You are on page 1of 33

Notes on Polymer Processing

Koushik Viswanthan

Fall 2015
1 Introduction
Conventional manufacturing process engineering has involved a detailed study of metal pro-
cessing. While this is important for many modern technologies — one can immediately think
of automobile bodies, aircraft wings, building structures, almost any technological innova-
tion (phones, computers,...) — there is a whole other class of materials whose existence is so
ubiquitous that we often taken it for granted. These constitute polymers.
Look around you and you will see how commonplace polymer materials really are. Starting
from plastic cereal bowls, to your clothes, from your (most likely plastic) key chain to your
car seat covers and even the tires. And the surprising thing is that polymers are not even
new — man has used polymers in some form or the other for thousands of years! Wood,
leather and wool are all examples of natural polymers.
In addition to these natural products, polymers were first artificially synthesized in the
lab in the nineteenth century with the large scale demand for rubber. The word polymer itself
refers to the material’s molecular structure, being composed of a long chain of many small
subunits (‘polymer’ == polus meros, i.e., Greek for ‘many parts’). Each subunit is called
a monomer and contains some unique chemical properties. These properties fundamentally
affect the mechanical behavior of the resulting polymer but a detailed study will take us far
afield into the realm of organic chemistry. A good overview and introduction may be found
in the book by Treloar [1].
Natural polymers include things like cotton fibers (mainly cellulose) and wool/ silk fibers
(composed of proteins), containing oriented crystallites, which have strong tensile strength
(along the fiber). For instance, fibers used in textiles are processed in such a way so as
to ensure strong cohesion between individual fibers (to ensure thermal conductivity) and
high tensile strengh per fiber (to withstand the harsh processing and usage conditions).
Another example of a natural polymer is natural rubber, which is now being substituted by
synthetic rubber. Of course, biological examples abound, including gelatin (made of collagen,
a protein), skin, muscles (made of bundles of fibers containing protein) etc.
Synthetic polymers can be broadly classified into two types — thermoplastics, which can
be melted, worked, then remelted and reworked; and thermosets, which cannot.
Some examples of thermoplastics are:

• polyethylene — most commonly used thermoplastic polymer, ∼ 80 million tons pro-


duced annually. Used in plastic bags, containers, packaging, synonymous with the name
plastic.

• polypropylene — used for making cups and cutlery, usually by injection molding. High
fatigue resistance (lids in a tic tac box).

1
• polyvinylchloride (PVC) — rigid form used as construction material and in credit
cards(!), can be softened by adding additives, and then used for inflatable products,
electrical cable insulation, etc.

• polystyrene — ‘styrofoam’ , sheets or molded polystyrene used in CD cases (rigid), in


foam form for packaging (good damping and thermal insulation).

• polymethylmethacrylate (PMMA) — transparent rigid polymer, most often formed


into sheets and used as window glass (perspex/ plexiglas).

On the other hand, some polymers are produced by forming cross–links in the structure
to enhance properties. These cannot be remelted and reworked once formed, and are called
thermosets or thermosetting plastics. Examples include the following:

• Bakelite (polyoxybenzylmethylenglycolanhydride) — one of the first synthetic plastics,


heat–resistant, used in firearms, children toys.

• Epoxy or polyepoxides — can form cross-links upon exposure to chemicals, used as


sealants.

• Polyurethanes — used in seals, high–performance adhesives.

In general, both thermoplastics and thermosets can show similar mechanical behavior, only
they arise from slightly different microscopic mechanisms. The main (and important) dis-
tinction is their ability (or lack thereof) to be remelted and reworked.

2 The importance of time scales


Almost all polymers show a remarkable variety of mechanical behaviors, perfectly illustrated
by the toy silly putty. A quantitative characterization of this range of behavior is provided
by the concept of a time scale. This is basically a value, with units of time, that measures
the duration of the important event in a problem. Depending on the exact problem at hand,
a time scale can either be a precise value or a rough estimate.
Every problem has two such time scales. One is a material relaxation time τ , i.e., an
approximate time it takes for an initially loaded sample of a given material to lose (or
dissipate) the stresses developed. The second is an experimental time scale T , i.e. a measure
of the time for a given experiment. This experimental time scale could be the duration for
which the material is loaded, for example.

2
The actual behavior of a material depends on the ratio (denoted De, or the ‘Deborah
number’) between these two time scales, defined as

τ
De = (2.1)
T

If De  1, the material relaxes very quickly compared to the duration of loading, hence
is characteristic of a fluid. On the other hand, De  1 implies that the material retains the
stresses on the time scale of the experiment, so behaves like a solid.
To illustrate this concept further, consider a typical relaxation time for water, τwater ≈
−12
10 seconds and for steel, τSteel ≈ 1012 seconds. A ‘normal’ experiment will last anywhere
between 10−9 seconds (nanoseconds) and 109 seconds (30 years). Hence for any such normal
experiment, Dewater has a maximum value of 0.001, which is much lesser than 1. Steel has a
minimum value DeSteel of 1000! So it is clear that water has to be a liquid and steel a solid
under normal considerations.
Now it is clear why polymer materials show mixed, complicated behavior. Their time
scales are of the order of 1 second. For instance, consider silly putty. Let us take τSillyP utty = 1
second. If we apply a load for T = 0.01 seconds (bouncing off the floor), De = 100 so it is
a solid at this experimental time scale. On the other hand, an experiment lasting T = 3000
seconds (a 50 minute lecture) results in De = 0.0003 so on this experimental time scale silly
putty shows fluid–like behavior!
Most other polymers have relaxation time scales similar to silly putty. For example,
consider loading a PMMA sample (a sample of plexiglas) at a strain rate of 104 s−1 (or a time
scale T = 10−4 seconds), it is a brittle solid and cracks. But by applying slow pressure, if
we technically wait on the sample for a thousand years (T = 3 × 1011 seconds), De = 10−11
it will lay out like a liquid! Alternatively, one can cause cracks to propagate in water (yes,
liquid water, not ice) if we load a sample of water at an extremely high strain rate (∼ 1015
sec−1 )!
So in summary, polymers, given their relaxation times and wide variety of experimentally
attainable Deborah numbers, can show either fluid–like or solid–like behavior. We have to
hence develop models that describe such a behavior and that will be our central occupation in
the coming lectures. Armed with these concepts, we will attempt to model industrial polymer
processes (which are innumerable) and estimate parameters like strains and stresses, analyze
failure, fracture etc., just as we did with plastic deformation in metals. Our approach will be
simple in scope, additional details, including an elaboration on the discussion above on time
scales, can be found in Refs. [2] and [3].

3
3 General overview of mechanical properties
Any discussion of the mechanical behavior of polymers broadly involves two aspects — the
macroscopic and the microscopic. The former is akin to what we did when studying metals,
namely writing down phenomenological laws based on experiments and laboratory tests. The
latter involves many mathematical nuances and only qualitative statements will be made in
this regard. Of course, for engineering applications, a quantitative macroscopic description
is of primary interest. This is the route we will pursue now.
As we’ve seen, the mechanical response of polymers depends crucially on experimental
time scales, or loading rates. In addition, it is observed that temperature and applied strain
also play a key role. At low temperatures (or short time scales), a polymer may behave like a
brittle solid (glassy) with a Young’s modulus of 109 − 1010 Nm−2 . They will hence break and
flow at very low strains (∼ 5 − 10%). At high temperatures (or long time scales) the same
polymer might behave like rubber, being highly elastic with very low Young’s modulus (106
Nm−2 ) and extremely high elongations (strains of 1 or more) without permanent deformation.
In between these glass–like and rubber–like regimes, a region commonly referred to as
the glass transition range, polymers how intermediate behavior. This cross–over regime
is characterized by a temperature Tg , the glass transition temperature, at which a sharp
change in mechanical properties (volume expansion coefficient, Young’s modulus etc.) is
observed.

Figure 1: Load–elongation curves for a typical polymer at different temperatures. A, B, C,


D represent brittle fracture, ductile fracture, cold–drawing and rubber–like elasticity respec-
tively.

4
The entire range of behavior can be summarized by considering a combined plot of a
load–elongation curve under these different conditions (temperature low or high, time scale
small or large), see Fig. 1. Curve A, typical of glassy polymers (such as PMMA at room
temperature), shows brittle behavior. A small elastic range exists, beyond which fracture
occurs instantaneously. In contrast, curve D shows rubber–like behavior typical at high
temperatures — the polymer undergoes very large reversible elongations and snaps beyond
a critical applied load. An important observation here is that the stress–strain relationship
is nonlinear, even though the material is elastic, implying that Hooke’s law is not a great fit
for rubbers.
At an intermediate temperature, somewhere below Tg , the stress–strain graph resembles
that of a ductile metal (curve B). As we approach Tg from below, the polymer can be cold–
drawn to very large strains (as high as 5 or more) before it necks and eventually fails. Just
like changing temperature, changing the time scale of observation also has a similar effect.
To get started, let us consider a sample in shear, in the xy–plane. The shear strain and
shear stresses are exy and σxy respectively. If the sample material were a viscous fluid or an
elastic body, we have the following relations:

dexy
σxy = η Viscous fluid (3.1)
dt
σxy = Gexy Elastic solid (3.2)

with η and G being the viscosity and shear modulus respectively. One possible formulation
for a viscoelastic material is to combine these characteristics.

dexy
σxy = Gexy + η (3.3)
dt

We will see later on, when studying viscoelasticity in more detail, that other descriptions are
also necessary. In order to demonstrate the concept of creep, the above model will suffice,
see fig. 2. To explain the related idea of stress relaxation, another model will be needed.

Creep:
When a sample is loaded at constant stress σ0 , the corresponding strain developed depends
on the nature of the material. For a perfectly linear elastic material, the strain e0 is given by
Hooke’s law. However, for a viscoelastic solid of the form Eq. 3.3, the strains developed are
time–dependent, and can be solved from this ODE. As shown in the figure, an immediate
elastic deformation e1 and delayed elastic deformation e2 are developed. In the middle (when
σ0 is kept constant), an increase is observed in the strain, even though the load is constant.

5
Figure 2: Comparison between the deformation of an elastic solid and a viscoelastic solid for
constant strain loading.

This is termed creep.


In general, we can write
e(t) = J(t)σ0 (3.4)

where J(t) is termed the creep compliance of the material. In general,

J(t) = J1 (t) + J2 (t) + J3 (t) (3.5)

each Ji (t) corresponding to ei (t). J3 (t) corresponds to the viscous flow part of the response.
J1 (t) and J2 (t) correspond to the instantaneous and delayed responses of the material.
We can now quantitatively describe the various material behaviors, depending on the
nature of J(t). This is shown in Fig. 3, as a function of time. We can now clearly distinguish
the regimes we outlined earlier. At low times, the material behaves like a glassy solid, with
very low compliance (or high modulus). At high times, the material behaves like a rubber
(high compliance or low modulus) and in between it behaves like a viscoelastic material. Of
course, for really large time scales, flow is observed unless the material is chemically treated
or cross–linked (eg., vulcanization). The typical material response time scale τ will fall in
one of these regimes, determining the character of the material’s response. This is, of course,
related to the microscopic properties of the particular polymer.
Cross–linked polymers (like thermosets) do not have an appreciable J3 (t) term, whereas
most other polymers do, particularly at high temperatures. The temperature dependence of

6
Figure 3: Creep compliance as a function of time for typical polymers.

J(t) is very similar to time diagram above.

Stress relaxation:
Stress relaxation is analogous to creep, but now for a constant strain loading. A very simi-
lar diagram can be constructed for the resulting stress relaxation modulus G(t), analogous
to J(t) showing distinct behavior for the three different mechanical response regimes.
To understand this entire spectrum of mechanical response, each regime is usually isolated
and treated differently. This is the approach we will follow, beginning with rubber elasticity.

4 Rubber–like elasticity
As we have seen, rubber–like elasticity is a nonlinear phenomenon and has its own unique
properties. One of the reasons for this is microscopic. In an ideal metal, the atoms are
arranged in regular pattern and any reversible or elastic deformation involves reasonably
small motions of these atoms from their original position. They are like balls attached by
springs, if pulled lightly they extend and return to their original locations when the pulling
load is removed. This is the origin of elastic behavior.
The onset of elastic behavior in polymers, on the other hand, is completely different. As
we’ve seen, polymers are made up of coiled up long-chained molecules. This coiled–up state
presents a low–energy configuration for the chains. Any applied load (like a pulling force)
uncoils them and causes a macroscopic extension. Once the load is released, the chains want

7
to coil back to their low energy configuration, hence producing a complete recovery. This is
the basic picture for elasticity in polymers.
Due to this coiling/ uncoiling, rubbers show a very peculiar effect during deformation,
called the Gough–Joule effect. We begin by discussing this first.
Let us take a piece of rubber (like a rubber–band) and stretch it under a constant dead
load. If the rubber is heated in the process, it contracts in length. Alternatively, if the
same sample is kept at a fixed extension and heated, the tension in the band increases.
This behavior is completely opposite that observed in say metals, where heating causes an
expansion of the material.
An exactly analogous effect is the following. If the rubber is suddenly stretched, a tem-
perature reduction occurs. The heating and cooling are both reversible and can be performed
one after another. This unique behavior is referred to as the Gough–Joule effect. It finds
applications in the design of O-rings and rotary seals in high–temperature applications. If
the seals are prestressed, an increase in temperature will cause them to ‘tighten up’ instead
of expand and lose contact.
The Gough–Joule effect is a remarkable demonstration that elasticity in rubber is fun-
damentally different from the elasticity usually observed in metals. We can write down the
common characteristics of all rubber–like polymers as follows:

1. High extensibility (strains > 5), nonlinear (doesn’t obey Hooke’s law)

2. Complete recovery after stress removal

3. ‘Entropy–induced’ elasticity, as demonstrated by the Gough–Joule effect

4. Volume constant (nearly) during deformation

We have already discussed the first three features above. The fourth one is another
common observation — under almost any (elastic) deformation of a rubber–like polymer
sample, the volume remains constant. One must be careful however, in comparing this with
volume constancy in plastic deformation in metals. In the latter case, the deformation is
permanent and no recovery occurs.
Again, in the big scheme of things, all of this behavior occurs above the glass transition
temperature Tg for the particular polymer, and is driven microscopically by an ensemble of
entangled chains. Having gained knowledge of the general features of rubber–like elasticity,
we must develop some model to be able to describe its effects quantitatively.

8
5 Modeling rubber–like elasticity
To develop a usable model for rubber–like mechanical behavior, we start by considering the
microscopic picture. Any sample of a rubber–like material contains hundreds of thousands
of coiled–up chains, entangled with one another, see Fig. 4 (left). Consider a single chain,
with end-to-end vector R,~ as depicted in Fig. 4 (right). The ends of this particular chain
correspond to crosslinks, that are either physical or chemical.

Figure 4: (Left) Many polymer chains entangled inside the rubber. Two chains are shown
here, between crosslink points (marked by crosses). (Right) Single rubber chain considered
~
in isolation, with end-to-end vector R.

This single chain, in isolation, is still a macroscopic object. So we can apply some ideas
from thermodynamics to describe its response to an external ‘disturbance’. In thermody-
namics, the Helmholtz Free energy A is given by

A = E − TS (5.1)

E is the internal energy, T and S are temperature and entropy respectively.


This quantity A can be thought of as analogous to the potential energy in mechanics. If a
ball is placed on top of a hill, it has high potential energy. Consequently, the force of gravity
attempts to bring it down to the ground, where it has minimum potential energy. Likewise,
if the ball is in a valley, it is already in a state of minimum potential energy, so it will not
get out.
In a similar manner, for a thermodynamic system that has energetic interactions with its
surroundings, E is the potential energy term, but A measures the effective energy available

9
for doing work. The presence of entropy (or say heat loss) reduces this, resulting in the
second term with a negative sign. So any system left to itself will attempt to reach a state of
minimum free energy A, just as our ball (an athermal system) attempts to lower its potential
energy by rolling off the hill.
For our single chain, Fig. 4, the free energy A is a function of the end-to-end distance
~ Left to itself, the chain wants to reduce this distance (equilibrium) and so any change
R.
~ from the equilibrium state causes a restoring force (again compare gravity in the ball
in R
example). From thermodynamics, this force is given by (temperature is kept constant)

F~ = ∇A = ∇E − T ∇S (5.2)

~ but the entropy S


It turns out that for metals, the internal energy E is a function of R
is not. Hence only the first term contributes to the force. For polymers on the other hand,
~ but S is, and so the restoring force
the opposite situation arises — E is not a function of R,
arises due to an entropy contribution.
With this knowledge, we can find an expression for the force if we have an expression for
~ There are multiple ways to obtain this information, the simplest example is that of
S(R).
an ideal Gaussian chain, for which

3R2
 
~ = −k
S(R) + const. (5.3)
2 < R2 >

~ · R,
where R = R ~ k = 1.38 × 1023 J/K is Boltzmann’s constant, T is temperature in Kelvin
~ ·R
and < R2 > is the average value of R ~ over all the chains in the material1 , and is given by

N
2 1 X ~ (i) ~ (i)
< R >= R ·R (5.4)
N i=1

~ (i) for the ith chain. The constant in Eq. 5.3 is irrelevant
for N chains, with end-to-end vector R
for all future calculations and can be set to 0.
Using this model for S(R), the force on changing the end-to-end vector of a single chain,
given by Eq. 5.2, simplifies to
3kT ~
F~ = R (5.5)
< R2 >
Let us now put N such chains together, with end-to-end vector R ~ (i) , i = 1, . . . , N .
Let us also fix a coordinate system, in which each end-to-end vector R ~ (i) has components
1
Remember, we are considering a single chain here. The material actually contains hundreds of thousands
of such chains.

10
(x(i) , y (i) , z (i) ). Assume homogeneous deformation, i.e. that each point in the material
(and consequently every chain) is deformed to the same extent. This is most often the case
as rubbers do not develop localized deformation zones, such as necks.
The work done W (i) in extending 1 chain from initial end-to-end vector (x(i) , y (i) , z (i) ) to
final value (λx x(i) , λy y (i) , λz z (i) ) is given by
Z (λx x(i) ,λy y (i) ,λz z (i) )
W (i)
= F~ (i) · d~l(i) (5.6)
(x(i) ,y (i) ,z (i) )

where the infinitesimal length change ~l(i) is (dx(i) , dy (i) , dz (i) ). Using the expression in Eq. 5.5,
this can be evaluated to

3kT
W (i) =
 2 (i) 2 2 (i) 2 2 (i) 2

(λx − 1)(x ) + (λ y − 1)(y ) + (λz − 1)(z ) (5.7)
2 < R2 >

For all N chains in the material, the total work done is a summation

N
X
W = W (i) (5.8)
i=1

Using the relations in Eq. 5.7 and the definition of the mean square for x(i) , y (i) , z (i) , we
can write Eq. 5.8 as

3N kT  2 2 2 2 2 2

W = (λ x − 1) < x > +(λy − 1) < y > +(λz − 1) < z > (5.9)
2 < R2 >

By definition R~ (i) · R
~ (i) = (x(i) )2 + (y (i) )2 + (z (i) )2 . For an isotropic rubber material, where
chains are distributed with equal probability in all directions, we can make the additional
2
approximation that < x2 >=< y 2 >=< z 2 > hence each being equal to <R3 > . Using this
simplification, the expression for W becomes

N kT  2
λx + λ2y + λ2z − 3

W = (5.10)
2
If we define
G = N kT /V (5.11)

11
where V is the total volume2 of the material, then writing W in terms of G gives us:

VG 2
λx + λ2y + λ2z − 3

W = (5.12)
2

which is an expression for the energy change when the body is deformed homogeneously with
stretches λx , λy , λz respectively.
All of the macroscopic information regarding the deformation is contained in Eq. 5.12,
which we have derived from a microscopic description. Remember that the crucial quan-
titative entity in our picture of entangled polymer chains (Fig. 4) is that of the entropy
expression, Eq. 5.3. Since we claimed that this expression holds for the so–called ideal Gaus-
~ For this reason, we will refer to any rubber
sian chain, which is one possible model for S(R).
that obeys Eq. 5.12 as an ideal Gaussian chain rubber.
If we forget our expression for G, and treat it as a mere constant, Eq. 5.12 by itself does
not contain any indications of having arisen from a microscopic model. It does, however,
contain all the information we need for obtaining macroscopic quantities, such as the stresses,
for given stretch ratios (the λs). Exactly how this is done will be explained with specific
examples in the next section. So, in principle, one could postulate an expression for W , and
proceed analogously without worrying about chains at all. In fact historically, a description
of rubber–like elasticity was first formulated in this manner, with subsequent advances in
microscopy and chemical analysis enabling a more detailed microscopic picture.
Taking this latter, purely phenomenological approach, we can postulate the following
energy function, instead of Eq. 5.12,
  
1 1 1
C1 λ2x λ2y λ2z

W =V + + − 3 + C2 2
+ 2+ 2 (5.13)
λx λy λz

this is a more general version of Eq. 5.12, which corresponds to C1 = G/2, C2 = 0. C1 and
C2 are general fitting constants, whose values are typically obtained from experiments.
A rubber that has an energy function given by Eq. 5.13 instead of Eq. 5.12 is called
a Mooney rubber. It turns out that the Mooney expression, Eq. 5.13 is a much better
approximation for rubber elasticity at very high strains (λ ∼ 5) whereas the ideal chain
approximation that led to Eq. 5.12 works well for moderate strains λ ∼ 2.
Just like the above, we can come up with a number of other energy function expressions,
subject to certain symmetry constraints, each describing some aspect of a particular type
of rubber. Examples of these are Gent rubbers, Blatz–Ko rubbers (such as foams) and
2
It does not matter which volume we consider here, since V is constant during deformation for rubber–like
materials

12
St.Venant–Kirchhoff rubbers which are anisotropic. We will only consider the ideal Gaussian
chain and Mooney rubbers in the rest of these notes.

6 Examples
Our fundamental recipe now for studying macroscopic behavior is simple. We start with an
expression for the energy, either Eq. 5.12 or Eq. 5.13, depending on the exact rubbery polymer
under consideration. These were derived for homogeneous deformation, with stretches in
three perpendicular directions, resulting only from normal stresses. One can reduce any
complicated loading condition, involving both normal and shear strains, to this configuration
by orienting the axes x, y, z along the three principal directions of the stress tensor3 . In light
of this, let us consider some simple loading configurations

Tensile test

Figure 5: Tensile test of a rubber sample.

Consider a rubber sample, with square cross section (area a2 ) and length L0 . Let the
final length, under application of a tensile load F , be λ0 L0 . We want to calculate the tensile
3
If you recall from linear algebra, for axes along the principal directions, the stress tensor becomes diagonal,
and shear stresses are zero. We have encountered the 2D analogoue of this when discussing principal directions
using the Mohr circle.

13
stress in the sample. First, since the cross section is a square, λs corresponding to the in–
plane deformation must be equal (isotropy). Denote this value by λ̃. We know from volume
conservation,
1
L0 a2 = λ0 L0 λ̃2 a2 =⇒ λ̃ = √ (6.1)
λ0
−1/2 −1/2
Using the values λ0 , λ0 , λ0 , we can write W as
 
GV 2
W = λ20 + −3 (6.2)
2 λ0

under the application of a force F , a change occurs in W . If the final length Lf = λ0 L0 , then
the force at any instant is given by the derivative of the energy

dW dW dλ0
F = = (6.3)
dLf dλ0 dLf

where we have used the chain rule in the second expression. Since the final length Lf = λ0 L0 ,
we can use Eq. 6.2 with this expression, to obtain
 
GV 1
F = λ0 − 2 (6.4)
L0 λ0

and since V = L0 a2 , the engineering stress σ0 = F/a2 is given by


 
1
σ0 = G λ0 − 2 (6.5)
λ0

For a linear elastic material obeying Hooke’s law, tensile the stress is proportional to the
strain (or the stretch λ0 ). Clearly, the expression in Eq. 6.5 is nonlinear in λ0 due to the
second term on the right hand side.

Linear approximation

To see the connection between this expression and Hooke’s law as we know it, we consider the
linear approximation next. This is valid when the strains are very small, and consequently,
the stretches λ0 → 1. We can expand the expression in Fig. 6.5 to the case when λ0 − 1 → 0,
by retaining only the lowest order terms. Doing this one arrives at

σ0 = 3G (λ0 − 1) (6.6)

where λ0 − 1 is, by definition, equal to the strain ∆L/L0 = (Lf − L0 )/L0 .

14
In linear elasticity, the tensile stress is related to the tensile strain by the Young’s modulus
E. Further, for volume conservation in linear elasticity, the Poisson ratio = 1/2 thereby
relating the shear modulus to the Young’s modulus by E = 3G. In view of Eq. 5.11, we have
an expression for the shear modulus G in terms of the polymer chain parameters N , its
temperature T and volume V .
Thus, our analysis has not only yielded an expression for the energy but in addition has
also provided insight into what physically governs the magnitude of the elastic modulus. By
increasing N , i.e. the number of chains between crosslinks in the polymer, we can increase
the stiffness of the polymer (G increases). This is precisely what happens when rubber is
vulcanized; the sulphur atoms create more crosslinks in the chain network increasing N .
This expression for G can also be used to explain the Gough–Joule effect. Rewriting
Eqs. 6.6 and Eq. 5.11 to expression λ0 in terms of σ0 ,
  
σ0 V 1
λ0 = +1 (6.7)
3N k T

For a constant applied σ0 , if the temperature T is increased, the stretch reduces or the rubber
contracts, which is precisely the Gough–Joule effect.
Before concluding this example, we compare the predictions obtained from linear elastic-
ity, Eq. 6.6 with that of the nonlinear stress, Eq. 6.5. Some sample stretch–stress values are
given below, obtained by solving for the stretch for a constant applied stress . N L stands for
nonlinear and L is linear.

σ0 = 0.03G =⇒ λN
0
L
= 1.01, λL0 = 1.01 (6.8)
σ0 = 1.75G =⇒ λN
0
L
= 2.00, λL0 = 1.38 (6.9)

Clearly, as λ0 goes farther away from one, naı̈vely assuming the material to obey Hooke’s
law will increasingly under -predict the stretch resulting from an applied load.

Biaxial loading of a rubber sheet

Our next example is the biaxial loading of a rubber sheet, assumed to be ideal Gaussian,
shown in Fig. 6. Let us solve the following problem. We know that the final desired stretches
in the x and y directions, both coinciding with the edges of the sheet as in the figure, are λx
and λy respectively. What are the required engineering stresses if λx = 3/2 and λy = 2/3?
Assume the shear modulus G = 1 MPa.
In general, we denote the stretch in the thickness direction as λz . From volume conser-

15
Figure 6: Biaxial deformation of a rubber sheet

vation, we have
1
λz = (6.10)
λx λy
Just as we did for the tensile test, the forces are given as gradients of the energy W , only
this time we have to take partial derivatives with respect to λx and λy for obtaining Fx and
Fy respectively.

∂W dλx ∂W dλy
Fx = Fy = (6.11)
∂λx dLx ∂λy dLy
Using the expression for an ideal Gaussian rubber, Eq. 5.12, and the substituting for λz
in terms of λx , λy , we obtain expressions for Fx , Fy as
   
Fx 1 Fy 1
σx = = G λx − 3 2 and σy = = G λy − 3 2 (6.12)
A λx λy A λy λx

for the given values, σx = 0.833 MPa and σy = −0.833 MPa.


Finally, a note of caution is in order when working on 2D problems such as this. From
the given λx , λy values, λz = 1 due to volume conservation. However, this cannot be put
into the expression for W directly, since partial derivatives are needed. Hence λz is kept a
function of λx , λy throughout and values are substituted only in the final step.

Other loading geometries and W expressions

We have seen examples of rubber behavior for ideal Gaussian chain rubbers. It has also been
mentioned that any complex 3D loading can be reduced to stretches in three orthogonal

16
directions x, y, z by changing directions to point along the principal axes. In solving the
tensile and sheet problems above, we have only utilized the expressions for W , so one can
also do the exact same thing for a Mooney rubber, only by starting with Eq. 5.13 instead of
Eq. 5.12.
The Mooney rubber relation, Eq. 5.13 better predicts stretches at large λ(' 2.5) when
compared to the ideal Gaussian model. There are additional models for rubber elasticity that
are commonly used, depending on the circumstances. Most of these are phenomenological
(like the Mooney rubber) and not based on a microscopic model (as was the ideal Gaussian
rubber). We list them here for completeness

• The Gent model is used for stretches even larger than those with the Mooney model

• The Blatz–Ko model is used for modeling foam rubbers, such as polyurethane foams

• The St.Venant–Kirchhoff model is used for modeling general anisotropic rubbers, which
we haven’t considered here

7 Engineering Rubbers
To conclude our brief study of rubbers, we make some comments about commercial rubbers
and their production. Historically, natural rubber, chemically termed polyisoprene, was the
most commonly used rubber. It was obtained from rubber trees that only grow in tropical
climates. Natural rubber, as obtained from the tree sap, has the undesirable property of
being mechanically weak and prone to flow over time. They are also swollen by many organic
liquids. In order to counter these properties, two primary processing routes are adopted:

• Vulcanization, as already alluded to earlier, is a process where the rubber (natural


or synthetic) is heated in sulphur vapor. The sulphur atoms create bridges (crosslinks)
between chains in the polymer, thereby increasing its modulus.

• Addition of fillers to rubber (natural or synthetic), most commonly carbon black,


has a number of advantages. Firstly, carbon black acts as a mechanical reinforcement
to the polymer chains thereby increasing stiffness. Secondly, it also reduces degradation
due to sunlight, ozone and liquid absorption. However, the addition of fillers to rubbers
results in a characteristic mechanical response known as the Mullins effect, wherein
the stress–strain curve behavior depends on maximum loading.

In order to meet the large demands for rubber following the second world war, synthetic
rubber production increased sharply in the 1950s and later. Examples of synthetic rubbers
include polybutadiene and synthetic polyisoprene. A particular form of synthetic rubbers are

17
produced as copolymers. These are polymers produced from two or more types of monomer
units4 . The most common examples of this are

• Styrene-Butadiene rubber (SBR) made from monomers butadiene + styrene.


Nearly half of all automobile tires are produced from SBR. Addition of styrene in-
creases the stiffness of the polymer, but also increases its glass transition temperature.
Hence, a balance between operating temperature (which will determine the allowed Tg )
and the % of styrene (to obtain suitable stiffness) must be used while producing SBR.

• NBR made from acrylonitrile + butadiene. It has high resistance to chemicals and
oils, and is often used to make surgical gloves.

Finally, a number of synthetic rubbers exist, depending on the application. They are
particularly useful and advantageous over natural rubber since their properties can be fun-
damentally altered by chemical modification.

8 Viscoelasticity
Going back to the creep function curve, Fig. 3, we now discuss the central viscoelastic zone
where J(t) exhibits significant time dependence. There are two general features of viscoelastic
polymers,

• They do not recover completely after stress removal

• They flow under constant stress (creep)

For all polymer materials in general, it is important to know what the operating conditions
are — the timescale for use as well as the intended temperature range — as these determine
the polymer’s behavior completely.
As far as mechanical loading is concerned, there are three types of loading conditions
from which all the mechanical responses can be derived:

1. Creep: A constant stress σ0 is applied and the strain e(t) = J(t)σ0 is a function of
time.

2. Stress relaxation: A constant strain e0 is applied and the stress decays as a function
of time σ0 = G(t)e0 where G(t) is the stress relaxation function. Just like the
J(t) graph (Fig. 3) G(t) exhibits plateaus in the glass and rubber regions, termed the
unrelaxed modulus GU and relaxed modulus GR respectively.
4
For instance, polyisoprene has a single monomer unit (isoprene) and polybutadiene has a sinlge one too
(butadiene). Such single monomer polymers are called homopolymers

18
3. Oscillatory loading: A sinusoidal stress (or strain) is applied and the resulting strain
(or stress) is observed. The response, in general, is a function of loading frequency
ω.

The most important property that is assumed to hold for viscoelastic materials is linearity.
By this, we mean that the functions J(t) or G(t) do not depend on the stresses or strains.
For this reason, the behavior we describe in the next few sections is referred to as linear
viscoelasticity. At higher strains/ stresses this assumption of linearity breaks down, leading
to nonlinearities, yielding, necking etc. We now discuss each of the above loading conditions
separately. The material introduced in these sections can be found in more detailed in
Refs.[3, 4, 5].

Creep and stress relaxation

The creep compliance J(t) and stress relaxation G(t) functions provide information on the
mechanical response of the polymer at constant loads. In order to determine them experi-
mentally, the torsional configuration in Fig. 7 is used.

Figure 7: Experimentally determining the creep or stress relaxation functions.

Consider the thin-walled cylinder shown in the figure, with radius r, length l and wall
thickness s. For creep, let a constant torque Γ be applied to the cylinder, the circumferential

19
force is F = Γ/r. The shear stress is given by
  
Γ 1
τ= (8.1)
r 2πrs

For the tube, the shear strain γ(t) can be related to the change in angle θ(t) by

rθ(t) = γ(t)l (8.2)

Using the relation between stress and strain (e(t) = σ0 J(t)), we have an expression for
J(t) as
2πr3 s θ(t)
 
J(t) = (8.3)
l Γ
Following the same sequence for a constant angular rotation θ, the stress relaxation func-
tion G(t) is given by
 
l Γ(t)
G(t) = 3
(8.4)
2πr s θ
This provides a method to calculate the stress relaxation or creep compliance functions
by measuring either the change in torque or the angular rotation with time respectively.

Oscillatory loading

When you strike a piece of metal, you hear a constant ring with a definite pitch that can last
for several seconds. On the other hand, striking a piece of polymer (like the chair you are
sitting on) with the same force (or any force for that matter) results in a muted nondescript
sound. The reason for this characteristic behavior lies in the natural damping behavior of
polymers. They therefore find a lot of use in damping applications (shock absorbers, artificial
tissue etc.).
Let us consider the application of an oscillatory strain e(t) = e0 sin ωt and characterize
the stress response; application of an oscillatory stress is treated analogously. The measured
stress is found to vary with time as

σ(t) = σ0 sin(ωt + δ) (8.5)

where σ0 and δ are apriori unknown. Expanding the sine function, we can write the stress as

σ(t) = e0 (G0 sin ωt + G00 cos ωt) (8.6)

20
where G0 = σ0 cos δ/e0 and G00 = σ0 sin δ/e0 are called the storage and loss moduli
respectively. Also, tan δ = G00 /G0 .
δ is an angle; for elastic solids δ = 0, for metals, δ ' 0 but for polymers, δ can be as
high as 30◦ . In general, δ is a function of the applied frequency ω, hence G0 and G00 are
also functions of ω. Further, any applied strain e(t) can be decomposed in a Fourier series
containing different ω-s and the stress response is given for each such ω by the G0 (ω) and
G00 (ω) functions.
To see the significance of the loss modulus, we can compute the energy dissipation per
cycle ∆E = σde = σ( de
R R
dt
)dt. Using Eq. 8.6 and the definition e(t) = e0 sin ωt, we can
evaluate ∆E as
∆E = πe20 G00 (8.7)

which clearly shows us that the loss modulus is directly proportional to the energy lost per
cycle of loading.

Figure 8: Frequency response for the storage and loss moduli.

As mentioned earlier, an analogous treatment for strain response in an oscillatory stress


loading gives the creep functions J 0 and J 00 where the latter can again be related to the
energy lost per cycle. The functions G0 (ω), G00 (ω), tan δ are shown in Fig. 8. Clearly, there
is a large difference in the behaviors at extreme and intermediate ω values. The graph is
qualitatively similar to the one in Fig. 3 (with due note of the change from J(t) to G(t))
and shows why losses are important in the viscoelastic regime. The G00 (ω) curve is non–zero
only in the viscoelastic regime, being nearly zero in the rubber–like and glassy regimes. This
is consistent with the definition of ∆E in the equation above as well as the elastic behavior
observed in these regimes. This explains the nature of the damping observed in viscoelastic

21
polymers. At extremely low and high frequencies, the mechanical response (characterized by
GU and GR ) is elastic (reversible with minimal damping) while the peak occurs in the middle
regime.

Zener model

Now that we know what the functions J(t), G(t), G0 (ω) and G00 (ω) mean, we need a model
to estimate them. The simplest model that can capture all three loading conditions (creep,
stress relaxation and oscillatory loading) is called the Zener model or the Standard Linear
Solid model. This model is often presented in terms of springs and dashpots, see for e.g.,
Ref. [3] but we will only use the governing differential equation. The Zener model has three
parameters JR , JU and τ or equivalently GR (= 1/JR ), GU (= 1/JU ) and τ̃ = τ JU /JR . These
constants a priori have no physical significance. The governing equation, relating the stress
σ(t) to the strain e(t) is given by

de dσ
e(t) + τ = JR σ(t) + τ JU (8.8)
dt dt

or equivalently by
 
de dσ
GR e(t) + τ = σ(t) + τ̃ (8.9)
dt dt

This 3-parameter model can be used to find expressions for J(t), G(t), G0 (ω) and G00 (ω)
in terms of the parameters JU , JR , τ (or GU , GR , τ̃ ).

1. Creep: Setting σ(t) = σ0 and solving Eq. 8.8 for e(t) with initial conditions e(0) = JU σ0
(glassy response at short times), we obtain an expression for J(t) = e(t)/σ0 as
  
t
J(t) = JU + (JR − JU ) 1 − exp − (8.10)
τ

where now the parameters JU and JR make sense as the unrelaxed and relaxed creep
compliances. τ is the relaxation time for the creep function.

2. Stress relaxation: Setting e(t) = e0 and solving Eq. 8.9 for σ(t) with initial conditions
σ(0) = GU e0 (glassy response at short times), we obtain an expression for G(t) =
σ(t)/e0
  
t
G(t) = GU − (GU − GR ) 1 − exp − (8.11)
τ̃
and the parameters GR , GU correspond to the relaxed and unrelaxed moduli. The

22
relaxation time now is τ̃ instead of just τ .

3. Oscillatory behavior: We insert the expressions for the applied strain e(t) = e0 sin ωt
and resulting stress σ(t) = e0 (G0 sin ωt + G00 cos ωt) into either of Eqs. 8.8 or 8.9, to
obtain expressions for G00 (ω) and G0 (ω) as

(GU − GR )ω 2 τ̃ 2 (GU − GR )ωτ̃


G0 = GR + G00 = (8.12)
1 + ω 2 τ̃ 2 1 + ω 2 τ̃ 2

and equivalently for a sinuousidal applied stress, the strain function gives expressions
for J 0 (ω), J 00 (ω) as

(JR − JU ) (JR − JU )ωτ


J 0 = JU + J 00 = (8.13)
1 + ω2τ 2 1 + ω2τ 2

A plot of these expressions for the various response functions looks very similar to those in
Fig. 8. The model parameter values are fixed to precisely fit the values for the particular
polymer considered.
The maxima in the curves for G00 (ω) and J 00 (ω) occur at ω = 1/τ̃ and ω = 1/τ respectively.
The maximum in the tan δ curve (obtained from the ratio of G00 /G0 above) occurs at ω =

1/ τ τ̃ . It is quite intriguing that the viscoelastic response, as described by the Zener model,
is very closely tied in with the rubber–like and glassy responses too. Hence, by just fitting 3
parameters, the Zener model gives us a very good analytical approximation for the response
functions in linear viscoelasticity.

9 Temperature dependence of polymer behavior


As mentioned at the beginning of the previous section, the mechanical response of polymers
is acutely dependent on the temperature. We now quantify this statement by recourse to the
Zener model; the ideas in this section are not a feature of the Zener model alone; polymers
whose response functions are not well approximated by the model still follow these general
principles, with a few exceptions. But for the purpose of illustration, we will use the relations
for the Zener model in the following.
Temperature dependence in the Zener model is reflected in the 3 parameters GR , GU and
τ̃ . The parameters GR and GU are practically independent of temperature (the glassy and
rubber–like responses are in general temperature-agnostic) and so temperature dependence
occurs only in the relaxation time τ̃ , or equivalently in τ . Let τ1 and τ0 be the relaxation

23
times at temperatures T1 and T0 . Experiments reveal that
  
∆H 1 1
τ1 = aT τ0 where aT = exp − (9.1)
R T1 T0

the coefficient aT is called the shift factor. In the equation above, T1 > T0 and ∆H > 0 is
called the activation enthalpy. The universal gas constant, R = 8.31 J/(molK).
From the expression for J(t) in the Zener model, Eq. 8.10, for the creep compliance
functions J 1 (t) and J 0 (t) at temperatures T1 and T0 respectively, we have to use τ1 and τ0
respectively, alongwith the same JR and JU values. It can be seen at once, using Eq. 9.1,
that
J 0 (t) = J 1 (aT t) (9.2)

so the creep at a lower temperature T0 is obtained from the creep function (if known) at
a higher temperature T1 by scaling the time with the shift factor aT . Hence we need only
1 J(t) curve, obtained at some suitable temperature, from which we can obtain suitable
response curves at other higher or lower temperatures using Eq. 9.2 and the expression
for aT in Eq. 9.1 once ∆H is known. This principle is commonly referred to as the time–
Temperature superposition. Detailed experimental temperature data for typical polymer
materials can be found in chapter 7 of Ref. [3].

10 Examples
We now present some examples for calculating viscoelastic behaviors in practical situations.

Zener model approximation

As stated in the previous section, the Zener model is a very useful way of describing response
functions once reasonable estimates are available for the 3 parameters. Consider for instance
the response for a certain polymer, fit to the Zener model

J(t) = 1.2 + 0.5(1 − exp(−0.1t)) GPa−1 (10.1)

where t is in hours. If this creep function is given at 35◦ C, what is the creep function at
40◦ ? Assume ∆H = 170 kJ/mol, and R = 8.31 J/(mol.K).
We can solve this problem using time–temperature superposition. Calculate aT from
Eq. 9.2 using the given data. We obtain aT = 0.346, a dimensionless constant. Using this,

24
we can obtain the behavior at higher temperature using Eq. 9.2 as
 
40◦ C t
J (t) = J = 1.2 + 0.5(1 − exp(−t/3.46)) (10.2)
aT

It is important to remember that even if the polymer does not obey the Zener model, but
just showed time–temperature correspondence, we can still use Eqs. 9.1 and 9.2 to obtain
the corresponding J(t) expression.

Superposition principle permits short–time experiments

One practical advantage of the time–temperature superposition principle is the ability to


perform short–time experiments to infer long–time behavior. Let us consider the following
scenario. Suppose we have a new polymer X which is supposed to go into a part. The part
itself is supposed to be usable for 10 years. We want to know how much creep the polymer
will show in 10 years so that we can evaluate its applicability for making said part. Let us
assume that the part is to be used under nearly constant temperature conditions of 300 K.
Assume ∆H for the polymer is 120 kJ/mol.
Clearly, since the polymer is new, we do not have creep data for 10 years so what do
we do? let us say we perform tests at a higher temperature, say 350 K, for which the shift
factor aT is 10−3 (T0 = 300 K and T1 = 350 K). So in order to obtain time dependence of
the creep for 10 years at 300 K, we only need to perform experiments at 10 × 10−3 years or
close to 90 hours. Therefore, the characterization can be completed at a higher temperature
in a week and time–temperature superposition can be used to obtain data for 10 years at
300 K. This is a very practical example of the utility of the time–temperature superposition.
Notice, finally, that we didn’t need to assume the polymer obeyed the Zener model at all!

11 Yielding
From the mechanical behavior in Fig. 1, we recall three distinct responses — rubber–like
elasticity, discussed earlier, the linear viscoelastic regime (roughly the linear part of curve
C) and the glass–like regime. Notice first that this figure is distinct from Fig. 3, as the axes
contain stress and strain, as opposed to stress or strain vs. time. For viscoelastic materials
that show significant strains, the deformation is no longer linear ( i.e. stress ∝ strain).
Under significant additional straining, the material begins to yield, just as with metals, see
the maximum point of curve C in Fig. 1.
For rubbers, even when the stress–strain curve is nonlinear, the deformation is completely
reversible. In contrast, the nonlinear regime for viscoelastic materials is not easily separated

25
into elastic and plastic regimes. Obviously, the recovery depends on other factors such as
time, temperature etc. as we have already noted. However, a common feature in such
polymers is the formation of a localized neck, which eventually propagates and finally leads
to fracture at large strains of ∼ 3 (end of curve C in Fig. 1).
Neck formation occurs usually at the end of the linear regime, and corresponds to a local
maximum in engineering stress, just as with metals. However, unlike metals, neck formation
is not followed by fracture. In polymers, subsequent to neck formation (maximum point in
curve C), elongation occurs as the neck propagates towards the two ends of the sample. This
process of neck propagation is accompanied by a sharp reduction in cross section all along
the length of the sample. This property of polymers is very unique and finds application in
fiber processing. The criterion for neck formation itself is easily obtained by considering the
maximum in the engineering stress curve, exactly as is done with metals (recall dσ/d = σ).
Yielding is often taken to be simultaneous with neck formation, but that is not necessarily
true. A commonly used criterion for polymer yielding, for a general multi-axial loading, is
the von Mises criterion, which states that

(σ11 − σ22 )2 + (σ22 − σ33 )2 + (σ33 − σ11 )2 + 6(σ12


2 2
+ σ23 2
+ σ31 ) ≥ 6C 2 (11.1)

This criterion is used for treating yielding in metals as well with one major difference. For
polymers, the constant C, which is related to the yield stress, is dependent on the hydrostatic
pressure applied. If we recall, this behavior is not observed in metals (C is independent of
hydrostatic pressure) and hence yielding in compression and tension is equal.
The hydrostatic pressure p is defined as follows

1
p = (σ11 + σ22 + σ33 ) (11.2)
3

Note that sometimes a negative sign is used for the definition of p, the resulting behavior of
C does not depend on this sign convention.
Yielding occurs when the LHS, called the von Mises effective stress, equals the RHS.
If the functional dependence of C(p) is known, we can calculate the corresponding stresses
at which yielding must occur by considering equality in Eq. 11.1. For example, in uniaxial
tension, with stress σ0 , the pressure p = σ0 /3. For uniaxial loading (again tensile stress is
σ0 ) with an additional compressive hydrostatic pressure p0 (say under water), the total value
of p = σ0 /3 − p0 (or −σ0 /3 + p0 depending on the sign convention chosen). Yielding then
occurs when the stresses given by

σ11 = σ0 σ22 = σ33 = σ12 = σ23 = σ31 = 0 (11.3)

26
or by
σ11 = σ0 − p0 /3 σ22 = σ33 = −p0 /3 σ12 = σ23 = σ31 = 0 (11.4)

are put into the LHS of Eq. 11.1 and equated to the corresponding value of C(p), using the
values of p outlined (either σ0 /3 or σ0 /3 − p0 ). In the second case, it will be seen that the
imposed hydrostatic pressure p0 does not appear on the LHS while it will affect the RHS,
via C(p).
The pressure dependence of the yield stress in polymers has two main consequences.
Firstly, the yield stresses in tension and compression are different and secondly, pressure de-
pendence of the yield stress can deter crazing. Crazing and fracture are important phenomena
that play a role in failure.

12 Crazing and fracture


We now move from discussing yielding and large strains to glassy behavior. In the glass
regime, the elastic behavior of polymers is not very interesting; polymer glasses are largely
linear elastic with a relatively high modulus. However, even at very low strains, they can fail
due to crazing and crack propagation (fracture).

Crazing

Crazing is a common precursor for polymer failure, particularly thermoplastic glasses (e.g.,
polypropylene). A craze itself is a long narrow region comprised of a network of fibrils, see
Fig. 9. Crazes nucleate at points of high stress concentration on the sample free surface or on
voids in the bulk. Once a craze, such as the one in Fig. 9 is formed, subsequent deformation
is concentrated around it. In the process, the fibrils themselves can extend by factors of 2 to
5, depending on the temperature and chemical nature of the polymer itself. Once this strain
is reached, the fibrils can no longer extend and fail leading to fracture.
Crazing and fracture caused by absorbed liquids forms a practical limitation on the use
of many glassy plastic materials. This phenomenon is often referred to as Environmental
stress cracking or solvent crazing. Additional examples of crazing and its prevention by
the addition of monomers (such as in polystyrene) can be found in Ref. [5].

Fracture

Material failure by crack propagation or fracture is a common property of brittle materials


and polymeric glasses are no different. We now analyze, by simple means, the conditions
under which cracks can propagate in a sample. In order to keep things simple, we assume

27
Figure 9: Schematic showing crazing behavior in polymers.

that the material is perfectly elastic and fractures instantly, i.e. without complications such
as plastic flow, crazing etc. Such an analysis provides an upper bound estimate on the
stresses that a particular sample can be subjected to in a praticular loading condition. A
more thorough treatment of the case presented here can be found in the book by Lawn [6].
Consider a thin, rectangular sample as shown in Fig. 10, loaded in tension remotely by
tensile stresses σ̄. Consider a narrow slit in the material, representing the crack, of length
2a as shown in the figure. Let the thickness of the material be B. We assume that a 
dimensions of the body and that B is very small compared to the other dimensions. A typical
value for a would be of the order of a few µm.
Taking recourse to energetics, we can argue that the crack grows only if the total energy is
lowered in the process. In the hypothetical situation where the crack expands, say a becomes
1.01a, how will the energy of this system change? Let us assume that we are working with
fixed grips loading (doesn’t matter). Two terms contribute to the energy

• Elastic strain energy arising from the strain due to the imposed σ̄. If the crack were
completely absent, the (linear elastic) material would have energy V σ̄ 2 /(2E), where V
is the energy and E is the Young’s modulus. This is obtained by assuming homogeneous
R
straining with  = σ̄/E and evaluating σ̄d. If the narrow slit is present, this value
is lowered by σ̄ 2 πa2 B/E so that the elastic strain energy, in the presence of the slit
becomes:
V σ̄ 2 σ̄ 2 πa2 B
EM = − (12.1)
2E E
• Surface energy is the energy needed to form a unit surface. Recall that the polymer

28
Figure 10: Plot of ∆U vs. a for a crack in mode I fracture.

chains in the bulk are kept intact due to cohesive forces. In order to separate them
and form two surfaces, we need to put in some energy, here denoted by GC per unit
surface area created. Hence, for creating the slit, we’d have to overcome cohesive forces
by providing an energy equal to
ES = 2aBGc (12.2)

The total energy in the process is therefore

V σ̄ 2 σ̄ 2 πa2 B
∆U = EM + ES = − + 2aBGC (12.3)
2E E

The individual terms, along with ∆U , are shown in Fig. 10 (right). The maximum in the
∆U curve, corresponding to the value amax is obtained from d∆U/da = 0, giving

 1/2
EGC EGC
amax = σ̄F = (12.4)
πσ̄ 2 πa

The expressions for amax and σ̄F are reciprocals of each other and can be thought of in the
following manner. Consider a sample with an initial crack, of fixed length a0 . For σ̄F = 0,
amax → ∞ and a0 < amax . As σ̄F is increased from 0, amax starts reducing until eventually

29
amax = a0 . At this junction, corresponding to the maximum in ∆U , it is more favorable for
the crack to grow (from the graph ∆U can be reduced by increasing a) so that the crack
starts propagating. Alternatively, for a fixed crack of size a in the material, increasing the
stress from 0 onward results in propagation when σ̄ = σ̄F . For these reasons, amax and σ̄F
are called the critial crack length and critical stess for fracture respectively.
Eq. 12.4 provides a criterion for evaluating when cracks propagate in a material. Yet
another way to look at this formula involves the following relations

K = (πa)1/2 σ̄ Kc = (πa)1/2 σ̄F (12.5)

called the stress intensity factor and critical stress intensity factor respectively. Notice that
they involve σ̄ or the applied tensile stress and σ̄F or the critical stress for fracture respectively.
The above description of the criterion can be written as

K < KC no fracture occurs (12.6)

We have a condition that is quite independent of actual a and σ̄F values. In fact Kc is
a material parameter (as can be checked using Eqs. 12.4 and 12.5) and can be recorded
in a handbook. To make sure that a given sample is safe under loading, calculate K using
Eq. 12.5 and compare it with KC . This provides a very simple and useful criterion for fracture
problems.
All of these relations were obtained assuming that the plate (Fig. 10) was very thin. For
thick plates, the solution changes a little bit. In addition, tensile loading is not the only type
encountered. The mode we have considered is commonly called Mode I fracture. Other
modes exist, these are Mode II or sliding mode and Mode III or tearing mode and
a similar analysis can be used for them. The only trick is in evaluating the corresponding
strain energies EM for these loading conditions; the rest of the calculation is the same in
principle. In general, the stress intensity factor can be written as

K = ψσ̄ a (12.7)

where ψ is a numerical factor, depending on which mode is being considered. Of course,



ψ = π for mode I as we have seen above. Hence the method for calculating K and
comparing it to a Kc is very useful for assessing the fracture safety of an engineered part.

30
13 Conclusion
In polymer processing, there broadly are two main types of problems that merit considera-
tion. The first is the mechanical behavior of polymers resulting from particular processing/
chemical methods. The second is the analysis of the manufacturing process itself. Since
most polymer parts are produced via forming from the melt, analysis of processes is common
to almost all polymers. Details of forming processes can be found in Chapter 7 of Ref. [7],
along with a brief introduction to melt analysis. The above notes have mainly dealt with
the first part and have hopefully introduced you to the common vocabulary used to describe
mechanical behavior of solid polymers.

31
Bibliography

[1] L. R. G. Treloar. Introduction to Polymer Science. Wykeham Publications London, 1970.

[2] G. I. Barenblatt. Flow, Deformation and Fracture. Cambridge University Press, 2014.

[3] I. M. Ward. Mechanical Properties of Solid Polymers. John Wiley & Sons Ltd., 1983.

[4] R. J. Young and P. A. Lovell. Introduction to Polymers. Chapman & Hall London, 1991.

[5] N. G. McCrum, C. P. Buckley, and C. B. Bucknall. Principles of Polymer Engineering.


Oxford University Press, 1997.

[6] B. R. Lawn. Fracture of Brittle Solids. Cambridge University Press, 1993.

[7] M. Daoud and C. E. Williams. Soft Matter Physics. Springer, Berlin, 1999.

32

You might also like