Sunteți pe pagina 1din 78

PRINCIPLES OF POLYMER CHEMISTRY PART A - THEORY OF POLYMERIZATION

Academic Year 2018-19


Chapter 1: Classification of monomers and polyreactions

The term monomer indicates a molecule with low molecular weight that
Part A: Theory of Polymerization generates a polymer by chemical reaction. In general, in order for an organic
• Chapter 1: Classification of monomers and polyreactions (p.2) molecule to be also a monomer it is necessary for it to bear at least two
• Chapter 2: Step-growth polycondensations (p.10) reactive functions suitable to give covalent bonds. In this way, a linear polymer
• Chapter 3: Free-radical chain polyadditions (p.21) (thermoplastic) is generated, while if the functions are more than two, the
• Chapter 4: Ionic chain polyadditions (p.43) result will be a branched or a crosslinked polymer.
• Chapter 5: Copolymerization (p.62) The chemical reaction resulting in the synthesis of the polymer is called
• Chapter 6: Industrial polymerization processes (p.78) polyreaction. The term “polymerization” will designate the set of operations
• Chapter 7: Crosslinking (p.104) that lead to the production of polymers. These operations consist of the
polyreaction itself, and various other ancillary operations (like distillation,
filtration etc.).
Part B: Rubber Science and Technology Not all chemical reactions are eligible to become a polyreaction. In general it
• Chapter 1: Introduction to rubbers (p.115) can be said that for a chemical reaction to become polyreaction it must be
• Chapter 2: Classical thermodynamics of rubber elasticity (p.119) repetitive and cumulative (the concept of base reaction). A repetitive reaction
• Chapter 3: Molecular theory of rubber elasticity (p.125) is one in which the reaction product of a former act becomes the reagent of the
• Chapter 4: Vulcanization (p.139) latter. A cumulative reaction is one in which the product of a given reaction
• Chapter 5: Mechanical reinforcement (p.149) transfers partially or totally to the following act of reaction, and therefore leads
to an accumulation of product. In fact, these conditions occur whenever the
monomers involved are at least bivalent with respect to the base reaction.
It is necessary that the conditions of a repetitive reaction and a cumulative
reaction occur simultaneously, otherwise it is not possible to obtain a polymeric
product. In this regard, it is useful to give a few negative examples that
illustrate this point.
For example the reaction between hydrogen and chlorine is repetitive
(specifically autocatalytic) but not cumulative (in fact it does not lead to an
increase of the molecular weight of the species – HCl is always formed):

H2 → 2H*
H* + Cl2 → HCl* + Cl*
Cl* + H2 → HCl + H*
etc

while on the contrary the hydrogenation reaction of alkynes is cumulative, but


not repetitive because after two steps the reaction stops:

1 2
In this case, the reactivity of the reaction intermediate is always high or very
high, and the system is not very sensitive to catalysis. Furthermore, the
polarity of the solvent has no effect on the reaction rate, and indeed the
Both condensation and addition reactions can become polyreactions. In
dependence of the reaction kinetics on the type of solvent is an indicator of the
condensation reactions there is formation of a new A-B bond between species
type of mechanism in the polyreaction.
A and B, plus the generation of a by-product molecule ab, which is absent in
In any case, for a correct prediction of reaction pathways, the stereoelectronic
the addition reactions:
effects of the molecule must be taken into account. The following examples will
help to clarify this effect.
Condensation: Aa + Bb → A-B + ab (by-product)
A first example of a condensation reaction with ionic mechanism is the
esterification reaction between alcohol and carboxylic acid, with elimination of
Addition: A + B → A-B
water as a by-product. When this reaction occurs between bifunctional
monomers (diols and diacids), a polyesterification takes place.
Therefore, in the case of polymer synthesis reactions, one speaks of
polycondensations and polyadditions.
Every chemical reaction occurs through breakage and formation of new bonds.
Similarly, the chemical mechanism through which the polyreaction takes place
can be ionic or radical.
In the ionic mechanism there is a heterolytic breaking of the bond with the The ionic nature of the reaction is not immediately evident by looking at its
formation of an ion pair, and with the electronic pair that moves to the more stoichiometric balance. However it should be stressed that this occurs through
electronegative atom: a general mechanism of addition-elimination, with the opening of the C=O
bond by the group -OH, and the consequent formation of an intermediate C+-
A-B → A+ + B- O- ion pair. The system will therefore be sensitive to acid catalysis, able to
decrease the energy of the reaction intermediate (the ion pair) and to promote
In this case, the reactivity of the reaction intermediate will depend strongly on the opening of the carboxylic bond.
the degree of charge separation, and on the greater or lesser localization of Another example is provided by the addition reaction of a free radical to
charges themselves. In these conditions catalysts and solvents play a alkenes, through a radical mechanism.
fundamental role. The former are often acid or basic substances (in a general
sense) that help delocalize the charges of the reaction intermediate, thus
lowering the energy level and consequently decreasing the activation energy of
the reaction itself. On the other hand, the polarity of the solvents is important
to determine the degree of separation of the ion pair; polar solvents such as
In this case the intermediate bears an unpaired electron on the chain end. The
alcohols, ketones etc... will be efficient in solvation and ionic species will be
degree of substitution of the alkene and the chemical nature of the radical
separated and become accessible to reaction collisions. Solvents of low polarity
influence the bond position and formation and, ultimately, the formation of the
such as hydrocarbons will be ineffective in the solvation of reaction
polymer.
intermediates, and the ion-pair will be strongly associated and less reactive.
In the common case of addition to vinyl monomers, i.e. monomers with the
In the radical mechanism there is a homolytic breaking of the bond with the
general structure CH2=CH-X, the most substituted radical is normally formed,
formation of species with high-energy unpaired electrons
therefore the radical attack occurs on the less hindered CH2= group.

A-B → A* + B*

3 4
In most macromolecular chemistry text-books polyreactions are classified on
the basis of their chemical or kinetic mechanism, so that one speaks of M + M → M2
polycondensations and polyadditions, or step-growth polyreactions and chain- M2 + M → M3; M2 + M2 → M4
growth polyreactions. In this text we will instead refer to the nature of M3 + M → M4; M3 + M2 → M5; M3 + M3 → M6
monomers from which we will consequently deduce the principles of M4 + M → M5; M4 + M2 → M6; M4 + M3 → M5; M4 + M4 → M8
macromolecular chemistry in a more comprehensive vision. etc
The former class is represented by monomers with electronically independent
reactive functions, or “explicit” functions. These monomers have reactive Figure 1. Scheme of a step-growth polyreaction
groups explicitly present in the structure, generally separated by at least 2 or 3
atoms. Examples in this respect are diols, diamines, diacids and other According to this scheme it can be shown mathematically and experimentally
structures such as Cl-R-Cl, HS-R-HS, etc. verified that the free monomer practically never goes to zero, but after a few
The latter class is represented by monomers with electronically interdependent reaction steps it rapidly becomes so scarce as to be considered negligible (that
functions, or “implicit” functions. These are typically unsaturated or cyclic means that purification of polymeric product from residual monomer is aften
monomers; in this case an initiation or ring opening reaction is required for the unnecessary).
chemical reactivity of the monomer. The base chemical mechanism of these polyreactions is normally ionic and only
in a few cases radicalic. The base reaction can be condensation (more
commonly) or addition and the kinetic mechanism is usually step-growth, as
the reactions are typically slow and involve the formation of stable and
identifiable intermediates after each reaction step.
Examples of polymers obtained from step-growth polyreactions are polyesters,
polyamides, polycarbonates and polyurethanes. Overall, they represent many
technopolymers of large industrial interest. The first three examples come from
On the basis of this description it is thus possible to create a schematisation,
polycondensation reactions, all involving a common chemical mechanism of
saying that monomers with explicit functionalities, or electronically
addition-elimination on the carboxylic derivative. The nucleophilic group may
independent functions, generate individual reaction acts, with the formation of
be an alcohol -OH (for polyesters), an amine –NH2 (for polyamides), a phenol -
stable and identifiable intermediates. The kinetic mechanism is thus of step-
Ar-OH (for polycarbonates), while the electrophilic group is given by the
growth type.
carboxylic acid -COOH or by its derivatives such as acylic chlorides -COCl and
On the contrary, cyclic monomers bearing implicit functionalities generate
esters -COOR.
interconnected reaction acts through the propagation of activated
intermediates. The predominant kinetic mechanism is that of a chain-growth
reaction, but it can also be step-growth in a few cases.
The two different types of monomers not only generate polymers which differ
in their chemical structure, but they also follow different growth statistics that
Polyesters – e.g. Poly Ethylene Terephthalate (PET) via transesterification between diethylene
can be modelled mathematically through specific and distinct relationships.
glycol (DEG) and dimethyl terephthalate (DMT).
The polyreaction scheme defined as "step-growth" originates from monomers
that are at least bifunctional and bear explicit reactive functions. According to
this scheme each species is able to react with any other. For example, two
monomers M react to yield a dimer M2, the dimer reacts with another
monomer or with another dimer to give a trimer M3 or a tetramer M4 and so on, Polyamides – e.g. Polyamide 66 via reaction of polyamidation between hexamethylene diamine
as shown in Figure 1. and adipic acid.

5 6
As will be described below, in chain-growth polyreactions the molecular weight
of the polymer can also grow significantly although often accompanied by the
presence of significant amounts of free and unreacted monomer, that means
that purification of polymeric products after polymerization is needed.
Polycarbonates – e.g. Polycarbonates PC from bisphenol A and phosgene. Examples of chain-growth polyreactions are the polymerization of ethylene to
yield polyethylene, propylene to yield polypropylene, styrene to yield
Polyurethanes are instead produced through an addition polyreaction between polystyrene, of vinyl chloride to yield polyvinyl chloride, of methyl
monomers with -OH hydroxyls and isocyanates –N=C=O, to yield a urethane methacrylate to yield polymethylmethacrylate and other cases reported in
bond -OCONH-: Figure 3.

No byproducts are formed.


On the contrary, the "chain-growth" polyreaction scheme involves monomers
with interdependent or implicit functionalities, and requires the action of an
external auxiliary substance, normally called initiator, which explicates the
chemical functions and starts the polyreaction.
The growth pattern of the polymer is substantially different from the previous
case, and in this case the polyreaction is a succession of chain reactions that
take place through the propagation of an active center M* capable of reacting
with the monomer M only (Fig.2) .

I + M → IM*
IM* + M → IM2*
IM2* + M → IM3*
IM3* + M → IM4*
IM4* + M → IM5*
etc.
Figure 2. Scheme of a chain-growth polyreaction.

In this type of polyreaction the chemical mechanism may be both of radical


type as well as ionic type. The base reaction is almost always addition, and the
kinetic mechanism is chain-growth type.
Figure 3. Examples of some chain-growth polymers.

7 8
Finally, it is essential to highlight how the mere presence of suitable monomers substantial cessation of the polymerization process, which is exothermic but
is not normally sufficient to guarantee the formation of the corresponding high also exoentropic (∆S <0), being each polymerization an inherently orderly
molecular weight polymer. There are, in fact, general chemical, process. Since ∆G = ∆H – T ∆S where both enthalpic and entropic terms are
thermodynamic and kinetic aspects of the process that need to be satisfied. negative, there will be a threshold temperature (ceiling temperature or TL)
A first condition is obviously related to chemistry and stoichiometry: it is above which the polymerization will be thermodynamically unfavoured (∆G >
necessary to select monomers with appropriate chemical reactivities (e.g. 0). Moreover, in these conditions the opposite reaction - depolymerization -
alcohol + acid etc.) and, when necessary, to provide them in appropriate will be favoured to ultimately yield monomer. More details will be explained in
stoichiometric ratios to increase the molecular weight. Chapter 3.
A second condition is related to the thermodynamic aspects of the process, and
here it is useful to distinguish between polycondensations and polyadditions.
Typically in a polycondensation, such as a polyesterification, some σ bonds of Chapter 2: Step-growth polyreactions
comparable strength break and form. Therefore, the polyreaction is mildly
exothermic or athermal from an energy balance point of view . As a result, Step growth polyreactions are obtained from monomers with electronically
from a thermodynamic point of view, low temperatures are more favourable as independent or explicit functionalities, and mainly concern polycondensations
they would allow to achieve equilibrium conditions more shifted towards the with few but important cases of polyadditions.
products. However this condition does not match the kinetic requirements of a Let us consider now in detail the case of bifunctional polycondensations, i.e. of
step-growth process, which is typically slow and therefore needs high polyreactions obtainable from bifunctional monomers to yield linear polymers.
temperatures and the use of catalysts to achieve polymerization in times that The first subcase, which is also the simplest case although not very frequent, is
are convenient in a technological perspective (hours). The problem is usually represented by the polycondensation of type aAb heterobifunctional monomers.
solved by favouring the kinetic aspect and actually carrying out the This polyreaction can be schematised as follows:
polycondensation in conditions close to thermodynamic equilibrium, but never
quite at that point as the by-product is constantly subtracted from the reaction aAmb + aAnb ↔ aA(m+n)b + ab
equilibrium through distillation, extraction and other techniques.
By expressing the equilibrium constant K as a function of the concentrations of The subscripts m and n can take any integer value, as in step-growth reactions
reagents and products such as: all species (monomers, oligomers) can react with each other.
A good example is given by the polycondensation of naturally occurring
Keq = [-AB-][ab]/[-Aa][-Bb] = C e(-∆H/RT) α−ω hydroxy acids to give polyesters as shown below:

the gradual decrease of [ab] by removal of the by-product steers the system n HO-R-COOH → H-[O-R-CO]n-OH + (n-1) H2O
towards the product according to the Le Chatelier’s principle (if a system at
equilibrium is perturbed it will react to contrast that perturbation). Thus, if the Conversely, the most important and widespread case in industry is that of
by-product is removed the reaction will tend to increase the concentration of polycondensation of two aAa + bBb type monomers. The oligomeric species
product to compensate for the loss. present in various moments of the polyreaction will be of the type a(ABA)na,
On the contrary, the thermodynamic condition for chain polyadditions is very a(AB)mb, b(BAB)rb. An example in this sense is the already seen case of
different. In this case each occurring reaction leads to the disappearance of a π polyethylene terephthalate (PET).
bond and to the formation of a much more stable σ bond. The process is Starting from the general case with two monomers of the aAa + bBb type, the
therefore strongly exothermic and kinetically rapid since it is a chain process stoichiometric ratio r is defined as the ratio of the initial number of reactive
and therefore is favoured by low temperatures, with no problems on the functions of type A and B
overall reaction rate. On the contrary, high temperatures can lead to a

9 10
r = Na0/Nb0≤ 1

The stoichiometric ratio is less than or equal to 1 by definition, i.e. the limiting
stoichiometric agent is always at the numerator.
The extent of reaction p of the reaction is then defined as the fraction of
reacted terminals at the generic time t, i.e.

p = (Na0-Na)/Na0
From the expression of Xn it is possible to immediately derive the value of the
Therefore the extent of reaction varies from zero (onset of reaction, Na=1) to 1 number average molecular weight Mn being Mn = Xn M0 (ignoring the
(all terminals consumed, i.e. Na=0). contribution of the terminals, acceptable for sufficiently high molecular
Through simple algebra, Na and Nb can be expressed as a function of the weights).
stoichiometric ratio and of the extent of reaction, with the hypothesis that the The result shown above is important because it shows that for step-growth
fraction of type “b” terminals at time t is given by the initial quantity Nb0 polyreactions, regardless of their chemical mechanism, it is possible to derive
reduced by the fraction of reacted type “a” terminals, i.e. Na0 p (since each b the number average molecular weight of the obtained polymer simply as a
reacts with an a) function of the stoichiometric ratio r between the reagents and of the extent of
reaction p. Process parameters such as temperature, pressure, concentration
Na = Na0-Na0 p = Na0 (1-p) etc. do not enter explicitly in this equation. At most, they influence the
Nb = Nb0 – Na0 p = Na0/r –Na0 p = Na0 (1/r – p) dependence of p on time.
It is also interesting to evaluate the significance of the above equation in a few
Having thus defined Na and Nb, it is possible to obtain a general expression particular limiting conditions. For example, if r=1 is assumed (the polyreaction
that correlates the average degree of polymerization Xn to r and p. This is perfectly balanced from the stoichiometry point of view), the following
relation is valid for all step-growth polyreactions, both polycondensations and applies:
polyadditions.
Assuming that the amount of residual monomer is negligible (a valid Xn = 1 / (1-p)
assumption for sufficiently high values of p), Xn can be defined as the number
of repeating units per molecule, i.e. as the ratio between the number of initial Therefore Xn (and the molecular weight) will tend to infinity for degrees of
monomer molecules (for the hypothesis written above) and the number of conversion equal 1.
polymer molecules formed at a given time t. The number of molecules is given On the other hand, if p=1, i.e. in conditions of complete consumption of
by the half-sum of the number of terminals for linear molecules, i.e. bearing reactive terminals, the equation is reduced to
two terminations.
Xn = (1+r)/(1-r)
Xn = [(Na0 + Nb0)/2] / [(Na + Nb)]/2 = (Na0 + Nb0)/(Na + Nb)
so the molecular weight will tend to infinity for values of r approaching one.
After appropriate substitution on the basis of the previous relations, the The graph in Figure 4 illustrates the general dependence of the maximum
following is obtained obtainable average degree of polymerization on the stoichiometric ratio
between the reagents, here expressed as the mole fraction of monomer B
(molar fraction of 0.5 is equivalent to r=1)

11 12
0.99); there be no monofunctional species. For example, one can easily
calculate that with r=1 if p=0.9 then Xn = 10, if p=0.99 Xn=100 while if
p=0.999 then Xn = 1000.
The production of polycondensation technopolymers at industrial-scale
normally requires to obtain the highest possible molecular weights, but
simultaneously fulfilling the three above mentioned requirements is quite
difficult. Each process has therefore developed specific strategies to overcome
these obstacles (as illustrated in some examples in Chapter 6) yet it is worth
mentioning that in polymers from polycondensation Mn does not normally
exceed 30,000-40,000.

It has already been mentioned that from a thermodynamic point of view


polycondensations are mildly exothermic polyreactions, favoured as such by
low temperatures. However, being kinetically step-growth processes with slow
reaction steps, it is usually necessary to keep the temperature high in order to
obtain acceptable reaction times. Furthermore, since often such polyreactions
are carried out in bulk, it is necessary to heat also to decrease the high
viscosity of the polymer melt that grows exponentially with the increase of
Figure 4. Dependence of the average degree of polymerization on the stoichiometric ratio
molecular weight. Under these process conditions Keq values are typically very
between the reagents
small, thus it is not possible to obtain high polyreaction conversions (high
extent of reaction) if the system is allowed to reach thermodynamic
equilibrium, that is, if one works in a closed system.
Another parameter that is able to influence the maximum obtainable molecular
This can be demonstrated by considering polycondensation as an equilibrium
weight in step-growth polyreactions is the purity of the reagents, in terms of
polyreaction. In this sense, for a generic condensation reaction of the type Aa
degree of monomer bifunctionality. This is the case of polymerization of a
+ Bb ↔ AB + ab, the equilibrium constant can be expressed as
system aAa + bBb + bB. It is in fact intuitive that the presence of
monofunctional molecules, i.e. bearing a single reactive chemical function,
Keq = [AB] [ab] / [Aa] [Bb]
works as a chain stopper thus reducing the maximum obtainable molecular
weight. In this case - defining with Nb0m the number of monofunctional groups
The degree of conversion and the stoichiometric ratio can be written as a
- it is necessary to redefine the stoichiometric ratio r as
function of the concentrations of the species involved

r = Na0 / (Nb0 + 2Nb0m)


p = (Na0 – Na) / Na0 = 1 – Na/Na0 = 1 – [Aa]/[Aa]0
r = [Aa]0/[Bb]0
to make the previous equation still applicable. The coefficient 2 in front of Nb0 m

is required because a bB molecule shows the same quantitative effect of an


Through simple algebrical steps it is also possible to rewrite the expressions for
excess of bBb molecules in limiting the growth of the polymer.
current concentrations of reagents, products and by-products as a function of
This fact implies that in order to obtain high molecular weights in bifunctional
measurable quantities of the polyreaction, namely initial concentrations of
polycondensations - and in general in all step-growth polyreactions – it is
reagents, p and r:
necessary that: the stoichiometric ratio between the reagents be as close as
possible to one; the degree of conversion be also very high (usually well above
13 14
[AB] = p [Aa]0 = [ab] individual reactions have equal rate, i.e. the reaction rate is independent of the
[Aa] = [Aa]0 (1-p) molecular weight and a single rate constant k can be considered.
[Bb] = [Aa]0 (1/r – p) According to the classical kinetic theory in the case of a reaction between aAa
and bBb, the reaction rate is given by the rate of disappearance of the reacting
The concentration of Bb was calculated as [Bb]0 (i.e. [Aa]0 1/r = [Aa]0 groups
[Bb]0/[Aa]0) reduced by the concentration of unreacted Aa, i.e. [Aa]0 p, since
each “b” reacts with an “a”. These relations can be inserted into the expression v = -d[a]/dt = k [a][b]
of the equilibrium constant, thus obtaining a quadratic equation that outlines In the frequent case of r = 1, the above equation can be simplified by setting
Keq as a function of p and r. [a]=[b]=c

p 2 ⋅ [Aa]0 r ⋅ p2
2 v = -dc/dt = kc2
[AB][ab]
Keq = = =
[Aa] [Bb]
(1− p) ⋅ [Aa]0 ⋅  1 − p  ⋅ [Aa]0 (1− p) ⋅ (1− r ⋅ p) The above relationship is a typical case of second order kinetic model, and is
r  followed for example by catalyzed polyesterification reactions as shown in the
example in Figure 5. In this case, the catalyst is normally a protic acid such as
Since r is a constant for a given polyreaction, it is interesting to see which H2SO4 or a Lewis acid (acceptor of electronic pairs) such as a transition metal
values of extent of reaction (and therefore of molecular weight) can be salt.
reached by the system at equilibrium conditions. In particular, setting r=1 it is By setting [COOH] = [OH] = c, then
easy to calculate that if Keq = 10 then peq = 0.759, while it is necessary to
have Keq = 100 to obtain p = 0.909. Since normally in polycondensation v = -dc/dt = k f(cat) c2 = koss c2
processes Keq at processing temperatures is on the order of 100, it follows that
it is not possible to obtain high degrees of conversion for the polymerization if where f(cat) is a kinetic expression depending on the type and concentration of
the system is closed. The problem is in fact solved by carrying out the catalyst, and that changes - as such - its analytical expression depending on
polymerization typically in non-equilibrium thermodynamic conditions, for the chemical mechanism of the catalytic process. By applying the second order
example by distilling progressively the by-product as this is formed (if it is model to the previous relationship, the observed kinetic constant koss will
sufficiently volatile). actually be the product of all elementary k’s and of the function of the catalyst.
By separating the variables and integrating the simple differential equation
Now let us derive the expressions of kinetic equations for polycondensations. between the limits of concentration c0-c and time 0-t, it is possible to obtain
Considering the reaction scheme, a polycondensation can be represented as a
set of reactions such as those shown in the following scheme. -dc/c2 = koss dt
-∫1/c2 dc = koss ∫dt = 1/c – 1/c0 = koss t

However, from the above-written relationships, by multiplying the numerator


and the denominator by the reaction volume V (c = N/V), the following
expression can be written
Theoretically the overall polyreaction rate depends on the rate of the individual
reaction steps, and the kinetic scheme is therefore very complex. However for N0/N = 1/(1-p) = c0/c
practical purposes the hypothesis of equal reactivity of the terminal groups
applies, which is also experimentally verifiable. According to this hypothesis, all

15 16
If the integrated kinetic model of the second order equation is multiplied by c0 The examples in Figure 5 show the graphical representation of the reaction
the following relationship is obtained kinetics of the polyesterification between (a) ethylene glycol and adipic acid
uncatalyzed, and between (b) decamethyleneglycol and adipic acid catalyzed
c0/c -1 = koss c0 t = 1/(1-p) -1 with p-toluene sulfonic acid. These data experimentally confirm the predictions
of the kinetic models written above.
which outlines the dependence of the conversion of the polyreaction on the
reaction time. Linearity is thus envisaged, and experimentally verified, in the
expression of 1/(1-p) as a function of t.
From an experimental point of view, in order to kinetically control the progress
of a step polymerization it is crucial to have a method that allows to quantify
the residual reactive groups at various reaction times. This can normally be
done either using chemical titration (e.g. titration with bases of the -COOH
groups in the polyester described in the above example) or using spectroscopic
methods (nuclear magnetic resonance NMR or infrared spectroscopy IR in
some cases).
Figure 5. Reaction kinetics of the polyesterification between (a) ethylene glycol and adipic acid
Going back to the example of the polyesterification, autocatalysis occurs even
uncatalyzed, and between (b) decamethyleneglycol and adipic acid catalyzed with p-toluene
in the absence of external catalysis due to the presence of the -COOH groups
sulfonic acid
of the monomer. In addition the analytical form of the kinetic equation
changes. In practice, f (cat) = [COOH] in the specific case, so that
For step-growth polymerizations the distribution of molecular weights can be
predicted on the basis of purely statistical considerations. From here on, the
v = -d[COOH]/dt = k [COOH]2[OH]
case of stoichiometric ratio r = 1 will be considered, always with the hypothesis
Assuming that r=1 and separating the variables as in the previous case, an
of an equal reactivity of the terminal groups.
analytic form of a third order kinetic equation is obtained.
Assuming to have all the N molecules of the polymer available, as if they were
arranged in a large abacus, the probability of "collecting" a molecule with a
-dc/dt = k c3
degree of polymerization x is equal to its mole fraction nx, i.e.:
-dc/c3 = k dt

nx = Nx/N
By integrating between the concentrations c0-c and in the time interval 0-t one
obtains
The probability that such a molecule is constituted by at least one monomeric
unit is 1 (since this molecule exists). Since the degree of conversion of the
1/c2 – 1/c02 = 2k t
polyreaction p is, in statistical terms, equal to the probability of reaction, then
the probability that the molecule is at least a dimer (i.e. that it has reacted
At this point, by multiplying by c02 and normalizing on the volume as in the
once) will be given by 1 p.
previous case the following expression is found
Similarly, the probability that the molecule be a trimer, (i.e. that it has reacted
twice), will be 1 p2.
1/(1-p)2 -1 = 2k c02 t
Generalizing this argument, one can define the probability that the molecule
has a length x, i.e. nx, as
Therefore, this relation predicts a linear dependence of 1/(1-p)2 on t.

nx = p(x-1) (1-p)
17 18
the analytical expression of the molecular weight distribution function at
where (1-p) represents the probability of undergoing no further reaction, equilibrium, typical of step-growth polymerization processes.
namely the fact that the molecule will not grow beyond the length x. It is possible to further develop these equations to obtain an expression of the
With similar arguments it is possible to derive the expression for the mass average degree of polymerization and of the polydispersity index.
fraction wx, neglecting the corrective effect of terminals, The average degree of polymerization is defined as

wx = Wx/W0 = (Nx Mx)/ (N0 M0) = (Nx x M0)/ (N0 M0) = (Nx x) / N0 Xn = ∑ xi ni = ∑ xi p(xi-1) (1-p) = (1-p) ∑ xi p(xi-1)

Being in the above relations As p < 1, it is a convergent series of type ∑ x px-1 = 1/(1-p)2, so the
expression of the number-average degree of polymerization is given by
Nx = nx N = nx N0 (1-p)
Xn = 1/1-p
(since normally the number of terminals or molecules at time t, i.e. N, is given
by the initial number N0 decreased by its reacted fraction, i.e. N0 p), by The same applies for the weight-average degree of polymerization
performing the appropriate substitutions of Nx in the expression of wx one
obtains Xw = ∑ xi wi = ∑ xi2 p(xi-1) (1-p)2 = (1-p)2 ∑ xi2 p(xi-1)

wx = nx x (1-p) By using the standard mathematical equation ∑x2 p(x-1) = (1+p) / (1-p)3, one
obtains
Yet as stated by above definitions
Xw = (1+p) / (1-p)
nx = p(x-1) (1-p)
So that the polydispersity index will be given by Mw/Mn = 1+p, i.e. for
by working on the last substitutions one arrives to the expression of the mass polymers obtained through step-growth processes the polydispersity index
fraction of the molecules of length x as a function of the probability of reaction approaches two when the degree of conversion tends to one. Analytically, this
p. is equivalent to saying that

wx = x p(x-1) (1-p)2 Mw/Mn → 2 for p → 1

Note how the above equation allows to estimate the fraction of unreacted This in fact occurs experimentally in the industrial production of different
monomer (w1) at each reaction time, a quantity which is thus already very low polymers obtained through polycondensation. It is interesting to note how the
at relatively small degrees of conversion. system will tend to such most probable distribution not only at the stage of
So both nx and wx, and therefore also the polydispersity index and ultimately synthesis of the polymer, i.e. when starting from monomeric molecules, but
the molecular weight distribution, depend only on x and p, and are entirely more generally each time that it will be allowed to do so, for example during
independent from classic polymerization process parameters such as the mixing at high temperature of polyesters of different molecular weight, as
temperature, pressure, concentration etc. occurs in recycling technologies.
The distribution defined by the definitions of nx and wx given above is referred Finally, the graphs presented in Figure 6 show the trends and the evolution of
to as the most probable distribution (or Schulz-Flory distribution). It represents the distribution of the number-fraction νx and the mass-fraction ωx of X-mers

19 20
(species with polymerization degree X) as a function of X, for different values As opposed to the previous case of step-growth polycondensations, chain-
of the degree of conversion p. growth polyadditions are characterized by a rather complex polyreaction
scheme which is divided into four different phases, namely initiation,
propagation, termination and chain transfer.
The following scheme (Fig. 7) presents some of the most important monomers
and industrial polymers that can be obtained from radical polyaddition
processes. Among these, it is worth mentioning polyethylene, polystyrene,
polyvinylchloride, polymethylmethacrylate, polyacrylonitrile and
polytetrafluoroethylene. As will be discussed in the next chapters, in addition
to these polymers, many other important industrial copolymers are also
obtained through this type of polymerization.

Figure 6. Distribution of the number-fraction νx and mass-fraction ωx of X-mers as a function of


X

As regards the performance of the number distribution, it is possible to


observe that the number of molecules of monomer is always higher than the
number of any other molecule for each value of p. By contrast, by analyzing
the mass distribution, the actual amount of monomer is very low, and there is
a maximum in the distribution that moves progressively to higher values as p
increases.

Chapter 3: Free radical chain polyadditions

This chapter will cover the principles of free radical polyadditions, a particular
class belonging to the large family of chain polymerizations.
Radical polyadditions are an important category of polyreactions that require
the use of monomers with electronically interdependent or implicit
functionalities (generally characterized by the presence of a double bond C=C).
The chemical mechanism is radical-type while the kinetic mechanism is
generally chain-type, i.e. characterized by the presence of very quick and
interconnected reaction acts.
Also controlled radical polyaddition processes exist (often indicated with the
acronym of ATRP, that is Atom Transfer Radical Polymerization), leading to a Figure 7. Some types of polymers obtained from free radical polymerization.
more controlled polymer microstructure and narrow molecular weight
distribution, but they will not be considered in this chapter.

21 22
As stated above, every free radical polyaddition process involves four different As will be shown in the following, t1/2 – T curves represent the most important
phases, the first one always being the initiation reaction, namely the onset of parameter for the choice of a specific peroxide as initiator in a free-radical
chain-growth reactions caused by the formation of the first free radical species. polymerization process, as they essentially determine time and temperature of
Theoretically, this initiation reaction can be a purely temperature-dependent the polymerization reaction.
process in which two monomer molecules impact against each other to form a Another class of initiators is represented by diazocompounds, having general
bi-radical: formula R-N=N-R. Among these, the most important is AIBN – azo bis
isobutyronitrile.
CH2 CH + CH2 CH R - CH• − CH2 − CH2 − CH• − R

R R

More frequently, at industrial level an external substance is added (the


initiator), which can decompose upon absorption of thermal or light energy Diazocompounds decompose thermally to yield two radicals and nitrogen N2.
thereby generating a reactive species able to attack the first monomer
molecule. In this case, the initiation reaction can be represented as follows:

I → I*
I* + M → IM* Although peroxides and diazocompounds are soluble in organic solvents, many
radical polymerization processes are actually carried out in heterogeneous
Among the most common radical initiators, it is worth mentioning organic phase using H2O as reaction medium. It is therefore useful to be able to
peroxides. These molecules are characterized by the presence of the -O-O- employ radical initiators which are soluble in water. This very important topic
bond, which is unstable upon thermal- or light-induced stresses. will be discussed in Chapter 6, where emulsion polymerizations will be
As an example, the following scheme shows the radical formation process in explained. Redox couples are often used as water soluble initiators: they
benzoyl peroxide, which leads to the decomposition of this molecule into two consist of combinations of chemical species that can generate radicals
phenylic radicals and carbon dioxide upon heating. (typically inorganic) via redox reactions, at ambient temperature or even lower.
The combination of hydrogen peroxide (H2O2) and soluble iron salts or the
persulfate-tiosulfate couple are sometimes used as water soluble initiators, as
shown in the following scheme.

H2O2 + Fe2+ → OH- + *OH + Fe3+

Several peroxides exist, differing from each other because of the type of their S2O82- + S2O32- → SO42- + *SO42- + *S2O3-
substituents, e.g. alkyl, aryl and acyl peroxides. Depending on the nature of
such substituents, solubility in different solvents can be achieved and specific Finally, a particular and very important category of initiators is that of
decomposition temperatures are needed. Each peroxide, as well as all thermal photoinitiators. These molecules are able to absorb UV-photons (as opposed to
initiators, is characterized by its half-life time t1/2, defined as the time required the monomer) and to generate radical species by homolitic bond breaking at
to yield half of the initial concentration of the peroxide. The half-life is a low temperature. In general, these photoinitiators are conjugated organic
function of temperature, and a negative linear dependence of t1/2 on T is molecules such as for instance various types of aromatic ketones, which
observed on a semi-log scale. decompose photochemically to yield reactive fragments:

23 24
R • + CH2 CH R CH2 C H• predominant
X X

R CH CH2•

CH2 C H• + CH2 CH CH2 CH CH2 C H•

X X X X

head-to-tail

Figure 8. Chain sequences in free-radical polyadditions.

Photoinitiators are industrially rather important in some polymerization


This is due to the reaction mechanism, which favors the formation of the most
processes at low temperature, such as photocrosslinking of unsaturated
stable radical, that is the most substituted. This radical will attack the less
monomers and oligomers in several painting processes as well as the ink
sterically hindered part of the monomer molecule.
technology.
Conversely, the free-radical reaction intermediate is highly reactive because of
For all cases presented above, it is worth mentioning that the general working
its high energetic level. Therefore the configurational control is poor, as the
mechanism of initiators is such that one of their fragments chemically binds to
thermal collisions are successful regardless of the direction of attack. The
the monomer and becomes one of the chain terminals of the newly formed
consequence is that radical polyreactions generally have poor configurational
macromolecule. This mechanism is therefore very different from that of
control and yield the formation of essentially non-crystallizable, atactic
classical catalysts, which do not undergo changes during the chemical process
polymers.
they are involved in.
The termination phase is characterized by the quenching of the macroradicals
The propagation phase represents the actual synthesis of the polymer, where
that are increasing in size, and the closing of the polyreaction kinetic chain.
the reaction of transformation from monomer to polymer occurs; it can be
Two distinct mechanisms are feasible, usually coexisting and competitive
represented as a set of identical reaction acts all characterized by a kinetic
depending on the process variables involved (Fig. 9). In the combination
constant Kp (that will be called K2 from here on).
mechanism, unpaired electrons coming from two different macroradicals
combine to form a new covalent bond and a single macromolecule of molecular
weight equal to the sum of the two. On the other hand, the radical
disproportionation mechanism involves abstraction of H* hydrogen by a
Since free-radical polymerizations often involve CH2=CH-R vinylic monomers, it macroradical and the simultaneous formation on the other macroradical of a
is interesting to note that the regiochemical control is generally high. Indeed, vicinal diradical which instantaneously evolves into a terminal double bond. In
the molecular arrangement is regular and the newly formed macromolecule this case two distinct macromolecules are obtained, and the molecular weight
primarily presents head-to-tail linkages (Fig. 8). does not change.

25 26
− combination
X X
• • CH2 CH CH CH2
CH2 C H + CH CH2

X X
one molecule

− disproportionation

X X

CH2 C H + CH • CH2 CH CH + CH2 CH2

X X two molecules

Figure 9. Termination mechanisms in radical polymerizations.


Figure 10. Mechanism of LDPE formation from chain transfer reaction to the polymer.

Finally, the chain transfer reaction is characterized by the transfer of the active
center from a growing macroradical to another generic molecule S, called chain The effect of chain transfer on the overall kinetics of radical polymerization
transfer agent. Similarly to the termination reaction, the macroradical is cannot a priori be generalized. In fact, it depends on the reactivity of the
quenched. However, in this case a new radical species is concurrently formed, radical S*. If the reactivity of S* is comparable to that of the macroradical then
able to generate another kinetic chain. there is control (reduction) of the molecular weight but no effect on the
reaction rate. Conversely, if the reactivity of S* is poor or nil, a termination
reaction occurs. Finally if the reactivity of S* is lower than or greater than that
Mn• + S Mn + S •
of the macroradical, there will be a retarding or accelerating effect,
respectively.
In many free-radical polymerization processes, especially emulsion processes,
In the presence of chain transfer reactions the molecular weight of the polymer
it is easy to obtain very high molecular weights that make the material
decreases.
practically impossible to process and treat from a technological and rheological
From a chemical point of view, the chain transfer is normally a reaction of
point of view (remember that viscosity increases exponentially with the
hydrogen removal by a macroradical from another molecule that may be
molecular weight). In these cases, a chain transfer agent is often intentionally
solvent, initiator, monomer or polymer already formed. In this last case, the
added to control the molecular weight. Among the most commonly used chain
new forming kinetic chain leads to branching in the polymer. The example
transfer agents are thiols or organic mercaptans, molecules of type R-SH that
shown in Figure 10 describes the radical polymerization of ethylene.
react via hydrogen abstraction according to the following scheme.
Polyethylene obtained from radical processes is called LDPE (low density
polyethylene), being indeed branched. Chain branchings can be considered as
structural defects in the macromolecule that cause a decrease of its Mn• + R − SH Mn − H + R − S•
constitutive regularity, thus hampering its tendency to crystallization. Since in
the amorphous phase the chains are less packed, the decrease of the degree of
Chain transfer agents should not be confused with inhibitors, molecules that
crystallinity is accompanied by a decrease of polymer density.
generate very stabilized - and therefore practically non-reactive - radicals. An
27 28
example is hydroquinone (Fig. 11), in which the radical can delocalize in Keq = K2/Kdp = [-Mn+1*]/([-Mn*] [M]) = 1/[M]
different positions on the oxygen or the carbon atom.
It is thus possible to express the dependence of ceiling temperature on
monomer concentration
OH O O . O

. Tl = ∆H0 / (∆S0 + R ln[M]l)


etc…
OH O OH O
The above relationship highlights that the ceiling temperature for a polymer is
not unique but depends on the monomer concentration. In particular, being
both ∆H0 and ∆S0 negative, the ceiling temperature grows with monomer
Figure 11. Hydroquinone
concentration. This means that there will always be a ceiling temperature
above which it is impossible to carry out the polymerization, also in the
Hydroquinone, as well as other inhibitors, is industrially used as a stabilizer of
presence of pure monomer. Beyond this ceiling temperature, the polymer is
monomers to prolong their shelf-life. The effect on the kinetics of
thermally unstable and degrades back to monomer. For many monomers that
polymerization is that it creates an induction time, proportional to the content
can be polymerized by a radical mechanism, the ceiling temperature is very
of inhibitor, after which the polymerization reaction re-starts at its regular rate.
high and this fact does not hinder the polymerization process. However, in
other cases this value is quite low (for example the Tl of methylstyrene is only
From a thermodynamic point of view, radical polyadditions are always
+61°C) and this results in considerable difficulties in the synthesis of the
exothermic, because a π−bond breaks and a more stable σ−bond is formed. As
corresponding polymer.
temperature increases, the following equilibrium gradually establishes:

The kinetic mechanism of radical polymerizations is of chain-growth type, i.e. it


∗ ∗ consists of a sequence of very fast reaction acts (low activation energy). The
M + M
n M n+1
fact that radical polyadditions follow a complex scheme with several possible
reactions makes it significantly difficult to obtain kinetic models.
This means that there will be a competition between propagation (with kinetic It is therefore necessary to obtain the expression that describes the rate of
constant K2) and depolymerization (Kdp), where depolymerization is the reaction through a series of simplifying assumptions, which are described and
reversion reaction of polymer into monomer. Both kinetic constants increase commented on below.
with temperature although with different sensitivity, and become equal at a The first hypothesis implies that the reactions are kinetically simple, i.e.
critical temperature value called ceiling temperature (normally indicated as Tl). reaction orders are equivalent to the stoichiometric ratios in the mass balance
The polymerization-depolymerization equilibrium constant can be expressed as equation.
The second hypothesis implies equal reactivity of the terminal groups, namely
∆G = ∆G0 + RT lnKeq the fact that the reactivity of the molecules is independent of the molecular
weight. This allows to represent the propagation phase by means of a single
where ∆G0 expresses the free energy content of the species in the standard kinetic constant K2.
state (pure liquid monomer, crystalline solid polymer). At equilibrium ∆G = 0 The third assumption is the above mentioned hypothesis of steady state. This
by definition. Therefore by applying the hypothesis of steady-state of free- implies that in several steps it has to be assumed that the concentration of
radical species (see below), the following can be written radicals, which is the unknown parameter in the kinetic equation, is small and
constant. This hypothesis is plausible when the steady-state of the process is
∆G0 = -RT lnKeq = ∆H0 – T∆S0

29 30
considered, i.e. sufficiently far from the beginning and the end of the
polyreaction. -d[I]/dt = Kd [I]
_
A final hypothesis implies a long kinetic chain ν , the latter defined as the Accordingly, it is also possible to express the rate of formation of the reactive
number of addition reactions between the generation and the quenching of the fragment I* as
macroradical. This assumption is also entirely acceptable when considering
polymers with sufficiently high molecular weight, as is the case of radical +d[I*]/dt = a ε Kd [I]
polyadditions.
The rate of disappearance of the monomer M is given by since on average a fragments (typically 2) are formed for each initiator
molecule that disappears. Note that the specific efficiency of each initiator
v = -d[M]/dt ≈ v propagation should also be considered and quantified. In fact, only a fraction of produced
radicals is capable of activating the propagation. Therefore in the
since the monomer essentially disappears because of the propagation reaction. decomposition process controlled by Kd a correction factor ε always <1 should
By applying the hypothesis of kinetically simple reactions, and of equal also be included.
reactivity of the terminal groups the following is obtained The initiation phase then includes the addition of the fragment I* to the first
molecule of unsaturated monomer M with a kinetic constant K'
v = -d[M]/dt = K2 [M] [M*]
I* + M → IM*
K'

In this way the problem is reduced to finding an expression of the


concentration of radical species [M*] as a function of experimentally According to this equation, the expression of the rate of disappearance of the
measurable quantities. It should also be emphasized that the term [M*] is fragment I* is given by
strictly to be intended as a summation extended to all macroradicals of
different lengths. However, thanks to the assumption of equal reactivity of the -d[I*]/dt = K’ [I*] [M]
terminal groups, it can be considered as a single term.
At this point, the mathematics varies depending on the type of initiation At this point the hypothesis of steady-state is applied for the first time: the
reaction. The case of free-radical polyaddition with thermal initiators will be concentration of reactive fragments I* remains constant during the
hereby considered, as it is the most important from the industrial point of view. polymerization process. This means that the rate of formation and
For clarity, the derivation will not consider the chain transfer reaction, the disappearance must be equal, i.e. in mathematical terms
effects of which will be analyzed separately.
The initiation reaction is given by the decomposition of the initiator I according +d[I*]/dt = -d[I*]/dt = a ε Kd [I] = K’ [I*] [M] = K1 [I]
to the kinetic constant Kd, which is followed by the addition of the first
monomer molecule. The initiator decomposes according to the scheme with K1 = a ε Kd

I →
Kd
aI* (with Kd ≡ kinetic constant) After defining the rate constant of the initiation phase K1 according to the steps
outlined above, it is now possible to quantify the effects of the rate constants
where a is the number of reactive radical fragments formed from one molecule of the propagation and termination phases. As already mentioned, the
of initiator. Usually, on average a ≈ 2 (such as in the case of AIBN). propagation is the main responsible for the disappearance of the monomer, so
The rate of disappearance of the initiator in the case of kinetically simple
reactions will therefore be given by -d[M]/dt = K2 [M] [M*]

31 32
are observed, where α is around 0.5 and β is about 1.
For the termination reaction, a kinetic constant Kt can be defined that Finally it should be stressed that these relationships are valid at the very
determines the disappearance of macroradicals, regardless of whether this beginning of the polymerization, and thus concentrations must be interpreted
occurs by combination or disproportionation. as initial concentrations (in many textbooks in fact, the symbols adopted are
[I]0 and [M]0). In practice, it is necessary to stop at low conversions of
monomer in the polymer (<5%) if accurate kinetic studies are needed.
At this point it is important to clarify the concept of yield in a radical
polyaddition, and in a linear scheme polymerization in general. Yield η is
defined as the ratio between the mass of the formed polymer and the initial
Therefore, again considering the stoichiometric coefficients equal to the mass of monomer, i.e.
reaction order
η = (W0-W)/W0 = Wp/W0

-d[M*]/dt = 2Kt [M*]2


where W indicates the mass, and the subscript p indicates the polymer.
The coefficient 2 indicates that two macroradicals are consumed in each Experimentally, the yield of a radical polymerization is easily determined
through gravimetric analysis, isolating the formed polymer by precipitation at
termination reaction (irrespective of the termination mechanism).
increasing reaction times. Alternatively, spectroscopic analyses can be
The hypothesis of steady-state is applied for the second time, in this case on
performed on the entire reaction mixture, without the need to separate the
macroradicals M*. As a result, the rate of macroradical formation during the
formed polymer. In fact, monomer and polymer usually show different signals,
initial phase and the rate of macroradical disappearance during the termination
phase are equal, i.e. for example in NMR analysis (nuclear magnetic resonance). Finally it is also
possible to apply methods of on-line monitoring of the yield, considering the
+d[M*]/dt = -d[M*]/dt = K1 [I] = 2Kt [M*]2 = K3 [M*]2 (to simplify, 2Kt = K3) progressive increase in the density of the reaction system.
Already at this stage, it is worth pointing out some differences between step-
The expression for the concentration of macroradicals as a function of and chain-growth polymerization processes described in the previous chapter.
experimentally accessible parameters can now be obtained For step-growth processes, it is not possible to speak about polymerization
yield in the same terms as those discussed here, since the free monomer
rapidly tends to zero and, strictly speaking, the entire reaction mass can be
[M*] = (K1/K3)0.5 [I]0.5
considered as "product". Instead, the extent of reaction p must be considered,
i.e. the fraction of reacted terminals, as the parameter that describes the
At this point, the expression of [M*] is substituted into the equation for the
conversion of the polymerization and therefore its molecular weight.
reaction rate, obtaining the final relationship
Conversely, it is conceptually easier to define the reaction yield in the case of
free-radical chain-growth polymerization as seen above. In this case, the
-d[M]/dt = K2 [M] [M*] = K2 (K1/K3)0.5 [I]0.5 [M]
amount of free monomer is almost unrelated to the molecular weight
obtainable by the system.
The model thus leads to a law showing that the reaction rate is proportional to
Figure 12 shows some typical yield(conversion)-time curves (or conversion
the square root of the initiator concentration, and to the first power of the
curves) for radical polymerizations.
concentration of monomer. In fact experimentally kinetic laws of the type

v = Koss [I]α [M]β

33 34
to meet in order to yield termination reactions. In other words, given the
relationship

v = K2 (K1/K3)0.5 [I]0.5 [M]

the rate of reaction will tend to infinity if K3 tends to zero. In this case, the
system self-accelerates, it no longer has the possibility to dispose of the
exothermal heat produced by the polyreaction, and in the presence of large
reaction volumes the process gets out of control and can also lead to explosion.

Figure 12. Yield-time curves for radical polymerizations

From the shape of the curve it is possible to calculate the reaction rate. In
particular, it can easily be shown that the tangent to the yield-time curve at
the origin is proportional to the reaction rate v

dη/dt = d(W0-W)/dt(W0) = -1/W0 dW/dt = -1/[M]0 d[M]/dt = 1/[M]0 v

The curve a) illustrates the typical behavior of a yield-time curve for a free-
radical polyaddition process. The slope of the curve gradually decreases, i.e.
the reaction rate tends to decrease as the concentration of monomer
Figure 13. Gel effect in free-radical polymerization
diminishes. Curve b) shows the effect of a retarding agent (such as a chain
transfer agent S* with reactivity lower than M*), while the curve c) shows the
Other examples of conversion curves are shown in Figure 14. By indicating
effect of the presence of inhibitors: an induction time appears after which the
with N the reference curve, curve A is obtained in the presence of accelerators,
conversion curve shows more or less the same pattern as that observed in a).
curve R in the presence of retardants, while curves I', I'' and I''' are
Yield-time curves can be considered approximately linear at low conversions
characterized by increasing amounts of inhibitor.
(5-10%). For higher values of conversion, the presence of a curvature
indicates a decrease of the reaction rate due to progressive dilution of the
monomer. However, the initial concentration [M]0 can have a significant effect
on the yield-time curve of processes carried-out in homogeneous phase (in
solution or in bulk). As M0 grows, the curvature is progressively less negative
up to a certain point beyond which abrupt positive variations of the yield-time
curve are observed, indicating that the polymerization undergoes a process of
self-acceleration. In Figure 13 this occurs for example when [M]0 = 60%.
This phenomenon is known as gel effect, and is caused by the high viscosity of
the system. In fact, the viscosity increases with concentration, and increases
Figure 14. Examples of yield-time curves for radical polymerizations in the presence of
exponentially with the molecular weight. However, the high viscosity scarcely
accelerators, retardants and inhibitor
affects the microbrownian diffusion motions of small monomer molecules,
while it greatly hinders the macrobrownian motions of macroradicals that need

35 36
As opposed to the case of step growth polyreactions, yields in radical simplification is possible by considering the case of low-conversion
polyadditions are often less than 90%, i.e. at the end of the process there is homogeneous-phase processes, where the parameters [M], [I] and the kinetic
always a quantity of unreacted monomer that must be removed and possibly constants k can be considered constant. Under these conditions the molecular
recovered before the polymer is used. weight does not change with conversion.
By considering first the situation where one polymer molecule is produced from
It can be assumed and experimentally verified that the average degree of each kinetic chain (termination by disproportionation and/or chain transfer but
polymerization Xn is not dependent on the yield if conversions are low. At this no combination), the molecular weight distribution can be derived exactly in
_
point, it is useful to define the length of the kinetic chain ν as the number of the same manner as for linear step polymerizations, and the following relations
monomers linked to each other during the lifetime of the radical. In hold:
mathematical terms this is equivalent to the ratio between propagation and
termination or initiation rates (the latter being equivalent because of the nx = p(x-1) (1-p)
hypothesis of steady-state).
wx = x p(x-1) (1-p)2

v propag. - d[M]/dt K 2 [M][M• ] Xn = 1/1-p


ν= = = →
v term. = vinit. + d[I• ]/dt K1[I] Xw = (1+p) / (1-p)

It is however necessary to re-define p, which for free-radical chain


K 2 [M] [M]
→ ν= = const ⋅ 0.5 polymerizations is conceptually given by the probability that a given
K1K 3 [I]0.5 [I] macroradical continues to propagate instead of terminating; analytically, this
can be expressed as

The length of the kinetic chain is thus directly proportional to the concentration p=v propagation / (v propagation +v termination +v transfer)

of monomer, and inversely proportional to the square root of the concentration


of the initiator. Moreover, both nx and wx concern only the polymer fraction (unreacted
The correlation between the kinetic chain length and the molecular weight (or monomer fraction is excluded). In these conditions, if the termination
average degree of polymerization) depends on the termination mechanism. In mechanism is by disproportionation, even in the presence of chain transfer, it
_ can be shown that
fact, if the disproportionation mechanism prevails Xn = ν , while if the

combination mechanism prevails Xn = 2ν . In most cases there will be an
Xw/Xn = 1 + p

intermediate situation due to the simultaneous presence of the two termination


i.e. the molecular weight distribution tends to 2 for values of p tending to 1, as
mechanisms.
in the case of polycondensations.
The description of the molecular weight distribution in free radical
For termination by coupling / combination where a polymer molecule arises
polymerizations is more difficult to generalize than in the case of
from the combination of two kinetic chains, the size distribution is narrower.
polycondensations, due to the simultaneous occurrence of different transfer
Actually consider the probability ny+z of formation of an x-sized polymer
and termination mechanisms. The situation is further complicated since the
molecule by the coupling of y- and z-sized propagating radicals. Then ny+z will
molecular weight of the polymer produced at any instant may vary with
conversion due to changes in the monomer and catalyst concentration. A first

37 38
be the product of probabilities, ny and nz, of forming the y- and z-sized For polymerization where termination occurs by both combination and
propagating radicals. disproportionation, one can obtain the size distribution by a weighted
combination of the above two sets of distribution functions.
ny = (1-p)py-1 However, in many practical cases of industrial interest, free-radical
polymerizations are pushed to medium-high or complete conversions. The
nz = (1-p)pz-1 distribution of molecular weights becomes much broader because the length of
the kinetic chain - and thus the molecular weight of the single chain - depends
or on the ratio [M]/[I]0.5 and will change progressively with conversion (because
[M] decreases faster than [I], and chain transfer phenomena to polymer can
nz = (1-p)px-y-1 occur). Generally, polymers characterized by an excessive polydispersity are
undesirable, and therefore the molecular weight distribution width can be
since y+z = x. Then ny+z will be given by: partially controlled by adding portions of initiator and possibly monomer.
Finally, still broader molecular size distributions are achieved when the
ny+z = (1-p)2px-2 polymerization process is carried out in heterogeneous phase (like in emulsion
polymerizations, see chapter 6).
The equation written above gives a probability of forming an x-sized polymer
molecule by only one of the many equally probable pathways (for ex. Let us consider now in more detail the effect of chain transfer reactions which
combination of a (z+1)- with a (y-1)-sized radicals, and so on). For high are always present to a certain extent in free radical polymerizations. As
degrees of polymerization (such as those normally achieved by free-radical already mentioned, the transfer reaction often involves hydrogen removal by
processes) there are about x- different pathways of producing an x-sized the macroradical from a second molecule which may be a monomer, solvent,
polymer molecule, and therefore the equation of probability can be rewritten already formed polymer etc. For example, if transfer to the monomer, the
as initiator and the solvent occurs the disappearance rate of the macroradical will
be influenced by the following equilibria
nx = x(1-p)2px-2
- d[M*]/dt = KtrM [M] [M*]
Following the usual statistical approach the other relations can be derived - d[M*]/dt = KtrI [I] [M*]
- d[M*]/dt = KtrS [S] [M*]
wx = 1/2x(1-p)3(x-1)px-2
The chain transfer affects the molecular weight of the produced polymer, and
Xn = 2/(1-p) thus on the length of the kinetic chain, whose expression must be changed to
take into account the new contributions to the disappearance rate of the
Xw = (2+p)/(1-p) macroradical

Xw/Xn = (2+p)/2

i.e. the width of the molecular weight distribution will tend to the limiting value
of 1.5 for p close to 1.

39 40
where Ktr are the kinetic constants related to the different transfer reactions.
_
K = A exp (-Ea)/RT
By increasing the denominator, the numeric value of ν decreases.
Since the length of the kinetic chain is directly correlated to the average where A is the pre-exponential factor and Ea is the activation energy of the
degree of polymerization, it is possible to establish a quantitative relationship process. Usually, E1 is of the order of 30 kcal/mol, E2 approx. 5 Kcal/mol and E3
between the characteristic parameters of the chain transfer process (Ktr and approx. 1 Kcal/mol. Therefore the ∆E relative to Kapp (apparent rate constant of
concentration of the transfer agent) and the molecular weight. the entire polymerization process) can be estimated as ∆E = E2 + ½ E1 – ½ E3
If a polymerization where termination is by disproportionation is considered for ≈ 20 Kcal/mol, a value that allows to quantify approximately the rate increase
_
the sake of simplicity, then Xn = ν . Considering the reciprocal of the
with temperature. Also note, as is typical of chain growth processes, that the
greater activation energy is relative to the first phase of the polyreaction
expression of the kinetic chain length, a linear relationship is obtained which
(initiation reaction).
correlates the inverse of the average degree of polymerization with the
As to the effect of concentration, being Xn ∝ [M] / [I]0.5, if the monomer
concentration of the species which give transfer:
concentration is increased both the reaction rate and molecular weight
increase. On the other hand, if the concentration of the initiator is increased,
1/Xn = (K3 [M*])
/ (K2 [M]) + CM + CI [I]/[M] + CS [S]/[M]
the reaction rate increases but the molecular weight decreases.
= (1/Xn)0 + CM + CI [I]/[M] + CS [S]/[M]
In practice, to moderate the molecular weight (radical processes often
generate high molecular weights that are then difficult to process
where Ci = Ktr-i/K2, i.e. the term C expresses the kinetic competition between
technologically) a small amount of initiator is generally used (because of its
transfer reaction to the i-th molecule and propagation reaction: the greater is
high costs, significantly higher than for monomers) and transfer agents are
C, the more significant is the effect of the chain transfer. By observing the
preferably added.
above equation, it is possible to notice how the term K3 [M*]/K2 [M] is
equivalent to the inverse of the ratio between propagation and termination
It is finally useful to summarize and compare the characteristic features of
rate in the absence of transfer, and is therefore equal to (1/Xn)0. This equation
polycondensations and free-radical polyadditions.
quantifies the effect of decrease of the molecular weight caused by chain
Polycondensations follow an ionic-type chemical mechanism with step-growth
transfer.
kinetics (i.e. slow); the molecular weight increases with the extent of reaction
p, which in turn depends on time and on the stoichiometric ratio between the
When running an industrial process, as will also be discussed in chapter 6,
reagents r. At each stage of the process the polymer is characterized by a
free-radical polymerizations require a high control of the process variables
negligible level of residual monomer (as its weight fraction). Therefore,
such as temperature, pressure (monomers are often gaseous) and
purification processes are generally not necessary or easy.
concentration to optimize reaction rate, heat management, control of the
Free-radical polyadditions follow a chain-growth-type polymerization kinetics
molecular weight and polymerization yield (minimization of the residual
(i.e. fast). The molecular weight obtained is generally high already at the
monomer).
beginning of the reaction with minor changes throughout the polymerization
When the temperature is increased, the overall reaction rate also increases,
time (unless extensive chain transfer reactions to products occur). Since the
which is given by
propagation of a single macroradical is almost instantaneous (reaction times of
the order of ms), over time only the yield grows, i.e. the conversion of
monomer into polymer. The level of residual monomer at the end of the
reaction is often high and requires purification procedures from the produced
polymer.
The individual rate constants K follow an Arrhenius-type law

41 42
Chapter 4: Ionic chain polyadditions

Many types of monomers with electronically interdependent or implicit


functionalities (thus unsaturated) can undergo polyaddition reactions with an
ionic-type chemical mechanism. The kinetic mechanism is usually chain-growth
but under conditions of controlled reactivity it can be step-growth. Figure 15. Active species in ionic polyadditions
As for the case of free-radical polymerizations, the ionic polyreaction consists
of several phases such as initiation, propagation and termination reactions, It is possible to distinguish two types of attack onto the unsaturated monomer
while the chain transfer reaction is less typical. In some particular cases the of C=C type.
termination reaction can be avoided. This allows to obtain highly controlled In cationic polymerizations with ionic attack, the π electron pair of the
architectures, as will be described hereinafter (see living polymers). monomer is added to the carbocation, generally leading to a regular head-to-
The overall polymerization kinetics - and therefore also the degree of tail linkage (in the - frequent - case of polymerization of vinyl monomers).
regiochemical and stereochemical control of the process - largely depend on Note that monomers that can undergo cationic polymerization possess R
the reactivity of the ionic pair, which is the chain end that propagates. This substituents with electron-donating effect, thus the formation of the most
reactivity is essentially correlated to the degree of dissociation of the ionic pair substituted carbocation is always favored.
which is in turn a function of the polarity of the reaction environment. Apolar In anionic polymerizations with ionic attack the carbanion electronic pair will be
solvents (like hydrocarbons) are ineffective in the solvation of ionic pairs that added to the less substituted carbon of the unsaturated monomer, with the
will thus remain highly associated and therefore relatively poorly reactive. subsequent formation - also in this case - of the most substituted carboanion.
Conversely, polar solvents such as ethers allow for an effective solvation of the Usually monomers that can undergo anionic polymerization differ from the
ionic pair, which will thus be more available for the monomer addition reaction. previous ones because they have R substituent groups with electron-attracting
effect.
In polymerization processes with counter-ionic attack an essential role is
played by the M+ catalyst, which is typically a salt of a transition metal (high
atomic number). In this particular class of anionic polymerizations, the
unsaturated monomer first undergoes a polarization of the double bond π by
The degree of dissociation of the ionic pair will therefore depend on the the transition metal M+, and then an insertion on the polymeric carbanion, with
dielectric constant of the solvent. In general it will be possible to have both a very good control of both regiochemistry (head-to-tail configuration) and in
cationic and anionic active centers, the latter being more important. The ionic many cases stereochemistry (pseudoasymmetrical carbon configuration).
species that are formed are essentially carbocations, carbanions, oxanions and Figure 16 shows the reaction chemical mechanisms in the various cases
less frequently oxocations (Figure 15). mentioned above.

43 44
to ensure an adequate lifetime of the carbocation. As initiators it is possible to
use protic acids such as sulfuric acid H2SO4, perchloric acid HClO4, phosphoric
acid H3PO4, and Lewis acids such as BF3 and AlCl3. Polymerizable monomers
are mostly alpha-substituted olefins and vinyl ethers of the type RO-CH=CH2.
Also epoxides can undergo cationic polymerization with a mechanism of
protonation of the oxygen and ring opening (Fig. 17). Such polyreaction can
also be conducted at room temperature and is quite used in the
photocrosslinkable ink and paint technology.

Figure 16. Ionic (i) and counter-ionic (ii) attack addition mechanisms

Figure 17. Mechanism of cationic polymerization of epoxides


A feature of ionic polyadditions is that the characteristics of the process are
strongly influenced by the experimental conditions adopted for the polyreaction The cationic polyaddition of unsaturated monomers of type C=C includes an
(type of initiator, type of solvent etc.). It is therefore less easy, compared to initiation phase, a propagation phase and a termination phase, as shown in
the case of free-radical polyadditions, to obtain a true generalization of the Figure 18. The termination can take place in different ways, for example it may
kinetic model. be a spontaneous dissociation of the ionic pair usually following a rise in
Furthermore it is always to be remembered that the reactivity of monomers temperature, or the carbocation can be easily quenched by traces of moisture.
(typically ethylenic, vinylic or cyclic) are influenced by the nature of the R Processes must therefore be conducted under strictly anhydrous conditions.
group: if this group is an electron-donating one such as an alkyl or ether group,
the corresponding olefin may be defined as electron-rich and cationic 1. initiation
polymerization is favored. Conversely, if the R group is an electron-attracting
one such as an acrylonitrile -CN or an ester -COOR, the olefin is defined H H
electron-poor and it will preferentially undergo anionic polymerization. In some C C + HX CH3 − CH+ X −
cases such as with styrene (R = phenyl) there will be a stabilizing effect both H R R
of the cationic and anion charge, and both polymerization mechanisms can 2. propagation
take place.
Steric effects are also important since ionic polyadditions compared to radical CH+ X − + CH2 CH R CH2 CH CH2 CH + X −

polyadditions are able to impart to the formed polymer a higher constitutive R R R

control and often also configurational control.

Cationic polyadditions are essentially based on the growth of carbocations, 3. termination


which are high energy - and therefore very reactive - species. The chemical
mechanism of these polyreactions is ionic, the kinetic mechanism is chain-
growth and the polymerization must be typically conducted in an inert
environment (anhydrous hydrocarbons) at low temperature (-40°C to -100°C)

45 46
by thermal decomposition of the ionic pair or as a result of the presence of
CH+ X − CH2 CH + HX (T effect) water, but also because of chain transfer to the solvent in some cases.
R R

1. Initiation
+ acqua
CH+ X − CH2 CH OH + HX (moisture)
R M +B − + CH2 CH B CH C H −M +
R
R R

Figure 18. Phases of cationic polymerization

There are relatively few cases of industrial production processes based on 2. propagation
cationic polymerizations. One of these is the process of polymerization of CH2 CH

CH2 C H M
+

CH2 C H−M+ + CH2 CH


isobutylene CH2=C(CH3)2, which leads to the synthesis of polyisobutylene, the R R
R R
so-called "butyl rubber". Such polymerization is very fast and is carried out at -
100°C in the presence of AlCl3 acting as catalyst. The industrial product is
actually a copolymer which contains small amounts (1-2%) of isoprene, a
monomer that provides the allylic hydrogens –CH2-CH= suitable for
3. termination
vulcanization with sulfur.

CH2 C H−M+ CH CH + MH

R R

CH2 CH H + M-OH
CH2 C H−M+ + H2O
R
R
Anionic polyaddition is based on the reaction of carbanions growth and also
oxanions by ring-opening reaction. These intermediates are, on average, less
Figure 19. Phases of anionic polymerization
reactive than the carbocations described above.
In these polymerizations the kinetic mechanism can be chain-growth but
An industrially relevant case of anionic polymerization is represented by the
sometimes also step-growth or intermediate between these two situations
synthesis of 1,4-butadiene initiated by butyllithium. The anion attack can lead
depending on process variables such as the type of monomer, solvent etc.
to two different constitutional isomers, namely with connections in positions
The process can be conducted in an inert apolar hydrocarbon environment, but
1,4 (a double bond remains in the chain) or in positions 1,2 (the double bond
also in highly polarized systems such as ethers, liquid ammonia etc. which
is outside the chain).
favor the complete dissociation of the ionic pair. As initiators it is possible to
use metals (such as sodium Na), sodium amide NaNH2, alkyllithium
compounds such as butyllithium (BuLi) etc. The anionic polymerization is
possible on alpha-olefins substituted with electron-attracting groups, dienes,
epoxides.
The phases of the anionic polymerization (Fig. 19) are similar to those
described for cationic polymerization. The termination reaction may take place

47 48
Mw/Mn = 1 + r/(1+r)2
In addition, polybutadiene also presents a level of steric isomerism relative to
the configuration of the double bond. By modulating the polarity of the solvent where in this case r is the ratio [M]0/[I]0.
and thus the degree of dissociation of the ionic pair it is possible to change the From the previous relationships it can be observed that in the limit conditions
composition of the cis-trans isomeric mixture. In particular if the anionic of r tending to infinity, the polydispersity index tends to 1. In practical
polymerization of butadiene is carried out in hydrocarbons - thus under conditions, with living polymerizations it is relatively easy to obtain polymers
conditions of moderate reactivity – good regiochemical and (partially) with a very narrow distribution of molecular weights, around 1.05, while the
stereochemical control is obtained with the prevailing formation of 1,4-cis- number average molecular weight can be predetermined from the ratio of the
polybutadiene. Yet it should be emphasized that the best results (high cis concentrations of monomer and initiator.
polybutadiene rubbers, with stereochemical purity higher than 90%) are An interesting industrial example of living anionic polymerization is the
obtained by using Ziegler-Natta type catalysts, which will be described later. synthesis of polybutadiene-polystyrene block copolymers, known by the
acronym SBS (styrene-butadiene-styrene).
The synthesis scheme is shown in Figure 20. The first step involves the
synthesis of the first polystyrene block through living anionic polymerization of
styrene. The molecular weight of this block can be predetermined from the
monomer/initiator ratio, and must anyhow be higher than 10000 in order to
obtain the desired morphology in the copolymer (see below). Under conditions
of inert and absolutely pure and anhydrous solvent, the polystyril carbanion is
A peculiar feature of chain anionic polyaddition processes is the fact that under
not subjected to termination once the addition of styrene (S) monomer is
controlled experimental conditions (temperatures, anhydrous conditions, inert
finished. At this point the formation of the polybutadiene block takes place,
solvents, purity of monomers) it is possible to avoid the termination reaction.
with further growth of molecular weight until the butadiene monomer (B) is
The polymers obtained by processes of this type are called living polymers. In
completely consumed: also in this case it is possible to control the length of
practice, such processes only present the initiation and the propagation
the segment through stoichiometry. Formally, it is then possible to
reactions, with the latter continuing until the monomer is completely consumed.
subsequently add S to obtain the final triblock copolymer.
In the absence of the termination reaction, the polymer thus formed remains
SBS is the most typical example among a class of materials known as
with "living” carbanion terminals, i.e. reactive and ready to resume the
thermoplastic elastomers; their very peculiar physical-mechanical behavior
polymerization with further increase of molecular weight if more monomer is
includes some of the characteristics of vulcanized rubbers (highly reversible
added, even of a different type: in this case block copolymers will be obtained.
deformations at low values of applied stress) and of thermoplastics (the
The peculiarities of living polymerizations are considerable in terms of control
possibility of being thermo-processed several times), and is fully related to the
of molecular weight, macromolecular architecture, and molecular weight
polyphasic morphology of the material.
distribution.
In block copolymers the length of the macromolecular segments is crucial to
The average degree of polymerization can in fact be estimated with good
ensure polyphasicity, i.e. the thermodynamic immiscibility of the blocks S and
approximation simply from the ratio of the initial concentrations of monomer
B, that are also chemically bonded. In fact, the miscibility between the
and initiator.
polymers is hindered by the high molecular weight of the components of the
mixture, unless there are strongly exothermic interactions such as hydrogen
Xn = [M]0/[I]0
bonds. The block copolymer thus formed, usually characterized by a content of
polybutadiene phase around 70%, is morphologically heterophasic with two
Furthermore, the molecular weight distribution that is obtained is not the most
amorphous phases each characterized, in first approximation, by its own Tg
probable, but rather it follows a much tighter Poisson-type distribution
corresponding to that of polybutadiene (rubber, with Tg at about -80°C) and

49 50
polystyrene (rigid thermoplastic, with Tg at about +100°C). From a
microscopic point of view the material can be represented by a soft elastomeric
continuous phase (poly-B) containing hard segregated glassy phases (poly-S)
chemically bound to each other, as shown in Figure 20. These rigid domains
act as reinforcement and as "physical crosslinking" of the material, which at
room temperature behaves like a typical rubber (highly reversible deformability
at low values of stress).
Obviously, the polymer is not chemically crosslinked since if it is heated above
the Tg of polystyrene it flows like any thermoplastic, and it can then be
processed with normal extrusion or casting technologies.

M+B − + n CH2 CH CH2 CH - M+


n
Figure 21. Heterophasic morphology of SBS-type block copolymers

Going back to the synthesis of SBS copolymer, it should be emphasized that,


CH2 CH - M+ + m CH CH CH CH while the poly-S- macrocarbanion shows good reactivity towards butadiene, the
2 2
n
opposite does not hold true, i.e. the poly-B- carbanion is added to the S
monomer with difficulty. This entails that, for the formation of the third block,
- M+ industrial processes make use of coupling agents, typically dichlorosilanes,
CH2 CH CH2 CH CH CH2
n m which can give equally good substitution reactions with butadiene and styrene
carbanions. As shown in Figure 22, dimethyldichlorosilane is used to link the
diblock polyS-polyB- to a second polyS- block synthesized separately. Moreover,
the use of chlorosilanes with different functionalities (such as
CH2 CH CH2 CH CH CH2 - M+ + n CH2 CH
n m
trichloromethylsilane or tetrachlorosilane), also allows to diversify the
architecture of the polymer, leading to branched or star structures that can
exhibit interesting rheological properties, and therefore different processability.

H
CH2 CH CH2 CH CH CH2 CH2 CH - M+ CH3
n m n
CH3 CH2 CH2 CH2 CH2 CH n CH2 CH CH CH2 n CH2 CH CH C + Cl Si Cl

CH3
H H

+ :C - CH2 CH CH2 n CH2 CH2 CH2 CH3

Figure 20. Synthesis of the block copolymer SBS

Figure 21 shows in a simplified way the particular morphology of SBS block


copolymer.

51 52
CH3

CH3 CH2 CH2 CH2 CH2 CH n CH2 CH CH CH2 n Si Cl

CH3

cross-linked for applications, thus generating a large class of sealants,


CH3
adhesives and special elastomers.
CH3 CH2 CH2 CH2 CH2 CH CH2 CH2
n CH2 CH CH CH2 n Si CH CH2 n CH2 CH3
It is interesting to note that from the thermodynamic point of view the anionic
CH3
polymerization of octamethylcyclotetrasiloxane is substantially athermic
(∆H=0) while ∆S is positive, approximately +1.6 cal/mol K. An exoentropic
Figure 22. Synthesis of the SBS copolymer with dimethyldichloro silane polymerization process is a very rare fact, and can be explained by the high
chain flexibility of the linear polymer (caused by the high atomic weight of
Among ionic polymerization processes there are also some important cases of silicon compared to carbon) with a consequent gain of degrees of freedom with
ring-opening polymerizations of cyclic monomers with interdependent respect to the initial cyclic monomer.
functionalities. Such monomers are cyclic ethers (epoxides), acetals, esters,
amides and siloxanes. Cases of anionic and cationic polymerization of epoxides
have already been described, while here the reaction for the synthesis of
polyamide 6 and siliconic polymers will be now discussed.
The polymerization of cyclic amides can be initiated by bases, acids and water.
If caprolactam is gradually heated up to +250°C in the presence of small
amounts of water or soda, and in an inert atmosphere, polycaprolactam or
polyamide 6 is obtained via a mechanism of anionic polymerization:

O H

C N
O H
H2C CH2 533K
C CH2 N
5
H2C CH2 N2
n C n
H2

Polyamide 6 is - together with polyamide 66 (obtained from


hexamethylenediamine and adipic acid) - the polyamide with the greatest
volume of production and use. Polyamides 6 and 66 have the same chemical
composition in terms of molecular formula, but different atomic arrangement
which also determines different physical-mechanical properties and
Figure 23. Synthesis of polydimethylsiloxane
morphology of the material. The process of polymerization of caprolactam is
practically intermediate between a step-growth and chain-growth scheme
Stereospecific polymerization
because once the polymer is formed by opening of the caprolactam, equilibria
Chain polymerization processes with counter-ionic attack mechanism are an
can establish between formation and breaking of the amidic bond that lead to
important class of processes which are generally known as stereospecific
the most probable distribution of molecular weights.
polymerization, or coordination polymerization, or polymerization with Ziegler-
Linear polysiloxanes or silicones, including primarily polydimethylsiloxane, may
Natta catalysis.
be obtained by anionic or cationic ring-opening polymerization of cyclic
These are particular processes of ionic polymerization through insertion, or
monomers such as octamethylcyclotetrasiloxane (Figure 23). Silicones are
progressive coordination of unsaturated monomers with strictly hydrocarbon
polymers characterized by high thermal and chemical inertness, as well as high
flexibility at low temperature with Tg that can reach -120°C. They are usually

53 54
nature (only ethylene, alpha-olefins and dienes) catalyzed by specific surface of the titanium crystals. In fact, inside the catalytic material the
organometallic compounds based on transition-metals salts. titanium presents octahedral coordination as it is surrounded by six chlorine
The core of this technology are the catalysts, known as Ziegler-Natta (ZN) atoms shared with other metal atoms. On the other hand, on the surface there
catalysts, which allow for extremely regular monomer linkages both in terms of are non-shared chlorine atoms, ethyl substituents shared with aluminum (in
connectivity (constitutive level) and in terms of arrangement of the groups in the dimeric form), and unoccupied coordination sites (empty d orbitals) to
the space (configurational level). In practice, ZN catalysts make it possible to satisfy the principle of electro-neutrality, as shown schematically in Figure 24.
obtain polymers with a high molecular weight and regular configuration, thus
with the possibility of obtaining high degrees of crystallinity.
Al
ZN catalysts are organometallic compounds obtained by reaction of salts
CH2 CH CH2 CH3
(typically halides) of transition metals of group IV-VIII such as Ti, Zr, V, Co, Hf,
R
Nb, Mo, Co, Ni with alkyl or aryl derivatives of non-transition metals (I-III Ti

group), often of Al, known as cocatalysts. They are very delicate compounds,
whose activity is easily poisoned and that are therefore compatible with only a
few monomers strictly with hydrocarbon composition (ethylene, alpha-olefins Figure 24. Formation of the catalytic sites on the surface of ZN catalysts
and dienes), inert solvents (hydrocarbons), mild process temperatures and
pressures. Figure 25 illustrates in a simplified manner the mechanism of initiation and
The types of polymer products that can be obtained are relatively few, but they propagation mediated by ZN catalysts: the unsaturated C=C monomer is
are materials with huge production volumes and countless possible coordinated by the d orbital of titanium, and undergoes insertion by the ethyl
applications: among these are the aforementioned HDPE and isotactic PP, 1,4- group, in turn leaving free another d orbital meant to accommodate a second
cis-polybutadiene, 1,4-cis polyisoprene, ethylene-propylene copolymers etc. monomer molecule. Since the reaction takes place on a three-dimensional
ZN processes typically lead to polymers characterized by a broad molecular surface of a crystal - and with an ionic pair that is strongly associated because
weight distribution. of the hydrocarbon solvent - there are strong geometrical constraints that
From a historical point of view this represents at truly "milestone" in the field affect the attack modes of the monomers. In practice, the progressive
of synthetic polymers: in 1952 Karl Ziegler synthesized high density linear insertion reaction of the monomeric units occurs in a highly stereospecific
polyethylene (HDPE) from a catalyst based on AlCl3 and TiCl4. Then Giulio mode, i.e. always on the same side, leading to a regular configuration of the
Natta obtained the stereospecific synthesis of isotactic polypropylene, now the pseudo-asymmetrical carbon of the obtained polymer. With these types of
most widely used plastic in the world together with polyethylene. Both catalysts isotactic and cis-tactic polymers are always obtained, with few
researchers won the Nobel Prize for Chemistry in 1963. exceptions (i.e. VCl4 based catalysts).
The basic mechanism involves various steps of alkylation and reduction of the
titanium salt:

TiCl4 + AlEt3 → TiCl3Et + Et2AlCl


δ+ δ+ CH3
TiCl3Et → TiCl3 + Et# Al Al
δ-
TiCl3 + AlEt3 → TiCl2Et + Et2AlCl CH2 CH3 CH2
TiCl2Et → TiCl2 + Et# δ+ δ+
Ti Ti δ- δ+
CH2 CH CH CH2
The polymerization takes place via the insertion of the olefin in the σ−bond R R
carbon-transition metal. In such catalysts, heterogeneous and therefore
insoluble in the reaction environment, the polymerization takes place on the
Figure 25. Initiation and propagation mechanism with ZN catalysis

55 56
There is not a true termination reaction; it is common practice to add a
mixture of monomers and hydrogen, which practically works as a
monofunctional molecule ending the polyreaction (Fig. 26).

δ+
Al Al
δ- Figure 27. Enantiomorphic structures of the catalytic complex (Cl=shared chlorine, Cl’=non-
CH2 CH3 + H2 H + CH3 CH3
shared chlorine)
δ+ δ+
Ti Ti From the point of view of an industrial process, ZN catalysts can be classified
as heterogeneous (insoluble) and homogeneous (soluble). Over the years
several generations of catalysts have been developed, with increasing
Figure 26. Termination mechanism with H2
performance in terms of yield and stereochemical purity of the product.
Heterogeneous catalysts can be unsupported or supported.
It is interesting to investigate in detail the mechanism of polymerization of Unsupported catalysts are the first generation of products developed by Natta
propylene in order to understand the reasons for the stereochemical control of for the production of isotactic polypropylene, and are represented by systems
the process. The initiation reaction involves possible structural and steric such as α-TiCl3/AlEt3, VCl3/AlEt3, δ-TiCl3/AlEt2Cl. In fact, already the classic
isomers as shown schematically below. Ziegler catalyst based on TiCl4 and AlEt3 forms β-TiCl3 (brown powder) by
reduction, which however showed low activity and stereospecificity with
respect to the α and δ forms of titanium trichloride (violet). Anyhow, with the
first generation heterogeneous catalysts it is possible to obtain isotactic PP
*
Mt C2H5 + CH2 CH Mt CH2 CH C 2H5
with relatively low yield and stereochemical purity (<92% isotacticity). It is
CH3 CH3 therefore necessary to wash off the residual catalyst and extract the fraction of
atactic polymer with hot heptane.
Mt C2H5 + CH2 CH Mt
*
CH CH2 C2H5
Second generation unsupported ZN catalysts have been developed since the
1970s and consist of smaller, porous, large surface area crystals (about 150
CH3 CH3
m2/g) that thus have a greater number of active sites on their surface.
Moreover, the isotacticity index can be increased by selectively poisoning the
With classical catalysts an initiation reaction with primary or antiMarkovnikov aspecific sites (more acidic) with Lewis bases such as ethers, with mild effects
addition is obtained (verified by NMR), head-to-tail linkage and an on the polymerization yield.
approximately 90% isotactic polymer. The stereospecificity of the Later, research has shifted towards the development of ZN catalysts supported
polymerization is ensured by the steric control of the catalytic site, which is on inert materials (Al2O3, SiO2, MgCl2) and the development of third and fourth
inherently asymmetrical. By way of example the two enantiomorphic structures generation catalysts (superactive). Examples are the catalysts obtained by
(mirror images) of the catalytic complex are shown in Figure 27. grinding anhydrous MgCl2 with TiCl4 in the presence of weak Lewis bases such
as ethyl benzoate (known as inner base), and subsequent addition of
cocatalyst (triethyl aluminum) often associated with a second weak base
(known as external base) such as an ether or ester. It should be noted that,
with catalysts of the latest generation, the obtained polymerization yields and
isotacticity indexes are so high as to make further operations of product
purification superfluous, with a considerable reduction of production costs.
57 58
which are alternatively occupied by the growing macromolecule and by the
A separate chapter is then represented by homogeneous catalysts (soluble) monomer molecule. Figures 28 and 29 illustrate the meaning of what discussed
based on metallocenes. Metallocenes are metallorganic complexes of transition above.
metals, with a general structure of type Cp2MtX2. They must be activated by
soluble cocatalysts based on methyl alumoxane (MAO), obtained by hydrolysis
of trimethyl aluminum.

(n+2) Al(CH3)3 + (n+1) H2O →


CH3-Al(CH3)-O-(Al(CH3)-O)n-Al(CH3)-CH3

Cp indicates the cyclopentadienyl anion (aromatic cycle with five carbon atoms
and six electrons), while X indicates a halogen, generally chlorine. Among the
main types of metallocenes it is worth mentioning titanocenes, zirconocenes,
and hafniocenes. Figure 28. Stereochemical equivalence of the two insertion positions of the monomer in chiral
The role of MAO seems to be that of activating the metallocene complex with metallocenes with C2 symmetry.
formation of cationic alkylated species, and consequent protection of the
resulting alkylcation.

Cp2MtX2 + MAO → Cp2MtX+MAOX-

Homogeneous metallocene/MAO catalytic systems exhibit high polymerization


yield, and it is often possible to obtain polymers with very high molecular
weight (106). Figure 29. Stereochemical non-equivalence (mirror images) of the two positions of insertion of

The main steps of the process are common to conventional polymerizations the monomer in chiral metallocenes with Cs symmetry

with ZN catalysts, namely:


- Regiospecific insertion of olefins Note. For a better understanding of the mechanism of metallocenes catalysis it is useful to

- Termination through hydrogen transfer recall some elements of molecular symmetry. A rotation axis, indicated with the symbol Cn, is

- Stereospecific propagation if asymmetric (thus not Cp-based) metallocenes an axis that passes through the object under examination (the molecule) and such that a

are used. rotation of 360°/n degrees around that axis brings back to a model of the object that cannot

In fact, to exert also a stereochemical control, it is necessary to use a ligand be distinguished from the previous one. All molecules possess the trivial symmetry axis C1 (i.e.

that is inherently asymmetric. Depending on the symmetry of the ligand and all are identical to themselves for rotations of 360° around an axis). Yet some molecules only

with high stereospecificity it will be possible to obtain not only isotactic but also have that one axis as element of symmetry (for example a chiral carbon, i.e. variously tetra

syndiotactic polymers, even with more complex regular structures. substituted).

In particular a metallocene with symmetry group C2v (such as Cp based ones) A plane of reflection (σ symbol) is a plane which divides an object in such a way that the two

generates atactic polymers, a metallocene with group symmetry C2 generates halves are reflected exactly into one another. A center of inversion (i symbol) is a point of the

isotactic polymers, and a metallocene with symmetry group Cs generates molecule such that by moving on a straight line in opposite directions from that point the same

syndiotactic polymers. The explanation can be found in the stereochemical atoms are met at equal distances.

equivalence or non-equivalence of the two insertion positions of the catalyst,

59 60
The molecules belonging to the punctual group C1 are asymmetrical molecules because they
only possess the rotation axis C1 (chiral molecules). The molecules belonging to the punctual
Chapter 5: Copolymerization
group Cn (with n> 1, thus C2, C3...) are known as dissymmetric molecules. Molecules belonging
to the group Cs also possess a symmetry plane. Molecules belonging to the punctual group CnV,
A copolymer is a polymer consisting of at least two monomers. The
very common, are symmetrical, possess a Cn axis and n σ−planes; examples of C2V molecules
copolymerization is the chemical method used to modulate the properties of
are H2O and also the bis-cyclopentadienyls complexes.
the polymeric material. The main variables to consider in a copolymerization
process are: 1. the composition of the copolymer, which depends on the
In particular, syndiotactic polypropylene can be obtained when the olefin is
composition of the monomer mixture and on the relative reactivity of the
forced to place the methyl in enantiomeric positions during polymerization as
monomers; 2. the arrangement of the monomeric units at constant
shown in Figure 29.
composition, which is essentially a function of the synthetic method (statistical,
For example, since bis-chloro zirconocene is a symmetrical molecule, it can
preordered). Indeed, the disposition of the monomeric units, i.e. the
only generate atactic polymers.
constitutive level, determines the types of the different copolymers.

In statistical copolymers the monomer units (hereinafter referred to as A and


B) are linked to each other randomly without any order. An example of
Cl Zr Cl
statistical copolymer is SBR rubber (styrene-butadiene rubber), a copolymer of
bis-
chlorozirconocene styrene and butadiene obtained by free-radical polymerization.
Alternating copolymers are instead characterized by an ordered sequence of
units of type AB; examples are given by ethylene-tetrafluoroethylene (ETFE),
If the ligand is instead a bis-indenyl group, asymmetric, the propylene but also in a broad sense by many polycondensation polymers obtained by
monomer must approach the catalyst with its –CH3 group always far away polymerization of monomers of type aAa + bBb, such as polyethylene
from the aromatic ligand because of steric hindrance. This determines the terephthalate.
formation of a polymer that is stereochemically regular, in particular isotactic The already mentioned block copolymers are characterized by a regular
polypropylene (control of tacticity) due to the C2 symmetry of this particular arrangement of long poly-A and poly-B type sequences, and can be typically
metallocene. obtained by living polymerization processes.
Finally, graft copolymers are obtained by non-regular grafting of poly-B type
chains on preformed poly-A polymers. Figure 30 summarizes the molecular
microstructure of the copolymers described above.

H
H2C C + CH2
Zr
CH3 CH2
control of tacticity
H3C

61 62
material is known as HIPS (high impact polystyrene) and it is the type of
polystyrene normally used for structural purposes. Figure 31 summarizes
schematically the structure and composition of styrene copolymers.

Figure 31: Structure and composition of styrene copolymers


Figure 30: Molecular architecture of various copolymer types

Similar percentages of the two monomers (around 30% styrene and 70%
Polystyrene is a good example on how the properties of a polymeric material butadiene) can be copolymerized via living anionic copolymerization, to obtain
can be modulated or improved through copolymerization. Polystyrene is in fact the already described SBS elastomer. This is also a polyphasic material with
an amorphous polymer (atactic because it is obtained by radical two Tg just like HIPS, although it shows very different mechanical properties
polymerization), with Tg around +100°C, good stiffness and transparency, but since the continuous phase in this case will be of elastomeric type.
with high fragility and poor chemical resistance which limit its use as a Finally, an interesting case is given by the statistical copolymerization of
structural engineering material. When styrene is statistically copolymerized ethylene and propylene performed with ZN-type catalysis. The resulting
with acrylonitrile CH2=CH-CN one obtains the copolymer referred to as SAN material is a statistical copolymer known as EPM or EPR, an amorphous
(styrene-acrylonitrile), with slightly higher Tg but with greatly improved rubbery polymer with Tg at about -50 °C. It should be emphasized that
chemical resistance: the material reduces the tendency to form micro-cracks although both ethylene and propylene form highly crystalline polymers (linear
when in contact with chemicals, a behaviour known as “environmental stress and stereoregular) by ZN-catalysis, in the specific process conditions used to
crazing” that usually causes a dramatic worsening of the fragility of the form EPR the corresponding copolymer is strictly amorphous due to the lack of
material. If styrene is statistically copolymerized with butadiene CH2 = CH-CH regularity in the composition, typical of statistical copolymers. The Tg is
= CH2 in molar ratios around 2/8 or 3/7 one obtains the above mentioned SBR intermediate between that of polyethylene (-100°C) and polypropylene (0°C).
rubber, a monophasic elastomer with a Tg around -40 ° C, widely used in tire
technology. Polymerizing and then grafting polystyrene from polybutadiene In statistical copolymerization, the direct feed of mixtures of at least two
latexes, in ratios around 8/2, one obtains a polyphasic material (i.e. monomers is required. In principle, statistical copolymerizations are possible
characterized by more than one Tg) consisting of a matrix – the continuous
phase of polystyrene - containing rubbery polybutadiene microinclusions. This

63 64
both as step-growth and as chain-growth reactions, yet the latter - especially The problem of modeling a chain copolymerization process (usually of radical-
with a free-radical mechanism - are by far the most important. type) can be addressed by first considering a propagation scheme for the two
In the step-growth processes one can speak about true statistical monomers M1 and M2 competing against each other. Such monomers are
copolymerization in the case of a polyreaction among at least three monomers generally characterized by a different reactivity which is also a function of the
of the type aA1a + aA2a + bBb. In fact, the classical polycondensation of aAa + chemical nature of the chain terminal (macroradical intermediate), as shown in
bBb already produces a perfectly alternating copolymer, or even better a Figure 32.
homopolymer with a AB repeating unit. Due to the intrinsic characteristics of
step-growth processes, the composition of the copolymer will correspond to
that of the mixture of the starting monomers (since it is all "product"), while M1∗ + M1 K→
11
M1∗
the arrangement will depend on the reactivity of the monomers. A completely
random arrangement will be obtained when all the monomers have equal M1∗ + M2 K
12
→ M2∗
reactivity (rare event), and the polyreaction can be considered in this case as a
sequence of bernoullian events. On the other hand, it will be possible to talk M2∗ + M1 K
21
→ M1∗
about statistical copolymerization when the reactivity depends on the nature of
the last or penultimate monomer unit in the chain. Moreover, in step-growth M2∗ + M2 K
22
→ M2∗
copolymerizations, typically polycondensations, interchange reactions (such as
Figure 32. Propagation reactions in the two-monomer copolymerization process
transesterifications) often occur, causing mixing of the sequences over long
reaction times with a tendency to reach the most probable distribution of
The reactivity of the macroradical may depend to a different extent on the type
molecular weights, and a completely random arrangement of the monomeric
of monomeric units entered at the beginning of the process. Herein only the
units.
terminal model will be considered, known as such because the reactivity of the
macroradical is assumed to depend on the nature of only the last monomer
As mentioned earlier, chain copolymerizations are more important industrial
unit inserted in the macromolecular chain.
processes. Their kinetic model will be obtained by generalizing the approach
In this kind of process there are four distinct propagation reactions competing
already outlined for the kinetic equation of radical polymerization.
with each other, represented in Figure 32 and quantified by the kinetic
In particular, it is assumed that the rate of the polymerization process is
constants K11, K12, K21 and K22.
substantially equal to that of monomer consumption, which in turn corresponds
The following assumptions are made: the individual reactions are kinetically
to the propagation reaction rate constant K2.
simple; the hypothesis of steady-state is valid; the process generates high
molecular weight chains (i.e. that the kinetic chain is long); the reactivity of
Mn• + M →
K2
Mn• +1
the macroradical is only function of the last inserted monomer unit and is
independent of molecular weight.
Let us first remember that in chain-growth reactions the rate of formation of The instantaneous composition of the copolymer formed as a function of the
the single macromolecule is much lower than the total process time (which relative rate of disappearance of the two monomers during the time interval dt
essentially must be long enough as to ensure a complete and progressive is defined as follows
decomposition of the initiator). On the other hand, in step-growth reactions the
time taken by the macromolecule to form corresponds to the time of the Π = -d[M1]/-d[M2]
polymerization process.
This equation is obviously valid for chain-growth processes where the
formation of the macromolecule is instantaneous. In contrast, in step-growth

65 66
processes where the chain is growing during all the process time, a more By dividing the expression of Π by K12 [M1*] [M2], and after appropriately
complex integral function will be employed, Π = -∫d[M1] / ∫d[M2]. substituting the terms [M1]/[M2] = µ, and [M1*]/[M2*] = K21 [M1] / K12 [M2], the
For the general case of a free-radical copolymerization with ultimate effect, the well-known copolymerization equation or Lewis-Mayo equation is obtained:
instantaneous composition of the copolymer formed can be expressed as

The meaning of the Lewis-Mayo equation is that in a chain copolymerization


process the instantaneous composition of the output copolymer is typically
At this stage, the application of steady-state hypothesis to both macroradicals
different from that of the input monomer mixture. Obviously bringing the
corresponds to the assumption of their constant concentration over time. This
polymerization yield to values close to one (something which is not always
also means that cross propagation reactions, i.e. of copolymerization in the
possible, or easy) will result in an average copolymer composition approaching
strict sense (macroradical M1* on monomer M2 and macroradical M2* on
that of the monomer mixture. However such copolymer composition may be
monomer M1) are characterized by the same rate, or in mathematical terms
characterized by a rather wide composition distribution, and effect known as
“compositional drift”.
K12 [M1*] [M2] = K21 [M2*] [M1]
In many cases the copolymerization equation is graphically represented by
plotting the molar fraction of monomer in the input mixture (f1) on the x-axis,
Therefore it is possible to obtain an expression of radical concentration as a
and the mole fraction of monomer in the copolymer (F1) on the y-axis; these
function of experimentally accessible parameters
are the so-called composition diagrams or f1-F1 diagrams. These quantities can
be related to the previously defined composition ratios µ and Π as shown in the
following equations

The composition of the inlet monomer mixture is defined as

µ = [M1]/[M2]

It is now necessary to define the reactivity ratios r1 and r2 as


With appropriate substitutions it is therefore possible to obtain an expression
r1 = K11/K12 for F1 as a function of easily quantifiable and measurable parameters, namely
r2 = K22/K21

In practice, the reactivity ratios express the competition between direct and
cross propagation, i.e. between homopolymerization and copolymerization in
the strict sense; low values of reactivity ratio indicate the tendency to form
alternating sequences. The resulting composition diagrams relate the instantaneous composition of
the copolymer to the initial composition of the monomer mixture.

67 68
Figure 33 shows the graphical representation of some compositional diagrams In general, for various situations (pairs of r1 and r2 values) the curve intersects
created for the special case in which r1 x r2 = 1; this case is referred to as ideal the diagonal; the value of f1 for which such an intersection occurs is called
copolymerization. The diagonal of the graph represents the case in which r1=r2, azeotropic composition, and it represents the point of equivalence between the
which implies that for the entire range of possible compositions f1 = F1 or composition of the monomer feed and the composition of the resulting
µ = Π (the compositions of the monomer mixture and the formed copolymer are copolymer. However, the azeotropic composition is of limited practical use,
identical). since unstable azeotropes are usually found, as small errors in the input
composition result in a gradual shift from the azeotropic point.
As a general consideration, values of reactivity ratio <1 will yield a prevalence
of alternating sequences. Conversely, values of reactivity ratio >1 will result in
longer sequences of monomers of the same type (blocks).
To summarise, the case r1 x r2 = 1 represents the so called ideal
copolymerization. This implies that K11/K12 = K21/K22, i.e. the same relative
reactivities are reported. Namely, each macroradical has equal probability of
adding the monomer M1 or M2 regardless of the nature of the last incorporated
monomer unit. The sequence of events is Bernoullian and this leads to a truly
random copolymer.
The frequent case for which r1 x r2 < 1 with both reactivity ratios < 1 leads
instead to composition diagrams that always show an intersection with the
diagonal, i.e. an azeotropic composition. If both reactivity ratios tend to zero,
Figure 33. Composition diagrams with r1x r2 = 1 normally there is the formation of an alternating copolymer.
The case in which r1 > 1 and r2 >1 is quite uncommon and the system under
A characteristic of ideal copolymerizations is that the curves never intersect these conditions tends to form two separate homopolymers of M1 and M2.
the diagonal, i.e. if r1 x r2 = 1 but r1 ≠ 1, the composition of the copolymer Instead, when r1 >> 1 and r2 <<1 there will be the formation of copolymers
obtained at low conversions is never coincident with that of the input monomer with long blocks of M1 and few units of M2; under these conditions it is
mixture. generally difficult to obtain a high yield of polymerization.
Figure 34 represents various examples of composition diagrams for different Finally, when the product r1 x r2 = 0 different cases are reported. If r1 = 0 and
pairs of r1 and r2 values. r2 < 1 there will be an azeotropic composition, while if r2 > 1 there will be no
azeotropic composition. The ideal case of both reactivity ratios equal to zero
leads to a perfectly alternating copolymer.
The considerations presented above apply when the instantaneous composition
of the copolymer is considered, or in practical terms at low conversions (<5%).
For each copolymerization, the composition of the input monomer mixture and
that of the formed copolymer will be different, except for the case of azeotropic
composition. While the conversion increases, there will be a drift of the
composition of the monomer mixture which is progressively enriched by the
less reactive monomer, and this in turn leads to a drift of the composition in
the resulting copolymer. The calculation of the instantaneous composition of
the copolymer as a function of conversion is made possible by integrated forms
of the Lewis-Mayo equation. A useful method for analyzing and predicting
Figure 34. Composition diagrams for different r1 and r2 values

69 70
copolymer composition as a function of conversion is that developed by Skeist. which relates the degree of conversion to changes in the comonomer feed
Consider a system initially containing a total of M moles of the two monomers composition.
and in which the copolymer formed is richer in M1 than the feed, i.e. F1>f1. For example the following image shows the behavior observed in the radical
When dM moles of monomer are consumed, the copolymer will contain F1dM copolymerization of styrene and methylmethacrylate, where the average or
moles of the species 1 (remember that F1 is dM1/d(M1+M2)) and the feed will cumulative composition of the copolymer as a function of conversion is shown.
contain (M-dM)(f1-df1) moles of M1. The mass balance for the species 1
requires that the moles of M1 copolymerized equal the difference in the moles
of M1 in the feed before and after reaction:

Mf1 – (M-dM)(f1-df1) = F1dM

The above written equation can be rearranged and converted in integral form
(neglecting the small term df1dM) between the limits of composition M0 - M,
and (f1)0 - f1:

The Skeist equation allows the calculation of F1 as a function of f1 for a given


set of r1 and r2 values. It is possible: to tabulate F1 as a function f1 for various The copolymer is slightly richer in methylmethacrylate than the feed because
values of r1 and r2 (see copolymerization equation); to calculate the term of its larger reactivity ratio. The feed therefore becomes richer in styrene with
within parentheses (F1-f1); and to solve graphically or numerically the integral conversion, leading to an increase in the styrene content of the copolymer with
equation written above within the limits (f1)0 and f1, thus obtaining the conversion. The compositional drift is however acceptably small in this case.
changes of composition of the monomers and of the copolymer as a function of
the degree of conversion of the copolymerization, expressed by the term 1- The determination or estimate of reactivity ratios is a key aspect in the study
M/M0. of copolymerization.
The Skeist equation has been integrated to a useful closed form Experimentally, it comes to performing tests of copolymerization at various
values of monomeric composition µ, and to measuring the composition of the
copolymer obtained at low yield (typically 5%). The copolymer is normally
separated from the unreacted monomer by precipitating it in a solvent mixture
that is such for the monomer, but not for the polymer. Finding a suitable
solvent mixture is generally not difficult because the solubility, for substances
with the same chemical nature, decreases significantly at increasing molecular
weight.
The composition of the copolymer can be determined by various techniques.
Among these, the one with broader applicability is the nuclear magnetic
resonance (NMR) spectroscopy. This technique is extremely useful not only in
determining the composition, but also the arrangement (sequences) of the
71 72
monomeric units in the copolymer. The reader is referred to specialized texts expressed in a scale relative to styrene, for which it is assumed that Q = 1 and
on spectroscopy of polymers for further studies. e = -0.8.
From a mathematical point of view, the various methods of reduction of the The reactivity ratios can be calculated as follows:
experimental data are based on different procedures to linearize the equation
of copolymerization.
According to the commonly used Mayo-Lewis method the copolymerization
equation can be rewritten as:

r2 = µ/π (1+r1µ-π) It is interesting to note how the reactivity ratios depend not only on the
monomer pair and on the temperature, but also on the chemical mechanism of
For each copolymerization test, given a pair of µ and π values, it is possible to the process. Figure 35 illustrates the case of the copolymerization of styrene
define a linear correlation between r1 e r2. By plotting in the graph the different (M1) and of methyl methacrylate (M2).
lines in the reactivity ratio space, a unique intersection should be obtained at a
point with coordinates r1 and r2. Actually, an area of intersection is normally
obtained whose amplitude is proportional to the size of the experimental errors, CH3
and therefore the method is considered inaccurate. CH2 CH + CH2 C
According to the Fineman-Ross method, the copolymerization equation can be COCH3
linearized as follows:
O

[(π-1)/π]µ = (µ2/π)r1-r2

By performing at least 4 or 5 low yield experimental copolymerization tests


with different composition, [(π-1)/π]µ is plotted as a function of µ2/π thus
obtaining a straight line with slope r1 and intercept r2. Figure 35. Composition diagrams of the copolymerization of styrene (M1) and of methyl
Other more recent methods exist based on linear regression of least squares methacrylate (M2) with radical, anionic and cationic mechanism.
(for example Kelen-Tudos and subsequent variations) characterized by greater
computational complexity. Curve 1 in Figure 35 represents the diagram of the composition for cationic
copolymerization: low values of input styrene (f1) correspond to high values of
For many monomer pairs, under specific conditions of temperature, solvent styrene in the copolymer (F1). This is due to the fact that the
etc.., the reactivity ratios are already tabulated (see Polymer Handbook) and methylmethacrylate carbocation is destabilized by the presence of electron-
easy to consult. attracting ester-COOCH3 groups, and thus it is not easily formed. In these
Various methods of approximate calculation also exist that allow for an conditions r1 >> r2, and in particular r1 > 1.
estimate of the reactivity ratios without the need of experimental tests. For If the process is anionic, curve 3 is obtained, in which the situation is reversed
instance, the Alfrey-Price Q-e parameters tabulate the stereoelectronic (r2 > 1) as the formation of the methyl methacrylate stabilized carbanion is
contributions to copolymerization for various monomers. In particular, the Q favored.
parameter indicates the possible resonance stabilization of the radical, i.e. it Finally, curve 2 represents the case of radical copolymerization. Both values of
quantifies how easily it forms. On the other hand, the e parameter indicates reactivity ratio are smaller than one, and thus a predominantly alternating
the polarizability of the monomer and therefore its reactivity. The values are copolymer is obtained.

73 74
Through classical radical processes it is practically impossible to obtain of the system, after the addition of the chain extender it is usually completed
structures with preordered blocks, the latter being an interesting strategy to in hot molds.
achieve special morphologies in the material and hence specific physical and Aromatic, aliphatic or cycloaliphatic monomers (the first type is more
mechanical performance. As a matter of fact, "living" radical processes also commonly used because of its lower cost and better mechanical properties of
need to be mentioned, although they will not be discussed here. the final polymer) are used as diisocyanates; examples of aromatic isocyanates
In addition to the already mentioned example of living anionic polymerization are the toluene diisocyanate (TDI, mixture of isomers) and the methylene-
yielding regular block copolymers such as SBS (thermoplastic elastomers), it is bisphenyl diisocyanate (MDI).
worth mentioning the possibility of obtaining segmented copolymers by means
of step-growth processes, as in the case of certain types of polyesters and – NCO

more often - of polyurethanes. CH3

Polyurethanes are a very large class of polymers that have in common a basic OCN CH2 NCO
reaction between alcohol and isocyanate to give urethane
NCO

At the morphological level, segmented polyurethane copolymers obtained in


this way show a polyphasic morphology with segregation of a flexible polyester
The term polyurethane does not indicate a specific polymer but rather a family or polyether phase (macrodiol) and a rigid polar urethane phase characterized
of materials characterized by the presence of the bond -OCONH-. Depending by a high melting point.
on the choice of monomers that compose them (bifunctional or trifunctional,
monomeric or macromeric) and on the strategy with which they are combined,
polyurethanes may be used for various applications such as rigid and flexible
foams, adhesives, paints, textile fibers, thermoplastic and crosslinked rubbers
etc.
Figure 36 illustrates the synthesis of a segmented polyurethane used as
thermoplastic elastomer, and thus somehow similar to the SBS described
above.
The first step involves the synthesis of the so-called prepolymer, which is
normally obtained by adding a macrodiol to a diisocyanate. Polyester
(polyadipates) or polyether macromers (polyethylene glycol, polypropylene
glycol, polytetramethylene) with a molecular weight around 2-3000 and
hydroxylic bifunctionality are often used as macrodiols. The prepolymer,
although represented in a simplified scheme as a triblock, is in fact
characterized by a specific molecular weight distribution that can be estimated
Figure 36. Synthesis of a segmented polyurethane
with the relations already described for step-growth polyreactions.
The second step is the "chain extension", namely the addition of a second diol
Finally, a last example of a copolymerization process is represented by polymer
of low molecular weight (such as 1,4 butanediol HOCH2CH2CH2CH2OH) in a
grafting reactions, which are a common method for toughening thermoplastics
stoichiometric ratio OH/NCO = 1 to obtain the highest possible molecular
(Figure 37).
weight. The polymerization is carried out in bulk and, due to the high viscosity

75 76
On an already formed high-molecular-weight polymer it is possible to hot-mix Chapter 6: Industrial polymerization processes
peroxides (e.g. in an tween-screw extruder) with formation of active radical
species. By adding another unsaturated monomer it is possible to start a From a process point of view, it is useful to classify the different ways in which
kinetic chain with lateral grafting. In this way, it is possible to graft an a polymerization process can be carried out, as shown schematically in Figure
elastomeric polymer (e.g., polybutadiene) to a thermoplastic matrix. Also in 38.
this case a morphology with segregated phases is obtained (caused by the Processes in homogeneous phase can be carried out in bulk or in solution of an
immiscibility between polymer chains of high molecular weight), where the appropriate organic solvent; the first method is obviously to be preferred on an
presence of rubbery microinclusions increases the toughness (energy industrial scale for economic and environmental reasons. Most of these
absorption before fracture) of the polymeric material. processes are conducted in liquid phase, but there are few examples of gas
phase polymerization or even solid phase ones.
Many production processes are instead carried out in heterogeneous phase or
dispersion: it is worth mentioning suspension (including those known as slurry),
emulsion and interfacial processes.
Homogeneous processes are characterized by the presence of a single phase in
the thermodynamic sense, namely by the absence of separation interfaces
between portions of materials with different composition. Bulk and solution
processes are the most popular in this category. In these processes,
polymerizations are conducted in liquid state between melted monomers or in
the presence of a solvent, typically at high temperatures above the Tg or Tm of
the polymer which will be formed. In this way it is possible to form mostly
polycondensation polymers (bulk) and partly polyaddition polymers.
Solid state polymerization represents a very special case of polymerization
process that, from an industrial point of view, mainly concerns the upgrade
reactions (increase of the molecular weight) of technopolymers such as
polyethylene terephthalate, PET. It is typically conducted at temperatures
between the Tg and the Tm of the polymer, and is related to the reactivity of
the amorphous phase.

Figure 37. Grafting reaction for the synthesis of grafted-copolymers

77 78
disadvantages, operational difficulties and product quality. It should be
emphasized that in most cases the choice of a process rather than another is in
fact largely determined by the requirements given by the chemical nature of
the monomers and of the base reaction. For example, it will not be possible to
conduct an ionic polyaddition in aqueous phase, or a polycondensation in gas-
phase. Whenever possible, polymerizations in aqueous environment are
normally preferred.

Process Bulk Solution Interfacial Suspension Emulsion


Viscosity
dramatic high high nil modest
Increase
Gel effect dramatic high ----- high high
Heat
poor difficult good good good
exchange
Figure 38. Schematic classification of polymerization processes very
Stirring difficult difficult difficult easy
difficult
On the other hand, gas phase polymerizations include medium-high monomer Product
good fair poor good poor
pressure processes and relatively high temperature, often using N2 as inert purity
thinner and operating at low conversion with continuous recirculation of the
unreacted monomers: it essentially concerns the production of commodities
from radical and ionic polyaddition. Important parameters to be evaluated are the increase of viscosity, connected
Heterogeneous processes are instead characterized by several phases of to the increase of the molecular weight, and the gel effect typical of chain-
different composition and with well-defined separation interfaces. The most growth polymerizations, in turn connected to uncontrolled increase of
important examples are represented by suspension and emulsion processes of viscosity: these will be maximal in bulk processes, high in in solution processes
hydrophobic monomers mechanically dispersed in H2O. Different additives are (but with a dilution effect related to the quantity of solvent used), while on the
typically used to stabilize the reaction environment (especially in emulsions), contrary they will be more controllable in heterophase processes such as
as well as low or medium temperatures and vigorous mechanical stirring. They emulsion and most of all suspension polymerizations. An explanation can be
are essentially free-radicalic polyadditions since from the point of view of the given by considering that in homogeneous processes the macromolecules form
chemical mechanism such polyreactions are among the few that do not suffer a typical random coil conformation, with entanglements progressively more
from water interference. efficient with increasing molecular weight and consequent exponential increase
A particular case is represented by "slurry" processes which are - strictly of viscosity (remember that in random coil polymers for sufficiently high
speaking – suspensions, yet in an inert non-aqueous solvent (hydrocarbons). molecular weights - i.e. able to generate entanglements - the viscosity is
In such cases the catalyst is suspended for ZN ionic type polyaddition related to molecular weight by the relation η = cost. Mw3.4).
processes. In dispersion processes, the macromolecules will always be present as particle
Finally interfacial processes include some special polycondensations, and the aggregates. Under these conditions the viscosity of the system is independent
polyreaction shows an electrophile-nucleophile behavior at the interface of the molecular weight while depends only on the concentration of the
(stirred or not) between water and an organic phase. dispersed phase, as stated by Einstein’s law (η0 = viscosity of the continuous
The following table shows a qualitative comparison between some features of phase, Φ = volume fraction of the dispersed phase) and its modifications.
the general processes briefly described above, in terms of advantages /

79 80
η = η0 (1 + 2.5 Φ) The target of difficult polycondensations is instead the synthesis of high
polymers, which represent the main classes of thermoplastic technopolymers
Other important aspects of the processes that have implications on the type of such as polyesters, polyamides and polycarbonates.
equipment to be used are heat transfer and stirring efficiency. Easier disposal Among the most important thermosetting polymers (resins) are phenolics or
of exothermic heat is achieved in heterogeneous processes that also allow phenol-formaldehyde resins, obtained by reaction of phenol (C6H5-OH) with
better stirring due to a lower viscosity of the system. formaldehyde (HCHO). This is the first historical example of a fully synthetic
By contrast the purity of the polymer product will be good in the case of bulk polymer (bakelite). Nowadays, these resins are still widely used due to their
processes (except for the potential occurrence of partial thermal degradation) good mechanical properties and low costs (yet the well documented toxicity of
and rather poor in heterogeneous phase processes that require drying of the formaldehyde will certainly cause use limitations in the future). The synthetic
product and purification from the different additives (surfactants etc.) present process requires the reactor to be loaded with phenol together with an excess
in the synthesis. of formaldehyde in aqueous solution at non-neutral pH. This phase is followed
Ultimately the choice of the type of process and of the appropriate reactors by heating up to +160°C with gradual distillation of the by-product of reaction
depends on several factors including: the chemical nature of the polyreaction (H2O), since this is a polycondensation between polyfunctional monomers. At
and its own requirements (as to thermal stability, disposal of exothermic heat, the end of the process the product, namely the phenolic resin, is obtained by
need for removal of by-products etc..); the characteristics and specifications of cooling and pulverization. From the point of view of the chemical mechanism,
the product (purity, molecular weight, functionality); and of course economic formaldehyde can be considered as a bifunctional monomer (equivalent to a
factors. dimethylolmethane). On the other hand, the reactive groups in phenol are on
In principle, although with several exceptions, continuous production processes the aromatic ring because positions 2 and 6 are activated with acidic pH and 2,
have been developed on a large scale for those polymers that can be classified 4, 6 are activated with basic pH with respect to the condensation reaction.
as commodities (polyethylene, polypropylene). Conversely, discontinuous
processes have been employed for the production of technopolymers, polymer
OH
specialties and often to carry out emulsion polymerizations because of their
stability problems. Activated positions 2, 6 (acidic
pH)
Some of the main industrial processes will be hereby examined starting from
polycondensation polymers, or better, from polymers that can be obtained OH OH

from step-growth processes. Most of the processes considered are conducted H H CH2 OH
C
+
in homogeneous liquid phase (especially in bulk) or they are less frequently O
interfacial processes. Both discontinuous and continuous processes are used.
OH OH

CH2 OH CH2 CH2


Polycondensation processes H
+

It is useful to distinguish between polycondensations defined as easy, and


polycondensations defined as difficult: the difference lies primarily in the
molecular weight that has to be obtained. The reaction path is actually very complex and it is practically impossible to
Easy polycondensations mainly include the synthesis of functionalized describe comprehensively all the variously branched oligomeric species that
oligomers to be subsequently used as reactive resins in different polymer are formed. The type of resin that is obtained by running the polycondensation
formulations, such as adhesives, paints and thermoset polymers in general with an acidic pH is called novolac, and it is characterized by phenolic
(phenolics, epoxies and also polyurethanes, although the latter are step- oligomers linked by CH2- methylene bridges. To obtain the thermosetting
growth polyaddition polymers). material the use of an external crosslinker is necessary (see chapter 7).

81 82
Conversely, the products obtained at basic pH are called resols; this reaction
mixture is constituted by partly oligomerized phenols substituted with methylol
groups. Resols are able to crosslink by self-heating with formation of
methylene and dibenzylether bridges and elimination of H2O, which is the by-
product of the polycondensation. In reality, more or less relevant emissions of
formaldehyde are also always present.

Figure 39. An autoclave for the synthesis of phenolic resins

An example of a polycondensation process for the synthesis of high polymers is


represented by the production of PET (polyethylene terephthalate). That
polyreaction can be represented by the balance between reagents (terephthalic
Figure 39 shows an autoclave that can be used as a reactor for the synthesis of
acid and ethylene glycol) and products such as the one shown below.
phenolic resins. It is a jacketed stirred reactor, capable of operating at
moderately high pressure (to raise the boiling point of H2O up to the
temperature required for the completion of the polycondensation), equipped
with a reflux condenser for the collection and quantification of the reaction by-
product.

As in other polycondensation processes with bifunctional monomers there are


several critical parameters for obtaining high molecular weights, including the
difficulty in achieving extent of reaction equal 1, the control of the
stoichiometry and the purity (degree of bifunctionality) of the reagents.
The so called DMT process (dimethyl terephthalate) is based on
transesterification reactions and has been historically developed to overcome

83 84
the low purity of the industrially available terephthalic acid. Today, however,
PET is obtainable mainly from the direct polyesterification of terephthalic acid.

Figure 41. Transesterification in the DMT process


The DMT process consists of two distinct phases, known as "monomer
synthesis" and "polycondensation".
In the first step (Figure 40), dimethylterephthalate is allowed to react in bulk The advantage of this process is that it allows to bypass the constraint of the
with a variable excess (2-4 times) of ethylene glycol, in the presence of metal exact stoichiometric balance between the reactants needed for obtaining high
catalysts (Zn, Co, Mn acetates) that favor the opening reaction of the methyl molecular weights. In fact the system is self-regulated by progressively
ester. The temperature is allowed to increase from +160°C to +190°C, with eliminating the more volatile fractions. The molecular weights that can be
distillation of the methanol by-product. Therefore, the so-called “monomer” is obtained are never very high: typically it is possible to extrude from the
actually an adduct formed by an average of two molecules of glycol and one of polycondensation reactor raw PET with Mn of approximately 20000. This limit is
terephthalate. In reality, a mixture of oligomeric species is formed which is imposed by the high viscosity of the melt which cannot be lowered by further
characterized by a relatively broad polydispersity. heating as it is already close to the thermal stability limit of the polymer. This
grade is suitable for various applications, such as for instance spinning - in the
field of textiles. However, the molecular weight is too low to allow
transformation of the polymer for example into bottles for blow molding. Since
under these conditions PET has not yet reached an extent of reaction equal to
1, i.e. some reactive terminals are still present, it is possible to perform an
additional post-polymerization phase in the solid state (directly on the
extruded granules, in stoves with forced ventilation in dry nitrogen) at a
temperature between the Tg (+80°C) and Tm (+250°C) of PET. In these
Figure 40. Synthesis of the “monomer” in the DMT process conditions the amorphous interphases regain thermal mobility and are able to
complete the polymerization, reaching Mn of up to around 40000.
In the second step (Figure 41) the catalyst (antimony oxide Sb2O3) is changed The bulk polymerization process usually leads to products with good purity.
and the temperature is gradually increased to +260 / 290°C, i.e. brought However, in this specific case potential problems may arise linked to the
above the melting point of PET. It is necessary to progressively apply vacuum possible presence of traces of toxic compounds (diethylene glycol and
(down to 1 mbar) to distill the high-boiling point glycol. In practice, during the acetaldehyde), and the presence of residues of –COOH terminal groups that
polycondensation phase the molecular weight of the system increases because are potential sites of thermal and hydrolytic instability of the polymer.
the catalyst allows the formation of a dynamic equilibrium regime (continuous
openings and closures of the ester bonds) that is perturbed by the distillation A simplified schematic diagram of a plant for the production of PET is shown in
of the by-product. Figure 42. In the heat exchange reactor the synthesis of the monomer takes
The process is mildly exothermic and is carried out in bulk typically by place, after which the product is progressively transferred to a number of
controlling the viscosity of the reaction system (that is proportional to the autoclaves placed in series where the actual polycondensation takes place. The
molecular weight). complexity of the process - caused by the high viscosity and consequently by
stirring, heat and mass exchange (distillation) problems - grows as the

85 86
polyreaction advances. The last reactors are actually extruders-reactors where
stirring is ensured by the continuous motion of a screw until the final product is
extruded in the form of pellets.
The polyreaction is controlled by means of transducers that are sensitive to
torque variations, which are related to the increase in viscosity. The possibility
of connecting in series a number of reactors makes it possible to manage the
process as a semi-continuous process (need for high production volumes
Figure 43. Formation of the salt between hexamethylene diamine and adipic acid
because of the low cost of PET).

As opposed to the previous case, here it is necessary to operate by


progressively increasing the pressure in the autoclave (up to 15-20 bar) to
raise the boiling point of water which is the by-product of the reaction.

Figure 44. Polyamidation reaction


Figure 42. A plant scheme for the industrial production of PET

Figure 45 also shows a simplified schematic diagram of a plant for the


Another important example of industrial polycondensation process is the
production of Nylon 6,6.
production of polyamide 6,6 (or nylon 6,6). The monomers are hexamethylene
diamine and adipic acid, and the process is known as "salt dehydration".

O O

HO C CH2 C OH H2N CH2 NH2 Hexamethylene


4 Adipic acid 6
diamine

The first step takes place in aqueous phase (where both monomers are
soluble), and the stoichiometry of the monomers is self-controlled by feeding
the two monomer solutions in H2O until the pH = 7.6, corresponding to the
formation of a double inner salt as shown in figure 43.
At this point the polyamidation process begins (Figure 44), taking place in an
autoclave under stirring and at temperatures increasing from +180° up to
+280°C, i.e. still above the Tm of the polymer.
Figure 45. Schematic diagram of a plant for the industrial production of Polyamide 6,6

87 88
Finally, a third important example is the interfacial polymerization process for phases as this is formed (as for some polyamides in conventional lab-scale
the production of polycarbonate PC. It is a type of process suitable for very experiments, see below)
fast, non-equilibrium polycondensations, based on polyreactions between very
reactive nucleophilic and electrophilic groups. Examples of this type of reaction
are the reaction between amino-terminals –NH2 and acyl chlorides -COCl, or
the reaction between alcoholates –O- (such as phenates) and again acyl
chlorides. In practice, it is possible to obtain mostly polycarbonates and
polyamides.

O
O
O M+ + C O C + M+ Cl −
Cl

The term nucleophilic identifies a group with an availability of electronic


The scheme for the interfacial process for the production of PC is shown in
doublets, which thus "looks for" positively polarized centres to form bonds. In
Figure 46. In the first step the bisphenol A monomer is dissolved in alkaline
contrast, electrophilic groups are groups depleted of electrons and thus
H2O by allowing the two -OH groups to react with soda. The second monomer,
positively polarized (as carbon acyl chloride), who "are sought" by negatively
phosgene (a gas) is instead dissolved in a solvent such as dichloromethane,
polarized centres to form bonds.
and the two phases are put in contact under vigorous stirring. The polymer
The interfacial processes present certain unique characteristics: they are
dissolves in the solvent, while the by-product NaCl dissolves in H2O and leaves
heterophasic polymerizations that take place at the interface between solvent
the reaction phase. Some excess of bisphenol is tolerated, as this can self-
(organic phase) and H2O and the polyreaction can occur even at low
regulate by diffusing into the organic phase.
temperatures. In addition, such polyreactions take place at the interface via a
It is interesting to note how the three polycondensation processes used for the
diffusive control so that the exact stoichiometric control of the reagents is not
production of PET, PA6,6 and PC have developed different strategies all aimed
necessary.
at overcoming the constraint of the exact stoichiometric balance between
Catalysts or auxiliary substances may be required for the completion of the
reactants.
polyreaction, such as alkaline agents to capture the acidity generated as a by-
product, or alkyl-chain quaternary ammonium salts that serve as phase
transfer agents coordinating to the nucleophilic reagent (the phenate dissolved
in basic water, for example) thus increasing its solubility in the organic phase.

+
O Na + + NR 4+ Br − O NR 4 + NaBr

where R = alkyl C4/C18 solubility in organic phase increases

When the polymer is soluble in the solvent (as in the case of the PC process)
stirred reactors can be used. On the contrary, if the polymer is insoluble it is
necessary to progressively eliminate the polymer from the interface between
Figure 46. Scheme for the synthesis of polycarbonate by interfacial polymerization

89 90
Free-radical processes
Figure 47 shows the plant diagram used for the production of PC. Also in this Free-radical polymerization processes may be conducted in both homogeneous
case more reactors are placed in series to increase the production capacity. and heterogeneous phase.
Centrifugation is also required to separate the organic and aqueous phases. As Polymerizations and copolymerizations in the homogeneous liquid state may be
a general rule, the operations necessary for polymer purification are much carried out with monomers such as styrene and methyl methacrylate, while
more laborious than in the previous cases. ethylene can be polymerized at high pressure (100-300 MPa) and temperature
(300°C), to obtain high-density polyethylene (HDPE). Such processes can be
both continuous and batch, and are typically operated at medium-low
polymerization conversions with reuse of the unreacted monomer.
In general, free-radical processes allow poor control on the macromolecular
architecture, and mostly atactic and branched polymers are typically obtained.
However in the case of symmetric monomers such as ethylene,
tetrafluoroethylene CF2=CF2, or vinylidene fluoride CF2=CH2 it is possible to
obtain partially crystalline polymers.
In homogeneous phase processes, polymerizations can also be conducted in
solvent: under these conditions chain transfer to the solvent is very significant
and medium-low molecular weight polymers are obtained that can be used as
resins in the field of paints and adhesives.
Free-radical polymerization processes in heterogeneous phase (aqueous
Figure 47. Scheme of a plant for the production of PC dispersion) are technologically important. They are often used because the
presence of water is compatible with the chemistry of free-radical reactions.
The obvious advantages of processes conducted in aqueous dispersion reside
in the typical characteristics of H2O: non-flammability, non-toxicity, low cost
and the ability to dispose of exothermal heat.
Suspension processes are based on the formation of drops of monomer
(hydrophobic) mechanically dispersed by stirring in water; the size of these
drops can be of the order of 10-2-101 mm. Colloidal stabilizers such as
phosphates or polyvinyl alcohol (PVA) are often used to prevent the
coagulation of these drops. Peroxy initiators - such as those commonly used in
bulk or solution polymerizations – are often used: the peroxide is hydrophobic
and dissolves within the drops of monomer, where a thermal polymerization
begins, entirely analogous to that typical of a homogeneous process. In
practice, the system consists of a set of microreactors dispersed in an aqueous
medium which allows to efficiently dispose of exothermal heat. The entire
polymerization process includes a number of other operations in addition to the
actual polyreaction. Among these are filtration, washing of the granules, and
their final drying.

91 92
Industrial polymers obtainable by processes of this type are polystyrene,
polyvinylchloride, polymethylmethacrylate, polyvinyl acetate,
polytetrafluoroethylene.
The process of PVC production in aqueous suspension is a good example. It
consists of three main phases: the polymerization of vinyl chloride monomer
(VCM); the stripping of the non-converted monomer; the final drying. Since
the monomer is carcinogenic (unlike the final polymer, which is safe),
technologies have been developed allowing the whole process to be carried out
"in a closed environment", and with a highly pure final product. The process is
conducted at temperatures between +50 and +75°C in conditions of liquefied
monomer at pressures of 8-14 atmospheres. Like all free-radical
polymerizations the process is strongly exothermic (360 Kcal / Kg at 60°C). A
Figure 48. Scheme of a plant for the industrial production of PVC in suspension
peculiarity of the PVC product consists in the fact that the polymer is insoluble
in its monomer, while the monomer has a solubility of about 25% by weight in
The diagram in Figure 49 illustrates temperature and pressure trends during
the produced polymer. These aspects influence both the macroscopic
the process of PVC suspension polymerization. At first, the reactor jacket
morphology of the PVC obtained (coarse and porous granules), and the
temperature is brought to 90°C, while simultaneously the pressure inside the
purification procedures of the polymer which must be very efficient because
reactor rises up to 1000 kPa and the internal temperature rises to 60°C. In
the level of residual monomer in the polymer must be below the ppm range.
these conditions the polymerization starts, and the heat generation must be
Figure 48 shows in a simplified way a PVC production plant. VCM, H2O and
counterbalanced by a progressive decrease of the external temperature down
various additives are loaded in a first reactor where the polymer synthesis
to 20°C or even less. In the last phase of the reaction, the pressure decrease
reaction takes place. The product is in the form of polymer granules suspended
indicates that the solubility limit of VCM in PVC is reached, therefore a new
in water. The suspension is sent to the stripping column, where the unreacted
increase of the external temperature must be ensured.
monomer is removed by countercurrent steam (the suspension progressively
goes through a series of sieve trays). The produced PVC granules are then
centrifuged and dried.
Because - as mentioned above - the polymer is insoluble in the monomer, the
morphology of the granule formed is typically incoherent and resulting from
the aggregation of submicron particles that correspond to the polymer
synthesized and phase segregated within the liquefied monomer drop. The
porosity of the raw PVC granules is at the base of the plasticization technology,
which consists in the absorption of quantities (that can be relevant) of natural
or synthetic oils added with the purpose of making the material more flexible.

Figure 49. Trend of temperature and pressure during the process of PVC suspension
polymerization

Emulsion polymerization processes are substantially different, and their


peculiarities and technological importance will be treated separately.

93 94
Emulsion processes The hydrophobic monomer will be poorly soluble in H2O and it will be present
A polymer emulsion or latex is a colloidal dispersion of particles in H2O, with in the system both in the form of mechanically dispersed macroscopic drops
diameters of 100÷300 nm, and with a concentration of the order of 1015 cm-3. and as molecules solubilised within the micelles.
These particles consist of aggregates of macromolecules, typically hydrophobic Figure 51 shows the various phases of an emulsion process.
but stabilized by the presence of surfactants.
The surfactant is a molecule consisting of a hydrophilic-polar head and a
hydrophobic-lipophilic tail; sodium lauryl sulphate is an example of a surfactant
commonly used in many industrial polymerizations.

CH3-(CH2)11-OSO3-Na+

Emulsion polymerizations are related to processes of free-radical chain


polymerization and copolymerization, and they have several advantages
compared to processes in homogeneous phase and in suspension: low viscosity
independent of the molecular weight and non-flammability (as in suspension),
possibility of increasing the molecular weight without altering the
polymerization kinetics (as opposed to processes in bulk and suspension).
The recipe of an emulsion polymerization can be quite complex and involves
adding in H2O, under vigorous mechanical stirring: hydrophobic monomers
(30-50%), surfactant (1-5%), water-soluble initiator (0.01-0.5%), other
colloidal components such as stabilizers, buffers, and optionally chain transfer Figure 51. The various phases of an emulsion polymerization process
agents.
The core of emulsion technology is the role of surfactants. The surfactant In phase a) a water-soluble initiator, such as hydrogen peroxide, ammonium
partly dissolves in H2O, partly adsorbs on the different system interphases, and persulfate or more often a redox couple, is added to the system. The
partly - once it reaches the critical micelle concentration (cmc, which depends hydrophilic radical fragments react with the fraction of monomer dissolved in
on the type of surfactant) – it forms aggregates called micelles. Such micelles H2O and begin to form oligomers, which rapidly become insoluble and migrate
consist of a core of lipophilic chains and a shell with the hydrophilic polar heads into the micelle, whose specific surface is larger by several orders of
oriented towards the exterior (as shown in Figure 50), containing monomer magnitude than that of the monomer droplets. Inside the micelle, radicals find
molecules and with a diameter of the order of 50 Angstroms. a higher concentration of monomer molecules and the actual propagation
reaction starts off. With this mechanism, also known as micellar or
heterogeneous nucleation, the micelles progressively become polymer particles,
and they grow in size (up to 102 nm) and number (indicated with Np). The
mechanism of micellar nucleation is characteristic of the early stages of the
process, up to conversions of 10-15%. Moreover, there is also a parallel
mechanism of homogeneous nucleation according to which the oligomeric
radical fragments self-combine to generate the first polymer particles; such
Figure 50. Formation of a micelle mechanism can prevail in the case of processes with little amount of surfactant.
In phase b) of the process, Np is constant and the particles increase in size at
the expense of the monomer droplets (conversion 30-40%). Finally in phase c)

95 96
the polymerization is completed with the consumption of the monomer
remaining in the particle. In this phase, the behavior is not dissimilar to that of
a suspension process and a gel-effect can take place. Figure 52 shows the
qualitative behavior of conversion and reaction rate as a function of time for a
typical emulsion process.

Figure 53. Schematic representation of the Smith-Ewart equations. The circles indicate the
class of particles and the arrows the kinetic event

Let us indicate the following quantities with the dimensions of a number/unit of


time,:
ρ = input rate of the active chains from the H2O phase
k = rate of desorption or monomolecular termination
c = rate of bimolecular termination
Figure 52. Trend of conversion (X) and reaction rate (dX / dt) as a function of time in an The meaning of figure 53 is the following:
emulsion polymerization. - the Nn particles increase because of bimolecular termination of the n+2
species with rate c, desorption from the n+1 species at rate k, and entrance in
Emulsion free-radical polymerization processes are peculiar from the point of the n-1 species with rate ρ. All these contributions are positive in the general
view of the kinetic model. Limiting the analysis to the propagation phase, the equation.
rate of reaction as in all free-radical processes can be expressed as - the Nn particles decrease because of bimolecular termination to yield n+2
species with rate c, because of monomolecular termination to yield n+1
v = K2 [M] [M*] species at rate k, and because of entrance in n+1 species with rate ρ. In these
cases, the terms will be negative.
The expression of [M*] was obtained from the mass balance equations derived The mathematical expression of the Smith-Ewart population balance equation
by Smith and Ewart, that consider the effect of all reactions such as radical under steady-state regime is:
entrance, desorption (monomolecular termination), bimolecular termination
(by combination and transfer) on the number Nn of particles containing on dNn/dt = ρ(Nn-1–Nn) +k [(n+1)Nn+1–nNn]+ c[(n+2)(n+1)Nn+2 – n(n-
average n active chains. From a schematical point of view, the meaning of the 1)Nn]=0
Smith-Ewart equations is shown in Figure 53.

After solving the system of differential equations, the expression of the


reaction rate for free-radical emulsion polymerizations is obtained, which can
be written in the following simple form:

v = K2 [M] (Np n)/NA

97 98
where NA is the Avogadro’s number. Unlike the analogous relation valid for progressively varying the composition of the input monomer mixture and
free-radical processes in homogeneous phase or suspension, this kinetic enriching it progressively with more reactive monomer (i.e. that is consumed
expression does not include the initiator concentration term [I]; the rate of faster). Useful relationships are obtained by taking the general
polymerization in emulsion processes depends mainly on the concentration of copolymerization equation and setting as boundary conditions the invariance of
monomer and on the number of Np particles, in turn related to the efficiency of the composition of the copolymer Π with respect to time or yield.
the surfactant. An important example of a radical emulsion copolymerization process is the
The average number of active chains per particle n tends to the value 0.5 synthesis of SBR (styrene-butadiene, Figure 54).
when the termination reactions are faster than the input ones. In fact, when
the first radical fragment enters the micelle the polymerization starts, while the
kinetic chain is terminated when the second radical enters. In practice, each
particle contains a radical species for about half of the total polymerization
time. Therefore, when conducting an emulsion polymerization with a system
dominated by bimolecular terminations, none or one radical will always be
present in each particle, but the rate will still be high because of the large
number of particles. The parameter that determines the length of the kinetic
chain will be the input rate ρ. By calibrating the rate of initiator decomposition
it is possible to adjust the molecular weight, and with low values of ρ very high
molecular weights are obtained.
Once the actual polymerization is finished the polymer latexes can be directly
Figure 54. Emulsion polymerization of the SBR copolymer
used as coatings, paints, adhesives. On the other hand, in several other cases
such as for thermoplastics and many types of rubber, it is necessary to isolate
The recipe for emulsion polymerization can be very complex. The following
and purify the polymer from the reaction environment. Latexes are
table describes the components of the “cold” (i.e. carried out al low
electrostatically stabilized by the electrical double layer of surfactant
temperature) emulsion process of SBR rubber: in this case poorly branched
surrounding the single polymer particle. To obtain the coagulation of the latex
linear copolymers are obtained, with improved processability characteristics in
the repulsive interactions between particles must be overcome through
the compound.
acidification or addition of soluble salts. Then the produced polymer is washed
from the residual monomer, surfactant, initiator, and finally dried out. In this
way a wide range of polymeric materials can be obtained, such as PVC (with a
finer morphology than that achieved in the suspension process), many
copolymers such as ABS (acrylonitrile-butadiene-styrene), synthetic SBR
rubbers (styrene-butadiene rubbers), polyacrylates, and also polymer
specialties such as most of fluoropolymers: polyvinylidene fluoride (PVDF) and
its copolymers based on monomers such as vinylidene fluoride,
tetrafluoroethylene, hexafluoropropene.
Emulsion processes are widely used in the synthesis of copolymers. For this
type of processes, the main issue is the control of the composition of the
obtained copolymer. In principle, this can be achieved by running the
polymerization at low conversion, and continuously recycling the unreacted In addition to monomers, water and surfactant, the presence of mercaptan
monomers. Otherwise one can operate with semicontinuous processes, allows control of the molecular weight (by means of a chain transfer

99 100
mechanism). The initiator is a redox couple consisting of a hydroperoxide and butene (to lower the degree of crystallinity and strengthen the material), and
iron sulphate, with the addition of EDTA (ethylenediamine tetraacetic acid) to finally heterophasic copolymers where an ethylene-propylene-based rubbery
increase the solubility of the iron salts in water. The process temperature is phase is grafted to the semi-crystalline polypropylene matrix (impact
around +5°C, and conversions of about 60% are obtained. copolymers), always with the aim of strengthening the material.
Figure 55 shows the plant diagram for the production of SBR in emulsion. Different industrial processes have been developed, such as suspension
Monomers, initiator (3, improperly called catalyst) and surfactant are processes of inert hydrocarbons (slurry processes), bulk processes (liquefied
measured and mixed in polymerization reactors (pressure autoclaves) placed monomer) and gas phase processes, as schematically summarized in Figure 56.
in series (5): the transfer of the latex from one reactor to another takes place
depending on the process conversion. In reactor 6 the termination occurs by
means of chemical deactivation of the radicals. The recovery of the butadiene,
which under normal environmental conditions is a gas, can take place simply
by depressurizing the autoclave. Styrene is instead a relatively highly-boiling
liquid and must be recovered by distillation in vacuum. The next phases
include: collection of the purified latex; its coagulation by addition of acids or
salts; drying; packaging of raw rubber slabs or bales.

Figure 56. Industrial grades of isotactic polypropylene and main production processes

Because of the very high production volumes and the low cost of the material,
the production processes of polypropylene (as well as of other commodities
such as polyethylene) are typically continuous processes, strongly integrated
with the oil processing cycle and with high productivity (plants producing more
than 150,000 tons/year).
Slurry processes are the oldest. They employ insoluble catalysts and take place
in mild conditions (60-80 °C, 5-15 atm), with mechanical stirring and inert
diluents such as anhydrous hexane. The most advanced process conditions in
Figure 55. Scheme of a plant for the production of SBR latexes
this field are represented by the Spheripol™ process, which employs liquid
Processes with ZN catalysis monomers and gas phase post-reactions in order to obtain impact copolymers.
The case of isotactic polypropylene and stereospecific polymerization is herby The plant scheme is shown in Figure 57: the homopolymerization of propylene
described, as an important example of industrial production based on linear in the presence of hydrogen as moderator of molecular weight is carried out
scheme ionic polymerization processes. continuously in tubular reactors, where the catalyst and liquid
Different types of polypropylene exist on the market, in particular monomer/polymer slurry is fed by means of a propeller pump. Cooling is
homopolymers with different molecular weight distribution (monomodal, ensured by the reactor jacket.
bimodal etc.), statistical copolymers with monomers such as ethylene or

101 102
Chapter 7 - Crosslinking

Crosslinking consists in the formation of covalent bonds between distinct


macromolecules, with the generation of a single three-dimensional network
(Figure 58).
Crosslinking takes place when the average functionality of the monomer
mixture is greater than two, i.e. when at least a fraction of the monomers is
trifunctional, tetrafunctional etc.
Crosslinked polymers present some unique characteristics, such as insolubility
in any type of solvent and impossibility to melt. They are known as
thermosetting polymers. On the one hand this involves the improvement of
some performance characteristics (such as chemical resistance and high-
Figure 57. Plant scheme of the Spheripol process for the production of PP
temperature mechanical properties), yet on the other hand it hinders the
processability of the material: in a thermosetting polymer the shaping of an
The subsequent copolymerization with grafting of ethylene-propylene rubber is
item takes place together with the completion of a chemical polyreaction, thus
conducted in a fluidized bed reactor by adsorption of monomer gases in the
being irreversible.
synthesized polypropylene particles. This second step is carried out at 70°C,
with monomer pressures of up to 4 MPa. It is possible to obtain copolymers
with a rubber content of up to 40%. The obtained material is polyphasic
because of the thermodynamic immiscibility between the semi-crystalline
continuous phase (isotactic polypropylene) and the amorphous dispersed
phase (ethylene-propylene statistical copolymer); the obtained morphologies
are finer and more stable than those obtained by direct mixing of
thermoplastic and rubber in extruder.
In the Spheripol™ process the polymeric product is in the shape of spherical
particles. This shape replicates the morphology of the catalyst used. The high
yield - both in terms of conversion and stereochemical purity - allows for a
polymer production without the need for removal of the catalyst and of the Figure 58. Three-dimensional network
atactic fraction.
A mild chemical crosslinking is also typical of rubber technology. It is
commonly called vulcanization: in this case crosslinking prevents viscous flow
at high-temperature and high deformations, thus allowing elastic recovery.
The basic difference between the crosslinking characteristics of thermosetting
polymers and rubbers does not lie in the chemical mechanism but rather in the
crosslinking density of the material, defined as the number of chemical nodes
per volume unit. Thermosetting polymers are characterized by high
crosslinking density and thus by short chain segments between nodes, while
rubbers have a low crosslinking density and a high molecular weight between

103 104
nodes (of the order of 103-104), the latter being necessary to preserve the
average random coil conformation.
Several methods exist to perform a crosslinking reaction. A first method
(Figure 59) consists in conducting a polyreaction of a mixture of bi and tri (or
poly) functional monomers, by means of a thermally activated process or by
supplying light energy.

Figure 60. Crosslinking of a polyurethane formulation based on toluene diisocyanate and


trimethylolpropane

From a technological point of view the “pot-life” is defined as the time during
which the formulation can be applied before the viscosity becomes too high.
This parameter is related to the achievement of a critical extent of reaction for
the polyreaction, as will be explained shortly.
Figure 59. Crosslinking process by means of trifunctional monomers Conversely, in single-component systems the viscosity of the formulation is
kept low until heat or light are supplied: phenolic resins are a good example of
Many important industrial cases are represented by polyfunctional step-growth this class.
polymerizations, including phenolic resins (phenol-formaldehyde), epoxy resins The gel point of a thermosetting formulation can be experimentally estimated
and polyurethane resins, while among polyfunctional chain-growth as the time at which the viscosity of the system tends to infinity. A simple
polymerizations it is worth mentioning unsaturated polyesters. experiment targeted to its determination consists in insufflating inert gas into
Figure 60 illustrates the crosslinking of a bicomponent polyurethane the reaction mixture and observe when the bubbles stop ascending in the fluid
formulation based on toluene diisocyanate and trimethylolpropane. The term mass. More sophisticated methods are based on dynamic-rheological
bicomponent is used because once the polyol and the isocyanate are brought measurements.
into contact, the system begins to react even at room temperature. As a From a theoretical point of view it is very useful to be able to estimate, at least
consequence, branching occurs together with a progressive increase in approximately, the gel time of a thermosetting formulation, in order to
viscosity until gelation is reached, a critical phenomenon that describes the evaluate the processability time of the product.
transition from a fluid state (sol) to an elastic gel state in which flow is In order to do so, polyfunctional step-growth polyreactions will be examined as
inhibited. a general case. Let us consider the general case of a polyreaction with three
monomers of the type A2 + A3 + B2, in which the chemistry of the system
allows A to react only with B and vice versa (for example, hydroxyl-OH with
isocyanate-NCO), something which – by the way - is not always true.

A
A A + A + B B
A

105 106
A
A first estimation method is quite simple and is based on the Carothers A A + A + B B
equation. By defining the extent of reaction p as the ratio between the number A
of reacted groups and the number of initial groups, with N0 and N being the A
number of initial and current molecules, and with fav being the average A A
functionality of the monomer mixture (total number of functions with respect A B B A A B BA
to the total number of molecules, by definition greater than 2), the Carothers A i
equation states that

Figure 61. Definition of degree of branching α (i = 0 ÷ ∞)


p = 2 (N0 – N)/ (N0 fav)

At this point the problem is to calculate the probability that at a certain point of
since the number of reacted groups is twice as much as the reacted molecules
the polyreaction a hypothetical molecule extracted from the reactor has
(A reacts with B), and N0 fav is the number of initial groups. The extent of
precisely the structure set forth above.
reaction at gel point pc is thus expressed as
With:

p = (2/f av) (N0-N)/N0 = 2/f av (1 – N/N0) = 2/ f av (1 – 1/Xn)


pA = extent of reaction of A, i.e. in statistical terms the probability that A has
reacted
The critical extent of reaction is thus pc = 2/f av in the limit case of molecular
pB = probability that B has reacted
weight tending to infinity. In reality, gelation always occurs to an extent of
ρ = the fraction of A groups belonging to A3 with respect to all A groups
reaction lower than pc calculated according to Carothers, because physically
present (attention, it is different from Carothers’ fav).
the system loses the fluid characteristics typical of liquid before the molecular
weight is infinity, i.e. when the polymer is still characterized by the presence of
the probability that B reacted with A3 is pB ρ, while the probability of reaction
a sol fraction (soluble and extractable with solvent).
with A2 is pB(1-ρ).
A more complex approach is the statistical one proposed by Flory and
Given the following sequence, it is possible to calculate the probability that this
Stockmayer. According to this model two assumptions are made: the equal
occurs.
reactivity of functional groups (regardless of their position and the degree of
The probability of a sequence of length i is thus:
branching of the macromolecule); the absence of cyclization reactions. The
latter hypothesis is actually quite unrealistic and, as will be shown, it is the
pA [pB (1-ρ) pA]i pB ρ
source of the systematic error generated by the theory of Flory-Stockmayer.
It is necessary to introduce a branching parameter α, defined as the probability
The same probability extended to all sequences of any length (i = 0÷∞) is
that a functional group A belonging to a specific branching (i.e. coming from
A3) is connected through bifunctional connections of the AB type to a second
α = ∑ [pA pB (1-ρ)]i pA pB ρ = pA pB ρ ∑ [pA pB (1-ρ)]i
branching unit A3. Visually, the situation is shown in Figure 61.

The above summation is a geometric series with argument X of the type ∑ Xk,
where if the argument is <1 (as in the present case) it converges to the value
1/(1-X), thus giving

α = pA pB ρ / [1-pApB (1-ρ)]

107 108
By introducing the stoichiometric ratio r, i.e. pB = r pA, the expression of the Naming i-1, i, i +1 etc. subsequent generations of branching chains, it is
branching parameter can be rewritten as possible to calculate the ratio between trifunctional units A3 of 2 successive
generations of macromolecular chains as Yi+1 / Yi = 2α (where Y is the
α = r pA2 ρ / [1-r pA2 (1-ρ)] branched termination of the chain). Since the branching parameter gives the
α = pB2 ρ / [r-pB2 (1-ρ)] probability of finding two A3 branches at the ends of a chain, the relationship
written above expresses the fact that the chains in the next generation are
The two above equations are useful for calculating the branching parameter twice as likely to end with a branching compared to those of the previous
from experimental data, as r and ρ are defined through the stoichiometry of generation. The condition for crosslinking can be expressed in statistical terms
the reaction mixture, while pA and pB may be quantified using appropriate as Yi+1/Yi > 1, i.e. the polyfunctional polyreaction must be able to move
analytical methods (chemical titration or spectroscopy). A particularly simple forward. This means that α > ½ for crosslinking and therefore that the critical
case is given by formulations where A = B (stoichiometric equivalence) and degree of branching is αc = ½.
hence r = 1 and pA = pB = p. In this case Generalizing to systems with any polyfunctional monomer of the type Af (with f
= average functionality), the expression can be rewritten as
α = p2 ρ / [1-p2 (1-ρ)]
αc = 1 / (f-1)
At this point what remains to be determined are the critical conditions of
gelation, i.e. appropriate expressions for αc and pc. and therefore the critical degree of branching will depend only on the
During the polyfunctional polyreaction of the system A2 + A3 + B2, each functionality of the polyfunctional monomer.
macromolecular chain which ends with a unit A3 is expected to generate 2 Finally, by substituting this relationship in the equation that correlates p and α,
chains through the subsequent reaction, then 4, then 8 and so on until a useful expression for the calculation of the critical extent of reaction of the
crosslinking occurs. All with a branching probability given by α system can be obtained. In fact, if at the gel point pAc = pBc = pc
The situation can be illustrated as in figure 62. (stoichiometric ratio = 1 between reagents), the following can be written

αc = 1 / (f-1) = pc2 ρ / [1-pc2 (1-ρ)]

And thus, solving the second order equation, the following is obtained

pc = [1 + ρ(f-2)]-1/2

As already mentioned, it is systematically observed that: pc (Flory-Stockmayer) < pc


reale < pc (Carothers). The error present in the statistical theory is neglecting the

possibility of cyclization reactions, which instead are always present to some


extent.
The calculation of pc according to the statistical theory and according to the
Carothers equation allows anyhow to estimate fairly accurately the range
within which the critical conditions of gelation occur.

Figure 62. Branching growth during successive generations of macromolecular chains


Let us consider now – from the point of view of the chemical mechanism – the
crosslinking reactions of the main types of thermosetting polymers used in the

109 110
production of composite materials, i.e. phenolic, unsaturated polyester and
epoxy resins.
Phenolic resins or phenol-formaldehyde resins are obtained by reaction
between phenol C6H5-OH and formaldehyde HCHO in non-neutral pH conditions
(Figure 63), as already described in chapter 6. For the purposes of the
polyreaction, the phenol is trifunctional and the reactive positions are the two
in meta and the one in para on the aromatic ring, while formaldehyde is
bifunctional and it is therefore similar to dimethylol methane HOCH2OH.

Figure 64. Structure of the crosslinked phenolic polymer

Phenolic resins allow to obtain very high crosslinking density and thus very
Figure 63. Synthesis of phenol-formaldehyde good high-temperature mechanical and thermovolumetric properties. However,
the emission of formaldehyde (carcinogenic) during the processing of the
The reaction of phenol with excess of HCHO with basic pH generates resols, material will determine an increasing restriction of use.
which are self-crosslinking resins. The reaction of phenol with formaldehyde in The name “unsaturated polyesters” refers to a series of oligomeric resins
defect at acid pH generates novolac resins, which instead require an external obtained by polyreaction of glycols with maleic anhydride (Figure 65). To
crosslinker to give the thermosetting polymer. Hexamethylenetetramine modulate the viscosity they are often diluted with styrene monomer, which
(CH2)6N4 is often used as crosslinking agent. Crosslinking mainly occurs with acts as a reactive diluent. The crosslinking mechanism is in fact of radical-type
the formation of methylene bridges and elimination of ammonia as a by- and it involves the in-chain double bonds of the oligomers and the vinyl ones in
product. styrene. Peroxides or redox couples such as hydroperoxides and cobalt salts
In both cases, mixtures of species are obtained that are different both in are used as initiators, allowing crosslinking even at room temperature. The
structure and in molecular weight. Figure 64 shows the simplified structure of radical process involves the same phases described for linear scheme
the crosslinked polymer. polymerizations; for the purpose of crosslinking, the useful step is termination
by combination.

111 112
trifunctional molecule, although the actual average functionality is in fact
between 3 and 6 and also depends on the crosslinking conditions (temperature,
catalysis).

Figure 65. Synthesis of unsaturated polyester resins

The most widespread epoxy resins are oligomers obtained by polycondensation


of bisphenol A (the same monomer used for the synthesis of polycarbonate)
and epichlorohydrin as shown in Figure 66. They can be crosslinked with
various reagents including anhydrides and polyamines, which are the most
commonly used. In the latter case, bicomponent formulations are obtained
that can be applied even at room temperature like in the case of polyurethanes.

Figure 67. Crosslinking of an OH-functionalized polymer with melamines

Finally, let us consider the process of vulcanization of rubbers or elastomers.


Vulcanization corresponds to the formation of chemical nodes (permanent)
between macromolecules, introduced to avoid viscous flow under stress. This
allows the rubber to store elastic energy with no dissipation phenomena due to
chain flow, which would hinder the complete recovery of the deformation once
the stress is removed. The crosslinking density obtained is low because the Tg
value of the polymer needs to be kept low. Vulcanization will be described in
Part B of these notes (Rubber Science and Technology).

Figure 66. Synthesis of epoxy resins and mechanism of crosslinking with polyamines

Another crosslinking method is the polyreaction between polyfunctionalized


polymers and external crosslinker molecules. This approach allows for the
development of both monocomponent and bicomponent formulations. It is
widely used for paints. An example is represented by the crosslinking of
copolymers with side -OH functionality (often polyacrylates or polyester
oligomers) with melamines. In particular, hexamethoxymelamine can be
considered as an hexafunctional resin for the purposes of crosslinking. Figure
67 illustrates the crosslinking mechanism of an OH-functionalized polymer with
melamines. In this case it is assumed that the melamine reacts as a

113 114
PART B – RUBBER SCIENCE AND TECHNOLOGY Rubber materials are conventionally divided in natural rubber, general purpose
synthetic rubbers, and specialty elastomers. The main industrial application is
in the manufacture of tyres (about 80% of the total rubber market), while
Chapter 1: Introduction to rubbers other important technical articles made of rubbers are seals, gaskets, hoses,
shoe soles, antivibration parts etc.

The natural rubber (NR) is very high molecular weight 1,4 cis polyisoprene, the
A rubber, or elastomer, is a natural or synthetic polymer which at room molecular formula of which is reported below:
temperature can be stretched repeatedly to at least twice its original length
and which after removal of the tensile load will immediately return to
approximately its original length (ASTM standard definition).

Typical values of Young modulus and strain at break for common elastomers
are 106 Pa and 103 % respectively. NR is an unsaturated rubber, therefore a polymer having C=C double bonds in
The high extensibility of rubbers is due to the fact that any single chain the main chain. All unsaturated rubbers can be vulcanized by sulfur. The raw
assumes a random coil conformation when no external perturbation is applied. NR polymer is produced by coagulation of the aqueous latex obtained by
This conformation can be easily deformed by the action of an even small several trees, like typically Hevea Brasiliensis. It is therefore a well
external stress thus generating highly elongated conformations. consolidated example of industrial polymer derived from renewable resources.
The stereochemistry is very well controlled (more than 99.9% 1,4 cis) which
The structural requirements for a polymer to behave as an elastomer are the implies particularly good mechanical properties at high elongation (see strain
following: induced crystallization mechanism). It is mainly used in the manufacture of
tyres.
• The chains must be of high molecular weight, with presence of many
entanglements among the macromolecules Synthetic elastomers containing only C and H are often classified as general
purpose rubbers. Examples of unsaturated synthetic rubbers are polybutadiene
• The chains must be flexible, which means that the polymer Tg must be (BR), synthetic polyisoprene (IR), styrene-butadiene random copolymers
low (well below ambient temperature typically), as well as their cohesive (SBR). They are all used in the manufacture of tyres. Saturated synthetic
forces (low polarity polymers) rubbers do not contain any C=C double bonds, therefore cannot be vulcanized
by sulfur. All saturated rubbers show generally better thermal resistance,
• Random coil conformation and high conformational disorder are
assumed when the polymer is undeformed (absence of or very small examples are ethylene-propylene random copolymers (EPR) and
polyisobutylene (IIR), the formula of which is reported below.
crystallinity)

• The final material must be slightly crosslinked, both chemically


(vulcanization) or physically (with phase segregation like in case of
thermoplastic elastomers)

Elastomers normally do not show cristallinity, because polymers in the


crystalline state assume only the minimum energy conformation. One partial
exception is represented by some thermoplastic elastomers where the Specialty elastomers are low Tg polymers containing atoms other than C and
segregated phase shows some crystalline order, like in the case of some H, typically Si, O, Cl, F, N. This means that the solubility parameter of the
segmented polyurethanes. polymer is different from that of hydrocarbons, and therefore the materials
show better chemical resistance (less affinity) towards fuels. Examples are
In principle any high molecular weight polymer may show a rubberlike
fluorocarbon elastomers (mainly copolymers made of tetrafluoroethylene,
behavior at a temperature well above its Tg. On a more practical way,
hexafluoropropylene, vinylidene fluoride), copolymers of butadiene and
elastomers must have a Tg at least below -20°/-40°C in order to show
acrylonitrile (NBR), polychloroprene (CR), and silicone based polymers like
rubberlike behavior at room temperature.
polydimethylsiloxane (PDMS). They are normally much more expensive

115 116
materials than rubbers described before, and are used in applications where direction of the imposed deformation. Only at the highest elongation bond
better thermal and chemical stability is required (i.e. car engine seals). lengths and angles are stressed, with effect on the internal energy.

In such a deformation process there is significant decrease of the state of


disorder with a negative change in entropy with the deformation (loss of
conformational degrees of freedom). The driving force to the elastic recovery is
therefore of entropic origin, since the material tends to return to its minimum
free energy state by increasing its entropy level. It is however needed that the
rubber is capable to store the mechanical energy upon deformation (i.e. to
stay in a small entropy state) without dissipation in viscous flow. This is
Enthalpic and entropic elasticity normally achieved by the vulcanization process, which is a weak crosslinking of
the rubber molecules.
Most solids (crystalline solids) show an enthalpic-elastic behavior, since the
imposed deformation normally causes an increase in free energy of the
material due to the increase of internal energy / enthalpy of the material
related to the displacement of atoms from their equilibrium (minimum internal
energy) positions. The effect on the entropic part is negligible in this case,
since atoms are only minimally shifted from their equilibrium position in the
crystal lattice. Accordingly, the driving force to the elastic recovery of the
deformation will depend on the fact that the system tends to return to its
minimum free energy state with a decrease in its internal energy (or enthalpy)
content.

The entropic-elastic behavior of rubbers is sketched in the following figure,


showing also the role of vulcanization. It is clear how in the absence of
crosslinking the stretched rubber can return spontaneously to its maximum
entropy state (random coil) by viscous flow and thermal motions (passing from
b to c). In this case the elastomers still shows a high deformability, but no
elastic recovery occurs after removal of the external load (see d). The
elastomer after vulcanization maintains a small entropy state under load
without dissipation (from f to g), allowing for a fast and complete deformation
recovery.
The elastic behavior of rubbers is quite different since in that case the
macroscopic deformation causes only conformational transitions in the
polymer, with the chains progressively passing from a compact, disordered,
random coil situation to a more ordered arrangement with chains aligned in the

117 118
If the system is an elastic body, it will be subjected to a reversibile
deformation by the action of an external applied force . The work done
on the system is given by

being the work done on the surrounding due to the volume change.

Therefore the change in internal energy E can be rewritten as

The Gibbs free energy is given by

Chapter 2: Classical thermodynamics of rubber elasticity

The elastic behavior of rubbers can be justified following a classical


thermodynamic approach. Consider a strip of rubber and the change in free
energy that occurs upon stretching.
Passing to differentials we obtain:
The change in internal energy E for a reversible process (such as the elastic
deformation of a piece of rubber) will be given by:

reversible process

where +dQ is the heat added by the system, and +dW is the work done on the
system.

Recalling the first principle of thermodynamics, Qrev is the heat exchanged Therefore the following fundamental thermodynamic relations are obtained:

reversibly, and S is the entropy, related by the following equation:

119 120
The work done on an elastic body raises its free energy content. The elastic
retraction force of the body can be related to such an increase in free energy It is however experimentally found that the slope is not constant but
at constant pressure, and divided into independent enthalpic and entropic progressively decreases with the deformation, eventually obtaining negative
contributions. slopes at very small deformations (<5%). This phenomen is known as
thermoelastic inversion, and it typically occurs when the thermal expansion of
the material overlaps the elastic phenomena. An example is shown below.

By considering the following property of partial derivatives (the inversion of the


order of derivation), the change in entropy with elongation can be related to
the change in the retraction force with the temperature, at constant length L

The equation described above relates the elastic retraction force to the sum of
two independent terms, namely enthalpic and entropic. It is called equation of
state of elasticity at constant pressure. In order to verify the reliability of such
a model, the experimental test to be used is based on the measurement of the
force needed to maintain a given deformation (L = constant) in a rubber
specimen, measured as a function of the temperature T. By the following
graphical representation of the f-T curve, one could obtain the enthalpic
component to the elastic retraction force fH as the intercept of the curve To avoid this effect, it is therefore needed to repeat the experiment at
extrapolated at T = 0, and the entropic component fS as the slope.
, applying an external pressure to counterbalance the thermal
expansion of the material. The equation of state at constant volume is derived
passing to Helmoltz free energy F
= slope
= Helmoltz free energy (
= intercept at T=0
The Helmoltz free energy is related to internal energy E and to entropy S by

121 122
Therefore passing to differentials, and remembering that dV = 0 by definition, counterbalance the thermal expansion would be so high to be experimentally
the following relations are obtained not accessible. It is therefore convenient to work at constant pressure, but
applying the following mathematical corrections to the equation of state of
elasticity at constant pressure:

being and = coefficient of thermal expansion of the rubber, to be

The elastic retraction force is related to the change in Helmoltz free energy determined separately with proper thermomechanical measurements in the
by the following relation: temperature range considered.

The best experimental procedure is therefore to measure the retraction force


at various temperature in isobaric experiments, for rubber specimen with
different (but fixed) extent of deformation. By knowing the thermal expansion
coefficient of the rubber αT, the term (∂f/∂T)V,L can be easily computed. This
allows for the determination of internal energy and entropic contributions fE
and fS to the actual retraction force. Some results concerning common
elastomeric materials are reported in the table below.
Again, changing the order of derivation one obtains:

fE/f fS/f

NR 0.18 0.82

BR cis 0.15 0.85

PDMS 0.15 0.85


from which the equation of state of elasticity at is eventually
derived. IIR -0.03 1.03

It is evident how the major contribution to the elastic retraction force of real
rubbers is due to the entropic part, and this is the classical thermodynamic
explanation of the concept of entropic elasticity normally associated to the
rubberlike behavior of elastomers.

By applying this analytical form of the equation of state, the entropic


component of the retraction force fS should be obtained at any deformation
without observing the thermoelastic inversion. However a vulcanized rubber is
a nearly incompressibile solid, and the external pressure needed to

123 124
Chapter 3: Molecular theory of rubber elasticity

Note that the entropy content of the single macromolecule will be indicated in

According to classical thermodynamic considerations seen before, for an “ideal the following as lowercase.
rubber” the retraction force is simply given by

f = -T(∂S/∂L)V,T

Now the objective is to obtain the constitutive equation of rubbers relating the
stress σ to the elongation ratio λ, starting from molecular mechanics
considerations. Let’s consider first a single macromolecule in a Cartesian
coordinate system. This is the case of a gaussian chain (i.e. a chain obeying
the Gaussian distribution function) with one chain end grabbed in the origin
(0,0,0), and the other end stretched uniaxially along the z direction.

The distribution of end-to-end distances P(R) is given by the Gauss probability


function assuming a phantom chain model (see conformational analysis):

Remembering from classical thermodynamics the case of retraction force for


entropic elasticity, therefore f = -T(∂S/∂L)V,T, and considering in the present
case that we obtain by derivation:

where

The above written equation predicts that force is linear with distance like in
case of Hook’s law, but with some important peculiarities worth to be
The entropy of the system according to the Boltzman relation is given by:
underlined.

1. linearity is valid up to very large deformations (z ~ 300%), that is until the


Gaussian approximation holds

2. the “modulus” is given by , therefore it increases with T which is


the opposite of the trend observed with conventional, enthalpic-elastic solids;
with = number of different conformations available to the system (the this behavior is actually specific of entropic elasticity due to the fact that the
random coil shrinks with the temperature (more compact conformations are
macromolecular chain).
less unfavoured at high temperature).
For uniaxial extension in the z direction we can input
After having obtained the analytical expression of the retraction force for the
single chain, the next step is to pass to a more realistic rubber system. The
material is modeled in form of a network made of several crosslinked sub-
. chains, the so-called Kuhn model.

125 126
The Kuhn model is still a simplified version of the real material, actually it lacks
of entanglements, loops, sol fraction, dangling bonds etc., which are always
present in a real, vulcanized rubber. Moreover it is considered that all the
individual subchains have the same chain length and same connectivity (three-
functional, four-functional, etc).

Let’s now consider a cubic block of slightly crosslinked ideal rubber (therefore
realized according to the Kuhn model) that undergoes a strain along a set of We can use the end-to-end vector (see conformational analysis for its
axes parallel to a Cartesian system. definition) as a descriptor of the conformational state of the subchain. For the
undeformed state we will have:
Extension ratios in the x,y,z system are given by λ1, λ2, λ3 (where ).

Deformation of rubbers always occurs at and therefore we can fix


the following condition:

The corresponding entropy content of the individual subchain will be:

Now we want to evaluate the change in entropy, and the state of strain in an
individual gaussian subchain belonging to the cubic block of Kuhn rubber.
According to the “affine deformation” assumption, the strain of an individual
subchain will be coherent and proportional to the macroscopic strain of the Imagine now to apply a mechanical deformation: under the affine deformation
cubic block. assumption the individual subchain strain ( ) and the cubic block
The deformation occurring to the subchain of the Kuhn rubber is represented in deformation ( ) will be proportional, therefore
the figure below; note that the subchain always has one end grabbed in
(0,0,0).

127 128
and The above written relation can be modified with some simplifying assumptions.
Actually if in the undeformed state there is a perfectly random distribution of
the end-to-end vectors in all the directions (isotropic material), we can write

The entropy of the deformed subchain can be rewritten as:

and

Now the change in entropy upon deformation can be calculated as the By definition:
difference between the entropy content before and after deformation:

In the Kuhn model it is assumed that all the subchains have the same
with = number of subchains in the cubic block.
length/molecular weight, which means:

Now the change in entropy can be rewritten as a function of

Therefore it now feasible to compute the change in entropy for the whole
material, indicating in this case the entropy as S uppercase:

That is the entropy change for all the subchains, i.e. for the whole rubber
specimen.

129 130
Remembering that , the final expression of is eventually
the change in free energy upon deformation for a piece of rubber is
obtained

The change in energy ∆F will be equivalent to the work W added for the
Since , with uniaxial deformation in direction 1 the deformation of the rubber
following relations hold:

and

In order to pass to intensive properties (J/volume), the rubber density ρ can be


introduced:
can be now expressed as a function of only one variable

The expression of the Helmolz free energy for the case of entropic elasticity at with = molecular weight of the subchain between two crosslinks, and =
constant volume is Avogadro number (constant). It is now feasible to find a suitable expression for
N’, number of elastically active subchains, in terms of more easily available
quantities like the rubber density, and the molecular weight between
crosslinks.

since by definition the enthalpic part is zero.

Now being

We put it in the expression of the work by volume w added for the deformation
of the system obtaining:

131 132
being

The quantity is also indicated with G, and it is often called “modulus” of Since the stress is and , it follows that
the rubber. It increases with T (see entropic elasticity model) and with
decreasing . The density of crosslinking is given by the ratio between ρ
and Mc.

= density of crosslinking in

which is the common form of the constitutive equation of elasticity of the


The elastic (reversibile) work is therefore a function of both temperature and rubber.
density of crosslinking through G

The molecular theory of rubbers is noteworthy for its ability to predict the
mechanical behaviour of the material on the basis of purely molecular
considerations (entropy of the macromolecular chain).

The following picture shows the comparison between the stress-elongation


behavior predicted by the equation of the rubber elasticity (continuous line),
Let’s consider now the mechanical deformation of a cubic block of rubber
with the experimental behavior observed for a vulcanized NR sample (dotted
having the following geometrical parameters:
line).
= initial section, = initial length, = change in length upon deformation,

= uniaxial tensile force

The total work done by the rubber specimen will be indicated as W (uppercase)

while the work of deformation per unit volume will be w (lowercase). Passing
to the differentials we have:

133 134
- The number of elastically active subchains should include also the
entanglements, which are not considered in the Kuhn model.
- The eventual strain-induced crystallization of stereoregular polymers is
not taken into account.

Other experimental behaviors not predicted by the model are the hysteresis
cycles: when a real rubber sample is subjected to cyclic loading/unloading runs
without reaching the strain at break, the forward and reverse curves never
overlap completely. The area given by the difference of the curves is called
hysteresis of the rubber, and it is an indication of the energy mechanically
dissipated. Actually part of the mechanical energy stored upon deformation is
dissipated by friction and viscous phenomena (with heating of the rubber).
Strain-crystallizing rubbers show a particularly high hysteresis.

It can be observed that the agreement between the model and the real
behavior is very good in compression and moderate elongation only (up to
around ). At higher deformation the model predicts a too high
strength, and doesn’t foresee at all the strain-hardening behavior typical of NR
and other stereoregular rubbers ( ). The very sharp increase in the
stress at break observed at high elongation corresponds to a strain-induced
crystallization and it is typical of NR and, to a less extent, of other
stereoregular elastomers. Natural polyisoprene 1,4 cis shows actually all the
prerequisites needed to crystallize, however its melting point is very low,
around -25°. At the equilibrium temperature (corresponding to the melting
point) we have

∆Gm = 0 = ∆Hm –Teq ∆Sm A better fit of the mechanical behavior of real elastomers can be obtained by
using more complex chain models (instead of Gaussian approximation), or with
semi-empirical relations. One of the most important examples of the latter
Therefore Teq = ∆Hm /∆Sm = -25°C case is represented by the Mooney-Rivlin equation, which is a two-parameter
equation derived by the molecular theory model.

In the stretched elastomers the entropy content is smaller because the chains
are more ordered, so ∆Sm decreases shifting the Teq to room temperature,
where the strain induced crystallization occurs.

The main limitations of the molecular theory approach are therefore the
following:

- The gaussian chain model is mathematically easy but too rough for
describing high deformations.
135 136
and are constants typical of the material under consideration. In The case a, with corresponds to the behavior of the ideal rubber
particular depends on T and density of crosslinking (like G), while (Kuhn rubber or molecular theory).
quantifies the shift from the ideal behavior and it is normally considered an The case b, with is typical of vulcanized but unfilled real elastomers,
expression of the role of entanglements. The Mooney-Rivlin equation shows a
better fit for real, vulcanized (but unfilled) elastomers. and it is due to the presence of entanglements.

The reduced stress can be defined as follows: The case c shows positive deviations at high , and it is due to the limited
extensibility of real chains

The case d is due to crystallization under strain, and it is typical of highly


stereoregular polymer like NR.

Consequently the Mooney-Rivlin equation can be rewritten in a more compact


form as follows

If we plot σ* (or f*) against λ-1, various cases can be seen as shown below.

137 138
Chapter 4: Vulcanization A rubber compound is always a mixture made of the base polymer, fillers, and
chemicals needed to promote crosslinking during the molding operations.
Vulcanization is a mild chemical crosslinking process which allows for a Commonly used vulcanization agents are sulfur, peroxides, or ionic substances
practical exploitment of rubber elasticity. It is an irreversible process, normally in some special cases.
obtained while molding the rubber manufact. The main difference with the
other crosslinking processes already found in thermoset polymers is that in Vulcanization with sulfur was historically the first method discovered by
case of rubbers the number of chemical crosslinks to be formed are very few Charles Goodyear towards the end of the 19th century, and it is still the
and far between, and very small values of density of crosslinking have to be method of choice to process unsaturated elastomers like NR in the tyre
achieved. manufacture technology. The reagent is the cyclic octamer S8 which
decomposes thermally by ionic mechanisms. The fragments are capable to
After vulcanization the polymer must maintain the random coil conformation bridge two different insaturated polymer chains. The reaction doesn’t involve
with minimal effects on its Tg and conformational degrees of freedom. At the directly the opening of C=C double bonds, but it is more a substitution reaction
same time dissipation of mechanical energy in viscous flow due to thermal of allylic hydrogens by sulfur (allylic hydrogens are H atoms in α-position with
effects, or by action of a imposed strain, is effectively eliminated. After respect to C=C double bonds, like for ex. –CH2-CH=CH- ).
vulcanization the polymer is crosslinked and therefore cannot be molded or
formed anymore. A vulcanized rubber doesn’t dissolve in any type of solvent, The direct thermal reaction of S8 with unsaturated polymers is very slow (24-
but can be significantly swollen by exposure to thermodynamically affine 48h), therefore other chemicals are used to improve the efficiency of the
solvents. The effect of vulcanization on dynamic-mechanical properties of a process. Other substances used in the base recipe are ZnO (to promote sulfur
rubber material is shown below. The elimination of high temperature viscous decomposition), stearic acid (to improve solubility of ZnO into the polymer),
flow and the positive T-dependence of Young’s modulus are evident, as also and organic accelerants (like benzothiazoles etc) to speed-up the crosslinking
predicted by the entropic elasticity model and by the molecular theory of reaction. An optimized recipe and process can allow for an efficient
rubber elasticity. vulcanization in 5 minutes at +140°C.

Vulcanization with sulfur tends to form monosulfide C-S-C, disulfide C-S-S-C,


and polysulfide C-Sx-C bonds, as well as loops and dangling bonds. Polysulfides
are quite effective elastically but they are also very weak limiting the thermal
stability of the vulcanized product. The formation of monosulfide bonds can be
favoured by using small amount of S and larger amounts of organic
accelerants.

Vulcanization is therefore a crucial step in order to use elastomers for


structural applications. Alternatively, thermoplastic elastomers (see living
polymerizations) based on block copolymers exploit the spontaneous phase
segregation between the immiscible blocks to generate a physical form of
crosslinking. In this case the process is reversible, but normally the
performances of the materials are much lower especially in the high
temperature range.

139 140
The following table reports the dissociation thermal energies of the bonds
involved.
To minimize the risk of degradative beta-scission reactions, special coadiuvants
of crosslinking can be used like triallylcyanurates (see below). In this case the
allylic C=C double bonds are very reactive and can work as a bridge improving
the efficiency of the combination reaction between macroradicals.

O CH2 CH CH2
C
N N
With saturated rubbers sulfur tends to be ineffective and other vulcanization
C C
processes are used. Vulcanization with peroxides involves generation of free N O CH2 CH CH2
macroradicals and following crosslinking with termination by combination. In CH2 CH CH2 O
this case more stable C-C bonds are formed. However the free radical
vulcanization is not a very well controlled reaction, and degradation side
reactions are always feasible. Degradation by beta-scission reaction is
particularly frequent, leading to lower molecular weight chains. The Special elastomers can be vulcanized with other methods which are specific of
vulcanization process with peroxides is shown below. the chemical nature of the rubber to be processed. For example
fluoroelastomers are normally vulcanized by addition of bisphenates or
diamines, siliconic rubbers can be processed as bicomponent products by Pt
ROOR ∆
→ 2 RO∗ catalyzed hydrosililation reactions and so on.

Monitoring the vulcanization process

The monitoring of the vulcanization process is a crucial step in the rubber


RO ∗ + technology to optimize material performances. It is normally done by dynamic
ROH + * + *
H * rheological methods (see the fundamentals of dynamic-mechanical analysis).
The raw compound containing the polymer, fillers, and vulcanizing agents, is
subjected to small amplitude and constant frequency sinusoidal stress while
heating with a programmed thermal cycle simulating the molding conditions.
The torque-time (or dynamic modulus-time, depending on the instrument)
The side reactions leading to lower molecular weight products are shown below trend is monitored with formation of a graph as shown below.
in the case of EPR vulcanization with peroxides.

CH2 CH2 CH2 C CH2 + * CH

CH3 CH3
141 142
of an Arrhenius type equation, the latter the dependence of the concentration
of the reacting groups or functions.

dα /dt = K(T).f(α)

with K(T) = A. exp(-Ea/RT) and Ea = activation energy of the process.

Different models are instead available for f(α), in particular

f(α) = (1- α)n n-order reaction

In this case the rate is highest at the beginning of the process.


The initial torque decrease after the start of the test is only due to a simple
thermal effect (sample heating) until a minimum is reached after a certain time
tp (thermoplasticity time). The following torque increase is due to the onset of f(α) = αm (1- α)n autocatalytic model, with n>1 and
vulcanization. The final plateau torque is related to the density of crosslinking. 0<m<1
The (eventual) reversion with torque decrease after reaching a maximum
involves thermal instability of some formed bonds (like for example polysulfide
linkages).
The autocatalytic model predicts that the rate is zero at t=0 and at the end of
The following process parameters are usually measured and reported. the process, and highest at intermediate α values like in case of some
monocomponent thermosets (epoxies). Rubber compounds generally follow an
ML = lowest achievable torque autocatalytic model, an example of graphical representation is shown below.
The analytical model is:
MH = highest achievable torque

ts x = scorch time (time needed to reach a torque ML+X)

Tx = time at X% of vulcanization

The so called scorch-time is the maximum time allowed for the processing and
forming of the compound in the chosen experimental conditions, and it is
roughly related to the gel-point of the rubber. The optimal vulcanization time is
normally given as T90, that is the time at that temperature after which 90% of
the maximum torque is achieved. The nearly isothermal part of the curve
comprised between ML and MH can be used for a kinetic modeling of the
process. If we put ML = M0 and MH = M∞, the conversion α of the vulcanization
process can be defined as

α = (Mt-M0) / (M ∞ -M0)

The rate of any reaction dα/dt can be expressed as the factor of two
independent terms, the former indicating the temperature dependence in form
143 144
Conversion data can be obtained by methods other than dynamic rheology; in Determination of the density of crosslinking
particular the vulcanization process is exothermic, and therefore the process
can be monitored also by differential scanning calorimetry (DSC), with The polymer Tg and the density of crosslinking are among the most
integration of the enthalpy peaks. If ∆Htot is the total enthalpy of vulcanization important physico-chemical parameters of a rubber. While the former property
(generally measured in dynamic scans), the conversion α and the extent of is strictly material-dependent, the latter largely depends on the processing and
reaction dα/dt can be expressed by: vulcanization conditions. Modulus and hardness increase with the density of
α = ∆Ht/∆H∞ crosslinking, while elongation generally decreases. Other properties show a
non-monotonous and not easily predictable trend, like mechanical strength,
dα/dt = (1/∆Htot) . d∆Ht / dt resilience, gas permeability, compression set, tear resistance, so that an
optimal density of crosslinking has to be found according to the application
needs.
Measuring the enthalpy peaks Tp in dynamic scans at various heating rates q+
allows also a more direct estimation of energy of activation of the process
according to the Ozawa equation

Ea = -2.3R d logq+ / d(1/Tp)

The DSC technique is experimentally very easy and versatile, the main
limitation being the low sensitivity since rubber processing is much less
exotermic than thermoset processing (less chemical reactions are involved).
An example of temperature dependence of the exothermal peak temperature
Tp in a crosslinking process is shown below.

The easiest way to measure the density of crosslinking is through the


estimation of the shear modulus G, normally by dynamic-mechanical testing,
since it should be remembered that G = RT = ρRT / Mc (being ρ the density
and Mc the average molecular weight between crosslinks).

Alternatively, the density of crosslinking can be estimated through swelling


measurements in organic solvents. Vulcanized rubbers doesn’t dissolve but
may actually swell considerably in good solvents. Swelling leads to an isotropic
extension of the network. According to the Flory-Rehner theory the
deformation of the polymer network is given by

λ1. λ2. λ3 = (L1.L2.L3)/(L01.L02.L03) = V / V0 = λv

Since the expansion is isotropic we have λ1=λ2=λ3 = λ and therefore λ = λv 1/3.

The volume fraction of polymer in the swollen gel is given by

v2 = V0 / V where V0 is the volume of the unswollen polymer


145 146
polymer solutions can be expressed in one of the forms of the Flory-Huggins
equation reported below
This leads to the following expressions

∆G1,mix = RT [ln(1-v2) + v2 +χv22]


λv = 1/v2

λ = v2 -1/3

where χ is the polymer-solvent interaction parameter, related to the difference


between the solubility parameters δ of the polymer and of the solvent.
Remembering from the molecular theory of rubber elasticity (see page 17) that
∆G = -T∆S (entropic elasticity) with
χ=Vm/RT (δ1-δ2)2

The equilibrium Flory-Rehner equation is therefore:

ρV1
it can be shown that a suitable mathematical expression for the increase in − ln(1 − v2 ) − v2 − χv22 = v12 / 3
free energy due to the network expansion (chemical work done by the solvent)
Mc
∆Gel can be obtained
Accordingly, if the isothermal swelling of a strip of rubber is measured
gravimetrically (with determination of v2), and if the polymer-solvent
∆Gel = (3ρRT)/(2Mc) (v2 -2/3 – 1) interaction parameter is known or calculated, the corresponding density of
crosslinking of the rubber = ρ / Mc can be determined. Qualitatively, the
relation between equilibrium swelling and average molecular weight of chain
Considering that segments in the polymer network is shown below.

1/v2 = (V0 + n1 V1) / V0

with n1 and V1 being the mole fraction and molar volume of the solvent
respectively, the molar free energy of dilution ∆G1,el is finally obtained by
taking the partial derivative with respect to n1.

∆G1,el = (ρRT/Mc) V1 v21/3

The equilibrium swelling is reached, according to the Flory-Rehner theory,


when this increase in free energy caused by the random coil expansion (with
loss of conformational degrees of freedom, just like in the case of a mechanical
work and deformation) is counterbalanced by favourable polymer-good solvent
interactions. In order to quantify such interactions the thermodynamics of

147 148
Chapter 5: Mechanical reinforcement 1. The size of elemental particles

Real elastomers are made of vulcanized polymer + reinforcing fillers. 2. The “structure” of aggregates
Reinforcement means increase of modulus, mechanical strength (ca. 1 order of
magnitude), tear and abrasion resistance etc. 3. The surface physics

Normally in filled thermoplastics we observe increase in modulus, hardness, 4. The surface chemistry
strength, but decreased elongation. In case of filled rubbers a synergistic effect
with simultaneous improvement of modulus, strength, and elongation at break
is often found. The elemental particle size strongly influences the interfacial area available for
interactions with the polymer matrix. Actually smaller particle carbon blacks
The mechanical reinforcement in filled rubbers is a function of particle size, are those more reinforcing. CBs are classified on the basis of their specific
aspect ratio, physical-chemistry of matrix-inclusion interactions. If we follow a surface area As, which is defined as follows
size-based classification criterion, we can divide the various filler grades in:

reinforcing = 0,01-0,1 micron


As = surface/mass = [(4πr2)/[(4/3 πr3) . ρ] = 6/ρ.r
semireinforcing = 0,1-1 micron

diluents = 1-10 micron


where ρ is the density and r the particle radius. Diameter d can be directly
degradants > 10 micron measured through Electron Microscopy, while the surface area As is determined
The fillers commonly used to produce reinforced rubbers are carbon black (CB) through gas, or iodine, or cetylammonium bromide (CTAB) adsorption tests. So
and silica. Both are made of spherical particles which tend to form cohesive, we will have fillers with high as well as low surface area, as shown below.
anisotropic aggregates. CB is still the most important and used. An electronic High As = 40-200 m2/g (CB, silica)
microscope image of CB aggregate is shown below.
Low As =1-20 m2/g (talc, calcite, clay)

The ASTM nomenclature of CB grades is mostly related to the size of elemental


particles; those with smallest particles are highly reinforcing fillers.

Several grades of CB are available; they can be effective reinforcing agents, or


The term “structure” concerns the anisotropy degree of elemental particle
simply intert fillers used to reduce cost of the compound. If a CB grade is
aggregates, which cannot be broken down by mechanical mixing and therefore
reinforcing or not depends by a series of factors, which in order of importance
survive as such in the final elastomers. The “structure” of CB grades is
are:
measured through oil absorption tests. Reinforcing grades show more

149 150
persisting anisotropic aggregates, and show higher values of oil absorption • The medium-range interactions are those explained by simple mixture
test. The void fraction of aggregates determines the occluded or bound rubber laws (like in case of a conventional composite)
phenomena: a polymer fraction with reduced molecular mobility filling
interstitial pores, it cannot be separated by solvent extraction techniques even • The long-range interactions are those due to the filler re-agglomeration
in the raw (unvulcanized) rubber. (like in case of thixotropy).

Surface physics. CB surface is made of graphitized layers, normally highly The presence of fillers strongly influences both the processing (viscosity) and
crystalline. There are however always “defects”, high energy amorphous the mechanical properties (modulus, strength) of filled rubbers. Viscosity of
zones, capable of stronger physico-chemical interaction with the unsaturated complex fluids containing solid, not interacting particles is predicted by the well
polymers. CB grades with lower surface crystallinity normally show better known Einstein equation, where the most important parameter is the filler
reinforcing effect. volume fraction Φ.

η = ηo (1+2.5Φ)

The Einstein equation is not followed by filler-rubber compounds, where


interactions cannot be neglected. A modified version, named Guth-Gold
equation, is normally valid for compounds with non-reinforcing CB.

η = ηo (1+2.5Φ + 14.1Φ2)

The Guth-Gold equation must be often corrected taking into account the
effective filler volume fraction Φc, therefore including the bound and occluded
fraction; Φc can be determined separately through oil sorption methods.

Surface chemistry. Many polar groups can be present (due to oxidation) which
mostly influence reactivity during vulcanization, but with small effect on
reinforcement.

CB is structurally suitable to give strong interactions especially with


unsaturated polymers (containing C=C bonds). The interaction is mostly
arising from the π cloud of the polymer and the CB surface “defects”. Such
interactions are stronger than simple physical interactions, but they are not
really covalent bonds. They can be classified in various types.

• The short-range interactions concern the phenomena of bound rubber,


and rubber shell. The bound rubber is the polymer fraction which cannot
be solvent-extracted (even on raw compounds before vulcanization). The
rubber shell is the polymer layer immobilized on the surface filler (with
higher Tg), and it will depend on temperature and state of stress.

151 152
Small deformation reinforcement The interaggregate interactions (see long-range interactions defined above)
are responsible for a strongly non-linear reinforcement effect called also Payne
The small deformation behavior is best studied through dynamic-mechanical effect. The intensity of the Payne effect is related to the CB surface area and
analysis, working at constant temperature and frequency and sweeping the amount, as well as to the state of dispersion of the rubber-filler mixture. It
shear strain amplitude. In case of reinforced rubber compounds it can be cannot therefore be predicted by any mathematical relation. Examples of
observed that the complex modulus G* and storage modulus G’ are related to Payne effect (E vs. double strain amplitude DSA) are shown below in case of
contributions which are both deformation-dependent and deformation- compounds with increasing CB amount.
independent, as shown in the classical picture shown below.

The pure gum or polymer network contribution is due to the presence of This effect can be attributed to filler-network interactions (see also tixotropy)
chemical crosslinks and to the entanglements blocked by chemical crosslinks. or to polymer-filler interactions (labile bonding).The filler network is reversibly
It is therefore independent on the filler type and content. destroyed with strain (3-10%), since with a high DSA the particles are too
The hydrodynamic effect is due to the presence of rigid fillers with a strain distant to rebuild the “structure”. It is interesting to observe that the sharp
amplification effect. It involves the effective CB content. decrease in G’ is always related to a peak in G’’, indication of highly dissipative
phenomena; an example is shown below.
The filler-rubber structure is due to the rubber occluded within the void space
of CB aggregates.

The elastic modulus containing the contribution of polymer network,


hydrodynamic effect and filler-rubber structure can be predicted by an
empirically modified version of the Guth-Gold equation, where f is the aspect
ratio of the structured filler aggregate, and Φc is the corrected filler volume
faction including the bound / occluded rubber.

E = E0 (1 + 2.5 f Φ + 14.1 f2 Φc2)

In other similar models the numerical parameters are changed.

153 154
High deformation reinforcement

An effective mechanical reinforcement involves a synergistic improvement of


both mechanical strength and strain at break. In general all the loading curve
is shifted to higher stress values as in the example reported below.

Another experimental evidence of labile bond rearrangement is the well known


Mullins effect, which can be seen during repeated loading-unloading cycles. It
is normally observed that the loading curves show a progressive softening
following the cycles, until a residual deformation may eventually appear. The
Mullins effect depends on the rearrangement of labile bonds, but it can be due
A = unfilled rubber; B = reinforced rubber; C = strain-crystallizing rubber
also to mechanical degradation (that is rubber molecular weight decrease) if
the loading cycles are carried out at very high elongations. In these conditions
the presence of free radical species (as markers of chemical degradation
A better examination of the stress-strain curve of a reinforced elastomer phenomena) were evidenced.
suggests that the slope of the curve (i.e. the tangent modulus) is decreasing at
low strain, then it starts to increase at higher deformation as shown below.

The initial modulus decrease is due to filler network breakup. The following
modulus increase is due to a progressive chain desorption with a strain
amplification effect. It is often observed an increase of elongation at break,
which can be attributed to the presence labile bonds among chain segments
and filler particles, and to a homogeneization of macromolecular segments
absorbed on CB surface as an effect of the mechanical work.

155 156

S-ar putea să vă placă și