Anelastic Behavior in Polymers Class Notes

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

7-1

Lecture 7
Anelastic Behavior in Polymers
We now move on to from linear elastic deformation to anelastic deformation. Although this type
of deformation does occur in metals and ceramics, it is most common and important in polymers.
Thus, we start by developing and understanding of polymer structure and how it leads to
anelastic behavior. We then show a functional form for anelastic deformation and discuss elastic
properties in polymers. In the next lecture, we will apply these concepts to anelastic deformation
in metals and ceramics

Anelastic Deformation
In Lecture 1, we said that anelastic deformation is time-dependent elastic deformation. We also
noted that, since all deformation is time-dependent on some scale, we can make the time-
dependence important by simply conducting our test in a time frame that was similar to the time
scale of the deformation response—i.e. a Deborah number, D, near 1. In principle, we could
achieve this for elastic deformation in a linear elastic crystalline metal or ceramic by simply
loading fast enough that the time of the test was similar to the time it takes elastic waves to
traverse the sample. However, although we might have D ≈ 1 in such a case, we would not
normally use the term “anelastic” to refer that deformation.
Instead there is a fairly clear distinction between linear elastic and anelastic deformation
based on the deformation mechanism. In linear elasticity, atoms do not change their nearest
neighbors. In anelastic deformation, at least some atoms do change their nearest neighbors.
These processes take more time than simply distorting bonds, resulting in much longer
deformation response timesa.
As it turns out, the molecular structure of polymers allows for elastic deformation where
atoms change nearest neighbors. We explore this deformation mechanism in the following
sections.

Structure of Polymers
Polymers are made up of long chain molecules that consist of repeating segments called “mers”.
The basis of most polymer molecules is a “backbone” of carbon bonded with primary carbon-
carbon bonds as shown in Figure 7.1. A carbon atom has 4 electrons in its outer shell and can
form covalent bonds with 4 other atoms. The C-C bonds along the backbone are about 153 pm
long and bond-bond angles are about 112˚. Since these bonds are covalent, they are very stiff; i.e.
resistant to both changes in length and changes in angle.
The bonds along the backbone satisfy two of the available 4 bonds as shown in Fig. 7.1. The
other two bonds are satisfied by other species. Several examples of some important polymer
molecules are shown in Figure 7.2.

a
Note that the term “anelastic” is not well defined. A linguist would say that it should mean “not elastic.” But it is
not used in this way. In the broadest sense, it refers to elastic behavior for which there is no definite relationship
between stress and strain; that is, when there is not a single strain state associated with a given stress state as in
linear elasticity. In anelastic deformation, the strain state evolves with time at a given stress state. Thus, to specify
the strain, one must specify the stress-time history starting from some defined reference state. Here we use a
narrower definition by reserving “anelastic” for the time-dependent elastic deformation that arises from deformation
mechanisms in which at least some atoms change nearest neighbors. This encapsulates most common usages of this
term.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-2

Figure 7.1: Carbon chain configuration in polymers.

Cl H Cl H Cl H
C C C
H H H H H
H
C C C
C C C
C C H H
H H H H H H H H
(a) Poly(ethylene) (b) Poly(vinyl chloride), PVC

H H H H H
H
H H H H H C C C
H
C C C
C C
C C CH CH
H H 3 3 O
O

O O

CH CH
3 3
(c) Poly(styrene), PS (d) Poly(methyl methacrylate), PMMA,
or “Plexiglas”

H CH 3

H H C C
H H H

C C C C
C
H H H H H
C C
H CH 3

(e) cis-Poly(isoprene), natural rubber

O O O O
Si Si Si
CH
3
CH
3 CH C H 3 CH C H
3
3 3
(f) Poly(dimethylsiloxane), silicone rubber

Figure 7.2: A variety of different polymer molecule architectures.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-3

The main features of polymer molecule architecture that determine mechanical properties are:
1. Molecular length: This is the number of repeat units (the number of mers) along
each polymer chain.
2. The species that are attached at the carbon atoms along the chain.
a. Polymers that have primarily hydrogen or another relatively small
hydrocarbon moiety in those sites are known as linear polymers.
Polyethylene (Fig. 7.2a) is an example.
b. Side groups are “branches” that extend from the polymer chain; for example,
in the PMMA and PS structures shown in Fig’s 7.2c & d.
c. Extent of branching. Side groups can be polymer chains themselves, which
can “branch” by having long side groups of their own.
3. Degree of crystallinity: Although polymers can have complicated structures, the
structure is usually periodic along the chain. If several chains align in parallel a
crystalline region may be formed. This is shown in highly schematic form in Figure
7.3. Polyethylene (Fig. 7.2a), for example, can be made in a highly crystalline state.

(a) (b)
Figure 7.3: Schematic showing (a) amorphous and (b) crystalline arrangements in a polymer.b
4. Degree of cross-linking: When two carbon chains are connected at a point, or when a
side group or branch extends from one molecule to a different molecule, the junctions
are said to be cross links and the connected molecules are said to be cross linked. An
example of sulfur cross links in polyisoprene rubber is shown Figure 7.4.

Figure 7.4: Schematic illustrating cross-linking in polyisoprene rubber.c

b
Source Wikipedia.
c
The process of creating these links is called vulcanization. The vulcanization process, invented by Charles
Goodyear in 1839, made modern rubber products like tires, belts, and seals possible. Image source Wikipedia.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-4

In most cases, polymer molecules are not straight as shown in Fig. 7.2. Instead they are coiled up
in a complex interpenetrating network. As we will see, there are two main modes of elastic
deformation. In one, the molecules maintain their positions relative to each other (first nearest
neighbors remain the same) and deformation proceeds by bond stretching and distortion, just as
in metals and ceramics. However, since the bonds between molecules are predominantly weak
bonds, polymers are much more compliant than metals and ceramics. In the other mode,
molecules uncoil, becoming more and more aligned with the tensile axis. This uncoiling process
takes time takes time and is the origin of the anelastic behavior in polymers. Polymers that
deform by molecular uncoiling are even more compliant.
To first order, the effect of structure on mechanical properties can be thought of in terms of
how much the structural features prevent molecules from uncoiling, i.e. from moving around
relative to each other. In general, the longer the molecule, the bigger the side groups, and the
larger the extent of branching, crystallinity, and cross-linking, the less deformable the polymer.
The effect of temperature on polymers can be dramatic. As temperature increases, the
amplitude of thermal vibrations in the polymer molecules increase and the molecules are more
able to move relative to each other. Generally, polymers without cross-links are thermoplastic;
that is, at high temperatures the polymer melts when there is enough thermal energy that the
molecules can wiggle around each other enough to act as a liquid. Cross-linked polymers do not
melt when heated and are referred to as thermosets.
It is not necessary to have a carbon backbone to have a polymer. An example can be seen in
Fig. 7.2f, where the structure of polydimethylsiloxane is shown. This molecule is the basis of
silicone rubber. Silicon is just below carbon in the periodic table and also tends also to form 4
covalent bonds.
It is evident that the structure of polymer molecules can be varied over a very wide range. As
we will see, this results in an equally wide range of mechanical properties.

Single Molecule Mechanics


To understand the anelastic behavior of solid polymers, we must first understand the mechanical
behavior of a single molecule. Consider a single polyethylene molecule as shown in Fig. 7.2a.
The carbon atoms are arranged in a zig-zag chain with the hydrogen atoms as shown in Figure
7.5.

Figure 7.5: A polyethylene molecule showing the locations of the C and H atoms. 360˚
rotations about each C-C bond are possible.
In any equilibrium configuration (bonds not distorted), the distance between adjacent carbons is
153 pm and the angle between adjacent main-chain C-C bonds is 112˚ as shown in Fig. 7.1d. If
the molecule were laid out with all the carbon atoms in a plane, the distance between two
carbons on the same side of the zig-zag would be 254 pm. The diameter of a hydrogen atom is

d
All of the values regarding equilibrium atom spacings and bond angles are known to several decimal places.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-5

239 pm, so for this molecule, any arbitrary rotation about a main chain C-C bond is allowed (Fig.
7.5).
We can thus determine the possible configurations of the molecule as illustrated in Figure
7.6. If two C atoms, C1 and C2 are connected by bond 1 and atom C3 is connected to atom C2 by
bond 2, then atom C3 is 153 pm away from atom C2, the angle between bond 1 and bond 2 is
112˚ and atom C3 can take any position given by rotation ϕ 1 of bond 2 about the bond 1 axis.
Similarly if we add atom C4 to the chain, then for every possible position of atom C3, atom C4 is
153 pm away from atom C3, the angle between bond 3 and bond 2 is 112˚ and atom C4 can take
any position given by rotation ϕ 2 of bond 3 about the bond 2 axis. This is the freely jointed
chain model. If we know the orientation of bond 1 and all the rotation angles ϕi, then we can
determine the positions of all the atoms in the chain.

Figure 7.6: Rotational freedom in an idealized chain of four carbon atoms. ϕ1 gives the
angle of rotation of C3 about the C1–C2 axis, etc..
We see that the chain can adopt many different configurations. One way to characterize the
molecular conformation is by the end-to-end vector,

(7.1)
,
Where n is the number of links in the chain and li is the length of each link. Some possibilities
are shown in Figure 7.7. Note that here we are drawing the molecular chains as smooth curves
because the typical number of links is high (n ≈ 105 typical for polyethylene) and the size of each
link is small. At the sub-nanometer scale, these lines have the zig-zag form seen in Fig. 7.1.
We can find the mean square value ⟨r2⟩ by averaging over all possible ϕi. One can show that
⟨r2⟩ = nl2 so that the RMS end-to-end length is R = ⟨r2⟩½ = n½l. Thus, we see that the RMS
separation of the ends is much less than the chain length. This means that configurations like

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-6

those shown with r = x1 and r = x2 in Fig. 7.7are much more common than those with longer r,
especially those where r approaches L where there is (in 2-D) only one configuration.

Figure 7.7: Some molecular conformations as a function of r. The top figure shows two
possible configurations for the molecule at a smaller end-to-end spacing x1. The bottom
figure shows the one configuration available when r = L.
Let’s consider the change in energy as r goes from x1 to L. For a closed system, we can write
a change in internal energy dU as
dU = δQ – δW ,
where δQ is a small amount of heat added to the system and δW is a small amount of work
performed by the system. For a thermoelastic system at constant temperature we can write this as
DU = DY – TDS , (7.2)

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-7

where DU is the change in internal energy for an increase in strain energy DY, and an increase in
entropy DS. Here T is the absolute temperature.
We start with the molecule in some configuration and apply uniaxial force F to the two ends
to stretch it out. It extends by simply rotating C-C bonds about the bond axes (the freely jointed
chain model). The molecule “uncoils” and straightens out as shown in Fig. 7.7 as we do so. Note
that this process does not stretch or distort the bonds, so the change in strain energy is zero; that
is
𝜕𝛹
=0 . (7.3)
𝜕𝑟
Thus, the force needed to extend the molecule is then given by
𝜕𝑈 𝜕𝑆
𝐹= = −𝑇 . (7.4)
𝜕𝑟 𝜕𝑟
It is helpful to think of entropy as a measure of the number of configurations available to the
system. We see that we must apply force F to stretch the molecule because we are reducing the
number of configurations available to the molecule, thereby increasing its entropy. Equivalently,
when we release the externally applied force, F provides an (internal) restoring force that returns
the molecule to its highest entropy state (largest number of possible configurations). This is
entropic elasticity. It is possible to show using statistical mechanics that
𝜕𝑆 3𝑘𝑇
𝐹 = −𝑇 = / !4𝑟 , (7.5)
𝜕𝑟 𝑛𝑙
where k is Boltzmann’s constant. Thus, we see that entropic elasticity is another form of linear
elasticity!
It turns out that the stiffness associated with uncoiling is much lower than that associated
with bond distortions. Suppose we conduct a uniaxial test of a single molecule that starts out in a
coiled-up configuration like those shown with r = x1 in Fig. 7.7. We pull the molecule until it is
straight with r = L, and then keep increasing the load. The molecule will continue to stretch, but
now the bond lengths and angles must change. That is, the deformation mechanism changes from
molecular uncoiling to bond distortions. The result would be as shown in Figure 7.8.

Figure 7.8: A hypothetical uniaxial tension test of a single polymer molecule. As the
molecule is initially extended from its most probable configuration, it uncoils with a
stiffness given by 3kT/nl2. Once the molecule is fully straightened out, it can only deform

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-8

further by bond distortion and has much higher stiffness. If the load is increased without
limit, the molecule will eventually break.e

Deformation of Solid Polymers


We must now go from single molecule mechanics to deformation of solid polymers. We will do
this only very schematically as shown in Figure 7.9. We imagine a thermoplastic (not cross-
linked) solid polymer comprised of many individual molecules as shown in Figure 7.9a. If we
subject this sample to uniaxial tension and give the molecules enough time to move around, they
will extend by uncoiling in the tensile direction as shown in Figure 7.9b. When we unload, and
again give the molecules enough time to move around, they coil back up to maximize entropy
and the sample recovers elastically as shown in Figure 7.9c. However, the molecules are also
free to slide relative to each other, so when they coil back up, they will not be in the same
positions or configurations that they were initially. Thus, some plastic deformation will remain
as shown in Fig. 7.9c.

(a) (b) (c)


Figure 7.9: A solid polymer consisting of non-cross-linked molecules tested in uniaxial
tension. Deformation occurs by stretching of bonds (linearly elastic), molecular uncoiling
(anelastic), and by sliding of molecules past each other (plastic).
Of course, there will always be some bond stretching! This can easily be seen in the time
after the load has been applied but before the molecules have had enough time to uncoil. The
weak bonds between the molecules will stretch, and even the bonds along the C-C chains in the
molecules will be distorted very slightly, but these strains will be small compared to the strains
due to molecular uncoiling.
It is important to realize that by changing the details of the molecular architecture (molecular
length, side groups, branching, cross-linking, crystallinity, etc.) it is possible to change the time
scale of the deformation due to molecular uncoiling over several orders of magnitude.

Elastomers
We are interested in elastic deformation, and indeed many polymers are used for their unique
elastic properties in situations where plastic deformation is not desired. Elastomers are a class of
e
This experiment has been done! Molecules are prepared with a chemical end group that attaches to the surface of a
specially prepared substrate on one end and a different chemical end group that attaches to the end of a specially
prepared atomic force microscope (AFM) tip on the other. The AFM tip is displaced away from the surface while
the force that it detects is recorded. Force-displacement curves like that shown in Fig. 7.8 are found.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-9

materials in which the structure is controlled so that we can take advantage of the large strains
made possible by molecular uncoiling without unwanted plastic deformation. To achieve large
strains due to uncoiling, long-chain molecules in an amorphous arrangement are used. To
prohibit plastic deformation, the molecular chains are cross-linked.
Deformation of an elastomer can be imagined as deformation of an elastic net as shown in
Figure 7.10 When an elastomer is pulled in uniaxial tension, the molecules between the cross-
links uncoil until they are straight. Once they are straight, they can only deform by bond
distortion. There is some distribution of the lengths of the segments between cross-links, so some
chain segments become straight at lower strains and others at higher strains. As the fraction of
the chain segments that are straight increases, the stiffness of the material increases dramatically
as shown in Figure 7.11. This is similar to a superposition of many different single molecule tests
as shown in Figure 7.9 where the molecules have a range of lengths. The behavior shown in Fig.
7.11 is well known as this is the behavior of the common rubber band and many other rubber
products.

Figure 7.10: Schematic of an elastomer in the rubbery regime in the relaxed and loaded states.
Crosslinks are represented as dots at the intersections between molecular chains.

Figure 7.11: Stress-strain behavior of an elastomer. Stiffness increases dramatically as


more and more molecular segments between cross-links become straight.
We see that elastomers have highly non-linear elastic behavior at higher strains. However,
for the remainder of this Lecture we are going to consider elastomers subjected to small enough
strains that bond distortions of straightened-out chain segments can be ignored. That is, we
consider deformations only in the early linear part of the curve in Fig. 7.11.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-10

Effect of Temperature
In an elastomer at finite temperature, the atoms in the molecules vibrate about their mean
positions. Occasionally, these vibrations lead one or more segments of molecular chains to
“jump” to new mean positions. If two configurations are likely, segments of each polymer chain
may “jump” back and forth between those configurations. A schematic example is shown in
Figure 7.12.

Figure 7.12: Molecules change configurations by a series of “jumps” between


different configurations.
The probability of a jump occurring depends on the amplitude of the vibrations and since the
vibration amplitude increases with temperature, the frequency of such jumps, fjump, is a very
strong function of temperature as shown in Figure 7.13. A jump that occurs every 0.1 ns at a high
temperature may occur less than once every 3000 years at a low temperature!

Figure 7.13: The jump frequency fjump is a strong function of temperature.


Now suppose that we subject our randomly oriented polymer to a uniaxial tension test. Under
load, the atom-scale jumps will be biased in the direction that allows the material to stretch by
molecular uncoiling. Since uncoiling occurs by a series of jumps between configurations, the rate
of uncoiling varies dramatically with temperature.
Furthermore, we note that the material exhibits a typical glass transition. If we plot the
specific volume as a function of temperature as in Figure 7.14, we see that it varies linearly at
low temperatures with a low slope where bond expansion is the mechanism (same as thermal
expansion in metals and ceramics). At higher temperatures, it varies linearly with a higher slope
because jump motion leads directly to lower density configurations, dramatically increasing the
rate of change of specific volume with temperature. As is common in glass-forming materials,
the intersection of the low- and high-temperature slopes is used to define the glass transition

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-11

temperature Tg. We observe that Tg is also the inflection point in the plot of 1/fjump vs.
temperature shown in Fig. 7.13.

Figure 7.14: The amorphous polymer that we are considering shows a glass transition
indicated by a change in the volume expansion coefficient.
We immediately see that the elastic behavior depends on the temperature! Suppose that we
conduct uniaxial tension tests in which we increase the stress at a constant rate to some
maximum value over a few seconds, and then immediately unload at the same rate. We then
define the Young modulus as E = σmax/e|smax, where σmax is the maximum applied stress and
e|smax is the elastic strain measured at σmax. The form of the variation of E with T will appear as
shown in Figure 7.15. We see that this curve is similar in form to the jump frequency curve in
Fig. 7.13 with three different regimes of behavior with temperature. The inflection point
corresponds with Tg as found in Fig’s 7.13 and 7.14. Young’s modulus varies by several orders
of magnitude over this range. (Don’t forget that we are limiting the maximum stress such that the
deformation never exceeds the range where it could be accommodated by uncoiling!)

Figure 7.15: Young’s modulus of an elastomer as a function of temperature.


Schematic stress-strain curves from each of these regions are shown in Figure 7.16. At low
temperatures, the configuration of the molecules will not change in the time of the test (following
Fig. 7.13, only 1 in 1010 possible configuration changes will have occurred during this time).
Thus, the only (elastic) deformation mechanism available is stretching and distortion of bondsf.

f
Both strong and weak bonds will stretch but since these may be thought of as springs in series, the vast majority of
deformation will be due to stretching of weak bonds.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-12

We will have linear elastic behavior with a high modulus as shown in Fig.7.16a. This is the
glassy regime. Young’s modulus decreases slightly with increasing T in this regime due to
thermal expansion and reduction of the curvature of the potential well between atoms, just as in
metals and ceramics.
At high (but not too high!) temperatures, the configuration of the molecules will change very
rapidly (again following Fig. 7.13, each possible configuration change will have occurred 1010
times during this time). Thus, molecular uncoiling and recoiling will occur very rapidly
compared to the time of the test and will provide virtually all of the deformationg. We will again
have linear elastic behavior (see Eq. 7.5), but this time with a low modulus as shown in Fig
7.16b. This is the rubbery regime. Young’s modulus increases linearly with T in this regime
since, for a given change in entropy (which is equivalent to a given change in elastic strain) the
contribution to the energy of the system scales with temperature as shown in Eq’s 7.4 and 7.5.
That is, the force needed to create the configuration change (entropy S) associated with a given
amount of strain (the extension r in Eq’s 7.4 and 7.5 is related to strain) increases linearly with T.
A rubber band stretched under constant load becomes stiffer and shrinks when it is heated.

(a) (b) (c)


Figure 7.16: Schematic stress-strain curves from the (a) glassy, (b) rubbery, and (c)
viscoelastic regimes.
In between the glassy and rubbery regimes is the so-called viscoelastic regimeh, where
molecular uncoiling is occurring on the time scale of the test. Strain accumulates during loading
at any given stress as time-dependent molecular uncoiling proceeds; and upon unloading, strain
recovers at any given stress as time-dependent molecular re-coiling proceeds. As a result, there is
a hysteresis in the stress-strain behavior during a load cycle as shown in Fig. 7.14c. As we have
seen, the area inside this loop represents energy that is lost to heat. This leads to damping, which
we will explore shortly.
If we heat beyond the rubbery regime, eventually the sample will break down. We will not
study this part of the curve.
We can summarize these results as follows:
• For T << Tg, 1/fjump > t (t = time of test) — Glassy regime
– Only bond stretching possible
– Linear “time independent” behavior
– Modulus high (Fig. 7.16a)
– E decreases slightly with increasing T due to thermal expansion as in metals and
ceramics.

g
Bond stretching of course also occurs, but the vast majority of deformation will be due to molecular uncoiling.
h
We will learn why it has this name soon!

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-13

• For T >> Tg, 1/fjump < t — Rubbery regime


– Deformation mostly uncoiling
– Linear “time independent” behavior
– Modulus low (Fig. 7.16b)
– E increases with T since a given strain corresponds to a given change in S and the
contribution of S to the energy of the system scales linearly with T (Eq’s 7.4 and
7.5).
• For T ≈ Tg, 1/fjump ≈ t — Viscoelastic regime
– Strains accommodated by bond stretching and molecular uncoiling
– Non-linear behavior with effective modulus in intermediate range (Fig. 7.16c)
– Energy absorbed during load-unload process ⇒ damping.
– E falls rapidly with T as molecular uncoiling provides an increasing fraction of the
strain in the time of the test.

Effect of Deformation Rate


In Fig. 7.15, we saw the effect of varying the time scale of the materials response (by varying the
temperature) relative to a fixed test time. We could also fix the time scale of the materials
response by fixing the temperature and vary the time scale of the test. The result is shown
schematically in Figure 7.17. If the test times are sufficiently short (fast tests), then the test will
be over before molecular coiling can occur and the materials response will be glassy. If the test
time is sufficiently long, then molecular uncoiling will be fast relative to the time of the test and
the materials response will be rubbery. As before, when the time scale of the test and the time
scale of the materials response are similar, viscoelastic behavior will be seen.

Figure 7.17: Young’s modulus of a random polymer as a function of the time of the test.

Note that Young’s modulus is constant in both the glassy and rubbery regimes in Fig 7.17
because the temperature is constant.
It is evident that we could change Tg in Fig. 7.15 by testing at a different rate. Similarly, we
could change the location of the transition region in Figure 7.17 by testing at a different
temperature. Thus, this kind of behavior can be equivalently described using variations in time or
temperaturei. This is a very good application for the Deborah number!

i
In fact, there are methods for converting from one to the other. This is known as time-temperature equivalence.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-14

Time-Dependent Elastic Deformation Near Tg


We want to quantify the response of in the viscoelastic regime. Imagine that we impose a
uniaxial stress pulse on an elastomeric material at a temperature near Tg. We impose a stress so
at time t = 0 and hold it constant until a later time tu, at which point we unload. The stress is
imposed and removed quickly—meaning fast relative to the uncoiling process but not so fast that
we have to account for elastic wave speeds. The resulting strain-time behavior is shown in Figure
7.18.

so
stress

(a) 0 tu time

e∞
bond (un)stretching
strain

molecular uncoiling
molecular re-coiling
eo bond stretching
0 tu time
(b)

Figure 7.18: Imposed stress pulse (a) and strain response (b) in a viscoelastic polymer.

Initially, we see the linear elastic response of bond stretching to a strain eo. Molecular uncoiling
takes time and at t ≈ 0, there hasn’t been enough time for molecular uncoiling to occur. But
molecular uncoiling does occur starting at some rate and slowing down as the strain increases,
asymptotically approaching a final value, e∞. The initial rate is determined by the average
uncoiling time and the rate slows down as molecules of different lengths become straight and
stop uncoiling. (Remember that we are not applying enough stress to significantly deform these
straightened molecules!) When we unload the sample, the strain due to stretching, eo, recovers
“instantaneously” and the molecules start to coil back up. The initial rate is just the negative of
the initial rate on loading. Again, the rate declines as different molecules reach their equilibrium
configurations. So, on loading we have linear elasticity from 0 to eo and anelastic behavior from
eo to e∞. On unloading we have linear elasticity from e∞ to e∞ – eo and anelastic recovery from e∞
– eo to 0.
It turns out that the anelastic part of the loading curve can be described using an equation of
the form
−𝑡
𝜀" = (𝜀# − 𝜀$ ) /1 − exp > A4 , (7.6)
𝜏%
where ts is the time constant for constant stress. (Note that e∞ – eo = ea∞ is the final strain for the
anelastic part alone.) We can then write an expression for e(t) during the entire loading as

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-15

−𝑡
𝜀 = 𝜀$ + (𝜀# − 𝜀$ ) /1 − exp > A4 . (7.7)
𝜏%
Similarly, the anelastic part of the unloading curve can be described by
𝑡 − 𝑡&
𝜀" = (𝜀# − 𝜀$ )exp >− A , (7.8)
𝜏%
which represents the entire unloading.
The time constant ts in these expressions represents the time at which the anelastic strain
change is (1-1/e) ≈ 62.3 % complete (loading or unloading). Thus, ts is our indicator for the time
scale of the deformation. If the time of the test is small relative to ts, we will have glassy
behavior, and if the time of the test is large relative to ts we will have rubbery behavior. If the
time of the test is similar to ts we will have viscoelastic behavior.
It is evident that if we plot the stress-strain behavior from the experiment shown in Fig. 7.18,
we will get a result like that shown in Figure 7.19.

so
stress

0 eo strain
e∞

Figure 7.19: Stress-strain behavior for the experiment shown in Figure 7.18.

The shaded area in the box represents energy that was put into the system and lost as heat during
this experiment. As in Fig. 7.16c, this leads to damping as we will soon see.

Elastic Properties of Elastomers (and other Polymers)


Now that we understand the deformation mechanisms and structure, we can understand the
general features of the elastic properties of elastomers compared with metals and ceramics. We
also get some insight into the elastic properties of other polymers.

Polymer elastic behavior is time and rate sensitive


This of course has been the theme of this lecture so far, but it must be emphasized. The Young
modulus of an elastomer (defined as E = σmax/e|smax as in Fig. 7.16) varies by orders of
magnitude as the temperature varies from below Tg to above Tg. This turns out to be very useful
in applications where compliance is desired but total deformation should be limited (seals, tires,
tie-downs, etc…). In any case, it must be accurately accounted for in product design and use.j By
contrast, the change in Young’s modulus of typical metals and ceramics is of order 10% from
low temperatures to temperatures near the melting point and occurs due to thermal expansion.

j
For example, the in-flight explosion of the space shuttle Challenger in 1986 was due to a failure to properly account
for the time-dependent behavior of the rubber O-ring seals in the external solid-fuel rocket boosters.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-16

Polymer elastic behavior is very sensitive to structure


The molecular atomic arrangements in polymers lead to striking variations in elastic behavior
with temperature. For mobility, there must be easy rotation about the “backbone” bonds along
the chain molecules. In polymers such as PS and PMMA, this bond rotation is hindered by the
bulky side groups (phenyl and methacrylate, respectively, see Fig. 7.2c and d). Thus, at common
testing rates, these polymers exhibit bond stretching to relatively high temperatures and only
show the low moduli associated with molecular uncoiling at temperature above ~100°C.
Similarly, Polyethylene (Fig. 7.2a) is not a good elastomer at room temperature because it is
highly crystalline. A polymer may still exhibit elastomeric properties if only a small percentage
of the material is crystalline (a small degree of crystallinity).
Crosslinks also play an important role. Without crosslinks, the molecules would slide past
each other and much of the deformation would be plastic. These crosslinks normally are
introduced chemically, as in the sulfur crosslinks between polymer chains in polyisoprene shown
in Fig7.4. Small isolated crystallites can serve as crosslinks (this kind of crosslinking is thought
to occur in poly(vinylchloride)). The number of crosslinks per unit volume Nc is the important
parameter characterizing the crosslinking. Young’s modulus of an elastomer is directly
proportional to Nc.
Even if no crosslinks are deliberately introduced, entanglements between neighboring chains
can act as effective temporary crosslinks. Certain uncrosslinked polymers such as PMMA and PS
show a modulus that varies with temperature in a way very similar to that of elastomers, with
glassy, transition, and rubbery regimes and a well defined glass transition. Even though these
molecules are not crosslinked, the side groups cause the molecules to physically entangle with
one another so that, for a certain temperature range above Tg (and a certain deformation rate), the
material responds in a rubbery manner (creating a “rubbery plateau”), at least at short times. At a
high enough temperature (well above Tg), the entanglements are no longer stable and the polymer
flows like a viscous liquid (it melts—recall that this is a characteristic of thermoplastics).
Molding operations, such as extrusion or injection molding, can be carried out above this
temperature. Even melts of uncrosslinked polymers with long enough chains show elastomeric
behavior for very fast load/unload cycles.
It is convenient, and common, to categorize elastomers by Tg. Elastomers with Tg > Troom will
be glassy and stiff at room temperature while elastomers with Tg < Troom will be rubbery and
compliant. The Tg of a polymer can be changed by changing the molecular structure of the
polymer or by changing the rate at which the polymer is deformed.
By contrast, the Young moduli in metals and ceramics are relatively insensitive to changes in
structure—except for texture! For a given texture, variations in defect density and even
composition play a relatively minor role. Young’s moduli in metals and ceramics are said to be
structure insensitivek.

Polymers are compliant


Even in their glassy state, polymers are almost always very compliant compared to metals and
ceramics due to the presence of weak (secondary) bonds and large intermolecular spacings.
Molecular uncoiling makes them even more compliant. A table comparing Young’s moduli for a
wide range of materials is shown in Figure 7.19.

k
As long as the phase does not change, E will vary “only” approximately linearly with composition between end
points.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-17

The Young moduli of common polymers vary from about 1 MPa to about 10 GPa (4 orders
of magnitude), while the Young moduli of metals/ceramics vary from about 10 GPa to about
1000 GPa (2 orders of magnitude). Note, however, that the actual range of polymer Young’s
moduli is much wider than that shown in Figure 7.20. For example, poly-paraphenylene
terephthalamide (branded “Kevlar”) fibers can have E > 100 GPa in the fiber axis direction. On
the other end, there is almost no lower limit to the Young moduli of polymer hydrogels.
For metals and ceramics, even those with relatively weak bonds, the Young modulus, E
usually exceeds 10 GPa and may be as high as 1000 GPa. In steel, for example, E is about 207
GPa. Elastomers, on the other hand, have much lower Young’s moduli. For example, rubber
typically has a modulus in the range of 0.1 to 100 MPa.

Figure 7.20: Comparison of Young’s moduli E for several polymers, compared to


ceramics, metals, and composites.l

Polymers are often anisotropic


Unsurprisingly given the anisotropy of molecules, polymers can also be extremely anisotropic.
Consider for example a polymer with random molecular chain orientations vs. one with oriented
molecular chains as might be found in polymer fibers formed by drawing or stretching the fibers
(Figure 7.21):
l
From Ashby MF and Jones DRH. Engineering Materials 1, 2nd Ed. Butterworth-Heinemann 1996.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker


7-18

Figure 7.21: (a) Random chain orientations (b) oriented molecular chains.

Similar to drawn polycrystalline metal wires having fiber texture or oriented composites, these
oriented polymer fibers have hexagonal elastic symmetry (Figure 5.7). Elastic compliance
constants for several common oriented polymer fibers are shown below.

From Ward IM and Hadley DW. An Introduction to the Mechanical Properties of Polymers.
Wiley 1993.
Other elastic symmetries are of course possible.

Polymer elastic deformation can be highly non-linear


While the elastic response of metals and ceramics is very nearly linear, at least up to the point
where plasticity or fracture become important, elastomers can show elastic stress-strain behavior
that is highly non-linear (7.11).

Polymer elastic strains can be very large


While metals and ceramics exhibit elastic behavior only to rather small strains, with typical
maximum elastic strains being 10-4 to 10-2 (0.01 to 1%). Rubber, in comparison, may be
stretched elastically to strains of 300 to 1000%. As an aside, because deformations in polymers
can be large, the extension ratio, l = 1 + e is sometimes used to describe the displacements.
There is a complete mathematics of large deformations that is appropriate here. However, we
will have to save that for a more advanced/dedicated course.

Cornell MS&E 5802 Lecture 7 Ó2024 Shefford P. Baker

You might also like