, although unfortunately the only emulsifier
for which sufficient data exists to allow this to be done is sodium dodecyl
sulfate.
However, although the absolute values of the aggregation numbers may
be incorrect, there is no doubt (Table V) that micelle size increases with
alkyl chain length. Consequently, at constant micellar concentration the
number of micelles must decrease as their size increases. Clearly the number
of micelles initially present do es not determine the number of latex particles
formed. However, at a given micellar concent~ation, the total surface area of
the micelles is constant because the increasing size of the micelle compensates for their decreasing number as alkyl chain length is increased.
Thus, it seems that the total surface area of the micelles may be theofactor
determining the rate at which initiator radicals are captured by micelles and
hence the number of latex particles formed.
TABLEV
Total Surface

Areas of the Micelles in Solutions of Sodium Alkyl Sulfates


Containing 0.012 mol dm-3 Micellar Surfactant
Alkyl chain length

Aggregation numbera
10 - 3 Total surface area m z dm - 3 b
a
b

From Aniansson et al., 1976.


From Al-Shahib, 1977.

27
4.4

CIO

CIZ

CI4

CI6

41
4.5

64
4.3

80
4.4

100
4.5

CIS

A. S. Dunn

236

Because of their greater size, the efficiency of capture of initiator or


oligomer radicals by the latex particles is greater than that of the micelles.
Consequently, when the area of the latex particles has increased sufficiently
to ensure capture of all initiator radical s, particle formation ceases because
there is no chance of any remaining micelles capturing radicals. The small
micelles of the lower members of homologous emulsifier series are less
efficient in capturing initiator radical s than the larger micelles of the higher
members of the series, which leads to a longer duration of Interval I which
should mean that the particle size distribution is much broader when the
critical surface area of latex particles is ultimately attained. The efficiency of
radical capture by micelles will differ for different emulsifiers and will be the
factor determining the breadth of the particle size distribution in batch
polymerizations.

.
V. Latex Agglomeration
The size of the latex particles obtained by an emulsion polymerization
using an ionic emulsifier is in the 3O-100-nm range. Particles exceeding 100
nm in diameter can be produced by multistage seeded emulsion polymerizations: and although this technique has been used in the laboratory to
obtain model colloids it is not suitable for use on an industrial scale,
although latexes with an average diameter in excess of 100 nm are required
for some applications. The viscosity of dispersions increases with decreasing
average particle size although it also depends on particle size distribution.
The solids content of a latex produced by the mutual recipe using 100 parts
by weight of monomer to 180 parts water carried to a 70% conversion is
28% after the residual monomer has been removed. This is much too low for
many applications in which the product is used as a latex: solids contents in
excess of 60% are often required. However, if the latex is concentrated by
evaporation it sets to a stiff paste when the solids concentration rises to
about 40%. This can be avoided if the average particle size is increased by
agglomeration. The principIe of agglomeration is that the latex is destabilized slightly when particle coalescence occurs until the total surface
area of the particles has been reduced sufficiently for stability to be restored.
This may be brought about by the addition of a solvent that swells the
particles and which has to be removed after agglomeration, by partial
neutralization of carboxylate emulsifiers, by addition of electrolytes, or by
freezing which increases the effective electrolyte concentration when some of
the water separates as ice. Unfortunately, all these processes are more or
less difficult to control and generally re'Sult in partial coagulation of the
latex as well as agglomeration. The process of pressure agglomeration

6.

Effects of the Choice of Emulsifier in Emulsion Polymerization

237

developed by the International Synthetic Rubber Company (1971) appears


to have overcome these difficulties; this process involves subjecting a stream
of latex in turbulent flow to a large drop in pressure. Although no investigations into the mechanism of this process have been reported, it may depend
on the surface of the latex being less than completely covered with emulsifier
at the end of the polymerization. The drop in pressure might cause rapid
desorption of some emulsifier permitting some coalescence to occur until
equilibrium is established again at the lower pressure. For the process to be
successful the initial solids content has to be raised to about 35% by
evaporation. After agglomeration the solids content may be raised to 60%
by further evaporation before the latex viscosity exceeds 20 P, becoming too
viscous to handle conveniently. Ir the critical factor in pressure agglomeration is, in fact, the rates of desorption and readsorption of emulsifier in
comparison with the rate of slow coalescence of the latex particles, the
result obtained should depend on the emulsifier used and might provide a
means for studying rates of emulsifier adsorption.

VI.

Other Effects of Emulsifiers

Various other effects of emulsifiers have been observed in particular


systems but systematic studies of these effects in relation to emulsifier
structure have not yet been undertaken.
A.

Monomer So/ubilization

In general the amount of hydrophobic monomer that can be solubilized


in emulsifier micelles increases with the alkyl chain length of the emulsifier
parallel to an increase in micelle size. In the case of styrene there is a
definite correlation between solubilizing power of the emulsifier and the rate
of polymerization during Interval I (Fig. 3), although the number of latex
particles ultimately formed and consequently the ultimate rate of polymerization during Interval 11 is independent of the amount of monomer
solubilized (Fig. 2). Although no investigation of particle size distribution in
relation to amount of monomer solubilized has been published it is to be
expected that the width of the particle size distribution would decrease with
increase of the amount of monomer solubilized and consequent decrease in
the duration of Interval I.
Although emulsifiers do also solubilize the more polar monomers, which
have a greater solubility in wath, to about the same extent as nonpolar
monomers. this amount is usually negligible compared with the amount of

A. S. Dunn

238
P,'I

~'" 1.2

I
(a)

""

..
~O.B
;Q

"
~

00.4

<
~
o

8
16
24
Ofopolymerizedin 3 hr

P,'I-;241(b)
""

32

'"
:o..
.1:1

~"

16

'O

..
~
8
..
..
>.
~
O
O

0.4
0.8
1.2
DMAB solubilized(gdm'3)

Fig. 3. (a) Correlation of the initial (Interval 1) rate in the eopolymerization of


butadiene and styrene with the amount of dimethylaminobenzene (DMAB) solubilized in the
surfaetant solution. (b) Correlation of the amount of styrene solubilized in surfaetant solutions
with the solubilization of DMAB. Surfaetants: O Rubber Reserve Soap; O sodium rosinate;
D. potassiumtetradeeanoate: O potassiumdodeeanoate: \l sodiumoleate. (From Kolthoffet
al., 195\. Reprodueed with permission of John Wiley and Sons, Ine.)

monomer present in true solution so that it has no discernible effect on the


course of the polymerization in such cases.
B.

Po/ymer So/uhi/ization

When the surfactant concentration is high relative to the polymer


concentration sufficient surfactant may be adsorbed on individual polymer
molecules to prevent their coalescence to form latex particles or indeed to
disperse preformed polymer to form clear solutions in which the solute
behaves as a polyelectrolyte. But the influence of such effects on the course
of emulsion polymerization reactions has not been elucidated. Sata and Saito
(1952) showed that poly(vinyl acetate) precipitated from acetone solution
with water could be solubilized in sodium dodecyl sulfate solutions after
removal of the acetone by dialysis. To obtain a clear solution at 20C, a
weight of surfactant 5-10 times that of pol~mer was required. Although this
greatly exceeds the surfactant concentrations normally used in emulsion

6.

Effects of the Choice of Emulsifier in Emulsion Polymerization

239

polymerizations this effect could mean that the lowest molecular weight
fractions of polymer could remain in the aqueous phase at high surfactant
concentrations.
C.

lnitiator Decomposition lnduced by Emulsifier

Medvedev and Sheinker (1954; Medvedev, 1957) advanced a theory of


emulsion polymerization in which the reaction was supposed to take place
within the adsorbed emulsifier layer. Induced decomposition of the initiator
by radical s derived from the emulsifier formed an essential step in his
kinetic scheme. Although Ryabova et al. (1975a,b) subsequently showed that
the experimental evidence on which Medvedev's theory was based was
unreliaqle, it is nevertheless important to know whether decomposition rate
constants for initiators derived from experiments in homogeneous solutions
are applicable (as is usually assumed) in the presence of emulsifiers and
emulsified monomer. It is plausible that the electrical double layer surrounding micelles and emulsified monomer droplets might affect the rate of
decomposition of ionic initiators, although if it did the rate of decomposition might also be expected to be affected by the ionic strength of
aqueous solutions whereas, in the case of persulfate at least, no effect was
found (Kolthoff and Miller, 1951). Nevertheless, there is a considerable
body of evidence in the literature that the rate of persulfate decomposition
may be increased by a factor of two or thereby in the presence of
monomers. Although this would mean that the rate of consumption of
initiator would be greater than otherwise ~xpected, it does not necessarily
mean that the rate of initiation would also be enhanced. Where titrimetric
methods are used' to determine persulfate concentrations, higher concentrations have to be used than would be used in polymerizations: the
effect of second-order reactions involving persulfate might be negligible at
lower concentrations. Induced chain decompositions of persulfate, e.g.,
SOi
R.

+ RH =
R.
+ S20~- = RSOi

+ HSOi
+ SOi

may be prevented in the presence of monomer if all the primary radicals


react with the monomer. Although such an induced chain decomposition
increases the rate of consumption of the initiator, it would usually affect the
radical concentration little. Dunn and Tonge (1972) used diphenylpicryl
hydrazyl as a radical scavenger to measure the rate of initiation (rather than
the rate of decomposition) of persulfate under polymerization conditions,
although 50% ethanol solutioQs had to be used beca use diphenylpicryl
hydrazyl is insoluble in watef'. No evidence was found for any large
enhancement of the initiation rate by persulfate due to the presence of
..

A. S. Dunn

240

additives as was implied by the results reported by Morris and Parts (1968):
their results refer to rates of persulfate decomposition but they seem to ha ve
been vitiated by interference by the additives in the titrimetric method used
for determining persulfate concentrations.
Yurzhenko and Brazhnikova (1956) reported that hexane or toluene,
dispersed in water without an emulsifier, increased the rate of decomposition of persulfate by a factor of two or three. van der Hoff (1967)
showed that the rate of decomposition of persulfate was much increased in
dispersions of octane or benzene. Vinogrado v et al. (1962) found that
whereas persulfate decomposed at a much enhanced rate in an emulsion of
xylene with potassium oleate as emulsifier compared with a similar emulsion using the same concentration by weight of sodium dibutylnaphthalene sulfonate, the rate of polymerization of styrene emulsions was
actuaIly greater with the latter emulsifier! Volkov and Kulyuda (1978)
found nOl1ionicsurfactants to enhance the rate of persulfate decomposition:
the effect of a polyethyleneoxyalkylphenol was grea~er than that of a
polyethyleneoxymonododecyl maleate, which exceeded that of poly(vinyl
alcohol), which in turn was greattr than that of polyethyleneoxymonotetradecyl maleate; but the rate of polymerization of styrene was greater with the
second of the above emulsifiersthan with the first. Ryabova et al. (1977)
found that persulfate decomposed 50 to 100% faster in potassium decano ate
solutions than in water, whether the concentration was above or below the
CMC of the emulsifier: emulsification of styrene resulted in a further 100%
increase in the decomposition rateo Liegeois (1971) found that the rate of
initiation of vinyl chloride in aqueous solution by persulfate was that
expected from the rate of decomposition in the absence of additives,
although the rate of decomposition was increased by a factor of three in the
presence of vinyl chloride. Similariy, AIIen (1958) found that the rate of
initiation of methyl methacrylate emulsions by persulfate (as inferred from
measurements of overaIl polymerization rate and the degree of polymerization of the polymer) did not vary as the concentration of sodium dodecyl
sulfate emulsifier was increased. The general conc1usion is that the presence
of emulsifiers and emulsified monomers does not increase the rate of
initiation by persulfate but that the induced chain decomposition may
increase the rate of consumption of the initiator by a factor of three or
more. Thus, the effect will usually be negligible in laboratory studies in
which conditions are usually chosen so that the decrease in the concentration of the initiator will be negligible in the duration of the polymerization reaction, but under industrial conditions wasteful consumption of
persulfate could lead to premature exhaustion of the initiator and incom-

pletepolymerization.

6.

D.

Effects of the Choice of Emulsifier in Emulsion Polymerization

241

Catalysis 01the lnitiation Reaction

From the discussion above, it is cIear that there is no evidence for


catalysis of persulfate initiation in emulsion polyrnerization systems.
However, many ionic reactions have been shown to be subject to large
catalytic elfects in the presence of emulsifier micelles (Fendler and Fendler,
1975) so that the question arises as to whether there are any radical
reactions that are subject to micellar catalysis and whether this phenomenon plays any part in any emulsion polymerizationsystems. Prima Jacie
evidence that micellar catalysis may be important when emulsified monomer is allowed to polymerize thermally is provided by the work of
Asahara et al. (1970, 1973)who find that several emulsifiersdecrease the
energy of activation for thermal initiation of alkyl methacrylate and styrene.
In particular, the energy of activation for thermal initiation of styrene
emulsified with sodium tetrapropylene benzene solfonate was reported as

53 kJ mol- 1, much lower than any value determined in bulk. Hui and
Hamielec's value of 115 kJ mol-I (1972) seems to be representative of the
data available on th(;:rmalinitiation in bulk. The concIusions of Asahara et
al. are based on observations of the temperature dependence of the degree
of polymerization and are open to several objections.
1. The viscosity-average degree of polymerization is measured when the
number.oaverage degree of polymerization is required, although this will not
lead to any error in deriving an energy of activation provided the molecular
weight distribution does not change with temperature.
2. It is assumed that all transfer reactins can be neglected so that the
kinetic chain length can. be taken as half the degree of polymerization: this
seems dubious when it is generally agreed that the degree of polymerization
of thermally polymerized styrene in bulk is limited by transfer to monomer
and when the possibility of using oil-phase initiation for emulsion polymerization is dependent on the escape of one of a pair of.radicals generated in
the oil phase to the aqueous phase which seems likely to involve transfer
either to monomer or to emulsifier.
3. In their calculation, they seem to have confused the energy of
activation for propagation with the overall energy of ~ctivation for polymerization: when this is corrected the energy of activation for initiation
becomes 80 kJ mol- 1 which is still significantly lower than the bulk value of
115 kJ mol- 1 but which may stiII be incorrect if the second objection above
is justified.
An alternative method of calculating the energy of activation for initiation in emulsion polymeriza60n was introduced by Bartholom et al.

A. S. Dunn

242

(1956). It depended on the variation of the number of latex partic1es formed


N with temperature. Unfortunately, they have overlooked the fact that the
partic1e growth rate J1which appears to the power -~ in the Smith-Ewart
expression for the number of latex partic1es formed contains the propagation rate constant kp which is temperature dependent. It has also
recentIy been realized that another factor on which N depends, the area
occupied by a surfactant molecule at the polymer-water interface as, is also
temperature dependent. Dunn et al. (1981) observed that the temperature
dependence of N in the thermal polymerization of styrene differed from
different emulsifiers. It seems unlikely that the differences can be wholly
explained by differing enthalpies of adsorption of the emulsifiers and, if not,
this observation implies that the energy of activation for thermal initiation
of styrene in emulsion depends on the emulsifier used. Participation of
emulsifiers in thermal initiation (and probably also in initiation by oilsoluble initiators) is most probably attributable to transfer to emulsifier and
desorption of the emulsifier radical from the micelle .or latex partic1e into
the aqueous phase: the rates of these processes are likely to differ with the

emulsifier.

'

E. Transfer lo Emu/sifiers
The extent and importance of transfer reactions to emulsifier molecules
in emulsion polymerization has not been extensively investigated. Tr.ansfer
to an emulsifier molecule that is subsequentIy desorbed from the micelle or
latex partic1e would seem to be an attractive mechanism that would
facilitate the escape of radicals of the less water-soluble monomers to the
aqueous phase, a process of greater importance than had previously been
thought. However, most emulsifiers do not contain any particularly active
hydrogen or other substituent atoms, and the topology of the system
implies that transfer constants determined from experiments in homogeneous solution do not necessarily apply to latex systems. Thus, Okamura
et al. (1958) found from experiments in homogeneous solution that the rate
constant for the transfer of poly(vinyl acetate) radicals to poly(vinyl alcohol)
(35 x 10-4 dm3 mol-l sec-l) was much higher than that to poly(vinyl
acetate) (1.5 x 10-4 dm3 mol-1 sec-1) so that it might be expected that
when incompletely hydrolyzed grades of poly(vinyl alcohol-acetate) are
used as emulsifiers in the emulsion polymerization of vinyl acetate, the
extent of formation of graft copolymers of poly(vinyl acetate) with the
emulsifier sho~ld be expected to decrease with acetyl content. In fact, the
contrary trend was observed. The acetyl blocks form the hydrophobic
portion of poly(vinyl alcohol-acetate) emulsifiers and are the only parts of
the molecule readily accessible to poly(vinyl acetate) radicals in the latex

6.

Effects of the Choice of Emulsifier in Emulsion Polymerization

243

particIes: consequently,most of the transfer that does occur must occur to


these acetyl blocks.
Extensive studies of the removal of adsorbed emulsifiers have been made
only with polystyrene latices. Exhaustive dialysis does not seem to be a
satisfactory procedure (Edelhauser, 1969) as it fails to remove all the
emulsifier, but it does seem to be possible to do this either by ion exchange
(Vanderhoff et al., 1970) provided the resin used has been properly prepared
or by the serum replacementtechnique (Ahmedet al., 1980).AftercIeaning,
the number of sulfate groups per polystyrene molecule is less than two.
When persulfate initiation is used two sulfate end groups per molecule
would be expected, but hydrolysis of the sulfate ester end group is quite
facile so that some of the original end groups may have been replaced by
hydroxyl. However, no critical experiment has yet been attempted. This
might involve the use of an initiator (e.g., sodium azobiscyanopentanoate)
that could only introduce carboxylate end groups in conjunction with
radioactively labeled sulfonate emulsifier (since there is no danger of sulfonate groups being hydrolyzed): radioactivity remaining in the polymer
after removal of adsorbed emulsifier would enable the extent of transfer to
emulsifier to be determined.
Kamath et al. (1975)showedthat the nonionic octylphenoxypolyethylene
oxide emulsifier Triton X-I00 contains a labile pro ton so that the molecular
weight of polystyrene prepared using this emulsifier is abnormally low
because of extensive transfer to the emulsifier.
In general, transfer to emulsifier does not appear to be very important in
the emulsion polymerization of styrene, a1though it has recently become
evident (cf. Hawkett et al., 1980) that even in this case, escape of radicals
from latex particIes' cannot be neglected. However, there is no evidence
available to show whether emulsifiers participate in the escape>process.
Transfer processes are likely to be more important in the case of those
radicals (e.g., vinyl acetate) that are more active than styrene but very little
data indeed pertaining to these has been published.

References
Adamson, A. W., (1976). "Physics and Chernistry ofSurfaces," 3rd ed. Wiley, New York.
Ahmed, S. M., EI-Aasser, M. S., Micale, F. J., Poehlein, G. W., and Vanderhoff, J. W. (1980). In
"Polyrner Colloids 11" (R. M. Fitch, ed.), p. 265. Plenum Press, New York.
Allen,P. W. (1958).J. Polym. Sci. 31, 206.
AI-Shahib, A. W. (1977). Ph.D. Thesis, Manchester.
Anacker, E. W., Rush, R. M., and Johnson, J. S. (1964).J. Phys. Chem.68, 81.
Aniansson, E. A. G. et al. (1976). J. Phys. Chem.SO,905.
Asahara, T., Seno, M., Shiraishi, S., and Arita, Y. (1970). Bull. Chem. Soco Jpn. 43, 3895.

A. S. Dunn

244

Asahara, T., Seno, M., Shiraishi, S., and Arita, Y. (1973). Bul/. Chem. Soco Jpn. 46, 249.
Bakker, J. (1952). Phi/ips Res. Rep. 7, 344.
Bartholom, E., Gerrens, H., Herbeck, R., and Weitz, H. M. (1956). Z. E/ektrochem. 60, 334.
Benetta, G., and Cinque, G. (1965). Chim. Ind. (Mi/an) 47, 500.
Blackley, D. e. (1975). "Emulsion
Polymerisation:
Theory and Practice," p. 308, Fig. 7.22.
Applied Science, London.
Bobalek, E. G., and WiIliams, D. A. (1966). J. Po/ym. Sci. Part A-/ 4, 3065.
Carr, C. W., KoltholT, J. M., Meehan, E. J., and WiIliams, D. E. (1950). J. Po/ym. Sci. 5, 20\.
Davies, J. T. (1957). Proc. InI. Congr. Surface Activity, 2ndVol. 1, p. 426. Butterworths,
London.
Doherty, J. V. (1978). Unpublished
work.
Dunn, A. S., and AI-Shahib, W. A. (1978). J. Po/ym. Sci. Po/ym. Chem. Ed. 16,677.
Dunn, A. S., and AI-Shahib, W. A. (1980). In "Polymer Colloids 11", (R. M. Fitch, ed.), p. 619.
Plenum Press, New York.
Dunn,

A. S., and Tonge,.c.

J. (1972). Po/ym. Preprints

13, 126\.

Dunn, A. S., Said, Z. F*,


and Hassan, S. A. (I98\). In A.C.S.
A. E. Hamielec, eds.), p. 11 \. American Chemical Society,
Edelhauser, H. A. (1969). J. Po/ym. Sci. Part C 27, 29\.
Eliassaf,J. (1973). Polim. Vehomarim P/ast. 3, 9 (in Hebrew); cf.
Fendler, J.~., and Fendler, E. J. (1975). "Catalysis in Micellar
Academic Press, New York.

Symp. Ser.ll,
Washington

(D. R. Bassett and


D.e.

Chem. Abstr. 80, 121619u(l974).


and Macromolecular
Systems,"

Fiteh, R. M., and Shih, L.-B. (1975). Prog. Col/oid Po/ym. Sci. 56, \.
Fitch, R. M., and Tsai, C. H. (1971), In "Polymer
Press, New York.

Colloids"

(R. M. Fitch, ed.), p. 73. Plenum

Greth, G. F., and Wilson, J. E. (l96\). J. App/. Po/ym. Sci. S, 135.


Griffin, W. C. (1949). J. Soco Cosmet. Chem. 1, 31\.
Griffin. W. C. (1954). J. Soc. Cosmet. Chem. 5, 249.
Griffin. W. e. (1980). Emulsions Kirk-Othmer
"Encyc1opedia
of Chemieal Teehnology"
3rd
Edn. 8, 900.
Hawkelt, B. S., Gilbert, R. G., and Napper, D. H. (1980). J. Chem. Soc. Faraday Trans./76,
1323.
HoplT. H., and Falka. \. (1965). Makromo/. Chem. 88, 54 (in German); Br. Po/ym. J. 2, 40 (1970).
Hui, A. W., and Hamielec, A. E. (1972). J. App/. Po/ym. Sci. 16,749.
lntemational
Synthetie Rubber Co. (D. A. Bennett) (I97\). U.S.P. 3 573 243.
lvanova. N. Y.. and Yurzhenko, A. l. (1958). Kol/oid Zh. 22, 37 [English trans.: Col/oid J. USSR
22,39J.
Jagodie, F., and Fajt. B. (1976). Kem. Ind. 25, 13 (in SJovenian); cf. Chem. Abstr. 86,44067g
(1977).
Jagodie, F., Abe. M., and Ogrizek, N. (1975). Kem. lnd. 24, 591, 659 (in Slovenian); cf. Chem.
Abstr. 85, 33460h, 63365y (1976).
Kamath, V., Morton, M.. and Piirma, l. (1975). J. Po/ym. Sci. Po/ym. Chem. Ed. 13,2087.
KoltholT, J. M., and MiIler. J. K. (1951). J. Am. Chem. Soco 73, 3055.
KoltholT, J. M.. Meehan, E. J., and Carr, e. W. (1951). J. Po/ym. Sci. 6, 73.
Kratohvil, J. P. (1980). J. Col/oid Interface Sci. 75, 271.
Liegeois, J. M. (1971). J. Po/ym. Sci. Part C 33, 147.
Maron, S. H., Elder, M. E., and Ulevitch, l. N. (1954). J. Col/oid Sci. 9, 89.
Medvedev, S. S. (1957). In Proc. InI. Symp. Macromo/. Chem.. Prague p. 174. Pergamon, Oxford.
Medvedev, S. S., and Sheinker, A. P. (1954). Dok/. Akad. Nauk. SSSR 97,111.
Medvedev~ S. S., Zuikov, A. V., Gritskova, J. A., and Dudukin, V. V. (1971). Vysokomo~. Soedin.
A13, 1397 [English trans.: Po/ym. Sci. USSR, 13, 1572J.
Morris, C. E. M., and Parts, A. G. (1968). Makromo/ Chem. 119,212.
Okamura, S., Yamashita, T., and Motoyama, T. (1958). Kobunshi Kagaku 15,170; ef. Chem.
Abstr. 56, 953i (1960).

6.

245

Effets of the Choice of Emulsifier in Emulsion Polymerization

Onischenko, T. A., Rabinovich, M. A., and Gershenovich, A. 1. (1970). Kol/od Zh. 32, 886
[English transl.: Colloid J. USSR 32,744.
Ono, H., Jidai. E.. and Fujii, A. (1974). J. Col/oid Interface Sci. 49, 155.
Ono. H.. Jidai, E.. and Fujii. A. (1975). J. Phys. Chem. 79. 2020.
Ottewill, R. H., and Shaw, J. N. (1966). Discuss. Faraday Soco 42. 154.
Paxton. T. R. (1969). J. Col/oid Interface Sci. 31, 19.
Paxton. T. R. (1977) B. F. Goodrich Co.. Brecksville. private communication.
Piirma, l., and Chang, M. (1980). Org. Coatngs Plast. Preprints 43, 104.
Piirma, l., and Chen, S.-R. (1980). J. Col/od Interface Sci. 74, 90.
Piirma, l., and Wang, P.-e. (1976). In "Emulsion Polyrnerization" (1. Piirma and J. L. Gardon,
eds.), Am. Chem. Soe. Symp. Series 24, p. 34. American Chemical Society, Washington D.C.
Rosen, M. J. (1978). "Surfactants and Interfacial Phenomenena," Wiley, New Y.ork.
Ryabova, M. S., Sautin, S. N., Beresnev, V. N., and Smirnov, N. 1. (l975a). Zh. Prikl. Khim. 48,
1101 [Enf{li.fh transl.: J. Appl. Chem. USSR 48, 1148].
Ryabova. M. S.. Sautin, S. N.. and Smirnov. N. 1. (l975b). Zh. Prikl. Khim. 48, 1577 [English
transl.: J. Appl. Chem. USSR 48, 1632).
Ryabova. M. S.. Sautin. S. N.. and Smirnov. N. 1. (1977). Zh. Prikl. Khim. SO. 1719 [English
transl.: J. Appl. Chem. USSR SO, 1648].
Sata, N., and Saito, S. (1952). Kol/oid Z. 128, 151.
Sawyer, W. M., and Rehfeld, S. J. (1963). J. Phys. Chem. 67, 1973.
Testa. F., and Vianello, G. (1969). J. Polym. Sci. Part C 27,69.
Ugelstad, J., and Hansen, F. K. (1979). J. Polym. Sei. Polym. Chem. Ed. 17, 3069.
Ugelstad, J., EI-Aasser, M. S., and VanderhotT, J. W. (1973). J. Polym. Sci. Polym. Lett. Ed. U,
503.

Urban, P. e. (1980).Polysar Ltd., Canada.Paperpresentedat 54th Col/oid Surfaee Sei. Symp.,


54th, Lehigh Univ., Bethleham, Pennsylvania, June15-18 Abstr. 84.
Van der HotT, B. M. E. (1967). In "Solvent Properties ofSurfactant Solutions" (K. Shinoda, ed.),
Chapter 7. Dekker, New York.
VanderhotT,J. W., Van den Hui, H.J., Tausk, R.J. M.,andOverbeek,J.
Tb. G. (1970). In "Clean
Surfaces" (G. Goldfinger, ed.), Chapter 2. Dekker, New York.
Vinogradov, P. A., Odintsova. P. P., and Shitova, A. A. (1962). Vysokomol. Soedin. 4, 98 [English
transl.: Polym. Sci. USSR 4. 33].
Volkov, V. A., and Kulyuda, T. V. (1978). Vysokomol. Soedin. B 20,862; cf. Chem. Abstr. 90,

122120t
(1979).

Woods, M. E., Dodge, J. S., Krieger, 1. M., and Pierce, P. E. (1968). J. Paint Teehnol. 40, 543.
Waite, F. A. (1977). 1.e.1. Ltd., Paints Division, Slough, private cornmunication.
Yurzhenko, A. l.. and Brazhnikova, O. P. (1956). Zh. Obs. Khim. 26,1311 [English transl.: J. Gen.
Chem. USSR 26, 1481].

7
Polymerization
of Polar Monomers
v. l. Yeliseyeva

1. Introduction.
11. Interface Characteristics of Polymeric Dispersions
111. Relationship between Emulsifier Adsorption and
the Difference in the Boundary Phase Polarity
A. Monomer-Aqueous Phase Interface
B. Aqueous Phase-Polymer and Aqueous PhaseLatex Particles Interfaces.
IV. Mechanism of Particle Generation .
V. Colloidal Behavior of Polymerization Systems.
VI. Kinetics of Emulsifier Adsorption
VII. Mechanism of Formation and Structure of Particles
VIII. Polymerization Kinetics
IX. Relationship between Polymerization Kinetics
and Adsorption Characteristic of Interface.
X. Nor,nenclature .
References .

l.

247
249
250
251
254
257
261
268
270
278
283
286
287

Introduction*

Experience has shown that each monomer has its own specific
features in emulsion polymerization and requires a specific technology. To a
certain extent this is also true of other components in the reaction system,
e.g., the emulsifier and the initiator. Attempts to fit various emulsion
reaction systems into one physicochemical model or theoretical scheme
have therefore failed.
Nevertheless, several features that are common to all polymerizations in
colloidal systems have been established. They were formulated in the
* There is a Nomenclature
chapter.

section on p. 286 which defines the major symbols used in this

247
EMULSION POLYMERIZATION
Copyright <t> 1982 by Academic Press. Ioc.
Al! rigbts or reproductioo io aoy rorm reserved.
ISBN 0-12-556420-1

248

V. 1. Yeliseyeva

lassical theory of Smith and Ewart (1948) which was created in the late
1940s and which is weIl known to researchers working in the field of
emulsion polymerization. The most important concIusion that foIlows from
the theory is that the process takes place in an equilibrium of swoIlen
polymer-monomer particIes (PMP) each of which may be "living" or
"dead" at any given time depending on the presence of a free radical.
Furthermore, Case 2 of the theory was developed for systems in which
styrene was used as the weakly polar model monomer. An anion-active,
miceIle-forming surfactant with a low CMC (critical miceIle concentration)
value (such as sodium dodecylsulfate) was used as the model emulsifier and
persulfate as the model water-soluble initiator. The process is characterized
by a period of constant rate, the rate being proportional to the number of
particIes and the average number of radical s per particIe and depending on
monomer reactivity. According to the theory based on the Harkins' (1947)
model oi emulsion polymerization [a similar model was independentIy
developed by Yurzhenko and Mints (1945)], PM:P are formed from
monomer-swoIlen emulsifier miceIles into which free radicals enter from the
aqueous phase; the number bf particIes remains constant during the
constant rate periodo
The derived equation of the rate dependence on the emulsifier (Ce) and
initiator (C)concentrations during the constant rate period

R '" C~.6 . C?-4

(1)

satisfactorily describes polymerization in this and similar systems, on the


strength of which the theory became quite popular.
Medvedev (1968) suggested that not the number of particIes and miceIles,
but the value of their overaIl surface area is the rate-determining factor.
Inconstancy in the number of particIes during the constant rate period is
allowed, e.g., flocculation with retention of the overall surface value.
On the basis of this suggestion a general equatc:m was derived which,
however, differs only slightIy from the equation of the cIassical theory
shown in Eq. (1)
R '" C~.5 . Cp.5
(2)
With the significant increase in the number of various monomers,
emulsifiers, and initiation systems used in recent years the lmited applicability of the quantitative aspect of both theories quickly became apparent.
Let us consider one of the reasons for this limitation.
It do es not foIlow from Eqs. (1) and (2) that the polymerization rate
being a function of emulsifier concentration also depends on the emulsifier
activity in the given polymerization system. It has been shown in several
studies, however, that emulsifier activity at the water-organic phase interface is determined by the structure of the emulsifier molecule and the

7. Polymerizationof Polar Monomers

249

polarity of the organic phase. Hence, the use of the emulsifier concentration
value in the rate equation without taking into account its activity .is
inadequate. Therefore, in order to extend the applicability of Eqs. (1) and (2)
to the polymerization of monomers of various polarity in the presence of
emulsifiers of various structure, a parameter characterizing the emulsifier
activity in the given system should be introduced. Whereas this parameter
may be neglected when considering polymerization of some hydrophobic
monomers (hydrocarbons) with close adsorption characteristics, it must be
taken into account in the case of polymerization of highly polar monomers
(or their copolymerization with hydrophobic monomers) where the colloidal behavior of the system, and consequently the polymerization kinetics,
depends on emulsifier activity. Since the practical importance of latexes
obtained by polymerization of polar monomers or monomeric systems is
rapidly increasing, a consideration of the regularities of their formation,
taking into account a11important factors, is of considerable interest.

D.

InterfaceCharacteristics of Polymeric Dispersions

Polymers in the form of latex (co11oidal dispersions) acquire a novel


quality not observed in bulk or solution and which is due to the strongly
developed interface with the aqueous phase. The properties of this interface
vary specifically with the nature of the polymer, the latter varying in a wide
range for different latexes and other polymer dispersions.
The most important characteristic of the interface is its energy, the
interfacial free energy, which depends on the molecular interactions between
boundary phases. Molecular interactions which are responsible for interfacial tension occur as a result of various we11-known intermolecular
forces. Many of these forces are determined by the specific chemica] nature
of the substance. On the other hand in a11 types of matter London
dispersion forces exist, and they always ensure attraction between neighboring atoms and molecules independent of their chemical nature. Fowkes
(1964) has suggested that the effect of dispersion interactions on interfacial
energy may be predicted from the geometric mean ratio of dispersion forces
of surface tension components of boundary phases. My~)1/2. Since the
interfacial tension is the sum of tensions in both regions and provided that
only dispersion forces exist between the phases, then the interfacial tension
at the boundary of the first and second phase is
Y12 = Yl + Y2 - 2M - y~)1/2

(3)

Fowkes analyzed the available experimental data on interfacial tension Y12


at the water-organic liquid interface and calculated the values of interfacial

250

V. 1. Yeliseyeva

tensions Y12 under the assumption that only dispersion interactions are
operative at the interface. For all the compounds studied Y12was found to
be higher than Y12' The difference between them may serve as a measure of
the energy, which includes various polar interactions. Fowkes established
that for most of the cases examined, interaction attributable to dispersion
forces makes the highest contribution to interfacial tension, although for
substances capable of forming hydrogen bonds the attraction due to
dispersion forces may be approximately equal to attraction due to hydrogen
bonds.
Wu (1974) obtained an equation for boundary phases with various
properties which relates Y12not only to dispersion interaction but also to
phase polarity. In his opinion, for an interface between high and low energy
phases, e.g., mercury or water and organic polymers, this dependence may
be represented by the geometric-harmonic mean ratio
Y12 = Yl + Y2 -

2[y1- Y~- 4y~y~/(y~+,Ym

(4)

This ratio is transformed to Fowkes' equation when the polar term is


negligibly small. Using Eq. (4) and experimental data on interfaCal and
surface tension of polymers, Wu calculated the polarity of some polymers.
He estimates the contribution of the polarity to the surface tension of
polymers to be considerable. Thus, for poly(vinyl acetate) it takes up 33%,
for poly(methyl methacrylate) 28%, and for polychloroprene 11%.
Since the polar contribution is the most important factor determining the
interfacial energy, the polarity of polymers in a dispersed phase strongly
influences (i) the adsorption of surfactants, (ii) the mechanism of particle
formation, (iii) the particle flocculation, and (iv) the configurational behavior of the forming macromolecules. Several experimental confirmations
of this have been reported elsewhere (Yeliseyeva, 1972).

III. Relationship between EmulsifierAdsorption and the Difference in the


Boundary-PhasePolarity
The energy state of the interface obviously determines the adsorption of
a given surfactant by this interface. On the other hand, a given interface
may adsorb with different energy surface-active substances of different
chemical nature. Rhebinder (1927) was the first to point out that the
difference in polarity between boundary phases, which affects the interfacial
energies, is the main factor determing adsorption; for adsorption of a third
component by the interface, the polarity of this component should lie
between the polarities of the two boundary phases.

251

7. Polymerizationof Polar Monomers

In order to analyse the colloidal processes occurring in colloidal polymerization systems containing monomer, polymer monomer and polymer
particles it is important to know about adsorption characteristics of the
emulsifier on the monomer-aqueous phase and polymer-aqueous phase
interfaces.
A.

Monomer-Aqueous Phase Interface

The structure of the liquid-liquid interfacial layer depends on the


difference in polarity between the two liquids (Kaeble, 1971). Asymmetric
molecules of some liquids display a molecular orientation on the interface
which is indicative of their structure. Thus, interfacial tension at the octanewater interface is 50.5 nm/m whereas at the octanol-water interface it is
only 8.8 nm/m. Reduction of interfacial tension in the latter case points to
the orientation of octanol hydroxyl groups toward water, in other words to
the structure and polarity of the interfacial layer. Because of such an
orientation, the stimulus for adsorption of other asymmetric molecules on
the interface is decreased. A similar pattern is typical of the homologous
series of lower alkyl acrylates: at the interface with water the carbonyl
groups of their asymmetricaI" molecules are oriented toward water; this
orientation is more effective the higher the polarization of the carbonyl, Le.,
the smaller the alkyl. Interfacial tension decreases in the same order from
27.2 nm/m for hexyl acrylate (Yeliseyeva et al., 1978) to 8 nm/m for methyl
acrylate (datum from our laboratory by A. Vasilenko).
Adsorption isotherms of n-butyric acid at various interfaces are shown in
Fig. 1 where it can be seen that the greater the difference in polarity
between boundary phases the higher the adsorption (differing by more than
fourfold).
Interfacial tension, however, cannot serve as the only criterion of adsorption. Typical surfactants with asymmetric molecules or ions consisting of a
polar group and a sufficiently long hydrocarbon chain are always active at
water-hydrocarbon interfaces. The greater the difference in polarity between boundary phases the steeper the orientation of surfactant molecule at
the interface and the larger the reduction in free energy of the system due to
adsorption.
Upon substitution of a non polar organic phase (aliphatic hydrocarbon)
by a polar organic liquid, with resultant decrease in the energy of interaction between the aliphatic hydrocarbon portion of surfactant and the
organic phase, the free energy of the system is reduced to a lesser extent and
the surfactant molecules are less oriented at the interface.
Davis et al. (1972) estimated the contribution of CH2 groups to the free
energy of substance transfer from aqueous phase into organic liquids of

252

V. 1. Yeliseyeva

0.1
0.2
0.3
Concmtration of n-lJutyrcacid
(mol/tim!)

Fig.1. Adsorption isotherms of'n-butyric acid at the interfaces: (1) water-air; (2) waterhexane; (3) water-benzene; (4) water-olive oil. T, 318 K.
various polarity (~G). It was established that this value decreases with
increasing polarity of the organic liquido For nonpolar liquids such as
cycJohexane, hexane, and heptane, this value is 3.6 kJ/mol, whereas for
polar liquids (alcohol s, nitro-compounds, ethers) it decreases to 1.892.98 kJ/mol.
.
Using the stalagmometric method, we determined the adsorption energy
(E'ds) of sodium dodecylsulfate (SDDS) at the water boundary with the
homologous series of alkyl acrylates at 293 K (Yeliseyevaet al., 1978).
Calculations were performed from the initial section of the surface tension
isotherm using the Langmuir equation. Table 1 lists values of the initial
interfacial tension Y12;the overall free energy of adsorption, E'ds>and the
TABLEI
Energy Characteristics of SDDS Adsorption on WaterAlkyl Acrylate Interface Depending on Alkyl Acrylate
Polarity
YI2

E.d'

Acrylate

(mJ/m2)

(kJ/mol)

t:.G
(kJ/mol)

Methyl
Ethyl
Butyl

8
13.7
23.2

20.0
22.1
25.6

1.69
1.85
2.10

253

7. Polymerizationof Polar Monomers

free energy of adsorption LlG per methyl group of the alkyl (assuming that
the entire alkyl portion of the emulsifier is immersed in the monomer
phase). Data presented in Table I indicate a decrease in the energy of
interaction between the hydrocarbon portion of the emulsifier and the
organic phase (alkyl acrylate) with an increase in the polarity of the latter.
Accordingly, the overall adsorption energy also decreases. Several important practical conc1usions may be drawn from these data. On the one hand,
dispersions of polar monomers are stabilized to a lesser extent than
nonpolar monomers by emulsifiers containing an aliphatic hydrocarbon in
the oleophilic part of the molecule. Although the overall adsorption energy
increases only slightly with increasing length of the hydrocarbon portion,
as follows from data presented in Table 11,the solubility of the emulsifier in
water is significantly reduced, which may hinder its practical use. Therefore,
on the other hand, in order to increase the energy of interaction of the
oleophilic part of the emulsifier molecule with the polar organic phase,
polar groups should be introduced into its structure. This has been confirmed in practice.
A consideration of the adsorption kinetics is very important in an
estimation of the effectiveness of surfactants under the dynamic conditions
of emulsion polymerization. In a stalagmometric study of dynamic and
static adsorption of emulsifiers of various structure at the air-water interface, it was established that adsorption values of micelle-forming surfactants differ significantly in the period of drop formation (Nikitina et al.,
1961). This was explained by the consider~ble period needed for establishment of adsorption equilibrium connected with the kinetics of adsorption
layer formation. The authors conc1uded that for usual concentrations of
surfactant solutions the period of establishment of adsorption equilibrium
can be taken as equal to 2 mino Figure 2 shows the adsorption isotMrms of
TABLE 11

Energy Characteristics of Adsorption


Alkylsulfates on Ethyl Acrylate-Water
Interface8

Alkyl su,lfate
C12H2SS04Na
C14H29S04Na
CI6H31S04Na
a The

E.d.
(kJjmol)
22.1
23.3
24.6

of

I:1G

(kJjmol)
1.85
1.69
1.55

absence of proportionality between I:1Gand

increase of the hydrocarbon


radicallength
indicates that for
polar organic liquids Traube's rule is not valid.

254

V. 1. Yeliseyeva

-10

rmxa =3.ifx!O

moZ/cm2

2;5 5.0
Concfntration

Z5

10.0

of lIerosol (tOsmoZ/ dm J)

Fig. 2.
Adsorption isotherms of Aerosol AT plotted according to data of (1) static
(2 min) and (2) dynamic (2 sec) values of surface tension.

Aerosol AT solutions (sodium dioctylsulfosuccinate) .plotted according to


both dynamic (drop formation period 2 sec) and static (drop formation
period 2 min) values. The diff~rence in the shape of the isotherms clearIy
indicates the significance of the time of surface formation for effective
emulsifier adsorption and stabilization at the interface. Apart from the
emulsifier structure, the process may depend on the difference in polarity
between the contiguous phases.
Stalagmometric determination of the SDDS adsorption at the aqueous
solution-ethyl acrylate interface dependence on the rate of drop formation
(volume 0.03 cm3) were carried out in our laboratory by Vasilenko. The
measurements showed that establishment of adsorption equilibrium at the
CMC occurs at drop formation periods of 15-20 sec, i.e., at surface
formation rates not exceeding 10-6 m2/sec. Adsorption kinetics acquires
considerable importance in analysis of the mechanism of particle formation
during emulsion polymerization, when the rate of organic phase formation
and the rate of adsorption layer formation may be commensurate.
B.

Aqueous Phase-Polymer and Aqueous Phase-Latex Particle lnterfaces

The general ideas developed for adsorption of surfactants on liquidliquid and liquid-air interfaces obviously cannot be completely transferred
to adsorption on interfaces with a solid phase.
Wolfram (1966) caIculated the packing density of surfactant molecules on
the polymer surface from the adsorption value at the CMC and showed
that it varies with polymer polarity. Thus, the adsorption area of a SDDS
molecule on various surfaces takes the following values (nm2): paraffin 0.41,

7. Polymerization of Polar Monomers

255

polyethylene 1.21, poly(methyl methacrylate) 1.32, poly(ethylene terephthalate) 1.42, and polycarbonate 1.48. On the basis of the results obtained
the author introduced the concept of "required area" of a surfactant
molecule and conc1uded that this area increases with increasing polarity of
the polymer. The established relationship he explains by the existence of
other interactions such as orientation, induction, and hydrogen bonds
between contiguous phases, in addition to dispersion forces. Because of
these interactions the molecules adsorbed on the solid phase-liquid interface may acquire a nonvertical orientation: they may be arranged with
greater planarity the more intensively the field (depending on polymer
polarity) acts upon the polar part of the surfactant molecule.
Several investigations have determined the absorption behavior of surfactant adsorption on partic1es of aqueous polymer dispersions by adsorption titration. The results have been similar to those observed by Wolfram
for adsorption on aplanar polymer surface determined from the wetting
angle. Thus, Paxton (1969) established that the area occupied by a sodium
dodecylbenzylsulfonate molecule in a saturated adsorption layer (AS1im)on
the surface of PMMA latex partic1es is 1.31 nm2, whereas on the surface of
polystyrene latex partic1es it is only 0.53 nm2. The author considers that
previous studies of adsorption of this emulsifier, which gave adsorption area
(ASlim)of 0.50 nm2, were carried out on interfaces with similar adsorption
characteristics. The intermolecular forces involved in adsorption between
water and air or water and hydrocarbon solvent or nonpolar polymer are
nearly the same. In his opinion the differences observed in surfactant
adsorption on polymers of different nature allows one to use the molecular
adsorption area value as a polarity characteristic of the polymer phase.
Dunn and Chong (1970) determined the value of ASlim for sodium
dodecylsulfate on poly(vinyl acetate) latex partic1es; it was fountl to be
1.24 nm2, i.e., 2.4 times higher than the minimal adsorption area of this
emulsifier. Zuikov and Vasilenko (1975) showed that the ASlim value for
sodium dodecylsulfate increases with polarity of the polymer and is inversely related to the interfacial tension at the interface of the corresponding
monomer with water. A correlation was established (Yeliseyeva and
Zuikov, 1976) between ASlimof a given emulsifier and its adsorption energy
at the corresponding monomer-water interface. The same correlation is
observed between ASlimof emulsifier at the given polymer-water interface
and its adsorption energy at the interface between the monomer and water
for the homologous series of alkyl sulfates (Table III).
It follows from the results obtained that the factors determining the
adsorption value of a given surfactant at the water interface with the
monomer and with the corresponding polymer are the same. Since with the
increase in the adsorption energy on the emulsifier-monomer interface, the

256

V. 1. Yeliseyeva
TABLE

111

Adsorption Characteristics of Surfactants on Water-Monomer


and Aqueous Phase-Polymer Particle Interfaces

Monomer

Emulsifier

MA
EA
EA
EA
BA
HA
Sty

C12H2SS04Na
C12H2SS04Na
C14H29S04Na
C16H31S04Na
C12H2SS04Na
C12H2SS04Na
C12H2SS04Na

aAt polymer-water

1'12
monomerwater
(mJjm2)

-8
13.7
13.7
13.7
23.2
27.2
33.8

E.d'
(kJjmol)

ASHm
(nm2)

20.0

1.51
0.92
0.82
0.74
0.67
0.52
0.48

22.1
23.3
24.6
25.6
-

interface.

value of the surfactant adsorption increases at the interface with the


polymer (Le., ASlimdecreases) which means that the packing density of its
molecule in the adsorption layer increases.
In a study of the adsorption characteristics of alkyl acrylate and alkyl
methacrylate latexes Stterlin et al. (1976) obtained values for adsorption
areas of sodium dodecylsulfate which support the values listed in Table III.
The authors, however, attributed the established regularity of ASlimvariation not to polarity but to hydrophilicity of the polymer, the increase of
which leads to reduction of the equilibrium emulsifier concentration on the
latex particle surface. Since in the homologous series of alkyl acrylates and
alkyl methacrylates the increase of hydrophilicity of the monomer occurs
parallel to the polarity of the interface, the established relationship may also
be attributed to the latter factor. In an attempt to apply Wu's (1974) ratios
to calculation of the polar component of interfacial tension at the interface
with aqueous phase for polymers of difIerent polarity, Vijayendran (1979)
estimated ASlimin latexes of the same polymers on the basis of the obtained
results. The values are in good agreement with reported experimental data
(Yeliseyeva et al., 1978; Stterlin et al., 1976). The coincidence of adsorption
areas for molecules of a given emulsifier on latex particles of polymers of
difIerent polarity obtained by difIerent authors points to the determining
role of this factor in adsorl'tion in the case of usual latexes prepared by
emulsion polymerization. However, in general when determining the adsorption areas of surfactant molecules other factors, such as electrolyte concentration, temperature, and particle size, should also be considered .(Piirma
'ind Chen, 1980).

7.

IV.

257

Polymerization of Polar Monomers

Mechanism of Particle Generation

Numerous experimental data are available which suggest that both the
solubility of the monomer in water and its polarity affect the mechanism of
particle generation. Whereas during a conventional emulsion polymerization of styrene (or other hydrophobic monomer) examined by Harkins
(1947) and Yurzhenko and Mints (1945) the most important sites of particle
generation are monomer-swollen emulsifier micelles; with the increase of
monomer solubility in water particle generation in similar systems may be
shifted to the aqueous phase. Thus, the possibility of preparing concentrated and stable latexes by polymerization of methyl acrylate and ethyl
acrylate in the absence of an emulsifier was demonstrated by Yeliseyeva and
Zaides (1965) and by Yeliseye~a and Petrova (1970). It is also well known
that styrene latexes may be obtained by emulsifier-free polymerization; such
latexes are characterized, however, by larger particles and much lower
concentrations.
In all such cases the process initiated by persulfate begins in the aqueous
phase with the formation of water-soluble, surface-active polymeric radicals
which, after growing to a certain critical size, precipitate to form particles;
subsequent polymerization proceeds mainly within these particles. The
higher the solubility of the monomer in water, the more surface-aciive
radical s and therefore primary particles are formed and the higher the
stability and the concentration of the latex. A kinetic curve of the
emulsifier-free polymerization of ethyl acrylate confirming this scheme is
shown by Curve 1 in Fig. 3 (Yeliseyeva anq Petrova, 1970). The process

15
t(mini

25

Fig. 3.
Conversion kinetics (g/100 g of aqueous phase) of ethyl acrylate (1) in the
absence of emulsifier and (2) in the presence of 0.5 mol % of emulsifier E-30. Ammonium
persulfate, 0.33%; tert-dodecyl mercaptan, 0.33% in aqueous phase; T, 60C; phase ratio, 1: 3.

258

V. 1. Yeliseyeva

beginning in the aqueous solution is characterized by a low initial rate


which rapidly increases, evidently due to the formation of partic1es and the
transfer of the polymerization mainly into these partic1es. At the initial
stages the surface tension of the system decreases and then increases with
conversion. Estimation of the total activation energy of emulsifier-free
polymerization of ethyl acrylate showed it to be 100 kJjmol initially, but it
markedly decreases in the course of the process in accordance with the
increase in the termination energy occurring in the partic1es typical of
emulsion polymerization. In the presence of emulsifier (Fig. 3, Curve 2), the
process proceeds from the beginning at a high rate and a low activation
energy (53.5 kJjmoI), suggesting formation of partic1es from the very beginning of the process and participation of the emulsifier in the mechanism of
partic1e generation.
The slight difference in the vaIues of the integral molecular mas s of the
poIy(et~1 acrylate) oDtained in the absence and in the presence of an
emulsifier supports the assumption that in both cases the process mainly
follows the mechanism of emulsion polymerization; the viscosity average
molecular masses of this polymer obtained under identicaI conditions, both
without emulsifier and in the presence of 2% (based on monomer) SDDS,
are approximately 4 x 106 and 7 x 106, respectively (Mamadaliev, 1978).
The slightly lower integral molecular mas s in the former case may be
explained by a larger contribution of polymerization in aqueous solution at
the initial stages of the process. Hence, the difference in the mechanisms of
partic1e formation of ethyl acrylate (solubility in water 2.5%) in the absence
and presence of emulsifier at concentrations above CMC consists in the first
case of poIymerization in the aqueous solution with the formation of" own"
surfactants which precede the generation of partic1es, whereas in the second
case partic1es are formed from the very beginning. Whether own surfactants
are formed in the latter case and the role they play is yet to be determined.
After generation of partic1es, when their concentration becomes quite high,
polymerization in the aqueous phase probably results in the formation of
low molecular weight oligomer radical s which are captured by the partic1es
before they grow to critical size. Their role in the subsequent process
consists of initiation and termination of polymerization in the formed
partic1es according to the usual behavior of water-soluble, active radicals.
In the case of their termination in the aqueous phase they may also act as
own surfactants contributing to the stabilization of partic1es.
Zuikov and Soloviev (1979) studied the effect of monomer polarity on
both the size and the size distribution of partic1es in the forming latexes. As
a criterio n of polarity they took the solubility of the monomer in water,
which in this case has an independent significance as well. Polymerizations
of styrene, methyl methacrylate, butyl methacrylate, and methyl acrylate

Polymerization of Polar Monomers

259

initiated by ammonium persulfate(0.05%in aqueous phase) in the presence


of SDDS at concentrations below the CMC were investigated. As can be
seen from the electro n microphotographs of the latexes (Fig. 4), with
increasing monomer solubility in water the size of particles decreases and
their distribution broadens. This is also illustrated by plots shown in Fig. 5.
The results obtained may be associated with the effects of both monomer
solubility in water and polarity of the forming interface on the mechanism
of particle formation. Since in all experiments the emulsifier concentration
was below the CMC, particle generation in the case of the water-soluble
initiator began in the aqueous phase. Because of the lower solubility of
styrene, the concentration of polymer radicals initially formed is lower than
for MMA and MA. Accordingly, a smaller number of large particles is
formed. The critical size of polymer radicals increases with the solubility of
the monomer in water (according to literature data it is 8 monomer units
for styrene and 65 for MMA). The period of particle formation also
increases, leading to higher polydispersity. Moreover, as previously discussed, the emulsifier adsorption decreases with increasing polarity of the
interface; and therefore for the formation of a given surface less emulsifier is
consumed, which also increases the period of particle generation and

...

'.
2

1.1"~

Fig. 4. Electron micrographs of latex particles obtained by polymerization of (1)


styrene; (2) methyl methacrylate; and (3) methyl acrylate. SDDS, 0.066%; ammonium
persulfate, 0.05% in aqueous phase; T, 80C; phase ratio, 1: 10.

260

V. 1. Yeliseyeva

250
200

~t::
150f4
4'

1J

0.7

100

1
2
J
'+
5
MonomerSolu.bilitgin water (%)
Monodispersity
coefficient
(K) and diameter(D)of latexparticlesformedby

Fig. 5.

polymerization
of monomers differing in solubility in water: (1) styrene; (2) butyl methacrylate;
(3) methyl methacrylate; (4) methyl acrylate. For conditions see legend to Fig. 4.

decreases the coefficient of monodispersity. This facilitates the increase of


the overall surface area, because of a decrease in particle size. As can be
seen from Fig. 5, the broadest particle size distribution is observed for the
most polar monomer, MA. Due to the increased critical size of radical s and
the smallest consumption of emulsifier by the forming interface, this corresponds to the longest period of particle formation.

""

0.8

"-

0.0'+

0.05

I
0.05

/3

0.7
I

0.01

0.02

0.03

[s1)1)8] <%)
Fig. 6.
The effect of SDDS concentration (% in aqueous phase) on the monodispersity
coefficient of latex particles obtained by polymerization of (1) styrene; (2) methyl
methyacrylate; and (3) methyl acrylate. Ammonium persulfate, 0.05% in aqueous phase; T,
80C; phase ratio, 1: 10.

7. Polymerizationof Polar Monomers

261

A comparative study of the dependence of the monodispersity coefficient


K on the SDDS concentration for polymerization of styrene, MMA, and
MA was carried out in the same work. The results obtained, similar to data
of Yakovlev el al. (1971), indicate that for styrene polymerization the
polydispersity of the latex sharply increases in the region where the initial
emulsifier concentration equals the CMC (Fig. 6). This is explained by a
change of particle-formation mechanism, a transition from homogeneous to
micellar, and accordingly, by an increase of the particle-generation periodo
In the case of MMA (solubility in water 1.5%) the transition of SDDS
concentration from below to above CMC affects the polydispersity coefficient to a lesser extent, since in the polymerization of this monomer
particle formation apparently follows both the homogeneous mechanism
and the micellar mechanism, In the case of polymerization of the most
soluble monomer (MA, solubility 5.2%), an increase in the emulsifier
concentration smoothly decreases the monodispersity coefficient, which is
not affected by passage over the CMC. The contribution of micelles to the
formation of particles is evidently negligible in this case.

V.

Colloidal Behavior of Polymerization Systems

The differences in the polymerization kinetics and colloidal behavior of


polymerization systems based on monomers of different polarity may be
illustrated (Bakaeva el al., 1966; Yeliseyeva and Bakaeva, 1968) by the
polymeri;ation of the model monomers, methyl acrylate and butyl methacrylate, at various concentrations of sodium alkylsulfonate (C15H31S03Na).
The fact that the solubility of the monomers in water differs by two orders
of magnitude (5.2 and 0.08%, respectively) was used as a criterio n of
polarity. An additional advantage to comparing these two monomers is
that their polymers have rather close glass transition temperatures which is
important for coalescence of particles at later stages of polymerization.
As can be seen from Fig. 7, variation of the initial emulsifier concentration from 0.5 to 5% in the aqueous phase results in an increase in the
BMA polymerization rate (conversion after 30 min) by about 15 times,
whereas the MA polymerization rate essentially does. not change with
changing emulsifier concentration. The data in Table IV show how emulsifier concentration affects particle size and, accordingly, number.
AJthough for both monomers an increase in the emulsifier concentration
results in a decrease in the size and an increase in the number of particles,
the effect is much weaker in the case of MA"polymerization. A comparison
of run 2 with runs 3, 4, and 5 in Table IV indicates that at the same initial
emulsifier concentration the number of particles in 1 dm3 of the latex is an

262

V. 1. Yeliseyeva
50

40

--

s:::

"Jt
....
'->
-b.

s:::
1::
20

/.

4.-

""
Q.. 10

2
,

Fig. 7.

[RSOaNa](%)

Dependence of conversion (%) over 30 min on sodium alkylsulfonate

concentration
in polymerization
of (1) methyl acrylate and (2) butyl methacrylate.
persulfate, 0.01% (1) and 0.1% (2) in aqueous phase; T, 60C; phase ratio, 1: 1.5.

Ammonium

order of magnitude higher for BMA polymerization. Since in both cases the
polymerization begins with the formation of very small particles, it may be
assumed that the flocculation of particles occurs during MA polymerization, even with excess emulsifier (run 4), and does not occur (or occurs on
a much smaller scale) during BMA polymerization. Morover, as can be seen
from Figs. 4 and 5, during the polymerization of 'polar monomers (MMA
TABLE IV
Colloidal Characteristics of MA and BMA Latexes Depending on Initial
Concentration of Sodium Alkylsulfonate in Aqueous Phase

No.

Emulsifier
concentration
in aqueous
phase (%)

Monomer

1
2
3
4
5

0.5
1.0
1.0
3.0
3.0

MA
MA
BMA
MA
BMA

y
(mJ/m2)
41.6
38.1
"
31.0 (-CMC)
-"

Partic1e
diameter
(nm)

Number of
partic1es
in 1 dm3

217
. 178
86
148
60

7.5 x 1016
1.4 xl017
1.25 x 1018
2.45 x 1017

3.7 X 1018

"We are unable to determine y precisely because of rapid film formation on the surface of
the p-BMA latex.

7. Polymerizationof Polar Monomers

263

and MA) a larger number ofparticles of lesser size is generated (and


retained in dilute latexes) than for the less polar monomer (styrene). This
also indicates flocculation of particles during MA polymerization (shown in
Table IV). This explains why Eq. (1) does not describe polymerization of
both investigated monomers equally well.
The relationship between colloidal behavior of polymerization systems
and monomer polarity has also been observed in other works. Thus, in an
adsorption titration study (Zuikov and Vasilenko, 1975) of alkyl acrylate
and styrene latexes synthesized with SDDS it was shown that the amount of
emulsifier additionally adsorbed by 1 g of polymer is sharply reduced by an
increase of interface polarity and is 40 times less for the methyl acrylate
latex than the styrene latex (Table V), the methyl acrylate latex particles
being much larger. The area occupied by an emulsifier molecule (ASllm)on a
particle surface at CMC increases with interfacial polarity in agreement
with the experimental data described above. Two characteristics of the
adsorption area of emulsifier molecules are given in Table V: As, the
calculated area per molecule on latex particles and ASlim'the area occupied
by molecule of the given emulsifier at saturation determined by adsorption
titration. The p value [(AS1m/As)100]may be used as a characteristic of the
degree of adsorption saturation but not of the degree of filling of the
adsorption layer and therefore not of latex stability, as usually assumed
when dealing with butadiene-styrene latexes. Because of the hydrophobic
nature of the particle surface, ASlimfor most rubber latexes is close to the
minimal .value corresponding to the most dense packing of emulsifier
molecules on surface, and therefore p may cparacterize the degree of filling
of the adsorption layer and thus of latex stability. But in the general case,
when ASlimis a variable dependent on polymer polarity and which may
acquire high values, even at p = 100%, the degree of filling of the adsorption layer and latex stability may be quite low. Thus, from data presented
in Table V it follows that on the one hand with an increase in interface
polarity the limiting adsorption of emulsifier decreases, whereas on the
other hand the degree of adsorption saturation of latex particles increases:
in the ~ethyl acrylate latex the p value is close to 100%. Allowing
flocculation of primary particles it may be considered that in the case of
MA it proceeds almost to the limiting degree of saturation which however is
not sufficient for the required stability of particles during polymerization
(formation of coagulum was observed). In the case of styrene latex the
degree of saturation at the given emulsifier concentration is '" 50%, but As
is lower (the degree of filling of the adsorption layer is practically twice as
high as in the methyl acrylate latex) which ensures a higher aggregate
stability of the particles during polymerization (absence of coagulum). It is
noteworthy that at a higher initial concentration of emulsifier (Table V) its
consumption during styrene polymerization is higher (surface tension of the
resultant latex is higher) than in alkyl acrylate polymerization.

TABlE V
Colloidal Characteristics of Acrylate and Styrene Latexes Synthesized with SDDS

Monomer

Initial
SDDS conc
in aqueous
phase
(mol/dm3 x 103)

Latex
conc
(%)

Coagulum:
monomer
(%)

MA
EA
BA
St

8.68
17.36
26.64
26.04

20
18.7
20
19.1

0.20
0.11
0.00
0.00

Particle
diameter
(nm)

)'
(mJjmZ)

As
(nmZ)

Additional
adsorbed
amount of
SDDS
(moljg of
polymer x 103)

49.5
47.5
53.6
63.9

1.92
1.24
1.06
0.95

0.347
2.01
5.78
13.23

125
96
88
71

TAlE

Degree of
adsorption
saturation
ASlim
(nmZ)

( )

1.75
0.86
0.61
0.49

90.6
78.2
63.8
51.7

A:m

VI

Colloidal Characteristics of Acrylate and Styrene Latexes Synthesized with AAEOS


Latex stability

Monomer

Solubility
in water
(%)

Latex
conc
(%)

Coagulum:
monomer
(%)

Viscosity
(cp)

)'
(mJjmZ)

MA
EA
BA
2-EHA
BMA
St

5.2
2.6
0.34
0.01
0.08
0.03

34.4
34.3
34.3
34.0
33.8
33.7

0.17
0.06
0.00
0.68
0.36
0.18

6.0
12.3
29.2
13.9
18.5

41.7
49.4
59.5
67.0
62.0
-

Mechanical

stability was determined

on a Marone-Ulevich

instrumento

Particle
diameter
(nm)
150
150
MO
97
120
100

Number of
particles
. in 1 dm3
(x 10-17)
1.66
1.67
2.19
6.17 .
3.58
6.10

Overall
surface
in 1 cm3
(m-Z)

As
(nmZ)

mechanical
(% formed
coagulum)a

eIectrolyte
(mi 0.1 M
CaClz)

11.7
11.8
13.5
18.5
16.1
19.1

1.05
1.12
1.31
1.87
1.51
1.80

2.1
1.6
O
0.2
1.0
0.5

0.75
0.95
1.25
0.15
0.05
0.02

7.

Polymerization of Polar Monomers

265

Yeliseyeva et al. (1975) investigated the effectiveness in emulsion polymerization of the alkyl acrylates, butyl methacrylate and styrene, in the
presence of another type of emulsifier, alkylac polyoxyethylenesulfate
(AAOES), obtained by exhaustive sulfonation of Triton X-100 (Rohm and
Haas, Philadelphia, Pennsylvania) with sulfamic acid and subsequent neutralization with ammonia. Polymerizations were carried out under identical
conditions by the batch method. Emulsifier concentration
in aIl runs was
26.4 x 10- 3 mol/dm3 in aqueous phase, ammonium persulfate concentration
1.1 x 10- 3 mol/dm3
for MA and EA polymerization
and
2.2 x 10-3 mol/dm3 for polymerization of the other monomers. AIl in dices
characterizing the latex stability have optimal values for the monomer with
medium solubility in water, namely for butyl acrylate (Table VI). However,
latexes based on more polar monomer are more stable when using this
emulsifir than latexes based on less polar, hydrophobic monomers. Surface
tension in the latter case is considerably higher, indicating a higher capacity
of the surface to adsorb the given emulsifier when formed during polymerization of less polar monomers. As in the case of alkylsulfonate (Table IV),
at the same initial concentration of the emulsifier a smaIler number of larger
particIes are formed in the case of polar monomers, where the surface is
better protected by the emulsifier. Thus, the area As occupied by the
emulsifier molecule on the surface of styrene latex particIes is 1.7-fold larger
than for the methyl acrylate latex, the diameter of the polystyrene latex
particIes being smaIler.
The l0l" stability to electrolytes of latexes based on hydrophobic
monomers obtained in the presence of AAOES play therefore be explained by
the rarefied adsorption layer on the particIe surface. Conversely, the
enhanced stability to' electrolytes of latexes of polar monomers may be
attributed to the denser packing of emulsifier molecules in the adsorption
layer at a given emulsifier concentration.
The somewhat higher mechanical
stability of latexes based on hydrophobic monomers may be associated with
their higher stability to orthokinetic fIocculation which is attributable to the
smaIler size of particIes, which to a certain extent compensates for the fact
that their surface is les s protected by the emulsifier.
Similar data on the adsorption saturation of acrylate latexes (based on
EA, BA, and 2-EHA) synthesized with the same emulsifier were obtained
recentIy by Snuparek and Tutalkova (1978). A lesser degree of filling of the
adsorption layer, 46.5 and 49.6%, was observed for more hydrophobic
monomers (2-EHA and BA): with a greater degree being observed in the
polymerization of the more hydrophilic EA, 66.3%.
Summarizing the described results, it may be concIuded that when using
the emulsifiers described above, the coIloidal behavior of polymerization
systems is mainly determined by monomer polarity. Latexes of polar

266

V. 1. Yeliseyeva

monomers are characterized by larger particles with surfaces close to


limiting saturation, which, due to high values of Aslim, correspond to low
degrees of filling of the adsorption layer. Since polymerization invariably
begins with the formation of very small particles, the observed difference in
colloidal behavior is associated with tlocculation to a certain size at the
initial stages of polymerization of polar polymer-monomer particles (PMP)
and with the formation of secondary particles. The better protected PMP of
hydrophobic monomers retain stability to higher conversions at which their
protection may become insufficient due to the consumption of most of the
emulsifier. Flocculation of primary PMP of polar monomers up to a certain
limit ("limited tlocculation") may be explained by the kinetics of the
adsorption layer formation under dynamic conditions of polymerization.
This leads to a reduction of the real filling of the adsorption layer of
primary particles in relation to equilibrium adsorption which is lower than
in the case of hydrophobic monomers. Equilibrium adsorption is attained
on particles of the resultant latex (or in the course. of the process when
formation of new particles is terminated) ensuring at sufficient emulsifier
activity the necessary aggregate stability of the system.
In the above-mentioned cases of polymerization of lower alkyl' acrylates,
because

of the high rate of the process

kp

1260 dm3/mol sec for MA

(Bagdasar'yan, 1966) and the high monomer concentration in particles


(Gerrens, 1964), the surface of the monomer-polymer phase increases at a
rate that may exceed the rate of attainment of equilibrium adsorption.
On the other hand, the 'PMP formed during polymerization of hydrophobic monomers are better protected by the emulsifier because of the
high adsorption capacity of the surface (low value of Aslim)' In addition,
their polymerization rate (kp = 190 dm3/mol sec for styrene) (Bagdasar'yan,
1966) and monomer concentration in particles (Gerrens, 1964) are lower,
which ensures a lower rate of interface growth and creates conditions for
achieving adsorption equilibrium during PMP formation, which in turn
hinders tlocculation up to high degrees of conversion. Flocculation in this
case occurs only when most of the emulsifier is consumed at high values of
As. It should be noted that particles of hydrophobic monomers may remain
descrete at lower degrees of filling of the adsorption layer (styrene, BA,
Table V) which is possibly connected with the reduced interparticle interaction (Hamaker's constant) and the lower monomer concentration in
them. For retention of stability, particles of polar monomers should be
better protected by the emulsifier, apparently because of stronger interparticle interaction and the higher content of monomer in them.
The lower stability of smaller particles which follows from the DL VO
theory may be one of the factors affecting limited tlocculation of particles
during polymerization of polar monomers. The importance of this factor
','

7.

267

Polymerization of Polar Monomers

was noted by Dunn and Chong (1970) in a study of the mechanism of


particle formation during polymerization of vinyl acetate in aqueous solution. They also concluded that flocculation throughout the process occurs
between smalI, newly formed particles and larger "oId" particles. It should
be noted, however, that such flocculation is not typical of hydrophobic
monomers.
The restricted adsorption of anion-active emulsifier during polymerization of polar monomers (vinyl acetate) has also been reported in other
work. Breitenbach et al. (1970) established that in the case of polymerization initiated by a,a-azobis(methyl butylonitrilesulfonate Na) the emulsifier (sodium dodecylsulfonate) content in the resuItant latex twice exceeded
the CMC value, Le., the emulsifier was not exhausted by the forming interface. The polymerization rate was found to depend on emulsifier concentration to the power of 0.1.
According to data obtained by Okamura and Motoyama (1953), adsorption of anion-active emulsifiers by the surface of vinyl acetate latex
particles is also limited: an increase in emulsifier concentration (sodium
dodecylsulfate and dioctylsulfosuccinate) above a certain limit does not
produce an increase in adsorption. The authors found that the ratio of the
amount of adsorbed emulsifier to the mas s of formed polymer remains
constants and concluded that part of the emulsifier is located inside
particles of the polar monomer, in contrast to hydrophobic polymers where
it is found only on the surface of particles. In our opinion, the presence of
emuIsifier jnside latex particles may be due to the flocculation of primary
particles from the surface of which not alI the.emulsifier is transferred to the
surface of secondary particles. Of the two emulsifiers studied by Okamura
and Motoyama (1953).the molar quantity of dioctylsulfosuccinate adsorbed
by 1 g of polymer is 1.8-fold higher than that of dodecylsulfate, pointing to
the higher adsorption energy of the former on the surface of poly(vinyl
acetate).
It was established by Yeliseyeva and Bakaeva (1968) that in the polymerization of polar monomers (MA) the decrease of emulsifier adsorption
depends on the structure of the latter and for some types of emuIsifiers may
reach limiting values. This was observed in the polymerization of MA in the
presence of a mixed type of emulsifier, partialIy sulfurated. with suIfuric acid
oxyethylated alkylphenol (emulsifier "C-lO"). Its adsorption on the particle
surface increases with the initial concentration and reaches 100% filling of
the adsorption layer, conditionalIy corresponding to 0.35 nm2 per molecule.
Stable, concentrated latexes with smalI particles are formed. Therefore,
emulsifier adsorption and the mechanism of particle formation associated
with it depends not only on monomer poIarity but also on the chemical
structure of the emulsifier.
~

..

V. 1. Yeliseyeva

268

VI. Kinetics of Emulsifier Adsorption


In view of the specificity of latex system formation, which is directly
related to the adsorption capacity of the forming interface, it is of interest to
examine the kinetics of effective adsorption of emulsifiers of various structures in the course of polymerization of monomers of various polarity
(Yeliseyeva and Petrova, 1972; Yeliseyeva et al., 1973, 1976). The consumption kinetics of SDDS and AAOES during the polymerization of
methyl, ethyl, and butyl acrylates was studied by Yeliseyevaet al. (1976).
Benzoyl peroxide was used as the initiator, and experiments were carried
out in a special dilato meter with platinum electrodes connected to a
conductometer so as to monitor the rate of emulsifier consumption with
increasing conversion via variation of electroconductivity. Since the initator
used is poorly soluble in water, the possibility of polymerization in the
aqueous phase and the formation of "own" surfactants that may mask
adsorption of the introduced emulsifier are exc1uded. In order to obtain
colloidally stable systems 0.1% (with respect' to monomer) tertdodecylmercaptan was added to all runs; the effectiveness of such addition
had been previously shown by Yeliseyeva and Petrova (1972~ The dependence of adsorption of both emulsifiers on conversion during polymerization of the three monomers is shown in Fig. 8. At a certain degree of
conversion (varying with the monomer used) a limiting adsorption value is
reached, a value which decreases with increasing interface polarity from BA
to MA. It differs only slightly for both emulsifiers during BA polymerization. With an increase in the polarity of the interface, both the effective

20

un

I (8)

..
.
....
'1::-,0

"':

20

40

20

40

COT!1rersion(%)

Fig. 8.

Adsorption kinetics of emulsifiers AAOES (A) and SDDS (B) in polymerization

of (1) methyl acrylate;

(2) ethyl acrylate;

and (3) butyl acrylate.

Recipe m.p.: monomer,

benzoyl peroxide, 0.2; emulsifier, 2; tert-dodecyl mercaptan, 1; water, 900. T, 45C.

100;

7.

269

Polymerization of Polar Monomers


(11')

30

(S)

90

30

90

t (min)
Fig. 9.
Conversion (%) kinetics during polyrnerization of (1) rnethyl acrylate; (2) ethyl
acrylate; and (3) butyl acrylate in the presence of SDDS (A) and AAOES (B). For recipe see
Fig.8.

rate and the equilibrium value of the adsorption of AAOES increases as


compared with SDDS. During the MA polymerization the amount of
adsorbed AAOES is twice as high as SDDS. This is apparentIy the reason
for the higher effectiveness of AAOES in emulsion polymerization of lower
alkyl acrylates: increase of the rate and equilibrium value of adsorption of
this type of emulsifier reduces the flocculation of primary particles dui-ing
polymerization and facilitates the formation of coIloidaIly stable systems
with smaIler particles.
Kinetic. curves of the polymerization of these monomers in the presence
of both emulsifiers with their initial concentration close to CMC are shown
in Fig. 9. As can be seen the initial polymerization rate, at the same rate of
radical formation (the same concentrations of initiator and the same
temperature), decreases from MA to BA in accordance with the decrease of
the chain propagation rate constant in these monomers (Bagdasar'yan,
1966). SubsequentIy, however, the polymerization rate of the MA becomes
much lower than the rate of other monomers, a fact which apparently is
associated with the coIloid behavior of the system (reduction of the number
of polymerization centers as a result of flocculation).
Similar resuIts were obtained by Yeliseyeva et al. (1973, 1975) when
studying the sodium alkylsulfonate emulsifier adsorptiQn kinetics during
polymerization of butyl acrylate and its copolymerization with such watersoluble monomers (3%) as methacrylic acid, N-methacrylamide, and Nmethylolmethacrylamide. In aIl cases the introduction of water-soluble
comonomers resuIted in a decrease of the effective rate and in the value of
the emulsifier adsorption. The authors attributed this to the hydrophilization and increase of polarity of latex particle surfaces.

270
VII.

V. 1. Yeliseyeva

Mechanism of Formation and Structure of Particles

Using the method of flow ultramicroscopy by Kudryavtseva and


Deryagin (1963), we investigated the variation in the number of partic1es
during polymerization of ethyl acrylate with ammonium persulfate in the
presence of three types of emulsifiers: anion-active sodium alkylsulfonate,
nonionogenic oxyethylated (30) cetyl alcohol ("C-30"), and mixed type
(AAOES)(Y eliseyeva and Zuikov, 1976; Yeliseyeva et al., 1975)(Fig. 10).For
all three emulsifiers, the number of partic1es value goes through a maximum,
indicating flocculation. Similar curves were observed by Fitch and Tsai (1971)
when studying MMA polymerization in dilute aqueous solutions of the
monomer. Judging from the number of partic1es in the resultant latex (Fig.
10), which in the case of AAOES is twice as high as that for alkylsulfonate,
partic1es flocculate to a smaller extent when using the emulsifier of the
mixed type, indicating their higher stability. Indeed, an estimation of the
adsorption energy of these emulsifiers on a water-ethyl acrylate interface
(see above) gave the following values (kJ/mol): 22.2' for sodium alkylsulfonate, 25.6 for oxyethylated (30) cetyl alcohol, and 25.6 for ammonium
alkylphenylpolyoxyethylenesulfate. These data indicate a higher adsorption
energy for emulsifiers which result in a smaller degree of flocculation.
However, because of their small size the number of partic1es formed at the
beginning of the process could not be determined by flow ultramicroscopy.
The nature of partic1e flocculation occurring during polymerization may be

J
12
o

4-

20

40

60

Con1rerson

(Dio)

.
-
80

Fig. 10. Variation of the number of particles during polymerization of ethyl acrylate in
the presence of emulsifiers: (1) sodium alkylsulfonate;
(2) oxyethylated (30) cetyl alcohol; and
(3) AAOES. Recipe m.p.: ethyl acrylate, 100; emulsifier, 4.0; ammonium persulfate, 0.5; water,
900. T, 60C.

7. Polymerization of Polar Monomers

271

assessed from electro n micrographs of ethyl acrylate latex at 38.0% conversion (Fig. 11A) (before preparation the latexes were irradiated by a 60Co
source, at 20 Mrad, in order to avoid particle deformation). A similar
pattern is observed during the polymerization of a mixture of MMA and
BA (55:45) at 60% conversion (Fig. llB). The complex structure ofparticles
of carboxyl-containing acrylate latex MMA-BA-MAA (55:40:5) may be
revealed by increasing the pH above 9 (Fig. llC). The desaggregation
which occurs is attributable to the electro static repulsion of dissociated
carboxyl groups present in the macromolecules of primary particles.
We were able to reveal the structure of latex particles and films of polar
polymers (acrylate latexes) using a special technique for preparing samples

Fig. 11.
Electron rnicrophotographs
of latex particles: (A) ethyl acrylate at 38 %
conversion; (B) rnethyl rnethacrylate-butyl
acrylate copolyrner (60:40) at 60% conversion; (C) rnethyl rnethacrylate-butyl acrylate-rnethacrylic acid terpolyrner (55:40:5) at
pH 9 and 100% conversion.

272

V. 1. Yeliseyeva

Fig. 12.
Electron micrographs of polystyrene latex particles (A) before etching and (B)
after oxygen etching.

for electron microscopy (Yeliseyeva et al., 1967, 1968). The method of lowtemperature etching with active oxygen formed in an electrodeless highfrequency discharge was used. Under such conditions the less ordered,
looser regions are" burned out" initially, revealing the denser morphological regions (Jacobie, 1960). The latexes were placed on quartz supports, .
subjected to oxygen etching and shaded with platinum at a 20 angle. The
structure of films was studied by the method of single-step carbon replicas
shaded with platinum. Optical periods of etching were determined empirically. The effectiveness of this procedure may be assessed by comparing
electro n micrographs of polystyrene latex particles which were subjected
(Fig. 12B) and not subjected (12A) to oxygen etching. As can be seen, these
particles, contrary to expectation, are not coils of tangled macromolecules
but an arrangement of rather clearly defined polymeric globules (this may
provide a new insight into the mechanism of particle formation during
styrene polymerization). On the other hand, examination of the structure of
poly(vinyl acetate) emulsion particles by the same method gave quite
different results (Fig. 13A and B).

7. Polymerization of Polar Monomers

273

A study of the structure of polar polymer particles was carried out on


MA-BA copolymer latexes; in some experiments methacrylic acid was
added in order to change polymer polarity. Latexes were synthesized at a
phase ratio of 1: 1.5, with the emulsifier concentration in aqueous phase
being 0.07%. Polymerization was carried out by gradual addition of the
monoiners and emulsifier solution into the reaction mixture at 75-78C.
Composition and properties of the latexes are listed in Table VII. The efIect
of oxygen etching on the particle structure is demonstrated in Fig. 14 which
shows electron micrographs of latex number 3 particles (Table VII) at
various periods of etching. As can be seen, they consist of a spherical shell
filled with smaller particles in the form of globules. After 2 min of etching
(Fig. 14A) the shell is ruptured by oxygen and forms a dense ring on the
substrate at the periphery of the particles. At longer periods of etching (10
min, Fig. 14B) the size of primary globules decreases and the peripheral ring
is found to consist of primary globules of the same sizes as those inside the
particles. Because of the den ser arrangement in the ring on the substrate

274

V. 1. Yeliseyeva

(B)

lefA)
Fig. 13.

Electron micrographs of poly(vinyl acetate) emulsion particles (A) before

etching and (B) after oxygen etching.

they are the last to disappear (after 20 min of etching, Fig. 14C). After
etching for an hour all trace of particles on the substrate disappears. In
order to estimate the size of primary globules and their content in latex
particles, etching was carried out of films obtained by drying of the same
latexes (Table VII).
Electron micrographs of replicas of films obtained by drying latexes
numbers 1, 2, and 3 (Table VII) at room temperature after 20 min oxygen
etching are shown in Figs. 15A, 15B, and 15C. In the comer of each
micrograph a medium-sized particle of the corresponding latex at the same
enlargement is shown. The main structural elements of films are primary
globules which are much smaller than latex particles. Taking the diameter
of primary globules as the distance between their centers, for latex number 3
(Table VIII) we conclude that these distances coincide in micrographs of
particles and films, i.e., the structural element of both particles and films is a

7.

275

Polymerization of Polar Monomers


TABLE VII

Composition and Properties of Latexes used in the Study of Particle Structure

MAA

Latex
conc
(%)

Molecular
mass of
polymer
(10- 3)a

O
2.6
4.3

37.95
39.34
39.6

214
203
227

Polymer composition
(%mol)
No.

MA

BA

1
2
3

65
63.3
62.3

35
34.1 .
33.4

Molecular

mass of copolymers

was calculated

Partic1e
diameter
(nm)

Latex
pH

118-210
120--220
118-200

from specific viscosity of their solutions

2.9
2.8
2.8
in

dioxane using Staudinger's coefficient for poly(alkyl acrylates) Staudinger and Trommsdorff,
1933).

globule that apparently constitutes th primary particle formed at the initial


stages of the process.
The size of primary particles is different for different latexes; they
decrease with increasing polymer phase polarity (Fig. 15). Data of approximate calculation of sizes of primary particles, the number of macromolecules in them, and the content of primary particles in particles of the
resultant latex are given in Table VIII; data for a butyl methacrylate latex
synthesized under similar conditions wherein primary particles were not
revealed by the same oxygen etching method are also presented in this
~~.

From this discussion it may be concluded that latex particles forming


during the polymerization of polar monomers are the result of limited
flocculation of primary particles. The mechanism of flocculation may be as
follows. In the case of a water-soluble initiator, the process begins in the
aqueous phase. Upon reaching a critical degree of polymerization polymeric radical s -S04(M)nM precipitate aggregating into nuclei of primary
particles which are stabilized partially by the polarity of the polymer itself
(polar groups such as S04, COO-, COOH, COOR) and partially by the
dynamic adsorption of the emulsifier. Increase in the size of primary
particles may result from two factors: polymerization of the absorbed
monomer and aggregation with other polyme~ic radicals. The latter is
increasingly slowed down with increasing surface charge density of polymeric radical s and with hydrophilization of their surface. Indeed, as can be
seen from data in Table VIII primary particles differ in size, and for a
polymer of medium molecular weight may contain from a few to several
dozen macromolecules, depending on the polarity of the polymer phase (the
size of primary particles decreases with increasing polarity).

276

V. 1. Yeliseyeva

Fig. 14.
Electron micrographs
of polyacrylate latex particles (Table
VII. run 3) after oxygen etching for:
(A) 2 min; (8) 10 min; and (C) 20 mino

Under dynamic conditions of emulsion polymerization the process of


formation and aggregation of primary particles is accompanied by establishment of an equilibrium adsorption of the emulsifier on the polar surface .
of particles. As a result of competition of these processes.. "secondary"
particles of complex structure are formed (Figs. 11 and 14). The rate of
attainment of equilibrium adsorption, depending on monomer polarity and
the nature of the emulsifier, determines the degr~e of limited tlocculation of
primary particles at which aggregate stability of the "secondary" latex
particles is reached. The formation of nuclei from micelles during the

7.

277

Polymerization of Polar Monomers


TABLE VIII

Structural Characteristics of Latex Particles of Acrylic Polymers of Various Polarity

No.

3
4

Polymer
composition
(%rnol)
BA-MA
(65:35)
BA-MA-MAA
(63.3:34.1:2.6)
BA-MA-MAA
(62.3:33.4:4.3)
BMA"

Partic1e
diameter
(nm)
latex

prirnary

Numbers
of primary
partic1es
in latex
partic1e

Number of
macromolecules
in
primary
partic1e

118-210

40

-120

-90-150

120-220

25-30

-530

- 25-40

118-200

15-20

-1000

15-40

no

no

-5-10

u In order to ensure necessary stability of the latex during polyrnerization of BMA the
initial ernulsifier concentration in aqueous phase was taken equal to 3.3%.

polymerization of hydrophobic monomers, which polymerize more slowly


and better adsorb the emulsifier, is accompanied by the establishement of
equilibrium adsorption of the emulsifier on account of which "organized"
flocculation does not occur and as a result of which much smaller particles
are formed. However, at a high adsorption capacity of the forming nonpolar surface, the emulsifier is initially rapidly consumed, and the formation

Fig. 15.
Electron rnicrographs of replicas frorn the surface of polyacrylate latex filrns:
(A) run 1, Table VII; (B) run 2, Table VII; (e) run 3, Table VII.

278

V. 1. Yeliseyeva

of particles at the end of the processwith poorly protected surfacemay lead


to an avalanche of unlimited flocculation and precipitation of coagulum.
When using more active emulsifiers of the C-10 type (Yeliseyeva, 1966)
the effective rate and the equilibrium value of adsorption increase in the
polymerization of polar monomers, which reduces the intensity of the
processes of initial flocculation. Moreover, in this case the charge density is
not the only stabilizing factor; it is supplemented by the enthalpy factor
which arises from hydration of oxyethylenic chains of emulsifier molecules.
The presence of these emulsifiers creates conditions that enable the obtainment of polar polymer latexes of high stability.
Generalizing, it may be concluded that the limited flocculation of
primary particles depends on the ratio of the rate of formation of their
overall surface (S) to the rate of establishment of equilibrium adsorption (r)
of the emulsifier by the surface.
For formation of particles stabilized by the emulsifier
dr/dt = K dS/dt

(5)

where K is the proportionality constant which takes into account the value
of equilibrium adsorption.
In reality however it may occur that
dr/dt < K dS/dt

(6)

,. Depending on whether or not the condition [Eq. (5)] is met the


mechanism of particle formation will be different. When this condition is
satisfied the forming surface of the polymer phase is stabilized by the
adsorbing emulsifier and the formed particles may remain discrete at high
degrees of conversion. In this case the polymerization rate is described by
Eq. (1) which is derived by assuming a constancy in the number ofparticles.
In the case of inequality, as expressed by Eq. (6) for example, at low surface
activity of the emulsifier on the polar polymer surface and at high polymerization rates, particles appear whose surface is insufficiently protected by the
emulsifier. Under conditions of high temperature and intensive agitation,
such particles rapidly flocculate with each other and an equilibrium adsorption of the emulsifier is established at the formed aggregates latex particles.
In such cases the polymerization rate does not obey Eq. (1).

VIn.

Polymerization Kinetics

In the study of emulsion polymerization kinetics which depend on


monomer polarity, the homologous series of alkyl acrylates and alk.yl
methacrylates are usually used because of their wide ranges of solubility in

279

7. Polymerizationof Polar Monomers

water, their reactivity, and the fact that they do not display any particularly
specific features of polymerization (as do vinyl chloride and vinyl acetate
which have a high constant of transfer onto the monomer or as do vinyl
chloride and acrylonitrile where there is low solubility of polymer in
monomer).
The differences in polarity of alkyl acrylates (alkyl methacrylates) is
understood as the difference in polarity at the interface with water, since
taken separately lower alkyl acrylates have practically equal dipole moments (Granzhan, 1969). However, the structure of the alkyl acrylate-water
interface layer changes with increasing alkyllength: the polarization of the
carbonyl group, which determines polarity of the asymmetric alkyl acrylate
molecule at the water boundary, decreases as a result of an increase of the
induction effect of the alkyl group. The degree of orientation of monomer
molecules decreases and interfacial tension increases. Medvedev (1956)
proposed that molecules of some acrylic monomers have a negative charge
in aqueous phase.
Yeliseyeva et al. (1976) studied the polymerization kinetics of polar
acrylic monomers using ethyl acrylate. The dependence of the polymerization rate on the initiator concentration (ammonium persulfate) was
examined at equal ionic strength in the presence of two emulsifiers, sodium
dodecylsulfate and oxyethylated (30) cetyl alcohol. Polymerization was
carried out in a dilato meter at 45C, monomer-water phase ratio 1:9, and
ammonium persulfate concentration 0.44-4.4 x 10-3 mol/dm3. The obtained loagrithmic dependence is shown in Fig. 16. In each cases the
1.5

0.5

1.0

1.5

2.0

-4 +tg Ct(mol/dm')
Fig. 16.
Dependence of ethyl acrylate polymerization rate (R mol/dm3 sec) on
a~monium persulfate concentration (mol/drn3) in the presence of (1) 13.8 x 10-3 mol/drn3
SDDS; (2) 2.55 x 10- 3 rnol/drn3 oxyethylated (30) cetyl alcohol. Phase ratio, 1: 9; T, 45C.

280

V. 1. Yeliseyeva

dependence is represented by a straight line with the same slope, the tangent
being equal to 0.5. It should be noted that such a clearcut dependence is not
observed when the ionic strength is not controlled. This may be associated
.

with the effectof this factor on polymer-monomer solubility [which was


established by Klein et al. (1973) for the vinyl acetate-poly(vinyl acetate)
system] and on the degree of particle surface protection by the emulsifier
(Piirma and Chen, 1980).
Typical curves of the dependence of ethyl acrylate polymerization rate on
conversion in the presence of sodium dodecylsulfate and initiated by
persulfate are shown in Fig. 17 together with the variation in the number of
particles determined by flow ultromicroscopy with increasing conversion
(Kudryavtseva and Deryagin, 1963). As can be seen, the number of particles
significantly varies during the constant rate period: after formation of a
large number of particles at the beginning, their number gradually decreases; the number of particles also decreases when emulsifier concentration in the resultant latex exceeds the CMC value ,(Curve 2). However,
within the range of concentrations studied, the resultant number of particles
increases as does the polymerization rate with emulsifier concentration.
The ethyl acrylate polymerization rate remains constant up to 38-40%
conversion (Fig. 17), whereas according to data on equilibrium swelling of
ethyl acrylate latex particles (Mamadaliev, 1978) the monomer phase disappears at 20% conversion. Hence, the polymerization rate remains constant
after the disappearance of the monomer drops. Since the monomer drops

.
o
Fig. 17.
Variation of polymerization rate (R mol/dm3 sec) (Curves I and 2) and of the
number of partic1es (N) (Curves l' and 2') with conversion (5) during polymerization of ethyl
acrylate in the presence of (1) 0.3% and (2) 0.8% SDDS with respect to the monomer. Phase
ratio, 1: O; T, 45C.

7.

281

Polymerization of Polar Monomers

maintain the equilibrium concentration of the monomer in the partic1es, in


the absence of a monomer concentration gradient in the partic1es a constant
rate should have been observed only untiJ disappearance of the monomer
phase. On the other hand, this inconsistency in continuation of the constant
rate period after the disappearance of the monomer phase does not take
place in the case of alkyl acrylate polymerization initiated by benzoyl
peroxide (Yeliseyeva et al., 1976), as can be observed from rate versus
conversion curves for polymerization of MA, EA, and BA initiated by
benzoyl peroxide in the presence of two emulsifiers (SDDS and AAOES).
Figure 18 shows that the end of the constant rate period for both emulsifiers corresponds to 15.8% conversion for MA, 26.6 for EA, and 32.4%
for BA, which coincides approximately with the moment of disappearance
of the monomer phase. Therefore, continuation of the constant rate period
after the disappearance of the monomer phase is typical only of persulfate
initiation. When using a poor1y soluble initiator this inconsistency is not
observed, and the polymerization of alkyl acrylates obeys in this respect the
regularities of c1assical kinetics.
A possible explanation for the observed difference in duration of the
constant rate period is concerned with the differences in structure and
formation mechanism of partic1es. In the case of a poor1y water soluble
initiator the process begins in emulsifier micelles from which PMP are
formed, the latter being stabilized from the very beginning. With an increase
in conversion and partic1e size the surface of partic1es is protected by the

(In

(8)

i:I:t 0.5

20

50

20

80

C07!Versio7! (%)
Fig. 18.
Rate (R %/min) versus conversion dependence during polymerization of (1)
methyl acrylate; (2) ethyl acrylate; and (3) butyl acrylate in the presence of SDDS (A) and
AAOES (B). The graphs are plotted from the kinetic curves in Fig. 9. Initiator benzoyl
peroxide.

282

V. 1. Yeliseyeva

adsorption of emulsifier and the rate of attainment of equilibrium adsorption value is sufficiently high for protection of partic1es from flocculation.
The protection is also enhanced by the absence in the surface layer of
negatively charged S04' groups. Therefore, the probable ratio between
living and dead partic1es is retained at 1: 1 as required by the Smith-Ewart
(1948) theory, and the duration of the constant rate period approximately
coincides with the period of the existence of the monomer phase.
Because of the considerable solubility of ethyl acrylate in water in the
case of persulfate initiation, the process begins in the aqueous phase with
the formation of charged polymer radicals which, after reaching a certain
critical size, precipitate forming primary PMP. Establishment of equilibrium adsorption of the emulsifier, which ensures stability of the partic1es, is
preceded by the flocculation of partic1es with other primary PMP which at
the beginning is accomplished by interpartic1e termination of radicals. Thus,
a new termination factor appears which is not considered in the above
theory and which leads to a higher relative content of d;:adpartic1es. This is
confirmed by data on the average number of radicals per partic1e which
drops to 0.07 for persulfate ibitiation of ethyl acrylate polymerization
(Mamadaliev, 1978). The large number of equilibrium monomer-swollen
partic1es creates, after the disappearance of the monomer phase, a reserve of
monomer in the system which, being redistributed via the aqueous phase
into the growing partic1es, facilitates maintenance of equilibrium monomer
concentration and thus continuation of the constant rate period after
disappearance of the monomer phase. This may also be facilitated by the
complex structure of partic1es and the existence of a monomer concentration gradient.
The logarithmic dependence of the ethyl acrylate polymerization rate on
concentration of emulsifiers of various chemical structure is shown in Fig.
19 (Yeliseyeva et al., 1976). Polymerization was carried out in a dilatometer

f.O

2.0

- i.+ Lg Cc (mol/dm3)
Fig. 19.
Dependence of ethyl acrylate polymerization rate (R moljdm3sec) on
concentration (moljdm3) of (1) SDDS; (2)sulfated polyoxyethylated (10) octadecyl alcohol; and
(3) oxyethylated (30) cetyl alcohol.

7.

Polymerization of Polar Monomers

283

at 45C at a phase ratio of 1: 9, at an ammonium persulfate concentration


of 2.2 x 10-3 moljdm3 in the aqueous phase. The polymerization rate was
determined during the constant rate periodo As can be seen, the rate
increases only up to a certain value of emulsifier.concentration, after which
the rate remains constant. The dependence is strongly affected by the nature
of the emulsifier: for the emulsifiers studied the reaction order is 0.4 for
C-30 [oxyethylated (30) cetyl alcohol], 0.25 for sulfatized oxyethylated
(10) octadecyl alcohol, 0.21 for sodium dodecylsulfate.
On the basis of the data presented the kinetic equation forethyl acrylate
polymerization in the presence of various emulsifiers may be written as
foIlows:
(7)
where x is the reaction order with respect to emulsifier which varies from
0.2 to 0.4 depending on the chemical structure of the emulsifier. Such
dependence must be associated with the energy of adsorption of the latter at
the interface.

IX.

Relationship between Polymerization Kinetics and Adsorption


Characteristic of the Interface

In order to establish whether Eq. (7) may be extended to other polymerization systems (Yeliseyeva and Zuikov, 1977; Yeliseyeva et al., 1976, 1978) a
study was conducted on the dependence of.the polymerization rate on the
concentration of dodecylsulfate and its homologs in the polymerization of
other homologs of alkyl acrylates. The logarithmic dependences of polymerization rate of alkyl acrylates on sodium dodecylsulfate concentration is
presented in Fig. 20. With the shortening of the alkyl chain, the slope of the
straight line is regularIy decreasing, indicating a weakening of the rate
dependence on emulsifier concentration. A similar dependence for a wide
range of monomers differing in water solubility was earlier observed by
Gershberg (1965). Data presented in Fig. 20 indicates the general trend of
the established dependence. This is further confirmed by the graph for
styrene which also conforms to this regularity. The value of x determined
from the slopes of the straight lines regularIy increase from MA to hexyl
acrylate (HA) to styrene (from 0.13 to 0.46 to 0.5, respective1y) and from
dedecylsulfate to hexadecylsulfate (from 0.29 to 0.33 respectively) (Table
IX).
In order to establish the correlation between the value of x and emulsifier adsorption characteristics in the given reaction system in latexes
prepared according to the same recipes, as was used in determination of
~eaction order, the adsorption areas of emulsifier molecules in the saturated

284

V. 1. Yeliseyeva

2.0

1.8

1.6

~gCe'.l11oZ
/dmJ)
Fig. 20.
Dependence of the polymerization rate (R mol/dm3 sec) of (1) methyl acrylate;
(2) ethyl acrylate; (3) butyl acrylate; (4) hexyl acrylate; and (5) styrene on sodium
dodecylsulfate concentration. Ammonium persulfate 0.025% in aqueous phase; phase ratio
1:4; T, 60%.

adsorption layer (ASlim)were determined (Yeliseyeva et al., 1978). As follows


from Table IX there is a relationship between these values and the energy of
emulsifier adsorption on the initial monomer-water interface, as determined
from adsorption isotherms.
As can be seen from Table IX the product
X. ASlim

=K

(8)

is approximately constant for most polymerization systems. Deviation from


the constant value toward lower values is observed for the two lower alkyl
acrylates and SDDS, Le., in systems characterized by the lowest adsorption
energy of emulsifier. In these systems an increase nf the real dynamic value
ASdynas compared with its equilibrium value ASlim(p. 263) is most probable.
As shown even for polymerization of EA in the presenC'e of SDDS, the
overall partic1e surface does not vary up to 60% conversion (Fig. 21)
(Mamadaliev, 1978) except for a smallleap at the earliest stage. This may be
considered as eV;idence in favor of the partic1e surface being the main
polymerization site. Since from Eq. (8) the value of the overall interface is

s = K 1 . ASlim

(9)

285

7. Polymerizationof Polar Monomers


TABLEIX

Adsorption and Kinetic Characteristics of PolymerizationSystems Based


on AlkylAcrylates, Styrene, and VariousAlkylSulfates
Monomer
MA
EA
EA
EA
BA
HA
Sty

Emulsifier

Ead.
(kJ/mol)

Asum
(nm2)

K (nm2)"

20.0
22.1
23.3
24.6
25.6
-

1.51
0.92
0.82
0.74
0.67
0.52
0.48

0.197
0.198
0.238
0.244
0.222
0.240
0.240

0.13
0.21
0.29
0.33
0.33
0.46
0.50

C12H2SS04Na
C12H2SS04Na
C14H29S04Na
C16H31S04Na
C12H2SS04Na
C12H2SS04Na
C12H2SS04Na

" K = x . Asum.

it follows that x

= K2/S.

Thus the role of emulsifier

in polymerization

kinetics decreases with decrease in its participation in the creating of the


overall interface area. This supports Med.vedev's theory that the main
zone of emulsion polymerization is the surface of the forming particIes. The
shortcoming of his scheme [Eq. (2)] is the assumption that the magnitude of
this zone is determined by the concentration of the emulsifier, whereas in
reality it also depends on the polarity of the polymer-monomer phase
itself.
Summing up the results presented here it may be concIuded that in the
polymerization rate equation the reaction o,rder with respect to emulsifier is
a variable and a function of the area occupied by the emulsifier molecule in
the saturated adsrption layer on particIes of the resultant latex
x = f(Aslim). The basic difference of this equation from the equations
proposed earlier [Eqs. (1) and (2)] is the presence of a parameter that
characterizes the properties of the interface (adsorption capacity), which is
necessary for a kinetic equation of polymerization occurring in a colloidal

6
20

5/.

'+0

o
60

Converston (0/0>
Fig. 21.
Dependence of overall surface of particles (5) on conversion during
polymerization of ethyl acrylate. SDDS 1.04 x 10-3 mol/dm3; ammonium persulfate
0.44 x 10- 3 mol/dm3; phase ratio, 1: 9; T, 45C.

286

V. 1. Yeliseyeva

system. Both characteristics of the interface discussed-the energy of the


emulsifier adsorption on the water-monomer interface and the degree of
protection of partic1e surface-are
important for processes of partic1e
flocculation and mass transfer between phases.
The kinetic scheme of emulsion polymerization of polar monomers
initiated by persulfate which follows from the experimental data described
above is presented in Table X in comparison with the c1assicalscheme.
TABLEX
Kinetic Se heme of Emulsion Polymerization during
the Constant Rata Perioda
Classical
(for hydrophobic

N = const
Am = ASlim

For polar monomers


(of the type of acryl acrylates)

monomers)

N = ((Ead"t)
S = const

= const

R = KC~_6C?_4

ASlim= I(Ead,)
R = kC~ . C?-5[X = ",(Ead,) =

tr.con~1 =

K/Aslim = K'/AsdyoJ
tr'con~1> !m'ph

t'm'ph

aAbbreviations: N, number of particles during the constant rate


period; Amand ASlim,areas occupied by emulsifier molecule in the micelle
and saturated particle surface; "-C0051
and 'm-ph' duration of the constant
rate period and polyrnerization period up to disappearance of the
monomer phase; t, polyrnerization period; S, overall partlcle surface.

Nomenclature
Am
As
ASlim
C.
C

D
Ead'
I:1G

area of emulsifier molecule in a micelle


calculated value of the area occupied by emulsifier molecule on the particle surface of
resultant latex assuming that all the emulsifier is adsorbed
adsorption area of emulsifier molecule at saturation of the adsorption layer
ernulsifier concentration
initiator concentration
diarneter of latex particles
overall free energy of adsorption
free energy of adsorption per CH2 group
particle monodispersity
coefficient

= DJD =
w

kp
N

propagation
rate constant
number of particles

p
R

(ASlim!As}-degree
of adsorption
polymerization
rate

n-D~

1/3

"~ n-D~

-1/3

~ni ) (LniD~)
~

of particles

of resultant

latex

7. . Polymerization of Polar Monomers

s
x

287

overall particle surface


polymerization
period
reaction order in emulsifier

1'12
interfacial tension at the boundary of the first and second phases
I'
surface tension at the ith phase
1'1,1'~2 surface or interfacial tension due to dispersion interaction
I'f, I'Y2 surface or interfacial tension due to polar interaction

r
adsorption value
'm.ph duration of monomer phase existence
'r'consl duration of the constant rate period

References
Bagdasar'yan, Kh. S. (1966). "Theory of Radical Polymerization," p. 102. Nauka, Moscow.
Bakaeva, T. V., Yeliseyeva, V. 1., and Zubov, P. 1. (1966). Vysokomol Soedin. 8, 1073.
Breitenbach, J. W., Kuchner, K., Fritze, H., and Tarkowsky, H. (1970). Dr. Polym. J. 2, 13.
Davis, J. T., Higuchi, T., and Rytting, J. H. (1972). J. Pharm. Pharmacol. 24, 30.
Dunn, A. S., and Chong, Z. C-H. (1970). Dr. Polym. J. 2, 49.
Fitch, R. M., and Tsai, C. H. (1971). "PolymerColloids" (R. M. Fitch, ed.), pp. 73,103. Plenum
Press, New York.
Fowkes, F. M. (1964). Ind. Eng. Chem. 56, 40.
Gerrens, H. (1964). DECHEMA Monogr., Frankfurt a/M 49, (859), 346.
Gershberg, D. (1965). J. Chem. Eng. Symp. Ser. Inst. Chem. Eng. N3, 3.
Granzhan, V. A. (1969). Zh. Vses. Khim. Obshch. 14, 223.
Gritsenko, G. M., and Medvedev, S. S. (1956). Dokl. Akad. Nauk SSSR 11, 235.
Harkins, W. D. (1947). J. Am. Chem. Soco 69, 522.
Jacobie, E. (1960). Electron Microsc. 1.
Kaeble, D. H. (1971). "Physical Chemistry of Adhesion," pp. 153-170. Wiley (Interscience),
New York.
Kiselev, M. R., Evko, E. t.. and Luk'yanovich, V. M. (1966). Zavodskaya Lab. 32, 201.

Klein,A.,Kuist,C. H., and Stannett,V.T. (1973).J. Polym.Sci. 11,211I.


Kudryavtseva, N. M., and Deryagin, B. V. (1963). Kolloid. Zh. 25, 739.
Mamadaliev, A. (1978). Candidates Dissertation. Institute of Physical Chemistry Akad. Nauk
SSSR, Moscow.
Medvedev, S. S. (1968). "Kinetics and Mechanism of Formation and Conversion of
Macromolecules," pp. 5-24. Nauka, Moscow.
Nikitina,S. A., Konstantinova,V.V.,Zakieva,S. Kh., and Taubman,A. B.(1961).Zh. Pricklad.
Khim. 34, 2658.
Okamura, S., and Motoyama, T. (1953). Kyoto Univ. 15, (4), 242.
Paxton, T. R. (1969).J. ColloidInterface Sci. 31, 19.
Piirma, l., and Chen, S-R. (1980). J. Colloid Interface Sci. 74, 90.
Rhebinder, P. A. (1927). Z. Phys. Chem. 129, 161.
Smith, W. V., and Ewart, R. H. (1948). J. Phys. Chem. 16,592.
Snuparek, 1., and Tutalkova, A. (1978). Int. Symp. Polydispersionen, Dresden, Kurtzreferate N3.
Staudinger, H., and Trommsdorff, E. (1933). Ann. Ed. 502, 201.
Stterlin, N., Kurth, H. J., and Markert, G. (1976). Makromol. Chem. 177, 1549.
van der Hoff, B. M. E., (1960).J. Polym. Sci. 48,175.
Vijayendran, B. R. (1979). J. Appl. Polym. Sci. 23, 733.

288

V. 1. Yeliseyeva

Wolfram, E. (1966). Kolloid Z. Po/ym. 211, 84.


Wu, S. (1974). J. Macromo/. Sci. C. 10, l.
Yakovlev, Y. M., Lebedev, A. V., and Fermor, N. A. (1971). "Prob1ems

ofSynthesis,

Study of

Properties and Treatment of Latexes," pp. 148-158. TSNITE Neftekhim., Moscow.


Ye1iseyeva, V. 1. (1966). USSR Patent No. 203899.
Yeliseyeva, V. l. (1972). Acta Chim. Acad. Sci. Hungar. 71, 465.
Yeliseyeva, V. 1., and Bakaeva, T. V. (1968). Vysokomo/. Soedin. Ser. A 11,2186.
Yeliseyeva, V. 1., and Petrova, S. A. (1970). Vysokomo/. Soedin. Ser. A 12, 162l.
Yeliseyeva, V. 1., and Petrova, S. A. (1972). Dok/. Akad. Nauk.SSSR
202 N(2), 374.
Yeliseyeva, V. l., and ZiJikov, A. V. (1976). Am. Chem. Soco Ser. 24, 62-82.
Yeliseyeva, V. l., and Zuikov, A. V. (1977). Vysokomo/. Soedin. Ser. A 19, 2617.
Yeliseyeva, V. l., Zaides, A. L., and Zubaryan, K. M. (1965). Dok/. Akad. Nauk. SSSR 162, 1085.
Yeliseyeva, V. l., Zharkova, N. G., and Evko, E. 1. (1967). Vysokomo/. Soedin. Ser. A 9, 2478.
Yeliseyeva, V. l., Zharkova,
178, 1113.
Yeliseyeva,
Ye1iseyeva,
Yeliseyeva,
Yeliseyeva,

V.
V.
V.
V.

l.,
1.,
l.,
l.,

N. G., and Luk'yanovich,

V. M. (1968). Dok/. Akad. Nauk SSSR

Petrova, S. A., and Zuikov, A. V. (1973). J. Po/ym. Sci. Part C 63.


Zuikov, A. V., and Kalnin'sh, B. Ya. (1975a). Kolloid. Zh. 37, 152.
Zuikov, A. V., and Mamadaliev, A. (1975b). Zh. Fiz. Khim. 49, 1044.
Mamada1iev, A., and Zuikov, A. V. (1976a). Vyso.komo/. Soedin. Ser. A 18,

114l.
Yeliseyeva, V. 1., Zuikov, A. V., and Lavrov, S. V. (1976b). Vysokomo/. Soedin. Ser. A 18,648.
Yeliseyeva, V. 1., Zuikov, A. l., Vasilenko, A. l., and Shehepetil'nikov,
B. V. (1978). Int. Symp.
Macromol. Chem. Tashkent 89-90.
Yuzenko, V. l., and Mints, S. M. (1945). Dokl. Akad. Nauk SSSR 47(2),106.
Zuikov, A. V., and Soloviev, J. V. (1979). Acta Polym. 30, 503.
Zuikov, A. V., and Vasilenko, A. 1. (1975). Kolloid. Zh. 37, 640.
Zuikov, A. V., Vasi1enko, A. l., and Yeliseyeva, V. l. (1976). Vysokomol Soedin. Ser. A 18,707.

CON/\.CYT
<i> PROYECTO

8
Recent Developments
and Trends in the
Industrial Use of Latex
Carlton G. Force

1. Introduction .
11. Factors in Adhesion .
A. Physical Adsorption
B. Diffusion.
C. Electrical Charge Distribution

D. Chemisorption .
E. Tack .
111. Bonding Applications .
A. Adhesives
B. Fiber Bonding
C. Tufted Carpet Applications
D. Paints and Industrial Coatings
E. Paper Coatings .
.'
IV. Construction Applications .
V. Rubber Goods
VI. Properties of Various Latexes
References

l.

289
291
291
296
297
298
298
300
300
302
306
309
310
312
313
314
316

Introduction

A uniqueness common to latexes which is advantageous in


many of their applications is the small size of the polymer particles (0.050.5 J.lm)suspended in the aqueous medium. Among the advantages of such
small particle size is easy penetration into the smallest crevices of the
substrates ~eing treated with thf; polymer, maximum efficiencyin deposition
of thin films of the polymer, and optimum optical control of the films
produced from the polymer particles because of the light scattering characteristics that occur over the particle size ranges available. The intense
Brownian motion of such small particles maintains the colloidal stability of
289
EMULSION POLYMERIZATION

Copyright:> 1982by AeademiePress. Ine.


Al! rights of reproduetion in any forro reserved.
ISBN 0-12-556420-1

290

Carlton G. Force

latexes for very long periods even when there is a considerable discrepancy
between the density of the polymer and that of the aqueous medium.
This small particle size is a characteristic necessity of the emulsion
polymerization reaction permitting good heat control through the aqueous
medium at rapid polymerization rates. However, technology is emerging
(Vanderhoff et al., 1979) which permits the preparation of emulsions in the
latex particle size range from polymers that have previously been unavailable as latexes primarily because of sensitivity to water in their synthesis.
This technology is expected to expand rather rapidly as the understanding
of methods for producing microemulsions in the latex particle size range is
transferred from monomeric systems, such as those being dealt with in
tertiary oil recovery research (Shah and Chan, 1980), to polymeric applications. This type of technology has already been used commercially for
more than a decade for the preparation of polyethylene emulsions by the
indirect pressure method (von Bramer and McGillen, 1966).
Availability of such a universally convenient me~ium of application
permits optimization of product properties through polymer designo The
most fundamental property of polymers deposited from latexes is their glass
transition temperature. This dictates the temperature conditions under
which they form continuous films and the degree of softness or flexibility
imparted to the item containing the latex polymer. The glass transition
temperature is characterized by the monomer making up the backbone of
the polymer molecule. However, it can be modified without overwhelming
other desirable characteristics by the copolymerization of such comonomers
as low glass transition temperature acrylic esters with high glass transition
temperature vinyl chloride to produce copolymers incorporating the advantages of poly(vinyl chloride) (PVC) but with glass transition temperatures
down to below room temperature. Comparable technology is used in the
production of many plastics and elastomers, for example, poly(vinyl acetate)
and styrene-butadiene (S-BR). Flexibility can also be imparted to films
developed through the fusion of high glass transition temperature polymers
when a phase inversion takes place in the polymer melt to incorporate
thoroughly the plasticizer in which the polymer had been previously
suspended. Again, this is common practice in the utilization of poly(vinyl
chloride) produced by the emulsion polymerization process (e.g., plastisols
and organosols).
In the utilization of emulsion polymers many specific characteristics are
imparted by the specific polymer to the product. Although cost is always a
factor in producing any commodity, achieving optimum characteristics for
end use at an acceptable price is of course the ultimate aim. Latexes have
found many of their opportunities through permitting achievement of this
goal by unique designs producing equal or better utility more cost ef-

8. Recent Developments and Trends in the Industrial Use of Latex

291

ficiently than the products they replace. Thus, the versatility permitted with
latexes for the ingenious incorporation of the proper polymer into a
construction allows tremendous latitude, with the price of the polymer often
being a minor consideration compared to the properties imparted. This
chapter is organized around current applications of latexes. The specific
properties imparted by the various latexes utilized for each application will
be pointed out. New and emerging technology in each field will be
emphasized. However, the main objective will be to pro vide sufficient
background and stimulation to enable the reader to begin synthesizing
innovative emulsion polymers.
11. Factors in Adhesion
Regardless of the end use, factors that have been recognized as important
in adhesion are usually involved to some extent. Therefore, an outline of
these factors is desirable. An excellent comprehensive review of this subject
was recently published by Wake (1978). The factors responsible for adhesion can be subdivided into four categories: (i) physical adsorption, (ii)
diffusion, (iii) electrical charge distribution, and (iv) chemisorption, all of
which i~teract to some degree in every adhesive bond. A fifth characteristic
relating primarily to the adhesive polymer is tack.
A.

Physical Adsorption

Physical adsorption is a universal phenomena, producing some, if not the


major, contribution t.o almost every adhesive contact. It is dependent for its
strength upon the van der Waals attraction between individual molecules of
the adhesive and those of the substrate. Van der Waals attraction quantitatively expresses the London dispersion force between molecules that is
brought about by the rapidly fluctuating dipole moment within an individual molecule polarizing, and thus attracting, other molecules.
Grimley '(1973) has treated the current quantum mechanical theories involved in simplified mathematical terms as they apply to adhesive
interactions.
For qualitative purposes, it is instructive to look at the equation
developed by Hamaker (1937) for calculating the van der Waals force of
attraction between materials.
FA = A/6nd"
(1)
From this equation, it is seen that the force of attraction FA is proportional
to a constant A commonly called the Hamaker constant. This constant is
dependent upon the molecular composition of the materials. For example,

292

Carlton G. Force

the Hamaker constants of metals and their oxides suspended in water


generally fall in the 1O-13-erg range. Organic polymers suspended in water
usually have Hamaker constants in the 10-16 to 1O-14-erg region, with the
lower surface energy materials, such as tluorinated hydrocarbons, having
the smaller values (Visser, 1972). Experimentally, polystyrene in latex has
been found to have a Hamaker constant on the order of 10-14 to 10-13 erg
(Ottewill and Shaw, 1966; Watillon and Joseph-Petit, 1963). In the author's
doctoral thesis work utilizing a commercial styrene-butadiene latex, the
Hamaker constant for this polymer was established, in water medium, to be
about 3 x 10-14 erg (Force and Matijevic, 1968). This value has since been
confirmed using a better defined latex (Green and Saunders, 1970) and by
theoretical calculations (Fowkes, 1967).
In the denominator of Eq. (1) is inc1uded the distance of separation of the
surfaces from each other d raised to a power n. Since the van der Waals
attraction is an inverse function of distance of separation d, the c10ser the
molecules are together, the greater the force of attraction. This effect is
accentuated even more by the power n to which d is raised; n also decreases
as the distance of separation becomes less. From a distance d of about
800 A separation between surfaces, n gradually decreases from 3 to 2 at
about 80 A separution (Brisco and Tabor, 1972).
This points out dramatically why, for good adhesion, there must be
intimate contact between the adhesive and the substract. Initimate contact
is also necessary for the diffusion and chemisorption mechanisms whereas
electrical charge distribution assists the van der Waals forces in assuring
this contact. Promotion of contact by methods such as applying pressure is
important in insuring optimum bond strength. In fact, application of
sufficient pressure sometimes enhances bonding even more than can be
predicted from these surface force interactions because deformation of the
adhesive into the micro roughness present on the surface of any solid
substrate produces an increased area of contact. This dynamic phenomena
is virtually impossible to calculate accurately, or even measure, because of
the minute distances and small forces involved. Therefore, theoreticians
have found it necessary to formulate their models on thermodynamic
considerations involving the equilibrium situations of liquids adhering to
solids and to ignore the kinetic happenings involved in placing a solid in
contact with, and then separating it from, another solido
The primary thermodynamic consideration involved is that of wetting or
spreading. For the adhesive to achieve the molecular c10seness to the
substrate required for strong van der Waals forces to develop, it must wet
the substrate. In order for this to happen, the spreading coefficient S
(Harkins, 1941)as defined by the equation
(2)
S = Ysv - (YSL+ YLv)

8. Recent Developments and Trends in the Industrial Use of Latex

293

must be positive. In Eq. (2), Ysvrefers to the surface free energy of surface
tension of the solid, YLv is the surface tension of the liquid, and YSLis the
interfacial tension between the solid and liquido In other words, the surface
free energy of the solid must be decreased by the combination of the surface
free energy of the liquid plus the interfacial free energy between the solid
and the liquid for spontaneous spreading and wetting of the solid by the
liquid to take place. For solids, neither Ysvnor YSLis directIy measurable.
However, two methods have been developed which permit determination of
whether this condition of wetting of the substrate by a proposed adhesive is
likely to be met. A strictIy experimental method deveIoped by Zisman and
co-workers (1968) at the Naval Research Laboratory is based on a value Ye,
defined as the critical surface tension of the surface. Ye is the surface tension
of the liquid which will just, and only just, wet the surface of the solido
In order to determine this value experimentalIy, use is made of the concepts developed by Young. The Young equation is the mathematical expression of the vector diagram of a drop (shown in Fig. 1) which only partialIy
wets a solid surface:
Ysv

(3)

YSL+ YLv cos (J

In Eq. (3), the angle (Jis the contact angle made by the liquid on the flat
solid surface. It is dependent on the surface tension of the liquido The lower
the surface tension, the greater the tendency to wet the solid, and the
smalIer the angle (J.Ye is the value obtained by extrapolating to zero contact
angle the results of contact angle measurements made with a series of
liquids having different surface tensions on tQe solid in question. The Naval
Research Laboratory group has actively pursued this approach for welI
over a decade and has recentIy published (Shafrin, 1977) Ye values covering
some 150 synthetic and naturalIy occurring polymers. The values for some
of the materials of interest to the emulsion polymers industry are shown in
Table 1.
A more theoretical approach to determining whether an adhesive can be
expected to wet the surface of a solid has been developed and promoted by
Girifalco and Good (1957) and Fowkes (1964). Utilizing the geometric mean

Fig.1.

Vector diagram

of a liquid drop which partially

wets a so lid surface.

Carlton G. Force

294
TABLEI
Critical Surface Tensions of Polymer Materials

Critical surface
tension (y c)

Polymer
Phenol/resocrinol
adhesive
Resorcinol adhesive
Wool
Cellulose (regenerated)
Nylon 2
Polyacrylonitrile
Casein
Nylon 7,7
Poly(ethylene oxide)
Poly(ethylene terephthalate)
Nylon 6,6
Nylon 11
Nylon 6
,
Cellulose, from cotton linters
Nylon
Poly(methyl acrylate)
Polyoxyphenylene
Polysulfone
Poly(vinylidene chloride)
Cellulose acetate
Poly(methyl methacrylate)
Poly(vinyl chloride)
Starch
Chloroprene
Polyisoprene, chlorinated
Poly(vinyl acetate)
Poly(vinyl alcohol)
Nylon 9,9
Poly(vinyl ethyl ether)
Rubber (natural)/rosin
adhesive
Cellulose, from wood pulp
Polyacrylamide
Polyacrylate
Polyepichlorohydrin
Nylon 8,8
Polyisoprene, cyclized
Poly(ethyl methacrylate)
Polystyrene
Nylon 10,10
cis- Poly-1,3-butadiene
Polyoxypropylene
trans- Poly-1 ,3,-butadiene

(dynecm-I)
52.0
51.0
45.0
44.0
44.0
44.0
43.0
43.0
43.0
43.0
42.5
'42.0
42.0
41.5
41.0
41.0
41.0
41.0
40.0
39.0
39.0
39.0
39.0
38.0
37.0
37.0
37.0
36.0
36.0
(36.0
35.5
3540
35.0
35.0
34.0
34.0
33.0
32.8
32.0
32.0
32.0
31.0
(continued)

8. Recent Developments and Trends in the Industrial Use of Latex

295

TABlE I (continuad)
Critical surface
tension (Yc)
(dyne cm -1)

Polymer

31.0
31.0
30.0
29.0
29.0
29.0
28.0
28.0
27.0
27.0
25.0
22.0

Polyethylene
cis- Polyisoprene
trans- Polyisoprene
Poly-4-tert- butylstyrene
Polyurethane
Poly(vinyl methyl ether)
Poly(butyl acrylate)
Poly(vinyl butyal)
Butyl rubber
Polyisobutene
Poly-l,2-butadiene
Silicone rubber
a

Adapted from Skeist (1977). Reprinted by permission of

Van Nostrand

Reinhold

Co.

to take into account the London dispersion forces between molecules across
the interface, the interfacial tension between two immiscible liquids, such as
mercury or water and a hydrocarbon, can be expressed by
Y12= Yl + Y2

2{r1Y1)1/2

(4)

Mercury itself is capable of interacting by two main interatomic forces, the


metallic bond and London dispersion forces. Similarly, water has the
potential for both hydrogen bond and London dispersion force interactions. However, hydrocarbons cannot interact with either the metallic bond,
in the case of mercury, or hydrogen bonds, in the case of water. Therefore,
the only primary interatomic force within hydrocarbons and across the
interface is due to the London dispersion interaction, and
Y2 = Y1

(5)

for the hydrocarbon. Various hydrocarbons have surface tensions covering


the range from 18.to 30 ergjcm2. From the interfacial tensions between
several of them and water, the London dispersion force field of water has
been found to be 21.8 :!: 0.7 ergjcm2. Since water, for example, only partially wets most polymeric materials, yd of the polymer can be calculated
from the contact angle water makes with it. Having established yd for the
solid, Fowkes (1964) has shown that not only can the strength of an
adhesive bond be predicted, but several other thermodynamic properties of

296

Carlton G. Force

the polymer can also be calculated, such as the heat and free energy of
adsorption. The van der Waals attractive constant A can also be calculated
from yd, as was done for S-BR latex mentioned earlier. These dispersion
force interactions are quite strong, being approximately equal in strength to
hydrogen bonds. In cases, where solid and liquid systems can interact only
by London dispersion forces, Fowkes has demonstrated that for wetting to
occur
yty~ > 2YL

(6)

From Eq. (6), it can be determined that y~ must exceed 243 dynejcm for
water to wet a solid surface.
From this, it can be seen why meticulously clean surfaces are usually
necessary for good adhesion. Although the contaminant may wet and thus
coat the substrate, the contaminant will often have sufficiently low surface
free energy so that the adhesive in turn cannot wet it. It should be noted,
however, that contaminants do not always lead to poor adhesion. In some
cases, a contaminant can improve adhesion, for example, by providing
improved wetting on one constituent and adhesion by another mechanism,
such as chemisorption or hydrogen bonding, to the other constituent.

B. Diffusion
This mechanism of adhesion advanced and originally promoted by the
Soviets (1963) is of the greatest importance to applications where a material
is being adhered to itself, such as in the self tack of rubber. Another familiar
example is the application of the same adhesive to two different surfaces
from a solvent system, allowing the solvent to dry partially or totally, and
then combining the treated surfaces, such as in the use of "contact cement"
or self-sellingenvelopes.
In the theory of diffusion it is postulated that high molecular weight
polymer molecules interdiffuse with each other across the interface. The
term "autohesion" is often applied to this process in the adhering of
portions of the same plastic material together. Since molecules of the same
material diffuse across the interface, making the two layers one, the original
joint disappears. Once the joint is completely healed, there is little chance
that adhesive failure will occur at the original interface.
For autohesion to occur, the polymer must be well above its glass
transition temperature. It is the primary mechanism operating in the heat
sealing of plastics and the coalescence of droplets from emulsions of the
film-forming plastics such as S-BR rubber during drying.

8. Recent Developments and Trends in the Industrial Use of Latex

297

Those who are familiar with the difficulties in formulating polymer


blends will quickly recognize, however, that care must be taken in trying to
apply this mechanism to the bonding of two different polymers. Very few
polymers exhibit good compatibility with each other. The concept of
solubility parameters (Hildebrand and Scott, 1950) is often utilized to help
predict the degree of compatibility to be expected. Voyutskii (1971) has
presented data that he interpreted as evidence that molecules of one
polymer, poly(vinyl chloride), can diffuse into another, poly(butyl methacrylate), having a solubility parameter nearly three units lower. In this case,
there was no indication that poly(butyl methacrylate) molecules had diffused back into the PVC. With poly(methyl methacrylate), whose solubility
parameter is within 0.8 units of PVC, an interfacial zone of nearly 4000 A
was shown by UV luminescence, where intermingling of the molecules
presumably existed. There is data relating peel strength to thermodynamic
compatibility, but further studies are necessary before quantitative predictions can be made on the contribution of diffusion to adhesion in
relation to polymer compatibility. The order of strength demonstrated
could also be contributed primarily by physical adsorption.
C.

Electrical Charge Distribution

Fowkes (1973) has recentIy provided an explanation for the electrostatic


theory of adhesion based on donor-acceptor relationships across the interface that seems quite plausible. This type of acid-base interaction is
responsible for the improved adhesion of carboxylated polymers to inorganic fillers, such as calcium carbonate, al~minum hydroxide, and other
familiar metal oxides' and hydroxides. The polymer surface is predominantIy negative in charge, whereas the metallic compound has a positively
charged surface. In more general terms, common polymer electron-donor
groups are oxygen atoms, amides, and double bonds. Chlorine atoms
in poly(vinyl chloride) and fluorine atoms in fluorinated polymers are
potential electron acceptors. In addition to the every present van der Waals
attractive force, carboxyl groups on rubber also pro vide a potential for
hydrogen bonding to strengthen adhesion to oxygen-containing substrates
such as cellulosic materials.
.
In inorganic systems, ionic bonds across interfaces must also be considered as electrostatic adhesion. In a number of instances, oxide or other
films are deliberately created on metal surfaces to promote adhesion with
other systems, such as polymers. The bond between the metal oxide and the
metal itself. is primarily electrostatic and is usually stronger than the
cohesive strength of the metal oxide.

298

Carlton G. Force

D. Chemisorption
Whereas the electrical charge distribution mechanism described in the
preceding section postulates a relatively nonspecific electrostatic attraction
or donor-acceptor interaction across the interface, chemisorption is considered to involve the formation of covalent or donor-acceptor bonds
between specific molecules on opposite sides of the interface. One of the
more popular current applications of this theory is the siloxane coupling
agents used for bonding glass or hydrated metal oxides to polymeric
adhesives. These coupling agents are provided with appropriately reactive
terminal groups on one of the organic linkages to the silicon atom. The
reactive group could be a double bond, for reaction with unsaturated
polyester resins dissolved in styrene; an amine or carboxyl group, for
reaction with epoxy resins; or a mercaptan, to interact in the vu1canization
reaction with rubber. Some materials which are blended with elastomers for
the purpose of reacting with the substrate, the elastomer, or both, to
produce chemisorptive adhesion, include isocyanates and resorcinolformaldehyde resins in the familiar RFL, or resorcinol, formaldehyde, vinyl
pyridine latexes used for tire cord dips. Patterson (1969) has written a good
survey of the technical and patent literature of fiber adhesion.
Chemisorption is strongly emphasized in this survey, although not to the
exclusion of other mechanisms.

E. Tack
Closely associated with the factors responsible for adhesion is the
tackiness of the polymer. An adhesive is said to possess tack if, under the
conditions of application, only light pressure is required to produce a bond
sufficiently strong to require work to restore the interfjlce to its original
separated state. This differs from the preceeding four factors in that the
objective in those four cases was to produce a sufficiently strong bond so
that failure would be cohesive in the substrate or adhesive, but not at the
original interface. Tack is generally measured by the energy required to
separate the bond after contact has been made, under carefully controlled
conditions, which should include pressure, time, and temperature. Two
primary factors are involved in such a measurement. First, the adhesive'
surface must readily wet the surface of the material with which it is brought
into contact in order to bond quickly with minimum pressure. Second, the
bulk of the adhesive should generally be capable of absorbing, through
deformation, a portion of the energy applied to separate the bond. Thus,
the viscoelastic properties of the adhesive are extremely important in

8.

Recent Developments and Trends in the Industrial Use of Latex

299

obtaining optimum strength without producing a bond that cannot be


separated when desired.
Certain materials, such as natural rubber, possess considerable tack
without modification. Many synthetic rubbers, such as S-BR, do not show
much tack in the dry state, although their latexes may be quite tacky while
water is still present during drying. Dry tack can be improved by blending
these polymers with other materials. The mechanism by which tackifiers,
such as rosin, terpenes and their esters, and phenolic and hydrocarbon
resins, operate is complex and obscure. However, they possess a sufficient
degree of incompatibility with the adhesive so that they migrate, to some
degree, to the air or substrate-adhesive interface. Here they may carry out a
number of functions, such as forcing low molecular weight polymer, which
is sometimes generated on or exuded to the surface, back into the bulk so it
will not produce weak bonds. The tackifier will also tend to separate
entangled polymer chains near the surface by virtue of its prescnce in the
area. Since there is a degree of incompatibility, some of these separated
chain heads may prefer to be at the air-tackifier interface for good bonding
while their tails remain firmly imbedded in the bulk polymer. AIso, the
effect the tackifier-adhesive combination has on wetting characteristics
toward the substrate cannot be ignored.
Materials generally considered to be tackifying resins for pressuresensitive applications have one characteristic in common, a softening point
well above room temperature. The glass transition temperature of the
PARTS RESIN PER HUNDRED PARTS RUBBER

;;-
ID
o
a:
CL

30

10

1400

50 70

200

100

400

900

1200
1000

<{

i5

800

'",
S

600

a:

UJ
CL

'"
'-'
:!

400
200

20

40

60

'Q

80

I
100

PERCENT TACKIFYING RESIN IN NATURAL RUBBER

Fig. 2.

Variation of taek with proportion of taekifyingresin in the elastomer (from

Skeist, 1977, eourtesy, Hereules,


Nostrand Reinhold Co.).

Ine., Wilmington,

Delaware.

Reprinted

by permission

of Van

Carlton G. Force

300

adhesive elastomer is considerably below room temperature. Apparently, in


some manner, this incongruity of properties synergizes tack benefits to
produce a large maximum in the work of separation at somewhere near
equal blends of resin and elastomer. A typical tack curve is depicted in Fig.
2.
As the region of maximum tack is reached, it has been demonstrated by a
variety of measurements that two phases actually exist on the surface. Such
a remarkable increase in strength, however, must be the result of a
combination of substrate bonding and viscoelastic changes in the bulk
polymer.

III.

Bonding Applications

The polymer particle size in latexes is far too small to be effective for
adhesive applications. Therefore, many of these very tiny building blocks
must coalesce through the diffusion adhesion mechanism to provide an
optimum sized critical mass of polymer for the particular adhesive requirement. Methods of bringing the crucial number of particles together at the
proper point of intersection for optimum bonding represents much of the
ingenuity and skill involved in the utilization of emulsion polymers in
adhesive applications.
Equally important is having the correct emulsion polymer for the
physical conditions available to create the bond. This involves such factors
as glass transition temperature of the specific polymer and molecular weight
of the polymer molecules. Too low a molecular weight do es not provide
sufficient chain length for adequate entanglement and diffusion across
particle boundaries. Too high a molecular weight makes the polymer so
viscous it cannot flow readily enough to pro vide maximum intimate contact
with the optimum amount of substrate during the time provided for
bonding under application conditions (Mallon et al., 1980). In emulsion
polymerization,molecular weight is generallycontrolled by the quantity of
chain-transfer agent (often an alkyl mercaptan) used in the recipe.

A.

Adhesives

The ever increasing emphasis on reducing air pollution and on health

considerations emphasized by OSHA has provided a strong impetus for


adhesive manufacturers to develop water-based emulsion polymer systems

to replace the organic solvent-based systems on which the industry was


fou!lded.This has been a difficultprocess.

8. Recent Developments and Trends in the Industrial Use of Latex

301

First, latex polymers had to be optimized for the specific adhesion


characteristics required in their primary areas of utility: pressure sensitive
and contact adhesives. Second, there had to be developed suitable aqueous
dispersions of auxiliary ingredients; most notable in this category are
tackifiers. Finally, there has been economic resistance on the part of those
who apply the adhesives because of the greater difficulty in evaporating
water compared to organic solvents. Thus, most adhesive manufacturers
have found it expedient to install solvent recovery equipment to meet air
quality standards.
The problems to be solved include minimization of adjuncts such as
surfactants in the latexes and auxiliary agents which interfere with adhesion. Application to water-sensitive substrates also presented a problem
now being solved by transferring a film of the adhesive to the substrate from
a continuous low-surface-energy belt such as mylar on which the latex
system was deposited and dried. Latex manufacturers are rising to these
challenges by designing additional properties into the polymer molecules
and developing systems which give both the desired adhesive characteristics
and a savings in application energy through higher solids and optimized
temperatures in suitably designed drying ovens (Mallon et al., 1980; Lee,
1980). The rapidly rising costs of energy and petroleum solvents is bound to
catalyze the utilization of emulsion polymers throughout the rapidly growing adhesives industry.
Because of the speed and ease of utilization, pressure-sensitive adhesives
have become an integral part of a multitude of items utilized in both home
and industry over the last generation. They can be designed to provide
either permanent or removable bonds and are applied to cloth, plastics,
aluminum, paper, and numerous other substrates for tapes, labels, wall
covering, packaging, and many other uses. Pressure-sensitive adhesives
require good tack and wetting properties for the substrate to which they
will be attached. Elasticity is also desirable, particularly for removables.
These properties are generally obtained from rubbery polymers such as
natural rubber and S-BR. For more demanding performance, such as light
stability and solvent resistance, acrylics, nitrile rubber, polychloroprene, and
other high-performance polymers may be required.
These so-called elastomers are characterized by low glass transition
temperatures and a coiled or folded molecular chain configuration capable
of great extension. Similar performance flexibility available from latex is
required in a broad variety of applications where adhesives are utilized in
manufacturing processes. Latexes are also finding wide application in foil
laminating for food and other types of packaging. In short, there are very
few items which a consumer purchases where adhesives have not been used
at some time in their manufacture and packaging. For example, ther~ are

Carlton G. Force

302

on the order of 10 different adhesive operations in producing a carton of


filter tip cigarettes.
Contact adhesives are another large market where rapid penetration is
being achieved by latexes. Most generally in this high technology application two types of material each, coated with a dried layer of the adhesive,
are brought together to form the bond. Often a requirement of the adhesive
is to allow some movement to permit redistribution of stresses in use due to
physical differences such as thermal expansion coefficients between the
materials being bonded. Examples where these types' of adhesives are
utilized include shoe manufacture, automotive product assembly, adhering
Formica to furniture, various construction applications, and do-it-yourself
home repair. The rapidly increasing cost of solvents as well as the emphasis
on environmental and health regulations are quickly phasing the solventbased contact adhesive systems previously used out of this market
worldwide.
. Government

regulations regarding flame spread. in construction ma-

terials have shaped another trend in latex adhesives. Suitably compounded


polychloroprene latex has Underwriter Laboratory approval for the bonding of fiberglass batting to its facing for home insulation. It is also used for
adhering insulation to the inside of heating and air conditioning ductwork
in commercial buildings.
Not all solvent adhesives will be replaced with latexes per se. For some
applications, hot melt adhesive systems are being developed. However, just
as ethylene-vinyl acetate copolymers produced by emulsion polymerization
are a major component of many hot melt systems, suitably designed
emulsion polymers to meet specific requirements will probably be present in
these systems, too.

B.

Fiher Bonding

Another area where emulsion polymers have found broad acceptance is


in the bonding of fibers. These fibers can all be of the same material, as in
nonwoven-type structures or of different materials, such as in the woven
and nonwoven composites utilized in tufted carpets. Optimized deposition
of the polmer at fiber intersections rather than overall fiber coating and in
some instances the copious use of inert fillers represent the art involved in
developing properties at minimum cost.
The earliest of these applications was natural rubber latex bonded paper.
This was patented in England in 1926 and commercially developed in the
United States before 1930 to produce products which effectively replaced
leather in applications such as shoe insoles and midsoles. Latex was

8. Recent Developments and Trends in the Industrial Use of Latex

303

incorporated into the fibrous stock before formation of the paper (beateradd) in this initial development. Arter World War n, synthetic rubber
latexes provided the smaIler partiele size, versatility, and broader range of
properties which permitted expansion of the industry into new methods of
rubber addition to pre-formed sheets (dry web and wet web saturation).
Applications also expanded to inelude replacement of eloth as weIl as
leather. The uniqueness of the products often developed their own special
applications, such as in asbestos and fiber reinforced gasketing. The area
has now expanded to inelude nonwoven synthetic fiber composites prepared from both water and air layed structures. Those bonded by latexes
are primarily dry web saturated in one manner or another.
There has been considerable development over the last few years in
methods for saturating these nonwoven structures. Additional new methods
would be expected to emerge periodicaIly to provide the physical and
aesthetic properties required to broaden the rapid expansion of this industry into different applications characteristic of its growth. Among the
saturating techniques presently used for synthetic fiber nonwovens are print
bonding (the desired amount of latex is transferred by a patterned gravure
roIl of a two roIl nip to the web), saturation bonding (the synthetic web,
often supported, is saturated in latex bath and excess squeezed out in
passing through a two-roIl nip), impregnation bonding (the desired amount
of latex is transferred to the unsupported web by the lower roIl of a two-roIl
nip the bottom of which, is rotating in a latex bath), foam bonding (the
foamed latex is metered onto the fibrous web with a knife coater and sucked
into the structure by passing the web ov.er a vacuum box), and spray
bonding (the latex is sprayed onto a high bulk to weight ratio web for end
uses such as fiber-fiIl'insulation while a vacuum box below the web control s
penetration) (Brooks, 1979).
The synthetic fiber nonwoven industry developed around air lay fiber
substrates. The hydrophobic character of synthetic fibers in water makes
them difficult to disperse satisfactorily to produce suitable substrate using
the more rapid water lay papermaking technology. However, this has been
overcome by developing synthetic fibers having tiny hairlike protrusions
(fibrils) extending, verticaIly from their surfaces in water. These fibrils keep
the fibers physicaIly separated from each other and in good dispersion for
substrate formation on papermaking equipment. In order to provide adequate strength in the substrate for drying on paper machine drying drums
or other unsupported drying systems, it is often desirable to incorporate
some latex prior to sheet formation to assist the fibrils in fiber bonding until
the sheet is finaIly saturated.
The primar"y advantage imparted to fibrous structures by latex bonding
is toughness in the form of very good hand tear strength along a smooth cut

304

Carlton G. Force

edge. This is provided by low viscosity chewing gum-like polymers which


stretch to distribute tear stresses over many fiber rubber junetions. Sueh
soft polymers are also useful in providing the drape or "hand" eharaeteristies of the leather and c10th substrates often imitated by latex bonded fiber
products. Other strength properties (e.g., tensile) whieh are dependent on
firmer bonding are sometimes poorer in latex bonded sheets than ordinary
kraft paper (Sweeney, 1958). Less dramatie but often as useful eharaeteristies provided by. latex bonding for specifie applieations are inereased
elongation, better internal bond, improved eompress"on reeovery, solvent
resistanee, wet abrasion resistanee and folding or flex enduranee (Heiser et
al., 1962).
Mueh teehnology has developed around methods for produeing efficient
latex bonding of fibers. Only that latex deposited within or around a fiber
junetion to pro vide a sufficient volume of rubber to allow adequate
extension before breaking eontributes to optimum efficiency. Therefore, in
addition to the requirements of good wetting and adhesion to the fibers of
the substrate, the rubber needs the proper viseoelastic charaeteristics and
the latex must coagulate andlor eoalesce to an adequate partic1e size. In
addition, its efficieney is governed by how much of the total rubber actually
deposits within or surrounding fiber junctions. In the beater addition of
rubber, the latex is coagulated in a eontrolled manner by some means such
as addition of multivalent counterions, ehange in pH, or reduction in
colloidal charge by other methods to produce rubber partic1es of a charaeteristie and optimized size for deposition in the forming sheet. Ir the charge
on the partic1es is of the same sign as that of the fibers, the rubber will be
deposited within the fiber strueture primarily by filtration. Bonding efficiency is improved by both wet and dry pressing of the sheet providing the
rubber possesses sufficient tack to bond to the fibers that are pressed into
contaet with it. The Dow Chemical Company (Heiser et al., 1962) and the
Rohm and Haas Company (Sweeney, 1958) have reported on the development of experimental latexes with positively charged functional groups
bound into their polymer structure. Such cationic latexes are attracted to
the normally anionic charge sites on eellulose and synthetic fibers in
aqueous medium to create an eleetrostatic bond. Again, it is necessary for
an optimum number of latex partic1es to eoagulate at the bonding sites to
develop satisfactory end use properties. Work has recently been published
(Alinee et al., 1978) demonstrating various methods whereby cationic latex
deposits on eellulose fiber furnish. This type of e1ectrostatie deposition
should pro vide greatly improved rubber bonding efficiency because of the
e1ectrical attraction between deposited rubber and neighboring fibers. To
aehieve such effieieney, methods must be developed for agglomerating
cationie latexes to optimal bonding partic1e size before introduction into the

8. Recent Developments and Trends in the Industrial Use of Latex

305

anionic fiber furnish. Dow points out that the many previous attempts to
utilize cationic latexes in fiber furnishes ha ve shown no pronounced advantage because the latexes were stabilized with migratable monomeric cationic
emulsifiers (Heiser et al., 1962; Wessling, 1979). In aqueous solution,
sufficient of these emulsifier molecules quickly migrate to the negative
charge sites on the fibers making them neutral before the bulkier rubber
particles can travel and attach to these sites. In this same vein, care must be
taken not to add materials such as alum to the fiber furnish which will
make it cationic (Goring and Mason, 1950; Goring et al., 1950) before
addition of the cationic latex.
In latex saturation processes of preformed fiber webs, the predominant
force responsible for drawing the rubber particles to fiber crossings and
coagulating these particles into a coherent mass is the surface tension of the
evaporating water medium. Current trends are to develop latexes whose
surface tension is close to water by utilizing oligomeric emulsifiers (White,
1976) or emulsifiers polymerized into the polymer chain (Sweeney, 1958) to
maximize the surface tension forces.
This industry has remained highly technological, producing on relatively
slow equipment and employing a disproportionately large technical staff
compared to other areas of papermaking primarily because of cost considerations involved in the relatively inefficient utilization of rubber. With
the possible exception of optimized systems in which the latex is of opposite
charge to the fiber at the time of addition, paper products with suitable tear
strength ordinarily contain from 20 to 40% rubber. Inefficiency in anionic
latex beater-add sheet s is due to the frac.tion of filtered rubber particles
attached to only one fiber. In saturated, preformed webs most fiber
junctions are encapsulated with rubber, whereas much less would be
required if the rubber were located only between the fiber crossings. Again
considerable rubber is deposited on only one fiber.
Fiber bonding is a market strongly dependent on latex properties. The
tear strength is dependent on the viscoelastic properties of the rubber.
Adhesive strength of the rubber to the fiber must only be greater than the
tensile strength of the rubber itself. Thus, the low viscosity rubber will string
out and break near the center between two fibers before the attachment to
either fiber separates. This may welI be why carboxylated latexes show little
or no advantage in properties over uncarboxylated versions in these
products. The additional bond strength provided by carboxylation in many
applications is simply not required in fiber bonding. Rubber cure can also
be undesirable in some latexes in that it shortens the stretchability of the
rubber. Any degree of sulfur cure of S-BR bonded fibers drasticalIy reduces
tear strength of the product. However, slight sulfur cure of a natural rubberbonded product is desirable for improving the rubber oxidative stability

Carlton G. Force

306

without harming tear strength. Cure of other polymers such as the metal
oxide-catalyzed cure of neoprene or the heat cure of acrylics is desirable to
improve properties such as solvent resistance and often develops the
optimum degree of stretchability for tear strength (Sweeney, 1958).
A myriad of opportunities exist for the latex producer to design specific
properties into new latexes that will allow the bonded fiber manufacturers
to develop new products. These can be based on polymer properties such as
solvent resistance,' flexibility characteristics, and stre~gth. New developments can also be based on the colloidal characteristics 'of the latex. The
cationic latexes of Dow and Rohm and Haas are examples. Another
opportunity is in the development of methods for reducing the tendency of
latex particles to migrate with the evaporating water from the center of
saturated products to the hotter surfaces during drying. Unique colloid
stability characteristics might also permit the development of new beater
addition and saturation techniques for improved latex efficiency and
properties.
C.

Tufted Carpet Applications

Another area of fiber bonding where latex is the key ingredient is in the
m~nufacture of tufted carpets. In this process, the yarn tufts making up the
surface of the carpet are needle punched through a fabric known as the
primary backing. This was originally jute fabric (burlap). However, it has
been replaced by nylon nonwoven fabric which in turn is currently being
displaced in many grades of carpeting by a coarse woven ribbon-like
polypropylene fabrico Another backing that will sit on the underlay pad or
floor is then attached to this basic carpet structure. This secondary backing
is often fabric (Klein, 1979). Jute, the secondary fabric workhorse until the
fall Qf 1979, is presently being replaced by woven polypropylene and spunbonded polypropylene because of economics and the uncertainty of jute
supply from the Far East. Roughly one-fifth of tufted carpeting in the
United States and almost all tufted carpeting in Europe has, as the
secondary backing, attached foam rubber made from latex.
A latex adhesive is necessary to bond the tufts into the primary backing,
adhere the individual fiber strands in the yarn tufts together so they don't
separate and "pill" at the carpet surface, and attach the secondary backing
to this primary structure. This is accomplished almost exclusively by
carboxylated styrene-butadiene latex containing 400 or more parts per
hundred parts of rubber of a mineral filler. In addition to economics, this
filler helps pro vide density and stiffness to the carpet structure for the
desired "hand." Feldspar used to be the filler of choice, but less expensive
calcium carbonate has replaced it in most instances.

8. Recent Developments and Trends in the Industrrnl Use of Latex

307

It is desirable to have a minimum of water to evaporate in carpet


manufacture for the obvious energy savings. Therefore, the dry filler makes
it possible to compound a 50% solids latex to about 85% solids. In current
practice for viscosity control purposes, it is necessary to reduce this to
about 82% solids with a small amount of water. Current trends aim to
eliminate the need for this dilution.
In this application, it is necessary for the adhesive to pro vide a quick,
strong bond to both the carpet components and the mineral filler.
Carboxylated latex is desirable because of the enhanced adhesion provided,
particularIy to the filler, by enactment of the e1ectrical charge distribution
mechanism.
Permanent grab earIy in the adhesive drying cyc1e is necessary when
applying a fabric secondary backing because any place it separates from the
primary carpet structure during drying will not readhere with adequate
strength unless additional adhesive is applied. This requires hand slitting of
the fabric in the defective area, application of the adhesive and sale of the
carpet as a "second."
In order to achieve this so-called green tack, the rubber partic1es must
coagulate into a large enough mass to pro vide adhesion between components as soon as possible after drying begins. This coagulation is brought
about by the surface tension forces in the reducing volume created by
evaporation of the water. The more rapidly the water evaporates, the more
quickly the partic1es are brought into intimate contact by the surface
tension of the aqueous phase. The poorer the stability of the adhesive
system the lower the solids at which the desired degree of coagulation will
occur.
Many factors are'involved in providing the optimum degree of stability
to the adhesive system. The carboxylated latexes generally have a sufficiently low concentration of monomeric emulsifier so that their surface
tension is no lower than the mid-forties in dynes per centimeter (White,
1976; Klein, 1979) (pure water has a surface tension of 72 dynejcm). The
filler competes with the latex rubber partic1e surface for this monomeric
emulsifier. Therefore, the more finely ground the filler, the more surface
area it has available to adsorb emulsifier. However, fineness of the grind can
also affect other properties of the adhesive such as its viscosity characteristics (Klein, 1979). Finally, a development that has moved rapidly through
the industry recently is frothing the adhesive to volumes of from half again
to more than twice its unfrothed volume just before application. This
produces at least three effects. Each bubble wall containing the solids has
two large interfaces with air (inside and outside) which strongly compete
with the rubber partic1es and filler for emulsifier, thus destabilizing the
latex. Second, this large surface area provided by the bubbles produces

308

Carlton G. Force

more rapid remo val of the water. Third, the thinner films of rubber within
the bubble walls assure thorough coverage of the two substrate surfaces
with a lower quantity of adhesive if other characteristics such as bond are
adequate. Adhesive frothing, which has now expanded to over 90% of
United States industry, has reduced green tack adhesion problems considerably allowing acceptable properties to be achieved without the excess
of adhesive often applied to insure good early bonding. Frothed latex
application has been important in the drapery and furniture fabric industries for at least a decade to provide strength, the adhesive for fIock
coating and desired "hand" properties to the fabrics utilized.
Westvaco has recently introduced a water-dispersed tackifier to the
carpet industry which improves both green tack and final adhesion. Since
these properties are currentIy adequate, it is possible to reduce the amount
of rubber in the adhesive compound by substituting this tackifier and still
maintain the required properties in the carpe1. Good success is achieved
replacing up to 15% of the rubber in fabric backed carpets and 20% of the
rubber in the precoat upon which latex foam backing is to be applied.
Because of the difficult wetting characteristics of polypropylene, opportunities exist to provide latexes and ingredients that stimulate better adherence between substrates made from this polymer.
The current no-gel process of applying foam rubber to carpet tuft bonded
with a precoat that might contain 700 or more parts of mineral filler per
hundred parts of rubber became predominant in 1975. Prior to that, and
still used for applications such as waffie-designed carpet backing and
underlay, the more technically demanding gel system was utilized. Gel foam
is compounded with ammonium acetate which is decomposed to acetic acid
and ammonia gas by infrared heat just beyond the foam coat metering bar.
This acidifies the system sufficientIy so that zinc ions present enter a pH
region where they are hydrolyzed to a + 3 species which then electrolytically coagulates the rubber (Force, 1971). In the no-gel system, a frothing
agent (tallow amine sulfosuccinamate) produces a foam of sufficient stability
that the water can be dried from it without foam collapse. This simpler
technique has permitted the carpet milis to reduce costs by compounding
their own products for foam backing instead of relying on technical experts
who supplied the gel formulation. However, they have accepted some
disadvantages in doing this. One is a reduction in coating speed to permit
sufficient drying before the first turn roll in multipass ovens so the coating
will not fall off the carpe1. The second is greater carpet drying difficulty after
water spills in use because of the more hydrophilic character of the
surfactant used in the no-gelsystem. Finally, tallow amine sulfosuccinamate
does not produce as firm a rubber foam structure as the gel system. This
softer resiliance provides a less cushiony feel under foo1. In spite of these

8. Recent Developments and Trends in the Industrial Use of Latex

-309

deficiencies, economics dictate that no process much more costly than the
no-gel system can gain prominence in this industry. However, the deficiencies also provide opportunities for latex producers and chemical suppliers
to develop new products and new processes for this application.
The predominant latex used for backing foam carpet is S-BR made by
the cold redox-initiated system. Smaller amounts of specialty latexes such as
poly(vinyl chloride) and polychloroprene are sometimes used for meeting
specific flame spread requirements where chlorine-containing polymers
compounded with suitable catalysts such as antimony salts pro vide distinct
advantages. A fraction of hydrated alumina in the filler portion of both the
adhesive and foam coating allow S-BR to meet government regulations for
flame spread for most applications. Urethane foam is making small inroads
into this market. A number of suppliers are currently developing urethane
foam systems. The cost and availability of natural gas normally used for
carpet drying will probably be the factor governing how quickly this
development matures. Materials that permit compounding of higher solids
latex foam systems should help deter penetration of urethane into this
industry. Systems that permit application of foam backing of greater
thickness than the roughly i in. limit currently practical could also expand
the United States market beyond the one-fifth of carpeting which it now
enjoys and utilize more latex per carpet.
D.

Paints and Industrial Coatings

Another large-scale application of latexes as adhesives is in the bonding


to continuous substrates of pigments, clays, ~mdmineral particles in coatings
that produce specific end use advantages. Major among these are paints
and paper coatings. There are also many functional coatings applied to a
variety of substrates to pro vide solvent resistance, to retard vapor and
liquid permeability, or pro vide a host of other benefits.
Following World War n, synthetic latexes made possible the development of a new concept in coating technology with the introduction of
water-based paints. Their first use was interior home walls where hiding
and color were qf primary importance. Such paints are formulated to cure
over a period of time to provide wa.shability and scrubability. A particularly
challenging problem has been the development of suitable viscosity characteristics to achieve good leveling to smooth out brush marks while maintaining sufficient body so the paint do es not splatter easily under application conditions or "run" on the surface. This has been achieved by
additives that pro vide optimized thixatropy to the paints so they thin
sufficiently under the shear conditions of roller or brush application for
quick leveling but remain sufficiently thick before and soon after application

310

Carlton G. Force

to prevent splattering and running. The capacity of latex polymer to adhere


quickly in the wet state to the substrate and to coalesce into a continuous
film during drying in order to maintain the hiding and color pigments welI
dispersed over the surface makes synthetic latexes welI suited for paint
purposes. From their beginning with S-BR latex paints on interior walIs,
continual performance improvement through additive and polymer development has expanded the markets to gloss woodwork coatings, durable
chalking and mildew-resistant exterior coatings, and currently to a degree
of penetration into the more demanding industrial coatings market. In
industrial and maintenance coatings, chemical resistance is of primary
concern in many applications. It is also necessary that industrial coatings
form a continuous film and cure to acceptable rust and chemical resistance
quickly after application. Mechanical and flexibility characteristics of the
film that pro vide durability against cracking are also of primary importance
in many industrial uses (Bierwagen, 1979).
Storage stability is a special problem with paints., The pressures de- .
veloped by the formation of ice crystals quickly coagulate latex particles.
Therefore, additives must be incorporated such as specific nonionic type
emulsifiers to pro vide good freeze-thaw stability. Maintaining adequate
dispersion of the pigments and other ingredients over long storage periods
so they can be easily reblended uniformly at their original particle size is
also a necessity. Specific opportunities always available for improvement by
unique formulating ingredients and advanced latexes are water sensitivity of
the film, viscosity stability and shear stability of the paint in storage, and in
general the whole realm of application characteristics involving leveling,
flow, and viscosity. As these are advanced for specific applications, latexbased products will gain larger shares of industrial and maintenance
coatings markets. Increasing cost of solvents and governmental restrictions
on their use pro vide a strong incentive for market penetration as quickly as
suitable properties are available.

E.

Paper Coatings

Paper and paperboard for printing are coated to improve both printability and appearance. The coating consists primarily of mineral pigments
bound to the celIulose surface with synthetic and natural polymers. A
variety of additives may also be incorporated into the coating such as
lubricants, antifoams, pigment dispersants, flow modifiers, and dyes. The
binder level is generalIy in the range of lOto 20% of the quantity of pigment
and is composed of a combination of synthetic latex and a natural polymer
such as starch, protein, or casein. In addition to bonding the pigment to the
substrate, each binder pro vides other desirable properties to the coating.

8. Recent Developments and Trends in the Industrial Use of Latex

311

For example, carboxylated S-BR helps produce gloss in the coating and seal
the surface for good ink hold out. In offset printing both water and inks are
applied to the paper. Therefore, the coating must be hydrophilic and porous
enough for the paper to absorb the water during drying. Starch for
hydrophilicity and poly(vinyl acetate) which pro vides strong pigment bonding but is nonfilm forming to maintain porosity is the binder combination
of preference for coating offset printing paper. For gravure printing, which
is currently gaining in popularity, the ink used is of sufficiently low viscosity
that surface bond strength of the coating is of little importance whereas
surface smoothness is of greater significance. Thus, film-forming synthetic
latex is often the exclusive binder (Heiser, 1979).
Pigment and binder help fill in voids between the surface fibers. The
pigments main purposes, however, are to reflect more light back to the eye
for improved brightness of the surface and to help scatter light passing
through the sheet for better opacity to prevent show-through of the
printing. Because of its high refractive index, titanium dioxide is an
extremely effective light scatterer for opacity improvement. Small particle
size clays, calcium carbonate, and hydrated alumina make good brightnessimproving pigments. Dow has introduced some synthetic polymer pigments
in the form of latex-like water dispersions. Just as polystyrene latex binder
pro vides high gloss in supercalendered paper, these 0.4 to 0.5 Jlm pigment
particles provide gloss, optimized scattering for opacity due to the specific
particle size, good brightness, and ink holdout. Economics will be important in determining how much penetration plastic pigments make into this
market.
Carboxylated latexes made with a small fraction of polymerizable fatty
acid in the monom'er mix pro vide many advantages as coating binders.
These improvements include better adhesion to both the mineral pigment
and cellulose through the bound carboxyls enhancement of the electrical
charge distribution mechanism and hydrogen bonding mechanism, respectively. This provides advantages in gluability and wet rub. Other
advantages include improved compatibility with starch and better coating
blade performance.
Among current trends is the compounding of paper coatings to higher
solids for faster drying at lower energy requirements. Future opportunities
in which the latex manufacturer can participate include providing better ink
holdout and more uniform ink adsorption, enhancing easier finishing (or
even elimination of the need for finishing), providing specific latexes for low,
medium, or high gloss coatings, enhancing faster ink drying, providing
improved basestock reinforcing and better coverage at faster coating speeds,
and improving coating blade runability (Heiser, 1979). These challenges will
be met through both polymer design and colloid chemistry of the latex and
coating systems.

Carlton G. Force

312

IV.

ConstructionApplications

A relatively new opportunity for latexes is in providing special benefits to


various types of pavement. Probably the fastest growing market worIdwide
in this respect is coatings for recreational and athletic surfaces. Here, a top
coat consisting of ground rubber scrap is bonded with latex to make a i-in.
surface over an asphalt or concrete subsurface. Its resiliency, durability, and
elasticity pro vide a. maintenance-free playing surface. Such surfaces have
been medically demonstrated to create fewer imfirmities to joints and other
parts of the body jolted in athletic endeavor, to help prevent playground
injuries, and to produce a lower degree of fatigue than the subsurfaces alone
or most other surfaces. These surfaces have required no repair in all types of
weather over the ll-year history since they were first introduced in 1969.
Thus, the additional cost of application is now recognized to be more than
compensated for by their minimal maintenance.
Small amounts of S-BR or the more oxidation-resistant polychloroprene
latex ih asphaltic concretes are beneficial in bonding crushed aggregate
firmly while leaving a sufficieritly open structure around the aggregate
particle for rain water to penetrate below the pavement surface. This
effective method for eliminating hydroplaning is being utilized for airport
runway construction in some of the western United States. However,
because of the additional cost imparted by the latex,"broaq penetration into
airport construction and carryover to general road construction has not yet
developed. Currently, in the more dangerous stretches, grooves are
mechanically impressed into pavement surfaces to help pro vide the drainage
required to overcome the hydroplaning problem.
Another paving system in which latexes provide unique properties is
alumina hydraulic cement. Here, 10 to 15% latex (based on the cement
fraction) produces a flex modulus which greatly reduces fatigue for persons
who work on concrete floors. Latexes such as polychloroprene also help
produce chemical resistance and nonporous structure in concrete. As a
result, latex is used in Terezo-type flooring in shower stalls and for
hospitals and dairies where sterilization is important. Along this same line,
latex in concrete pavement prevents fizzures from developing due to the
expansion of absorbed water upon freezing. In climates where icing is
common in winter, concrete bridges are particularIy vulnerable. Water
rapidly penetrates the concrete sufficiently to rust the reinforcing rods
requiring costly bridge repair in a relatively short periodo This problem is
especially evident through the northern half of the United States on the
interstate road system .which is now about 20 years old. Unfortunately,
latex is still not sufficiently cost effective to be utilized in much of this
first round of bridge surface replacement. However, it is finding appli-

8. Recent Developments and Trends in the Industrial Use of Latex

313

cation in parking garages, recreational are as, and industrial areas where
the resiliency property is also of prime importance.
A large application of latex is in blends with asphalt to produce a
waterproof membrane on cinder block and concrete walls. Here, as well as
when incorporated into concrete mortars, the rubber fraction produces
better adhesion, durability, and temperature expansion tolerance than the
asphalt or mortar base alone.
With the decreasing availability and increasing price of petroleum,
methods are being implemented to convert more of it to higher value
distilled fractions. As a result, the residual asphalt fraction, is changing in
properties. There are also vast differences in asphalt properties depending
upon the section of the world from which the crude oil is obtained. An
active program is just beginning among petroleum refiners to investigate
the potential for achieving and retaining optimum properties in asphalt
through the incorporation of a fraction of latex in their products.

v.

RubberGoods

A large proportion of solid polymers are made by emulsion polymerization covering the entire range from tire rubber to high technology
thermoplastic resins such as ABS. However, these solid polymers and their
production is too vast a subject for a single chapter. Thus, this section will
be confined to a couple of product applications whose production technology is dependent on the colloid chemistry of latexes, in addition to being
a convenient medium for production of the .rubber.
Elastic thread used in many garments such as pajama waists and sock
tops is made either by cutting from solid rubber or extruding from latex.
Cut thread has the disadvantage in some applications of having square
edges. Extruded thread is produced by running a continuous stream of latex
into a coagulation bath through a circular hole that pro vides cylindrical
thread of the proper diameter. For natural rubber latex, the coagulant is
usually acetic acid whose density is adjusted to allow the thread to remain
suspended whilt: being drawn through the bath. The rubber thread thus
formed is then drawn through a leach bath to remo ve salts and passed
through an oyen for drying and curing of the rubber. This method of elastic
thread production is less costly than cut thread. However, the dry heat
resistance, particularly under tension, of this extruded thread is not as good
as cut thread. Development of better rubber cure technology in the
extruded thread process should allow latex to take over the majority of this
market. Synthetic latexes can provide special properties to the thread (e.g.,
polychloroprene can pro vide good solvent resistance under stress). In such

314

Carlton G. Force

cases,the coagulation bath density must be balanced to the density of the


polymer.
The second area utilizing latexes for production is dipped goods products. These include articles such as toy balloons, rubber gloves, finger cots,
and automotive floor gear shift rubber boots dipped onto a metal support.
The form on which the article is to be produced can be coated by straight
dip where the latex is simply allowed to dry or a coagulation dip as dictated
by processing and application requirements including e!!-seof removal from
the formo In coagulation dipping, the form is often first dipped in a
coagulant bath of 20 to 35% calcium nitrate dissolved in water, then into
the latex bath and sometimes back into a coagulant bath to set the outer
surface. Dipped goods articles can be either unsupported or supported such
as the gear shift boot or a rubber coated textile glove, respectively. Often
such a textile article. first is straight dipped followed by one or more
coagulation dips. After deposition of the proper thickness of rubber from a
latex bath of suitable viscosity to maintain uniform films for the item being
manufactured, the articles are leached to remove residual salts, dried, and
cured in appropriate apparatus. .
The latexes upon which this industry developed were natural rubber and
polychloroprene for solvent resistance. However, technology is advancing
to permit penetration of carboxylated nitrile latex for optimized solvent
resistance and tougher abrasion resistance. Among the competition to
latexes in this field are poly(vinyl chloride) plastisols. As technology develops in producing small particle size latexes from polymers whose synthesis is too water-sensitive for emulsion polymerization, the dipped goods
industry will quickly convert to their utilization from the solvent-based
cements of these polymers now employed. Prime candidates include butyl
rubber, EPDM, hypalon, and viton.

VI.

Properties of Various Latexes

In Table 11,common characteristics of a variety of polymer latexes are


listed. Throughout the text, the more important requirements for each
application were alluded to. Initial work on a specific problem should
quickly establish the properties needed in both the polymer and its latex.
The table is organized to permit easy determination of the capability of
different classes of latexes to satisfy property requirements. Latex producer's
specific product literature should be consulted to determine which of the
many products within a polymer class would be most likely to satisfy the
desired application. Although this book is devoted to emulsion polymerization, natural rubber latex has been included in this chapter for two

TABLE

Application

Natural rubber
Styrene-butadiene
Nitrile
Polychloroprene
Poly(vinyl chloride)
Poly(vinyl acetate)
Acrylic resin
Adapted

Properties of Commonly

Used Latexes8,b

Glass
Typical
transition
Uncured
Organic
Price
cohesive
Polymer Oxidative Color
solvent
Water
temp
Compounding Specific Gravity
(0C)
requirements latex solids ranking
strength Resiliency tack
stability stability resistance resistance

Latex

11

E
F
S
E
G
G
G

E
S
E
E
G
F
F

from Industrial Adhesives

E
F
P-G
P-G
P-G'
F

F
F
G
G-Ed
G-E
E
E

F
F
F
F
. S
E
E

P
P
E
S-Ed

G-S
F-G
F-G

edited by G. W. Koehn. Courtesy,

E
E
E
E
E
F
G-S

Armstrong

-67
-48
-25
-40
78'
30
-55 to 451

Extensive
Extensive
Average
Extensive
Minimal
Minimal
Minimal

Cork Co., Lancaster,

0.95
1.00
1.00
1.12
1.17
1.09
1.06

Pennsylvania.

b Code: E, excellent; S, superior; G, good; F, fair, P, poor.


, Partial hydrolysis to poly(vinyl alcohol) produces good tack.

.
d

Cured systems.

Copolymerized
with acrylate or vinylidene chloride
1 Depending on acrylate monomers utilized.

monomers;

glass transition

temperature

may be as low as -25C.

Major
markets
(by topic
outline)

0.91
High B-I,3,D
0.99 Moderate B-l,2,3,4,5,C
1.00
High B-l,2,3,5,D
1.23
High B-l,2,3,C,D
1.41
Low B-2,3,4,5,C,D
1.18
Low B-l,2,4,5,C
1.10 Moderate B-l,2,3,4,C

316

reasons. Many of
latex and natural
uncured cohesive
synthetic emulsion

Carlton G. Force
the industries were developed utilizing natural rubber
rubber possesses a few important properties such as
strength and tack that have yet to be matched by
polymers.

Acknowledgment
The author wishes to express his appreciation to the following persons who graciously
provided information on the subjects dealt with in this chapter: William de Vry and David
Weaver of B. F. Goodrich Chemical Company; B. R. Vijayendran of Celanese Polymer
Specialities Company; Evans Townsend of the Norton Company; and J. J. Gorman, C. H.
Gilbert, and J. C. Fitch of E. I. duPont de Nemours and Co. Appreciation is also expressed to
Westvaco Corporation for permission to publish this chapter and to numerous colleagues at
Westvaco for their suggestions and assistance.

References
Alinee, B., Robertson, A. A., and Inoue, M. (1978). J. Co/loid Interface Sci. 65, 98.
Bierwagen, G. P. (1979). Adv. Emul. Polym. Latex Technol., Ann. Short Course Papers, 10th,
Vol. n. Lehigh Univ., Bethlehem, Pennsylvania.
Briscoe, B. J., and Tabor, D. (1972). Faraday Special Discussion, "Solid/Solid Interfaces,"
Chem. Soco No. 2, p. 7. Chemical Society, London.
Brooks, B. A. (1979). Adv. Emul: Polym. Latex Technol., Ann. Short Course Papers. 10th, Vol. n.
Lehigh Univ., Bethlehem, Pennsylvania.
Force, C. G. (1971). Paper presented at American Chemical Society, Rubber Div., Southern
Rubber Group Meeting, Memphis, Tennessee, October.
Force, C. G., and Matijevi, E. (1968). Ko/loid Z. Z. Polym. 224, 51-62.
Fowkes, F. M. (1964). 1nd. Eng. Chem. 56, (12) 40-52.
Fowkes, F. M. (1967). "Surface and Interfaces" (J. J. Burke, ed.), Vol. 1,p. 199. Syracuse Univ.
Press, Syracuse, New York.
Fowkes, F. M. (1973). "Recent Advances in Adhesion" (L. H. Lee, ed.), pp. 39-43, Gordon and
Breach, New York.
Girifalco, L. A., and Good, R. J. (1957). J. Phys. Chem. 61, 904.
Goring, D. A. l., Biefer, G. J., and Mason, S. G. (1950). Can. J. Res. 288, 339-344.
Goring, D. A. l., and Mason, S. G. (l950a). Can. J. Res. 288, 307-322.
Goring, D. A. l., and Mason, S. G. (l950b). Can. J. Res. 288,.323-338.
Green, B. W., and Saunders, F. L (1970). J. Co/loid Inte~race Sci. 33, 393-404.
Grimley, H. (1973). "Aspects of Adhesion" (D. H. Alner and K. W. Allen, eds.), Vol. 7, p. 11.
Transcripta Books, London.
Hamaker, A. C. (1937). Physica 4, 1058.
Harkins, W. D. (1941). J. Chem. Phys. 9, 552.
Heiser, E. J. (1979). Adv. Emul. Polym. Latex Technol., Ann. Short Course Papers, 10th, Vol. n.
Lehigh Univ., Bethlehem, Pennsylvania.
Heiser, E. J., Morgan, R. W., and Reder, A. S. (1962). TAPPI 45,588.
Hildebrand, J. H., and Scott, R. L. (1950). "Solubility of Non-Electrolytes," 3rd ed. Van
Nostrand-Reinhold, Princeton, New Jersey.

8. Recent Developments and Trends in the Industrial Use of Latex

317

Klein, A. (1979). Adv. Emul. Polym. Latex Technol., Ann. Short Course Papers, 10th, Vo!. 11.
Lehigh Univ., Bethlehem, Pennsylvania.
Lee, Y. S. (1980). Polym. Con! Polym. Improved Energy Efficiency, 11th, Akron, Ohio.
Mallon, C. B., Hagan, J. W., and Hamer, A. D. (1980). Polym. Con[ Polym. Improved Energy
Efficiency, 11th, Akron.
Ottewill, R. H., and Shaw, J. N. (1966). Discuss. Faraday Soco 42, 154.
Patterson, H. T. (1969). "Treatise on Adhesion and Adhesives," (R. L. Patrick, ed.), Vol. 2,
Chapter 6. Dekker, New York.
Shafrin, E. G. (1977). "Handbook of Adhesives" (l. Skeist, ed.), 2nd ed., pp. 67-71. Van
Nostrand-Reinhold, Princeton, New Jersey.
Shah, D. O., andChan, K. S. (1980). J. Dispersion Sci. Technol. 1, 55.
Skiest, 1. (ed.) (1977). "Handbook of Adhesives," 2nd ed. Van Nostrand-Reinhold (Litton
Educational Pub!.), Princeton, New Jersey.
Sweeney, E. J. (1958). TAPPI 41,304.
Vanderhoff, J. W., EI-Asser, M. S., and Ugelstad, J. (1979). U.S. Patent 4,177,177, December 4.
Visser, J. (1972). Adv. Colloid Interface Sci. 3, 331.
von Bramer, P. T., and McGillen, W. D. (1966). Soap Chem. Spec. 42(12), 123.
Voyutskii, S. S. (1963). "Autohesion and Adhesion of High Polymers" (S. Kaganoff,
Translator), (lnterscience), Wiley, New York.
Voyutskii, S. S. (1971). J. Adhes. 3, 69.
Wake, W. C. (1978). Polymer 19, 29.
Watillon, A., and Joseph-Petit, A. M. (1963). Am. Chem. Soco Div. Water and Waste Preprints
61-65.
Wessling, R. A. (1979). Adv. Emul. Polym. Latex Technol.. Ann. Short Course Papers, 10th, Vol.
11. Lehigh Univ., Bethlehem, Pennsylvania.
White, W. W. (1976). J. Elastomer. Plast. 8, 475.
Zisman, W. A. (1968). "Contact Angel, Wettability and Adhesion" (R. F. Gould, ed.), Adv.
Chem. Ser. No. 43. American Chemical Society, Washington, D.C.

9
Latex Reactor PrincipIes:
Design, Operation, and Control
A. E. Hamielec and J. F. MacGregor

1. Introduction .
11. Batch Reactors .
A. Introduction .
B. Monomer Consumption Rate .
C. Polymer Quality .
D. Initiation Systems .
E. Heat Removal
111. Continuous Stirred-Tank Reactors: Steady-State
Operation .
A. Introduction .
B. Polymer Quality .
C. Optimal Reactor Type and Operation
for Continuous Emulsion Polymerization
IV. Continuous Stirred-Tank Reactors: Dynamic
..
Behavior .
A. Introduction .
B. A Dynamic Model Based on Age Distribution
Functions.
C. Consideration of Dynamics in Reactor Design .
V. On-Une Control of Continuous Latex Reactors
A. Introduction .
B. On-line Instrumentation
C. On-line Estimation .
D. On-line Control.
VI. Summary.
Nomenclature
References

319
320
320
322
325
329
330
333
333
334
338
339
339
340
344
345
345
346
348
349
351
351
353

l. Introduction
The principIes of latex reactor design, operation, and control
will be illustrated by a consideration of the homopolymerization of styrene
and vinyl acetate. Emulsion polymerization of vinyl acetate follows Case 1
319
EMULSION

POLYMERIZATION

Copyright
~ 1982 by Aeademie Press, Ine.
AIl rights of reproduetion
in any form reserved.
ISBN 0-12-556420-1

320

A. E. Hamielec and J. F. MacGregor

kinetics with the production of polymer with long-chain branching, and


styrene follows Case 2 kinetics forming linear polymer chains. The choice
of these systems will permit a consideration of the control of long chain
branching as well as polymer particle number and size distribution, all of
which are important characteristics of paints, adhesives, and rubber products. Another consideration, of course, is the considerable number of
experimental and theoretical investigations published on the emulsion
polymerization of styrene and vinyl acetate. The principIes of latex reactor
design, operation, and control illustrated here with eniulsion homopolymerization should largely be applicable to copolymerizations.
ll.

Batch Reactors

A. lntroduction*
We consider a polymerizing system with Np polymer particles per unit
volume of latex. Np is usually of the order 1016-1018 particles per liter of
latex. Early in their life when polymer particles are relatively small (1001000 ).and when the polymer concentration is relatively low, termination
with an entering radical will be effectivelyinstantaneous. Thus, particles will
contain either zero or one radical. Later in the polymerization particles may
contain two or more radicals. The polymerization rate in a single particle
with n free radicals is given by

= kp[Mp]n

R
pp

(1)

NA

The total polymerization rate is obtained by summation over all polymer


particles
<1J

Rp

"

(2)

kp[Mp] '--' nNn


NA n=1

The average number of radicals per particle is given by


<1J

= n=1 n;np

(3)

= kp[Mp]Np

(4)

and finally
Rp

* Symbols are defined on pages 351-353 in the Nomenclature

section.

9. Latex Reactor Principies: Design, Operation, and Control

321

The conversion of monomer X is defined as


(5)
and therefore
dX
dt

= VRp

(6)

NMn

To get a feel for a typical polymer production rate, let us employ reasonable
numerical values for the parameters in Eqs. (4) and (6). These follow:
kp
[Mp]
Np

= 103 liter/Gmol

v=

sec

= 5 Gmol/liter

= 1017 partic1es/liter

latex

1 liter

NMn= 5 Gmol
= 0.5 (Case 2 kinetics)

from which Eq. (6) yields dX/dt '" 30%/hr. This is certainly a desirable rate
of polymerization for commercial production. The production rate in
practice would likely be limited by the heat removal capacity of the reactor.
The number of partic1es Np is a function of emulsifier type and concentration and initiator level, although for monomers that obey Case 1
kinetics such as vinyl chloride and vinyl acetate, Np is almost independent
of initiator level (Ugelstad et al., 1969; Friis and Nyhagen, 1973). The
calculation of Np for various reactor operations will be discussed later. The
monomer concentration in the polymer partic1e [Mp] can be obtained using
a simple mas s balance. Assuming the monmer and polymer volumes are
additive, one obtains. the following relationship for the conversion interval,
Xc :<:::
X :<:::
1.0 where Xc is the monomer conversion at the end of Stage 11.
(7)
To calculate

[Mp] in Stages 1 and 11 set X

= Xc

in Eq. (7). The deter-

mination of requires, in general, solution of the Smith-Ewart recurrence formulas (O'Toole, 1965; Ugelstad et al., 1967). The general
solution for the case of negligible water-phase termination (Y = O)is shown
graphically in Fig. f. The parameters are given by
a = p1 V/(Npk,P)
a' = R.V(Npk,P)
m = kdeV/k,P

(8)

Y = 2Npk,pk'w/(k; V)

For most emulsion polymerizations, water-phase termination can be neglected (Y = O) (Ugelstad et al., 1969). However, when an active chain-transfer

322

A. E. Hamielec and J. F. MacGregor

y=o

-3 L..L
-8

-6

-4

-2

log a'
Fig. 1.

General solution of tbe Smitb-Ewart recurrence equations. Negligiblewater-

pbase termination (adapted from Ugelstad et al., 1967).

agent is employed, higher rates of radical desorption may result giving


significant water-phase termination and lower polymerization rates. It has
also been shown that at high initiator levels the free radical capture
efficiency may be appreciably less than 100%, presumably due to significant
water-phase termination (Gilbert et al., 1980).
B.

Monomer Consumption Rate

1. Case 1 Kinetics
Consider a monodispersed latex, where water-phase termination is pegligible and termination is instantaneous when a radical enters a polymer
particle containing one radical. By definition, N2 = N3 = ... = O and the
total radical entry rate per liter of latex equals pl. Application of the
stationary-state hypothesis gives
R)

= 2pl(N/Np)

(9)

In the limit of very rapid radical desorption


(10)
and therefore
(11)
and
(12)

9. Latex Reactor Principies: Design, Operation, and Control

323

Ugelstad et al. (1967, 1969) have derived a more general expression which
accounts for the possibility that two radicals can coexist in a polymer
particle:

R
p

kp[Mp]

(N ) (
A

1/2

RI12 VpNi + Np
2k.p
2kde

(13)

A useful expression for the radical desorption coefficient fo11ows(Nomura et


al., 1976).
kde = (12~w/mD;(;)(kf.Jkp)
(; = (1 + ~w/m~p)

(14a)
(14b)

In Stage III, the reduction in Dp and increase in (; lead to a reduction in


desorption coefficient with an increase in polymer-phase viscosity. For vinyl
chloride emulsion polymerization the separate monomer phase disappears
at about 70% conversion (Xc = 0.7). Therefore, as soon as Np reaches a
constant value, the only parameter that changes for X < Xc is Vp.In fact, it
is mainly the increase in Vpthat causes the acceleration in rateo For X > Xc
the situation is more complex, with both k.p and kde fa11ingas polymer
concentration increases. For vinyl acetate, the separate monomer phase
already disappears at 20% conversion. For X > Xc, Vpis almost constant;
however [Mp], k.p, and kdea11decrease with conversion. These effects will be
discussed in more detaillater.
2. Case 2 Kinetics
Consider a monodispersed latex, where water-phase termination and
radical desorption are negligible and termination is instantaneous when a
radical enters a polymer particle containing one radical. By definition,
N2 = N3 = ... = O and the total radical entry rate per liter of latex equals
R" the rate of radical generation per liter of latex via initiator decomposition. The rate of radical termination per liter of latex is equal to twice
the rate at which radicals enter polymer particles containing one radical
(two radicals are consumed per radical entry). Application of the stationarystate hypothesis gives
R, = 2R,NdNp

(15)

= No = tNp

(16)

In other words
NI

and
-1.
-2

(17)

324

A. E. Hamielec and J. F. MacGregor

One normally assumes that systems such as styrene and methyl methacrylate, where transfer to monomer is not prominent, follow Case 2 kinetics
when latex particles are small and termination in polymer particles is
instantaneous. It has recently been shown that at low initiation rates radical
desorption can be significant relative to radical absorption, and as a
consequence values appreciably smaller than 0.5 were found (Gilbert et
al., 1980).At higher initiation rates = 0.5 was approached. The use of
chain-transfer agents would of course increase the desorption rate and
lower .
For Case 2 kinetics the rate of polymerization is given by
Rp

= kp[Mp]Np/2NA

(18)

with the rate independent of initiator concentration for fixed Np. Ir the
initiation rate is reduced to a great extent, will fall below 0.5 as radical
desorption becomessignificant.
3. Stage III PolymerizatiolJs
In Stage III, the polymer concentration in the polymer particles increases
with ever increasing chain entanglements, and if the polymerization tem-:
perature is below the glass-transition temperature of the polymer being
syntht;sized, a glass-state transition will occur with the diffusion coefficients
of small molecules falling by several orders of magnitude. The pro paga tion
reaction becomes diffusion-controlled and a limiting conversion of less than
100% is reached where the rate of polymerization is effectively zero (Friis
and Hamielec, 1976). Marten and Hamielec (1978) pro po sed the following
models for the effect of polymer molecular weight and free volume on the
termination and propagation constants.
k
~=

M:

2a

1
---

VF

VFcrl

( ) [ (
~ = exp[- B(~ - ~ )]
k,o

kpo

a exp(b/RT)

~C<I

exp -A

Mw

VF

)]

(20)

VFcr2

(21)

= M~c:' exp(A/VFcr.)

VF = [0.025 + Cip(T-

(19)

Tgp)]<I>p

+ [0.025 + Cim(T- Tgm)]<I>m

(22)

where a, b, Ci,A, B, VFcr!'and VFcr2are adjustable parameters found by


fitting rate and molecular weight data.

9.

325

Latex Reactor Principies: Design, Operation, and Control

TABlEI
Diffusion-ControlledTermination and Propagation Model Parameters
Polyrner
PMMA
PS
PVC

Tgp(0C)

ex

110
95
85

1.05
0.48
0.25

1.0
0.7
0.2

1.75
1.75
1.75

VFcr1

0.066
0.043
0.085

0.563
0.926

8900
3830

Parameters for three homopolymer systems are tabulated in Table 1.


Using Eq. (20), Sundberg et al. (1980) observed values of B = 0.6 and
VFcr2= 0.061 for the emulsion polymerization of both styrene and methyl
methacrylate. The volume fraction of monomer in the polymer partic1e for
X > Xccan be ca\culated using
(23)
To ca\culate <I>m
for X < Xc set X

= Xc in Eq. (23).

C. Polymer Quality
1. Molecular Weight
Polymer can be produced via the followjng reactions: transfer to monomer to chain transfer agent and to polymer, termination reactions, and
terminal double bona reactions. Transfer to polymer and terminal double
bond reactions produce long-chain branches. Transfer to small molecules
and termination reactions produce linear polymer chains. A major simplifiGation in the modeling of molecular weight and branching development
results when it can be assumed that a negligible amount of polymer is
produced via termination reactions. Fortunately, in most emulsion polymerizations, transfer reactions are relatively more important than termination
reactions in the production of polymer. This is a direct consequence of the
compartmentalization of radical s in polymer partic1es which permits commercial polymerization rates at relatively low radical initiation rates. When
the amount of polymer produced by termination reactions is negligible,
molecular weight and branching development is independent of initiator
and emulsifier levels (or number and size of the polymer partic1es). The
appropriate equations that describe molecular weight and branching development, which have been proved valid for the emulsion polymerization

326

,
A. E. Hamielec and J. F. MacGregor

of vinyl chloride and vinyl acetate in a batch reactor, follow (Friis and
Hamielec, 1975).
For X < Xc

MR
KXc
1+CM
1 - Xc

)
2MR 1 + KXc
(
)
CM+-1+1 - Xc (1 - Xc)(
1 - XC)
EN= ~
( - )( + K)

MN=-

(25)

M=

l-Xc'

Xc

CM = krm/kp,

KXc

2CpXc

CpXc

CpMn
MR

Cp = kr~/kp,

K = k;/kp

(26)

(27)
(28)

For X > Xc
dQo = CM- KQo
dX
1- X

(29)

dQl = 1
dX

(30)
1

dQ2 =2
dX

KX
+ 1- X

)(

d(QoEN) - CpX + KQo


dX
1- X
IN = 1000MREN/MN
MN

= MRQdQo

Mw = MRQ2/Ql

CpQ2

KX

+1=X+1=X
C

CpX
M+I-X

(31)

)
(32)
(33)
(34)
(35)

Equations (29-32) can be solved using standard numerical techniques. The


initial conditions may be found using Eqs. (25-27).
When polymer produced by termination reactions is appreciable, molecular weight and branching development depend on the initiation rate and
emulsifier concentration (number of particles and their size distribution).
This more general problem has never been solved and is perhaps intractable
because of the long-chain branching reactions. Solutions have been obtained for the case where long-chain branching reactions are neglected;

..

9. Latex Reactor Principies: Design. Operation. and Control

327

these involved the use of population balance equations and sophisticated


numerical solution techniques. The most general model that predicts particle size and molecular weight distributions was developed by Min and Ray
(1974). Work of a more fundamental nature on the details of the microscopic events that control particle size and molecular weight development is
being done by Lichti et al. (1977; Lichti, 1980).
2.

Particle Number and Size Distribution

The prediction of particle number and size distribution has been far less
successful than the prediction of conversion time histories and molecular
weight development, given these parameters as initial conditions. The
nucleation and early growth of polymer particles is even today, after the
comprehensive investigations of Ugelstad and Hansen (1976), not. well
understood. There are several reasons for this. First, the measurement of
number

and size of polymer

particles

smaller than

100

A presents

rather

difficult experimental problems. Many complex processes occur simultaneously including radical capture by micelles and polymer particles,
precipitation of growing radicals from the aqueous phase, and finally
particle flocculation and coagulation. To illustrate how certain variables
such as initiation rate and emulsifier concentra tion might affect particle
number, we will refer to a highly idealized model for particle number after
Smith and Ewart (1948). To simplify the analysis, let us assume that the
area occupied by an emulsifier molecule is the same on the micelle as on the
polymer particle. In other words
(36)
For typical emulsifiers, as has values of 30-100 A2jmolecule or 1.8-6 x 107
dm2jGmol. The detailed derivation may be found in the original reference
and the final result has the form
(37)
where (X= 0.53 when radical capture by polymer particles is neglectedand
= 0.37 when micelles and polymer particles both capture radicals and
assuming surface area on both kinds of particles are equally effective in the
capture process. The diffusion theory of radical capture would predict that
the micelles, being smaller, would provide more effective surface area for
radical capture and thus give an exvalue greater than 0.37. According to
Ugelstad and Hansen (1976), the capture process is more complex with the
possibility of both collision and diffusion theories of radical capture applying under different limiting conditions. To develop effective control strategies for polymer particle concentration, it is of considerable importance to
(X

328

A. E, Hamielec and J. F. MacGregor

establish the time scale for micellar nucleation of polymer particles. The
effectiveness of any control scheme would of course depend on the
nucleation time. Np is given by
Np

= NAR.tr

(38)

where tr is the nucleation time (time to end of Stage 1). A value of


2fkd

10-6 sec-1 (K2S208

=6

at 50C) and [/]

X 10-3

Gmol/liter

gives

an initiation rate' of 6 x 10- 9 Gmol/sec liter latex. The time required to


nucleate 1018 particles per liter of latex is thus only abou't 5 mino It is clear
that manipulation of initiator and emulsifier concentrations to control
polymer particle nucleation would not be an easy task. Equation (37) has
been shown to be valid for the emulsion polymerization of styrene (Gerrens,
1959). It does not apply to Case 1 systems such as vinyl chloride and vinyl
acetate where in fact Np is independent of initiator concentration. This
disagreement is not unexpected as .J.is a function of initiation rate and the
total radical entry rate into micelles and polymer particles is much greater
than the initiation rateo Parti~le nucleation has been successfully modeled
for vinyl chloride polymerization by Min and Gostin (1979). The particle
size distribution (PSD) of a fully converted latex depends on the conditions
in both Stages 1 and 11. It depends on the original PSD at time tr (end of
Stage 1) and the subsequent growth rates of various sized particles in Stage
11.Gerrens (1959) has shown that the cumulative PSD at time tr is, for Case
2 kinetics, given by

v
cum F(v) =

.J.tr

v
= O.53.J.2/5(as[S]/R.NA)3/5

(39)

It is clear that the PSD at time tr narrows with increasing initiation rate
and broadens with increasing emulsifier concentration. Ir the original PSD
at tr is very narrow, then all particles will grow at the same rate in Stage 11
and the fully converted latex will have a narrow PSD. On the other hand, if
the original PSD is broad, the final PSD will be broad. Data on the final
PSD after Gerrens (1959) showing the effect of initiator concentration on
the breadth of the PSD are shown in Fig. 2. As anticipated by Eq. (39), the
distribution broadens with decrease in initiator concentration. To predict
the PSD to the end of Stage 11 for a broad distribution requires a
knowledge of the radical capture mechanism. According to the collision
theory, surface area on a small particle is equally effective in capturing
radicals to that on a large particle. On the other hand, according to the
diffusion theory surface area on a small particle is more effective.

9.

Latex Reactor Principies: Design, Operation, and Control

329

1.0

0.8

~ 0.6
IL

E
::J
()
0.4

0.2

1.0

0.5

1.5

2.0

V (LlTERS X 1016)

Fig. 2.
Cumulative partic1e size distributions for styrene emulsion polymerization at
different initiator concentrations [1]0 (in g/liter of water) (adapted from Gerrens, 1959):
A, 5.41; B; 2.88; C, 1.80; D, 1.08; E, 0.541; F, 0.361.

D. lnitiation Systems
Hydrophilic initiators such as the persulfates and hydrogen peroxide and
initiators with varying solubilities in the polymer particles such as the
hydroperoxides with 'or without reducing agents are employed in emulsion
polymerization. A recent review of redox systems in emulsion polymerization was published by Warson (1976). The most efficient manner of
adding the components of a redox system is gradually in stages or by
adding the peroxide at the start and then the reducing agent gradually
(Warson, 1976).
A typical redox system for cold SBR production employs sodium
formaldehyde sulfoxylate (reducing agent), a hydroperoxide (oxidant), and
ferrous sulfate, plus' a chelating agent or a chelated irol) salt. A simplified
kinetic mechanism for this redox couple follows (Wright and Tucker, 1977).
ROOH + FeH
Fe3+ + XH

--

RO'+M-

ROM'

330

A. E. Hamielec and J. F. MacGregor

The function of the chelator is to complexthe ferrousion and thus limit the
concentration of free iron. Redox systems appear very versatile, permitting
polymerization at ambient temperatures and the possibility of control of the
rate of radical initiation versus polymerization time. This would thus permit
control of heat generation and the minimization of reaction time. The use of
the redox system ammonium persulfate (2 mmol) together with sodium
pyrosulfite (Na2S20S 2.5 mmol) together with copper sulfate (0.002 mmol)
buffered with sodium bicarbonate in 1 liter of water form an effective redox
system for vinyl acetate emulsion polymerization. The reaction was started
at 25C and run nonisothermally to 70C. The time to almost complete
conversion was 30 min (Warson, 1976; and Edelhauser, 1975).
The rational for the use of finishing initiators that have appreciable
solubility in the polymer particJe follows. At high conversions, the concentration of monomer in the aqueous phase is very low and water-phase
termination of hydophilic radical s becomes excessive. The rate of radical
entry into polymer particJes is thus greatly reduced an,d the polymerization
rate falls to a very low level prematurely.

E. Heat Removal

1. Techniquesfor Heat Removal


An energy balance for a batch reactor may be written as
V( - LlHp)Rp = UAw(T - Te)+ QE
(40)
As reactors are increased in size it is usually necessary to add supplemental
heat removal capacity since the heat transfer area of the jacket only
increases with reactor volume to the 0.67 power whereas heat generation
rate increase is proportional to volume. Internal cooling coils and baffies
are often used to increase heat transfer and in this way heat transfer area
may be increased 40-50%. In some processes external heat exchangers are
used where the reaction mixture is continuously pumped through the heat
exchanger cooled and ret.urned. Reflux cooling is very effective in increasing
heat removal capacity and is extensively used with large latex reactors.
It is of interest to investigate the effect of reactor size on the magnitude of
QE' The results of sample calculations indicate that for reactor volumes
greater than 5000 gal additional cooling capacity would likely be required
to achieve commercial production rates.
2.

Techniquesfor the Reduction of Reaction Time

To minimize the polymerization time, it is necessary that full utilization


of the heat removal capacity of the reactor be made at all times during the

331

9. Latex Reactor Principies: Design. Operation. and Control

batch. In other words, the polymerization rate and associated heat generation rate should equal the heat removal capacity at all times during the
batch. For an isothermal polymerization this can best be done by controlling the initiation rate using a redox system. Of course, the sinallest reaction
time occurs with adiabatic polymerization. The temperature rise can be
reduced by controlled heat removal to moderate the spread in molecular
weights. Typical conversion and heat generation rate profiles for an isothermal polymerization with constant radical initiation are shown in Fig. 3. The
heat generation rate over the batch is far from optimal as the maximum
heat remo val capacity of the reactor is about 220 arbitrary units. The
polymerization time could be significantly reduced by employing a higher
initiation rate early in the polymerization and by letting it fall with time so
as not to exceed a heat generation rate of 220 arbitrary units. A similar
reduction in polymerization time could be achieved with semibatch monomer feed using monomer starvation to give a polymerization rate equal
to Rpmax'A cold monomer feed would also assist in heat removal. The
shortest batch time would be obtained using semibatch monomer feed
giving Rpmaxand permitting the temperature to rise.
In semibatch emulsion polymerizations the polymer particles are kept
monomer-starved to obtain higher rates of polymerization and to permit
easier control of the rate and particle size distribution. There are two
aspects to the control of PSD. The controlled addition of emulsifier during
particle growth stabilizes the particles without further particle nucleation.
The second aspect is related to the particle sticky stage which often occurs
1.0

2.5

0.8

1.5 ~

0.6

"el.
:J:
<1
1.0 I

X
0.4

0.5

t (hrl

Fig. 3. Conversion and instantaneous heat generation rate for a typical isothermal
batch emulsion polymerization.

332

A. E. Hamielec and J. F. MacGregor


5

.O

>C
U
Q)
lit
I

..
Q)
::

......

o
E

Q.
Q:

1.0

. 2.0

3.0

4.0

ff4] mol/ 1Iter


100

50
Conversion

Fig. 4.
semibatch

J
o

(%)

Polymerizationrate curve showingstable and unstable operating points for a

emulsion polymerization.

at intermediate conversion levels and certain polymerjmonomer concentrations. During the sticky stage wall fouling and formation of coagulum
often occur. The use of semibatch operation under monomer-starved
conditions effectively eliminates the sticky stage. To illustrate that it is
easier to control a semibatch emulsion polymerization under monomerstarved conditions, consider the rate curve shown in Fig. 4. Point P1 at low
monomer concentration is a stable operating point for semibatch feed of
monomer, whereas P2 is unstable. A consideration of perturbations in
monomer feed rate clearly illustrates this fact.
Semibatch operations are commonly employed in copolymerizations to
maintain a more uniform copolymer composition. The monomers are fed at
rates equal to their consumption rates in the reactor.

9. Latex Reactor PrincipIes: Design, Operation, and Control

333

111. Continuous Stirred-Tank Reactors: Steady-State Operation

A. lntroduction
In this section, emulsion polymerization in continous stirred-tank
reactor(s) (CSTRs) operating at steady state will be treated, with emphasis
on polymer quality, Le., molecular weight, long-chain branching, and
polymer particle number and size distribution. The commercially important
problems associated with oscillations in conversion (Le., particle number
and size, weight average molecular weight, and long-chain branching) and
their control will be considered in following sections.
A single CSTR or a train of CSTRs may offer several advantages over
batch reactors both with regard to production rate and polymer quality.
The copolymer composition distribution obtained in a CSTR is generally
narrower than that for a batch reactor. To obtain fairly narrow copolymer
composition distributions without appreciable loss in productivity, one has
the flexibility to increase the number of CSTRs in series to approach plug
flow.
Because of diffusion-controlled termination and propagation with concomitant Rpmaxat a particular conversion, it is possible to operate a CSTR
at considerably higher production rates. Because of the additional beneficial
effects of cold monomer and water feeds on heat removal, much higher
production rates are possible than with a batch reactor of the same volume.
It should be remembered that polymer production rates are usually limited
by heat removal capacity.
Quality variations in the polymer produced in batch reactors are often
caused by slight vaI:iations in the reactor start up procedure. Furthermore,
the polymerization rate may change considerably during the batch and this
may give temperature variations that are difficult to reproduce causing
batch-to-batch variations in quality. These problems would be minimized
with CSTRs if the continuous reactor system could be operated for at least
several weeks before wall fouling and coagulum build up become critical
and require reactor shutdown for cIeanup. Ir an effective start-up procedure
for a continuous reactor train is not available, the costs associated with
offspec material' could make continuous operation uneconomical. In addition, with a continuous reactor system one loses the flexibility of batch
reactors when a multiproduct operation, with its short productions runs, is
involved.
During start-up and operation of CSTRs there is a potential safety
hazard created by the possibility of more than one stable steady-state

334

A. E. Hamielec and J. F. MacGregor

operating point (Ley and Gerrens, 1974). The start-up procedure should
ensure that the desired steady state is reached.

B. Po/ymer Qua/ity
1. Particle Number and Size Distribution
The frequency distribution of partic1e radius F(R) is. related to the exit
age distribution E(t) with the identity
F(R) dR

= E(t) dt

(41a)

which states that the number fraction of polymer partic1es in the exit stream
of a CSTR train with radius in the range R to R + dR equals the fraction of
fluid in the exit stream with age in the range t to t + dt. For a single CSTR
F(R) dR

= rr1 exp( -

tlO) dt

(41b)

Equation (41b) is valid when there are no polymer partic1es flowing into the
reactor with aIl the partic1es nudeated within the reactor. lt is assumed thar
density changes can be neglected and that partic1es foIlow the streamlines.
These are reasonable assumptions in view of the smaIl size of partic1es and
the smaIl density difference between particIe and water. When two or more
CSTRs are employed in series, however, one must remember that the total
residence time of a polymer partic1e is made up of different times in each
reactor in the train. The relative amounts of time spent in each reactor wilI
not matter when the volumetric growth rate of a partic1e is the same in
each. This would require that the temperature, monomer concentration,
and average number of radicals per partic1e be the same for each reactor, an
unlikely possibility. This idealization is useful, however, when ilIustrating
the effect of increasing the number of CSTRs in series on the breadth of the
partic1e size distribution.
The exit age distribution function for N tanks in series is given by
E(t) =

(~)((N ~ 1)!)(~r-1 exp(-tIO)

(42)

The change in partic1e radius with time is obtained as


dRldt

= KlR2

(43a)

where
K = kp[Mp]MR/4nNAdp

(43b)

To solve Eq. (43a), a relationship between and time or partic1e radius is


required. The simplest situation is to treat Case 2 kinetics where is

9.

335

Latex Reactor Principies: Design, Operation, and Control

independent of t and R and equals 1- Using = 1, integrating Eq. (43a) and


then substituting for t in Eq. (41b) for a single CSTR, one obtains the
differential frequency distribution.
F(R)

= (2R2/K()

(44a)

exp[ -(2/3K()R3J

and the cumulative frequency distribution


cum F(R)

= 1-

(44b)

exp[ -(2/3K()R3J

In a similar manner the following PSD may be derived for Case 1 kinetics
for a single CSTR (Friis and Hamie1ec, 1976).
(45a)
F(R) = K~() exp[ -(2~1() )R2 J
cum F(R)

K
1

(45b)

= 1 - exp[ -C~I()

(4/3n)2/3 R,

kp[MpJMR

dp

)R2 J
1/2

(45c)

J[ 16n2NANpke J

The cumulative distribution for Case 2 kinetics [Eq. (44b)] is shown in


Fig. 5 plotted for different values of the parameter K(). It is c1ear that the

0.8

~0.6

=' 0.4
u

R ( Al(
Fig.5.

Cumulative

16

12

partic1e size distributions

20

24

10-3)
for a single CSTR. Case 2 kinetics.

336

A. E. Hamielec and J. F. MacGregor

PSD broadens with increasing K(). For situations where a separate monomer phase exists, [Mp] is a constant independent of (/ and perhaps
varying slightly with temperature. K will therefore depend on temperature
alone and increase with an increase in kp. Thus, increase in temperature and
() broaden the PSD. The PSD from a single CSTR is excessively broad for
most applications and it is thus desirable to add additional CSTRs in the
train to narrow the PSD. Eq. (42) shows the effect of increasing the number
of CSTRs on the .exit age distribution which also narrows and approaches
that of a plug flow reactor (PFR) (or a batch reactor). The narrowest PSD
would be obtained in a PFR. The use of 5-6 CSTRs in series should give a
PSD which is insignificantly narrower than that for a PFR.
The total number of polymer particles per liter of emulsion (Np) in a
single CSTR may be calculated by
[S]o

[S]

Np

(4nNp/as) Loo R2F(R)

= ()RNA

as[S]

Ap

dR

(46)

=O

[S]

+ aseS] = ()RNA [S]o

(47)

The last term in Eq. (46) gives the amount of emulsifier covering the
polymer particles in the exit stream. Equation (47) is obtained by equating
the rate of particle outflow to the rate of particle nucleation within th

reactor employingthe collisiontheory. For Case 2 kinetics N p is givenby


()RNA
Np

(48)

= 1 + NAR(4n/as[S]0)(3K/2)2/3()S/3r(5/3)

This equation is identical to the one derived by Gershberg and Longfield


(1961). Equation (48) is shown graphically in Fig. 6. It appears that for a

Q.
Z

10
81hrl

Fig. 6.

Number of polymer particles as a function of emulsifier and initiator levels for a

single CSTR. Case 2 kinetics.

9.

Latex Reactor Principies: Design, Operation, and Control

337

given set of operating conditions there is a O which gives a maximum Np.


Npm.xhas been observed experimentally by Veda et al. (1971) and Nomura
et al. (1971) for continuous styrene emulsion polymerization. Gerrens and
Kuchner (1970) have furthermore confirmed the zero-order dependence on
initiator, the first-order dependence on emulsifier at long residence times,
and that Np oc 0-2.3 at large O. The mechanisms of particle nucleation for
commercial monomer systems such as vinyl acetate, vinyl chloride, methyl
acrylate, and chloroprene have not been as extensive1y studied experimentally as has styrene. It is likely, however, that an Npmaxwould occur with
these more complex systems and this fact should be considered when
designing a reactor train with high volumetric efficiency (Nomura, 1980).
2.

Molecular Weight and Long-Chain Branching

For the calculation of molecular weights in CSTR emulsion reactors, a


useful classification comes to mind. This includes those monomer systems
whose molecular weight and branching development depends on particle
size and those that do no1. Styrene falls into the former class and vinyl
chloride and vinyl acetate into the latter class. Thus, in vinyl chloride
emulsion polymerization where LCB is neglected, the instantaneous molecular weight distribution is given by
W(M)

= (MjM~)

exp(-MjMN)

(49)
-1

M
N

1.fJ:
2

=M

krm + krx[xJ

R kp.

kp[M]

It should be noted that at high conversions some branched PVC may form
and it may therefore be necessary to consider the transfer to polymer
reaction (Hamielec et al., 1980b). When transfer reactions control molecular
weight development, all polymer particles in the first CSTR in the train will
have the same molecular weight distribution (MWD) (neglecting inflow of
polymer particles) given by Eq. (49). It there is no CTA and each reactor in
the train is at the same temperature then the MWD will be the same in all
the reactors in the train. Ir a CT A is employed for an isothermal train, the
contribution from the CT A will increase in downstream reactors as the
monomer concentration falls, and therefore molecular weight averages will
fall. Of course, the MWD in each reactor will be of the same form and given
by Eq. (49), but the final whole polymer will be a composite of the polymer
produced in each reactor. This can be easily calculated if the conversion in
each reactor is known. It should be remembered that less than one CTA
molecule is consumed per polymer chain and therefore the CT A consumption by reaction can often be neglected.

338

A. E. Hamielec and J. F. MacGregor

With vinyl acetate the situation is more complex since transfer to


polymer and terminal double bond reactions are significant at all conversion levels. If one calculates the MN, Mw, and BNfor a nucleated polymer
particle with different lifetimes in a single CSTR, it is found that these
quantities are independent of the particle age and are given by

M = MR
N.

Mw

CM

where

KWp

(50)

KW
2MR 1 + 1 -

(
= ~
(c

BN

( + 1- W )
(

~)

M+1-W

Wp

)(

1-W

CpMn

+K

MR

(51)

CpWp- 2CpWp 1
p

)(

K~

+l-W

(52)

is the weight fraction of polymer in the polymer particle. Friis et

al. (1974) have measured the following parameters for vinyl acetate using
batch emulsion polymerization.
kp = 1.89 x 107 exp(- 5650/R T)
krm= 3.55 x 106exp(-9950/RT)
krp = 1.430 x 106exp(-9020/RT)
(k:)o = 1.07 x 107 exp(-5650/RT)

liter/mol sec
liter/mol sec
liter/mol sec
liter/mol sec

(53)

For Wp> 0.2, a decrease in k: was observed and this decrease is given by

k: = (k:)o - (l69.6~ + 479.9W; + 1014.0W;)

(54)

At 60C with a conversion of 60% in a single CSTR, BN= 4.86 long


branches per polymer molecule. Increasing the number of equal-sized
CSTRs in series so as to approach plug ftow (or batch reactor) behavior,
one would obtain a BNof about 1.5 long branches per polymer molecule. In
other words the long-chain branching frequency can be greatly reduced by
increasing the number of CSTRs in a train. This is clearly the strategy that
is used for the production of SBR and polychloroprene. Even with a CSTR
train it is necessary to limit the conversion to keep branching at a low level.
C.

Optima/ Reactor Type and Operation lor Continuous Emu/sion


Po/ymerization

This section is a review of a recent paper by Nomura (1980) who has


made some interesting suggestions about the design and operation of

9. Latex Reactor Principies: Design, Operation, and Control

339

continuous latex reactors with stable operation at steady-state conditions.


The conventional CSTR train is operated with a11the recipe ingredients
being fed into the first reactor and the product latex removed from the last
reactor in the train. For a given production rate, the size and cost of the
CSTR train can be reduced by maximizing the nuc1eation of polymer
partic1es in reactor one. This can be done by increasing the emulsifier and
initiator concentrations in the feed stream and by lowering the temperature
of the first reactor. The use of higher emulsifier levels may have deleterious
effects on the polymer product, however, Other techniques of maximizing
the number of polymer partic1es in the first reactor are given. The optimal
choice of mean residence time will give Npm.xas already discussed earlier in
this chapter. Another approach to increasing the number of polymer
partic1es early in the train is to use a plug flow type reactor (PFR) as reactor
one. A larger number of polymer partic1es can be produced in a PFR than
in a CSTR at the optimal value of (J.The volume of the PFR could then be
reduced substantia11y by reducing monomer in feed and feeding rest of the
monomer into reactor two. It is also c1aimed that a PFR as first reactor in
the train would substantia11y reduce oscillations and thus increase the
stability of operation of a CSTR train. These design and operation suggestions by Nomura are based on calculations performed with a steady-state
model. More realistica11ythe design, operation, and control of a CSTR train
of latex reactors should be based on calculations with a transient CSTR
model. This subject will be considered in the fo11owingsections.

IV.

Continuous S~rred-Tank Reactors: Dynamic Behavior

A. lntroduction
Continuous emulsion polymerization is one of the few chemical processes in which major design considerations require the use of dynamic or
unsteady-state models of the process. This need arises because of important
problems associated with sustained oscillations or limit cyc1es in conversion, partic1e'nlJmber and size, and molecular weight. These oscillations
can occur in almost a11commercial continuous emulsion polymerization
processes such as styrene (Brooks et al., 1978), styrene-butadiene and vinyl
acetate (Greene et al., 1976; Kiparissides et al., 1980a), methyl methacrylate,
and chloropene. In addition to the undesirable variations in the polymer
and partic1e properties that will occur, these oscillations can lead to
emulsifier concentrations too low to cover adequately the polymer partic1es,
with the result that excessive agglomeration and fouling can occur.
Furthermore, excursions to high conversions in polymer like vinyl acetate

340

A. E. Hamielec and J. F. MacGregor

and styrene-butadiene where transfer to polymer reactions are important


can lead to excessive long-chain branching and thereby result in poor
processability of the rubber.
Although these oscillations can be avoided by operating at sufficiently
high emulsifier concentrations, the concentrations needed are often high for
most commercial processes. Even if, under the conditions used, a nonoscillatory steady-state exists, it is not unusual that during start-up or other
disturbance transients an oscillation is induced which qamps out only very
slowly (20 or more residence times). Furthermore, on is never certain
whether steady-state models actually apply in any given situation unless
one either has experimental confirmation or verifies it by first solving a
dynamic model.

B.

A Dynamic Model Based on Age Distribution Functions

In this section we consider a model for the continuous emulsion


polymerization of vinyl acetate based on one presented by Kiparissides et
al. (1978). Latex properties such as average particle size, number of particles, total particle surface area and volume, molecular weight averages,
and average number of long-chain branches per molecule will be considered. The approach is similar to that used by Dickinson (1976) and
focuses on the residence time or age distribution of particles in the reactor
rather than on the size distribution. The latter approach (Min and Ray,
1974) has recently been used successfully by Min and Gostin (1979) to
model the semibatch emulsion polymerization of poly(vinyl chloride)~
However this approach could lead to difficulties in modeling reactors where
the growth rate of a particle is not only dependent on its current size and .
the current conditions in the reactor but also on the conditions prevailing
during its previous history. Using an age distribution approach if we define
the function n(t, -r)d-r to be the number of particles in the reactor at time t
that were born in the time interval (-r,-r+ d-r),and p(t, -r)tobe a property of
the latex associated with this cIass of particles, then the total property
P(t) = J~ p(t, -r)n(t,-r)d-rfor all particles in a CSTR reactor will be given by
dP(t)
-=--+
dt

P(t)
()

1 op(t, -r)
) ( t) +
f( tpt
-nt
( -r) d-r
,
'
1o ot

(55)

where () is the reactor residence time and f(t) is the net rate of particle
generation in the reactor. Using this equation, property balances for the
total number of particles, the total diameter, the area and volume of

9.

341

Latex Reactor Principies: Design, Operation, and Control

partic1es,and the conversion can be written as follows (Kiparissides et al.,


1978; Chiang and Thompson, 1979):
dN(t) = - N(t) + f(t)
dt
()

(56)

dD (t)
--f-

(57)

D (t)

= ---t-

+ 2~(t)N(t) + dp(t,t)f(t)

dAp(t)
+ 4nW)Dp(t) + tIp(t,t)f(t)
dt = - A~(t)
(7
dVit)
dt

=-

dX(t)
--

= (1 - e-I/6){

V~t)
(7

(58)

(59)

+ ~(t)Ap(t) + vp(t,t)f(t)

- X(t)
()

+ Rp(t)

(60)

[MFJ }

The exponential term in the conversion equation arises from the assumed
start-up condition of a reactor filled only with degased water and then fed
at time zero with a stream having monomer concentration [MFJ.
Assuming both micellar and homogeneous nuc1eation the partic1e generation ratef(t) can be written as
(61)
and following Fitch and Tsai (1971), the homogeneous rate constant is
assumed to decrease with partic1e area as k.h= kho(1- ApL/4). Applying the
stationary-state hypothesis to the radical balance gives the radical concentration in the wter phase as
(62)
where the rate of appearance p(t) of free radicals in the water phase is given
by

(63)
Assuming that radical termination takes place exc1usively in the polymer
partic1e,the average number of radicals in a partic1eof volume vp(t,-r)can
be obtained as (Ugelstad and Hansen, 1976)
q(t, -r) =

1
2kde(t, -r)n(t,-r)d-rJ

1/2

a (t -r) n(t -r) d-r

p"

Ap(t)

1/2

(64)

342

A. E. Hamielec and J. F. MacGregor

Using Eq. (14) the desorption rate constant can be expressed as


kde(t, ,) = kge/ap(t, ,)
which when substituted

into Eq. (64) yields

q(t,,)

R,

1/2

2kodeJ

1/2

a;(t,')

A p(t)

(65)

Substituting these i~to Eq. (63) and integrating gives


p(t) = R,(t) + (!kgeR,)1/2(N(t)/A~/2(t))

(66)

The partic1e generation rate f(t) can now be evaluated using Eqs. (61), (62),
and (66),in conjunction with the property balance equations (56-60).
The rate of polymerization Rp [see Eq. (2)] can be expressed here as
Rp(t) = (kp[Mp]/NA)
LtJ(t,,)n(t,,) d,
= (kp[Mp]/NA)(RJ2kge)1/2
A~/2(

(67)
(68)

The time function ~(t) in the property balance equations (57-59) is given by
kpdm R,
~(t) = NAdp 2kge

1/2

( )

cI>m(t)

A~/2(t) 1 - cI>m(t)

(69)

where cI>m(t)
is the monomer volume fraction in the polymer partic1es given
earlier in Eq. (23).
The property balances must be coupled with the balances for the
initiator and emulsifier given, respectively, by
d[1]/dt

= e-1([1F]- [1]) - R,

(70)

where R, = 2fkd[1], and

d[S]/dt = e-1([SF] - [S])

(71)

The micellar area Am(t) needed in Eqs. (61) and (62) can then be obtained as
Am(t) = ([S]

[Scmc])as- Ap(t)

(72)

where as is the coverage area of 1 mole of emulsifier.

The set of simultaneous nonlinear differentialequations defined by Eqs.


(56-60) and Eqs. (70)and (71)therefore provides a dynamic model for the
total latex properties of PV Ac being produced in a single eSTRo This
model has been shown to provide an excellent representation of the
conversion-time experimental data of Greene et al. (1976) under conditions
of sustained oscillations (see Fig. 7) and of the extensive data of Kiparissides
et al. (1980a,b) under both steady-state and sustained-oscillation conditions

343

9. Latex Reactor Principies: Design, Operation, and Control

30

t520
;

Q:
ILI

~
15
o

<..>

10
1]= 0.005

6
8
10
12
14
DIMENSIONLESS TIME (t / e)

16

18

20

Fig. 7. CSTR conversion transients. Comparison between experimental results of


Greene et al. (1976) and model predictions. Model (---):
Greene's experimental results
(-0-0-).

using sodium lauryl sulfate as an emulsifier and potassium persulfate as


initiator. Figure 8 illustrates the behavior of the number of particles, the
particle area, and the free soap or micellar area predicted by the model for
one of Greene's runs. The reasons for the s4stained oscillation phenomenon
are quite apparent from these figures. In periods where kmAm or
khO(1 - ApL/4) are greater than zero, rapid generation of particles occurs
leading to a large surface area and the subsequent depletion of free
emulsifier. Micellar particle nucleation therefore stops and, with the explosive increase in particle are a (Ap), so does homogeneous nucleation.
There follows a long period, the duration of which depends on the
emulsifier feedrate and residence time of the reactor, in which particles are
not generated. However, as the washout of existing particles continues, the
emulsifier concentration builds up again to exceed the CMC, and a new
.generation of partcles is formed. This periodic nucleation leads to the
formation of discrete particle populations with concomitant oscillations in
polymerization rate, conversion, and latex properties.
In other situations where the emulsifier feedrate to the reactor is
sufficient to produce a steady-state concentration in the CSTR that is above
the CMC, one observes damped oscillations upon start-up followed by an
eventual attainment of a steady-state, nonoscillatory condition (Kiparissides

et al., 1980a,b).

344

A. E. Hamielec and J. F. MacGregor

2.0
re
TQ
..
.....
..
...

= 1.5
......
C/)
UI

.J
o
~

1.0

el.
la.
O
a:
UI
m
:!! 0.5
:)
z
.J

~
O

1-

O
O

6
8
10
12
14
DIMENSIONLESSTIME (tIa)

16

18

Fig. 8. CSTR transients.Model predictions for number of particles, free emulsifier


concentration and total particle surface area. 6 = 30min, 1 = 0.01molJliter,S = 0.01molJliter,
MjW = 0.43.

C.

Cons;derat;on 01 Dynam;cs ;n Reactor Des;gn

These unsteady-state experiments and the dynamic models developed to


explain the observed behavior ha ve a number of important implications for
the design of continuous emulsion polymerization reactor trains. In order
to avoid this oscillation phenomenon and the varying latex quality that
results from it, one should not design continuous flow stirred-tank reactor.
trains along previous lines, namely with a number of equal or nearly equalsized CSTRs in series with all the feed streams of the receipe entering the
first reactor. Most continuous industrial systems designed in this manner
exhibit oscillations in the early reactors of a magnitude comparable to those
shown in Fig. 7. Although the oscillations may be largely damped out by
the later reactors in the train certain inhomogeneities will exist in the latex.
A common industrial practice used to avoid oscillations is to seed the
first reactor with small-diameter seed particles produced earlier in batch
reactors. By then keeping the emulsifier concentration below the CMC one
avoids further micellar generation and simply grows the seed particles.
Although in some cases this may introduce additional flexibility in that a

9.

Latex Reactor Principies: Design, Operation, and Control

345

cheaper or otherwise more desirable polymer may be used for the seed
partic1es, it usually results in costlier operations resulting from the need for
batch reactors to produce the seed partic1es.
This oscillation phenomenon, however, can be avoided quite easily
without the need for seeding. By employing as the first reactor in the train a
small, optimally sized CSTR reactor and splitting the emulsifier, initiator,
water, and monomer feeds to enter at various points along the train it is
quite possible to ensure that all partic1e generation occurs in the first small
reactor under nonoscillatory
or steady-state conditions, followed only by
growth of these partic1es in the subsequent reactors (Pollock, 1981;
Nomura, 1980). As a simple illustration one could feed most of the
emulsifier to the small first CSTR "seeding" reactor along with part of the
initiator and part of the water and monomer, thereby ensuring continuous
partic1e generation in condition of excessemulsifier. By feeding much of the
remaining water and monomer to the second reactor, the emulsifier is
diluted to a value below the CMC at which partic1e generation is avoided
and partic1e growth only is promoted. A good dynamic model is the key to
sizing the reactors determining an optimal split of the feed streams. As will
be discussed later, the dynamic model is also the key to designing an on-line
monitoring and control system that will ensure that the desired conditions
of latex quality and quantity are maintained.
It is also possible, and in some cases more desirable,to use a continuous

tubular reactor as the first "seeding" reactor (Nomura, 1980). Oscillations


should not be a problem in such a reactor, but reactor fouling might be a
more important consideration.

V. On-Line Control of Continuous Latex Reactors

A. Introduct;on
As indicated above a number of fundamental control problems with
cOQtinuous emulsion polymerization reactors are associated with the initial
.

reactor systemdesign,and such problems should, therefore,be solved at the


design stage. However, even with an efficient design some on-line monitoring and control are needed to ensure that the design conditions are indeed
being maintained in the face of numerous disturbances which can affectthe
system.In this section we first consider some of the on-line instrumentation
and techniquesfor monitoring reactor condition, and then we look at some
of the techniques that are being used fOr"the control of polymer latex
properties in these reactors.

346
B.

A. E. Hamielec and J. F. MacGregor

On-Line Instrumentation

The most useful on-line instruments would be those capable of measUfing the latex properties given in Eqs. (56-60), the free soap concentration
(Am), and some molecular weight characteristics. In this section we will not
attempt on exhaustive survey, but will only concentrate on a few of the
more rugged and more promising on-line measurements that have been
used.

1. Reactor Heat Balances


One of the earliest attempts at monitoring the progress of emulsion
polymerization reactors was to perform an on-line enthalphy balance
around the reactor or the reactor cooling jacket in order to monitor the
heat release by reaction. Using temperature measurements in the reactor, in
the cooling jacket entrance and exit, and at any ~ther essential points,
together with relevant cooling water flows one can perform steady-state or
unsteady-state enthalpy balances using an on-line minicomputer. From the
computed rate of heat release, the rate of reaction can be evaluated, and the
conversion in the reactor can be followed using Eq. (60).
A major problem in successfully implementing this scheme is to overcome the effects of the numerous measurement errors that propagate into
the ca1culated reaction rate and conversion. For any given situation a
simple propagation of error analysis will reveal the precision possible.
Usually, significant filtering of the measurements is necessary in order to
obtain worthwhile results. Two approaches are possible, both requiring an
on-line mini- or microcomputer. In the first approach instantaneous steadystate balances can be performed over short periods using averaged or
filtered values of many measurements. Alternatively, less-frequent measurements can be utilized with the unsteady-state balance equations in the form
of a Kalman Filter (Astrom, 1970; Jazwinski, 1970).
2. Density
A number of fairly rugged on-line instruments are available to follow the
emulsion density variations. Examples include nuclear instruments and
instruments based on mechanicaloscillator techniques(Kratky et al., 1973).
By utilizing the density difference between the unreacted monomer and the
polymer (providing a reasonable difference exists) the reactor conversion
ca,n then be ca1culated via

x = (Pe- p~)/(p:oo- p~)

9.

347

Latex Reactor Principies: Design, Operation, and Control

where the emulsion densities at O and 100% conversion (p~ and


be approximated

P:OO)

may

by the weighted averages

o_

Pe

{'
- 1.mPm + JwPw

100 _
Pe
- 1.mPp

{'

+ JwPw

Here fm and fw are the feed stream weight fractions of monomer and water,
and Pm, Pw, and Pp are the densities of, respectively, monomer, water, and
polymer. Using the new on-line densitometers which rely on vibrational
frequencies, a precision of approximately :t 0.5% on conversion can be
reasonably attained in the polymerization of methyl methacrylate (Schork
and Ray, 1980) or vinyl acetate (Pollock, 1981).

3. Turbidity Spectra
Light transmission has been a standard method for the measurement of
the size of spherical particles for many years. Mie and Rayleigh scattering
theories show how the measured turbidity can be related to particle
diameter and number. In the Mie regime (1 < nD/Am < 10) information on
the number of particles and the size distribution could be obtained,
particularly if used in conjunction with other instruments such as a
densitometer (Maxim et al., 1969). In the Rayleigh regime (nD/Am< 1) only
a turbidity-average particle diameter is easily obtained. From the point of
view of on-line process measurements, turbidity instruments are rugged and
have been demonstrated to follow easily the oscillations in average particle
diameter in continuous emulsion polymerizations (Kiparissides et al.,
1980a). For example, Fig. 9 shows a plot ofthe absorbance at 350 nm and the
0.4
o~
o
o

10
0.31<')

o;t 3000
a:

IIJ
1IIJ
~ 2000
~
iS
IIJ
..J
U
: 1000
a:
~
Q.

w
u
0.2~
al
a:
o
0.1

10

12

14

(f)
al
~

o
16

DIMENSIONLESS TIME (t181

Fig. 9.

Particlediameterand absorbancemeasuredduring the production of poly(vinyl

acetate) in a eSTRo O = 20 min; 1

= 0.01

mol/liter;

= 0.01

moljliter.

348

A. E. Hamielec and J. F. MacGregor

average particle diameter as determined by size-exclusion chromatography


measured on poly(vinyl acetate) latices being produced under conditions
of sustained oscillations in a single eSTRo
4.

Surface Tension

Recently, Schork and Ray (1980) developed an on-line tensiometer based


on the maximum blibble pressure technique and demonstrated its usefulness
in following surface tension in methyl methacrylate batch-emulsion polymerization. The value of this surface tension measure for monitoring emulsion
polymerization reactors lies in being able to combine it with a measure or
knowledge of the monomer concentration in the water phase to give a
measure of the free emulsifier concentration (Am) in the water phase. As
discussed previously (Sections IV,B and C), the free emulsifier concentration
Am is of the utmost importance in the dynamics of particle generation in
continuous emulsion polymerizations, and being able to monitor it on-line
would be a very important step ip.the control of these reactors.
C.

On-Line Estimation

Realistically, one could never hope to have on-line measurements of all


tge major properties of lnterest but would have to rely perhaps on one or
two of the above. However, herein les the utility of a good dynamic model
for the process, in that, by using the measurements that are available, the
remaining properties or states of the system can be inferred froIDthe on-line
solution of the model equations. Such state estimation can be accomplished
by using an extended Kalman Filter (Jazwinski, 1970), which essentially
combines the model equations with the measurements to yield a set of
recursive equations providing estimates of all the states. The state measurements that are made serve to keep the model tracking those particular
states closely, thereby forcing reasonable estimates of the other states,
providing the model is good.
Using an earlier version of the dynamic model given in Section IV,B,
Kiparissides et al. (1980b) illustrated the use of such an extended Kalman
Filter to infer N(t), Vp(t),Ap(t), and X(t) from measurements taken only on
conversion [X(t)] using UV turbidity spectra. Jo and Bankoff (1976) used
these filters to track some of the moments of the MWD of PVAc in a
solution polymerization process using measurements made on refractive
index and viscosity.

9. Latex Reactor Principies: Design. Operation. and Control


D.

349

On-Line Control

A number of different approaches to the computer control of emulsion


polymerization reactors can be taken, depending on the level of modeling
and control theory sophistication one wishes to use. Most batch or
semibatch emulsion polymerization control systems consist of a preprogrammed recipe addition, a start-up and shut-down scheduler, temperature
and pressure controls, and perhaps a heat balance control for reaction rate
(Amrehn, 1977), but no feedback control on the latex properties. ExactIy
what can be done in the way of feedback control in these systems is an area
of current research. One is essentiaIly limited by the fact that most of the
particIe generation occurs in the first 5 to 15 min and there is very little one
can do over the remainder of the batch to compensate for things that went
wrong at this stage. In continuous reactor trains particIe generation is
continuous, or at least periodic, and it can be monitored and adjustments
can be made in the recipe ingredient ftows to the system and in their splits
between the various reactors of the train. Two possible approaches are
discussed below.
1. Advanced Control Based on Theoretical Dynamic Models
To ilIustrate this approach we consider the emulsion polymerization of
vinyl acetate in a single eSTRo The dynamic model given by Eqs. (56-60),
and Eqs. (69) and (70) can be represented more concisely by the set of
nonlinear differential equations

. dx(t)/dt

= f[x(t), u(t)] + w(t)

(73)

where the vector of states x = (N, Dp, Ap, Vp,X, 1, S)Tand the vector of
manipulatable variables u = (IF, SF)' The vector of measurements can be
expressed generaIly as
y(t) = f[x(t), u(t)] + v(t)

(74)

where the vector y(t) and the function f[x(t), u(t)] wilI depend upon which
on-line instruments are available. The vectors x(t) and w(t) represent,
respectively, measurement noise and process disturbances arising from
other variations in the process not accounted for in the modeI. The
variance-covariance structure of these noise sources can be identified from
process data taken from the system.
Using this model of the reactor system, usuaIly discretized for some
sampling interval, near-optimal control algorithms can be formulated which

350

A. E. Hamielec and J. F. MacGregor

will manipulate the input vector at time instant (UI)'in such a way as to try
to optimize some performance index involving the system states (XI)and the
manipulations (UI)over some interval of time. The most commonly chosen
performance index is the quadratic one
N

J(x,u)= E

L [(XI 1=1

Xdes)TQl(XI

xdes)

+ (UI-l- Udes)TQiul-l- Udes)]

where E denotes the expectation, xdes and Udesare the desired operating
levels, and Ql and Q2 are positive semidefinite matrices. For example, this
would inelude the case where it is desired to minimize a weighted sum of the
mean squared errors of the states about their desired values subject to a
constraint on the magnitude mean squared errors of the manipulated
variables about their target values.
Kiparissides et al. (1980b) illustrated the use of this approach, using a
simulation of their single CSTR for the emulsion polymerization of VAc.
They looked at control during reactor start-up, during switching from one
operating condition to another, and at trying to reduce the magnitude of
the oscillations in partiGR: properties while operating in the oscillation
region (i.e., at low emulsifier concentrations). Note that in the latter
situation one cannot entirely eliminate the oscillations since partieles must
be generated sometime.
To implement such near optimal control algorithms would require
considerable space and time on a process minicomputer because of the
severe nonlinearities in the process, and therefore it is probably impractical
in most situations. However, off-line simulations of this control policy
shows some upper-limits as to what could possibly be accomplished and the
nature and magnitudes of the feedrate manipulations that are necessary.
This can then lead to the development of simpler heuristic-type control
algorithms which capture the important aspects of the optimal control but
are more easily implemented on-line.
Some simpler approaches to conversion control using material and
energy balances on a train of CSTRs are described by Amrehn (1977).

2. Control Based on ldentified Empirical Models


In some situations where one or more of the latex properties are
measured either directly or indirectly through their correlation with sur.
rogate variables and where extreme nonlinearities such as the periodic
generation of polymer partieles does not occur, one can use much simpler
modeling and control techniques. Linear transfer function-type models can
be identified directly from the plant reactor data. Conventional control
devices such as PID controllers or PID controllers with dead-time compensation can then be designed. If process data is also used to identify

9. Latex Reactor Principies: Design, Operation, and Control

351

models for the disturbances in the system, optimal controllers can also be
developed (Box and Jenkins, 1970; Astrom, 1970).
MacGregor and Tidwell (1979) iIIustrate some of the steps involved in
running plant experimentation, building these process and disturbance
models, and implementing simple optimal controllers on some continuous
condensation polymerization processes. A number of similar applications to
continuous emulsion polymerization processes have also been made.
VI.

Summary

In this chapter an attempt has been made to review the state of the art of
the design, operation, and control of latex reactors and to introduce some
promising new approaches to the design of latex reactor trains using
dynamic models. Highly sophisticated dynamic models which can accurately predict particIe number and size distribution and also molecular
weight averages and long-chain branching frequency are being developed.
These dynamic models should permit the design of a new generation of
commerciallatex reactor systems.
Nomenclature
as
ap(t,T)
Ap, Ap(t)
Am
Aw
A,B
a,b
BN
CM,Cp, K
dM
dp
dp(t,T)
!?)w
!?)p
DpDp(t)
E
E(t)
F(R)
fm.!",
f(t)
[l]
[/]0
[/F]

area occupied by an emulsifier molecule


area of polymer particles born at time T arid present at time t
total area of polymer particles per liter of latex
total area of micelles per liter of latex
jacket heaf transfer area
parameters in Eqs. (19) and (20)
parameters in Eq. (21)
number average number of long branches

per polymer molecule

kinetic parameters defined by Eq. (28)


monomer density
polymer density
diameter of polymer particles born at time T and present at time t
diffusion coefficient of monomer transfer radicals in the water phase
diffusfon coefficient of monomer transfer radicals in polymer particles
total diameter of polymer particles per liter of latex
xpectation operator
exit age distribution function
frequency distribution of polymer particle radius
weight fractions of monomer and water in feed stream
particle nucleation rate
initiator concentration
initial initiator concentration
initiator concentration
in feed

352
K
KI
kp
k,p
k,w
krm
kfx
krp
k*p

k,o
k.
k~.
kd.
km
kv
kh
L
m
M
MR
MN
Mw
Mw.,
[MpJ
[MFJ
Np
NA
Nn
NMo
NM

N,N(t)
n

. n(t, t) dt
p{t, t)
P(t)

A. E. Hamielec and J. F. MacGregor


defined in Eq. (43b)
defined in Eq. (45b)
propagation constant in the polymer particles
termination constant in the polymer particles
termination constant in the water phase
rate constant for transfer to monomer
rate constant for transfer to chain transfer agent
rate constant for tran~fer to polymer
rate constant for terminal double bond reaction
terminaton constant when reaction is chemically controlled
rate constant

for radical

absorption

into

a particle

of radius

defined in text
kd.(t, t) is the.rate constant for radical desorption from a particle of volume v (or
born at time t and present time t)
rate constant for micellar particle generation
volume of particle phase over volume of water phase
rate constant for homogeneous particle generation
distance a radical will dilfuse before precipitating
partition coefficient in Eq. (14)
molecular weight
monomer molecular weight
number average molecular weight
weight average molecular weight
weight average molecular weight when termination

becomes dilfusion controlled

monomer concentration in the polymer particles in GmolJliter


monomer concentration in the feed
number of polymer particles per liter of latex
Avogadro's number
number of polymer particles containing n radicals(n ~ O)
initial moles of monomer
moles of monomer at time t
number of polymer particles per liter of latex
number of radicals
average number of radicals per particle
number of particles born at time t and present at time t
property associatedwiththe n(t,,) d,
total property

class of

particles

Qo, Ql' Q2
Ql'Q2
QE
ij(t, t)

zeroeth, first, and second moments of the molecular weight distribution


are positive semidefinite matrices in the control performance index
heat removed by coils, condenser or other device

Rpp
RI

monomer

Rp
Rpma.
R

[8]0

polymerization
rate in Gmoljsec liter latex
maximum polymerization
rate
polymer particle radius and gas constant
radical concentration
in water phase
concentration
of emulsifier
initial concentration
of emulsifier

[SmJ

concentration of emulsifier present as micelles

[KJw
[8]

average number of radicals in particles born at time, present at time t


consumption

rate in Gmoljsec

particle

rate of initiation of radicals in the water phase in Gmolejsec liter latex

9. Latex Reactor Principies: Design, Operation, and Control


[Sp]
[SCMcJ
T
T.
t
tr
Tam,1'.p
U
u" u(t)
Vp
VF
VFe"
VF..,
Vp(t,T)
V
V(t)

concentration
concentration

of emulsifier adsorbed
of emulsifier at CMC

353

on polymer partic1es

absolute temperature, polymerization


temperature
coolant temperature
polymerization
time
time to the end of Stage I
are glass transition temperatures for pure monomer
overall heat transfer coefficint

and polymer

vector of manipulatable
variables (Ud.. is desired operating level)
volume of polymer partic1es per liter of latex
free volume fraction
free volume fraction when termination becomes diffusion controlled

~
W(M)
W(t)
X
X.
[X]
x"x(t)
y(t)
c<,o:l,m,Y

Cl
1>m,1>m(t)
1>p, 1>p(t)

S(t)
f1
T
lJ

P., p~, p:OO


p'
p(t)
-Hp
IN

free volume fraction when propagation becomes diffusion controlled


volume of latex partic1es generated time T and present at time t
volume of latex, reactor volume
vector of measurement noise
weight fraction polymer in polymer partic1es
weight fraction of polymer of molecular weight M
vector of process disturbances
monomel' conversion
monomer

conversion

at the end of Stage 11

concentration
of chain-transfer agent in polymer partic1es
vector of states (xd.. is the desired operating level)
vector of measurements
kinetic parameters defined in Eq. (8)
differences between coefficients of volumetric expansion above
(polymer and monomer)
exponent in model for diffusion-controlled
termination [Eq. (19)]
volume fraction monomer in polymer partic1es
volume fraction polymer in polymer partic1es.
defined in text

and

below

1'.

volumetric growth rate


time of birth of a polymer partic1e
defined by. Eq. (14a)
emulsion densities (defined in text)
is the total radical entry rate into polymer partic1es of volume v per liter of latex
rate of appearance of radicals in the water phase
heat of polymerization
is the number average number oflong branches per 1000 monomer units

References
Amrehn, H. (1977). Automatica 13, 533-545.
Astrom, K. J. (1970). "Introduction
to Stochastic Control Theory." Academic Press, New York.
Box, G. E. P., and Jenkins, G. M. (1970). "Time Series Analysis, Forecasting and Contro1."
Holden Day, San Francisco.
Brooks, B. W., Kropholler,
H. W., and Purt, S. N. (1978). Polymer 19, 193-196.

354

A. E. Hamielec and J. F. MacGregor

Chiang, A., and Thompson,


R. W. (1979). Am. Inst. Chem. Eng. J. 25, 552-554.
Dickinson, R. F. (1976). Ph.D. Thesis, Univ. ofWaterloo.
Edelhauser, H. (1975). In "Polymer
Norway.

Colloids."

NATO Advanced

Fitch, R. M., and Tsai, C. H. (1971). In "Polymer


Plenum Press, New York.
Friis, N., and Nyhagen,
Friis, N., and Hamielec,

Colloids"

Study lnstitute,

Trondheim,

(R. M. Fitch, ed.), pp. 103-115.

L. (1973). J. Appl. Polym. Sci. 17, 2311-2327.


A. E. (1975). J. Appl. Polym. Sei. 19,97-113.

Friis, N., ap.d Hamielec, A. E. (1976). In Polymer Reaction Engineering-Principles


ofPolyrner
Reactor Design. McMaster Univ. Professional Development Course.
Friis, N., and Hamielec, A. E. (1976). In" Emulsion Polymerization"
(l. pirma and J. L. Gardon.
eds.), (Am. Chem. Soco Symp. Ser.) Vol. 24. pp. 82-91.
Friis, N., Goosney, D., Wright, J. D., and Hamielec, A. E. (1974). J. Appl. Polym. Sci. 18,
1247-1259.
Gerrens,
Gerrens,

H. (1959). Fortschr. Hochpolym. Forsch. 1, 234-246.


H., and Kuchner, K. (1970). Br. Polym. J. 2, 18-30.

Gershberg,
AIChE
Gilbert,

D. B., and Longfield,


Preprims.

R. G., Napper,

J. E. (1961). In Polymerization

Kinetics and Catalyst

D. H., Lichti, G., Ballard, M., and Sangsterr

Systems.

D. F. (1980). In Emulsion

Polymerization, pp. 527-530. Am. Chem. Soco Preprints, Las Vegas.


Greene, R. K., Conzalez, R. A., and PQehlein, G. W. (1976). In "Emulsion Polyrnerization"
(l.
Piirma and J. L. Gardon, eds.), pp. 341-358. Am. Chem. Soco Ser. 24, Washington, D.C.
Hamielec, A. E., and Marten, L. (1980a). In "Emulsion Polyrnerization," pp. 515-519. Am.
Chem. Soco Preprints, Las Vegas.
Hamielec, A. E., Gomez-Vaillard,
R., and Marten, L. (1980). Diffusion-controlled
free radical
polyrnerization-effect
on polymerization
Symp., 3rd, Cleveland, Ohio.
Jazwinski,
York.

A. H. (1970). "Stochastic

rate and molecular

Processes

and Filtering

properties

Theory."

of PVC, 1m. PVC

Academic

Press, New

Jo, J. H., and Bankoff, S. G. (1976). Am. Inst. Chem. Eng. J. 22, 361-369.
Kiparissides,

c., MacGregor,

J. F., and Hamielec, A. E. (1978). J. Appl. Polym. Sci. 23, 401-418.

Kiparissides, C., MacGregor, J. F., and Hamielec, A. E. H. (1980a). Can. J. Chem. Eng. 58,48-71.
Kiparissides,
(5).

c., MacGregor,

J. F., and Hamielec,

A. E. H. (l980b).

Am. Inst. Chem. Eng. J. 26,

Kratky, O., Leopold, H., and Stabinger, H. (1973). Methdds Enzymol. 27, 98-110.
Ley, G., and Gerrens, H. (1974). Makromol. Chem. 175, 563-575.
Lichti, G., Gilbert, R. G., and Napper, D. H. (1977). J. Polym. Sci. 15, 1957-1969.
Lichti, G. (1980). PhD Thesis, Univ. of Sydney.

MacGregor, J. F., and Tidwell, P. W. (1980). In "Computer Application to Chemical


Engineering" (1. Squires and l. Reblaitis, eds.), pp. 251-268. Am. Chem. SocoSymp. Ser. 124.
Marten, L., and Hamielec, A. E. (1978). In "Polymerization Reactors and Processes" (J. N.
Hendersonand T. C. Bouton, eds.), pp. 43-70. Am. Chem. Soco Soco Ser. 104.
Maxim, L. D., Klein, A., Mayer, M. E., and Kuist, C. H. (1969). J. Polym. Sci. Part C 27,
195-207.
Min, K. W., and Gostin, H. 1. (1979). Ind. Eng. Chem.-Prod. Res. Dev. 18,272-284.
Min, K. W., and Ray, W. H. (1974). J. Macromol. Sci.-Rev. Macromol, Chem. Cll, 177-255.
Nomura,

M. (1980). In "Emulsion

Polyrnerization,"

pp. 822-823. Am. Chem. Soco Preprints,

Las

Vegas.
Nomura, M., Kojima, H., Harada, M., Eguchi, W., and Nagata, S.(1971).J. Appl. Poly~. Sci.15,
675-695.

9. Latex Reactor Principies: Design, Operation, and Control

355

Nomura, M., Harada, M., Eguchi, W., and Nagats, S. (1976). In "Emulsion Polymerization" (1.
Piirma and J. L. Gardon, eds.), pp. 102-121. Am. Chem. Soco Symp. Ser. 24.
O'Toole, J. T. (1965). J. Appl. Polym. Sci. 9, 1291-1302.
Pollock, M. (1981). PhD Thesis, McMaster Univ.
Schork, F. J., and Ray, W. H. (1980). In "Emulsion Polymerization," pp. 823-828. Am. Chem.
Soco Preprints, Las Vegas.
.
Smith, W. V., and Ewart, R. H. (1948). J. Chem. Phys. 16, 592-606.
Sundberg, D. c., Soh, S. K., and Hsieh, J. Y. (1980). In" Emulsion Polymerization," pp. 520-525.
Am. Chem. Soco Preprints, Las Vegas.
Ueda, T., Omi, S., and Kubota, H. (1971). J. Chem. Eng. Jpn. 4, 50-63.
Ugelstad, J., and Hansen, F. K. (1976). Rubber Chem. Technol. 49, 536-561.
Ugelstad, J., M0rk, P. C., and Aasen, J. E. (1967). J. Polym. Sci. A-I 5, 2281-2296.
Ugelstad, J., M0rk, P. c., Dahl, P., and Rangnes, P. (1969). J. Polym. Sci. C 27,49-60.
Warson, H. (1976). In "Emulsion Polymerization" (1. Piirma and J. L. Gardon, eds.), pp.
228-235. Am. Chem. Soco Symp. Ser. 24.
Wright, D. E., and Tucker, H. (1977). In Polymer Reaction Engineering-Process TechnologySynthetic Elastomers. McMaster Univ. Professional Development Course.

10
Emulsion Polymerization
in Continuous Reactors
Gary W. Poehlein

1. Introduction .
A. Background Information .
B. Reactor Configurations and Flow Possibilities .
C. Differencesbetween CSTRsand BatchReactors
D. Scope of This Chapter .
11. Smith-Ewart Case 2 Model for a CSTR .
A. Particle Formation .
B. Particle Growth .
C. ParticleAge and Size Distributions .
D. Number of Particles and PolymerizationRate .
E. Comparison with a Batch Reactor
F. Effect of Inhibitors in Feed Streams .
G. Use of a Seed-Particle Feed Stream .
111. Deviations from Smith-Ewart Case 2.
A. Large Particles or Slow Termination Rates .
B. Radical Desorption .
.. .
C. Other Particle Formation Mechanisms .
D. ExperimentalResults
IV. Transient Behavior of CSTR Systems.
A. ExperimentalObservations.
B. Physicochemical Mechanisms
C. Control Methods
.
V. Strategies for Process and Product Development .
A. Reactor Design Considerations .
B. Product Development Considerations
VI. Summary.
References

357
357
358
359
361
361
361
362
363
363
366
367
367
367
368
369
370
374
375
375
375
378
378
378
380
381
381

l. Introduction

A.

Background lnformation

Three major types of chemical reactor systems are used to


produce emulsion polymers; batch, semicontinuous, and continuous. Batch
reactors usually consist of stirred tanks with various forms of heat removal
357
EMULSION POLYMERIZATION
Copyrigbt ~.1982 by Aeademie Pres" Ine.
All rigbts of reproduetion in any form rescrved.
ISBN 0-12-556420-1

358

Gary W. Poehlein

devices such as jackets, reflux condensers, internal cooling surfaces, and/or


external heat exchanger loops. All of the recipe ingredients are added at or
near the beginning of the batch cycle, the reaction is carried out to the target
conversion, and the reactor contents are discharged for further processing.
Most kinetic studies and models are based on batch reactor data and batch
operation.
Semicontinuous (semibatch is probably a better name) reactors are
designed like batch reactors, but with the former Qot all of the recipe
ingredients are added at the beginning of the reaction 'cycle. Part of the
monomer, and sometimes other ingredients as well, are added in a programmed manner after the reaction has started. Semicontinuous operation
can be used to control reaction rate, copolymer composition drift, particle
size distribution, particle morphology, and perhaps other product propertieso Semicontinuous reactors are widely used for coatings and adhesive
products.
Continuous reactors are operated with continuous, and it is hoped,
steady input flows of reagents and output flows of products. Such reactors
are generally economically advantageous when high production rates of
closely related products are required. Continuous systems are not used if
rather long run times cannot be achieved. Thus, latexes that foul badly and
cause frequent shutdowns are usually produced in batch or semicontinuous
reactors. Likewise, continuous reactors are not practical for product distributions which require frequent, significant recipe changes.
Commercial continuous processes have been reported for synthetic
rubbers [SBR: Wolk, 1959; Owen et al., 1947; Lundrie and McCann, 1949;
Feldon et al., 1953 and CR (Chloroprene Rubber): Calcott and
Starkweather, 1954; Aho, 1958] and a number of other polymers (Poehlein
and Dougherty, 1977). As sales volumes have increased a number of
industrial organizations have become interested in continuous processes
and current activity in this area is significant.
B.

Reactor Configurations and F/ow Possibilities

Commercial continuous reactor systems generally consist of a number of


continuous stirred-tank reactors (CSTRs) connected in series. The reagents
are pumped into the first reactor and the product is removed from the last.
Heat is exchanged through reactor jackets and internal cooling surfaces.
Commercial SBR systems sometimes consist of as many as 12-15 equalsized reactors connected in series. Reactor sizes are 2500-3500 gal with
total mean residence times of about 8 hr. More recent SBR systems consist
of fewer reactors and some commercial systems contain CSTRs that are not
all the same size.

10. Emulsion Polymerization in Continuous Reactors

359

Continuous tubular reactors can also be used to produce emulsion


polymers. Such reactors have been used in series with CSTRs (Gonzalez,
1974), as flow-through reactors (Rollins et al., 1979; Ghosh and Forsyth,
1976) and in a continuous loop process (Lanthier, 1970) in which material is
fed and removed from a tubular loop with a circulating flow greater than
the throughput.
A tubular prereactor, in series with CSTR system, can offer stability
advantages, which will be discussed later. A number of other flow alternatives are also possible with a CSTR-series system but these alternates are
not widely utilized. An obvious flow alternative for a reactor system
consisting of a series of CSTRs would be to introduce some portion of the
total recipe at places other than the front end of the reactor train. These
intermediate feeds would, in many respects, be analogous to semicontinuous
operation of batch reactors.
Intermediate feeds of monomer, water, emulsifier, etc., could be used to
control particle generation, copolymer composition, particle morphology,
particle size distribution, and molecular structure. The use of intermediate
feed streams andjor recycle streams has not been studied extensively. This is
an area in which process advances might be made.
C.

Differences hetween CSTRs and Batch Reactors

The performance of a single CSTR can be quite different from that of a


batch reactor for a number of reasons. First, the distribution of reactor
residence times in a CSTR is quite broad. This leads to broad size and age
distributions of the latex particles. By contnist, the polymer particles in a
batch reactor are usually all formed near the beginning of the reaction and
the particle size and age distributions of the product latex are narrow.
A second important difference concerns reactor behavior in response to
inhibitors andjor retarders in the feed streams. In a batch reactor these
compounds react near the beginning of the cycle and then the polymerization proceeds in a near normal manner. In a CSTR, inhibitors in the feed
streams are continuously added to the reactor and serve to reduce the rate
of initiation of poly,mer molecules.
Copolymer compsition is a third area of contrast. Copolymer composition in a batch reactor tends to change with time. The first polymer
formed is rich in the more reactive monomer and the final polymer contains
more of the least reactive monomer. This drift in composition can lead to
polymer particles with nonuniform composition in the radial dimension.
The copolymer product formed in a single CSTR, however, should be
relatively uniform in composition if the reactor is operated at steady state.
Ir several CSTRs are connected in series, polymers of several different

360

Gary W. Poehlein

compositions can be formed. Sincemost partic1eswould be nuc1eatedin the


first reactor, the final partic1esfrom a multireactor system would also be
nonuniform in composition. The consumption of chain-transfer agents can
be influenced by the same phenomena that effect comonomers. Thus, their
performance may differ with reactor type.
A fourth concern in operating a CSTR involvesthe addition of the feed
streams. All ingredients are added near the beginning of a batch reaction
cyc1e,before the partic1es are formed. Recipe ingredients are usually divided
among sveral feed streams in continuous systems. In this case the mixing
should be sufficient to blend these streams rapidly with the polymerizing
mixture in the reactor. The addition of the initiator stream can be especially
critical. The water-soluble initiation systems normally used for emulsion
polymerizations are electrolytes, and they can cause significant flocculation
if they are highly concentrated andjor if they are not introduced to the
reactor in a proper manner and mixed quickly. Monomer and monomer
emulsion streams must also be added and mixed properly in order to
prevent situations in which mass transfer factors may control the polymerization rateo
A fifth, and commercially important difference between CSTRs and
batch reactors is heat removal. Heat generation in a batch polymerization
can be highly nonuniform. Sometimes the reactor productivity is dependent
entirely on the capacity to remove heat during the peak reaction periodo
Heat removal in steady-state, continuous systems is uniform and, for this
among other reasons, higher productivity can often be achieved.
Reactor fouling is a sixth area wherein continuous systems can perform
differently. The batch cyc1e involves filling and emptying the reactor
frequently. A disadvantage of this operation is that the latex can dry or
stick to the walls and initiate the formation of wall polymer. The fact that
the reactor is emptied frequently, however, does provide an opportunity for
c1eaning. A properly designed continuous reactor system should operate full
of latex, without a vapor space. This helps to prevent or retard the
formation of wall polymer. In fact the absence of a liquid-vapor interface
may slow the formation of coagulum in the latex. Heller and Peters
(1970a,b, 1971; Heller and De Lauder, 1971a,b) have c1early demonstrated
that interfacial flocculation can be significant. One disadvantage of a
continuous system is that operation in a particular reactor must be stopped
for c1eaning.If frequent disruptions are required continuous reactor systems
will probably not be economically feasible.
A seventh difference between batch reactors and CSTRs is concerned
with molecular weight, branching, and cross-linking characteristics. In
general, branching, cross-linking, and gel-effect phenomena become more
significant at higher conversions. The final product from a batch reactor

10.

Emulsion Polymerization in Continuous Reactors

361

consists of molecules formed over a rather broad range of conversion; this


is less true in emulsion polymerization than in bulk or solution polymerization since the conversion at the reaction site in the polymer particles may
be quite high even through the overall reaction conversion is low. In a
CSTR, however, the polymer is formed at a constant conversion, that
existing under the state of operation. Usually this will be higher than the
average conversion over a batch reactor cycle. Thus, branching and crosslinking reactions are normally more pronounced in CSTR systems.

D.

Scope of This Chapter

Other parts of this book contain detailed discussions of emulsion


polymerization mechanisms, kinetics, and reaction engineering. Thus, the
intent of this chapter is to introduce the reader to some of the basic
conceptsof continuous reactor systems.A rather simplesteady-state reactor
model will be presented for a single CSTR system. This model will be
compared with a batch reactor model for the same reaction mechanism and
the model differences highlighted.
Deviations from the simple model will be reviewed and transient reactor
behavior will be discussed. Possible methods for control of continuous
reactors will be suggested and strategies for process and product development will be considered.

II. Smith-Ewart Case 2 Model Cora CSTR


Models for emulson polymerization reactors must account for particle
formation and particle growth. Ir these two phenomena can be handled in a
satisfactorymanner one can predict the polymerizationrate, the number of
particles formed, and the particle size distribution. The model presented
below was first developed by Gershberg and Longfield (1961). It is based on
the concepts developed for batch reactors by Smith and Ewart (1948) in
their "Case 2" modelo
A.

Particle Formation

Free radical s, initiated in the aqueous phase, can form polymer particles
by a number of mechanisms. Smith and Ewart (1948) considered two
quantitative models for predicting the number of particles formed. One of
these models is shown as Eq. (1) below.
(1)

362

Gary W. Poehlein

where N is partic1e concentration in number per volume of aqueous phase, t


is time, R is the rate of initiation of polymer oligom~rs in the aqueous
phase (assuming no aqueous phase termination), NAis Avogadro's number,
Af is the area that could be covered by the emulsifier that is not adsorbed
on the surface of the polymer partic1es, as is the area occupied by one
molecule of adsorbed emulsifier, and S is the total emulsifier concentration.
Note that the term (asSNA) represents the total surface covering capability
of the emulsifier charged. Thus, Eq. (1) is based on the assumption that the
partic1e formation rate is directiy proportional to the fraction of emulsifier
not adsorbed on the surface of the polymer partic1es. Since no polymer
partic1es exist at the beginning of the reaction, the rate of formation is
dN/dt = RNA. Partic1e formation stops when AF = O,that point where the
total partic1e surface is (asSNA).
In a single CSTR the number partic1es does not change with time and
the partic1e number equation analogous to Eq. (1) is
N

= RfJNA[Arf(asSNA)]

(2)

where () is the mean residence time of the reactor, the working volume
divided by the volumetric effluent flow rateo In order to calculate N one
needs to determine Af from equations for partic1e size distributions. These
equations in turn, come from partic1e growth and reactor residence time
relationships.
B.

Particle Growth

The basic tenet of the Smith-Ewart Case 2 partic1e growth model is that
each partic1e will contain an active free radical one half of the time. Thus,
the average number of free radicals per partic1e is 0.5. The rate of
polymerization Rp is given by
Rp

= kp[M][R.] = kp[M](N/NA)

(3)

where kp is the propagation rate constant and [R -J is the total radical


concentration at the reaction site in the polymer partic1es. The polymerization rate per partic1e Rpp is
(4)
Rpp = kp[M](/NJ
If excess monomer is available in the form of monomer drops the monomer
concentration in the partic1es remains relatively constant. Thus, particle
growth occurs when monomer is converted to polymer, and additional
monomer for swelling the polymer diffuses into the partic1e from the

10.

Emulsion Polymerization in Continuous Reactors

363

monomer droplets. Under these conditions the particIe growth relationship


is given by

dv/d7:= 4nr2 dr/d7: = K1[M]

(5)

where K 1 is a constant dependenton the monomer swelling of the latex


particIe and on the propagation rate constant. Ir = 0.5, Eq. (5) can be
integrated to yield the following relationships for particIe size as a function
of age 7:.

v=

Vo +

0.SK1[M]7:

(6a)

r3 = r~ + 3K[M]7:/Sn

(6b)

where ro and Voare the radius and volume respectively, of the particIe at
nucIeation.

C. Particle Age and Size Distrihutions


The particIe age distribution in a single CSTR is given by
f(7:)

= 0-1 exp(-7:/0)

(7)

The particIe size density function U(r) is obtained by combining the growth
equation with the age distribution.
U(r)

= f(1:)/(dr/dt)

U(r)

= [Snr2/(K

The cumulative distribution

U(r)

(Sa)

I[M] O)] exp[

Sn(r3 - r~)/(3K 1 [M] O)]

(Sb)

function is given by Eq. (9).

= J: U(r) dr = 1.0 - exp[-'Snr3/3Kl[M]0]

(9)

The last equation has been derived by assuming that ro, the nucIeation
particIe radius, is equal to zero. Ir the conversion in the reactor is modest
and the particIe sizes are not too large, Eq. (9) can be used for accurate
predictions of particIe size distributions in styrene latexes. Figure 1 is a
typical comparison of Eq. (9) with experimental data.
D.

Numher 01 Particles and Polymerization Rate

Equation (2) for the number of particIes per unit volume of aqueous
phase has one undetermined parameter Af which is the difference between
the surfactant charged and the surfactant adsorbed on the surface of the
polymer particIes.
(10)

364

Gary W. Poehlein
1.0

;:
=>
al

0.8

,~
~,,/

~ 0.6
(J)

15

eI '

ILI

~ 0.4

~
~

=>
:::!:

e'I
I

a 0.2

,,

e,'
,

,I

" ,/
.,,,,,;I

200

400

600

PARTlCLE

800

1000

DIAMETER

(Al

1200

1400

Fig. 1. Particle size distribution for a styrene latex (Poehlein and De Graff, 1971).
T = 70oC: 0= 15min; (S)= 2.79g/lOOg H20; (/) = 0.35g/IOOg H20. ---, Dn ($-E Case
2) = 690 ; -,
Dn(De Graaf) = 700: O, Dn(Exper.) = 720.

Combining Eqs. (10), (8b), and (2) yields the following relationship for N
(RfJNAfN) = 1 + (C<oRO/asS)[kp[M]O/(1

C(1[M])]2/3

(11)

or
1t1 = 1

+ 1t2

where the 1t1 and 1t2 are the groups indicated, c(o= 3.85(MoVp)2/3,
C(l= (MoVrnx 10-3), Mo is the monomer molecular weight, and Vp and Vrn
are specific vqlumes of polymer and monomer, respectively. Equation (11) is
written in dimensionless formo The second term on the right hand side 1t2
represents the ratio of surfactant adsorbed on the polymer particIes to the
free surfactant Ar. This ratio is normally much greater than 1 and Eq. (11)
reduces to

(12)
Combining Eqs. (3) and (11) permits the calculation of polymerization rateo
(13)
or
1t3

= 1+ 1t2

Figures 2 and 3 show a comparison of Eqs. (11) and (13) with data for
styrene emulsion polymerization over a wide range of experimental conditions. The rate data fit the theory quite well and the particIe number data

10.

Emulsion Polymerization

in Continuous

DEGRAFF

"

GERSHBERG a

GERRENSa KUCHNER

365

Reactors

LONGFIELD

CALCULATED FROM
R~ ASSUMING

1\

1/2

STYRENE

lO'

102
I

Fig.2.
1971).

103
+ lf2

Particle number: Theory-experimental

'"
o

comparisons (Poehlein and De Graff,

DEGRAFF
GERSHBERG a LONGFIELD
GERRENS a KUCHNER

STYRENE

101

Fig.3.
1971).

loS

102
+ 1T2

Po1ymerization rate: Theory-experiment

comparisons (Poehlein and De Graff,

Gary W. Poehlein

366

also tit in a reasonable manner. The particle number data are usually
somewhat lower than model predictions. There are at least two reasons that
the pro po sed theoretical model would predict more particles than actually
measured. First, the procedures used for counting particles would normally
error on the small side. Second, the model equations are based on the
assumption that the particle nucleate size ro is zero. Thus, the model would
overestimate the number of particles.
E.

Comparison with a Batch Reactor

The Smith-Ewart Case 2 model for the number of particles formed in a


batch reaction is given by
N

= kR?4(asSNA)0.6

(14)

where k is a constant. The polymerization rate relationship given by Eq. (3)


is also valid for the batch reactor when N is given by Eq. (14). The
important point here is that the same physicochemical mechanisms yield
model equations for the two reactor types in which even the exponents on
the important recipe parameters are different (see Table 1). In such cases the
use of batch data for design of continuous reactor systems would involve
considerable risk.
A second important difference between the batch reactor model and the
CSTR model is the prediction of a maximum number of particles for the
CSTR as a function of mean residence time. This maximum is readily
determined by setting the derivative (dN/dO),obtained from Eq. (11), equal
to zero and solving for N. The result of this computation is
NCSTR'MAXIMUM

(15)

O.577NBATCH

Thus, fewer particles will be produced from the same recipe in a eSTRo The
particle concentration in a eSTR can be increased by increasing the
emulsitier concentration or by using seed particles in the feed streams.
TABLEI
Model Equation Exponents for
Smith-Ewart Case 2 Models
Equation exponents

Parameter
R;
S

Batch reactor

0.4
0.6
NA

CSTR
O
1.0
-0.67

10. Emulsion Polymerization in Continuous Reactors


F.

367

Effect ollnhibitors in Feed Streams

Inhibitors are normally present in stored monomers and sometimes in


other feed streams. In batch reactors inhibitors delay the start-of polymerization, but in a CSTR they cause a reduction in the rate of initiation. The
effective initiation rate for the first CSTR in a series system is given by
(16)
where f is the initiator efficiency factor, kd is the decomposition rate
constant, [/]0 and [H]o are the concentrations of initiator and inhibitor,
respectively, in the combined feed streams, and fH is the number of free
radicals consumed per inhibitor molecule.
Equation (16) is valid only if the quantity within the brackets is positive.
If this term is zero or negative no polymerization will take place in the first
reactor. The rate of initiation in the second CSTR in a series system is
(17)
where e2 is the mean residence time in reactor 2. R.2 can be larger than R.l
and this fact can sometimes cause problems with heat removal and temperature control in the second reactor. This effect can be strong enough in
some situations to require that additional inhibitor be added to the second
reactor.
G.

Use 01 a Particle Feed Stream

Emulsion polymerization reactions are sometimes carried out with small


seed particles formed in another reaction system. A number of advantages
can be derived from using seed particles. In a batch reactor seed latex can
be helpful in controlling particle concentration, polymerization rate, particle
morphology, and particle size characteristics. In a CSTR the use of a feed
stream containing seed particles can also help to prevent conversion andjor
surface tension oscillations, which are caused by particle formation phenomena. This factor will be discussed in more detaillater in this chapter.
III.

Deviations from Smith-Ewart Case 2

The key assumption in the Smith-Ewart Case 2 theory can be stated


mathematically by the simple expression = 0.5. The physicalbasis for this
assumption involves several phenomena. First is the fact that free radicals
react with one another to terminate polymerization very rapidly. Second,
the latex particles, which are the reaction sites, are very small. Third, the

368

Gary W. Poehlein

mobility of free radicals within the polymerizing partic1es is considered to


be high. Fourth, the free radical oligomers within the partic1es are assumed
to be incapable of escaping from the partic1es. Factors that alter any of
these phenomena will cause deviations from Smith-Ewart Case 2 kinetics.
A.

Large Particles or Slow Termination

Stockmayer (1957) was the first to present a genera) solution to the


Smith-Ewart recursion equation for . This solution is reproduced below
for the case in which free radical escape from partic1es is not possible.

= (a/4) (Io[a]/l

(18)

l[a])

where a = (Pa/N)(v/kt), lo and 11 are Bessel Functions, and kt is the


termination rate constant in the latex partic1es. Figure 4 is a plot of
Eq. (18). Higher values of a can be caused by large partic1es (large v), slow
termination reactions (low kt), or a high radical flx per partic1e (large
pJN). When becomesgreater.than 0.5 the larger partic1esgrow faster,in a
volumetric sense, than the 'smaller partic1es. This causes the partic1e size.
1.5

1.0

0.5

IC:
O>

!2

-0.5

- 1.0

-4

-3

-2

-1
log 01

Fig. 4. Stockmayer's equation plotted as logn as a function of logO( (Ugelstad and


Hansen, 1976). m = O; O(= Pav/Nk,.

10.

369

Emulsion Polymerization in Continuous Reactors


1.0

08

z 0.6
o
::
:o
ID
:
....

'"
o 0.4
"'
>
::
<t
..J
:o
;;

0.2.

/
0/
/
/
/
/

500

1000
PARTICLE

1500

2000

2500

3000

3500

4000

OIAMETER (A)

Fig. 5. Particle size distribution for a styrene latex made at a mean residence time of
59min (Poehlein and De Graff, 1971). T=50C:
O=59min;
(S)=0.6g/100g
H20;
(1) = 1.5 g/100 g H20. ---, Dn (S-E Case 2) = 858 ; -,
Dn (De Graaf) = 1018 ; 0, Dn
(Exper.) = 1090 .
.

distribution to become broader. Figure 5 shows this effect for a styrene


emulsion polymerization. The dashed curve is based on the Smith-Ewart
Case 2 model and th~ solid curve includes the Stockmayer correction.
B.

Radical Desorption

Ugelstad and co-workers (1967, 1976) presented a more general solution


to the problem of computing . Their work was presented in the form of a
series of graphs based on the parameters listed below
(X'= (RJNkt/v)

(19)

m = k~v/kt

(20)

y = 2Nktktw/k;v

(21)

where k~ is the rate constant for radical desorption from particles, ktw is a
termination rate constant for the water phase, and ka is arate constant for
radical absorption. Ugelstad's theory allowed for the reabsorption of
previously desorped radials; a factor not handled properly by Stockmayer.

370

Gary W. Poehlein
2

IC

'"
.2

-1

-2

-6

-8

-o

-2

-4

logo'-

Fig. 6.

Average number

of radicals

per partic1e as a function of el and m for Y

= 10-2

(Ugelstad el al., 1967).

Ugelstad's results were presented as a series of graphs such as shown by


Figure 6.
The parameters defined by Eqs. (19-21) need to be defined differently
when dealing with particle populations that are not monodisperse. The
basic result, however for a system where #- 0.5, is that particle growth is a
positive function of particle size, i.e.,
dv/dt = k'vr

r>O

(22)

This causes particle size distributions in the latex from a CSTR to be


broader. Rate of polymerization is directly proportional to (N). Since
can decrease as N increases the rate of polymerization may not depend
strongly on N.
C.

Other Particle Formation Mechanisms

Smith and Ewart (1948) assumed that particles are formed when free
radical s diffuse into monomer-swollen micelles. Roe (1968) demonstrated
later that the S-E Model, in a mathematical sense, did not depend on the
concept of a micelle. Roe referred simply to the stabilizing capability of the

10.

371

Emulsion Polymerization in Continuous Reactors

HYDROPHILlC

-S04'

+M

---

-S 04M

SURFACE
ACTIVE

--

I
I +M

,-

-S04Mn

HYDROPHOBIC
I
HOMOGENOUS I+M
NUCLEATION

ADSORPTION
ONEXISTING
R\RTICLES OR
MONOMERDROPS
CAPTUREBY
MICELLES

so;
FLOCCULA

FLOCCULATION
OF PRIMARY
R\RTICLES

TION

ONTO
MATURE
R\RTICLES

+MaS
POLYMERIZATION
GROWTH

CONTINUED GROWTH BY

POLYMERIZATION
I FLOCCULATION,
OR

Fig.7.

BOTH

Paths for free radicals initiated

in the aqueous

phase.

system or the free emulsifier concentration. We know that particles can be


formed by a number of mechanisms. Figurs: 7 shows some potential ways in
which a sulfate ion-radical might become associated with a latex particle.
The rate at which 'various mechanisms proceed depends on a number of
factors, including those listed below:
1. Monomer solubility in water.
2. Propagation rate of the polymerization reaction in the aqueous and
organic phases.
3. Size and concentration ofmonomer drops.
4. Size and concentration of micelles if any are present.
5. Size and 'concentration of polymer particles.
6. Solubility of the growing polymer oligomers in the water phase.
7. Concentration of free emulsifiers in the water phase.
8. Concentration of electrolyte in the water phase.
The monomer solubility in the water phase, discounting that located in
any swollen micelles, will influence the rate at which the oligomer chains
grow. Monomer solubility will, of course, also be related to oligomer
solubility, or more precisely to how long the oligomer chains might grow

372

Gary W. Poehlein
TABLE

11

Solubility of Monomer in Water and Polymer


Concentration in (M):

Monomer

Water

Styren"e
Butadiene
Vinylidene chloride
Vinyl chloride
Methyl methacrylate
Vinyl acetate
Ethylene
Acrylonitrile
Acrolein

0.005
0.015
0.066
0.11
0.15
0.3
0.3-0.6
1.75
3.1

Polymer
partic1es

5.4
6.5 .
1.1
6.0
7.0
7.6
5.0

before they precipitate. Table 11gives monomer solubilities in water and in


the polymer particles for several common monomers. It should be noted
that in copolymer systems the oligomers formed in the water phase would
be expected to have a higher concentration of the more soluble monomer if
the reaction rates are of the same order.
The size and number of monomer droplets are usually ignored in
emulsion polymerization kinetic stud~es. The basis for this comes from
several observations. First, in CIassical recipes the monomer is located in a
relatively few, large monomer droplets that represent a small total surface
area when compared to the micelles or free emulsifier. Second, large
particles are not usually seen in electronmicrographs of latex particles. Ir
one or two are found they can be conveniently forgotten.
Ugelstad et al. (1973) cIearly demonstrated that if the monomer droplets
can be made small enough, they can efTectivelycompete for free radicals and
form particles. Considerable work has been published (see Ugelstad and
Hansen, 1976, for other references) on the preparation of finely dispersed
monomer droplets for the formation of latex particles by direct
polymerization.
Recently,Durbin et al. (1979)have shown that a few large particles are
formed by monomer droplet polymerization in cIassical styrene recipes.
Their work showed that the number of oversized particles increased with
increased intensity of preemulsification. Thus, in some circumstances the
installation of preemulsification systems may be counterproductive.
The size and number of micelles and particles will influence the path
taken by free radicals in the water phase. As sizes andjor numbers increase
the probability for absorption increases.

10. Emulsion Polymerization in Continuous Reactors

373

PRIMARY
PARTICLE
Fig. 8.

Schematic representation of the growth and precipitation of an oligomeric

radical (Fitch and Tsai, 1971).

The vertical path shown in Fig. 7 iIIustrates a rnechanisrn of particle


formation which Fitch and Tsai (1971) have called "hornogeneous nucleation." Figure 8 shows this process of diffusion and addition of rnonorner
units schernatically.
As the oligorneric radical grows it becornes strongly hydrophobic. In this
state, it rnay precipitate to forrn a prirnary particle or deposit on an existing
particle, micelle, or rnonorner drop. Estirnates of the degree of polymerization necessary to cause precipitation have been given as 54 for vinyl
acetate (Priest, 1952),65 for methyl rnethacrylate (Fitch and Tsai, 1971), and
30 for styrene (Peppard, 1974).
Another rnechanisrn for particle forrnation involves the scheme shown in
Fig.7, but it oCGurslater in sorne batch polyrnerizations. This secondary
nucleation is caused by free ernulsifier which is libenited from the particle
surface. Figure 9 shows sorne data of Gerrens for a batch ernulsion
polyrnerization of MMA. These data were successfully modeled by Ray and
Min (1976). The population of srnaller particles was forrned late in the
reaction because ernulsifier was desorped beca use of the crowding of radical
end groups on the surface and because of particle shrinkage caused by .
rnonorner conversion to polyrner.

374

Gary W. Poehlein
.30

'"

"

.20

>1j
Z
LrJ
o
.J
~
~
a:
o
z

.10.

40

80

DIAMETER

120
(nm)

Fig. 9. Experimental and predictedparticle sizedistribution for a MMA latex (Ray and
Min, 1976).0,Gerren'sdata: -,
Model prediction (97.8%conversion).

D.

Experimental Results

The model based on S-E Case 2 kinetics has been quite successful in
handling steady-state data for styrene emulsion polymerizations in a CSTR.
One or more of the mechanisms described above, however, generally cause
other monomer systems to deviate from this simple model. The nature of
these deviations varies among the different monomers. Ir published literature data are fitted to equations of the type listed below one can obtain
values for the exponents a, b, and c.
Rp = kRRf[S]bOc

(23)

= kNRf[S]bOC

(24)

Table III gives the results of this fitting for styrene, methyl acrylate, methyl
methacrylate, vinyl acetate, vinyl chloride, and ethylene. Only the styrene
data agree with the simple theory.
The exponents on R are usually greater than 0.0 (the theory prediction).
This can be explained by radical desorption from the particles, by slow
t.erminatiop in the particles, or by a combination of these factors. Data on
the other variables ([S] and O)are incomplete but scattered results show
significant deviations from the S-E Case 2 Model. More complete models,
such as those presented in this book by Hamielec, have been successful with
nonstyrene monomers.

10.

375

Emulsion Polymerization in Continuous Reactors


TABLE

Steady-State

111

Behavior of CSTRs
Approximate exponents

a
Theory (Rp and N)
Styrene (Rp and N)
MA(N)
MA(Rp)
MMA (Rp)
VA (Rp)
VC(Rp)
Ethylene (Rp)

IV.
A.

0.0
0.0
0.0
0.65
0.8
0.8
0.5 or less
0.5

1.0
1.0
0.85
0.0
0.9
SmaIl
?
0-0.3

-0.67
-0.67
-0.67
+0.43
?
0.0
SmaIl
Small

Transient Behavior of CSTR Systems


Experimental Observations

An understanding of the transient behavior. of continuous reactors is


important for start-up and reactor control considerations. Continuous
oscillations have been observed by a number of workers. Figures 10 and 11
show data for styrene and methyl methacrylate. Gerrens and Ley (1974)
reported continuous, undamped oscillations in surface tension during a
styrene emulsion polymerization run which lasted for more than 50 mean
residence times. Nearly five complete cycles were observed during this run.
Berens (1974) conducted experiments with vinyl chloride in which the
measured particIe size changed with time. No steady state was achieved with
the data shown in Fig. 12.

B.

Physicochemical Mechanisms

The reason for oscillations in conversion and surface tension become


clear when one considers particle formation and growth phenomena. If a
single CSTR is started empty or by adding initiator to a full vessel of
inactive emulsion, a conversion overshoot occurs. The first free radicals
generated are almost entirely utilized to form new particles. Since these
particles do not grow rapidly to the steady-state size distribution, radical

376

Gary W. Poehlein
70
60

STYRENE
40
z
O
C
a: 30
UJ

>
Z
O
u

-~
- -

-~

80

120
RUN

Fig. 10.

Conversion transients

160
TIME

at

start-up

200

7O'C

_STAGE2
- ~
7O'C

.: STAGEI
.100'C

20

10
O

StAGE 3

240

280

{min}
for

styrene

emulsion

polymerization

(Gershfield and Longfield, 1961).

50

40

30

z
020
C
a:
UJ
>

10
METHYL

METHACRYLATE

O
O

Fig. 11.

TIME

10

12

14

(t le)

MMA conversion transent (Greene, Gonzalez, and Poehlein, 1976). T

40C: IJ= 20 min:


(NaHC03)

DIMENSIONLESS

(S) = 0.01molj\iter;

= O.OL (NaCI)

= 0.02.

(1) = 0.01molfliter; Monomer:H20

= 0.43;

377

10. Emulsion Polymerization in Continuous Reactors


30
i
10

-,

',

,1 \

"'

"

"

-,

'1

\
\
\

\,

\,

"
i,
I

::>

i
I

\'

11
11

Q: 1.0
w
ID

I
I
I

"-

1Z
w
<.)
Q:
W
a.

/'.,

'

'Y..

'

I "

'-

'-

" '-. '-

'
I
I

\",

\
\

0.3

/ /

/-..

0.1
O

500

1000

1500

2000

2500

3000

~RTICLE DIAMETER
(A)
Fig. 12. PVC particle size distribution transients (Berens,1974).-,
'-, t = 120mins;---, t = 240mino

t = 60 min; -

efficiency in particle formation rema ins high and the particle concentration
pass es its steady-state value. As these parti~les grow they genera te too much
surface to be saturated by the emulsifier and the particle formation rate
become much smaller, perhaps zero.
The effiuent from the reactor carries some of this unsaturated surface and
new emulsifier is being continuously added with the feed. Eventually, the
surface will be saturated and free emulsifier will be available to stabilize new
particles. If the particle formation and growth rates are appropriate,
continuous limit-cycle oscillations can be observed.
The particle formation rate depends on the rate of production of free
radicals and on the relative rates of the various mechanisms in which these
free radicals participate. A high rate of particle formation in the presence of
small amounts offree emulsifier will contribute to unstable cyclic behavior.
The rate of growth of particles depends on the concentration of the
reagents in the particles, radicals and monomer, and on the propagation
rate constant. The gel effect, which causes the termination rate constant to
be lower at higher conversions, can cause higher free radical concentrations
in the particles and thus higher particle growth rates. This effect also
contributes to conversion oscillations.

378

Gary W. Poehlein

C. Control Methods
A number of methods have been considered for controlling transients in
eSTRs. One simple method for trying to handle transients involves the use of
a smalI eSTR as a prereactor. The natural oscilIations and transients from
this reactor are damped by backmixing in the larger reactors that folIow.
Greene el al. (1976) demonstrated that strong initial overshoots and
conversion oscilIatiQns could, in some cases, be avoided by careful start-up.
The reactor was started either with a seed latex from a previous run or fuIl
of distilIed water. Both procedures were successful in achieving steady
operation in some cases. In other cases, however, significant transients were
obse!ved after a smooth start-up.
Hamielec el al. (1980) and Ray and Schork (1980) have worked on
potential feedback control schemes. Hamielec and co-workers have developed an on-line light scattering device that dilutes the latex and measures
scattering at several waveIengths. This data is then used to predict changes
in particIe concentrations and size characteristics.
.
Ray and Schork (1980) have developed an on-line device for measuring
surface tension. A bubble rise technique is used. Ir this method or the light
scattering technique of Hamielec were used in a control system one could
presumably alter the feed rate of a key ingredient, probably emulsifier, to
prevent undesired transients. More work is needed to demonstrate the
potential of these methods.
Berens (1974) and Gonzalez (1974) have shown that the use of a seed
latex in the feed stream can prevent oscilIatory behavior. Berens used a
latex seed generated in a separate reactor. He demonstrated that the
particIe size distribution reached a constant state in a pve polymerization.
Steady states were not observed in unseeded reactions.
Gonzalez placed a continuous tubular reactor in front (upstream) of the
eSTRo In this case the particIe seed was formed in the tube from a recipe
that did not contain seed. Gonzalez found the tube-CSTR system to be
quite stable so long as the conversion in the tu be was adequate to prevent
significant particIe formation in the eSTRo
The use of a particIe seed has the added advantage of aIlowing more
careful control of particIe concentration and reaction rate in the eSTRo

v.

Strategies for Process and Product Development

A.

Reactor Design Considerations

A continuous emulsion polymerization system needs to operate for


significant periods between shut-downs. Thus, the reactor system should be
designed and operated to prevent, as much as possible, the formation of
waIl polymer. The causes of and cures for waIl polymer are not completeIy

1
~
~
1

10.

Emulsion Polymerization in Continuous Reactors

379

understood. The factors listed below need to be considered to help minimize the wall polymer problem.

1. Liquid level: The reactor should be operated full. A vapor space


almost always causes problems. Latex can dry on the wall and present a site
for growth, sometimes accelerated, of more deposits.
2. Nozzle locations: The location of the feed and effiuent nozzles should
insure rapid mixing of the feed streams and prevent short circuiting between
feed and effiuent. Electrolyte streams should be as dilute as feasible to
prevent electrolyte flocculation of the latex.
3. Reactor surface: The internal surface of the reactor should be
smooth. Rough places pro vide sites for polymer nucleation andjor deposition. Glass and polished stainless steel surfaces are common. Stainless .
steel is preferable if it can be used because the glass coatings reduce heat
transfer capability.
4. Agitation: The agitation needs to be designed to provide adequate
mixing of the feed streams and adequate heat transfer. Good mixing usually
involves a balance between flow and turbulence. Since turbulence or high
shear can sometimes cause coagulation this balance usually involves more
flow and less turbulence.
Good mixing normally requires internal baffies in the reactor. These
baffies, however, represent potential sites for wall polymer and thus should
be minimized or, if possible, eliminated. Good mixing can be achieved in
unbaffied reactors with axial flow impellers mounted at an angle to the
vertical. Such reactors are described in G~rman Patent 2,520,891 assigned
to Bayer, A. G.
5. Operationjsurface tension: The reactor should be operated in such a
manner that large transients in surface tension are avoided. Conversion and
surface tension oscillations will tend to contribute to wall polymer formation. Start-up policies, system design, and control procedures should be
selected to insure steady, free emulsifier levels in the particle formation
reactor. In some cases it may also be desirable to add more emulsifier to
downstream reactors.
6. Operationjcleaning: The reactor should be cleaned very well. The
growth of polymr deposits are usually non-linear functions of time in
which the size of deposits increase rapidly. During cleaning one must also
avoid damaging the surface. Cracks, roughness, and surface imperfections
represent potential nucleation sites for wall polymer buiidup.
7. Wall temperatures:Some practitioners of the art and science of
emulsion polymerization believe that nonuniform wall temperatures can
playa role in wall polymer formation. The reaction in any deposit could be
faster if the temperature is higher. Thus, some reactors are designed with
cooling jackets that cover as much of the reactor as possible, including the
reactor topo

380

Gary W. Poehlein

Other reactor design considerations may be necessary in special cases.


Monomer mass transfer, not normally a problem, can be important if the
monomer-aqueous phase interface is small. This is more likely in systems
involving gaseous monomers in which the large surface area of the monomer emulsion is not presento In such cases special attention must be paid
to gas dispersion and transporto Other factors that can have a significant
effect on reactor design include latex viscosity, heat transfer rates, reaction
pressure, and control mechanisms.

;;

~
~
~
~
~
~

B.

Product Development Considerations

A number of factors need to be considered when developing a product


that will be produced in a continuous reactor system. The initial development work on most new products is carried out in batch reactors. Bottle
polymerizers are often used for this purpose because a large number of
experiments can be completed quickly. These early experiments provide a
product that can be tested against fundamental standards (molecular
weight, particle size, rheology, etc.) and in proposed applications.
Preliminary recipes evolve from such tests.
Further development of any product usually involves more specialized
reactors and product testing. The results of these tests include more refined
recipes, reactor operation procedures, and engineering data for use in design
of the final commercial process.
Ir a continuous process is to be used for commercial production, a
similar small-scale reactor system should be utilized in this second stage of
product development. There are a number of reasons for this recom.mend~tion. The earlier discussion of the difference between batch reactors
and CSTRs lists some of these reasons. If, for example, engineering data are
to be obtained for design of a commercial unit, the variable relationships
might be quite different for the different reactors. The Smith-Ewart CSTR
model predicts a linear relationship between Rp or N and the surfactant
concentration [S]. The same mechanistic model for a batch reactor predicts
a 0.6 power relationship between Rp or N and [S].
Products from continuous, batch, and semicontinuous reactors can differ
in a number of ways. Some of the more obvious characteristics that might
be influenced are listed below:
1. Particle characteristics: particle number, average size, and size
distribution.
2. Distribution of the composition of molecules in copolymer products.
3. Particle morphology or the distribution of copolymer composition
as a function of radius. '

10. Emulsion Polymerizationin Continuous Reactors

381

4. Molecular weight averages and distributions.


5. Frequency and distribution of branching andjor grafting sites on the
molecules.
Batch, semicontinuous, and continuous reactors that are designed intelligently can be utilized to manipulate almost any fundamental andjor
application property. Suitable products for most applications can usually
be manufactured in several different reactor systems. One may be required,
however, to adjust the latex recipe and reaction conditions to produce a
satisfactory product in one reactor type if that produce has been developed
in another reactor system. Thus, if a reactor system has been chosen for the
commercial unit one should plan on some, and in most cases considerable,
product development work in a similar, small-scale system.

VI. Summary
Some of the fundamentals of reactor systems have been presented and
discussed in this chapter. This material should assist the reactor engineer in
planning appropriate pilot studies for a new or revised process. Equally
important, the concepts reviewed should help the product development
personnel in devising reactor systems and operating procedures that will
produce satisfactory products.
More complete discussions of polymerization kinetics and reactor
modeling are contained in other chapters of this book. These factors will
also be of great value to the reactor engineer and the development chemist.
As the kinetic mechanisms and reaction models become better developed
we can look forward to more efficient research and development programs
for new products and processes.

Acknowledgment
This material is based, in part, upon work supported by the National Science Foundation
under Grant No. CPE-8011455.

References
Aho, C. E. (1958). U.S. Patent 2,831,842 (to E. 1. duPont de Nemours and Co., Wilmington,
Delaware), April 22.
Berens, A. (1974). J. Appl. Polym. Sci. 18,2379.
Calcott, W. S., and Starkweather, H. W. (1954). U. S. Patent 2,384,277 (to E. l. duPontde
Nemours and Co., Wilmington, Delaware), September 4.

382

Gary W. Poehlein

Durbin, D. P., EI-Aasser, M. S., Poehlein, G. W., and Vanderhoff, J. W. (1979). J. App/. Po/ym.
Sci. 24, 703.
Feldon, M., McCann, R. F., and Lundrie, R. W. (1953). Indian Rubber Wor/d 128(1), 51.
Fitch, R. M., and Tsai, C. H. (1971). In "Polymer Colloids" (R. M. Fitch, ed.), Chapters 5 and 6.
Plenum Press, New York.
Gerrens, H., and Ley, G. (1974). Private cornmunication.
Gershberg, D. B., and Longfield, J. E. (1961). Symp. Po/ym. Kinet. Cata/o Syst., 45th AIChE
Meeting, New York Preprint 10.
Ghosh, M., and Forsyth, T. H. (1976). In "Emulsion Polymerization" (1. Piirma and J. L.
Gardon, eds.), pp. 367-378. American Chemical Society Symp. Series 24, Washington, D.C.
Gonzalez, P. R. A. (1974). M.S. Thesis, Chem. Eng. Dept., Lehigh Univ., Bethlehem,

Pennsylvania.

Greene, R. K., Gonzales, R. A., and Poehlein, G. W. (1976). In "Emulsion Polymerization"


(1. Piirma and J. L. Gardon, eds.), pp. 341-358. American Chemical Society Symp. Series 24,
Washington, D.C.
Hamielec, A. E., Kiparissides, G., and MacGregor, J. F. (1980). In "Polymer Colloids 11"(R. M.
Fitch, ed.), pp. 555-582. Plenum Press, New York.
HelIer, W., and De Lauder, W. B. (197Ia). J. Col/oid Interface Sci. 35(2), 300--307.
Heller, W., and De Lauder. W. B. (197Ib). J. Col/oid Interface Sci. 35(2), 308-313.
Heller, W., and Peters, J. (1970a). J. Col/oid Interface Sci. 32(4), 592-605.
Heller, W., and Peters, J. (1970b). J. C9l/oid Interface Sci. 33(4), 578-585.
Heller, W., and Peters, J. (1971). J. Col/oid Interface Sci. 35(2), 300--307.
Lanthier, R. (1970). U.S. Patent 3,551,396 (to Gulf Oil Canada, Ltd.), December 29.
Lundrie, R. W., and McCann, R. F. (1949). Ind. Eng. Chem. 41, 1568.
Owen, J. J., Steele, C. T., Parker, P. T., and Carrier, E. W. (1947). Ind. Eng. Chem. 39, 110.
Peppard, B. D. (1974). PhD Dissertation, Chem. Eng. Dept., Iowa State Univ., Ames, Iowa.
Poehlein, G. W., and DeGraff, A. W. (1971). J. Po/ym. Sci. Part A-2 9, 1955-1976.
Poehlein, G. W., and Dougherty, D. J. (1977). Rubber Chem. Techno/. 50(3), 601.
Priest, W. J. (1952). J. Phys. Chem. 56, 1077.
Ray, W. H., and Min, K. W. (1976). In "Emulsion Polymerization" (1. PiirmaandJ. L. Gardon,
eds.), pp. 359-366. American Chemical Society Symp. Ser. 24, Washington, D.C.
Ray, W. H., and Schork, F. J. (1980). News artic1e, Chem. Eng. News. 58(37), 44-46.
Roe, C. P. (1968). Ind. Eng. Chem. 60, 20.
Rollins, A. L., Patterson, W. l., Archambault, J., and Bataille, B. (1979). In "Polymerization
Reactors and Processes" (J. N. Henderson and T. C. Bouton, eds.), pp. 113-136. American
Chemical Society Symp. Ser. 104. Washington. D.C.
Smith, W. V., and Ewart, R. H. (1948). J. Chem. Phys. 16,592.
Stockmayer, W. H. (1957). J. Po/ym. Sci. 24, 314.
Ugelstad, J., EI-Aasser, M. S., and VanderholT, J. W. (1973). J. Po/ym. Sci. Po/ym. Leu. Ed.
11, 505.
Ugelstad,
Ugelstad,

J., and Hansen, F. K. (1976). Rubber Chem. Techno/. 49(3), 536-609.


J., Mork, P. c., and Aasen, J. O. (1967). J. Po/ym. Sci. Part A-I 5. 2281.

Wolk, l. L. (1959). U.S. Patent

2,458,456 (to Phillips Petroleum

Co.), January

4.

11
Effect of Additives
on theFormation
of M onomer Emulsions
and Polymer Dispersions
J. Ugelstad, P. C. M0rk, A. Berge. T. Ellingsen, and
A.A. Khan

1. Introduction .
11. Thermodynamic Treatment of Swelling
and Phase Distributions
111. Rate of Interphase Transport .
IV. Preparation of Polymer Dispersions .
A. Polymerization with Initiation in Monomer
Droplets Formed by the Diffusion Process .
B. Emulsificationof PolymerSolutions .
V. Effectof Addition of Water-Insoluble Compounds
to the Monomer Phase .
VI. Emulsificationwith Mixed EmulsifierSystems .
List of Symbols .
References

383
384
392
396
396
399
401
408
411
412

l. Introduction*
Quite recently, new methods for the preparation of emulsions by
diffusion processes have been described in a number of papers. The methods
in volved the diffusion of slightly soluble compounds' (ZI) through the
continuous phase to become absorbed into droplets of relatively low
molecular weight; highly insoluble compounds (Z2) or into particles consisting of polymer and highly insoluble compounds (Z2)' The swelling capacity
of such systems was in several cases shown to exceed that of pure polymer
particles by a factor of more than 1000.
* For an explanation

of symbols see p. 411.

383
EMULSION POLYMERIZATION
Copyright @ 1982 by Academic Preso. loc.
All righto of reproductioo io aoy forro reserved.
ISBN 0-12-556420-1

384

J.Ugelstad et al.

A section of the present chapter is devoted to a rather detailed description of the basic thermodynamic and kinetic principIes involved in the
various swelling procedures. This seems justified in view of the present, and
potential, applications of these principIes to the preparation of emulsions
and polymer dispersions. The swelling procedures developed so far have led
to the first successful methods for preparation of large, monodisperse
partic1es.
Addition of small amounts of Z2 to solutions of polymers has been
shown to facilitate formation of aqueous emulsions of 'such solutions. A
very recent application involves the addition of Z2 to the monomer phase in
an ordinary emulsion polymerization. This resuIts in a decrease in the
concentration of monomer in the polymer partic1es and thereby in a change
in the kinetics of polymerization.
Another, more specific method for the preparation of emulsions of ZI
involves the addition of ZI to a preformed mixture of an ionic emulsifier, a
long-chain fatty alcohol, and water. In this way, the. rapid formation of a
stable emulsion may be obtained at ordinary stirring with relatively modest
amounts of emulsifier. The mechanism of this process is still not satisfactorily explained. AIso, subsequent polymerization (in the case where ZI is a
monomer) may lead to polymerization with initiation in monomer droplets.
The present chapter reviews the most recent work on the abovementioned topics and inc1udes some hitherto unpublished theoretical and
experimental material.
11. Thermodynamic Treatment of Swelling and Phase Distributions
The formation of emulsions and the swelling of polymer-oligomer
partic1es involve the mixing of amorphous substances of different molecular
weights. These processes will be discussed in terms of the original FloryHuggins theory for the free energy of mixing. For a mixture consisting of n
components, the free energy of mixing may be expressed as
n

I1G/RT

n- 1

" N In +" "

L...
i=1

A..
'rt

i=lj=i+1

~.
J.NA..
I ro/) Xt,)

(1)

where X:,j is the interaction parameter per segment of component i with


component j, Ji is the number of segments in component i, Ni is the number
of moles and <Pi
the volume fraction of component i. If, for example, a fourcomponent system is considered and the free energy of mixing is given per
unit volume, one obtains from Eq. (1)
4

I1G/RTV =

i=1

(<Pi In <Pi/VM)

+ (1/Y.)(<pI

<P2X~,2

+ <P2<P3X;.3 + <P2<P4X;,4 + <P3<P4X3.4)

<PI<P3X~,3

<PI <P4X~,4

(2)

11. The Formation of Monomer Emulsions and Polymer Dispersions

385

where V.Mis the partial molar volume of component i and V. is the volume
of one segment. According to Eq. (2) the terms representing the combinatorial entropy are inversely proportional to V.M' Therefore, with large
values of V.M,as in the case of the mixing of two polymer compounds, the
combinatorial entropy makes a very small contribution to the free energy of
mixing.
Expressing the interaction energy in terms of Hildebrand solubility
parameters (Hildebrand, 1964),one obtains
Xl.)~. = Vs (>.I

2
>.
J)

/RT

(3)

Accordingly, polymers will only be miscible when they have almost equal
values of '>,or if specific interactions between the two types of polymer
chains are present, in which case the values of xi.j may be negative. In
practice it turns out that the value of xi.j is not only a function of
temperature but also varies with the composition. More recent treatments
applying the lattice theory of Flory-Huggins have taken this into account.
AIso, the original Hildebrand approach has been refined to take into
account the contribution of polar groups and hydrogen bonds to the
solubility parameters. These modifications of the Flory-Huggins theory and
of the solubility parameter concept have made these methods an even more
useful tool in the description of solutions, especially of mixtures containing
polymer compounds. A comprehensive treatment of these extensions of
Flory-Huggins' and Hildebrand's theories, as well as the new "equation of
state" approach of Flory (1965), has recently been published (Shinoda,
1978; Olabisi et al., 1979).
.
The systems considered in the present context may consist of several
phases, some or all of which may be present as droplets. The differential
molar free energy f9r an n-component system is given by
n

dG

= V dP - S dT + L GndNn +

y dA

(4)

where Gn js the partial molar free energy for component n, y is the


interfacial tension, and A is the surface area. Assuming monodisperse
droplets and subst"ituting dA = 2 dV/r, it follows from Eq. (4) that the
partial molar free energy of mixing for any component i when pure i in bulk
(plane surface) is used as the reference state is
(5)
where LlGr is the partial molar free energy of mixing of component i for the
case that one operates with aplane surface (r = ct:.and r is the droplet
radius.

386

J. Ugelstad el al.

The condition for equilibrium in the system is that the activity of any
component i is the same in all phases. Choosing the pure component in
bulk as the reference state, the equilibrium conditions may be written
l:1G ia

= l:1G

ib

= l:1G ic = ...

(6)

where a, b, c oo.refer to the phases that make up the system. In practice, the
equilibrium state of one or several of the components may be so slowly
established that a semiequilibrium state is reached. In this state, the
compounds that are more easily transferred between the' phases will have
attained their semiequilibrium distribution. Throughout the following discussion, the low molecular weight compounds that are equilibrated between
the phases will be denoted by Zl, the relatively low molecular weight
compounds which are effectively hindered from being transferred between
the phases are denoted by Z2, and the polymer by Z3'
Emulsions of Zl that are slightly soluble in the continuous phase may be
stabilized toward degradation by diffusion by the presence in the droplets of
Zl of a small amount of a Z2 that is insoluble in the continuous phase. This
idea was first suggested by Higuchi and Misra (1962). Later, a more general
thermodynamic discussion was presented by Ugelstad (1978) and Ugelstad
et a/. (1977, 1978a, 1980a,b; Ugelstad and Hansen, 1976). At the instant the
droplets in such a system are formed, their composition, with respect to Zl
and Z2, is identical. Consider two droplet sizes raoand rg where rao> rg. The
initial driving force for the diffusion of Zt from b to a droplets
expressed as (l:1GlbO - l:1G1ao)is then given by
l:1G tbO

l:1G taO

= 2yV1M(l/rg-

l/raO)

(7)

As it is assumed that Z2 is almost completely insoluble in the continuous


phase, the transport of Zl from the smaller to the larger droplets leads to an
increase in the volume fraction CP2bin the b droplets and a corresponding
decrease in CP2a'A concentration potential working in the direction opposite
the interfacial free energy difference in Eq. (7) will thus build up. At
equilibrium, the activity of component i is the same in the a and b droplets,
and furthermore equal to the activity in the continuous phase. The semiequilibrium distribution of component i is therefore determined by
l:1Gra

+ 2YVMlra
= l:1Grb + 2YVMlrb= RTlnaiw

(8)

where the values of the interfacial tension and radii are those at equilibrium (see note on p. 413). The index w is used to denote the continuous
phase, most commonly water. It should however be noted that it mayas well
be a mixture of water and nonaqueous compounds or pure nonaqueous
solvents. The partial molar free energy differences may be obtained from

11. The Formation of Monomer Emulsions and Polymer Dispersions

Flory-Huggins expression. Substituting


semiequilibrium state discussed above
In(4)la/4>lb)+ (1 - JjJ2)(4)2a

387

in Eq. (8) one obtains for the

4>2b)

+ 2V1My(1/ra - 1/rb)/RT = O

+ (4)1a -

4>1b)x12

(9)

where X12 = J1X~2is the interaction parameter per mole of Zl'


Calculations based on Eq. (9) show that even minor quantities (-1%) of
Z2 may prevent degradation by diffusion to the extent that the size of the
resulting emulsion droplets deviate only slightly from their "instantaneous"
values. As Eq. (9) describes a semiequilibrium situation, the emulsion is
subject to further degradation by diffusion. The rate of this "secondary"
degradation process is however determined by the rate of transport of Z2
from b to a droplets. For many practical purposes, this rate may be almost
negligible. In systems that are stabilized toward coalescence by means of
conventional emulsifiers, the capability of small amounts of Z2 to prevent
degradation by diffusion has been demonstrated by several authors
(Higuchi and Misra, 1962; Hallworth and Carless, 1974; Davies and Smith,
1974).
It has also been demonstrated that the stabilizing effect of different
compounds is independent of the chemical structure and determined only
by those compound's solubility in the continuous phase (U gelstad et al.,
1980a). Thus, the effect of straight- and branched-chain alkanes was found
to be the same for compounds with equal solubility. Long-chain alcohols
also stabilize aqueous emulsions toward degradation by diffusion to an
extent determined mainly by their water slubility. This observation seems
to contradict the suggestion by Hallworth and Carless (1974) that the
formation of a condensed layer of emulsifier and fatty alcohol at the droplet
surface plays a significant stabilizing role.
The effect of relatively water-insoluble additives has been utilized by
Ugelstad (1978) and Ugelstad et al. (1978a,b, 1979a, 1980a,b) for the
preparation
of emulsions by diffusion. The main principIe of these methods
is that particles that consist wholly or partly of low molecular weight,
insoluble Z2 are .capable of absorbing much larger quantities of Zl than are
pure polymer partitles of comparable size. In practice, tbe preparation of an
emulsion by diffusional swelling is carried out by first producing an
aqueous dispersion of the water-insoluble compounds. This dispersed phase
may consist either of pure Z2 or of polymer particles into which Z2 has
been introduced in a first step. During the subsequent swelling of these
preformed dispersions with a slightly water-soluble Zl, conditions are such
that the only transport process possible is the diffusion of Zl through the
aqueous phase to become absorbed into the preformed droplets or particles.

388

J. Ugelstad el al.

Any diffusion of Z2 from the particles to the original phase of Z1 is hindered


by the low solubility of Z2' The swelling capacity of such particles may be
calculated from Eq. (8). Since in this case Zl is present as apure phase, the
appropriate equation describing the semiequilibrium state is
In cPla+ (1 - JjJ2)cP2a+ (1 - JjJ3)cP3a+ cPiaX12+ cPiaX13
+ (X12 + X13 - X23JjJ2)cP2acP3a + 2V1My(1/ra

l/rb)/RT

=O

(10)

where raand rb are the radii of the swollen particles and the droplets of Zl
at equilibrium. In case no polymer is present, the terms '1Vithindex 3 should
be omitted.
Figure 1 shows the volume (V1)of Z1 which, according to Eq. (10), may
be absorbed per unit volume of polymer (V3)+ Z2 (V2)for different values of
J2fJ1 as a function of y/ro. Here, y is the interfacial tension at equilibrium
and ro is the radius of the polymer-oligomer particles prior to swelling with
Zl' The values of the interaction parameters are arbitrarily chosen to be
X12 = X13 = 0.5 and X23 = O. Furthermore,
l/rb is considered negligible
when compared to l/ra.

o
3

4
Lag I ~/ro)

(nm-2)

Fig. 1.
Swelling capacity of polyrner-oligomer partic1es as a function of 'Ilrofor various
values of J2fJ. as ca1culated fram Eq. (10): ro = radius of polymer-oligomer partic1e prior to
swelling with ZI, V2 = V3= 0.5, X12= XI3= 0.5, X23= O, V.M= 10-4 m3/mol, T = 323 K.

11. The Formation of Monomer Emulsions and Polymer Dispersions

389

Figure 1 reveals that the swelling capacity of polymer particles containing 50% of a relatively low molecular weight, water-insoluble compound is
drastically increased compared to the case with pure polymer particles of
the same size (J2fJI = 00). It also appears that the swelling capacity of

polymer-Z2 particles is far more dependent on the value of y/re than is the
case for pure polymer particles. This has been verified experimentally by
Ugelstad et al. (1978b).Calculations of the swellingof pure droplets of Z2
with ZI give similar curves (Ugelstad et al., 1980a).
Figure 2 illustrates the swelling capacity dependence on V2 at a constant
value of J2fJI = 5. It appears that even low amounts of Z2 should give a
substantial increasein swellingcapacity at low values of y/re. This effecthas
been verified experimentally (Ugeistad et al., 1979a).
In the course of the swelling process, LlGlasteadily increases. This means
that

the driving

force, LlGlb

LlGla

--

LlGla (in case rb =

00), steadily

decreases. Figure 3 gives LlGla/RT as a function of the swelling ratio


V/(V2+ V3)for different values of y/re. It is seen that LlGla/RT is close to
zero even at values of VI that are far from the equilibrium value. This means
6
5

V:1.0
0.67
0.50

0.20

f'r 3

0.10
0.05
0.02

o
..J

21

0.01
O

o
3

t.

Log(~/ro)
Fig. 2. Swellingcapacity of polymer-oligomerparticles versus y/ro for different values
of v2: ro = radius of polymer-oligomer particles prior to swelling with Zt> V2+ V3 = 1,
J2/J, = 5, X12= X13= 0.5, X23= O, V1M= 10-4 m3/mol, T = 323 K.

390

J. Ugelstad et al.

0.002
0.000

- 0.002
- O.OO~

1~:;:-0.006
-0.008

-0.010
-0.012
-0.014
10

1000

10,000

VI/IV2>V3)

Fig. 3. Partialmolar free energy of ZI in swelling partic1es versus swelling ratio, for
various values of "t/ro: XI2 = XI3 = 0.5, X23= O, J2/J1= 5, V2= 2/3, V3= 1/3,
V1M= 10-4 m3/mol,T= 323K.

that the maximum obtainable swelling is very sensitive to factors that


would cause a deviation in the entropy of mixing from that given by the
Flory-Huggins expressions.
At equilibrium the activity of Zl in the continuous phase is equal to its
activity in the a and b droplets. Ir an excess of pure Zl is present and if the
radius of the b droplets is considered to be infinite, as in the case discussed
above, this activity is equal to unity. The swelling equilibrium is in this case
the same as the one that would be obtained for swollen particles in
equilibrium with a saturated solution of Zl in the continuous phase.
Ir Zl is present .as droplets of finite size, its activity at equilibrium (equal
in all phases) is larger than unity because of the interfacial free energy of the
b droplets. This in turn may lead to an increase in the degree of swelling. Ir
the b droplets are smaller than the a droplets, or become smaller during the
swelling, one may in principie obtain "infinite" swelling of the a droplets.
Finally, it may be desirable to emphasize some points regarding the general
application of Eq. (8) for calculations of swelling and equilibrium distribution of Zl between the various phases.
1. Ir one operates with an excess of Zl, the swelling of each phase may
be treated independently of the other phases. Ir there is a shortage of Zl,

11. The Formation of Monomer Emulsions and Polymer Dispersions

391

several cases of competitive swelling may be encountered in practice. They


may easily be dealt with by applying the correct expression for dGl to the
various phases. A typical example is the swelling of polydisperse droplets of
Z2' The final distribution is determined by the initial droplet sizes only. The
swelling of a mixture of monodisperse droplets of various Z2 cO'nstitutes
another example. In this case the final size distribution will be determined
by the value of J2 for the various compounds. A third possibility is the
swelling of a system consisting of two types of monodisperse polymer
particIes containing different amounts of the same Z2' The appropriate
equilibrium equation in this case is
In

<Pla/<Plb

+ (1 - JjJ2)(<P2a - <P2b)+ (1 - JjJ3)(<P3a - <P3b)

+ (<P2a<P3a- <P2b<P3b)(X12 + X13


+ (<P1a- <P1b)x12 + (<Pla

+ 2yV1M(1/raIn this equation,

X23JlfJ2)

<Plb)X13

1/rb)/RT= O

(11)

a and b denote the two different particIe sizes. Combining

Eq. (11) with a material balance for Zl allows the caIculation of the radius
of the two types of particIes as a function of the total amount of Zl
absorbed. Excellent agreement between caIculated and experimental results
has been obtained for such systems (Ugelstad et al., 1979b, 1980a).
2. Every compound that may be transferred between the phases within
the time scale of the experiment will equilibrate between the various phases.
This means that if for instance one type of Zl has been incorporated in a
polymer particIe in a first step, and another type of Zl is added in a second
step, the equilibrium' swelling of the particIe will not deviate noticeably from
the one obtained in the presence of either Zl' The reason for this is that the
type of Zl first absorbed will diffuse out of the particIes and become mixed
with the second type of Zb with the concomitant gain in entropy that this
mixing process involves. Similarly, if the system consists of droplets of the
insoluble Z2, or of polymer particIes containing Z2, and if a mixture of two
types of Zl with similar J values is added, the resulting swelling will be
approximately the ~ame as that obtained with either of the two types of Zl'
3. Equation (10) may easily be extended to incIude the swelling of
particIes containing any number of different Z2' Again it should be noted
that for a given particIe size, the presence of a mixture of two or more types
of Z2 will not in crease the degree of swelling beyond that obtained with the
same amount of one type of Z2 as long as the various types have similar J2

values. Points 2 and 3 (above) emphasize the fact that the driving force for
the swelling process is mainly the gain in entropy obtained by mixing of the
various components of the system.

392

J. Ugelstad el al.

111. Rate of Interphase

Transport

The rate of transport of ZI from phase b to another phase a may in


principie be governed by one of several mass transfer steps. These inc1ude
diffusion from the interior of b to the interface, transport across the
interface of phase b, diffusion through the continuous phase, transport
across the interface of phase a, and finally diffusion inside phase a.
Transport within a phase may be rate determining where a polymer
constitutes a major part of the phase and the temperature.is below the glass
transition point. Diffusion across the interfaces may be rate determining in
cases where polymeric emulsifiers are applied (Netschey el al., 1969; Napper
et al., 1971), whereas conventional surfactants probably do not constitute
any interfacial barrier to micron-sized droplet systems (Higuchi and Misra,
1962). For such a case, these authors derived an expression for the rate of
degradation of emulsions by molecular diffusion. Several papers on the
mechanisms of interphase transport (Goldberg el al.,. 1967; Goldberg and
Higuchi, 1969) and on the effect of interfacial barriers on interphase
transport (Ghanem et al., 1969; 1970a,b) have appeared.
For the present purpose, discussions will be restricted to cases where the
transport of ZI through the continuous phase is the rate-determining step.
The rate of transport from the bulk of the continuous phase to the surface
of the swelling partic1es (phase a) is 'then given by (Smoluchowski, 1918)
R

la

= 4nraNaDw(CI- Clas)

(12)

where CI is the concentration of ZI in the bulk of the continuous phase,


Clas is the concentration at the partic1e interface, Dw is the diffusion
constant of ZI in the continuous phase, and Na is the number of partic1es
with radius ra, Assuming that Henry's law may be applied, one has for low
values of CI
Clas

= Clooexp(AGla/RT)

(13)

where Cloo is the concentration in the continuous phase when this is


equilibrated with pure ZI in bulk (plane surface), Substituting for Clas in
Eq. (12),one obtains for the rate of swelling of a partic1es
dVIJdt

= Rla = 4nDwraNa[CI -

Cloo exp(AGla/RT)]

(14)

where the concentrations are given in v/v units. In the case where the
solution is saturated with ZI (relative to pure ZI in bulk), Eq. (14) may be
written
dVla/dl

= 4nDwraNaCloo[1- exp(AGIJRT)]

(15)

11. The Formation of Monomer Emulsions and Polymer Dispersions

393

If the swelling system consists of a solution of Zl and a number of a

particles, then C1 will decrease as the swellingproceeds unless a very large


excess of the solution of Zl is employed, in which case Cl may remain
approximately constant.
Usually, however, Zl is dispersed as droplets (phase b) and the rate of
transport from the droplets to the bulk of the continuous phase (Rlb) must
be taken into account. Setting Rla = Rlb, the followinggeneral expression
for the rate of swelling is obtained (Ugelstad et al., 1980a).
dVla/dt = 4nfaNafbNbDwClOO
[exp(dGlb/RT)
faNa

fbNb

- exp(dGla/RT)]

(16)

The values of dGla and dGlb may be calculated using the appropriate forms
of Eq. (5). Two limiting cases of Eq. (16) will now be treated.
Case 1
Equation (16) is then simplified to
dVlJdt

= 4nfaNaClooDw[exp(dGlb/RT)

exp(dGla/RT)]

(17)

This implies that the concentration of Zl in the continuous phase is kept


constant, equal to the solubility of ZI from the b droplets. Even if fb > fa
this sitnation may be realized if the swelling is carried out with a very large
excess of ZI' If f~ > 10 f.1.m,Eq. (17) is for all practical purposes equal to
Eq. (15).
Case 2

(faNa ~ fbNb)

This condition is \.1sually fulfilled for fa ~ fb when a moderate excess of


phase b is applied. In many cases the magnitude of fb is such that

exp(dGlb/RT) may be set equal to unity (with


dGlb
4 x 10- 5J/mol. Equation (16) may then be written
dVla/dt = 4nfbNbClct:,Dw[1- exp(dGla/RT)]

fb = 10 f.1.m,
(18)

Equation (18) illustrates the importance of subdividing phase b in order to


obtain high rates.of swelling. Since fbNb = 3Vb/4nf;,
therate is proportional
to the volume of monomer and inversely proportional to the square of the
droplet radius.
Figure 4 shows the results of some calculations of the degree of swelling,
expressed as V1alv.,as a function of time for some selected.cases. The
general Eq. (16) was used with Dw = 10-10 m2/sec and Cloo =
6 X 10-4 m3/m3, Le., values that approximately would be applicable to
styrene in water. It was further assumed that the radius of the b droplets f~
remained constant during the swelling, as this seems to be the more

394

J. Ugelstad el al.

180
160
140

o.
>

....

120
100

->

80
60
40
20

10

20

30

40

50

80

10
Tim_1

10

10

100

110

120

130

minI

Fig. 4;
Volume of 21 absorbed per unit volume of polymer-oligomer (v.O) versus time,
calculated
from
Eq. (16): Dw = \O-10 m2jsec,
CI", = 6 X \0-4 m3jm3,
r~ = 2 x \0-7 m,
V2a = V3a = 0.5 V.o,
Ya= Yb= 5 mNjm,
X12 = X23 = 0.5,
X23 = O,
J2fJI = 3,
VIM= \0-4 m3jmol, T=

323 K, rg and Vbv. as indicated.

realistic approximation when the b droplets consist of 21 only. The


decrease in the size of the b droplets, which might be expected to result from
transport of 21 from b to a droplets, is probably counteracted by coalescence and by diffusion among the b droplets. In practice, a given stirring
intensity will tend to give the same average droplet size independent of
"conversion."
Curve A in Fig. 4 illustrates a case where a large excess of small b
droplets

are assumed

to be present (r~

1 jlm). This would correspond

to

the case exemplified by Eq. (17). Curve B is caIculated for a case where
the concentration of 21 in the aqueous phase is constant, equal to the
saturation concentration relative to aplane phase (CI",).
In curve C, a large excess (Vb;v..= 105) of lO-jlm droplets

of 21 is

assumed to be present. The higher rate of swelling obtained in the case A as


compared to B and C is due to the interfacial free energy term which,
because of the lower value of r~ in case A, gives a higher value of L\Glb, i.e.

395

11. The Formation of Monomer Emulsions and Polymer Dispersions

a higher solubility of Zl in the aqueous phase. Curve B corresponds to the


case leading to Eq. (15), and as is seen from a comparison of curves B and C
in Fig.4, this equation is a good approximation when rg> 10 p.m and a
large excessof b droplets is present.
Curves F and D illustrate the effect of decreasing the diameter of the b
droplets at constant value of VbO
fV..0(= 100). Comparison of curves C, E,
and F shows the effect of decreasing the volume of the b phase at a constant
value of rg (= 10 p.m). For curve E, the general Eq. (16) must be applied,
although the term exp(LlG1b/RT) may approximately be set equal to unity.
Curve F may be described by the limiting case given by Eq. (18).
Figure 5 shows some experimental results (Mfutakamba et al., 1979) of
the swelling of 2.3-p.m polystyrene-dioctyladipate
particles (V3afV2a= 0.5)
with chlorobenzene at 308 K. When the swelling is carried out with
ordinary stirring and an excess of chlorobenzene corresponding to
16

D
14

12

.=

10

...

..

..
8
E
111
e
6

Time (hrl

Fig. 5. Swelling of polystyrene/dioctyladipate


particles
with
chlorobenzene:
r~=1.l5xl0-6m,
V2.JV3a=0.5, T=308K.
(A): V,v.=49, ordinary stirring (140rpm);
(B): V,0/V.o
= 49, Ultraturrax; (C): V,0v.0
= 290, ordinary stirring (140 rpm); (D): Vbv.= 290,
Ultraturrax.

396

J. Ugelstad el al.

VbO/Ya0 = 49, curve A is obtained. If the mixture of chlorobenzene, water,


and emulsifier is treated with a high-speed mixer (Ultraturrax) droplets of
chlorobenzene in the micron-size range are formed. When the polystyrenedioctyladipate particIes are added to this dispersion, curve B is obtained. As
expected, the reduction in droplet size leads to a substantial increase in the
rate of swelling.
Curves C and D are obtained with a chlorobenzene-particIe volume
ratio of 290: 1, the other conditions being the same as for curves A and B,
respectively. A comparison of curves A and C or of curv~s B and D cIearly
shows the effect of increasing the total volume of the b phase, in accordance
with what would be expected from Eq. (18),which should apply to this case.
It thus seems quite evident that in this case, the rate-determining step in
the swelling process is the mass transfer from the droplets of Zl to the water
phase.

IV.
A.

Preparation of Polymer Dispersions


Polymerization with lnitiation in Monomer Droplets Formed by the
Diffusion Process

Monomer emulsions prepared by the methods described above have


been applied to the preparation of polymer dispersions. The crucial point in
such applications is to establish conditions that ensure that the initiation
takes place in the monomer droplets. In practice, this requires that the
concentration of emulsifier in the aqueous phase during polymerization be
as low as possible and certainly below the CMC (Hansen and Ugelstad,
1979).
Polydisperse PVC latexes with particIe sizes in the range of 0.3-1.5 ,um
have been prepared by homogenization of different Z2 with water and
emulsifier to give stable emulsions of Z2 with droplet sizes from 0.050.3 ,um. When water and monomer are added to t~is preemulsion, a
monomer emulsion is formed by rapid diffusion of vinyl chloride into the
preformed droplets of Z2' Polymerization may be carried out by using
either water- or oil-soluble initiators (Ugelstad et al., 1978a, 1980a,b). A
similar method has been applied to produce acrylate latexes of very low
particIe size (Ugelstad et al., 1978a).
Another interesting application of this method is the preparation of
large, monodisperse polymer particIes. In this case the starting point is
small, monodisperse particIes (seed) in the size range of 0.3-1 ,um, produced
by ordinary emulsion polymerization according to procedures known to
give mono disperse particIes (Vanderhoff et al.,1956, 1970; Goodwin et al.,

11. The Formation of Monomer Emulsions and Polymer Dispersions

397

1973, 1976, 1978). Low molecular weight, water-insoluble compounds (oligomers) may be incorporated into the partic1es during their preparation by
addition of a chain-transfer agent during polymerization (Ugelstad, 1978;
Ugelstad et al., 1978b), or may alternatively be introduced into the preformed polymer partic1es. One way of doing this is to swell the polymer
particles with monomer and initiator and carry out the subsequent polymerization under conditions that lead to the formation of oligomers.
Another procedure requires the seed in a first step to be swollen with a
low molecular weight, water-insoluble Z2' (Ugelstad et al., 1978a,b, 1979a,

Fig.6.

Scanning

electron micrograph

of monodisperse

5.3-Jlm polystyrene

particles.

398

J. Ugelstad el al.

1980a.) One way of facilitating this process is to add a water-soluble solvent


to increase the solubility of Z2 and thereby make possible its diffusion
through the continuous phase to become absorbed in the seed particIes. In
a next step the monomer is added to the dispersion of polymer-oligomer or
polymer-Z2 particIes. The experimental conditions in this second step are
adjusted so that no transport of oligomer or Z2 out of the particIes is
possible during the swelling with monomer. The mono disperse monomer
droplets formed by this procedure are subsequently polymerized, preferably
with an oil-soluble initiator. Figure (6) shows a scanning electron micrograph of mono disperse 5.3-j.lm polystyrene particIes produced according to
this method. ParticIes up to about 15 j.lm have been prepared with a
standard deviation of about 0.6% (measurements performed by Becton and
Dickinson with an "Ultra-Flow" high resolution particIe analyzer).
Monodisperse particIes may also be produced with a cross-linked structure, and monodisperse porous particIes may be obtained (Ugelstad el al.,
1980a) by applying methods known from suspe,nsion polymerization.
ParticIes with functional surface groups have been prepared by chemical
modification of the surface' of cross-linked monodisperse particIes of
styrene-di.vinylbenzene or by copolymerization with monomers containing
the desired functional groups.
In principIe, the two-step swelling procedure discussed above may be
applied to swell polymer particIes with a quite different type of monomer to

Fig. 7.
Electron micrograph of particles resulting from polymerization of a polystyrene
seed (500 nm) swollen with a mixture of acrylic monomers (magnification 4080 x ).

11. The Formation of Monomer Emulsions and Polymer Dispersions

399

obtain a homogeneousmixture of polymer Z2 and the second monomer. By


the subsequent polymerization, however, a phase separation will normally
take place. The reason for this, as was discussed in connection with Eq. (2),
is the low entropy of mixing in the case of two polymer substances. Most
interesting in this context is the fact that the original polymer particle may
be restored as a separate phase in the course of the polymerization. This
particle may be partly expelled from the final particle which consists mainly
of the polymer formed from the second monomer. The phenomenon has
been observed in the case of polystyrene swollen with a mixture of styrene
and methacrylic acid and even more so when polystyrene is swollen with
methyl methacrylate or mixtures of methyl methacrylate and hydrophilic
acrylates. An example of this phenomenon is shown in Fig. 7. Applying the
two-step swelling procedure, a monodisperse polystyrene seed was swollen
with a mixture of acrylate monomers. The droplets formed appeared to
consist of a homogeneous mixture of the monomers, Z2 and polystyrene.
After polymerization the original polystyrene particle is almost completely
expelled from the acrylic polymer particle.
B.

Emulsification of Polymer Solutions

The recent trend to avoid organic solvents has led to substantial elforts
to prepare aqueous dispersions of polymers that cannot be prepared
directly by emulsion polymerization. This includes such important polymers as epoxy resins, polyurethanes, and silicones. One method. to achieve
this is by emulsifying solutions of the polym~r by homogenization, followed
by evaporation of the solvent. Most often this method gives particles in the
range of 1 to 211m. 1'his may be a disadvantage in that the dispersion is
unstable toward settling of the particles on storage, and furthermore, large
particles tend to give inferior film properties. Vanderholf et al. (1979)
showed that addition of small amounts of Z2 to the solution of a polymer
before homogenization led to much more finely dispersed droplets of the
polymer solution. A number of dilferent polymer dispersions have since
been prepared (EI-Asser et al., 1977a,b; Vanderholf et al., 1978; Miscra et
al., 1978).
Ugelstad (1978) and Ugelstad et al. (1978a) have explained the elfect of
adding small amounts of Z2, along the same lines as the thermodynamic
treatment above. In the present discussion the previous explanation of the
effect of Z2 is somewhat extended.
The solvents applied in practice are all slightly water soluble, i.e., Zl
types. At a given composition of the solution of polymer (Z3) and solvent
(Zl) there is a theoretical minimum in the droplet size that may be obtained
before phase separation should occur. A droplet of this minimum size

400

J. Ugelstad el al.

contains a given amount of polymer. The minimum droplet size corresponds to the swelling capacity of a partic1e consisting of this amount of
polymer. Therefore, the minimum droplet size will in crease with increasing
amount of solvent (ZI)' If a small amount of Z2 is added before homogenization, the droplets formed contain a small amount of Z2 in addition to Z3'
The swelling capacity of a partic1e in which even a small amount of Z3 is
replaced by Z2 is considerably higher than that of pure Z3' Therefore, in the
presence of Z2 the. minimum diameter that may be obtained with a given
amount of solvent (ZI) may be drastically reduced. 'Calculation of this
minimum diameter may be carried out by applying Eq. (10). An example of
such a calculation is given in Table 1. It appears from the table -that the
minimum diameter in creases with increasing amounts of ZI' It also appears
that even with small amounts of Z2 it is possible to obtain small partic1es.
This is also the case where without Z2 one gets phase separation even at
infinite diameters of the partic1es.
During homogenization of a solution of polymer (Z3)' solvent (Z), and
water, droplets of different sizes are formed. Becase of the interfacial
energy difference, a transport of ZI from the smaller to the larger droplets
will take place. If droplets smaller than the critical size are formed, these
droplets, especially, will rapidly lose ZI to the surrounding larger droplets.
The viscosity within these small droplets will increase and a further
degradation may be mechanically hindered. In the presence of Z2, two
effects that will facilitate subdivision of the emulsion may be encountered.
1. The degree of diffusion between small and large droplets will be
reduced because the presence of Z2 means that a counteracting concentration potential is more rapidly established.
TABLEI
The Minimum Diameter That May Be Obtained by Homogenization of a Polymer
Dissolved in a Solvent (Z,) as a function of Volume V, of the Solvent
and Volume V2 of a Z28.b

X = 0.5

= 0.3

O
0.01
0.02
0.10

10

10

21
20
19
16

110
103
96
66

415
370
334
190

48
44
40
24

454
338
269
105

2980
1500
1005
280

147
108
85
32

From

Ugelstad

b XI2 =

O, XI3

T= 300 K.

X =0.6

(1978)

= X23

with

10

2545 3 x 105
146
356

permission.

= X, V2+ V3= 1, J2fJI = 5, V1M= 10-4 m3/mol, y.= 5X10-3 N/m,

11. The Formation of Monomer Emulsions and Polymer Dispersions

401

2. The presence of Z2 tends to decrease the critical minimum size. In


practice this means that the formation of droplets with a higher viscosity,
due to transport of ZI from small to large droplets, is reduced. The presence
of Z2 makes possible the formation of small droplets with relatively high
amounts of ZI, and therefore low viscosity, without getting below the
critical value of the droplet size.

V.

Effect of Addition ofWater-Insoluble Compoundsto the MonomerPhase

If in an ordinary emulsion polymerization, a water-insoluble Z2 is added


to the monomer phase, the etfect will obviously be to decrease the activity of
the monomer in this phase and accordingly to decrease the concentration of
the monomer in the partic1es (Azad et al., 1980; Ugelstad et al., 1980b,c).
The appropriate form of the equilibrium equation for the case in which one
has ZI equilibrated between Z3 partic1es and monomer droplets containing
Z2 will be
In 4>13 + 4>33

4>f3X13

= In 4>lb + (1 -

+ 2VMy/r3RT
JjJ2)4>2b

4>ibXl2

+ 2V1My/rbRT

(19)

where index a and b refer to polymer partic1es and monomer droplets,


respectively, and J3 is set equal to infinity. If the monomer droplet radius
(rb)is very large the last term may be neglected.
It appears that if y/r3is assumed to be constant, any equilibrium value-of
4>2b will correspond to a value of 4>13 that is independent of the initial
monomer-polymer
nitio.
Such a situation was considered by Azad el al. (1980) in their discussion
of the etfect of additives to the monomer phase on the degree of swelling
and on the kinetics of polymerization. A more relevant approach to this
problem might be to consider the etfect of a given amount of Z2, V2, or a

given initial volume fraction of Z2, 4>Rb' in the monomer phase on 4>13/4>13
(V2= O)(Le.,the ratio of the volume fractions of monomer in the partic1es
with and without addition of Z2 to the monomer phase) for various values
of vNV30,the initial.ratio of monomer to polymer.
In order to illustrate the etfects one may expect and the problems in
interpretation of experimental results, some ca1culationsshowing the etfect
of addition of Z2 to the monomer phase have been carried out using
Eq. (19).
In Figures 8 and 9, 4>13/4>13
(V2= O)is given as a function of VIO/V30for
various values of 4>fb = V2/(V2+ VIO)and V2, respectively.In these ca1culations, the radius of the polymer partic1es is assumed to be r~ = 5 x 10- 8 m

400

J. Ugelstad el al.

contains a given amount of polymer. The minimum droplet size corresponds to the sweIling capacity of a particIe consisting of this amount of
polymer. Therefore, the minimum droplet size wiII increase with increasing
amount of solvent (21), If a smaIl amount of 22 is added before homogenization, the droplets formed contain a smaIl amount of 22 in addition to 23'
The sweIling capacity of a particIe in which even a smaIl amount of 23 is
replaced by 22 is considerably higher than that of pure 23' Therefore, in the
presence of 22 the minimum diameter that may be obtained with a given
amount of solvent (21) may be drasticaIly reduced. Calculation of this
minimum diameter may be carried out by applying Eg. (lO). An example of
such a caIculation is given in Table 1. It appears from the table 'that the
minimum diameter increases with increasing amounts of 21, It also appears
that even with smaIl amounts of 22 it is possible to obtain smaIl particIes.
This is also the case where without 22 one gets phase separation even at
infinite diameters of the particIes.
During homogenization of a solution of polymer (23), solvent (21), and
water, droplets of different sizes are formed. Because of the interfacial
energy difference, a transport of 21 from the smaIler to the larger droplets
wiII take place. If droplets smaIler than the critical size are formed, these
droplets, especiaIly, wiII rapidly lose 21 to the surrounding larger droplets.
The viscosity within these smaIl droplets wiII increase and a further
degradation may be mechanicaIly hindered. In the presence of 22, two
effects that wiIIfacilitate subdivision of the emulsion may be encountered.
1. The degree of diffusion between smaIl and large droplets wiII be
reduced because the presence of 22 means that a counteracting concentration potential is more rapidly established.
TABLEI
The Minimum Diameter That May Be Obtained by Homogenization of a Polymer
Dissolved in a Solvent (Z1) as a function of Volume V1 of the Solvent
and Volume V2 of a Z2a.b
X = 0.3

x = 0.5

x = 0.6

>\'

10

10

O
0.01
0.02
0.10

21
20
19
16

110
103
96
66

415
370
334
190

48
44
40
24

454
338
269
105

2980
1500
1005
280

147
108
85
32

From Ugelstad

b X12

10

2545 3 x 105
146
356

(1978) with permission.

= O, XI3= X23= X, V2+ V3= 1, J2/11= 5, VIM= 10-4 m3/mol, y = SX10-3 N/m,

T= 300 K.

11. The Formation of Monomer Emulsions and Polymer Dispersions

401

2. The presence of Z2 tends to decrease the critical minimum size. In


practice this means that the formation of droplets with a higher viscosity,
due to transport of ZI from small to large droplets, is reduced. The presence
of Z2 makes possible the formation of small droplets with relatively high
amounts of ZI, and therefore low viscosity, without getting below the
critical value of the droplet size.

v.

Effect of Addition ofWater-Insoluble Compoundsto the MonomerPhase

If in an ordinary emulsion polymerization, a water-insoluble Z2 is added


to the monomer phase, the effect will obviously be to decrease the activity of
the monomer in this phase and accordingly to decrease the concentration of
the monomer in the partic1es (Azad et al., 1980; Ugelstad et al., 1980b,c).
The appropriate form of the equilibrium equation for the case in which one
has ZI equilibrated between Z3 partic1es and monomer droplets containing
Z2 will be
In I/Jla

+ 1/J3a +

I/Jfax13+ 2VIMy/raRT

= In I/Jlb+ (1 - JjJ2)1/J2b+ l/JibX12+ 2VIMy/rbRT

(19)

where index a and b refer to polymer partic1es and monomer droplets,


respectively, and J3 is set equal to infinity. If the monomer droplet radius
(rb) is very large the last term may be neglected.
It appears that if y/ra is assumed to be constant, any equilibrium value-of
1/J2bwill correspond to a value of I/Jla that is independent of the initial
monomer-polymer ratio.
Such a situation was considered by Azad et al. (1980) in their discussion
of the effect of additives to the monomer phase on the degree of swelling
and on the kinetics of polymerization. A more relevant approach to this
problem might be to consider the effect of a given amount of Z2, V2, or a
given initial volume fraction of Z2, I/J~b'in the monomer phase on I/JIJI/JIa
(V2 = O) (Le., the ratio of the volume fractions of monomer
in the partic1es
with and without addition of Z2 to the monomer phase) for various values
of vNv30, the initiaf ratio ofmonomer to polymer.
In order to ilIustrate the effects one may expect and the problems in
interpretation of experimental results, some ca1culations showing the effect
of addition of Z2 to the monomer phase ha ve been carried out using
Eq. (19).
In Figures 8 and 9, I/JlaNla (V2= O) is given as a function of VIIV30for
various values of I/Jfb = V2/(V2 + VIO)and V2, respectively. In these ca1culations, the radius of the polymer partic1es is assumed to be r~ = 5 x 10-8 m

402

J. Ugelstad et al.

1.0
0.9
0.8

-o

0.7

N
>

0.6

"

o
02b

"
61
.....
!!
61

0.2
o
02b =0.3

0.5

0.1,

0.3

0.2

0.1

1.0

2.0

5.0

3.0
VI/V3

Fig. 8.

Ratio of volume fraction of ZI in a partic1es to the same volume fraction when

V2 O versus initial ratio of ZI to Z3 for various values of the initial volume fraction of Z2 in
the b phase: r~ = 5 x 10-8 m, X12 = X13 = 0.5, V1M= 10-4 m3/mol, V2M= 3 X 10-4 m3jmol,

y = 5 mNjm,J2jJ = 3, T = 323K.

and the monomer droplet radius is set equal to infinity. The values of X12
and X13are arbitrarily chosen to be 0.5 and the molar volumes of 21 and 22
are set equal to 10-4 and 3 x 10-4 m3/mol, respectively. The interfacial
tension at equilibrium is assumed to be 5 mN/m and T = 323 K.
It would

appear

from Fig. 8 that

<PlaNla

(V2

= O)

passes

through

minimum as VIo/V3oincreases. This minimum occurs at a value of VIO


/V30
corresponding to that of maximum swelling of the polymer particles in the
absence of 22 and is less pronounced as the value of <Pgbincreases. This
position of the minimum is to be expected since the transpurt of monomer
from the droplets to the polymer particles wilI lead to the highest relative
increase in <P2bat this ratio of VIO
/V30.At increasing values of V?/V30the ratio
of <Pla/<Pta
(V2= O)wilI approach an upper, constant value. It also appears
that at low values of vNv30 the effect of <Pfbon <PlaNta (V2 = O) is markedly

11. The Formation of Monomer Emulsions and Polymer Dispersions

0.9

V2=0.S
V2=1.0

0.8

403

0.7

11
N

0.6
ni
SI
-ni
SI

o.S
0.4
03

0.2
01
2

IC

1,

14

16 18 20

v / V3
Fig. 9.
cP.Jc/>1a
(V2= 0.)versus vNvJoat various 'values of the volume (V2) of Z2 in the b
phase. V~ = 1. Parameter~ equal to those in Fig. 8.

reduced. AIso in Fig. 9 a minimum in the volume fraction ratio is observed


at low values of V2. As expected, the minimum disappears as V2increases.
When the etTect of Z2 on the rate of polymerization is determined
experimentally, it should be kept in mind that with a given value of </Jfb'the
concentration of monomer in the particles will, in principIe, vary over the
entire concentration range, even in the range that without addition of Z2
would correspond to Smith-Ewart lnterval n.
The relevant equation in the case of simultaneous sweHingand polymerization is obtained from Eq. (19) by introducing
V3a = V30

+ (VIO -

ra = r~[(V3a
VI

= V1a+

Vlb)(djd3)

+ Vla)/V30]

1/3

V1b

where di and d3 are the densities ofmonomer and polymer, respectively.

404

J. Ugelstad el al.
8
7
6

-...
E S
't:I
...

o
E 4
..
u 3

0.1

0.2

0.3

0.4

0.5

Q6

0.7

0.8

0.9

10

Fig. 10.
Concentration of monomer in the polymer particJes during polymerization as
a function of conversion when Z2 is present in the monomer phase: V2= 0.1, V30= 1, VIOas

indicated, r~= 5 x 10-8 m, VIM= 10-4 m3/mol, J2/JI = 3, X12= X13= 0.5, y = 5 mN/m,
T= 323K.

Figure 10 gives CIa as a function of the conversion of monomer, defined


VI)/VIO for V2 = 0.1 and V30= 1 at variousvaluesof vt The
ratio d/d3 is set equal to 0.82, the other parameters being the same as those
used in Figs. 8 and 9. It would appear that the value of CI. also decreases in
Interval n, even in the case where y is kept constant. In the case where y/r.
remains constant, the decrease in CI. will be more pronounced.
The effect of the additive on the rate will of course depend upon whether
we operate in Interval n or In, and moreover on whether Smith-Ewart
Case 1, 2, or 3 is operating. Figure 11 gives an example of experiments
with seeded styrene polymerization with different amounts of hexadecane

as 1J= (V? -

added. The value of VIO/V30 is in this case

20, the value of 4>fbis 0.08 and

0.16, respectively.
It would appear that the initial rate is lower the higher the value of 4>fb'
as would be expected under Smith- Ewart n conditions from the lowering of
the concentration of monomer in the particles. At increasing conversions it

405

11. The Formation of Monomer Emulsions and Polymer Dispersions


195,- -

- -

- -

150
o'"

J:
....
E

<f)
o..

I
100

DI

-~

50

.::.

0.6g K2S20a/dm3 H20

H20

10"10Hexadecane
O

0.6 g K2S20a/dm3
20"10 Hexadecane

H20

Seed: 10g /dm3 H~


N

10

~--3
0.6 g K2S20a/ dm

3.010'6 part.l dm3H20

~-l
60

120

lao

21,0

300

360
TIME

420

I,ao

51,0

600

(min)

Fig. 11.
Conversion versus time curves for seeded emulsion polymerizations
with hexadecane present in the monomer phase. T = 333 K.

of sty~ene

turns out that when hexadecaneis present, the rate of polymerization


increasesand becomeshigher than in the casewithout hexadecane.This is
most probably due to a Trommsdortf etfect.
In the caseof vinyl chloride, the additiori of Z2 to the monomer phase
leads to an increase in rate from low conversions on (Fig. 12). This result is
in accordance with the results of Ugelstad et al. (1967)who found that the
rate of polymerization of vinyl chloride, at constant particle number,
increased when the reaction was carried out at subsaturation pressures.
In the above calculations we ha ve assumed that the monomer is evenly
distributed in the polymer particles. In this case ole finds that at low values
of V?fV30 the value of <P'a/4J'a(V2 = O) is relatively high, even at relatively
high values of 4>fb' From Fig. 8 it is seen that at a value of V,ojV30= 0.05,
the value of

<P'a/4J,a(V2

= O) is

0.95 at <Pfb = 0.5. In cases where monomer

is

added continuously, Le., semicontinuous or continuous polymerization, one


has "starving" conditions with respect to monomer concentration. It has
been suggested that in such cases, polymerization may take place in an
outer shell of the particles. Figure 13 gives the result of a calculation of the
swelling in an outer shell of thickness 50 A, expressed as 4>la/4J'a(V2
= O),
versus the initial monomer-polymer volume ratio for various values of <Pfb'
This assumption implies that the etfective polymer particle volume is

406

J. Ugelstad et al.

300

oN

:x:
PI

e
~ 200
(.)

>
o.
el

100

100
Fg. 12.
chloride

200
TIME (min)

300

Conversions versus time curves for seeded emulsion polymerization of vinyl

with hexadecane

present

in the monomer

H20. Seed: 3.6x 1016 particles/dm3 H20,


hexadecane.

phase.

[K2S20SJ

= 2.7

x 10-3 moljdm3

T= 323 K. A: without hexadecane: B: 20%

reduced by a factor that decreases as the particle radius increases. In the


present example (r~ = 0.05 Jlm) this factor is initially 0.27. The relevant
equation for calculating the curves in Fig. 13 is obtained from Eq. (19) by
inserting the following relationships
r.

= raO[(V3o + V1s)fV30] 1/3

Vs = V30(3r;c5

- 3rac52 + c53)/(r~)3

= V1sIVs

<Pla

where VIS is the volume of monomer in the shell of thickness c5and total
volume Vs. The other parameters used in calculating the curves in Fig. 13
are the same as in Fig. 8.
Comparing Fig. 13 with Fig. 8 it appears that in this case one may
expect a considerably larger effect of the additive. Thus, for VIO
fV30= 0.05
and <Pfb= 0.5, the value of <Plal<Pla(V2
= O) is 0.75 when the swelling is

11.

The Formation of Monomer Emulsions and Polymer Dispersions

1.0

407

..

0.9
0.8
a7

-o

"
N
>
..'"

ii
.....

02b

0.2

0.6

0.5

SI
0.4
0.3

02

0.1

05

1S

10

2.0

V~/V3

Fig. 13.

Swellingof a 50- outer shell of a

Ratio of volume fraction of monomer

polymer particle with radius 5 x 10- 8 m.


in the shell with and without Z2 present in the monomer

phase versus initial volume ratio of monomer


given in Fig. 8.

to polymer.

Other parameters

are equal to those

limited to a sheIl of thickness 50 A, as compared to 0.95 when the monomer


is evenly distributed in the polymer particIe.
Azad et al. (1980) applied a ratio of VNV30 3.5 and measured the rate
of polymerization with different additives in the monomer phase. With
addition of octadecane in an amount corresponding to <P~b= 0.10 they
found that the rate of reaction was constant up to -40% conversion. This
result is surprising since with the given X12and X13values of 0.8 and 0.43,
respectively, and with the assumption of a constant y/ra, one should expect a
decrease in concentration of about 20 %according to the above calculation.
In reality the value of y/ra may decrease somewhat during polymerization.
::::::

408

J. Ugelstad el al.

This will tend to counteract the effect of an increasing CP2b.AIso, as they


pass into Interval III they may have had a Trommsdorff effect counteracting
the decrease in the concentration of monomer.

VI.

Emulsification with Mixed Emulsifier Systems

One of the metl.ods described above for preparing stable emulsions of ZI


involved homogenization of Z2 with water and emulsifier. to produce small
droplets of Z2 which served as loci for the subsequent preparation of an
emulsion of ZI by diffusion. A different method which makes the use of a
homogenization process unnecessary involves the use of a mixture of ionic
emulsifier and long-chain fatty alcohols. This procedure seems to have been
known industrially for quite a long time.
The first paper that clearly demonstrated that this method could be used
to prepare stable emulsions of monomer and that the subsequent polymeri~
zation might take place with initiation in monomer di-oplets was presented
by Ugelstad el al. (1973a). Since then a number of papers on this subject
have been published (Ugelstad el al., 1973b, 1974; Hansen et al., 1973, 1974;
Lange et al., 1974; Azad et al., 1975). The process is carried out by mixing
water, fatty alcohol, and emulsifier at a temperature above the melting
point of the fatty alcohol. ZI is then added at ordinary stirring and within a
few minutes, afine emulsion with droplet sizes in the range of 0.2 to 1.5 11m
is obtained. In the case where ZI is a monomer, subsequent polymerization
may lead to initiation in the monomer droplets. A typical recipe may
consist of 0.2-0.5 g ionic emulsifier, 0.2-1 g fatty alcohol, 50-100 g
monomer, and 100 g water. The following points should be noted:
1. The fatty alcohol and the ionic emulsifier should be mixed with
water at a temperature above the melting point of the fatty alcohol before
addition of monomer. If such small amounts of fatty alcohol are dissolved
in the monomer before addition to a mixture of water and ionie emulsifier,
only very coarse and unstable emulsions are formed.
2. The fatty alcohol that is to be used should have a chain length of at
least 16 carbon atoms.
3. The ratio of fatty alcohol to ionie emulsifier should be approximately
in the range of 1:1 to 4: 1 on a molar basis.
4. If fatty alcohol s with chain length equal to or less than 16 carbon
atoms are applied, the emulsion formed is relatively unstable. In order to
get initiation in monomer droplets, the polymerization should then be
carried out immediately after preparation of the monomer emulsion.
This emulsification process thus exhibits the interesting phenomenon of a
rapid and easy formation of an emulsion that is relatively unstable toward

11. The Formation of Monomer Emulsions and Polymer Dispersions

409

degradation by diffusion. The fact that with fatty alcohols below a certain
chain length one gets very unstable emulsions indicates that also in this case
the degradation by diffusion is a major cause of instability and that a fatty
alcohol may act as a "Z2" whose stabilizing effect is determined by its
solubility in the continuous phase. This does not exc1ude the possibility that
a condensed layer of emulsifier and fatty alcohol is formed at the interface.
Such a complex will in crease the stability toward coalescence and may also
decrease the rate of degradation by diffusion by forming a barrier to
transport of Z through the interface (HalIworth and Carless, 1974). Such a
hindered transport through the interface was also suggested as an explanation for the decrease in the rate observed in the seeded polymerization of
vinyl chloride when mixtures of ionic emulsifier and fatty alcohols were
applied (Ugelstad el al., 1973b).
The ease of formation of these emulsions is still not satisfactorily
explained. When an ionic emulsifier and a fatty alcohol are dissolved in
water, mixed micelIes, liquid crystals, and crystalline, rodlike partic1es (1-2
..mlength, 0.1-0.2 ..mdiameter) are formed (Chou el a/., 1980). Chou el a/.
carried out an extensive study of the formation of emulsions with mixed
emulsifiers. The emulsifier systems were prepared by heating water with
fatty alcohol and cationic emulsifier. To this mixture styrene was added
gradualIy, and the resulting emulsion was examined spectroscopicalIy and
by mesurement of the electrical conductivity. The conductivity of the mixed
emulsifier system was found to be considerably lower than that of an
equimolar solution of pure ionic emulsifier, and more so the higher the fatty
alcohol-emulsifier
ratio. This result is expected since the complexation of
the emulsifier with the fatty alcohol leads to a reduction in the amount of
free emulsifier in the' aqueous phase. When styrene was added, the conductivity at first decreased and then increased on further addition of styrene
before the curve leveled off or started to decrease slightly. With high fatty
alcohol-ionic
surfactant ratios, when the conductivity before addition of
styrene was at its lowest, the conductivity increased from the start when
styrene was added.
The most stable emulsions were formed with a molar ratio of fatty
alcohol to ionic emulsifier of about 3: 1, whereas higher ratios gave more
unstable emulsions. Similar results were obtained using systems with anionic emulsifier (Hansen et a/., 1973).
Chou el al. observed that addition of relatively smalI amounts of styrene
caused the rodlike crystals to disappear. They did not come to any
conc1usion about the mechanism of the emulsification process but suggested
that the droplets may have been formed by diffusion of styrene into the
crystalline, rodlike partic1es.
Chou el al. ascribed the stability of the emulsions to the formation of an
emulsifier-fatty alcohol complex at the surface of the droplets. As discussed

410

J. Ugelstad el al.

above, it seems plausible that even if such a complex may be formed and
even if it increases stability toward coalescence, the major stabilizing effect
is the prevention of degradation by diffusion, caused by the fact that a longchain fatty alcohol may act as a Z2'
Another explanation that has been advanced by Azad et al. (1975)
implies that when stirring the mixture, the mixed micelles collide with
droplets of Zt with the result that fatty alcohol is transferred to the
droplets. During this process, a transient high concentration of fatty alcohol
in the outer layer of the droplets of Zt may cause a spontaneous emulsification of parts of the droplet. It is well known that spontaneous
emulsification may be obtained with 20-30% fatty alcohol (Schulman et al.,
1940).
Another explanation that has more recently been advanced by Ugelstad
and Fitch (to be published) is based on the principIe of formation of
emulsions by diffusion. During stirring, droplets of Zt with fresh surface
and therefore a high interfacial tension are formed. Diffusion of Zt from the
surface of these droplets into neighboring mixed micelles takes place
rapidly. The swelling of the mixed micelle is determined by its fatty alcohol
content, acting as a Z2'
As the mixed micelles are very small (10-20 nm) the initial swelling must
be very limited. To be able to absorb more of Zt, the initial small droplets
must be furnished with more emulsifier, and even more importantly, with
more fatty alcohol (Z2)' This may be achieved by coalescence of initial
droplets or by absorbtion of mixed micelles from the surroundings. The
assumption that the emulsification takes place by a diffusion process seems
to be supported by experiments with mixed systems of ionic emulsifier and
fatty alcohol and various dispersed phases, showing that a necessary
condition for a rapid emulsification is that the compound to be emulsified
have slight water solubility. Furthermore, it has been observed that if even
small amounts of Z2 are added to Zt before addition to the water-mixed
emulslfier system, the extent of emulsification is reduced and the resulting
emulsion becomes less stable.
Another interesting phenomenon is that post-addition of small amounts
of a highly insoluble compound to an emulsion leads to a rapid degradation
of the emulsion (J. Ugelstad, unpublished results). It has also been observed
that if the molar ratio of fatty alcohol to emulsifier is increased above four,
at a constant concentration of emulsifier, the stability of the emulsions
decrease (Lange et al., 1973; Chou et al., 1980). In all these cases, a bulk
phase containing Z2 will be present, even if an emulsion of Zt is formed.
Thus, the emulsion may degrade by diffusion of Zt from the emulsion
droplets to the bulk phase containing Z2'
The increase in conductivity observed by Chou et al. at increasing

411

11. The Formation of Monomer Emulsions and Polymer Dispersions

additions of styrene may be explained as resulting from a dissolution of


fatty alcohol in the interior of the styrene droplets, leading to a liberation of
free emulsifier.

List o Symbols
ZI
Z2

aw
a and b
A
CI
Cta.
Clo:>
di
Dw
dG
t:.G
t:.Gi

low molecular weight compound(s) that are equilibrated between phases


relatively low molecular weight compound(s) that may be effectively hindered
being transferred between phases
polymer(s) insoluble in the continuous phase
activity of component i in the continuous phase
type of phases or partic1es
surface area of partic1es or droplets
concentration
of ZI in the bulk of the continuous phase
concentration
of Z. in the continuous phase at the droplet interface of phase a
concentration
of Z. in the continuous
bulk (plane surface)

phase when this is equilibrated

density of component i
diffusion constant of Z. in the continuous phase
differential molar free energy
free energy of mixing
partial molar free energy of mixing for any component
surface) as reference state
partial molar
surface

free energy of mixing of component

from

with pure ZI in

i, with pure i in bulk (plane

i when the mixture

has aplane

t:.Gix
t:.GxO
i
j
Ji
N.
Nb
Ni
P

type of
number
number
number
number
pressure

rO
x
R

partic1e or droplet radius


partic1e or droplet radius indexed (x = a or b) to distinguish between two different
sizes andjor partic1e compositions
initial rx befQre any mass transport has taken place in or out of the partic1es
gas constant
.

Rix

(dV;Jdt)

S
t
T
V;
"Ix
"1M

V;~

t:.Gi in phase x (x = a or b)
initial t:.Gix before any mass transport
type of component
component
of segments
of partic1es
of partic1es
of moles of

rate

has occurred

between phases

in component i
with radius r.
with radius rb
component i

of transport

ZI from

the bulk

of the continuous

phase

of the swelling partic1es


entropy
time
absolute

temperature

volume of component i
volume of component i in phase x
partial molar volume of component
initial "Ix before any mass transport

i
has taken place between phases

to the

surface

412
v.

J. Ugelstad et al.

Vs

segment volume
total volume of outer shell (of thickness

vIS
i'
b

volume of compound
interfacial tension
shell thickness

bi
r
rPi
rPix

Hildebrand solubility parameter for component i


degree of conversion of monomer, defined by (V.o V)!Vo
volume fraction of component i
volume fraction of component i in phase or particle type x

b) of a particle

21 in outer shell of particle

rPi~

initial

Xi.;

interaction parameter per mole of component i with component j

xi.;

Xi)Ji

rPix

before any mass transport

= interaction

parameter

has taken place between t~e phases

per segment of component

i with component

References
Azad, A. R. M., Ugelstad, J., Fitch, R. M., and Hansen, F. K., (1975). Paper presented at Symp.
Emul. Polym. Philadelphia, Pennsylvania Vol. 1. American Chemical Society Ser. 24,
Washington, D.C.
Azad,A. R. M., Nomura, M, and Fitch, R. M. (1980). Org. Coat. Plast. Chem. Preprints43, 537.
Chou, Y. J., EI-Asser, M. S., and Vanderhoff, J. W. (1978). Paper presented at Symp. Phys. Chem.
Properties Colloidal Partides, Miami.
Chou, Y. J., EI-Asser, M. S., and Vanderhoff, J. W. (1980). "Polymer Colloids" (R. M. Fitch,
ed.), Vol. 11.Plenum Press, New York.
Davies, S. S., and Smith, A. (1974). Paper presented at Symp. Theory Practice Emul. Technol.
Brunel Univ. p. 325. Academic Press, New York.
EI-Asser, M. S., Vanderhoff, J. W., and Poehlein, G. W. (1977a). Preprints Am. Chem. Soc. Div.
Organ Coating Plast. Chem. 37, 92.
EI-Asser, M. S., Miscra, S. C., Vanderhoff, J. W., and Manson, T. A. (1977b). J. Coatings
Technol. 49, 71.
Flory, P. J. (1965). J. Am. Chem. Soco 86, 1833.
Ghanem, A. H., Higuchi, W. l., and Simonelli, A. P. (1969). J. Pharm. Sci. 58, 165.
Ghanem, A. H., Higuchi, W. l., and Simonelli, A. P. (1970a). J. Pharm. Sci. 59, 232.
Ghanem, A. H., Higuchi, W. l., and Simonelli, A. P. (1970b). J. Pharm. Sci. 59, 659.
Goodwin, J. W., Hearn, J., Ho, C. C., and Ottewill, R. H. (1975). Br. Po/ym. J. 5, 347.
Goodwin, J. W., Hearn, J., Ho, C. C., and Ottewill, R. H. (1976). Colloid Po/ym. Sci. 60,173.
Goldberg, A. H., and Higuchi, W. 1. (1967). J. Pharm. Sci. 56, 1432.
Goldberg, A. H., Higuchi, W.I., Ho, N. F. H., and Zografi, G. (1969). J. Pharm. Sei. 58,1341.
Hallworth, G. W., and Charles, J. E. (1974). Paper presented Symp. Theory Practice Emul.
Technol., Brunel Univ. p. 305. Academic Press, New York.
Hansen, F. K., Ugelstad, J., and Lange, S. (1973). Abstr. Seand. Symp. Surface Chem., 5th, Abo.
Fin/and.
Hansen, F. K., Bauman Ofstad, E., and Ugelstad, J. (1974). Paper presented at Symp. Theory
Practice Emul. Techno/., Brunel Univ. p. 1. Academic Press, New York.
Hansen, F. K., and Ugelstad, J. (1979). J. Po/ym. Sci. Po/ym. Chem. Ed. 17, 3069.
Higuchi, W. J., and Misra, J. (1962). J. Pharm. Sci. 51, 459.
Hildebrand, J. H. (1964). In "The Solubility of Non Electrolytes" Dover, New York.
Lange, S., Ugelstad, J., and Hansen, F. K. (1973). Abstr. Scand. Symp. Surface Chem., 5th, Abo,
Fin/and.
Mfutakamba, H., Mork, P. C., and Ugelstad, J. (1979). Unpublished results.

11. The Formation of Monomer Emulsionsand Polymer Dispersions


.

413

Miscra, S. c., Manson,J. A.,and Vanderhoff,J. W (1978). Preprints Am. Chem. SocoDiv. Organ.
Coatings Plast. Chem. 38(1), 213.
Napper, D. H., Netschey,
Netschey, A., Napper, D.
Olabisi, O., Robeson, L.
Academic Press, New

A, and Alexander, A. E. (1971). J. Polym. Sei. Part A-I 9, 81.


H., and Alexander, A E. (1969). J. Polym. Sci. Polym. Lett. 7, 829.
M., and Shaw, M. T. (1979). In "Polymer-Polymer
Miscibility."
York.

Schulman, J. A., and Cockbain, E. G. (1940). Trans. Faraday Soco 36, 651.
Shinoda, K. (1978). In "Principies of Solutions and Solubility." Decker, New York.
Smoluchowski,
M. V. (1918). Z. Phys. Chem. 92, 129.
Ugelstad, J., Merk, P. c., Dahl, P., and Rangnes, P. (1967). J. Polym. Sci. Part C 27,49.
Ugelstad, J., EI-Asser, M. S., and Vanderhoff, J. W. (1973a). J. Polym. Sei. Polym. Letl. Ed.
11, 505.
Ugelstad, J., Flegstad, H., Hansen, F. K., and Ellingsen, T. (1973b). J. Polym. Sei. Part C42, 473.
Ugelstad, J., Hansen, F. K., and Lange, S. (1974). Makromol. Chem. 175,507.
Ugelstad, J., and Hansen, F. K. (1976). Rubber Chem. Technol. 44(3), 536.
Ugelstad, J., Hansen, F. K., and Kaggerud, K. H. (1977). Faserforsch. Textiltech.-Z.
Polym.
Forsch. 28, 309.
Ugelstad, J. (1978). Makromol. Chem. 179, 815.
Ugelstad, J., ElIingsen, T., and Kaggerud, K. H. (1978a). Int. Conf Organ. Coatings Sci. Technol.
Preprillts, p. 425.
Ugelstad, J., Kaggerud, K. H., and Fitch, R. M. (1978b). Symp. Phys. Chem. Properties Colloidal
Particles. Miami.
Ugelstad, J., Kaggerud, K. H., Hansen, F. K., and Berge, A (1979a). Makromol. Chem. 180,737.
Ugelstad, J., Merk, P. C., and Kaggerud, K. H. (1979b). Paper presented at Int. Conf Surface
Co/loid Sci. 3rd. Stockholm Abstracts, p. 344.
Uge1stad, J., Merk, P. C., Kaggerud, K. H., ElIingsen, T., and Berge, A. (1980a). Adv. Co/loid
Interface Sci. 13, 101.
Ugelstad, J., Merk, P. C., Hansen, F. K., Kaggerud, K. H., and Ellingsen, T (1980b). J. Pure
Appl. Chem. (in press.)
Ugelstad, J., Merk, P. c., and Khan,

A. A. (1980c). Organ. Coatings Plast. Chem. Preprints

43,514.
Ugelstad, J., Ellingsen, T., and Kaggerud, K. H. (1980). Adv. Organ. Coatings 2, 1.
Ugelstad, J., Kaggerud, K. H., and Fitch, R. M. (1980). "Polymer Colloids" (R. M. Fitch, ed.),
Vol. n. Plenum Press, New York.
Vanderhoff, J. W., Vitkuske, J. F., Bradford, E. B., and Alfrey, T. (1956). J. Polym. Sei 20,225.
Vanderhoff, J. W., van den Hui, H. J., Tausk, R. J. M., and Overbeck, J. Th. G. (1970). In "Clean
Surfaces. Their Preparation
and Characterization
for Interfacial Studies" (G. Goldfinger,
ed.), Dekker, New York.
Vanderhoff,

J. W., EI-Asser,

M. S., and Hoffman,

J. D. (1978). U.S. Patent 4,070.323.

Vanderhoff,

J. W., EI-Asser,

M. S., and Ugelstad,

J. (1979). U.S. Patent 4,177.177.

Note Added in Proof


In equations describing rates of transport and equilibria between two phases (a and b) (see
page 386), it has been assumed that the interfacial tension between them and water is the same
for the two phases. In so me cases this assumption may not hold, for instance, if the two phases
have different compositions.
In these cases it would be more appropriate
to inelude separate
terms for the two phases, for example, Yaand ')lbfor the interfacial tensions.

12
Radiation- lnduced
Emulsion Polymerization
Vivian T. Stannett

415
1. Introduction .
418
11. LaboratoryResults with Different Monomers
418
A. Styrene
424
B. Acrylonitrile .
425
C. Methyl Methacrylate
428
D. VinylAcetate.
429
E. VinylChloride
431
F. Miscellaneous Monomers .
433
111. Copolymerizations .
433
A. Random Copolymers .
434
B. Block and Graft Copolymers .
IV. Radiation-Induced Emulsion Polymerization Using
436
Electron Accelerators
437
V. Pilot Plant and Related Studies
437
A. Introduction .
437
B. VinylAcetate.
440
C. Ettylene .
442
D. Tetrafluorethylene .
E. Tetrafluorethylene-Propylene and Other
444
Copolymers .
447
References

l. Introduction
The fitst.reported work on the radiation-inducedpolymerization
of a vinyl monomer in emulsion appears to be that of Ballantine (1954) at
Brookhaven National Laboratory in 1953. This work was confined to
styrene and utilized a tantalum y source with a dose rate of only 60,000
radjhr. Smooth conversions at high rates and high molecular weights to more
than 90% were obtained at 25C with about 2 Mrad. Soap concentration,
monomer to water ratio, and temperature were varied. These and other
results obtained with styrene, the classical monomer for such studies, wil\ be
described later. The motivation for the work was the considerable industrial
415
EMULSION POLYMERIZATION
Copyright ~ 1982 by Aeademie Pros,. Ine.
An right, or reproduetion in any form reserved.
ISBN 0-12-556420-1

416

Vivian T. Stannett

interest in emulsion polymerization and the high yield of free radical s from
the radiolysis of water. This latter feature in addition to the high kinetic
chain lengths associated with emulsion po!ymerization kinetics has led to a
continued interest in radiation as the initiator for a number of monomers.
There are a number of other practica! advantages associated with the use of
radiation-induced emulsion polymerization. These will be described in the
next few paragraphs.
Radiation, particularly using isotopes such as 60Co, can give an essentially unlimited range of radical fluxes from zero to those equivalent to
many moles per liter of chemical initiators, which would c1early be impracti. cal. Associated with this range is the ease with which the fluxes can be
monitored during the course of the polymerization reaction as the kinetics
and other considerations demando The fluxes can be programmed and built
into modern process control techniques leading in principIe to the orderly
control of molecular weight and partical size distributions and to the
elimination of residual monomer. Alternatively, there need be no change in
the radical flux during the reaction, Le., no initiator exhaustion, and the
initiation process can be started and stopped at will by simply dropping
away or raising the radiation source. This could be useful for changes in the
feed in copolymerizations and, in principIe, for block and graft polymerization in emulsion. Initiation with radiation is essentially temperature
independent. This leads to comparatively low temperature dependencies for
the overall reaction, the activation energy dropping from about 20 for
chemical initiation to only about 7 kcaljmol. This difference, especially
when coupled with the ease of removing the initiation source, makes the
possibility of exothermic, runaway reactions extremely low indeed.
Furthermore, initiation is not a function of polymerization temperature and
hence is completely uncoupled, in direct contrast to chemical initiation.
The lack of any activation energy with the initiation reaction has already
been emphasized. This also means that polymerizations can be conducted
at will at any temperature and at any initiation rateo In principIe, this can
be, and is indeed, accomplished with chemical initiation inc1uding redox
systems. In practice, however, especially low-temperature initiation is not
easy to achieve and control, particularly with polar monomers. The lowtemperature polymerization of vinyl acetate is of particular interest. It is
well known that monomers such as vinyl chloride and vinyl acetate, whose
degree of polymerization is mainly governed by chain transfer to monomer,
have a negative dependence of molecular weight on temperature. To
achieve high molecular weights, therefore, low-temperature polymerizations
are necessary.
The radicals produced by the radiolysis of water are hydrogen atoms
and, mainly, hydroxyl radicals. These are neutral and highly reactive

1-2. Radiation-Induced Emulsion Polymerization

417

radicals which lead to efficient initiation and there are no electrolytes such
as arise from most chemical systems; this leads to somewhat higher surface
tension latices with lower ionic strengths and higher pH values, often
eliminating the need for adding buffers as, for example, is the case with
potassium persulfate initiation. No contamination with residual initiator
frag~ents occurs. Another advantage that could conceivably become of
considerable importance is that radiation is ideal for initiating the polymerization of systems using cationic soaps. which often interact with chemical
systems. This feature was pointed out and investigated many years ago in
Japan by Inagaki et al. (1960a,b,c); such cationic lattices were found to be
quite stable. There are disadvantages to the use of radiation; for example,
there are no ionic end groups such as arise from persulfate initiation. These
could, in principie, lead to some stabilization of the resulting latex. In
addition, the radiation attacks all the components including the emulsifier
and the polymer as it is formed. The former could lead to a very small
amount of grafting but is probably a negligible factor. The latter, however,
leads to branching and a very small loss of acetate groups. It is doubtful,
however, under the conditions that would prevail in an industrial process
where low total doses would be used, that any of these problems could
cause difficulties. A more important problem, which could arise in flow
reactors or even kettle systems, is possible build-up of polymer on the walls
of the reaction vessel caused by diminishing flow rate approaching the
walls. This could also arise because the radiation is at its most intense near
the radiation elements. However, with suitable reactor designs this problem
could also be eliminated, as well as the general coating problems that can
arise with chemical initiation. In fact, with kettle reactors these effects have
not been observed, at least over the limited time periods involved in pilot
plant studies. With wide-tube flow reactors some coating of the walls has
been observed, but again it is probably not a major problem, at least over
reasonable time periods.
In general, for most investigations, 60Co }' radiation has been used
because of its high degree of penetration and the comparative ease of
estimating dose-depth characteristics and because radical fluxes comparable
to those used witr. chemical initiation can easily be achieved. There have
also been a few, comparatively brief, studies using electr.on accelerators to
initiate emulsion polymerization in emulsion. These have mainly been
conducted in Japanese laboratories.
From a fundamental point of view it is interesting to speculate on the
differences that could exist between the kinetics of emulsion polymerization
initiated by }' radiation and those of a conventional chemical initiator with,
for example, potassium persulfate. At dose rates giving a free radical flux
comparable to those achieved with chemical initiation any differences

418

Vivian T. Stannett

should be minoro The aqueous radicals produced by the radiolysis of the


water are essentially hydrogen atoms and hydroxyl radical s, both of which
are neutral, fast moving, and efficient compared with the sulfate ion radical s
produced by persulfate ions or various redox systems. In addition, however,
radiation can attack the surfactant, the monomer, and the polymer being
produced, causing some additional differences. The surfactant is usually
present in small amounts and should not cause too much perturbation of
normal kinetics. Radiolysis of the monomer and polymer could lead to
more significant changes, particularly with polar materials such as vinyl
acetate, vinyl chloride, and the acrylates and methacrylates. In these cases,
the radical yields can approach those of water itself, Le., 5-7 radicals per
100 eV of absorbed radiation. Unfortunately, this group of monomers has
extremely complex and large unresolved kinetics and it is difficult to make a
clear differentiation between radiation and chemical initiation. The polymers, as they are produced, however, should be attacked to varying
degrees by the radiation, leading to branching and even cross-linking and
other changes in the resulting polymers. Styrene and plystyrene have very
low radical yields, (~0.7) compared to water and should be relatively
immune to such differences. The dienes such as butadiene and isoprene have
low radical yields but could be subject to branching and cross-linking
because of the residual double bonds in the polymer. It is pro po sed to
discuss some of the various monomers that have been studied on a
laboratory scale with radiation. Since styrene should be the least subject to
special effects introduced by the use of radiation, it will be the first
monomer to be discussed.

11. Laboratory Results with DitIerent Monomers

A. Styrene
In his early work, Ballantine (1954) used 1% of an amine long-chain
alcohol sulfate (Duponol G) as the emulsifier and monomer to water ratios
of 1:7 and 1:9 at 25, 35, and 45C. The overall activation energy for the rate
was found to be 3.7 kcaljmol. The molecular weights decreased slightly with
conversion but were very high, close to or more than one million. The rates
were quite high reaching, for example, 54% per hr at 45C with molecular
weights up to 2 million. Abkin et al. (1959) presented the next published
work, followed closely by Inagaki et al. (1960a,b,c) and Bradford et al.
(1961). Hummel et al. (1962) conducted a long series of studies with styrene
and other monomers beginning in 1962. There was an interesting paper by
Acres and Dalton in 1963. The present author and his -colleagues began

12. Radiation-Induced Emulsion Polymerization

419

work on the radiation-induced polymerization of styrene and other monomers in 1964, with the first published report in 1967.
Styrene is one of the best-behaved of monomers in that it tends to follow
most closely classical Ewart and Smith (1948) kinetics. There are many
exceptions, however, even with potassium persulfate, the standard chemical
initiator. The best agreement was found with potassium persulfate as the
initiator and sodium lauryl sulfate as the emulsifier. Until recentIy (Garreau
et al., 1979), the various studies were mainly conducted with other systems
and for reasons other than to check whether there are, indeed, any real
differences between radiation and chemieal initiation in such an ideal
system.
Abkin et al. (1959) found the dependence of rate on the dose rate to be
close to 0.4, the Classical Smith-Ewart (1948) Case 2 value. The molecular
weights wc-reclose to those caIculated from the chain transfer to monomer
values. However, they found close to zero activation energy for the rates
and substantial post-irradiation polymerization. It was pointed out that the
radiation could also generate hydrogen peroxide as an additional initiation
source. One cationic and three anionic emulsifiers were studied, and the
cationic emulsifier gave a very low rateo Inagaki et al. (1960a,b,c) concentrated their studies on the elfect of different emulsifiers: anionic, nonionie, alld cationie. In all cases the rate increased with increasing emulsifier
concentration. The rate, in general, was greatest with anionic emulsifiers
and least with cationic. The rate and molecular weight dependencies (on the
soap concentrations) were in the range 0.14-0.40. Cationie lattices are not
easy to prepare with chemical initiators, and this earIy work pioneered the
use of y irradiation for their production. Stable lattiees with up to 50%
solids content were made with an overall activation energy of 6.2 kcaljmoI.
Bradford et al. (1961) conducted a detailed study using the dihexyl ester
of sodium sulfosuccinate as the emulsifier. Their main emphasis was on the
competitive growth of particles using seeded systems and was related to
their development of highly uniform particle size polystyrene lattices (e.g.,
Bradford and Vanderhoff, 1955). The paper also included, however, a rather
detailed kinetic study leading to values of the termination rate constant kl'
the average nUII).berof radieals per particle , and the activation energy
associated with die molecular weights produced (7.2 kcaljmol). Rather
similar results were obtained with 60Co y radiation as with potassium
persulfate initiation. However, the number of particles decreased with
temperature using radiation but increased with potassium persulfate. This
was attributed to the increase in the rate of initiation with temperature in
the case of persulfate initiation. In general, their values for the activation
energy, the efficiencies of initiation, and the rate constants for termination
werein good agreementwith those obtained later by Garreau et al. (1979).

420

Vivian T. Stannett

The work of Hummel et al. (1962,1967; Hummel, 1963) was fundamental


in character involving not only styrene but methyl methacrylate, methyl
acrylate, ethyl acrylate, vinyl acetate, and a number of other monomers.
Unique dilatometer systems were developed which enabled both continuous
and intermittent determination of rates, with sodium lauryl sulfate as the
emulsifier. The molecular weight was found to decrease with increasing
dose rateo In general, in the case of styrene, standard Smith-Ewart Case 2
kinetics approximated the experimental findings. The rate-conversion
curves showed an initial increase in rate followed by a c'onstant period and
an eventual decline. In some cases there was an increase in rate at high
conversions, presumably due to a small gel effect. These results are in
agreement with results obtained with potassium persulfate initiation and
the Case 2 Smith-Ewart picture.
Acres and Oalton (1963a), using the dioctyl ester of sodium sulfosuccinate as the emulsifier, found the intensity exponent of the rate to vary with
the monomer to water ratio from 0.22 to 0.34 but to reach the classical
Smith-Ewart value of 0.4 when extrapolated back' to zero monomer
concentration. They interpreted this result in terms of the special role of the
hydrogen atoms arising from the radiolysis of water.
Araki et al. (1967, 1969) conducted some rather straightforward polymerizations of styrene, and the properties of the resulting latices using 60Co y
radiation were determined in a simple dilatometer. Again smooth, rapid
polymerizations to high conversions were observed. A typical conversion
curve is shown in Fig. 1. The shape of the conversion curves agreed with
those observed by Hummel et al. (1962, 1967) and were similar to those
found with potassium persulfate under normal Case 2 Smith-Ewart conditions. The overall activation energy was 3.6 kcal/mol, in excellent agreement with the results of Ballantine (1954). The viscosity average molecular
weight decreased with decreasing temperature in agreement with Ballantine
and consistent with a constant rate of initiation and a reduced rate of
propagation with temperature lowering. Compara tive studies indicated that
the molecular weight distribution was somewhat sharper with radiation
than with potassium persulfate initiation. Radiation Mw/Mn values, for
example were about 3 as compared with 6 for persulfate. Some other
properties are presented in Table 1 for both radiation and chemical
initiation. The surface tensions are somewhat higher with radiation and the
pH values much higher, 8-9 compared with 3.8 with persulfate. Low pH
values are typical with initiation by persulfate ion decomposition and, as
discussed in the introduction to this chapter, indica te an advantage offered
by radiation. The particles sizes and molecular weights are similar at 60C.
As with molecular weight, particle size distribution is narrower with radiation, some typical differential distributions are shown in Fig. 2.

12.

Radiation-Induced

421

Emulsion Polymerization

100

75
c:
.o
Q)

>
c:
o

50

eQ)
(,)

...

,f

25

Fig. 1. Typical styrene conversion curve at 30C (dilatometer trace). Dose rate
0.02 Mradjhr, 25%solids, 6.7%sodium lauryl sulfateon total volume.

Recently, Garreau et al. (1979) reported a careful and rather detailed


study of the kinetics of the radiation-induced polymerization of styrene in
emulsion with sodium lauryl sulfate under conditions found earlier to lead
to c10se agreement with simple Case 2 Smith-Ewart kinetics (Smith, 1948).
Most of the normal reaction variables were studied, and the rates of
polymerization were found to be independent of the monomer-to-water
TABLE I
Chemical and Radiation Polymerization

of Styrene in Emulsion
Preparation
y-Radiation

Property
pH
Surface tension (dyne/cm)
Particle size (Jm)
Molecular weight (viscosity)

K2S20S
at 600C.
3.8
61.1
0.13
2,700,000

9.2
68.9
0.10
1,664,000

7.9
69.0
0.07
413,000

.25% solids and 6.7% sodium lauryl sulfate based on water content

(Araki el al., 1969).

422

Vivian T. Stannett
B

24....
22
20
18L

16
14
e

:o

12

10
o
C 8
6,

:1

250

-Diameter

2000

(,.\.)

Fig. 2
Differential particle size distribution of polystyrene lattices initiated by: A,
radiation OC; B, radiation 600C; C, potassium persulfate 60C (Araki et al., 1967).

ratio and to the 0.4 power of the dose rate (Fig. 3). The activation energy
(Fig.4) associated both with the rate and the molecular weight was
7.9 :t 0.6, in good agreement with Van der Hoff et al. (1958) and with the
literature values for Ekp' This is reasonable since little or no activation
energy is involved in the initiation and termination steps. The dependence
of the rate and the molecular weight on the emulsifier concentration was
about 0.7, in reasonable agreement with the Smith-Ewart value of 0.6. The
number of particles, however, was of the correct order of 0.6 with respect to
the emulsifier concentration. The termination rate constants were caIculated
using the method of Van der Hoff (1958). The values found were in excellent
agreement with those of Van der Hoff for persulfate initiation and Bradford
et al. (1961) for y initiation. The propagation rate constant kp was caIculated from Case 2 Smith-Ewart kinetics and found to vary between 32 and

68 M- 1 sec- 1, well within the range of the reported literature values at


30C. The caIculated efficiencies of initiation were low, about 40%, but
similar to those discussed by Bradford et al. with y irradiation and
determined by Van der Hoff (1958) with persulfate initiation. No seeded
polymerizations were conducted but the above results indicate clearly that the

423

12. Radiation-Induced Emulsion Polymerization


~
.s::.

90
80
70
60

<D 50
& 40
30

.~
<D

20

E
>-

o
a. 10

.2

.3

.4 .5.67.8.9

4 5 6 789

Dose Rate (mradjhr x10)


Fig. 3.
Dependence of the rate of polymerization of styrene on the dose rate at 300C
and 2.0% sodium lauryl sulfate concentration (Garreau et al., 1979; reproduced with permission of Journal of Colloid amI Interface Science.)

radiation-induced polymerization of styrene under the correct conditions


cIosely approximate the kinetics found with potassium persulfate.
Karpov et al. (1968) reported a number of experiments concerning the
post polymerization of styrene in emulsion using y irradiation. The polymerization' continued long after the removal of the source, reaching cIose to
100% conversion and very high viscosity average molecular weights (1528 x 106) in about 20 hr at 22C. The conversion curves were similar to
those normally found with continuous irradiation, Le., linear until the
separa te monomer phase disappeared and then tapering off. The molecular
weight continued to increase during post-polymerization. It was demonstrated experimentally that very few radical s disappeared during the
process. The polymer had about 85% of narrow molecular weight distribution. The half-life times of the propagating radical s were determined in a
large number ofmonomer systems by Hummel et al. (1969). Styrene had the
highest value, up to 80 min in some cases. Karpov et al. (1969) determined
the kinetic parameters of the radiation polymerization of styrene with
potassium laurate as the emulsifier. The rates were found to be of approximate orders of 0.5~with respect to dose rate and emulsifier concentration.
The number of particles was proportional to the dose rate to the 0.4-0.6
power and the emulsifier to the 0.6 power. The activation energies were 4.6
and 3.9 kcaljmol with respect to the rates and number of particles, respectively. Cetyl trimethyl ammonium chloride gave slower rates and an
activation energy of 7.7 kcaljmol for the rateo The same group (Karpov et
al., 1973a) found the rate of polymerization to be independent of the
number of particles up to 1016particlesjper milliliter. This is contrary to the
resuIts found with chemical initiation and is attributed to chain transfer to

424

VivanT. Stannett

':"

le

f" 20 'ol-e

'-..)

G
..
I

t.

lO

lO
9
8
7
6

o 9
-E 8
7
Z 6
O
e

.
...
N

Z
O
e
N
...

!
...
O

A.
...
O

...
...
111:

o
...
Q

...

O
A. 3
...
O
...
e

111:

3.1

3.2

3.3

3.4

3.5

3.6

Fig. 4. Temperature dependenceof the rates and degreesof polymerization for the
radiation induced polymerization of styrene. Dose rate 0.093MradJhr (Garreau et al., 1979;
reproducedwith permissionof Journal of Col/oid and Interface Science.)

the initiator. There was considerable activity by this group of Soviet


workers and additional references may be found in the articIes quoted in
this chapter.

B. Acrylonitrile
The emulsion polymerization of acrylonitrile is quite complex because of
the high solubility of the monomer in water (7.4% at 25C) and the
insolubility of the polymer in the monomer. Hummel et al. (1967), using a
sophisticated recording dilatometer, showed that the rate-conversion
curves

12. Radiation-Induced Emulsion Polymerization

425

were very complex, with an initial rise followed by two periods of almost
constant, but different rates, and finally a first-order reaction rate at high
conversions. Partic1e formation in the aqueous phase was assumed. A few
studies reported by Glazkova et al. (1971) were conducted below the critical
micelle concentration of the emulsifier used. Karpov et al. (1974) reported a
rather thorough study using three different emulsifiers: potassium laurate,
sodium alkyl sulfonate, and a cationic emulsifier, cetyl pyridinium chloride.
Contrary to the general experience with styrene, vinyl acetate, and methyl
methacrylate, the rates were markedly higher with the cationic emulsifier.
True latices were formed when the monomer concentration was below the
solubility limit, Le., no monomer droplets were present. With potassium
laurate, however, only coagulum was formed in every case. The number of
partic1es formed in the latex systems were found to be independent of the
dose rate but proportional to the emulsifier concentration to about the 0.6
power. With high concentrations (0.5%) of cationic emulsifier the number of
partic1es increased linearly with conversion, but at lower concentrations and
always with the sulfonate the number remained constant. The molecular
weight did not vary more than about twofold over a wide range of
emulsifier, with monomer concentration and dose rate being of the order of
106 viscosity average. There is much useful practical data in the paper but
the results are confounded by the presence of heavy concentrations of
coagulum.
O'Neill and Stannett (1974) also found, with sodium lauryl sulfate as the
emulsifier, that when a separate monomer phase was present a milky slurry
that was 100% filterable was formed even ~t low conversions. However, in
the water-soluble regio n low solid latices were obtained. The initial rates
were much higher than in the absence of emulsifiers as was previously
reported by Karpov et al. A plot of the initial rates versus the monomer
concentration is shown in Fig. 5. A first-order dependence on monomer
concentration was found. Again, as with Karpov et al., considerable useful
data is presented for systems where a separate monomer phase is present.
Since these are not truly emulsion polymerizations, further details will not
be discussed in this chapter. The molecular weights were essentially independent of the emulsifier concentration and in the same range, about 106,as
found by Karpov et al. Izumi et al. (1967; Izumi, 1967) obtained similar
results with ammonium persulfate as the initiator.
C.

Methyl Methacrylate

Methyl methacrylate is a monomer that behaved "nonideally" with


conventional initiators (Zimmt, 1959). This is due mainly to the larger size
partic1es which lead to a larger average number of radical s per partic1e than

426

Vivian T. Stannett
400

200

e
.2100
'"
Q;

O
u

80

60

'<
IX
40
.....

20

10
0.1

/8
0.2

0.4

0.6 0.8 1.0

2.0

[M]c!ml-')
Fig. 5.
Dependenee of the rate of the radiation-indueed polymerization of aerylonitrile
in aqueous solution, linear portions of rate eurves, on the monomer eoneentration at 25C 8
in the absenee of sodium lauryl sulfate, O with 0.21% (O'Neill and Stannett, 1974: reprodueed
with permission of Journa/ of Macromo/ecu/ar Science-Chemistry AS, p. 949, by eourtesy of
Mareel Dekker, Ine.)

the ideal value (Smith-Ewart Case 2) of one-half and to the prevalence of


the gel effect at comparatively low conversions. The first published report of
the radiation polymerization of this monomer was by Abkin et al. (1959).
Very high rates, up to 6% per minute at room temperature were observed
with high molecular weights. Inagaki et al. (1960a,b,c) polymerized with 'Y
radiation methyl methacrylate and anionic, nonionic, and cationic emulsifiers, and as with styrene the rates were highest with anionic and lowest
with cationic emulsifiers. The molecular weights were similar with anionic
and nonionic and somewhat lower with cationic emulsifiers. Both the rate
and the molecular weight increased with increasing emulsifier concentration
in all cases. The order of reaction for the rate and the molecular weight was

12.

Radiation-Induced

Emulsion Polymerization

427

found to be 0.3, 0.4, and 0.2 for anionic, nonionic and cationic, respectively.
The dose rate dependence of the rate was 0.3 with the cationic emulsifier,
stearyl trimethyl ammonium chloride, and the activation energy was
6.7 kcal/mol. Allen et al. (1962) conducted a few experiments in connection
with their work on block and graft polymer synthesis in emulsion. An
estimate of 92 M-1 sec-1 at 25C for the rate constant for propagation was
obtained, in reasonable agreement with the published solution and bulk
values. As with Abkin et al. the possible contribution to the initiation step
from peroxides arising from the irradiation was stressed and some supporting evidence presented.
A more detailed fundamental study was reported by Hummel et al.
(1962; Hummel, 1963). Using a sensitive recording dilatometer of their own
design they could follow the variation of the rate of polymerization with
time and conversion. There was an abrupt rise in the rate for the first few
percent conversion attributed to particle formation. There followed a
constant rate period up to about 30% when the free monomer phase
disappeared. An increase in rate to a maximum was then observed, ascribed
to the gel effect, followed by a steady decrease in rate due to monomer
depletion and slower diffusion of the monomer to the active sites in the
highly viscous particles.

The particle sizes were rather small, about 500

diameter, compared with those often obtained with chemical initiation (e.g.,
Zimmt, 1959). This presumably explains the comparatively close adherence
to the simple Smith-Ewart picture up to the appearance of the gel effect.
Acres and Dalton (1963a) also studied the emulsion polymerization of
methyl methacrylate initiated by 60Co y raqiation using a recording dilatometer. Only the conversion-time curves were measured with constant dose
rate, varying monomer concentration, and with constant monomer concentration at different dose rates. Except at the lowest monomer concentration a clear gel effect was observed, with linear rates up to that point.
The linear rates increased with increasing monomer concentration up to
.

about 0.4 mol/liter and then leveledoff. The dependenceof the rate, before
the gel effect, on the dose rate was 0.4 and, unlike their findings with
styrene, not dependent on the monomer concentration. Their results w~re
consistent with those of Hummel et al. that methyl methacrylate follows,
with y radiation, the generally accepted Smith-Ewart Case 2 kinetics
except for the marked gel effect.
Hoign and O'Neil (1972) studied several features of the y radiationinitiated polymerization in emulsion. Sodium lauryl sulfate, dioctyl and
dibutyl sodium sulfosuccinates, and two nonionic, polyoxyethylene-type
emulsifiers were used. Sodium lauryl sulfate gave, by far, the highest rates
and most stable lattices and was used for all of the results reported. The
rates were found to be 0.43 order on the dose rate and 0.53 on the

428

VivianT. Stannett

emulsifier, above the critical micelle concentration. The activation energy


was 4.8 kcal/mol. Molecular weights of about one million were obtained at
22C, but no systematic study of this property was made.

D.

Vinyl Acetate

Vinyl acetate, in marked contrast to styrene, has a most complex


mechanism (Klein et al., 1975). This monomer was among the earliest
investigated with y-radiation initiation in emulsion (Allen.et al., 1956, 1960,
1962; and Inagaki et al., 1960a,b,c). Rapid rates to high conversions were
obtained, and the curves were linear from about 20 to 70% conversion,
similar to chemically initiated systems. Molecular weights as high as one
million were obtained. The rates, although high, were far less than those
calculated with conventional Smith-Ewart theory and the accepted values
for the propagation rate constant. This was correctly ascribed to the
possibility that the number of radicals per particle averaged less than 0.5
and that perhaps the Smith-Ewart Case 1 would be a 'better description. As
with styrene and methyl methacrylate, Inagaki et al. (1960a,b,c) studied
anionic, nonionic, and cationic emulsifiers. The relative rates, as with the
other two; were found to be anionic > nonionic > cationic. The principIe
additional findings were that the activation energy for the rates of polymerization was 5.7 kcal/mol, the rates increased with increasing dose rate but
with an unspecified dependency. The dependence of the rate and molecular
weight on the emulsifier concentrations was found to be between 0.14 and
0.40 order. Stable cationic lattices up to 50% by weight were obtained. A
more thorough investigation of the stability of radiation-initiated poly(vinyl
acetate) lattices has also been reported by Ohdan et al. (1970; Ohdan and
Okamura, 1969) and by Kamiyama et al. (1970).
Araki et al. (1967, 1969) carried out a more systematic study of the
kinetics and other features of the y-initiated emulsion polymerization of
vinyl acetate using sodium lauryl sulfate as the emulsifier. This system had
been thoroughly investigated with potassium persulfate as the initiator (Litt
et al., 1960, 1970). Some post effects have been observed with vinyl acetate,
particularly above 50% conversion (Friis, 1973; Sunardi, 1979). These effects
had been used by Allen et al. (1960, 1962)for the possible synthesis of block
and graft polymers and will be described later in this chapter. The half-life
of the radicals in a vinyl acetate latex polymerization was determined by
Humme! et al. (1969) as 0.8 min at 53.8% conversion. Araki et al. (1967,
1969) determined all the normal rate dependencies and included some
seeded latex studies. Their results and those of other investigators are
summarized in Table II together with those found with potassium persulfate
initiation and those predicted by the Smith-Ewart Case 2 theory. The

429

12. Radiation-Induced Emulsion Polymerization


TABLE IJ

Orders ot Reaction tor the Rates with Potassium Persultate


and Radiation Polymerizationot VinylAcetatea
Order
Variable
Initiator

or dose rate

Ditto with seeded system


Number of partic1es
Monomer-water
ratio
Emulsifier (sodium lauryl sulfate)
a

0.5- 1.0
0.8
0.2
0.35
0.1-0.3

Radiation

Smith-Ewart 2

0.5-0.9
0.26
0.7
0.2
0.3-1.0

0.4
O
1.0
O
0.6

Data taken from various sources.

radiation results tend to approach the Smith-Ewart theory more dosely


than the persulfate-initiated polymerizations. As indicated by the ranges
reported in Table n, there are considerable variations in the literature
results with both radiation and chemical initiation. In particular, there is a
wide range of dependence of the rate on the dose rate, the initiator
concentration, and the emulsifier concentration.
The molecular weight, in general, was found to be insensitive to dose rate
and emulsifier concentration, consistent with the determining factor being
chain transfer to monomer. The radiation-produced poly(vinyl acetate) had
a somewhat higher molecular weight even at the same temperature, perhaps
due to branching initiated by direct radiolysis of the polymer chains.

E.

Vinyl Chloride

Wang (1962) was the first to report studies on the y-radiation initiated
polymerization of vinyl chloride in emulsion at room temperature. Rapid
rates to high conversions were obtained after rather long induction periods
of 1 to 3 hr. The degrees of polymerization were constant within experimental error at ab0l!t 2000, in keeping with termination being dominated by
chain transfer to monomer. Little or no dependence .of the rate on the
emulsifier concentration or the monomer concentration was observed.
However, the rates were proportional to the 1.22 power of the dose rateo
Some preliminary results were presented by Barriac et al. (1969) and
followed by a more detailed report (Barriac et al., 1976) wherein two
emulsifiers were used. Again at 25C, for example, the degree of polymerization was essentially independent of the dose rate or the emulsifier
concentration and similar to that obtained by Wang. A strong post effect

430

VivianT. Stannett

was observed, and even at low conversions these effects were eliminated
experimentally by freezing in liquid nitrogen immediately after irradiation.
The conversion curves were linear up to about 80% conversion with a small
gel effect apparent at the lowest emulsifier, Le., largest size particles,
concentration. The rates of polymerization and the number of particles
were proportional to the 0.4 power of the dose rateo The rate was
dependent on the 0.3 power, and the number of particles on the 1.0-1.3
power of the emuls.ifier concentration and independent of the monomer to
water ratio. The activation energy for the rate for both emulsifiers was
4.0 kcal/mol and for the number of particles the unusualIy high value of
17.0 kcal/mol. A detailed kinetic analysis of these and the other results was
made. The mechanism and kinetics were in accord with those derived by
Gardinovacki et al. (1971) and Hansen and Ugelstad (1976), modified
somewhat due to the very smalI particles obtained with y initiation. The
smalI size and correspondingly larger number of particles is often found to
be a feature of radiation initiation. It has been explained by the lower ionic
strength of the aqueous phase leading to. the stabilization of more and
smalIer particles in addition to the faster diffusing and neutral hydroxyl
radicals and hydrogen atoms compared with the negatively charged, larger
sulfate-type radicals.
Karpov et al. (1977) studied the polymerization in the presence of a
number of anionic emulsifiers below the critical micelIe concentration. High
molecular weight polymers with low concentrations of impurities were
obtained at high rates. The overalI rate was found to be about 0.5 order
with resp~t to the dose rate with an activation energy of 5 kcal/mol, in
reasonableagreementwith those reported by Barriac et al. (1976).
A number of interesting and unusual phenomena were reported by the
same group of workers. With cationic emulsifiers suspensions rather than
latices were obtained although similar systems gave stable latices with
chemical or UV initiation (Karpov et al., 1971). It was hypothesized that
the radiation conferred some negative charges on the particles, reducing
their stability. Karpov et al. (1972)reported that higher rates were obtained
with cationic emulsifiers below their critical micelIe concentration but that
the reverse was true with hydrogen peroxide. With anionic sulfonate
emulsifiers, Karpov et al. (1973b) found that with monomer concentrations
below 35% stable lattices were obtained but that above 35% free flow
suspensions were obtained. Latex inversion was suggested as the explanation. Kinetics similar to those found with anionic emulsifiers was reported by Karpov et al. (1973c) when cationic emulsifiers below the critical
micelIe concentration were used.

12. Radiation-Induced Emulsion Polymerization


F.

431

Miscellaneous Monomers

1. Vinylidene Chloride
Panajkar and Rao (1979) have reported a rather extensive study of the y
radiation-initiated polymerization of vinylidene chloride in emulsion. With
sodium lauryl sulfate as the emulsifier smooth polymerization-time curves
at high rates were obtained, up to more than 90% conversion. Between 45
and 60% conversion, the linear region, the rate was 0.3 order with respect to
the emulsifier concentration. The molecular weights were found to increase
with conversion and values up to 79,000 were obtained. Some reasons for
the departure from Smith-Ewart behavior were suggested. Earlier, Hummel
et al. (1967) had presented some interesting data on a closely related system,
a similar rate-time behavior was observed and a tentative explanation proposed. Both discussions were based on the insolubility of the polymer in its
own monomer.
2.

Butadiene

This monomer has been investigated by Ishigure et al. (1974) using


sodium lauryl sulfate as the emulsifier. The polymerizations were found to
proceed slowly in the initial stages with rates increasing gradually with
conversion. Diffusion of the monomer appeared to be an important factor,
higher agitation rates giving higher rates. Four different anionic emulsifiers
gave essentially the same rate at the same molar concentrations. Since the
conversion plots were curving for most of the polymerization, it was difficult
to get a clear picture of the order of the raction. The rates were very low
with maximum G (Q1onomer)values of less than 4000 at 40C. The addition
of n-dodecyl mercaptan caused increases in the rates but only by 20-30%.
An interesting

feature was the small particle sizes, ranging from 122-537

in diameter, obtained with radiation. This was partly due, however, to the
rate of radical production at the dose rates used since styrene under similar
conditions had only about double the diameters.
3.

Methyl Acrylate

Methyl acrylate behaves differently from the other acrylates because of


its comparatively high solubility in water, 5.2% at 30C. This monomer has
been extensively studied by Hummel et al. (1962,1967; Hummel, 1963). The
rate first rises steeply and then drops off as the conversion proceeds, with
two roughly linear slopes. Rapid polymerization to stable latices was
obtained and the molecular weights were high, greater than one million.

432

Vivian T. Stannett

"Cross-linking" occurred in some experiments; this could be partially due


to entanglements. There was little temperature effect on the rate, molecular
weight, rate-time slope, or number of particles. The dose rate dependences
of the linear portions and the maximum rates were about 0.55. It was
suggested that a strong gel effect was present and that prolonged particle
formation in the aqueous phase was occurring. The work of Hummel et al.
appears to be the most detailed study of the radiation-induced polymerization of methyl acrylate and the original references should be consulted for
further details. Ohdan and Okamura (1969) reported some further details
including the stability and other properties of the resulting latices.
4.

Ethyl Acrylate

This monomer also appears to have been studied only by Hummel et al.
(1962, 1967; Hummel, 1963). The rate-time curve shows an initial sharp rise
followed by a slow decline, with no zero-order portion.. The curve can be
adequately explained by the assumption of a strong gel effect and the high
solubility of the polymer in the monomer; presumably the separate monomer phase disappears early in the reaction whereas the creation of new
particles continues longer than, for example, in the case of styrene because
of the comparatively high solubility of the monomer in water, 1.8 compared
with 0.012% with styrene.
5. n-Butyl Acrylate
Hummel et al. (1967) reported a brief study of this monomer with
sodium lauryl sulfate as the emulsifier. Behavior similar to ethyl acrylate
was observed. The rate increased rapidly during the first 10% conversion,
presumably because of particle formation. There followed a long first-order
decline in the rateo This behavior, like ethyl acrylate, was ascribed to the
high degree of swelling of the polymer by the monomer leading to the early
(-15% conversion) disappearance of the separate monomer phase. A more
detailed study with the same emulsifier has been described by Hoign and
O'Neill (1972). The overall activation energy of the rate was very low, only
0.5 kcaljmol; that of the molecular weight, however, was 3.9 kcaljmol. The
dependence of the rate was about 0.7 order with respect to dose rate and
0.48 with respect to the emulsifier concentration. The molecular weight was
essentially independent of the emulsifier concentration and the dose rateo
The values were high, 2.6-6.2 million, viscosity average, but they could be
systematically reduced by the addition of n-butyl mercaptan. The rate
dependencies are similar to those found with potassium persulfate
initiation.

12. Radiation-Induced Emulsion Polymerization

433

6. Other Monomers
Hummel et al. (1967) also conducted brief studies of a number of other
monomers. Chloroprene behaved similarly to styrene. Methacrylonitrile
behaved somewhat like methyl methacrylate with a detinite gel effect. Butyl
methacrylate behaved somewhat like styrene but with two small maxima in
the rate-conversion curves; the reasons for this are unknown but the
second small peak could arise from the gel effect. Decyl methacrylate showed
only one maximum rate at about 50% conversion. Again the reasons for
this behavior are unclear. Isoprene did not polymerize in emulsion at either
low or high dose rates. Kalyadin et al. (1975) have presented a study of the
radiation-induced polymerization of vinyl fluoride in emulsion.

111. Copolymerizations

There have been a few reports on the use of .radiation to initiate


copolymerization in emulsion. Earlier than the conventional random copolymerization work there were a number of studies on the formation of block
and graft copolymers in latex formo

A.o Random Copo/ymers


Butadiene has been successfully copolyperized with styrene (Ishigure et
al., 1974) and with acrylonitrile (Ishigure and Stannett, 1974). In addition,
vinyl acetate has been successfully copolymerized in a few experiments with
dibutyl maleate and with methyl, n-butyl, and 2-ethylhexyl acrylates (Araki
et al., 1969); all the copolymerizations proceeded smoothly with good
yields. The butadiene-styrene system polymerized slowly but with the rate
increasing with increased styrene content. They were considerably slower,
however, than comparable polymerizations using potassium persulfate as
the initiator, whereas styrene behaved normally. G(-M) values ranged from
only 1,000 for 21.5 mol

% styrene

to 3,800 with 60% styrene, compared

with

87,000 for pure styrene under similar conditions. The reasons for the low
rates, as with butadiene itself, are not clear. Acrylonitrile, however, as the
comonomer with butadiene behaved normally with easy copolymerizations
at rapid rates. G(-M) values reached 100,000 with 90% acrylonitrile. In the
range used for nitrile rubbers G values were still around 25,000, a quite
practical yield from the industrial viewpoint. The copolymer reactivity
ratios were normal. Strong post effects were observed showing a clear tirstorder termination process. Kamiyama and Okamura (1969) copolymerized

VivianT. Stannett

434

styrene and acrylonitrile in emulsion using y radiation.


two nonionic emulsifiers were used. A maximum in the
close to a 50: 50 molar proportion of the two monomers
nonionic emulsifiers, in contrast to bulk and chemically
polymerizations. Explanations based on the mechanism
presented.

B.

One anionic and


rate was observed
in the case of the
initiated emulsion
of initiation were

Block and Graft Copolymers

Block and graft copolymer formation in latices using radiation as the


means of initiation have been studied since early times. The basic idea, first
apparentIy suggested by Allen et al.{1956, 1958),is to use the trapped radicals in
the latex particles either during the polymerization ofthe first monomer or by
irradiation of a completed latex and adding an additional different monomer.
The former procedure should lead to block polymers an~ the latter to either
blocks or grafts. Hummel et al. (1969) have measured the half-lives of the
propagating radicals using intermittent 60Co irradiation. Such values depend
to various extents on the conversion and were found to vary from 20 min for
styrene at 28.8% conversion to 0.8 min for vinyl acetate at 53.8% conversion.
The early work of Allen et al. involved adding methyl methacrylate to
poly(vinyl acetate) at 70% conversion. It was clearly shown that block
formation did indeed occur. More detailed reports of similar systems were
published in subsequent papers (Allenet al., 1960,1962).It was shown that graff
copolymers were also formed by transfer reactions. Styrene radicals which have
a low transfer activity did not yield graft polymers, showing the great
importance of such reactions in the overall reaction. Mukoyama and Toriaki
(1962) successfully grafted vinyl acetate to polypropylene latices by irradiation, followed by the addition of monomer. The effect of monomer
concentration, monomer purity, and type of emulsifier were studied. Ito and
Watanabe grafted methyl methacrylate to polyethylene latices and a number
of details of the reaction were ellucidated in their 1965 publication.
Acres and Dalton (1963b) studied the graft andjor block copolymerization of methyl methacrylate onto irradiated polystyrene latex and styrene
onto irradiated poly{methyl methacrylate) latex. A number of experimental
variables and conditions that give good yields were investigated. It was
difficult to separa te the copolymers from the homopolymers but it was clear
that good yields of nonrandom copolymers were indeed obtained.
The direct radiation grafting of vinyl monomers to natural rubber latex
was studied by two groups of workers. Cockbain et al. (l958, 1959) grafted
methyl methacrylate using both y radiation and a chemical redox system.

12. Radiation-Induced Emulsion Polymerization

435

The grafting efficiencies and molecular weights of the side chains were both
much higher with radiation initiation. An interesting observation was that
the y-grafted latices formed strong continuous films on drying whereas the
redox grafts gave cracked or crazed films. This difference was attributed to
the chemical grafts being located mainly on or near the rubber particIe
surfaces, producing a heterogeneous polymer system. Evidence to support
this idea was provided by adding oil- and water-soluble retarders in the
radiation grafting experiments. The oil-soluble ones also gave cracked films
since they presumably reduced the amount of homogeneous grafting
whereas excellent films were obtained in the presence of water-soluble retarders. Cooper et al. (1959a-c; Cooper and Vaughan, 1959) also carried
out extensive studies of the direct radiation grafting of methyl methacrylate.
A general kinetic scheme was developed for the grafting system and the
activation energy determined as 3.1 kcal/mol for the rate, whereas the
molecular weight of the grafted side chains was little affected by temperature. A number of other interesting and useful observations were reported
in these papers.
More recentIy, Ishigure et al. (1973) and Garreau et al. (1980) reported
rather detailed investigations of the direct radiation grafting of styrene to
polybutadiene latices. The motivation for the work was for the possible
synthesis of such latices for use in the production of high-impact polystyrene. Both papers described direct comparisons between radiation- and
potassium persulfate-initiated systems. The earlier paper used gel-free smallparticle-size latices whereas the second used a larger particle-size crosslinked latex of the type used commercial~y for high-impact polystyrene
production. The former types of polybutadiene latex yielded propagation
rate constants and activation energies in good agreement with the literature
values for styrene. The grafting efficiencies decreased with conversion but
were similar in both radiation and chemical initiation. A simple model was
presented of the basis for the grafting efficiencies, and it gave a cIearer
picture of the various reactions involved.
The second paper (Garreau et al., 1980) with the larger cross-linked
polybutadiene latex gave a much more complicated pattern of behavior
with significant departures from the simple Smith-Ewart picture. It was
shown that the extracted homopolystyrene had molecular weights similar to
the grafted side chains after destruction of the polybutadiene backbone
polymer. Radiation was found to lead to more grafted side chains of lower
molecular weight than was observed for persulfate initiation. This was
explained by the higher rate of initiation in the former case and led to
somewhat lower grafting efficiencies. Nevertheless, under proper conditions
up to 80% efficiency was obtainable. The conversions also produced

436

Vivan T. Stannett

economically favorable yields. These results, when coupled with the ease of
control and other radiation features described in the introduction, make
radiation a viable and effective alterna tive to chemical initiation for grafting
processes on an industrial scale.

IV. Radiation-InducedEmulsionPolymerizationUsing ElectronAccelerators


There have been a number of brief reports, all Japanese, of the use of
electron accelerators to affect radiation initiation in emulsions. Vinyl acetate has been studied in most detail although brief reports on other
monomer polymerization, inc1uding styrene in emulsion, have been published. Kamiyama (1974) used pulsed electrons (6 MeV) from a linear and
continuous 3-MeV electro n Van de Graaff accelerator for the emulsion
polymerization of vinyl acetate using sodium lauryl sulfate as the emulsifier.
Close to 90% conversions were obtained in a few secopds with degrees of
polymerization of about 2500, almost independent of the dose rateo The
latter observation could be due to termination by chain transfer to monomero The resulting latices had a semitransparent appearance suggesting
very small partic1e sizes. The rates were proportional to about the 0.5
power of the dose rateo The continuous irradiations were several times
slower than the pulsed experiments with a weak dependence on the pulse
frequency.
Kamiyama and Saito (1975) extended these studies to methyl and ethyl
acrylates and styrene. The conversions leveled off at about 90 %. The latices
were less turbid than those obtained by y radiation. Again, those obtained
at the highest dose rates were almost transparent, with partic1e sizes as low
as 23 nm in diameter. In the case of styrene, the rates were almost
independent of the dose rate but were found to be to the 0.3, 0.5, and 0.7
powers for methyl and ethyl acrylates and vinyl acetate, respectively. At
dose rates higher than 2000 rad/sec the dependency dropped to 0.2. The
acrylates were so reactive that limiting conversions were reacted in a few
seconds.
Kamiyama and Shimizu (1975) also studied vinyl propionate and vinyl nbutyrate. The dose rate dependencies were 0.55 and 0.46, respectively, even
at the maximum dose rates. The rate-determining process at the high dose
rates associated with electron accelerators was suggested as being the rate
of diffusion of the monomer from the droplets to the aqueous phase in the
case of styrene, whereas with vinyl acetate the competition between polymer
nuc1eation and radical recombination of radicals in the aqueous phase
could be important. A further brief study of the vinyl acetate system was
presented by Kamiyama (1975).

12. Radiation-Induced Emulsion Polymerization

437

Finally, a very recent study (Hayashi and Okamura, 1980) was conducted in a flow system with electro n beam irradiation. Styrene gave very
low rates with broad distribution molecular weights (mn) averaging about
1000. Vinyl acetate, on the other hand, polymerized at much higher rates
reaching 60% conversion in 200 sec, in some cases at a dose rate of
0.1 Mradjsec at 40C. Trimodal molecular weight distributions were obtained at peak s of about 400, 40,000, and 400,000. The highest peak may
have been cross-linked or highly branched. Further studies of emulsion
polymerization at high dose rates could be quite rewarding and possibly of
industrial interest in certain cases.
V. Pilot Plant and Related Studies

A. lntroduction
The previous sections discussed the results of a large number of laboratory experiments concerning radiation-induced emulsion polymerization.
In this section the results obtained from a number of investigations using
small-scale pilot plant type equipment will be described. A full discussion of
these, more engineering oriented experiments, is beyond the scope of this
chapter. It was considered important, however, to summarize these studies
and to inelude those references which present the full details of the work.
Engineering flow systems and small scale autoelaves, designed largely for
batch experiments were developed for these investigations. Experiments of
this kind are necessary preludes to the eventuallarge-scale use of radiation
for industriallatex production.
.

B.

Vinyl Acetate

The basic laboratory studies described earlier were with batch systems;
in addition, one larger size and two micro pilot plants have been constructed and operated. AII were based on the general principIe of recirculation from a stirred vessel away from the radiation source. The
emulsions were ~ontinuously circulated through tubing to an in-so urce
"plug-flow" reactor and back to the kettle. The two micro pilot plants were
largely based on the originally designed and operated pilot plant described
by Stahel and Stannett (1969, 1971). The larger pilot plant, constructed by
Neutron Products, Inc., was designed and constructed simultaneously but
quite independently of the micro plants (Allen et al., 1969). The radiation
emulsion pilot plants described by Stahel et al. were investigated for
styrene, vinyl chloride-vinyl acetate copolymers (Stahel et al., 1979), and for
grafting to latex (Memetea et al., 1977).

438

VivianT. Stannett

The first micro pilot plant for vinyl acetate to be described was that of
Hoign el al. (1972), which was essentially an all glass apparatus. The
semicontinuous equipment consisted of a stirred, constant temperature
3000-ml glass reservo ir from which the emulsion was pumped through a
double helix 240 mI-reactor situated within the radiation cavity. On leaving
the reactor the emulsion flowed through a cooler and back into the
reservoir. The reservoir has inlets for monomer, soap solution, and the pure
nitrogen with which the apparatus was purged prior to and during the run.
The conversion curves were found to be linear between 1,5 and 60%. No
post effects were observed although this could have been caused by traces of
oxygen in the system beca use of the diffusion of air through the thin-walled
polyethylene transfer lines. Increasing the flow rate decreased the rate of
polymerization, and at very rapid rates of flow the polymerization ceased
entirely. This effect again was attributed to traces of oxygen in the equipment since the residence time in the oxygen-free reactor itself became too
short for the oxygen to be consumed. The dependence of the rates was
found to be about 0.3 and 0.5 order on the emulsifier' concentration and
dose rate, respectively.
An experimental flow equipment was also built by Friis (1973). There
were two important differences from the Hoign el al. system. First, after
initial purging of the system with nitrogen the polymerizations were conducted in the absence of any gas phase, and second the conversion-time
curves were measured by means of a dilatometer rather than sampling. In
the absence of any oxygen, increasing the flow rate increased the rate of
polymerization, confirming perhaps the explanation of Hoign el al. by the
opposite effect of the flow rateo A very small post effect was found.
A large-scale pilot plant constructed by Neutron Products, Inc., (Allen el
al., 1969) was built as a model for a commercial-scale facility, i.e., it had
features that would be important in an industrial process such as heat
exchangers and pumps capable of handling high-viscosity polymeric emulsions, multiple-stage processing, and the capability for continuous operation. Furthermore, all materials used in the latex formulations were of
commercial grade with no purification prior to reaction. The pilot plant
consisted of four complete reactor loops. The reactors varied in total
volume from 3 to 35 gal and were all interconnected to permit multistage
processing.
The reactants were pumped from the surge tank through a heat exchanger embedded in a 6 ft concrete floor that separated the radiation zone
from the rest of the pilot plant. The reactants were then circulated around
the source in a specified configuration, through another heat exchanger,
then back to the surge tank. All or part of the emulsion returning from the
radiation zone could be drawn off for continuous operation.

12. Radiation-Induced Emulsion Polymerization

439

The majority of the efforts of AlIen et al. at the Neutron products facility
were directed toward producing a poly(vinyl acetate) latex suitable for
production of high-quality paints. They employed a practical commercial
formulation containing a comonomer, a complex emulsifier system, a
stabilizer, and a buffer. The highest quality latexes were produced with a
two-stage batch or semibatch process. The first stage, for initiation and
propagation, was operated on either a batch or semibatch basis with
continuous monomer addition. Generally, the first stage was operated to
about 12% residual monomer content at which time the product was
transferred to the second stage for finishing under different conditions. The
second stage was operated in either a batch or a continuous manner.
Experiments were also performed with the plant perating on an entirely
continuous basis and a good quality latex was obtained.
The properties of the latices and the paints produced from them were
evaluated by several commercial firms. The latex properties varied somewhat from sample to sample but, in general, the paints produced from the
latexes were comparable to those produced from high-grade commercial
latexes made by conventional means. One property of the paint produced
from the radiation-catalyzed latex, enamel holdout, which is one measure
of the capability of a paint to serve as a primer for enamel on unpainted
wood, was superior to that of paint made from conventionally produced
latex.
The results showed that production of latices by radiation catalysis is a
commercially feasible process. However, the authors did not appear to have
resolved two process-development problems: the reduction of residual
monomer to commercially acceptable levels and the eIimination of polymer
build-up inside the process lines. The authors noted, however, that these are
problems of formulation and operation rather than of the basic process
itself.
Using the data obtained in the pilot plant, a conceptual design and cost
analysis was made for a commercial-scale radiation catalysis plant (AlIen et
al., 1969, 1971). The conceptual design was based on a two-stage semicontinuous process. The first stage is operated on a semibatch basis with
continuous monomer addition and intermittent feed to the second stage
which is operated on a batch basis. Economic analysis indicated that
capital and operating costs for radiation catalysis and conventional catalysis plants would be about the same; but the authors felt that the greater
enamel holdout of the radiation-initiated product would justify the use of a
radiation catalysis plant. In addition, the early work of Araki et al. (1969)
showed that better wet-scrub resistance was obtained with paints formulated with high molecular weight radiation-produced homovinyl acetate
latices compared with their conventional commercial counterparts.

440

Vivan T. Stannett

The results of the laboratory and pilot plant studies show clearIy that a
radiation-induced emulsion polymerization process for vinyl acetate is of
considerable interest. In addition to the general advantages discussed in the
introduction, the smooth and easy polymerization at low temperatures is of
special interest for vinyl acetate and its copolymers since it leads to very
high molecular weight products. High molecular weight-polymerized acetate latices are known to lead to paints and coatings with better ink and
enamel holdout and superior wet-scrub resistance.
C.

Ethylene Polymerization in Emulsion

These investigation s were all conducted, using 60Co y radiation, in the


laboratories of the Japan Atomic Energy Research Institute in Takasaki
(Konishi et al., 1974). The flow sheet of the apparatus used is shown in
Fig.6. The polymerization vessel was a 500-ml stainless steel autoclave
equipped with a variable motor speed driven propeller. The effect of the
reaction conditions on the rate of polymerization was studied using a
number of emulsifiers. The potassium salts of a number of fatty acids gave
the highest rates, potassium stearate being the best. The rate was found to
be linearIy proportional to the ethylene pressure.
Oxygen was found to inhibit the reaction in the sense that an induction
period was introduced. After this the steady-states rates were not affected.
Increasing the rate of stirring first increased th~ reaction rate which then
levelled off aboye about 600 rpm. No effect of the monomer to water ratio
was found. Both the rate and the number average molecular weight were

Fig. 6.
Flow sheet of apparatus: (1) ethylene cylinder, (2) air-operated automatic
intensifier, (3) autoclave, (4) temperature-controlled oi! bath, (5) 60Co radiation source, (6)
buret for flushing of medium, (7) sampling line, (8) shielded room (Senrui et al., 1974a:
reproduced with permission of Journal of Polymer Science.)

12.

Radiation-Induced

Emulsion Polymerization

441

found to follow Arrhenius law with an activation energy of 4.6 kcal/mol up


to 80C but to become negative at -4.5 kcal/mol after this maximum
temperature. The change was attributed to a near melting of the polyethylene particIes causing an increase in the rate of termination by the
increased mobility of the growing radicals. Chain transfer to the emulsifiers
was found and carboxylic acid groups were introduced in this way to the
polyethylene. The transferred radicals were thought to escape from the
particIes to the aqueous phase.
A more detailed kinetic study (Senrui et al., 1974a) with potassium
myristate as the emulsifier showed that the rate increased only slightIy with
increasing emulsifier concentration and was proportional to the 0.5 power
of the dose rateo Seeded polymerizations showed an increasing rate with
increasing seed volumes but only to a small extent. The solubility of
ethylene in water at the pressures used (200 kg/cm2) is comparatively high
( 0.5 mol/liter) compared with 5 mol/liter in the particIes themselves. In
some ways, therefore, the polymerization mechanism could be compared to
that of vinyl acetate polymerization. An elaborate kinetic scheme for the
polymerization was presented.
A further study substituted ammonium perfluorooctanoate as the emulsifier (Senrui et al., 1974b). This was to avoid the complications introduced by the chain transfer of the growing polyethylene chains to the
emulsifier. Indeed, no fluorine could be detected in the polyethylene produced using this emulsifier. The rate of polymerization was proportional to
the 0.6 power of the emulsifier concentration, in marked contrast to the
fatty acid saIts which had a very low dependence, reminiscent of vinyl
acetate emulsion polymerizations. This tends to confirm the mechanisms
proposed for the fatty acid saIt systems. The number of particIes was found
to be approximately proportional to the 0.6 power on the emulsifier
concentration. The dose rate dependence of the rate was 0.5, similar to that
found with the hydrocarbon emulsifiers. No post polymerization was
observed, even at room temperature. This was confirmed by two separate
types of two-stage experiments. It was postulated that any remaining
radicals soon transferred to monomer and escaped and terminated in the
aqueous phase. The resuIt is in marked contrast to the long-lived radicals
found with the radltion-induced bulk polymerization of ethylene at room
temperature (Gotoda et al., 1965). Seeded polymerizations showed arate
dependence on the number of seeds to be about 0.4 and on the dose rate to
be 0.35. A kinetic scheme was developed that was in reasonable agreement
with the experimental results.
The effect of pressure and temperature on the rate of polymerization was
studied in more detail by using a seeded system developed by Senrui and
Takehisa (1974) and by considering the effect of added alcohols and

442

VivianT. Stannett

electrolytes. With seeded polymerization the rate was found to be proportional to the 1.8 power of the ethylene pressure and to the 2.5 power of
the fugacity. The temperature dependence again slowed a maximum in the
rate, but at 87 rather than at 800C. Mixed emulsifiers showed that the
overall rate was close to the sum of the separate rates. It was believed that
the two, very different emulsifiers, behaved independently due to the big
difference in the solubility parameters
of n-dodecane and nperfluoroheptane (7.8 and 5.6, respectively). Alcohols depressed the rate,
presumably because of chain transfer since they should increase the solubility of the ethylene in the aqueous phase. The addition of salts increased
the rate of polymerization, and this effect was shown to be attributable to
an increase in the number of particles.
Experiments using no emulsifier were conducted in the same stainless
steel autoclave equipment described above (Machi et al., 1975). Stable
latices were obtained, believed to be achieved by hydroxyl end groups and
adsorbed hydroxyl ions. As with a number of the experiments with emulsifiers the polyethylene had a considerable cross-linked gel contento
Finally, the same group of workers studied the radiation-induced emulsion
polymerization of ethylene in a ftow system (Kodama et al., 1974). Both
potassium myristate and ammonium perfluorooctanoate were used as emulsifiers. At longer residence times (above 0.2 hr) the rate of polymerization
was essentially constant. As with the batch system it was assumed that the
number of particles remained constant. In this region the rate was found to
be proportional to the 0.3 power of the potassium myristate concentration
and the 0.5 power of the dose rate, not too different from the batch systems.
The kinetics was developed and estimates of the propagation rate constants
obtained. Despite other similarities between the two systems, these were
quite different, however, from those extracted from the batch experimento
These extensive and important studies of ethylene polymerization have
only been summarized in this review. The original papers should be
consulted for the full details.

D. Tetrafluoroethylene
Machi et al. (1974) first reported an investigation of the radiationinduced emulsion polymerization of tetraftuoroethylene, with ammonium
perfluorooctanoate as the emulsifier. A 200-ml stainless steel autoclave,
equipped with a magnetically driven propeller-type stirrer, was used. The
standard recipe used was 28 gm of monomer in 150 mI of water with 1%
emulsifier (based on the water). n-Hexadecane (2.0 mI) was added to inhibit
any gas-phase polymerization. The polymerizations were conducted at

12. Radiation-Induced Emulsion Polymerization

443

25C. The rate of stirring had a considerable effect on the reaction rateo The
dose rate exponent of the reaction rate was found to be 0.8, with little effect
on the molecular weights. The emulsifier concentration hardly affected the
rates but greatIy changed the molecular weights, which increased with
decreasing emulsifier concentration. The particIe shapes were also affected,
being spherical at less than 0.5% and rod-like and eventually fibrillar above
this concentration. There was considerable post-irradiation polymerization
with increasing molecular weight. Molecular weights were in the 105-106
range. In the absence of emulsifier much higher molecular weights were
obtained, presumably due to lower chain-transfer reactions.
There followed a very important series of papers concerning the
emulsifier-free system (Machi et al., 1978, 1979a-d). The equipment used
was a modification of that used earlier, but the tetrafluoroethylene pressure
was continuously recorded with the use of a strain gauge. In the first paper
of the series (Machi et al., 1978) the rate of polymerization was shown to be
proportional to the 1.0 and 1.3 powers of the dose rate and the initial
pressure, respectively. The activation energies were 0.8 above and - 5.2
kcal/mol below 70C. There was a maximum in the molecular weights at
about the same temperature. This behavior is reminiscent of the behavior of
ethylene and was again attributed to the increased mobility of the growing
chains above the maximum temperature. The very low mobility would also
account for the first-order dependence of the rate on the dose rate below
70C. As before, n-hexadecane pro ved to be an excellent inhibitor of
polymerization in the gas phase. ParticIe sizes in the range of 0.1-0.2
microns were obtained.
A detailed study of the size, distribution, and number of particIes was
presented in the secorid part (Machi et al., 1979a). It was found that the
particIes grow faster at higher dose rates in concert with the higher reaction
rates. At the higher dose rates the particIe size levels off with reaction time
but continues to grow at the lower dose rates. The size and number of
particIes are essentially independent of the temperature from 30-100oC. The
effect of pressure was complex but the particIe size tends to increase and the
number of partic1es to decrease with increasing pressure. The partic1e size
distribution also becomes broader, perhaps due to increased flocculation
effects at higher pressures. It was concIuded that after the 'generation of the
particIes, which tends to be in the first five minutes, they grow by propagation at the surfaces. The particIes are believed to be stabilized by
hydroxyl and carboxyl end groups and possibly by adsorption of ions. A
discussion of the mechanism of particIe formation and .the loci of the
polymerization was also presented.
Machi et al. (1979b) also investigated the concurrent formation of
hydrofluoric acid during the polymerization due to the radiolysis of the

444

Vivian T. Stannett

monomer. It was found to be unaffected by the agitation rate or the


presence of n-hexadecane. G values varied from 27 to 180, being larger at
high pressures and increasing with increasing dose rateo Possible mechanisms for the formation of the acid were discussed. The molecular
weights were normally in the 105--106range but a few samples from the
emulsifier-free grade were higher than 107. This was in the conventional
molding powder grade region and therefore of considerable interest. A
number of particles to decrease with increasing pressu~e. The particle size
control the molecular weights. Oxygen was found to decrease the rate and
the molecular weight; this was attributed to its capacity to inhibit radical s
in the aqueous phase and to diffuse into the particles to retard the growing
chains. A number of additives were studied but the conventional radical
scavengers were the most effective. Hydroquinone and benzoquinone are
highly soluble in water but barely soluble in the polymer particles. Their
effect was, therefore, to inhibit the rate of polymerization but to increase the
molecular weight. Molecular weights of 2 x 107 were obtained in the
presence of hydroquinone. Ethylenediamine and triethylamine were both
found to retard the rate, but the molecular weight was decreased by the
latter and increased by the former additive. The decrease was ascribed to
the power of triethylamine to absorb on the particles inhibiting the chain
growth. Further details of these and other additives were described, including their effect on particle size and distribution.
In the final paper of this important series (Machi et al., 1979d) the effect
of the reaction conditions on the stability of the tetrafluoroethylene latices
was studied. The stabilization was believed to be due to either carboxyl end
groups or to the adsorbed hydrofluoric acid formed by radiolysis of the
monomer or to both. The surface charge densities, electrophoretic mobilities, and infrared spectra were all measured. The storage stability of the
latices, measured by the appearance of coagulum, was found to be proportional to the total dose and the polymer concentration and not to
depend on the dose rateo Since the stabilization results from the radiolysis
of the monomer and water this is quite reasonable as is the effect of
concentration. The presence of monomer was necessary in the sense that
irradiation of the latex in the absence of monomer did not improve the
storage stability.

E. Tetrajluorethylene-Propylene and Other Copolymers


Ishigure et al. (1964) reported that the radiation-induced polymerization
of tetrafluoroethylene and propylene in the liquid phase produced an
essentially alternating rubbery copolymer. Ito et al. (1974a) and Matsuda et
al. (1974a) studied the same system in emulsion using a variety of mixed and
pure emulsifiers. The experiments were conducted in a stirred 200-ml

12. Radiation-lnducedEmulsion Polymerization

445

stainless steel autoclave with 60Co radiation. Ammonium perfluorooctanoate gave the highest rates and molecular weights and was therefore
chosen for the detailed investigation. Increasing the stirring speed first
increased the rates and stability but then essentially leveled off. The
emulsion system gave a higher rate and molecular weight than the corresponding polymerizations in bulk. Increasing the temperature slightly
decreased the rate and increased the molecular weight. The copolymers
were essentially alternating, as were the bulk copolymerizations previously
mentioned. Oxygen caused a small induction period which was followed by
a resumption of the normal rates. The effect of the emulsifier concentration
was to cause an increase in the rate and molecular weight above 1% and
then to have little further effect. The number average molecular weight
ranged from 3 to 8 X104. Matsuda et al. (1974b) continued their detailed
studies using ammonium perfluorooctanoate as the emulsifier. The rate of
polymerization was found to be 0.9 order and 0.26 order with respect to the
dose rate and the emulsifier concentration, respectively. Both the rate and
the molecular weight increased with increasing tetrafluoroethylene content
of the feed. The copolymer composition, however, remained alternating
over a very wide range of compositions. The dose rate dependence indicated
that most of the termination was by degradative chain transfer to the
propylene monomer.
Danno et al. (1974) presented additional work showing that above the
critical micelle concentration (CMC) of the emulsifier the dose rate dependence of the rate was actually 0.7 and below the CMC as 0.6. The molecular weight was higher above the critical micelle concentration and independent of the dose rate and lower below the CMC and decreasing with
increasing dose rateo These and other details of the polymerization reaction
were discussed further in terms of the particle sizes, which are larger and less
numerous at zero or low-emulsifier contents and at lower dose rates and
degradative chain transfer to propylene monomer. Okamoto and Suzuki
(1974) presented the results of a brief study of the effects of a number of
additives on the polymerization. Hexachlorethane and carbon tetrachloride
were found to decrease markedly the molecular weight; the former had no
effect whereas .the latter decreased the rateo 1,1,2-Trichloro-l,2,2trifluoroethane had no effect on either the rate or 'molecular weight.
Alcohols reduced the rate in the order isopropyl > ethyl > methyl, in the
same order as the ease of hydrogen abstraction but had no effect on the
molecular weight. This indicated that the primary radical s were reacting
with the alcohols in the aqueous phase but not in the particles themselves.
Diphenyl picrylhydrazyl and p-benzoquinone, as expected, completely inhibited the polymerizations above 3 x 10- 3 molar concentration. Okamoto
and Suzuki (1975) also investigated the hydrofluoric acid yields during the
polymerization. The yields were first order with the dose rateo The presence

446

Vivian T. Stannett

ve
mor.omer

VA
monomer

Metering

pump

Metering

pump

Soap

solution

tank

Solvent
tank

Liquid
level
gauge

EmuHifying
tank

Sample

Temperature
bath

Metedng
pump

Vent

Rotameters

Off produc.t
drum

Fig. 7.
Schematic diagram for radiation-induced copolymerization pilot plant (Stahel et
al., 1979; reproduced with permission of Journal 01 Applied Polymer Science.)

.
12. Radiation-Induced Emulsion Polymerization

447

of oxygen increased the yield, but was decreased greatly when more than
1% of emulsifier was used. G values were much lower than with pure
tetrafluoroethylene emulsion polymerization. Some tentative explanations
were presented.
Ito et al. (1974b) also studied the copolymerization of ethylene and
hexafluoropropylene with the same autoclave batch system used with
ethylene and with ammonium perfluorooctanoate as the emulsifier. The
rate of polymerization was found to be proportional to the 1.6 power of the
ethylene fugacity. The ratio of ethylene units in the copolymer increased
linearly with the reaction rate, extrapolating to zero at zero rateo This
agrees with the fact that hexafluoropropylene does not homopolymerize. A
number of thermal and other properties of the copolymers were determined.
Properties varied from semicrystalline polyethylene types to amorphous
rubbers. Memetea et al. (1977) attempted to graft styrene to a previously
radiation prepared, poly(vinyl chloride) latex. A recirculating flow system
was used, very similar to those used earlier by Stahel et al. (1969) and others
and shown schematically in Fig. 7. In contrast to earlier results published
by Wang (1963) very little grafting occurred if the correct extraction
procedures for the homopolystyrene were used. Interestingly, the polymerization of styrene on the poly(vinyl chloride) seeds ciosely approximated
Smith-Ewart Case 2 behavior. The less than 3 %grafting was attributed to
the low total dose needed to effect the essentially complete polymerization
of the styrene, only about 0.1 Mrad. Parallel experiments with styreneswollen films and the literature results did give reasonable grafting yields at
higher doses (e.g., 33% with 3.6 Mrad).
Finally, Stahel et al. (1979) studied the radiation-induced emulsion
copolymerization of vinyl chloride and vinyl acetate in an engineering flow
system with a l-gallon reactor, shown schematically in Fig. 7. The copolymerization proceeded smoothly to high conversions. The rate was found to
be proportional to the 0.17 power of the emulsifier concentration, within
the range reported for the individual monomers. The activation energies
were 5.5 and -2.0 kca1/mol for the rate and molecular weight, respectively,
in good agreement with literature results for similar, chemically initiated
systems. The res.ults showed that radiation does represent a viable and
practical method for producing stable lattices with high conversions and
solids content in engineering systems'.
References
Abkin, A. D., Mezhirova, L. P., Iakovleva, M. K., Matveeva, A. V., Khomikovskii, P. M., and
Medvedev, S. S. (1959). Vysokomo/. Soedin. 1, 68.
Acres, G. J. K., and DaIton, F. L. (I963a). J. Po/ym. Sci. A 1, 3009.

448

Vivian T. Stannett

Acres, G. J. K., and Dalton, F. L. (1963b). J. Po/ym. Sci. A 1,2419.


AIIen, P. E. M., Downer, J. M., Hastings, G. W., MelviIle, H. W., Molyneux, P., and Urwin, J. R.
(1956). Nalure (London) 177,910.
Allen, P. E. M., Bumett, G. M., Downer, J. M., Hardy, R., and Melville, H. W. (1958). Nalure
(London) 182, 245.
AIIen, P. E. M., Bumett, G. M., Downer, l. M., and Melville, H. W. (1960). Makromo/. Chem.
38,72.
AIIen, P. E. M., Burnett, G. M., Downer, J. M., and Majer, J. R. (1962). Makromo/. Chem.
58, 169.
AIIen, R. S., RansohotT,. l. A., and Woodard, D. G. (1969). USAEC Rep. ORO 673 (October
1969); See also Proc. Jpn. Conf Radioisolopes 9, 598.
AIIen, R. S., RansohotT, J. A., and Woodard, D. G. (1971). Isolopes and Radial. Techno/. 9, 92.
Araki, K., Stannett, V., Gervasi, J. J., and Keamey, J. J. (1967). USAEC Rep. TID-24281,
December l.
Araki, K., Stannett, V., Gervasi, J. A., and Keamey, J. J. (1969). J. App/. Po/ym. Sci. 13, 1175.
BaIlantine, D. S. (1954). Brookhaven National Laboratory, New York, Rep. BNL-294 (T-50),
March.
Barriac, J., Oda, E., Russo, S., Stahel, E. P., and Stannett, V. (1969). Proe. Jpn. Conf.
Radioisolopes 9, 605.
Barriac, l., Knorr, R., Stahel, E. P., and Stannett, V. (1976). Adv. Chem. Ser. 24, 142.
Bradford, E. B., and VanderhotT, J. W. (1955). J. App/. Phys. 26, 864.
Bradford, E. B., VanderhotT, J. W., Tarkowski, H. L., and Wilkinson! B. W. (1961). J. Po/ym. Sci.
50, 265.
Cockbain, E. G., Pendle, T. D., and Turner, D. T. (1958). Chem. Ind. (London) 759.
Cockbain, E. G., Pendle, T. D., and Tumer, D. T. (1959). J. Po/ym. Sci. 39, 419.
Cooper, W., and Vaughan, G. (1959). J. Po/ym. Sci. 37, 241.
Cooper, W., Vaughan, G., Miller, S., and Fielden, M. (1959a). J. Po/ym. Sci. 34, 651.
Cooper, W., Vaughan, G., and Madden, R. W. (1959b). J. App/. Po/ym. Sei. 1, 329.
Cooper, W., Sewell, P. R., and Vaughan, G. (l959c). J. Po/ym. Sci. 41, 167.
Danno, A., Matsuda, O., Okamoto, J., Suzuki, N., and Ito, M. (1974). J. Po/ym. Sci.-Chem. Ed.
12, 1871.
Ewart, R. H., and Smith, W. V. (1948). J. Chem. Phys. 16,592.
Friis, N. (1973). Danish A.E.C. Riso Rep. 282.
Gardinovacki, B., Lervik, H., Ugelstad, J., and Sund, E. (1971). Pure App/. Chem. 26, 121.
Garreau, H., Stannett, V., Shiota, H., and Williams, J. L. (1979). J. Co/loid Interface Sci. 71,130.
Garreau, H., Yoshida, K., Ishigure, K., and Stannett, V. (1980). J. Maeromo/. Sci.-Chem.
A14, 739.
Glazkova, K. G. el al. (1971). Vysokomo/. Soedin. BI3, 173.
Gotoda, M., Machi, S., Hagiwara, M., and Kagiya, T. (1965). J. Po/ym. Sci. A 3, 2931.
Hansen, F. K., and Ugelstad, J. (1976). Ruber Chem. Teehno/.49, 537.
Hayashi, K., and Okamura, S. (1980). Ann. Rep. Osaka Lob. JAERl M-9214, 113.
Hoign, J., and O'Neill, T. (1972). J. Po/ym. Sei. A-110, 581.
Hoign, J., O'Neill, T., and Pinkava, J. (1972). Proe. Tihany Symp. Radial. Chem., 3rd, Tihany,
Hungary. 1971 1, 713.
Hummel, D. (1963). Agnew. Chem. In l. Ed. 2, 295.
Hummel, D., Ley, G., and Schneider, C. (1962). Adv. Chem. Ser. 34, 60.
Hummel, D., Ley, G. J. M., and Schneider, C. (1967). Adv. Chem. Ser. 66, 184.
Hummel, D., Ley, G. J. M., and Schneider, C. (1969). J. Po/ym. C 27, 119.
Inagaki, H., Okamura, S., Motoyama, T., and Manabe, T. (1960a). .. Large Radiation Sources in
Industry," Vol. I. pp. 368-373. IAEA, Vienna.

12. Radiation-Induced Emulsion Polymerization

449

Inagaki, H., Yagi, K., Saeki, S., and Okamura, S. (1960b). Chem. High Polym. (Jpn.) 17,37.
Inagaki, H., Yagi, K., Saeki, S., and Okamura, S. (l960c). Chem. High Polym. (Jpn.) 17, 135.
Ishigure, K., and Stannett, V. (1974). J. Macromol. Sci. Chem. AS, 337.
Ishigure, K., Tabata, Y., and Sobue, H. (1964). J. Polym. Sci. A2, 2235.
Ishigure, K., Yoshida. K., and Stannett, V. (1973). J. Macromol. Sci. A7, 813.
Ishigure, K., O'Neill, T., Stahel, E. P., and Stannett, V. (1974). J. Macromol. Sci.-Chem. AS, 353.
Ito, l., and Watanabe, T. (1965). J. Chem. Soc. Jpn.-/nd. Chem. Sect. 68, 552.
Ito, M., Matsuda, O., Okamoto, J., Suzuki, N., and Tabata, Y. (1974a). J. Macromol. Sci.-Chem.
AS, 775.
Ito, M., Senrui, S., and Takehisa, M. (1974b). J. Polym. Sci. 12, 627.
Izumi, Z. (1967). J. Polym. Sci. A-l 5, 469.
Izumi, Z., Kuichi, H., and Watanabe, M. (1967). J. Polym. Sci. A-l 5, 455.
Kalyadin, V. G., Sirlibaev, T. S., and Tirkashev, 1. (1975). Nauch Tr. Tashkent Un-T (462) 100;
(1974) CHAB. 83, 193858f.
Kamiyama, H. (1974). Ann. Rep. Osaka Lab. JAER15029, 33.
Kamiyama, H. (1975). Ann. Rep. Osaka Lab. JAER/ M-6260, 23.
Kamiyama, H., and Okamura, S. (1969). Ann. Rep. Osaka Lab. JAER/5022, 33.
Kamiyama, H., and Saito, K. (1975). Ann. Rep. Osaka Lab. JAERI 5030, 147.
Kamiyama, H., and Shimizu, Y. (1975). Ann. Rep. Osaka Lab. JAER/503O, 148.
Kamiyama, H., Kitayama, M., and Okamura, S. (1970). Ann. Rep. Osaka Lab. JAER/5026(3),
26.
Karpov, V. L., Lukhovitskii, V. l., Polikarpov, V. V., Lebedeva, A. M., and Lagucheva, R. M.
(1968). Vysokomol. Soedin. AI0, 835.
Karpov, V. L., Lukhovitskii, V. l., Polikarpov, V. V., Lebedeva, A. M., and Lagucheva, R. M.
(1969). Khim. Vys. Energ. 4, 173.
Karpov, V. L., Smirnov, A. M., and Lukhovitskii, V. 1. (1971). Khim. Vys. Energ. 5, 470.
Karpov, V. L., Smirnov, A. M., and Lukhovitskii, V. 1. (1972). Vysokomol. Soedin. B-l 4, 6.
Karpov, V. L., Lukhovitskii, V. l., and Lebedeva, A. M. (1973a). Vysokomol. Soedin. A15, 2465.
Karpov, V. L., Smirnov, A. M., and Lukhovitskii, V. 1. (1973b). Vysokomol. Soedin. B15, 726.
Karpov, V. L., Lukhovitskii, V. l., and Pozdeeva, R. M. (1973c). Vysokomol. Soedin. B15, 907.
Karpov, V. L., Polikarpov, V. V., Lukhovitskii, V. l., and Pozdeeva, R. M. (1974). Vysokomol.
Soedin. A16, 2207.
Karpov, V. L., Smirnov, A., Lukhovitskii, V. l., and Pozdeeva, R. M. (1977). Plast. Massy
(USSR) 3, 18.
Klein, A., Stannett, V., and Litt, M. (1975). Br. Polym. J. 7, 139.
Kodama, A., Senrui, S., and Takehisa, M. (1974). J. Polym. Sci. 12, 2403.
Konishi, K., Senrui, S., Suwa, T., and Takehisa, M. (1974). J. Polym. Sci. 12,83.
Litt, M., Patsiga, R., and Stannett, V. (1960). J. Phys. Chem. 64, 801.
Litt, M., Patsiga, R., and Stannett, V. (1970). J. Polym. Sci. A-l S, 3607.
Machi, S., Suwa, T., and Takehisa, M. (1974). J. Appl. Polym. Sci. lS, 2249.
Machi, S., Suwa, T., N'akajima, H., and Takehisa, M. (1975). J. Polym. Sci. Polym. Lett, Ed.
13, 369.
.
Machi, S., Suwa, T., Watanabe, T., and Okamoto, J. (1978). J. Polym. Sci. Chem. Ed..16, 2931.
Machi, S., Suwa, T., Watanabe, T., Seguchi, T., and Okamoto, J. (1979a). J. Polym. Sci. Chem.

Ed. 17, 111.

Machi, S., Suwa, T., Watanabe, T., and Okamoto, J. (1979b). J. Polym. Sci. Chem. Ed. 17, 129.
Machi, S., Watanabe, T., Suwa, T., and Okamoto, J. (1979c). J. Appl. Polym. Sci. 23, 967.
Machi, S., Suwa, T., Watanabe, T., and Okamoto, J. (l979d). J. Polym. Sci. Chem. Ed. 17,503.
Matsuda, O., Suzuki, N., and Okamoto, J. (1974a). J. Macromol. Sci.-Chem. AS, 793.
Matsuda, O., Suzuki, N., and Okamoto, J. (l974b). J. Polym. Sci. Chem. Ed. 12, 2045.

450

Vivian T. Stannett

Memetea, T., Mitri, K., Stahel, E. P., and Stannett, V. (1977). J. Macromol. Sei.-Chem. All, 337.
Mukoyama, E., and Toriaka, S. (1962). Chem. High Polym. (Jpn.) 19, 337.
Ohdan, K., and Okamura, S. (1969). Ann. Rep. OsakaLab. JAERI 5022(2),37.
Ohdan, K., Kamiyama, H., and Okamura, S. (1970). Ann. Rep. Osaka Lab. JAERI5026(3), 15.
Okamoto, J., and Suzuki, N. (1974). J. Polym. Sci.-Chem. Ed. 12, 2693.
Okamoto, J., and Suzuki, N. (1975). J. Macromol. Sci.-Chem. A9, 285.
O'Neill, T., and Stannett, V. (1974). J. Macromol. Sci.-Chem. AS, 949.
Panajkar, M. S., and Rao, K. N. (1979). Radial. EjJecls 41,71.
Senrui, S., and Takehisa, M. (1974). J. Polym. Sci. 12, 535.
Senrui, S., Suwa, T., and Takehisa, M. (1974a). J. Polym. Sci. 12,93.
Senrui, S., Suwa, T., and Takehisa, M. (1974b). J. Polym. Sci. 12, 105. .
Smith, W. V., (1948). J. Am. Chem. Soco 70, 3695.
Stahel, E. P., and Stannett, V. (1969). "Large Radiation Sources for Industrial Processes," pp.
135-150. IAEA, Vienna.
Stahe1, E. P., and Stannett, V. (1971). USAEC Rep. ORO. 3687-1.
Stahe1, E. P., Tsai, J. T., and Stannett, V. (1979). J. Appl. Polym. Sei. 23, 2701.
Sunardi, F. (1979). J. Appl. Polym. Sci. 24, 1031.
Van der HofT, B. M. E. (1958). J. Polym. Sci. 33, 487.
Wang, U. P. (1962). J. Chinese Chem. Soco (Taiwan) Seco II 9, 195.
Wang, U. P. (1963). J. Chinese Chem. Soco (Taiwan) 4, 171.
.
Zimmt, W. S. (1959). J. Appl. Polym. Sci. 1,323.

lndex
D

A
Diffusion, 296

Acrylonitrile, 372, 424


Adhesion, 291
Adhesives, 300

Dispersion interaction, 250


DLVO theory, 8, 17,85

tufted carpet applications,

306

Adsorption, 250, 253, 291


of anion-active emulsifiers, 267
area, 232, 254
characteristics of surfactants, 256

E
Electrostatic
Electrostatic

isotherms, 251, 254


kinetics, 253, 268
Application properties of latexes, 315
Attractive interactions of particles, 8, 12

Electrostatic repulsion,
Emulsification

with mixed emulsifier systems, 408


of polymer solutions,
Emulsifier

B
Butadiene,

effects, 6
forces, 3

399

adsorption, 250, 252, 268


anionic and nonionic, mixtures of, 227
area occupied by saturated monolayer,
231-233

372, 431

tran~fer, 242

Emulsion polymerization
choice of emulsifier, effect of, 221
distribution of components and phases, 53
emulsifier-free, 257
initiatibn mechanisms, 51

Chemisorption, 298
Coagulation, 2, 7, 24
with aluminum salts, 19-22

with ionic emulsifiers, 230-233


kinetics~ 51, 145, 191,278,283,319,357
with nonionic emulsifiers, 229

as kinetic process, 14
measurements of rate, 15
Colloidal behavior of polymerization
261
Colloidal characteristics

systems,

particle stability, 45
radiation-induced, 415

of latexes, 264

Ethylene, 440

Colloidal stability
definition, 2

theory, 8
Constant rate period, 286
Copolymerization
block and graft, 434
random, 433

Fiber bonding, 302 .


FIocculation, 2, 7, 24
of particles, 262

Critical coagulation concentration, 15


determination of values, 17-19
Critical micelle concentration, 51, 63

Critical surface tensions of polymeric materials,


294

Hamaker constant,

451

10

452

Index

Hamaker equation, 291


Heterocoagulation, 36
Hydrophile-lipophile
balance, 225

Molecular weiglit distributions,


in bulk, 116

concepts in emulsion, 118, 183


experimental deterrninations, 139-141
generaltheory,
120-124, 183
polydispersity ratio, 132-133
Monodisperse latexes, 228
Monomers

1
Initiation
emulsifier-induced decomposition,
in monomer droplets, 396
radiation-induced, 415
systems, 329
Interface characteristics, 249
liquid-liquid,
251
liquid-polymer,
254
Interfacial tension, 249, 251
monomer-water,
256
Ions

115

239

concentration within reaction loci, 151


emulsification, 224
polarity effects, 258

solubility in water and polymer,


stabilization, 237

372

water insoluble compounds, effect of addition


of, 401

interaction with water, 19


Nonionic surfactants,
Nucleation

organic, 26

31

homogeneous, 52, 73
micellar, 54, 63

in monomer droplets,
Latex application properties,
Latex particles
.
colloidal behavior, 2

86

315

effect of electrolyte on, 6-8


formation mechanisms, 51
nature of, 2
nucleation, 51
sterically stabilized, 5
surface charge, 3
Lyophobic colloids, 8, 14
theory of stability, 8

M
Macromolecules, oligomeric, 53
adsorbed on particles, 5, 42, 43
grafted on particles, 5, 42, 43
Micellar nucleation
newer methods, 63
Smith'- Ewart theory, 54
Micellar size, 234
effect of polymerization, 235
Mixed electrolyte systems, 35
Molecular weight, 325
of styrene, 421
theoretical predictions, 93

o
Oligomeric radicals, 53

P
Paints and industrial coatings, 309
Paper coatings, 310
Particle generation
homogeneous nucleation, 73
mechanisms, 51, 257, 270, 370
micellar, 54, 63
in monomer droplets, 86
Particle number, 327, 334, 363
effect of emulsifier, 233
initial emulsifier concentration,
of,69
rate of initiator decomposition,
70

as function
as function of,

in styrene emulsion polymerization,


Particle size, 93, 258
Particle size distribution
363
approximate

90

(PSD), 258, 327, 334,

approaches,

95

453

Index
in batch polymerization, 99
in continuous polymerization,
cumulative, 335
experimental investigations,
functions, 102

mechanisms,

56

reabsorption, 199
Radical desorption, 191, 199, 204, 369

105
109-114

monodispersity index, 260


population balance model, 96
of radiation-induced polystyrene, 422
theoretical predictions, 93
Partic1e stability, 1-48
in absence of added emulsifier, 45
Peptization, 40
Poly(alkyl acrylates), 265
adsorption of C12H2SS04Na, 285
colloidal characteristics of latexes, 264
structural characteristics of latexes, 277

Radical number per locus


theoretical prediction, 181
theory for generation of pairs, 185
Rate of interphase transport, 392
Reactors
batch, 320, 359
continuous stirred-tank,

latex ccc values, 19,31


partic1e number dependence
concentration

on emulsifier

experimental, 90
theoretical, 69

Secondary minimum effects, 9, 22


Smith-Ewart theory, 54
Case 1, 322
Case 2, 248, 323, 361,419-425
derivations, 156-160, 194-198
deviations from Case 2, 367
solution of nonsteady state, 167-176
solution of steady-state, 164-167
solutions of equation, 160-164
Sodium dodecyl sulfate, 82, 261
adsorption
area, 256, 264

energy characteristics,
Solvation effects, 6

polymerization,

Poly(vinyl acetate), 428, 437


Potential energy
coagulation versus repeptization,
diagrams, 9-13,16
'
primary minimum, 9
secondary minimum, 9, 22
effects, 22-26

R
Radical absorption
inverse absorption efficiency,

energy, 252
kinetics, 268

partic1e size distributions


experimental, 110
theoretical, 1l1
radiation-induced

of, 344

Rubber goods, 313

Poly(butyl acrylate), 42, 281, 432


Poly(butyl methacrylate), 262
Poly(ethyl acrylate), 42, 279, 432
Polymer colloids, definition, 2
Polymer dispersions, preparation, 396
Poly(methyl acrylate), 42, 259, 261,431
Poly(methyl methacrylate), 82, 259, 425
PSDs, experimental and theoretical,
comparison of, 112
Polystyrene, 259
adsorption of C12H2SS04Na, 285
coagulation by lanthanium nitrate, 34
colloidal characteristics of latexes, 264

333, 339, 359

design considerations, 378


dynamics in design, consideration
heat-removal techniques, 330
on-line control, 345, 349
reaction time, reduction of, 330

61

418

41

252, 253

Stability of Iyophobic colloids, 8


adsorbed or grafted macromolecules,
of,42
Steric effects, 6
Steric stabilization, 4
Surface coagulation, 39'
Surface groupings, 3
Surface potential, 3
Surface properties of latex partic1es. 2
Surfactants
adsorption characteristics, 252- 256
Swelling
of outer shell of partic1e, 407
and phase distributions, 384

effect

454

Index
T

Taek, 298
Tetrafluoroethylene,

442

polymer latex cee values, 27, 31

v
Vinyl ehloride, 429
latex panicle size distribution,
Vinylidene ehloride, 431

377

Potrebbero piacerti anche