Você está na página 1de 24

1

VI. ELEMENTS OF TECHNOLOGY


OF CHAIN POLYMERIZATION

VI.1. Generalities
Chain polymerization is one of the most used methods for the industrial
synthesis of polymers and copolymers used as thermoplastic materials, rubbers,
synthetic fibres, adhesives, resins, etc. All the mechanisms (free-radical, cationic,
anionic and coordinative) have practical applications, but the majority of the
industrial processes are based on free-radical polymerization, since it requires less
restrictive reaction conditions (especially in what concerns the purity of the
reagents and the reaction medium).
As a general rule, if a monomer can be polymerized by free radicals, this is
the method chosen in industry, except for the cases when the microstructure
required for the final product must be obtained by other mechanisms available.
Such a case is the industrial synthesis of rubbers via anionic and coordinative
polymerization or the Ziegler-Natta polymerization of olefins.
Whatever the nature of the active centre, some features are common for all the
additive polymerizations: existence of the characteristic stages of a chain reaction
(initiation, propagation, termination and transfer) and presence of polymer with
high molecular mass in the system in the early stages of the polymerization. The
polymerization degree does not depend on conversion (except for ome spacial
cases of anionic polymerization) but is determined (in absence of autoacceleration
or chain transfer) by the ratio between the concentrations of the monomer and the
initiator.
Another important feature of chain polymerization is its exceptional
exothermicity, that imposes to find out suitable methods to eliminate efficiently the
reaction heat.
The structure of the polymer depends on the mechanism of polymerization:
linear and stereospecific chains are obtained using more selective mechanisms
(such as coordinative polymerization) while via free-radicals the reaction is less
specific and branched (or even cross-linked) chains are obtained in many cases.

From a practical point of view, there are several polymerization techniques,


different by the composition of the reaction mass, the number of phases and the
properties of the final polymer. The classical chain polymerization techniques are
bulk polymerization, solution polymerization, suspension polymerization and
emulsion polymerization. The latter two techniques are heterogeneous ones, while
the first two may be either homogeneous or precipitant.
In the following, each technique will be detailed and some practical examples
will be presented.

VI.2. Bulk Polymerization


VI.2.1. Main Features of Bulk Polymerization
Bulk polymerization (also called mass polymerization) is the simplest
technique in what concerns the number of components and phases of the reaction
system.
At the beginning of the process, the reaction mass consists only of monomer
and initiator, or even only monomer (if initiation is achieved by physical methods).
The chemical initiator must be monomer-soluble.
Depending on the solubility of the polymer in the monomer, there are two
ways in which the reaction may progress. If the polymer is soluble in the monomer,
then the polymerization system is homogeneous (homogeneous bulk
polymerization). If the monomer is not a solvent for the polymer (case of some
monomers as for instance, vinyl chloride or ethylene) then the polymer, once
formed, separates as a second phase. This is called precipitant bulk
polymerization and will be treated separately.
VI.2.1.1. Homogeneous Bulk Polymerization
In the homogeneous bulk polymerization, the systems remains homogeneous
until the end of the process; at 100% conversion, the reaction mass will consist
only of polymer (that incorporates the initiator at the end of the macromolecular
chains), that forms a compact block, hence the name of the technique. Obviously,
in an industrial system, the temperature must be raised high enough to keep the
polymer flowing, in order to be able to ensure the systems mixing. In some
practical applications however (dental surgery for instance) the system is not
stirred and the polymerization is photochemically initiated in a stationary layer. If

no additives (pigments, fillers, etc.) are added, then the polymer does not contain
any impurities, so it will be completely transparent.
Homogeneous mass polymerization has the advantage of the simplicity and of
the high purity of the reaction product, but at the same time there are some
significant problems raised when scaling up the process for industrial plants.
The most important problems that have to be solved (and that are difficult
enough to make bulk polymerization a process less used as compared to the other
techniques) are the following:
a. Viscosity of the reaction mass
Viscosity of polymer solutions increases with both the molecular mass and the
concentration of the polymer dissolved. In the homogeneous bulk polymerization,
the reaction mass consists of monomer that dissolves polymer with high molecular
mass and the concentration of the polymer increases with the conversion.
Polymer solutions are extremely viscous (105 to 108 times more viscous than
water). Moreover, during a high conversion polymerization, viscosity increases
several orders of magnitude (starting from the monomer viscosity, around 10-3 Pas
to the viscosity of the molten polymer, 102-105Pas). Since stirrers are
constructively adapted to the viscosity of the mixed fluid, the same stirrer cannot
be used for both the initial and the final stage of the process. Usually the reaction is
performed in several reactors, each one with a stirrer adapted to the range of
viscosities corresponding to a given interval of conversion: helical stirrers and
anchor-type mixers for lower viscosities, planetary stirrers or rotating profiled
cylindrical mixers for the late stages of the polymerization. In some cases,
stationary mixing devices are used (elements that ensure by their architecture - a
spiral flow of the reaction mass that produces a certain degree of mixing). There is
also a difference in what concerns the rotating speed of the stirrers, that diminishes
significantly with the viscosity (while the power consumption increases).
Another factor to take into consideration is that for viscous monomer-polymer
mixtures, achieving turbulent flow is impossible; the values of the Reynolds
criterion for concentrated polymer solutions are very low (100 to 102) while Prandtl
values are 10 to 1000 times higher than for low-molecular compounds. This has a
significant impact on the removal of the reaction heat (by convection) that
becomes extremely difficult when the conversion exceeds a certain level.
The high viscosity has to be taken into account also when polymer solutions
(reaction mass with a given conversion degree) have to be transported through
pipelines. Not only the energy consumption for pumping is very high, but also the
viscous solution may block the pipes, fittings and pumps. Classical centrifugal and
volumetric pumps cannot be used; they are replaced with screw pumps (similar

with the body of a screw extruder) or peristaltic devices (where the moving parts
are outside the hose that contains the polymer); another solution is not to use
pumps at all, but to take advantage of the gravitational flow when transporting the
reaction mass from one reactor to another. Moreover, to avoid blocking, diameters
of the pipes are larger than the usual for low-molecular compounds.
b. Autoacceleration
Autoacceleration is a phenomenon typical for the bulk polymerization, that
consists of the increase of the reaction rate without any outside intervention (any
change in the reaction parameters). As previously detailed (see chapter V.3.10),
autoacceleration leads to a characteristic S shaped conversion time curve (figure
V.1) and to a peak of the polymerization rate.
The autoacceleration is due to the massive increase of the viscosity that
accompanies the advancement of the reaction; this leads to a lower diffusion rate
that affects mainly the macromolecular species. Consequently, the termination rate
decreases and the overall polymerization rate increases, until the monomer
concentration becomes small enough to decrease the rate of the process. Another
consequence of the autoacceleration is an increase of the average molecular
weight, which, in turn, leads to a supplementary raise of the viscosity.
The practical consequence of the autoacceleration is that at a given moment
the reaction becomes very difficult to control, since the reaction rate (and the
corresponding reaction heat evolved) increase is of at least one order of magnitude.
Any computation of a chemical reactor has to take this phenomenon into account,
and estimation of the heat transfer area and stirring power must be done for the less
favourable case, i.e. the autoacceleration peak.
c. Polymerization heat
As already stated, the thermal effect of chain polymerization is higher than the
usual value for reactions between organic compounds or step-growth processes
(see table V.1.). Moreover, the heat is generated inside a reaction mass with a low
thermal conductivity (3-4 times less than water) and with a very high viscosity, that
makes efficient stirring impossible. Therefore, removal of the reaction heat by
either convection or conduction mechanisms is impossible in homogeneous mass
polymerization. Only approximately 10% of the reaction heat can be removed by
circulating a cooling agent through the shell or coil of an industrial reactor, and this
amount reduces to less than 1% for the autoacceleration peak. The extra heat
remains in the reaction mass and would lead, for a homogeneous mass
polymerization, to a huge increase of the temperature for high conversion. This
temperature does exceed both the boiling point of the monomer and the destruction
temperature of the polymer, therefore it would produce explosive decomposition of
the reaction mass.

Thermal control of the homogeneous mass polymerization is the most


important problem that has to be solved in industrial plants, and several solutions
are applied, that will be detailed below, tailored according the particular features of
each case of polymerization.

d. Monomer conversion
Most chain polymerization may be considered as irreversible, so 100%
conversion can be obtained at least in theory. In practice a total conversion might
not be useful in industrial plants, for several reasons:
the viscosity increase might limit the stirring efficiency; in a reactor with a
deficient mixing there may develop temperature gradients that may affect the
properties of the product (overheating may lead to partial decomposition affecting
the colour and the mechanical properties of the polymer);
the thermal control is so difficult during last stages of the polymerization that
limiting conversion to lower values might be the only solution for ensuring an
efficient heat removal;
in systems where chain transfer reactions have a high intensity, cross-linked
polymer is obtained at high conversion due to the transfer towards the polymer; to
limit the amount of insoluble product, conversion must be limited at lower than the
unit values.
some polymers have melting points higher than the decomposition temperature, so
the fluidity of the reaction mass cannot be maintained at high concentrations of the
polymer.
In all the above cases, the final product contains a given amount of unreacted
monomer; this needs to be quasi-totally eliminated from the polymer, since the
high toxicity of the monomers imposes a residual monomer content in the
commercial polymers of the order of ppm. Consequently, a demonomerization
stage must be included in all the flows where the conversion is not total. The
removed monomer is then purified and recycled in the synthesis.
To solve the above described problems, several technological solutions have
been adopted for the homogeneous bulk polymerization:
1. Low conversion polymerization
This methods aims at limiting the heat evolved during the process by limiting
the conversion at a value low enough not to enter the autoacceleration range. This
way only a fraction of the reaction heat is generated, and the reaction rate is low
enough for the heat flow-rate to be controlled. Moreover, at low conversion the
viscosity is relatively low, with the advantage of a better mixing. Part of the
reaction heat is removed using a cooling agent, part remains in the system and it is
used for heating the raw material to the reaction temperature. The reactor is

partially autothermal (hence has a lower specific energy consumption) but the
separation and recycling of the residual monomer increases both the investment
and the operation costs.
The method is usually used for monomers that can be easy separated (gaseous
monomers) and it applies mostly for precipitant systems (see below).
2. Two stages polymerization
The method consists of two separate stages of the polymerization. In the first
one, that occurs at low conversion (bellow the autoacceleration limit, that may
range between 10 and 40% depending on the nature of the monomer) the reaction
is performed in classical stirred reactors, with a cooling shell or coil. The viscous
monomer-polymer solution (sometimes called prepolymer, or polymer syrup due
to its viscosity) is then poured in rectangular or cylindrical shaped recipients that
have a very large specific surface and where polymerization continues until
reaching total conversion. The large specific surface ensures that an efficient
cooling may be achieved without stirring, by circulation of cooling water (or even
air) outside the shaped polymerization forms. In such processes, conversion may
reach 100% so demonomerization is not needed.
The best known examples are block polymerization of methyl-methacrylate to
obtain sheets with a 3-6 mm thickness and surfaces up to several square metres or
anionic polymerization of -caprolactam, to obtain polymer rods that are further
processes via mechanical methods, similar with metals.
3. Polymerization methods tailored for specific monomers
Bulk polymerization of styrene is a process that takes into account some specific
features of this monomer: its polymerization can be initiated by heating (purely
thermal initiation), the polymerization rate is lower than for other vinyl monomers
(due to the conjugation of the double bond with the aromatic ring in the monomers
structure) and both the monomer and the polymer are practically immiscible with
water.
Bulk polymerization of styrene is performed in continuous stirred tubular
reactors, disposed horizontally, at temperatures above the flowing range of the
polymer, to maintain the fluidity of the reaction mass. Initiation is purely thermal,
which ensures a quite low reaction rate even at high temperature. The reaction heat
is eliminated by injecting cooling water in the reactor and eliminating it as vapour.
Removal of the heat reaction as latent evaporation heat allows maintaining the
temperature of the polymer bellow the thermal destruction range. To ensure a good
mixing of the viscous reaction mass, shaped cylindrical stirrers are used. The
product thus obtained is very pure, so it had a high degree of transparency.

The bulk polymerization of vinyl chloride - which is not a homogeneous process


but a precipitant one (see bellow) but for which the above mentioned practical
problems apply as well - combines the solutions above described. The
polymerization is separated in two stages (with adapted systems of stirring and
heat removal) but the conversion is limited to 60-65% (the main reason being to
prevent branching and cross linking). The first stage, up to 10% conversion, uses
classical stirred tank reactors, while the second critical in what concerns the
viscosity, the reaction rate and the heat generation) proceeds in reactors with a
special construction, similar with the internal mixers with helical stirrers, that
allow homogenisation of the viscous mixture. The reaction heat is eliminated by
circulating cooling water (5C) in the reactor shell; the excess heat is used to heat
the reaction mixture this facilitates the demonomerization stage that occurs by
vaporization of the monomer (that is liquefied inside the polymerization reactor,
that works at high pressure).
VI.2.1.2. Precipitant Bulk Polymerization
In the homogeneous bulk polymerization, the polymer is dissolved by the
monomer to give a viscous solution. However, there are a few monomers that do
not dissolve the corresponding polymer: the best known example is the
acrylonitrile.
In this case, bulk polymerization evolves differently from the homogeneous
case. Initiation occurs in the monomer phase (that dissolves the initiator) but
after a few elementary propagation acts the length of the chain becomes large
enough for the macroradical to become insoluble; the polymer precipitates and
propagation continues in the solid phase (that may be, in fact, either glassy or
biphasic polymer). Therefore, a few moments after the beginning of the
polymerization, the system becomes heterogeneous: the continuous, liquid phase is
the unreacted monomer (with the dissolved initiator) and the dispersed phase
consists of very fine polymer particles. The growth of the chains continues on the
surface of the dispersed particles using the monomer from the liquid phase, with a
corresponding increase in the size of the particles.
Termination may happen both in the liquid phase and in the particles. In the
liquid phase possible reactions are between two chains short enough to be soluble,
between soluble chains and primary radicals or by transfer towards the precipitated
polymer particles. While possible, all these reactions have a low impact on the
process, since most of the active centres are in the dispersed phase. In the polymer
particles termination occurs by reaction between two precipitated growing chains.
However, in the precipitated phase the mobility of the growing chains is lower;
some of the active centres may become trapped inside the particles so the

termination rate diminishes with the increase in conversion (when particles are
larger than in the initial stage). This leads to a significant increase of the
polymerization rate (autoacceleration).
Note: due to the above features, the kinetic equations valid for homogeneous
polymerization are not anymore valid; several mathematical models for describing
the polymerization rate are found in the literature.
The high reaction heat coupled with autoacceleration imposes a very good
thermal control of the polymerization, since overheating can lead easily to
explosive decomposition of the reaction mass.
However, as compared to the homogeneous bulk polymerization, the viscosity
of the polymer-monomer suspension is much lower (comparable with the viscosity
of the monomer, so while there is still enough monomer to maintain the fluidity of
the reaction mass an efficient stirring can be achieved, thus allowing heat removal
by convection (using a cooling agent circulating through the shell or coil of the
reactor).
The stirring requirements limit the total conversion possible to be obtained in
practice, since if all the monomer would be consumed the reaction mass would
transform into polymer powder. Therefore, the polymerization must be stopped at
lower than the unit conversion and the unreacted monomer must be separated.
Demonomerization is however a simple process in this situation, since the
monomer can be separated by filtration. The polymer powder is then dried to
eliminate the last traces of residual monomer.
Other cases where the polymerization is precipitant is the one where
monomers are gases under normal conditions (such as ethylene and vinyl chloride).
Polymerization of gaseous monomers occurs at high pressure (to ensure a high
enough monomer concentration and thus the productivity of the plant). The
unreacted monomer is separated by decompressing the system; thus the monomer
passes in the gas state.

VI.3. Solution Polymerization


VI.3.1. Main Features of Solution Polymerization
Solution polymerization is the technique in which the monomer and the
initiator are dissolved in a suitable solvent before the reaction.
Depending on the solubility of the polymer in the monomer-solvent mixture,
there are two types of solution polymerizations: the process in homogeneous
solution, where the polymer is soluble in the reaction medium and the precipitant

solution polymerization (also called solution-suspension technique) in which the


insoluble polymer precipitates as a separate phase.

VI.3.2. Homogeneous Solution Polymerization


In this process, the monomer, initiator and polymer are soluble in the solvent.
If the reaction proceeds at total conversion, at the end the product will be a
(concentrated) solution of the polymer.
The presence of the solvent has a dilution effect on the monomer and initiator,
which leads to lower polymerization rates as compared to the homogeneous bulk
process.
The viscosity of the reaction mass also decreases with the amount of the
solvent. It still remains orders of magnitude higher than the one of a low-molecular
compound and it increases with the polymer concentration (thus with the
conversion) this phenomenon usually requires to split the process in two or
several separate stages, each performed in a reactor provided with a different
system of mixing (more powerful stirrers for the final stages).
However, the dilution by the solvent increases the mobility of the
macroradicals, therefore the auto-acceleration effect attenuates with the amount of
solvent added and, for a concentration of the solvent about a certain limit (that
depends on the polymers nature and its molecular mass) it disappears completely.
In these cases, the characteristic S-shaped conversion-time curve for
autoaccelerated systems is replaced by a platformed one (see figure VI.3.). The
reaction rate diminishes gradually in time, due to the consumption of the monomer.
conversion

temps

time

Figure VI.3.
Conversion-time curve for homogenous solution polymerization.

10

Technically there is no limitation of the conversion that can be reached in


solution polymerization. However, some cases may require lower than the unit
conversions, due to the particular features of the chemical process (excessive
branching or cross-linking due to chain transfer) in all these situations, a
demonomerization stage is needed after polymerization.
Thermal control of the reaction in solution polymerization is easier as
compared to the bulk technique, due to several factors: lack of autoacceleration,
better mixing of a lower viscosity reaction mass and especially the possibility to
eliminate the reaction heat as latent heat by solvents vaporization. In fact, the
solution polymerizations almost always occur with vaporization of the solvent (or
of the azeotropic mixture monomer-solvent); the vapour is passed through a heat
exchanger to be condensed and then recycled in the reactor. This has the advantage
that the cooling surface is not limited to the shell/coil area of the reactor but
imposes an upper limit to the reaction temperature (boiling point of the solvent
or of the monomer/solvent azeotropic mixture).
The presence of the solvent introduces some supplementary factors that may
affect the reaction and the product, depending on the nature of the monomer, the
solvent and the active centre. The solvent may participate in transfer reaction (for
free-radical processes), with several consequences:
If the radical obtained by transfer towards the solvent is able to reinitiate
polymerization, then the conversion may reach 100% but the average molecular
mass of the polymer will decrease, proportional with the intensity of the chain
transfer (and thus with the solvent concentration). In the polymerization of
monomers that produce very active radicals in chain transfer, a compromise has to
be reached between the need for dilution and the lowering of the polymerization
degree. Usually, this problem may be solved by gradually adding the solvent
(reaction starts in concentrated solution and subsequently, more solvent is added in
the following stages, corresponding to higher conversions.
If the radical resulted by transfer is more stable than the monomeric active centre,
the degradative chain transfer may limit conversion, together with a severe
reduction of the polymerization degree. These factors narrow the choice in what
concerns solvent as compared to the reactions in organic chemistry (not involving
polyreactions).
In ionic polymerization the dielectric constant of the solvent strongly influences
the reaction rate, through is solvation capacity (thus determining whether the
polymerization occurs via free ions or ionic pairs).
Solvent selection
The ideal solvent for a homogeneous solution polymerization should dissolve
the monomer, the initiator and the polymer, be inert in chain transfer reactions, and

11

have a boiling point in the range of temperatures at which the polymerization rate
is considered acceptable. High values of the vaporization latent heat and specific
heat will ensure a better thermal control. The viscosity of the solution should be the
minimum possible for a given polymer concentration (note that good solvents for
the polymer produce more viscous solutions due to a more advanced
disentanglement of the molecular coil as compared to the poor solvents).
Other characteristics for an acceptable solvent are the low cost, availability,
absence of toxicity and flammability, low corrosion and facility of purification.
The ideal solvent, according to the above criteria, is water; however, few polymers
are hydrosoluble. Most organic solvents are, toxic and flammable, so the plants
should be provided with efficient ventilation and the equipment should be sealed to
prevent solvent loss in the atmosphere.
Polymer separation
The final product in a solution polymerization is a concentrated polymer
solution that may contain a variable amount of unreacted monomer. Separation of
the residual monomer is done by distillation or stripping. Separation of the polymer
from the solution is more difficult and can theoretically be achieved using two
different approaches:
Solvent evaporation
If the solvent has a high volatility, it can be eliminated by vaporization, either
under vacuum or using a flow of heated gas (usually nitrogen singe the mixtures of
air and organic solvents have a high explosion risk). The solvent vapour should
then be condensed, and the solvent will be recycled in the process after
purification. The process has some major drawbacks: the need for perfectly sealed
equipment (due to the toxicity of most of the solvents), explosion risks, and the
difficulty of an advanced separation. The last is the most important problem, since
the viscosity increases during evaporation. To reach a high degree of separation,
the last stages have to occur in thin-layer equipment, that is expensive and
supposes long operational times or high surfaces.
The process is used mostly when concentrated solutions of polymers are used
for further processing: solution spinning of the synthetic fibres or manufacture of
paints and adhesives that dry after application on the support. In both cases, the
specific surface is large enough to allow the total removal of the solvent in an
acceptable time range. Note also that if drying is made in atmospheric conditions
(as for paints) the solvent is completely lost by evaporation, with a cost increase
together with an environment and health risk.
Polymer precipitation

12

The method supposes separation of the polymer by adding a non-solvent to


the polymer-solvent mixture (the non-solvent has to be miscible with the solvent).
It is used mainly for situations where the solvent has a high boiling point. In order
to ensure a complete precipitation of the polymer, the volume of the non-solvent
added to the solution must be 5 to 10 times larger than the one of the solvent. The
polymer precipitates as a fine powder and it can be separated by filtration or
centrifugation; afterwards it is dried under a current of heated inert gas and
granulated (to reduce the manipulation losses).
The liquid phase is a mixture between the solvent and the non-solvent that
must be separated usually by column rectification and recycled in the process.
The method has several disadvantages: using large volumes of solvent and
non-solvent, the need for rectification of the liquid phase and the risks associated
with working with organic solvents.

a.
b.

c.

The problems listed above, linked to the separation of the polymer from the
solvent at the end of the solution polymerization make this method extremely
inefficient form the point of view of the costs involved. Therefore, the use of the
method in industrial plants is limited to some particular cases:
Polymers that do not melt (that have the destruction temperature below the melting
point or the flowing range). For these, the bulk methods cannot be used, since the
last stages suppose a molten polymer.
Cases where the solution is used as such.
There are some applications where polymer solutions are the final
(commercial products) or represent the feed for another processing line. Some
examples are:
Paints some paints are polymer solutions in an organic solvent. The solvent
evaporates after application on the support. Some adhesives are used in a similar
manner.
Polymer analogous reactions. These are reactions in which the polymer chain is not
affected but the pending groups are transformed, using suitable reagents. All
reactions on polymers must be performed in solution since the solvation of the
macromolecular coils allows penetration of the reagents to all the pending groups.
The best known example is the polymerization of vinyl acetate in methanol
solution, that is used for polyvinyl alcohol synthesis.
Solution spinning the method can use both evaporation of the solvent (dry
spinning) and precipitation of the polymer (wet spinning) in this case, both
evaporation and precipitation are facilitated by the high specific surface of the
filaments.

VI.3.3. Precipitant Solution Polymerization

13

Precipitant solution polymerization (also called solution-suspension


polymerization) starts from a solution of the monomer and initiator in the solvent,
but the polymer is not soluble in this mixture so it precipitates as a fine powder
after the first moments of polymerization (initiation and the first elementary
growth acts occur in solution but after reaching a certain length, the chains
precipitate and the subsequent growth proceeds on the surface of the particles).
The technique is similar to the bulk precipitant polymerization; the difference
is that the reaction can reach 100% conversion since the presence of the solvent
will ensure the fluidity of the reaction mass even when all monomer is transformed
into polymer.
Note that separation of the polymer as a different phase implies a significant
reduction of the macroradicals mobility, the phenomenon that is the basis of
autoacceleration. However, the reaction heat can be removed efficiently by
evaporation of the solvent (as evaporation latent heat). The solvent is then
condensed outside the reactor (with no geometrical limitation of the heat transfer
area) and recycled in the reactor; this allows a good thermal control of the reaction.
At the end of the process the reaction mass is a suspension of the polymer in
the solvent; the mixture is easy to separate by filtration or centrifugation followed
by drying of the polymer particles in a current of heated inert gas. Note also that
since the polymer is not soluble in the solvent, drying will be facilitated by the
presence of the solvent only on the surface of the particles and not in their bulk.
If, for reasons linked to the chain transfer phenomena, the reaction is limited
to a lower than the unit conversion, then the unreacted monomer has to be removed
before polymer separation; this is made either by distillation or by steam stripping.
The precipitant solution polymerization is cost effective and it uses relatively
simple installations, therefore it is preferred if there are solvents that do not
dissolve the polymer. The main practical application is the synthesis of the
acrylonitrile copolymers in aqueous solution.

VI.4. Suspension Polymerization


VI.4.1. Main Features of Suspension Polymerization
Suspension polymerization is a heterogeneous technique in which the
monomer is dispersed in a continuous liquid phase. For a suspension
polymerization, the condition is that the monomer must be immiscible with the
dispersion medium (or to have a negligible solubility).
Usually the dispersion medium (the continuous phase) is water, since most
monomers are organic substances insoluble in water. In a typical suspension

14

polymerization, the discontinuous phase (the monomer) is also called organic


phase while the continuous dispersion medium (water) is designated as aqueous
phase.
There are, however, some monomers with polar molecules that have a nonnegligible water solubility. In these cases, the dispersion medium would be a nonpolar organic solvent (linear or cyclic hydrocarbons); such systems are called
inverse suspension systems. Except for the composition of the two phases, the
polymerization occurs according to the same laws, and there is not any difference
between a polymerization in aqueous suspension and a polymerization in an
organic dispersion medium. Therefore, the following presents polymerization in
aqueous suspension that represents the majority of the practical industrial and
laboratory applications.
The initiator soluble in the monomer must be dissolved in the monomer
phase previous to its dispersion (preparation of the suspension). The organic phase
may contain also other components soluble in the monomer, such as chain transfer
agents or additives (plasticizers, lubricants, stabilizers) that are inert in what
concerns polymerization, do not decompose at the reaction temperature and reduce
the energy used for mixing in the processing stage.
After reaching the polymerization temperature, the reaction proceeds in the
monomer droplets similarly to the bulk polymerization (homogeneous or
precipitant). In fact, each monomer droplet may be considered a mini-reactor for
bulk polymerization, which gave to the technique the alternative name of microbulk polymerization. The advantage of the suspension technique is that the
droplets have a high specific surface, therefore they can be efficiently cooled by
contact with the aqueous phase. Due to this feature, suspension polymerization
may be considered a bulk polymerization cooled with water.
The reaction is usually carried out to total conversion, and the final product
consists of polymer particles (called beads) that are suspended in the continuous
phase.
Since the beginning of the process (dispersion of the monomer in the aqueous
phase) until the end, the reaction mass must be maintained under intense stirring, to
ensure the dispersion of the organic phase in the aqueous one. In absence of
mixing, the monomer droplets will form a continuous layer and the polymerization
will continue in bulk, with a high risk of losing the thermal control of the reaction.
This imposes the use of suitable stirrers (helical or impeller-type), to reach a
turbulent regime of mixing and to prevent any accidental loss of power for the
stirrer (usually the stirrers for the suspension reactors have an alternate power
source power generator in case of accidental loss of the electrical power form
the grid).

15

Even under intense mixing, there is a strong tendency towards coalescence of


the monomer droplets (the driving force is the lower superficial tension of larger
drops, that have a lower specific surface). This tendency increases when part of the
monomer is transformed into polymer (if the polymerization inside the droplet is
homogeneous) due to the corresponding raise of the viscosity. At a specific
conversion the monomer-polymer particles become sticky, so to prevent
agglomeration of the droplets, suspension (or protective) agents are added to the
polymerization system. Suspension agents are substances soluble in the continuous
phase that tend to cover the surface of the droplets forming a protective layer that
reduces the tendency of the droplets towards agglomeration (prevent coalescence).
Another effect of the suspension agents (especially when they have a
macromolecular structure) is to increase the viscosity of the aqueous phase, thus
reducing the frequency of the collisions between the droplets.
Substances used as suspension agents are:
Natural macromolecular compounds: starch, gelatine, etc.
Synthetic or artificial water-soluble polymers: poly(vinyl alcohol),
polyethylenoxide and its derivatives, copolymer vinyl acetate maleic anhydride,
water-soluble cellulose derivatives (such as caroxymethylcellulose), etc.
Hydrophilic inorganic powders such as: Mg(OH)2, BaSO4, CaCl2, Na3PO4, etc.
The suspension agents do not play any role in the chemistry of the process
(they are chemically inert) so they can be separated at the end of the
polymerization by washing the beads with a suitable detergent (emulsifier). Some
traces of the suspension agents may be found however in the final product,
affecting its properties (mainly its transparency, but also in a lower degree the
processing parameters and the mechanical properties).
The mechanism of the polymerization is a typical chain mechanism, that
evolves inside the droplets similar with any bulk polymerization, with all the
features previously described. However, the presence of the water as a dispersion
medium leads to several practical consequences:
Viscosity of the reaction mass
The reaction mass is heterogeneous and its overall viscosity (s) depends on two
parameters: the viscosity of the continuous phase (water in which the suspension
agent is dissolved), l, and the volume concentration of the suspension, :
s l 1 2.5
0.3
(6.1)

16

s 0.59

l
0.77 2

0.3

(6.2)
Therefore, the viscosity of the system will be close to the water viscosity.
Even if the dissolved suspension agent is a polymer, its concentration (of the order
of magnitude of 0.5-1.5%) is low enough not to register a significant increase of
the viscosity, that remains close to the values typical for low molecular
compounds, orders of magnitude lower than the one registered in homogeneous
bulk or solution polymerization. Moreover, the viscosity remains constant during
all the reaction (opposite to the aforementioned techniques, in which it increases
with the concentration of the polymer, hence with the conversion).
A first consequence of this feature is that the polymerization can be carried
out from 0 to 100% conversion within the same reactor, without the need to adapt
the stirrer to the raise of the viscosity.
Another consequence is the intensive mixing is possible to be achieved, with
turbulent flow. This fact has direct consequences on the intensity of the heat
transfer phenomena. Typical rotational speeds for the helical or impeller type
stirrers used for suspension polymerization are between 3 and 5 rpm.
Specific heat and thermal conductivity
Water has high specific heat (double than the one of the organic phase) and
thermal conductivity (around 3 times higher than the polymer); the values for the
overall reaction mass will be computed as averages, depending also on the
concentration of the suspension (that usually ranges between 30 and 50%). As a
result, heat transfer will be facilitated, allowing the removal of the reaction heat by
circulating a cooling agent in the shell or coil of the reactor.
Particle size and specific surface
The average size of the droplets in the suspension polymerization ranges
usually between 0.1 and 2 mm; the droplets are not equal in size, there is a
distribution of the diameters that follows a Gaussian curve, with values between
100 nm and 5 mm. The polymer beads have a lower volume due to the 15-20%
shrinkage typical for any polymerization (polymer densities are lower than the one
for the corresponding monomers). This leads to a very high specific surface.
Suppose 1 m3 of monomer, as a single sphere; this means that the radius will be
0.43 m (according to the relation V=4R3) and the corresponding surface
(according to the relation S=4R2) of 2.32 m2. If this volume of monomer is
divided in droplets with the diameter of 1 mm, then the volume of a droplet will be
1.5710-9 m3 and there will be 6.4108 droplets. Since the surface of a droplet is
3.1410-6 m2 the total surface will be of about 2000 m2, an increase of 3 orders of
magnitude (860 times). This means that the reaction heat, generated inside the

17

droplets, will be easily transferred to the aqueous phase even if the process is
always autoaccelerated. From the water, due to the high thermal conductivity and
specific heat and to the turbulent regime of mixing the heat can be transferred to
the cooling agent in the reactors jacket.
The above features make the suspension polymerization very efficient in what
concerns the thermal control; this allows the reaction to be carried out in relatively
large reactors, even giant ones for some monomers, like styrene, where the volume
of a batch can reach 100 m3 of suspension.
Note: usually the volume of the chemical reactors is limited by the ratio
surface/volume, since the reaction mass must be heated/cooled through the shell.
An increase in the specific surface of heat transfer can be achieved by replacing the
shell with an internal coil, but always there is an upper limit of the volume,
determined by the need of an efficient heat transfer.
Densities of the two phases
Usually the monomers are lighter than water while polymers are heavier (the
exception is polyethylene, that has a density slightly lower than the water).
Therefore, with the advance of the reaction and the raise of the conversion, the
density of the droplets (computed as average value between the polymer and the
monomer) increases. For each monomer there is a given critical concentration at
which the density of the monomer-polymer particles equals the one of the
dispersion medium. At this point (immersion point) the tendency of the droplets to
migrate (in absence of stirring) towards the surface of the reaction mass will be
reversed (tendency towards sedimentation). This is the critical point in what
concerns mixing, since the risk of coalescence is maximum. Excessive
agglomeration of the particle will result in a lower specific surface. The
consequence would be that the reaction heat will not be completely eliminated so
temperature will increase, leading to the raise of the reaction rate and
correspondingly of the generated heat flow. If the phenomenon continues, it may
lead to overheating of the reactor, with the accompanying risks (degradation of the
polymer).
Another consequence of the presence of agglomerated particles is a higher
erosion of the installation (especially the pallets of the centrifugal pumps) and the
risk of sedimentation inside the pipes and fittings.
Polymer separation
Polymer separation in suspension polymerization is easy to be carried out,
since the beads are not soluble in the continuous phase. A first operation is the
removal (by washing) of the suspension agent. The beads tend to sediment, so they
can be easy separated from the aqueous phase by filtration or centrifugation,
followed by drying in a current of heated gas (air or nitrogen). Note that the drying

18

agent temperature must not exceed the glass transition temperature (for the
amorphous polymer) or the melting point of the crystalline phase (for biphasic
polymers) because if the beads loose rigidity there is the risk of agglomeration
inside the dryer. The moisture must be completely eliminated (except for polymers
that absorb a certain amount of water vapour) since the presence of volatile
compounds may lead to problems in processing (bubbles or other superficial
defaults).
Polymer quality
The polymer synthesised by the suspension process has properties similar to
the ones of the product obtained in bulk: high molecular mass (limited only by
chain transfer reactions towards the monomer and the polymer) and purity.
However, due to the traces of suspension agent that may be still present on the
surface of the beads, the product has a lower transparency (it is translucent).
Conversion of the polymerization
Due to the biphasic nature of the reaction medium, the unreacted monomer is
difficult to be removed from the beads. Therefore, suspension polymerization is a
technique used for monomers that can be polymerized at total conversion (without
the risk of cross-linking). To ensure the complete consumption of the monomer
(that cannot be found in the final products in amounts higher than the order of
ppm) the usual procedure consists in raising the temperature, at the end of the
process, above the glass transition temperature of the polymer, and maintaining the
heating for 15-30 min. This increases the mobility of both macroradicals and
monomer molecules and favours the complete consumption of the monomer.
Suspension polymerization in aqueous media has many advantages, such as
high productivity, facility of thermal control, purity of the product, low cost of
polymer separation.
Installations are always discontinuous; a continuous operation would suppose
the uninterrupted flow of the suspension through the pipes and fittings, that cant
be accepted due to the risk of blocking the transport lines with sediment particles.
Usually a discontinuous plant has a lower productivity than a continuous one, but
for the suspension polymerization this limitation is balanced by the possibility of
performing the reaction in high-volume reactors. Moreover, batch polymerization
adds the advantage of flexibility: the same installation can be used for
polymerizing various monomers or using different synthesis formulations (such as
adding a volatile inert liquid to the organic phase to obtain particles that can be
expanded in a further stage, manufacturing pre-plasticized products by adding the
plasticizers in the organic phase previous to the synthesis, adding pigments for
coloured products, etc.). Note that mixing the additives with the monomer before

19

obtaining the macromolecular compound ensures a better mixing and reduces the
energy consumption in the processing stage.
The drawbacks of the suspension polymerization are relatively few: a higher
investment cost (volumes of the equipment have to consider a suspension, hence
the water also) and the fact that not all monomers can be polymerized in aqueous
solution: either due to their water solubility of because the reaction must be
stopped before reaching 100%, to prevent excessive branching or cross-linking due
to the chain transfer.
Polymerization in inverse suspension is less cost-effective, since it replaces
the water with organic solvents. These have lower specific heat, thermal
conductivity and boiling points, so the reaction has to be carried out at lower
temperature both to ensure the removal of the reaction heat and to prevent
evaporation of the solvent. This leads to a lower polymerization rate, larger batch
times and lower productivity. Moreover, using an organic solvent as dispersion
medium supposes supplementary precautions in what concerns manipulation,
ventilation, sealing of the equipment and higher costs for polymer separation
(purification and recycling of the solvent separated in the filtration stage,
condensation and recycling of the solvent evaporated during drying. Drying has to
be performed using exclusively inert gases, to prevent the risk of explosion of
mixtures air-vapour of organic solvents. Therefore, while suspension
polymerization in water is a method widely used in industrial plants,
polymerization in inverse suspension remains mostly a laboratory method.

VI.5. Emulsion Polymerization


VI.5.1. Main Features of Emulsion Polymerization
The name of the technique is given by the initial state of the reaction system,
an emulsion of the monomer in water.
The size of the initial droplets of monomer is between 1 and 10 , while the
final polymer particles, at the end of the process, are much finer (0.05-1.5 ). The
stability of the final emulsion (polymer latex) is also better than the one of the
initial reaction mixture.
The main components of an emulsion polymerization system are:
The monomer should exhibit a very low water solubility; it represents the initial
dispersed phase.
The water is the reaction medium, the continuous phase. Usually, the
monomer/water ratio is between 30/70 and 60/40 (weight).

20

The initiator that must be, in the emulsion technique, a water soluble one (opposite
to the suspension polymerization, where the initiator was soluble in the organic
phase). The most used initiators are hydrosoluble peroxide compounds or redox
systems.
The emulsifier. Emulsifiers (surfactants) are substances with molecules containing
both a hydrophilic and a lyophilic (hydrophobic) group.
Due to the hydrophilic/lyophilic character, the water solubility of emulsifiers
is very low (less than 0.25% ). The solubility limit is also called CMC critical
micelle concentration, since if this concentration is exceeded, the emulsifier
molecules are not dissolved anymore but form aggregates (of 50-100 molecules)
called micelles. In the micelles, molecules are oriented with their lyophilic groups
toward each other and with the hydrophilic ones towards the water, forming either
parallel layers or even spherical structures.
According to their chemical compositions, emulsifiers can be:
Anionic emulsifiers: metallic salts of fatty acids (soaps), salts of alkyl-arylsulphonic acids, alkyl sulphates, etc.
Cationic emulsifiers: ammonium quaternary salts.
Non-ionic emulsifiers: ethhylenoxide oligomers esterified with hydrophobic
substances (fatty acids), polyvinyl alcohol, ethoxylated alkyl-phenols, etc.
Mechanism of the emulsion polymerization
The initial reaction system (see figure VI.8) consists of monomer droplets
with a diameter between 1 and 10 . A certain (small) amount of monomer
(depending on its water solubility) is also physically dissolved in water.
If the emulsifier concentration in water is superior to the CMC, then the
emulsifier molecules will be partially soluble and partially dispersed, as follows:
An amount equal to the CMC will be physically dissolved in water; since CMC is
very low, the solubilised emulsifier molecules will be surrounded only by water
(not forming aggregates in solution).
Part of the emulsifier molecules will be disposed at the interface between the
monomer droplets and the water, with their hydrophobic parts oriented towards the
organic phase, thus forming a protective layer that prevents the coalescence of the
monomer droplets.
The rest of the emulsifier will form micelles; these micelles may contain monomer
molecules, that is thus colloidally dispersed in water. Due to the much smaller size
of the micelles as compared to the monomer droplets, the number of micelles will
be much higher than the number of monomer droplets.

21

The initiator is soluble in water. When the reaction temperature is reached,


radicals are obtained that will migrate towards the organic phase. Since there are
many more micelles than droplets, and they have also a much higher specific
surface, there is a high probability for the radicals to penetrate the micelles (and
not the monomer droplets). The addition of the primary radicals to the monomer
molecules and the propagation will thus occur inside the micelles (with the
monomer that was colloidally dispersed). As a result, the number of monomer
molecules in the micelles will diminish, until it will be below the limit of colloidal
solubility. At this point, the monomer molecules that are solubilized in water will
migrate towards the micelles and, by the mechanism of solubility, part of the
monomer in the droplets will pass in the aqueous phase.
Since the monomer is continuously consumed in propagation, this
phenomenon will continue, with an increase of the size of the micelles at the same
time with a consumption of the monomer in the droplets.
A similar equilibrium is established in what concerns the emulsifier. Since the
size of the micelles increases, more emulsifier molecules that are dissolved in
water will migrate towards the micelles, disposing themselves at the interface
between the organic phase (monomer-polymer) and the water, while the emulsifier
molecules on the surface of the monomer droplets (that gradually shrink) will be
dissolved in water and thus transported towards the surface of the micelles.
Another source for emulsifier molecules (to cover the surface of the polymer
particle) is the dismantling of the micelles that have not been penetrated by any
radicals.

Figure VI.8.

22

Topochemistry of the emulsion polymerization.


The mechanism described above therefore supposes the existence of three
different stages (see figure VI.8.):
I - Formation of the monomer-polymer particles by penetration of the radicals in
the micelles, followed by growth of particles and absorption of both monomer and
emulsifier molecules from the empty micelles that have not been penetrated by
radicals. The first stage ends when all the empty (not penetrated by radicals)
micelles have disappeared.
II - Stationary stage when the increase of the monomer-polymer particles occurs by
absorption of monomer from the droplets, intermediated by the solubility
equilibrium in water. During this stage the number of monomer-polymer particles
remains constant; this stage corresponds to conversions between 10 and 80% and
its end is marked by the complete consumption of the monomer droplets. The
constant polymerization rate is due that the process is diffusion cntrolled during
this stage.
III Final stage, where polymerization continues inside the monomer-polymer
particles until all the monomer is consumed.
conversion, %

stage I

stage II

stage III
temps

time

Figure VI.8.
Conversion-time curve typical for an emulsion polymerization.

23

Polymer separation
When the reaction is completes, a disperse system (called latex) is obtained. It
is composed of very fine polymer particles that are covered by an emulsifier layer.
The emulsifier does not only physically cover the particle, but may be also
chemically bonded (grafting reactions that occurs as a result of the chain transfer).
Therefore, the removal of the emulsifier by washing is not possible, it will remain
mixed with the polymer (and it will play a secondary role, as a plasticizer;
however, the emulsifier concentration is low enough not to affect significantly the
properties of the final product).
Due to the dual (hydrophile - hydrophobe) nature of the emulsifier, all the
polymer particles will have the same superficial charge (or polarity, for non-ionic
emulsifiers). If ionic emulsifiers have been used, the latex particles will be
surrounded by a double electrically charged layers: the ionic groups of the
emulsifier and the corresponding counter-ions.
The fact that particles have the same electrical charge (or polarity) makes the
emulsion very stable; even in absence of stirring the latex will not separate
immediately (particle will not sediment, as it happens for suspension
polymerization). The system is in fact a colloidal solution, it will separate
spontaneously but in a long time interval.
Therefore, for separating the system, at the laboratory or industrial scale, two
methods are used:
Latex coagulation consisting of neutralising the double electrical layer by adding a
strong electrolyte to the polymer emulsion, thus compensating the electrical
charges and eliminating the repulsion forces that keep the particles from
sedimenting. The coagulated latex is then filtrated or separated by centrifugation.
Note that the separation process is more difficult to be achieved as compared to the
suspension polymerization due to the much smaller size of the particles, that may
clog the filters.
Water removal (drying of the latex); this method supposes heating the latex at a
temperature high enough to allow rapid evaporation of the aqueous phase without
any thermal destruction of the polymer. The rate of evaporation may be raised by
increasing the specific surface; the process is called atomization and consists in
dispersing the latex in very fine drops (of both water and polymer particles), by
contact with a mechanical system (method used in the emulsion polymerization of
vinyl chloride).

24

Emulsion polymerization is one of the most used polymerization techniques,


due to its advantages:
Using water as a reaction medium leads to the same advantages as outlined for
suspension polymerization: low viscosity (as compared with the bulk and solution
techniques), higher specific heat and thermal conductivity as compared to organic
substances (approximately double values), easy removal of reaction heat (due to
the possibility to achieve turbulent stirring regimes and to transfer the heat from
the polymer particles to the water and from here to the cooling agent in the shell or
coil of the reactor). Moreover, the thermal control of the reactor is easier since the
reaction rate is constant for a large interval of conversion and there is no
autoacceleration peak (as for suspension or bulk polymerization).
The polymerization advances at high reaction rate even at low temperatures (using
water soluble redox initiator systems that have a low activation energy for the
radical generation reaction); this is especially convenient when monomers are
gases under normal conditions and are liquefied previously to polymerization to
increase concentration and hence the reaction rate.
The molecular mass of the polymer is high and its characteristics can be controlled
through the nature and concentration of the components of the reaction system.
If for various reasons the final conversion does not reach 100%, the unreacted
monomer can be easy removed, due to the high specific surface of the latex
particles.
There are a number of applications (paints, adhesives, impregnation of tissues, etc.)
where the latex is used as such, with no need for separation of the polymer.
The main drawback of the process is the impurification of the polymer with
the emulsifier, that is grafted onto the surface of the latex particles. For usual
processing applications this does not influence significantly the polymer properties
(due to the low concentration of the emulsifier) and may also act as an advantage
since the emulsifier acts like a plasticizer for the final polymer. However, the
presence of other substances affects transparency, polymers obtained in emulsion
are either translucent or white and cannot be used for optical applications.
Another limitation of emulsion polymerization is the fact that water cant
always be used as a continuous reaction medium. Some monomers, with water
solubility, can be polymerized using the emulsion technique only if the dispersion
medium is an organic solvent (inverted emulsion); in these cases, some of the
advantages of the water are lost (organic solvents have lower heat transfer
properties) and there are supplementary costs linked to the manipulation of organic
substances (toxicity and flammability risks, need for recycling and purification).

Você também pode gostar