2.1 Structure and Property Relationship in Organic Polymers
For a very large proportion of polymeric
materials in commercial use, mechanical properties are of paramount
importance, because they are used as structural materials, fibers,
or coatings and these properties determine their usefulness.
Properties that also determine their utilization are compressive,
tensile, and flexural strength, and impact resistance. Hardness,
tear, and abrasion resistance are also of concern. In addition,
polymers may be shaped by extrusion in molten state into molds or
by deposition from solutions on various surfaces. This makes the
flow behaviors in the molten state or in solution, the melting
temperatures, the amount of crystallization, as well as solubility
parameters important.
The physical properties of polymer molecules are
influenced not only by their composition, but also by their size
and by the nature of their primary and secondary bond forces. They
are also affected by the amount of symmetry, by the uniformity in
their molecular structures, and by the arrangements of the
macromolecules into amorphous or crystalline domains. This, in turn
affects melting or softening temperatures, solubilities, melt and
solution viscosities, and other physical properties [1].
Due to the large sizes of polymeric molecules,
the secondary bond forces assume much greater roles in influencing
physical properties than they do in small organic molecules. These
secondary bond forces are van der Waal forces and hydrogen bonding.
The van der Waal forces can be subdivided into three types:
dipole–dipole interactions, induced dipoles, and time varying
dipoles.
2.1.1 Effects of Dipole Interactions
Dipole interactions result from molecules
carrying equal and opposite electrical charges. The amounts of
these interactions depend upon the abilities of the dipoles to
align with one another. Molecular orientations are subject to
thermal agitation that tends to interfere with electrical fields.
As a result, dipole forces are strongly temperature dependent. An
example of dipole interaction is an illustration of two segments of
the molecular chains of a linear polyester. Each carbonyl group in
the ester linkages sets up a weak field through polarization. The
field, though weak, interacts with another field of the same type
on another chain. This results in the formation of forces of
cohesion. Because polymeric molecules are large, there are many
such fields in polyesters. While each field is weak, the net effect
is strong cohesion between chains. The interactions are illustrated
in Fig. 2.1.
Fig.
2.1
Intramolecular forces
2.1.2 Induction Forces in Polymers
Electrostatic forces also result from slight
displacement of electrons and nuclei in covalent molecules from
proximity to electrostatic fields associated with the dipoles from
other molecules. These are induced dipoles. The displacements
cause interactions between the induced dipoles and the
permanent dipoles
creating forces of attraction. The energy of the induction forces,
however, is small and not temperature-dependent.
There are additional attraction forces that
result from different instantaneous configurations of the electrons
and nuclei about the bonds of the polymeric chains. These are time
varying dipoles that average out to zero. They are polarizations
arising from molecular motions. The total bond energy of all the
secondary bond forces combined, including hydrogen bonding, ranges
between 2 and 10 kcal/mole. Of these, however, hydrogen
bonding takes up the greatest share of the bond strength. In
Table 2.1
are listed the intramolecular forces of some linear polymers
[2, 3].
Table
2.1
Molecular cohesion of some linear
polymersa
Polymer
|
Cohesion/5 Å chain length
|
Polymer structure
|
---|---|---|
Polyethylene
|
1.0
|
|
Polybutadiene
|
1.1
|
|
Polyisobutylene
|
1.2
|
|
Polyisoprene (cis)
|
1.3
|
|
Poly(vinyl chloride)
|
2.6
|
|
Poly(vinyl acetate)
|
3.2
|
|
Polystyrene
|
4.0
|
|
Poly(vinyl alcohol)
|
4.2
|
|
Polyamides
|
5.8
|
|
Cellulose
|
6.2
|
As can be seen in Table 2.1., polyethylene
possesses much less cohesive energy than does a polyamide. This
difference is primarily due to hydrogen bonding. A good
illustration is a comparison of molecules of a polyamide, like
nylon 11, with linear polyethylene. Both have similar chemical
structures, but the difference is that nylon 11 has in its
structure periodic amide linkages after every tenth carbon, while
such linkages are absent in polyethylene. The amide linkages
participate in hydrogen bonding with neighboring chains. This is
illustrated in Fig. 2.2.
Fig.
2.2
Hydrogen bonding in (a) nylon 11, and absent in (b) polyethylene
Due to this hydrogen bonding, nylon 11 melts at
184–187°C and is soluble only in very strong solvents. Linear
polyethylene, on the other hand, melts at 130–134°C and is soluble
in hot aromatic solvents.
The energy of dipole interactions,
(?k) can be calculated from the equation [1]:
where, μ represents the
dipole moment of the polarized section of the molecule,
r is the distance between
the dipoles, T is the
temperature in Kelvin, and R is the ideal gas constant.
Intermolecular forces affect the rigidity of all
polymers. Should these forces be weak, because the cohesive energy
is low (1–2 kcal/mole), the polymeric chains tend to be
flexible. Such chains respond readily to applied stresses and can
exhibit typical properties of elastomers. High cohesive energy, on
the other hand, (5 kcal/mole or higher) causes the materials
to be strong and tough. These polymers exhibit resistance to
applied stresses and usually possess good mechanical properties.
The temperatures and the flexibility of polymeric molecules govern
both the sizes of molecular segments, the motion, and the frequency
at which that occurs. This in turn determines the rate at which the
polymer molecules respond to molecular stresses. In flexible
polymers, if the thermal energy is sufficiently high, large
segments can disengage and slip past each other quite readily in
response to applied stress. All elastomers possess such properties.
Properties of polymers depend also upon morpholog y or upon the arrangement of their
polymeric chains. This arrangement can be amorphous or crystalline.
The term amorphous
designates a lack of orderly arrangement. Crystalline morphology,
on the other hand, means that the chains are aligned in some
orderly fashion. Generally, the freedom of molecular motion along
the backbones of polymeric chains contributes to lowering the
melting temperature. Substituents that interfere with this motion
tend to raise the melting point. For instance, isotactic
polypropylene melts at a higher temperature than does linear
polyethylene. If the substituent is bulky or rigid it raises the
melting point because it interferes with molecular motion. Dipole
interactions, as discussed above, have a similar effect
[4]. A good illustration is a
comparison of poly(ethylene terephthalate) that melts at 265°C with
poly(ethylene adipate) that melts at only 50°C. In the first
polyester, there is a rigid benzene ring between the ester groups
while in the second one there is only a flexible chain of four
carbons.
This flexibility of the four carbon segment in
poly(ethylene adipate) contributes significantly to the lowering of
the melting point.
Linear polymers that possess only single bonds
between atoms in their backbones, C–C, or C–O, or C–N, can undergo
rapid conformational changes [5].
Also ether, imine, or cis-double bonds reduce energy barriers
and, as a result, “soften” the chains, causing the polymer to
become more rubbery and more soluble in various solvents.
The opposite is true of cyclic structures in the
backbones, as was shown in poly(ethylene terephthalate). Actually,
cyclic structures not only inhibit conformational changes but can
also make crystallization more difficult. Among the polymers of
α-olefins the structures of the pendant groups can influence the
melting point [6]. All linear
polyethylene melts between 132 and 136°C [7]. Isotactic polypropylene, on the other hand
melts at 168°C [8].
As the length of the side chain increase,
however, melting points decrease and are accompanied by increases
in flexibility [7] until the length
of the side chains reached six carbons. At that point, the minimum
takes an upturn and there is an increase in the melting points and
decrease in flexibility. This phenomenon is believed to be due to
crystallization of the side-chains [9].
Alkyl substituents on the polymers of α-olefins
that are on the α-carbon yield polymers with the highest melting
points. Isomers substituted on the β-carbon, however, if
symmetrical, yield polymers with lower melting points.
Unsymmetrical substitutions on the β-carbon, on the other hand,
tend to lower the melting points further. Additional drop in the
melting points result from substitutions on the γ-carbon or further
out on the side-chains. Terminal branching yields rubbery polymers
[10].
Copolymers melt at lower temperatures than do
homopolymers of the individual monomers. By increasing the amount
of a comonomer the melting point decreases down to a minimum (this
could perhaps be compared to a eutectic) and then rises
again.
The tightest internal arrangement of
macromolecules is achieved by crystallinity. As a result, the
density of a polymer is directly proportional to the degree of
crystallinity, which leads to high tensile strength, and to stiff
and hard materials that are poorly soluble in common solvents
[11]. The solubility of any
polymer, however, is not a function of crystallinity alone, but
also of the internal structure and of the molecular weight. The
solubility generally decreases with increases in the molecular
weight. The fact that crystalline polymers are less soluble than
amorphous ones can be attributed to the binding forces of the
crystals. These binding forces must be overcome to achieve
dissolution. Once in solution, however, crystalline polymers do not
exhibit different properties from the amorphous ones. One should
also keep in mind the fact that crosslinked polymers will not melt
and will not dissolve in any solvent. This is due to the fact that
the crosslinks prevent the chains from separating and slipping past
each other.
2.2 The Amorphous State
A number of macromolecules show little tendency
to crystallize or align the chains in some form of an order and
remain disordered in solid form. This, of course, is also the
condition of all molten polymers. Some of them, however, due to
structural arrangement, remain completely amorphous upon cooling.
The crystalline polymers, which crystallized from the melt, on the
other hand, while containing areas of crystallinity, also contain
some amorphous material. All crystalline polymers are also
amorphous above their melting temperature. When sufficiently
cooled, amorphous polymers can resemble glass. Above this glassy
state, long-range segmented motions are possible and the molecular
chains are free to move past each other. On the other hand, in the
rigid, glass like state, only short-range vibrational and
rotational motions of the segments are possible. At temperatures
above the glassy state, amorphous polymers resemble rubbers, if
crosslinked. If not crosslinked, amorphous polymers resemble very
viscous liquids in their properties and there is molecular
disorder. Tobolsky suggested [12]
that polymer molecules in an amorphous state might be compared to a
bowl of very long strands of cooked spaghetti. When molten, such
molecules are in a state of wriggling motion, though the amplitude
and speed depend very much on the temperature [13]. The important thing is to know that the
chains possess random conformations. These conformations were
characterized with the aid of statistical analysis [14].
2.2.1 The Glass Transition and the Glassy State
When the polymer cools and the temperature
lowers, the mobility in the amorphous regions of the polymer
decreases. The lower the temperature, the stiffer the polymer
becomes until a point of transition is reached. This transition is
called glass transition
or second-order
transition [15]. The
temperature at that transition is called the glass transition temperature, designated by
T g. Beyond
stiffness, a change is manifested in specific volume, heat content,
thermal conductivity, refractive index, and dielectric loss.
Bueche illustrated glass transition as follows
[16]. In measuring the force
necessary to force a needle into a polymer, like polystyrene, at
various temperatures, there is a relationship between the force
required to insert the needle and the temperature [16]. As the temperature is being lowered,
maximum resistance to penetration is reached at T g.
As stated, above T g chains undergo
cooperative localized motion. It is actually estimated that above
T g segmental
motion of anywhere between 20 and 50 chain atoms is possible. Below
the second-order transition temperature, however, there is
insufficient energy available to enable whole segments of the
polymeric chains to move. The structure is now stiff and brittle
and resists deformation. When, however, sufficient amount of heat
energy enters the material again and the temperature rises above
T g larger
molecule motion involving coordinated movement returns. This
requires more space, so the specific volume also increases and the
polymer is in a rubbery or a plastic state. Above T g, because large elastic
deformations are possible, the polymer is actually both tougher and
more pliable. Chemical structures of the polymers are the most
important factors that affect the glass transition temperatures.
Molecular weights also influence T g, as it increases with
the molecular weight. In addition, T g also varies with the
rate of cooling. Table 2.2. shows the structures of and lists relative
T g values of
some common polymers.
Table
2.2
Glass transition temperatures, T g, of some common
polymersa
Repeat unit
|
Polymer
|
T
g (°C)
|
---|---|---|
Polystyrene
|
81; 90
|
|
Poly(vinyl chloride)
|
75
|
|
Poly(methyl methacrylate)
|
57–68
|
|
Cellulose nitrate
|
53
|
|
Nylon 6,6
|
47
|
|
Poly(vinyl acetate)
|
30
|
|
Poly(butyl methacrylate)
|
22
|
|
Poly(ethyl acrylate)
|
−22
|
|
Poly(vinyl fluoride)
|
−39
|
|
Poly(butyl acrylate)
|
−56
|
|
Polyisoprene
|
−70
|
|
Polyethylene
|
−80; −90
|
|
Silicone rubber
|
−123
|
One way to obtain T g is by studying thermal
expansion of polymers. It is generally observed that the thermal
coefficient of expansion is greater above the glass transition
temperature than below it, though the magnitude of the change
differs from one polymer to another. By plotting volume vs.
temperature for a polymer, one can obtain T g as shown in
Fig. 2.3,
which illustrates obtaining T g from specific
volume-cooling temperature curves [17].
Fig.
2.3
The relationship of the specific volume of
a polymer to the temperature
Polymers with bulky, tightly held side groups or
stiff bulky components in the backbone have high T g values. This is due to
the fact that such side groups or bulky components interfere with
segmental motion. Such polymers require higher temperatures to
acquire sufficient free volume for segmental motion. This can be
observed in Table 2.2. which shows that the glass temperature of
polystyrene with stiff benzene ring side groups is much higher than
that of polyethylene.
Also polymers with high attractive forces between
chains will require more heat energy to go from a glassy to a
rubbery or a plastic state. On the other hand, polymeric chains
with loose hanging side chains that tend to loosen the polymer
structure and increase the free volume for segmental movement will
have lower T g.
For instance, the glass transition temperature of poly(methyl
methacrylate) is higher than that of poly(n-butyl methacrylate) as can be seen
from Table 2.2.
The transition to the glassy state from an
equilibrium liquid results in changes in enthalpy, H, and volume, V. The specific heat is related to the
enthalpy by definition:
The glassy state is nonequilibrium in nature and
exhibits a tendency to undergo structural relaxation toward
equilibrium. This tendency of the glassy state to relax
structurally toward equilibrium is often referred to as
structural recovery. It
was observed, however, that the progress towards structural
recovery with time varies significantly between a down-quench and
an up-quench. This is referred to as asymmetry of structural recovery. The nonlinearity of
the process is described by the following equation [16]:
where, τ is the relaxation
time and υ is the specific
volume. The value of T
g is also affected by other parameters, such as
molecular weight, diluent concentration, tacticity, external
pressure, crosslinking, chain stiffness, and mechanical
deformation. For linear polymer, the Fox–Flory equation [18]described the relationship between
T g and the
molecular weight:
This equation, however, is limited in scope. It
is actually neither valid for very high molecular weight polymers
nor is it applicable to low molecular weight polymers.
The transition in a polymer from a molten state
to a glassy one actually occurs over a temperature range. This
range also includes T
g. At the glass transition temperature, however, the
change in viscosity is rapid, from very viscous to a glassy one.
Polymeric materials that undergo rearrangements in response to
outside stimulus, like light, are becoming increasingly important
in various industrial application (see Chap.
10). Urban and coworkers [19] studied stimuli-responsive (T SR) transitions and
correlated them to the glass transition temperatures (T g). Based on their
empirical data obtained from a copolymer, they concluded that the
relationship between T
g and T
SR is
where the V 1
and V 2 are the
copolymer’s total volumes below and above the T SR, respectively,
T g is the glass
transition temperature of the copolymer, and P 1 and P 2 are the fraction of the
free volume (f
free) at T
g (P
1) and (T
g, midpoint − T
SR)50/50 for each random copolymer
(P 2),
respectively. They feel that this relationship can be utilized to
predict the total volume changes as a function of T SR − T g for different copolymer
compositions.
2.2.2 Elasticity
The phenomenon of elasticity of rubber and other
elastomers is a result of a tendency of large and very flexible
amorphous polymeric chains to form random, thermodynamically
favorable, conformations [18]. If
a certain amount of crosslinking is also present, then these random
conformations occur between the crosslinks. In a vulcanized
(crosslinked) rubber elastomer, the crosslinks may occur at every
five hundred to a thousand carbon atoms. The distance between the
ends in such polymers is much shorter than when these elastomers
are fully stretched. Deformation or stretching of rubber
straightens out the various conformations in the molecules. They
tend to return to the original state, however, when the forces of
deformation are removed. So each segment behaves in a manner that
resembles a spring. Some elastomeric materials are capable of high
elongation and yet still capable of returning to the original
conformation. Some soft rubbers, for instance, can be extended as
high as 800% and even higher with full recovery. There is a
preference for trans
conformation, a planar zigzag at high extension. Rigidity of the
chains, however, or crystallinity would hinder extension and,
particularly, the recovery. High viscosity and a glassy state would
do the same.
The high degree of elasticity of rubbers is due
in part to the effects of thermal motions upon the long polymeric
chains. These motions tend to restrict vibrational and rotational
motions about the single bonds in the main chain. Such restrictive
forces in the lateral direction, however, are much weaker than are
the primary valence forces in the longitudinal direction. Greater
amplitudes of motion also take place perpendicular to the chains
rather than in the direction of the chains. These increased motions
in the perpendicular direction result in repulsive forces between
extended or parallel chains. Such forces cause them to draw
together after stretching. So, the stretched rubber molecules
retract due to longitudinal tension until the irregular arrangement
of molecules is achieved again. This more random conformation is
actually a higher entropy state.
When unstretched rubber is heated it increases in
dimension with an increase in the temperature, as one might expect.
At higher temperatures, however, rubbers, upon elongation, have a
higher tendency to contract. This can be written as follows:
In summary, polymeric materials exhibit rubber
elasticity if they satisfy three requirements: (a) the polymer must
be composed of long-chain molecules, (b) the secondary bond forces
between molecules must be weak, and (c) there must be some
occasional interlocking of the molecules along the chain lengths to
form three-dimensional networks. Should the interlocking
arrangements be absent, then elastomers lack memory, or have a
fading memory and are not capable of completely reversible elastic
deformations.
2.2.2.1 Thermodynamics of Elasticity
Stretching an elastomer reduces its entropy and
changes its free energy. The retractive force in an elastomer is
primarily the result of its tendency to increase the entropy
towards the maximum value it had in the original deformed state
Current explanations of rubber-like elasticity are based on several
assumption [20]. The first one is
that rubber-like elasticity is entirely intramolecular in origin.
The second one is that the free energy of the network is separable
into two parts, an elastic one and a liquid one. The liquid one is
presumably not dependent on deformation. When an elastomer is
stretched, the free energy is changed, because it is subjected to
work. If we consider the stretching in one direction only, the work
done W el is
equal to f∂l, where f is the retractive force and
∂l is the change in length.
The retractive force is then [19–21].
where F is the free energy,
H is the enthalpy, and
S is the entropy of the
system. An ideal elastomer can be defined as one where
(∂H/∂L)T,p = 0 and f = −T(∂S/∂L)T,p. The negative sign is
due to the fact that work has to be done to stretch and increase
the length of the elastomer. This description of an ideal elastomer
is based, therefore, on the understanding that its retractive force
is due to a decrease in entropy upon extension. In other words, the
entropy of elasticity is the distortion of the
polymer chains from their most probable random conformations in the
unstretched condition. The probability that one chain end in a unit
volume of space coordinates, x, y, z is at a distance r from the other end is [21, 22]:
where b
2 = 3/2xL
2. The number of links is x and the length is L. The entropy of the system is
proportional to the logarithm of the number of configurations.
Billmeyer expresses it as follows [7]:
where k is the Boltzmann’s
constant. The retractive force for a single polymer chain,
f′ stretched to a length
dr at a temperature
T is, therefore,
It is generally assumed that the total retractive
force of a given sample of an elastomer is the sum of all the
f′ forces for all the
polymeric chains that it consists of. This is claimed to be
justified in most cases, though inaccurate in detail
[22].
Tobolsky wrote the equation for the entropy
change of an unstretched to a stretched elastomer as depending upon
the number of configurations in the two states [12]:
where Ω and Ωu represent the number of configurations.
The evaluation of these configurations by numerous methods allows
one to write the equations for the change in entropy as:
where N 0 and
L and L u are the relative lengths
of the unstretched and stretched elastomer. Tobolsky derived the
tensile strength as being [12]:
By dividing both sides of the equation by the
cross-sectional area of the sample, one can obtain the
stress–strain curve for an ideal rubber.
The retractive force of an elastomer, as
explained above, increases with the temperature. In other words,
the temperature of elastomers increases with adiabatic stretching
[21, 22]. The equation for the relationship was
written by Kelvin back in 1857 [22]:
where C p is
specific heat and ∂l again
is the change in length of the elastomer. Experimental evidence
supports this, as the temperature of elastomers, like rubber, rises
upon stretching. This equation can also be written in another form:
In actual dealing with polymers, stretching
rubber and other elastomers requires overcoming the energy barriers
of the polymeric chains with the internal energy of the material
depending slightly on elongation, because
∂V is the change in volume.
At normal pressures the second term on the right becomes
negligible. It represents deviation from ideality. The contribution
of the internal energy E to
the force of retraction is
Bueche [16]
expressed differently the work done on stretching an elastic
polymeric body. It describes deforming an elastomer of x length, stretched to an increase in
length α in a polymeric
chain composed of N freely
oriented segments. The other dimensions of this polymeric chain are
y and z. These are coordinates that become
reduced as a result of stretching to 1/(α)0.5.
where the chain ends are initially r 0 distance apart
[16],
Bueche [16] also
described the average energy per chain as
The modulus of elasticity, G, was shown to be related to strain in
elongation of polymeric elastomers. For up to 300% elongation, or
more, the following relationship [22] applies:
Where k is Boltzmann’s
constant, m represents the
number of polymeric chains in the sample, and γ is the strain. The relationship of
stress to strain is:
There are several molecular theories of
rubber-like elasticity The simplest one is based on a Gaussian
distribution function for the end to end separation of the network
chains: [23] (the dimensions of
the free chains as unperturbed by excluded volume effect are
represented by (r
2)0)
Within this Gaussian distribution function
(r
2)0 applies to the network chains both in the
unstretched and stretched state. The free energy of such a chain is
described by a Boltzmann relationship [23]:
C(T) is a constant at a specified
absolute temperature T. The
change in free energy for a stretched elastomer can be expressed as
follows:
where α is the molecular
deformation ratio of r
components in x,y,z
directions from the unstretched or elastomer to one that was
stretched and deformed. Additional discussions of this theory and
other theories of elasticity are not presented here because
thorough discussions of this subject belong to books dedicated to
physical properties of polymers.
2.2.3 Rheology and Viscoelasticity of Polymeric Materials
When an amorphous polymer possesses a certain
amount of rotational freedom, it can be deformed by application of
force. Application of force will cause the polymer to flow and the
molecules will move past each other. The science of deformation and
flow is called rheology. In the event that the
force is applied quickly, and then withdrawn rapidly, the polymer
molecules will tend to revert back to their previous undisturbed
configuration. This is called relaxation. Thus, the amorphous
polymers exhibit some elastic behavior due to disruption
of thermodynamically favorable arrangements. If, however, the force
is applied gradually and consistently, the molecules will flow
irreversibly. Due to chain entanglement that increases with
molecular weight and due to frictional effects, the viscosity of
the flowing liquid will be high. Thus, molecular weight control is
very important in polymer processing. In a way this is a
paradoxical situation. Higher molecular weights usually yield
better mechanical properties. On the other hand, higher molecular
weight materials are harder to process. The molecular weight
control, therefore, is quite critical. The combination of
properties of polymeric liquids, elasticity, and viscous flow is
referred to as viscoelasticity. It means
reversible uncoiling of the polymeric chains. By the same
terminology, viscous
flow means irreversible slipping of the chains past each
other. Thus, viscoelastic materials exhibit simultaneously a
combination of elastic and viscous behavior. This type of behavior
is particularly prominent in polymeric materials. The flow behavior
of polymeric liquids is also influenced by the molecular weight
distribution.
At proper temperatures the mechanical properties
of many amorphous polymers may approach the physical properties of
three idealized materials individually. These are [20, 22,
24]:
1.
A
Hookian or an ideal elastic solid, whose small
reversible deformations are directly proportional to the applied
force.
2.
A
Newtonian liquid that
flows with a viscosity independent of the rate of shear.
3.
An
ideal elastomer that is
capable of reversible extension of several hundred percent, with a
much smaller stiffness than that of a Hookian solid.
In the elastic response for a Hookian liquid the
stress–strain relationship is: σ(t) = Gγ(t). For the Newtonian liquid it is
σ(t) = ηγ′(t). In these equations, γ(t) and γ′(t) represent shear strain and shear
rate at time t, while
σ(t) is the shear stress.
Many forces can be applied to polymer
deformation. The force that rheologists are particularly concerned
with is tangential
stress or shear.
This is due to the fact that many polymers are extruded and forced
to flow into dies for shaping and commercial use. If a shearing
force f is applied to a
cube of molten polymer per unit area it causes the top layer of the
liquid to move a distance x
from a fixed point at the bottom of the material with a velocity
v. Shear causes the
molecules to move past each other. This is illustrated in
Fig. 2.4.
Fig.
2.4
Illustration of the movement of the upper
layer due to applied force
The above assumes that the viscosity of the
material is sufficiently small so that the cube is not very
distorted during the process. The viscosity of the material,
η is defined as the ratio
between the applied force and the velocity gradient, ∂v/∂x or as the rate of shear γ, where
The above equation can be rearranged into another
form,
Fv is the energy used up
per second on the cube.
The shear
stress, τ is
defined as
where γ is the shear rate. In this instance it
should be equal to v. If
the viscosity η is
independent of the shear rate, the liquid exhibits ideal flow and
is a Newtonian liquid.
Tobolsky commented [12] that
probably all Newtonian liquids, even those like water and benzene,
that are very fluid, possess some elastic as well as viscous
behavior. However, flow of most polymer liquids deviates strongly
from an ideal behavior and either the viscosity decreases with the
rate of shear, or no flow occurs until a certain minimum force is
applied. The decrease of viscosity with the rate or shear is called
shear thinning. It is
believed to be a result from the tendency of the applied forces to
disturb the long chains from their favored equilibrium
conformations. In this case, the polymer is at an yield point or at an yield value. A liquid with an yield
point is called a Bingham
Newtonian fluid,
provided that it exhibits Newtonian behavior once it starts
flowing. It is defined by
where τ c is the
critical shear stress, or threshold stress that is needed to
initiate the flow. The Bingham fluid behavior might be attributed
to structural arrangements of the molecules that give rise to
conformational and secondary bonding forces. The non-Newtonian behavior occurs when shear
stress does not increase in proportion to shear rate. In addition,
there are thixotropic
liquids that exhibit high viscosity or even resemble gelation at
low shear rate but flow readily and exhibit low viscosity upon
application of high shear. High shear rates can cause chain rupture
and result in loss of molecular weight. In some cases, the shear
rate may increase due to increase in molecular entanglement. In the
case of flexible chain polymers, there is a critical molecular
weight at which chain entanglement may show an increase. For most
common polymers this may be in the molecular weight range of
4,000–15,000. Flow is also influenced by chain branching. The
higher the degree of branching in a polymer, the lower will be the
degree of entanglement at a given molecular weight. In general, the
viscosity is higher with linear polymers than with branched ones at
a given shear rate and molecular weight. Flow behavior is also
influenced by molecular weight distribution (see
Sect. 2.7.1). Usually, the broader the molecular
weight distribution in polymers the lower is the shear rate that is
needed to cause shear
thinning. On the other hand, for polymers with narrow
molecular weight distribution, shear thinning, once it starts, is
more pronounced. This is due to absence of chain entanglement of
the higher molecular weight polymeric chains.
The viscosity of low molecular weight polymers is
related to their temperature by an Arrhenius-type relationship
[19, 22, 25]:
where E is the activation energy for viscous flow, and A is a constant related to molecular
motion, and M
η is the
viscosity average molecular weight. For branched polymers, the
larger or bulkier the side chains the greater is the activation
energy, E. The same is true
of polymers with large pendant groups. The activation energy of
flow, E does not increase
proportionately to the heat of vaporization for polymers but rather
levels off to a value that is independent of molecular weight. This
implies that for long chains the unit of flow is much smaller than
the whole molecule. Instead, segments of the molecules, no larger
than 50 carbon atoms move by successive jumps. This is accomplished
with some degree of coordination, but results in the whole chain
shifting. Based on experimental evidence, the viscosity can be
defined as,
where C is a constant.
M w is weight
average molecular weight (see Sect. 2.7) Chain length,
Z, or the molecular weight
of polymers is an important variable that influences flow
properties of polymers. The relationship of a Newtonian viscosity
of an amorphous polymer to the chain length when shear stress is
low can be expressed as [19,
22, 25],
where the constant k is
temperature-dependent. By the same token, based on experimental
evidence, the relationship of viscosity to temperature and to chain
length at low shear rates, for many polymers can be expressed as
follows:
where constant k′ is
specific for each polymer and T g is the glass transition
temperature (T g
is discussed in Sect. 2.2.3). Although linear molten polymers exhibit
well-defined viscosities, they also exhibit elastic effects. These
effects are present even in the absence of any crosslinks or a
rubber network. It is referred to as creep. This creep is attributed to
entanglement of polymeric chain to form temporary physical
crosslinks: This is illustrated in Fig. 2.5.
Fig.
2.5
Illustration of a physical crosslink in
molten polymers
Deviations from Newtonian flow can occur when
shear stress does not increase in direct proportion to shear rate.
Such deviation may be in the direction of thickening (called
dilatent flow) and in the
direction of thinning (called pseudo
plastic). Related to non-Newtonian flow is the behavior of
thixotropic liquids when subjected to shear, as explained above.
Flow behavior can be represented by the following equation:
where A is a constant and
B is the index of flow. For
Newtonian liquids A = η and B = 1. All polymers tend to exhibit
shear thinning at high shear rates. This is commonly explained by
molecular entanglement, as mentioned above. Certainly, in the
amorphous state there is considerable entanglement of the polymeric
chains. Low shear rates may disrupt this to a certain degree, but
the chains will still remain entangled. As the shear rates
increases, disruption can occur at a faster rate than the chain can
reentangle. The decreased amount of entanglement results in lower
viscosity of the liquid, allowing the molecules to flow with less
resistance. Actually, two factors can contribute to chain
entanglement. These are high length of the chains for very large
molecules and/or bulky substituents. Stress applied to a Newtonian
liquid, outside of an initial spike, is zero. Stress, however
applied to a viscoelastic fluid starts at some initial value. This
value decreases with time until it reaches an equilibrium value due
to the viscoelastic property of the material.
Figure 2.6 illustrates what a plot of the modulus of
elasticity G(t), which depends on the temperature,
when plotted against time, looks like:
Fig.
2.6
Illustration of a plot of modulus of
elasticity against time
The equation for shear–stress relaxation modulus
that varies with temperature can be written as follows:
With constant stress, σ(t) = Gγ 0, where creep strain
γ(t) is constant [γ(t) = σ 0/G] for a Hookean solid. It would be
directly proportional to time for a Newtonian liquid [(γ(t) = σ 0/η)t]. Here t is the initial time at which recovery
of the viscoelastic material begins. For a viscoelastic fluid, when
stress is applied, there is a period of creep that is followed by a
period of recovery. For such liquids, strains return back toward
zero and finally reach an equilibrium at some smaller total strain.
For the viscoelastic liquid in the creep phase, the strain starts
at some small value, but builds up rapidly at a decreasing rate
until a steady state is reached. After that the strain simply
increases linearly with time. During this linear range, the ratio
of shear strain to shear stress is a function of time alone. This
is shear creep
compliance,
J(t) The equation of shear creep
compliance can be written as follows:
The stress relaxation modulus and the creep
compliance are both manifestations of the same dynamic process at
the molecular level and are closely related. This relationship,
however, is not a simple reciprocal relations that would be
expressed as G(t) = 1/J(t), but rather in an integral equation
that is derived from the Boltzman superposition principle. It
relates recoverable compliance, to η 0, zero shear viscosity
[22].
In relaxation back to equilibrium, the polymer
assumes a new conformation. At first, the response is glassy. The
modulus for such an organic glass is large, G
g ~ 109 Pa. This modulus decreases with
time as the polymer begins to relax and continues along the whole
length of the chain. For short chains this relaxation to zero takes
place at a fairly constant rate. For very long chains, however, the
relaxation rate tends to be in three stages. It starts at a certain
rate at stage one, but after a while, noticeably slow down at some
point, and at stage two the modulus remains relatively constant
over some period of time. After that, at stage three, the
relaxation is resumed again at a rapid rate until full equilibrium
is reached. At stage two there is a period of relatively constant
modulus that is not affected by the chain architecture and the
material resembles a rubber network. The length of the third stage,
however, is profoundly affected by the molecular weight, by the
molecular weight distribution and by long-chain branching of the
polymer.
Koga and Tanaka [23] studied the behavior of normal stresses in
associated networks composed of telechelic polymers under steady
shear flow. They showed numerically that the first and second
normal stress coefficients reveal thickening as a function of shear
rate and that the sign of the second normal stress coefficient
changes depending on the nonlinearity in the chain tension, the
dissociation rate of the associative groups from junctions and the
shear rate by analytic calculation they showed that in the limit of
small shear-rate, the sign inversion occurs by the competition
between the nonlinear stretching and dissociation of associative
groups. Thus, the molecular mechanism of the sign inversion is
shown to be similar to that of thickening of the shear
viscosity.
In the behavior of polymeric liquids two
quantities are important. These are steady-state recoverable shear
compliance, (as shown above) and steady-state
viscosity at zero shear rate, η 0. These quantities are
related:
where γ′ss is
the shear rate and γ
r is the total recoil strain. Both shear compliance and
shear viscosity can be obtained from creep studies. The product of
the two, zero shear compliance and zero shear viscosity is the
characteristic relaxation time of the polymer:
There are various techniques for determining the
viscosity of molten polymers. One commonly used piece of equipment
is a cone and plate rotational viscometer. The equipment is
illustrated in Fig. 2.7:
Fig.
2.7
Illustration of a cone and plate
rheometer
The molten polymer is placed between the bottom
plate and the cone, and the cone is rotated at constant speed.
Shear stress is obtained from the following relationship:
where M is the torque in
dynes per centimeter, and R
is the cone radius in centimeters (or meters). The shear rate can
be obtained from the following equation
where α is the cone angle
in radians or in degrees, and Ω is the angular velocity in radians
per second or in degrees per second. The viscosity is obtained from
the following relationship [15,
16]:
k is a constant, specific
for the viscometer used. It can be obtained from the relationship
[7]:
The cone and plate rheometers are useful at
relatively low shear rates. For higher shear rates capillary
rheometers are employed. They are usually constructed from metals.
The molten polymer is forced through the capillary at a constant
displacement rate. Also, they may be constructed to suit various
specific shear stresses encountered in commercial operation. Their
big disadvantage is that shear stress in the capillary tubes varies
from maximum at the walls to zero at the center. On the other hand,
stable operation at much higher shear rates is possible.
Determination, however, of η0 is usually not possible
due to limitations of the instruments. At low shear rates. one can
determine the steady-state viscosity from measurements of the
volumetric flow rates, Q
and the pressure drop:
where, P 0 is
the ambient pressure. A capillary viscometer is illustrated in
Fig. 2.8,
where the diameter of the capillary can be designated as
D. For Newtonian liquids
the viscosity can be determined from the following equation:
where L represents the
length of the capillary. The shear stress at the capillary wall can
be calculated from the pressure drop:
Fig.
2.8
Illustration of a capillary
viscometer
Also, the shear rate at the walls of the
capillary can be calculated from the flow rate [22]:
Wang et al. studied the homogeneous shear, wall
slip, and shear banding of entangled polymeric liquids in
simple-shear rheometry, such as in capillary viscometry, shown
above [20]. They observed that
recent particle-tracking velocimetric observations revealed that
well-entangled polymer solutions and melts tend to either exhibit
wall slip or assume an inhomogeneous state of deformation and flow
during nonlinear rheological measurements in simple-shear
rheometric setups.
It is important to control the viscoelastic
properties of confined polymers for many applications. These
applications are in both, microelectronics and in optics. The
rheological properties of such films, however, are hard to measure.
Recently, Chan et al. [27]
reported that the viscoelasticity can be measured through thermal
wrinkling. Thermally induced instability develops when polymer
films are compressed between rigid, stiffer layers. This is due to
differences in coefficients of thermal expansion between the
polymer and the inorganic layers. A net compressive stress develops
at the polymer–substrate interface when the composite layers are
heated to temperatures that promote mobility of the polymer layer.
Such wrinkling substrate surface is characterized by an isotropic
morphology that can be approximated as a sinusoidal profile. Chan
et al. utilized the thermal wrinkling to measure the rubbery
modulus and shear viscosity of polystyrene thin films as a function
of temperature. They used surface laser-light scattering to
characterize the wrinkled surface in real-time in order to monitor
the changes in morphology as a function of annealing time at fixed
annealing temperatures. The results were compared with a
theoretical model, from which the viscoelastic properties of the PS
thin film are extracted.
2.3 The Crystalline State
Mendelkern [5]
pointed out that essentially all properties of polymers are
controlled by the molecular morphology. In contrast to the
amorphous or liquid state, the crystalline state is relatively
inelastic and rigid. In the crystalline state, the bonds adopt a
set of successive preferred orientations, while in the liquid state
the bond orientation is such that the chains adopt statistical
conformations. In the crystalline state the properties of the
polymers differ considerably. If a polymer molecule is coiled
randomly, then it cannot fit readily into a geometrically arranged,
regular crystalline lattice. So the molecules must change into a
uniform shape to fit into a crystal pattern. In many cases they
assume either a helix or a zigzag conformation. Such arrangements
are more regular than in a random coil [29–40]. This
can even be detected by spectral and thermodynamic studies. During
crystallization, polymers with bulky substituents that are spaced
close to each other on the polymeric chains tend to form helical
conformations that remain in the crystalline phase. The arrangement
allows close packing of the substituents without much distortion of
the chain bonds. That is particularly true of many isotactic
polymers that crystallize in helical conformations, taking on
gauche and trans positions. For the
gauche position steric
hindrance always forces the rotation to be such as to place the
substituents into a juxtaposition generating either a right-hand or
a left-had helix. The helical conformations of isotactic vinyl
polymers were illustrated by Gaylord and Mark [41] as shown in Fig. 2.9:
Fig.
2.9
Illustration of the helical conformation of
isotactic vinyl polymers
When macromolecules possess a certain amount of
symmetry, then there is a strong accompanying tendency to form
ordered domains, or crystalline regions. Crystallinity, however, in polymers
differs in nature from that of small molecules. When the small
molecules crystallize, each crystal that forms is made up of
molecules that totally participate in its makeup. But, when
polymers crystallize from a melt, which means that certain elements
of the polymeric system or segments of the polymeric chains have
attained a form of a three-dimensional order. Complete
crystallization, from the melt, however, is seldom if ever achieved
The ordered conformations may be fully extended or may be in one of
the helical forms as shown above. This resembles orderly
arrangement of small molecules in crystals. The crystalline
domains, however, are much smaller than the crystals of small
molecules and possess many more imperfections.
2.3.1 Crystallization from the Melt
Certain basic information was established about
the crystallization
from the melt
[5]: The process is a first-order
phase transition and follows the general mathematical formulation
for the kinetics of a phase change. Equilibrium conditions,
however, are seldom if ever attained and as a result complete
crystallinity, as stated above, is not reached. The tendency to
crystallize depends upon the following:
1.
Structural regularity of the chains that leads to
establishment of identity periods.
2.
Free rotational and vibrational motions in the
chains that allow different conformations to be assumed.
3.
Presence of structural groups that are capable of
producing lateral intermolecular bonds (van der Wall forces) and
regular, periodic arrangement of such bonds.
4.
Absence of bulky substituents or space
irregularity, which would prevent segments of the chains from
fitting into crystal lattices or prevent laterally bonding groups
from approaching each other close enough for best
interaction.
Natta and Carradini [30] postulated three principles for determining
the crystal structures of polymers These are:
1.
It is possible to assume that all mer units in a
crystal occupy geometrically equivalent positions with respect to
the chain axis. This is known as the Equivalence Postulate.
2.
The conformation of the chain in a crystal is
assumed to approach the conformation of minimum potential energy
for an isolated chain oriented along an axis. This is the
Minimum Energy
Postulate.
3.
As many elements of symmetry of isolated chain as
possible are maintained in the lattice, so equivalent atoms of
different mer units along an axis tend to assume equivalent
positions with respect to atoms of neighboring chains. This is the
Packing
Postulate.
X-ray diffraction studies of polymers
crystallized from the melt show recognizable features in some of
them. The Bragg reflections, however, appear more broad and diffuse
than those obtained from well-developed single crystals. Such
broadening could be the result of the crystals being small. It
could also be the result lattice defects. Because diffraction
patterns are too weak, it is impossible to reach a conclusion. The
majority opinion, however, leans toward the small crystal size as
being the cause of the broadening [5]. The crystals from the melt are approximately
100 × 200 × 200 Å in size. Rough estimates from these
diffraction studies indicate that the size of crystals, or
crystallites, rarely exceeds a few hundred angstroms. The fact that
there is a substantial background of diffuse scattering suggests
that considerable amorphous areas are also present. Because the
chains are very long, it has often been suggested that individual
chains contribute to several different crystalline and amorphous
domains. This resulted in a proposal of a composite single-phase
structure [5], a fringed micelle or a fringed crystallite model. This is
illustrated in Fig. 2.10. The fringes are transition phases between
the crystalline and the amorphous regions. Some analytical studies,
however, failed to support this concept.
Fig.
2.10
Illustration of a fringed micelle
The proportion of crystalline matter in a polymer
is defined as the degree
of crystallinity. It can be expressed
in volume or in mass. By expressing it in volume, it would be:
where, V c,
V a, and
V represent the respective
volumes of the crystalline and amorphous regions and the total
volume.
When expressed in terms of mass, the equation
would be:
where, m c,
m a, and
m are the masses of
crystalline, amorphous phases, and the total mass.
2.3.2 Crystallization from Solution
For many polymers, crystal growth can also take
place from dilute solutions and from such solutions they yield
single crystals. Crystal
formations in polymers were studied intensively almost from the
time of recognition of their existence in macromolecules.
Single crystals of organic
polymers were recognized as early as 1927 [26]. It became the subject of intensive
investigations after observations that linear polyethylene can
crystallize into single crystals. The observations made on
polyethylene were followed by observations that it is possible to
also grow single crystals of other polymers. Some of them are
polyoxymethylene [30], polyamides
[31], polypropylene
[32], polyoxyethylene
[33], cellulose [34], and others [35]. The electron diffraction pattern and a
single crystal of polyacrylonitrile are shown in
Fig. 2.11.
Fig.
2.11
Electron diffraction pattern and single
crystal of polyacrylonitrile (from ref [50])
There is good correlation between the structural
regularity of a specific polymer and the appearance of its single
crystals. Relatively larger single crystals with smooth, sharp
edges, and little random growth are formed only from polymers that
are known to have high regularity in their structures.
To explain the arrangement of the chains in the
single crystals a theory of folded chain lamella was proposed in 1947
[36]. It states that the basic
units in single crystals are lamellae about 100 Å thick.
The evidence for the existence of lamellae-like crystallites comes
from microscopic observations. Direct microscopic observations,
however, do not yield information about chain structure on the
molecular level. The thickness of the lamellae relative to the
chain length led some to postulate that the molecules are arranged
perpendicular or nearly so to the layers. Also that sharp, folded
configurations form with the fold length corresponding to the layer
thickness. In this view, one polymer molecule is essentially
constrained to one lamellae and the interface is smooth and
regularly structured with the chain folding back and forth on
itself like an accordion during crystallization. Two models of
chain folding are visualized. In one model, the chain folding is
regular and sharp with a uniform fold period. This is called the
adjacent-reentry model.
In the other one, the molecules wander through the irregular
surface of a lamella before reentering it or a neighboring lamella.
This mode is called switchboard or nonadjacent-reentry model. The two
are illustrated in Fig. 2.12. Some early experimental evidence failed
to support the lamellae concept [37]. Since, much evidence was presented to
support it, today it is accepted as an established fact
[5]. Recently, Hosoda and coworkers
[37] reported studies of the
morphology of specially synthesized polyethylene with n-butyl branches precisely spaced on
every 39th carbon. This was compared with a commercial
ethylene/1-hexene copolymer possessing the same branching
probability. The goal of their work was to elucidate the effect of
the intramolecular sequence length heterogeneity on the lamella
crystal thickness and its distribution. The commercial copolymer
was found to have an orthorhombic crystalline polymorphism, which
is normal for such polyethylenes and different from that of the
specially synthesized material. The synthesized one exhibits a
narrow lamella thickness distribution with the average thickness
corresponding exactly to the space length between two consecutive
branches. This suggests to them. complete exclusion of n-butyl branches from the crystal stem.
On the other hand, the 1-hexene copolymer forms much thicker
lamellae and a broader thickness distribution than the synthesized
polyethylene with butyl branches. Here, the average thickness is
1.5 times larger than that calculated from the most probable
ethylene sequence length obtained from NMR, or for a theoretical
ethylene sequence Length distribution, indicating that the lamellae
are composed predominantly of the sparsely branched longer ethylene
sequences.
Fig.
2.12
Models of chain folding. (a) adjacent reentry model, and
(b) nonadjacent reentry
model
Linear polyethylene single crystals often exhibit
secondary structural features that include corrugations and pleats.
It was suggested that the crystals actually grow in form of
pyramids, but that these pyramidal structures collapse when the
solvent is removed during preparation for microscopy
[42]. Various investigators
described other complex structures besides pyramidal. Typical among
these descriptions are sheaf-like arrays that would correspond to
nuclei. Also there were observations of dendridic growths, of
clusters of hollow pyramids, of spiral growths, of epitaxial
growths, of moiré patterns, etc. [42, 43].
2.3.3 Spherulitic Growth
For polymers that crystallize from the melt, an important parameter in the
characterization of the two-phase systems, is the weight fraction
of the crystalline regions. The degree of crystallinity that can be
reached is dependent on the temperature at which crystallization
takes place. At low temperatures one attains a much lower degree of
crystallization than at higher temperatures. This implies that
crystallization remains incomplete for kinetic reasons
[7].
Normal conditions of cooling of the molten
polymer establish the crystalline texture of the polymer and
usually result in formation of very tiny crystals. These crystals
are part of a closely spaced cluster called spherulite. The formation of a
single nucleus in a polymer cooled below its melting point favors
the formation of another nucleus in its vicinity due to creation of
local stresses.
Microscopic examinations with polarized light of
many polymeric materials that crystallized from the melt show the
specimen packed with spherulites. Often these appear to be
symmetrical structures with black crosses in the center
[38]. It is believed
[39] that these spherulites grow
from individual nuclei. Ribbons of crystallites grow from one
spherulitic center and fan out in all directions. Initially they
are spherical but because of mutual interference irregular shapes
develop. The diameters of spherulites range from 0.005 to
0.100 mm. This means that a spherulite consists of many
crystalline and noncrystalline regions. The black crosses seen in
the spherulites are explained [39]
by assuming that the crystallites are arranged so that the chains
are preferentially normal to the radii of the spherulites.
Spherulitic morphology is not the universal mode of polymer
crystallization. Spherulitic morphology occurs usually when
nucleation is started in a molten polymer or in a concentrated
solution of a polymer. Spherulitic growth is illustrated in
Fig. 2.13.
Fig.
2.13
Spherulitic growth (from ref
[50])
The size and number of spherulites in the polymer
tends to affects the physical properties. Thus, the impact strength
of polymer films or their flex life usually increases as the
spherulite size decreases. On the other hand, there does not appear
to be any correlation between the yield stress and ultimate
elongation and the size of the spherulites.
Rhythmic crystal growth is generally encountered
in thin films of semicrystalline polymers. This is believed to be
due to formation of ring-banded spherulites and attributed to the
periodical lamellae twisting along the radial growth direction of
the spherulites [42]. Recently, Gu
and coworkers [43] reported that
by using mild methylamine vapor etching method, the periodical
cooperative twisting of lamellar crystals in ring-banded
spherulites was clearly observed.
When the melt or the solutions are stirred
epitaxial crystallinity
is usually observed. One crystalline growth occurs right on top of
another. This arrangement is often called shish-kebab morphology. It contains
lamella growth on long fibrils. Drawing of a crystalline polymer
forces the spherulites to rearrange into parallel arrangements
known as drawn fibrilar
morphology.
In order for the ordered phase to crystallize
from an amorphous melt a nucleation barrier must be overcome. This
barrier is a result of interfacial energy between the ordered phase
and the melt that causes super cooling. Sirota [44] suggested that in order for the nucleation
barrier of the stable phase to be sufficiently high to form out of
the melt, another phase with a lower nucleation barrier and a free
energy intermediate between that of the stable phase and the melt
must form. This, he points out, is implied by Oswald’s rule
[45] and evidence presented by
Keller [35] that crystallization
in semicrystalline polymer systems is mediated by a transient
metastable phase [47,
48].
Stroble and coworkers demonstrated that lamellar
thickness is determined by a transition between the metastable
phase and the stable crystal phase [46–49]. In
addition, by relating the crystallization temperature, melting
temperature, and crystalline lamellar thickness, he suggested that
lamellar growth fronts are thin layers of a mesomorphic phases. He
feels that these phases thicken until such thickness is reached
that stable crystal phases are favored. The conversion occurs in a
block wise fashion but results in granular structures that were
observed in many polymers [46–49]. This
conversion is a stabilization process that lowers the free energy
of the newly formed crystallites and prevents them from returning
to the mesomorphic phase upon subsequent elevation of the
temperature [46–49]. Such a concept of crystallization, however,
is not universally accepted. Sirota [49], pointed out, however, that this picture and
the thermodynamic framework are generally correct in many
cases.
Sirota [49]
believes that the origin of granular structures, mentioned above,
can be understood in the following way. The initial nucleation and
growth take place by stem addition. into mesophases. Lamellae
thickness occurs while the chains are in the more mobile mesophase.
When the thicknesses grow large enough to allow conversion from
mesophases to crystals, the average densities in the lamellae have
been set and the crystals break up into blocks. The transitions
from mesophases to crystals also involves increases in lateral
packing densities. In semicrystalline polymers, the entanglements
in the amorphous regions, as well as the folding of the chains and
the lamellar spanning the chains, will also have an effect. These
effects limit the allowable lateral displacements of the
crystalline stems during the rapid solid-state transformation from
the mesophases to the crystal forms. As a result, over long
distances, on molecular scale, and short times, the stems are
kinetically constrained to be close to the lateral position that is
maintained as they grew in the mesophase. At the same time, the
local packing now favors the more stable crystal phase with a
higher lateral density. The constraints on the chain ends favoring
the lateral density of the mesophases, and forces competition with
the density of the thermodynamically favored crystal structures.
This causes the crystal to break up into domains. Sirota then
suggests that the granular structures observed in many polymer
crystals are fingerprints of the transitions from the transient
intermediates from which the crystals have formed [49].
It was reported recently, that polymeric can also
form quasicrystals. Hayashida et al. [50] demonstrated that certain blends of
polyisoprene, polystyrene, and poly(2-vinylpyridine) form
star-shaped copolymers that assemble into quasicrystals. By probing the
samples with transmission electron microscopy and X-ray diffraction
methods, they conclude that the films are composed of periodic
patterns of triangles and squares that exhibit 12-fold symmetry.
These are signs of quasi-crystalline ordering. Such ordering differ
from conventional crystals lack of periodic structures yet are
well-ordered, as indicated by the sharp diffraction patterns they
generate. Quasi-crystals also differ from ordinary crystals in
another fundamental way. They exhibit rotational symmetries (often
five or tenfold). There are still some basic questions about their
structure.
2.3.3.1 Thermodynamics of Crystallization
The free energy change of a polymer from liquid
to crystal is
At the melting point T m the crystals and the
liquid polymer are at equilibrium. The change in free energy,
therefore, ΔF = 0. That
means that at the melting point:
If the heat of fusion is large, then the melting
temperature of the polymer crystals will be high (due to high
intermolecular attraction) or if the entropy of fusion is
small.
Mendelkern [42]
noted that there are three different interfacial free energies that
are characteristic of crystallites. One, σ e, is for the equilibrium
extended chain crystallite, a second one σ ec represents the mature,
but non-equilibrium crystallite, and the third one is σ en is the interfacial free
energy involved in forming a nucleus. These quantities cannot be
identified with one another. Because only portions of the polymeric
chains participate in the formations of crystallites, the section
or sections of the chains of x length that participate in
crystallite formation can be designated as ζ e′ and the sections of the
chains that remain in disorder and amorphous, as x − ζ e′.
The dependence of the melting temperature of such
a system upon chain length is [42]:
where, ΔH u is
the enthalpy of fusion per repeat unit and ζ e is the equilibrium
crystalline length. The effective interfacial free energy
associated with the basal plane of an equilibrium crystalline
length ζ e′′
designated as σ
e is [42]:
where is the equilibrium melting
temperature for an infinite molecular weight chain and T me is the corresponding
melting temperature for a fraction that contains x repeat units. From the stand point of
thermodynamics, T
me is a first-order
transition temperature [15]. The melting points, when measured very
carefully, can in many cases be sharp. On the other hand, melting
points of ordinary crystals may melt over a range
(Table 2.3). Table 1.2 shows some first-order
transition temperatures, commonly designated as T m.
Table
2.3
Melting points, T m of some crystalline
polymersa
Repeat unit
|
Polymer
|
Melting point (°C)
|
---|---|---|
Poly(tetramethylene suberate)
|
45
|
|
Poly(ethylene oxide)
|
66
|
|
Poly(propylene oxide)
|
70
|
|
High-density polyethylene
|
132–138
|
|
Polyacrylonitrile
|
317
|
|
Poly(vinyl chloride)
|
212
|
|
Poly(vinylidene chloride)
|
210
|
|
Polypropylene
|
168
|
|
Polychlorotrifluoroethylene
|
210
|
|
Polystyrene
|
230–248
|
|
Poly(hexamethylene adipamide)
|
250
|
|
Polytetrafluoroethylene
|
327
|
For systems that are polydisperse, with a most
probable chain-length distribution, the melting temperature
molecular weight relation is expressed as [42]:
where is the number average
degree of polymerization. In this equation, the quantity
2⁄ represents the mole
fraction of non-crystallizing units. This equation is based on the
stipulation for conditions for phase equilibrium. It is specific to
and valid only for polymers that have a most probable molecular
weight distribution. This relationship for the melting temperature
of each polydisperse system has to be treated individually
[42]. By applying the Clapeyron
equation and from measurements of applied hydrostatic pressure the
value of ΔH u
can be determined.
This equation, T m = ΔH m/ΔS m applies to very high
molecular weight polymers. For polymers that are medium or low in
molecular weight, the degree of polymerization (X) has to be included:
Mandelkern points out [38] that the experimental data shows that there
is no correlation between the melting temperature of a polymeric
crystal and the enthalpy of fusion, as is found in many small
molecules. The ΔH values of
polymers generally fall into two classes. They are ether within a
few thousand calories per more or about 10,000 cal/mole.
Polymers that fall into the category of elastomers have low melting
temperatures and high entropies of fusion. This reflects the
compacted highly flexible nature of the chains. At the other
extreme are the rigid engineering plastics. These materials possess
high melting points and correspondingly lower entropies of fusion
[42].
2.3.3.2 Kinetics of Crystallization
The rate of crystallization in polymeric
materials is of paramount importance. For some polymers, like
atactic polystyrene or some rubbers, rapid cooling can lead to the
glassy state without any formation of crystallites. The amount of
crystallinity actually depends very much upon the thermal history
of the material. The amount of crystallinity, in turn, influences
the mechanical properties of the material. Microscopic observation
of the growth of the spherulites as a function of time will yield
information of the crystallization rate. The rate is a function of
the temperature. As the temperature is lowered, the rate increases.
This growth is usually observed as being linear with time. Presence
of impurities will slow down the growth rate. When the growth rate
is plotted against crystallization temperature, a maximum is
observed. This is due to the fact that as the temperature is
lowered the mobilities of the molecules decrease and the process
eventually becomes diffusion-controlled. According to the Avrami
equation, the fraction that crystallizes during the time
t, and defined as
1 − λ(t), can be written as [42]:
where, N(τ) is the nucleation frequency per
untransformed volume, V(t,τ) is the corresponding volume of the
growing center, and ρ
c and ρ
l represent the densities of crystalline and liquid
phases. Based on that, the rate constant for crystallization
kinetics can be described [42]:
where, V ∞,
V t, and
V 0 are specific
volumes at the times shown by the subscripts, and w c is the weight fraction
of the polymer crystallized. k is the rate constant for
crystallization It was found, however, that crystallization
continues in polymeric materials for much longer periods of time
than the Avrami equation predicts.
For all homopolymers the rate of crystal growth
increases linearly with time, or G = dr/dt. Mandelkern defines the steady-state
nucleation rate, N as
follows [42]:
where E D is the
energy of activation for transporting the chain segments across the
crystal–liquid interface.
If the crystallization takes place over an
extended temperature range, most if not all homopolymers display a
maxima in rates of spherulitic growth and in the overall
crystallization. The equation for spherulitic growth is written as
follows [42]:
where T ∞ is the
temperature at which all molecular and segmental motion
stops.
2.4 The Mesomorphic State, Liquid Crystal Polymers
The state of mesomorphism is a spontaneously
ordered liquid–fluid crystalline state. Liquid crystals were
discovered as early as 1888. They are materials that exhibit order
in one or two dimensions but not in all three. By comparison, the
amorphous materials lack any order, while the crystalline ones
exhibit order in three dimensions. All liquid crystalline polymers
exhibit some degree of fluidity. They were investigated extensively
in the 1900s and became commercially important in 1960s.
These are macromolecules that can align into
crystalline arrays while they are in solution (lyotropic) or while in a molten
state (thermotropic).
Such liquids exhibit anisotropic behavior
[51, 52]. The regions of orderliness in such liquids
are called mesophases.
Molecular rigidity found in rigid rod-shaped polymers, for
instance, is the chief cause of their liquid crystalline behavior.
It excludes more than one molecule occupying a specific volume and
it is not a result of intermolecular attractive forces. Some
aromatic polyesters or polyamides are good examples, like
polyphenylene terephthalate:
Because the molecules posses anisotropy, they are
aligned while still in a fluid form. This differs from ordinary
liquids, that are isotropic,
where the molecules lack any kind of arrangement. Anisotropy is not
affected by conformational changes. Generally, molecules that are
rigid rod like and elongated or disc like in shape are the type
that can form liquid crystal arrangements. Some biological polymers
exhibit liquid crystalline behavior due to their rigid helical
conformations. Among synthetic polymers, on the other hand, rigid
rod structures, mentioned above are the ones that exhibit most of
the liquid crystalline behavior. Polymers that form liquid crystals
may exhibit multiple mesophases at different temperatures. Based on
the arrangement of the liquid crystals in the mesophases, they are
further classified as nematic,
smectic, and cholesteric [51, 52].
Both, smectic and nematic are parallel
arrangements along molecular axes. The smectic liquid crystals are
more ordered, however, than the nematic ones. This is a result of
differences in the orientations of the chain ends. In smectic
liquid crystals the chain ends are lined up next to each other. In
nematic ones, however, they lack any particular orientation. Also
the smectic liquid crystals are layered while the nematic ones are
not. Microscopic observations [51]
can help distinguish between the two forms.
Smectic elastomers, due to their layered
structure, exhibit distinct anisotropic mechanical properties and
mechanical deformation processes that are parallel or perpendicular
to the normal orientation of the smectic layer. Such elastomers are
important due to their optical and ferroelectric properties.
Networks with a macroscopic uniformly ordered direction and a
conical distribution of the smectic layer normal with respect to
the normal smetic direction are mechanically deformed by uniaxial
and shear deformations. Under uniaxial deformations two processes
were observed [53]: parallel to
the direction of the mechanical field directly couples to the
smectic tilt angle and perpendicular to the director while a
reorientation process takes place. This process is reversible for
shear deformation perpendicular and irreversible by applying the
shear force parallel to the smetic direction. This is illustrated
in Fig. 2.14.
Fig.
2.14
Illustration of the arrangement of liquid
crystal into nematic and smectic orders
If the mesogens are chiral, a twisted nematic,
suprarmolecular, cholesteric (twisted) phase can
form [51, 52]. The achiral nonlinear mesogens can also
form chiral supramolecular arrangements in tilted smectic
phases.
Recently, Tokita and coworkers [54] reported a direct transition from isotropic
to smectic arrangement in a liquid crystalline polymer and
determined experimentally the existence of metastable nematic
orientational ordering that preceded the formation of translational
smectic ordering. A polymeric material was used that exhibits very
slow liquid crystalline transition dynamics [55]. This enabled use of conventional methods to
study the transitions, such as of polarized light scattering and
synchrotron wide-angle X-ray diffraction analyses. It was observed
that at high quench rates or super cooling, metastable nematic
(orientational) ordering occurs preceding full smectic
(orientational and translational) order. Also, the occurrence of
nematic preordering (high super cooling) resulted in morphological
changes of growing liquid crystalline domains compared to solely
smectic growth. Specifically, samples cooled at rates high enough
to exhibit nematic preordering formed well-oriented or “neat”
tactoidal smectic domains. Samples cooled at lower rates, where
only smectic ordering was observed, formed radially oriented or
textured spherulitic domains [55].
In commenting on this observation, Abukhdeir and Rey
[56] point out that through a
simulation model the isotropic to smectic liquid crystalline
transition experimental observations of preordering of smectic
liquid crystalline transitions can be studied. Phase transition
kinetics results presented by them show that nematic preordering
results from both thermodynamic potential and dynamic differences
in phase-ordering time scales.
The chemical structure of the polymers determines
whether the molecules can form rigid rods. If the backbone of the
polymer is composed of rigid structures then it tends to form
main chain liquid
crystals. If, however, the side chains are rigid, then the polymer
will tend to form side
chain liquid crystals. From practical considerations, these
two properties are of prime interest. The structures are
illustrated in Fig. 2.15.
Fig.
2.15
Illustration of arrangements of liquid
crystal structure
Liquid crystalline behavior affects the melt
viscosity of the polymer and the ability of the polymer to retain
the ordered arrangement in the solid state. Thus, liquid
crystalline behavior during the melt results in lower viscosity
because the rigid polymeric mesophases align themselves in the
direction of the flow. As a result, the polymer is easier to
process. Also, retention of the arrangement upon cooling yields a
material with greatly improved mechanical properties. Several
thermotropic liquid crystalline copolyesters and polyamides are
available commercially.
Samulski [57]
gives examples of molecules that can typically form liquid crystals
These are
1.
A discotic liquid crystal
The above shown structure has a mesogenic core, (hard central
segment) correlated with dynamic packing of anisometric shapes. The
flexible tales, often hydrocarbon chains, extend from the mesogenic
core and facilitate the transformation from the solid state to the
liquid crystalline phase.
2.
A calamitic liquid crystal
In this case there is a prolate mesogen axis and
flexible hydrocarbon chains that extend from it.
3.
A nonlinear liquid crystal
Many liquid crystal polymers tend to exhibit
multiple mesophases at different pressures and temperatures
[57]. When heated, these polymers
will go through multiple first-order transitions. Such transitions
are from more ordered to less ordered arrangements. This is
referred to as the clearing
temperature with the last one resulting in isotropic
melts.
A lyotropic liquid crystalline aromatic
polyamide, sold under the trade name of Kevlar, is an example of
such a polymer that is available commercially:
The polyamide forms liquid crystals in sulfuric
acid solution from which it is extruded as a fiber. After the
solvent is removed, the remaining fiber possesses greater uniform
alignment than would be obtained by mere drawing. This results in
superior mechanical properties.
There are polymers, however, that exhibit liquid
crystalline behavior, but are very high melting and insoluble in
most common organic solvents. This is a drawback, because such
materials are hard to process.
A preparation of new liquid crystal polymers with
bilaterally linked mesogens in main-chain was reported
[57]. Such materials exhibit
biaxial fluctuation in the nematic phase. This is interesting
because most commonly encountered polymeric liquid crystals have
mesogens linked at their ends to the polymer backbone by flexible
spacers. On the other hand, liquid crystal polymers with mesogens
linked bilaterally by flexible spacers are not common and only a
few examples were reported [58].
One such material can be illustrated as follows:
It was also shown that it is possible to
synthesize polymethacrylate liquid crystal polymers with
mesomorphic properties that contain ferrocenes with two flexible
chains at the l,l′-positions [59].
Based on dilatometric measurements, a head-to-tail molecular
arrangement of the monomeric units occurs within the smectic A
phase. Because of special electrochemical properties of ferrocene,
these materials are of interest for developing electroactive
mesomorphic polymers. The structure of the polymer can be shown as
follows [59]:
Finkelmann reported synthesis of a novel
cross-linked smectic-C main-chain liquid-crystalline elastomer that
was formed by polycondensation of vinyloxy-terminated mesogens,
tetramethyldisiloxane, and pentamethyl-pentaoxapentasilicane. The
introduction of the functional vinyloxy group allows the synthesis
of well-defined networks with good mechanical properties due to
elimination of side reactions as in the case of vinyl groups
[60].
Large amplitude oscillatory shear is frequently
capable of generating macroscopic alignment from an initially
random orientation distribution in ordered polymer fluids.
Berghardt and coworkers [61]
reported that by combined rheological and in situ synchrotron X-ray
scattering to investigate of such induced alignment in smectic
side-groups pf liquid
crystalline polymers. In all cases, they found that shear promotes
anisotropic orientation states in which the lamellar normal tends
to align along the vorticity direction of the shear flow
(“perpendicular” alignment). Rheological measurements of the
dynamic moduli by them revealed that large amplitude shearing in
the smectic phase causes a notable decrease in the modulus. They
also observed that increasing strain promotes higher degrees of
orientation, while increasing molecular weight impedes development
of smectic alignment.
Ahn et al. reported [62] preparation of a smetic liquid crystalline
elastomer with shape memory properties. Shape memory polymeric
materials can recover their equilibrium, permanent shapes from
nonequilibrium, temporary shapes as a result of external stimuli,
like heat or light. Such materials have application in medical
practice. Main-chain polynorbornenes were linked with three
different side-chains, cholesterol, poly(ethylene glycol), and
butylacrylate.
2.5 Orientation of Polymers
There is no preferred direction or arrangement in
the manner in which the macromolecules align themselves in a
polymeric mass during crystallization. If, however, after
crystallization an external stress is applied, the crystalline
material undergoes a rearrangement. From the X-ray diffraction
patter it is surmised that the chains realign themselves in the
direction of the applied stress. Polymeric films and fibers usually
show considerable increase in strength in the direction of that
stress. Oriented fibers are considerably stronger along their
length then perpendicularly to them. Generally, if the degree of
crystallinity in crystalline polymers is well developed prior to
the drawing, the process does not change the amount of
crystallinity appreciably.
The orientation in the direction of applied
stress occurs also in amorphous materials. The amorphous polymers,
like the crystalline ones, also exhibit increased strength in the
direction of orientation. If there is a small amount of
crystallinity in polymers, the crystallinity often increases as a
result of cold drawing.
Orientation of fibers and films is generally
carried out above the glass transition temperature. At the same
time, there usually is a desired limit on T m. Thus it is preferred
that the crystalline melting temperature does not exceed
300°C.
2.6 Solutions of Polymers
Polymers usually dissolve in two stages. First
the materials tend to swell and form gels. The gels then tend to
disintegrate into true solutions. Agitation only speeds up the
process in the second stage. Chains held together by crosslinks
will only swell. Crystallinity or strong hydrogen bonding might
also keep some polymeric materials from dissolution and the
materials only swell.
Some general rules about polymer solubility are:
(a)
Solubility is favored by chemical and structural
similarity between the solvent and the polymer.
(b)
Solubility of the polymer tends to decrease with
an increase in molecular weight.
When a polymer dissolves, the free energy of
solution, ΔF is negative,
while the entropy change, ΔS is positive, because of increased
conformational mobility of the polymer. The magnitude of the
enthalpy of solution determines whether ΔF is positive or negative. The heat of
mixing, ΔH mix
for binary systems was suggested to be related to the concentration
and energy parameters by the expression [7]:
where V mix is
the total volume of the mixture, the molar volumes of the two
components are, V
1 and V
2, respectively, and Φ1 and Φ2 are
their volume fractions. The energies of vaporization are
ΔE 1 and
ΔE 2, while the
cohesive densities
(energy required to remove a molecule from its nearest neighbor)
are ΔE
1/V 1
and ΔE
2/V
2.
2.6.1 Radius of Gyration
In solution, polymeric chains can form different
conformations, depending upon the solvent. When the solvent is such
that the chains are fully solvated, they are relatively extended
and the molecules are randomly coiled. The polymer–solvent
interaction forces determine the amount of space that the molecular
coil of the polymer occupies in solution. While quite extended in a
good solvent, if the solvent is a “poor” one, the chains are curled
up. A measure of the size of the polymer molecule in solution, or
the amount of space that a polymer molecule occupies in solution is
determined as radius of
gyration., or
root mean square radius of gyration, S. Qualitatively, it is the average
distance of the mass of the molecule from the center of its mass
(from its center of gravity). The following equation defines this
relationship [7]. To put it in
other words, it is the square of the distances between various
masses and the center of the mass:
where m is the mass
associated with each of the N chain bonds, and S is the vector distance from the
center of the mass to the terminal chain bond. The size of randomly
coiled polymer molecules is commonly designated by the root-mean
square distance between the ends, R 2. A molecular coil is
illustrated in Fig. 2.16.
Fig.
2.16
Illustration of a molecular coil
The distance between the chain ends is often
expressed in terms of unperturbed dimensions
(S 0 or
R 0) and an
expansion factor
(α) that is the result of
interaction between the solvent and the polymer
The unperturbed dimensions refers to molecular
size exclusive of solvent effects. It arises from intramolecular
polar and steric interactions and free rotation. The expansion
factor is the result of solvent and polymer molecule interaction.
For linear polymers, the square of the radius of gyration is
related to the mean-square end-to-end distance by the following
relationship:
This follows from the expansion factor,
α is greater than unity in
a good solvent where the actual “perturbed dimensions” exceed the
unperturbed ones. The greater the value of the unperturbed
dimensions the better is the solvent. The above relationship is an
average derived at experimentally from numerous computations.
Because branched chains have multiple ends it is simpler to
describe them in terms of the radius of gyration. The volume that a
branched polymer molecule occupies in solution is smaller than a
linear one, which equals it in molecular weight and in number of
segments.
The volume that these molecules occupy in
solution is important in determinations of molecular weights It is
referred to as the hydrodynamic volume. This volume
depends upon a variety of factors. These are interactions between
the polymer molecule and the solvent, chain branching,
conformational factors arising from polarity, restricted rotation
due to resonance, and the bulk of substituents. The above, of
course, assumes that the polymer molecules are fully separated from
each other.
2.6.2 The Thermodynamics of Polymer Solutions
Solutions of polymers deviate to a great extent
from Raoult’s law, except at extreme dilution. In extremely dilute
solutions the ideal behavior is approached as an asymptotic limit.
These deviations arise largely from small entropies of mixing. That
is mostly due to the large difference in size between the solute
and the solvent. The change in the entropy of mixing, according to Flory–Huggins theory of polymer
solutions [63] is:
where subscript 1 denotes the solvent and subscript 2 the solute.
v 1 and
v 2 are
volume fractions. They
are defined as follows
where x is the heat of
mixing. The change in the heat content of mixing of polymer
solutions is similar to that of other solutions
where x 1
characterizes the internal energy per solvent molecule. The change
in free energy of mixing, according to Flory–Huggins
[63], is
The Flory–Huggins treatment overlooks the fact
that dilute solutions of polymers consist of domains or clusters of
polymeric chains that are separated by regions of pure solvent that
is free from the solute. Flory–Krigbaum treatment assumes a model
of dilute polymeric solutions where the polymeric clusters are
approximately spherical and their density reaches a maximum at
their centers and decreases in an approximately Gaussian function
away from the center. The volume that is occupied by the segments
of each molecule excludes the volumes of all other molecules. Long
range intramolecular interactions take place within such excluded
volumes. The thermodynamic functions of such interactions can be
derived, such as the free energy change, the enthalpy change, and
the entropy change:
where Ψ is an entropy parameter and can be expressed as,
In addition, Flory temperature, Θ is treated as
a parameter, Θ = k
1 T/Ψ1 At Θ temperature the
partial molar free energy due to the solute–solvent interaction is
assumed to be zero and deviations from ideality (ideal solution)
become zero. What this means is that as the temperature of the
solution of a polymer approaches Θ, the solvent becomes
increasingly poorer and the excluded volume effect becomes smaller
and approaches zero with the molecules interpenetrating one another
with zero net interaction. The solvent is referred to as a Θ,
solvent. Below Θ temperature the polymer molecules attract each
other, the excluded volume is negative, and the polymer
precipitates. This can be expressed as [63]:
When the chains are extended, their conformations
may be considered as being determined by equilibrium between the
forces of expansion due to excluded volume and the forces of
contraction due to chain segments expanding into less probable
conformations. Based on random flight statistics, the chains are
extended linearly by a factor α over their dimensions. The actual
root-mean-square end-to-end distance is equal to α(R
02)0.5. The change in the elastic
part of free energy is
where the parameter α can
be expressed in terms of thermodynamic quantities:
In the above equation, C m represents a combination
of molecular and numerical constants. Based on the above equation,
at Θ temperature α = 1. It
has been stated that the Flory–Krigbaum treatment must be treated
with some reservations, because it predicts that α increases
without limit with increasing molecular weight [63].
2.7 Molecular Weights and Molecular Weight Determinations
The physical properties of polymers are also
related to their molecular weights. Melt viscosity, hot strength,
solvent resistance, and overall toughness increase with molecular
size. Table 2.4 illustrates the effect of molecular weights
(size) upon physical properties of polyethylene [64].
2.7.1 Molecular Weight Averages
Random events govern the process of synthetic
polymer formation, whether it is by a chain propagating process or
by a step-growth reaction. The result is that the chains vary in
lengths. (There are special methods available, however, in
chain-growth polymerizations that lead to formation of polymer
molecules that are almost equal in length. This is discussed in
subsequent chapters) As a result, most polymeric materials cannot
be characterized by a single molecular weight, but instead must be
represented by statistical averages [64]. These averages can be expressed in several
ways. One way is to present an average as a number average molecular weight. It
is the sum of all the molecular weights of the individual molecules
present divided by their total number. Each molecule contributes
equally to the average and can be obtained by averaging the
measurements of all the colligative properties. If the total number
of moles is N
i , the sum of
these molecules present can be expressed as, ΣN i . The total weight ω of a sample is similarly the sum of
the weights of all the molecular species present
By dividing the total weight of the molecules by
their total number we have the number average molecular weight,
Another way to express a molecular weight average
is as a weight average.
Each molecule in such an average contributes according to the ratio
of its particular weight to that of the total,
The above can be illustrated quite readily by
imagining that a sample of a polymer consists of five molecules of
molecular weights of 2,4,6,8, and 10, respectively. To calculate
the number average molecular weight all the weights of the
individual molecules are added. The sum is then divided by the
total number of molecules in the sample (in this case 5):
To calculate the weight average molecular weight
of the above sample, the squares of each individual weight are
divided by the total sum of their molecular weights, in this case
it is 30:
is more sensitive to the
higher molecular weight species, while is sensitive to the lower
ones. This can be seen by imagining that equal weights of two
different sizes of molecules are combined, M 1 = 10,000 and
M 2 = 100,000.
The combination would consist of ten molecules of M 1 and one molecule of
M 2. The weight
average molecular weight of this mixture is
(108/2 × 105 + 1010/2 × 105) = 55,000
while the number average molecular weight is only 18,182. If,
however, the mixture consists of an equal number of these
molecules, then the weight average molecular weight is 92,000 and
the number average molecular weight is 55,000.
The ratio of the weight average molecular weight
to the number average molecular weight is important because it
affects the properties of polymers. This ratio is called the
molecular weight
distribution:
When all macromolecular species are of the same
size, the number average molecular weight is equal to the weight
average molecular weight. On the other hand, the greater the
distribution of molecular sizes, the greater is the disparity
between averages. The ratio of this disparity, M w/M n is a measure of
polymeric dispersity. A
monodisperse polymer
has a ratio of:
In all synthetic polymers and in many naturally
occurring ones the weight average molecular weight is greater than
the number average molecular weight. Such polymers are polydisperse.
Two samples of the same polymer equal in weight
average molecular weight may exhibit different physical properties,
if they differ in the molecular weight distributions. Molecular
weight distribution can affect elongation, relaxation modulus,
tensile strength, and tenacity of the materials [65]. An additional average molecular weight that
is obtained with the aid of ultracentrifugation is referred to as
sedimentation average molecular weight, or Z-average molecular
weight, M z This
was more often used in the past, particularly with naturally
occurring polymeric materials. It is not used, however, as much
today.
In solutions of polymers the viscosities are more
affected by the long chains than by the short ones. A correlation
of the viscosity of the solution to the size of the chains or to
molecular weight of the solute, allows an expression of a
viscosity average
molecular weight:
where, β is a constant.,
usually less than unity. When β = 1, then M η becomes equal to
M w. Actually,
though, the values of M
η are usually
within 20% of the weight average molecular weights. For
polydisperse polymers M
w is larger than M η and M η in turn is larger than
M n.
Most of the methods for determining the molecular
weights and sizes of polymers (with the exception of small-angle
neutron scattering) require dissolving the polymers in proper
solvents first. The measurements are then carried out on dilute
solutions.
Solution viscosities of linear polymers relate
empirically to their molecular weights. This is used in various
ways in industry to designate the size of polymers. The values are
obtained by measuring the efflux time t of a polymer solution through a
capillary. It is then related to the efflux time t o of pure solvent. Typical
viscometers, like those designed by Ubbelohde, Cannon–Fenske,
(shown in Fig. 1.3), and other similar ones are
utilized and measurements are carried out in constant temperature
baths. The viscosity is expressed in following ways:
Common name
|
Symbol
|
Definitions
|
---|---|---|
1. Relative viscosity
|
η
rel
|
η/η o = t/t o
|
2. Specific viscosity
|
η
sp
|
(η − η o)/η o = η rel − 1 ∝ (t − t o)/t o
|
3. Reduced viscosity
|
η
red
|
|
4. Inherent viscosity
|
η
i
|
ln η rel/C
|
5. Intrinsic viscosity
|
[η]c→0
|
(η
sp/C)
c = 0 = (η i) c = 0
|
2.7.2 Methods for Measuring Molecular Weights of Polymers
To determine the intrinsic viscosity, both
inherent and reduced viscosities are plotted against concentration
(C) on the same graph paper
and extrapolated to zero. If the intercepts coincide then this is
taken as the intrinsic viscosity. If they do not, then the two
intercepts are averaged. The relationship of intrinsic viscosity to
molecular weight is expressed by the Mark–Houwink–Sakurada equation
[66]:
where K and a are constants. Various capillary
viscometers are available commercially. Figure 2.17 illustrates a
typical capillary viscometer.
Fig.
2.17
Cannon–Fenske capillary viscometer
The logarithms of intrinsic viscosities of
fractionated samples are plotted against log M w or log M n. The constants
a and K of the Mark–Houwink–Sakurada equation
are the intercept and the slope, respectively, of that plot. Except
for the lower molecular weight samples, the plots are linear for
linear polymers. Many values of K and a for different linear polymers can be
found in the literature [66].
Actually, the Mark–Houwink–Sakurada equation
applies only to narrow molecular weight distribution polymers. For
low molecular weight polydisperse polymers this equation is useful,
because the deviations due to chain entanglement are still
negligible. On the other hand, chain entanglement in high molecular
weight polydisperse polymers affects viscosity and this equation
does not really apply.
The determinations of molecular weights of
polymers rely, in most cases, upon physical methods. In some
special ones, however, when the molecular weights are fairly low,
chemical techniques can be used. Such determinations are limited to
only those macromolecules that possess only one functional group
that is located at the end of the chain ends. In place of the
functional group, there may be a heteroatom. In that case, an
elemental analysis might be sufficient to determine the molecular
weight. If there is a functional group, however, a reaction of that
group allows calculating the molecular weight. Molecular weights
above 25,000 make chemical approaches impractical. In chemical
determination each molecule contributes equally to the total. This
determination, therefore, yields a number average molecular weight.
With the development of gel permeation chromatography (discussed
below), this method is hardly ever used today.
There are various physical methods available. The
more prominent ones are ebullioscopy, cryoscopy, osmotic pressure
measurements, light scattering, ultracentrifugation, and gel
permeation chromatography (also called size exclusion
chromatography). All these determinations are carried out on
solutions of the polymers. Also, all, except gel permeation
chromatography, require that the results of the measurements be
extrapolated to zero concentrations to fulfill the requirements of
theory. The laws that govern the various relationships in these
determinations apply only to ideal solutions. Only when there is a
complete absence of chain entanglement and no interaction between
solute and solvent is the ideality of such solutions approached. A
brief discussion of some techniques used for molecular weight
determination follows
Ebullioscopy, or boiling point
elevation, as well as cryoscopy, or freezing point
depression, are well-known methods of organic chemistry They are
the same as those used in determining molecular weights of small
molecules. The limitation to using both of these methods to
determine the molecular weight of macromolecules is that
ΔT b and
ΔT f become
increasingly smaller as the molecular sizes increase. The methods
are limited, therefore, to the capabilities of the temperature
sensing devices to detect very small differences in temperature.
This places the upper limits for such determinations to somewhere
between 40,000 and 50,000. The thermodynamic relationships for
these determinations are:
The above two determinations, because each
molecule contributes equally to the properties of the solutions,
yield number average molecular weights. How much this technique is
used today is hard to tell.
A method that can be used for higher molecular
weight polymers is based on osmotic pressure measurements. It can be applied to
polymers that range in molecular weights from 20,000 to 500,000
(some claim 1,000,000 and higher). The method is based on van’t
Hoff’s law. When a pure solvent is placed on one side of a
semi-permeable membrane and a solution on the other, pressure
develops from the pure solvent side. This pressure is due to a
tendency by the liquids to equilibrate the concentrations. It is
inversely proportional to the size of the solute molecules. The
relationship is as follow:
where π is the osmotic
pressure, C is the
concentration, T is
temperature, and R is the
gas constant, A
2 is a measure of interaction between the solvent and
the polymer (second viral coefficient).
A static
capillary osmometer is illustrated in Fig. 2.18. Rather than rely on
the liquid to rise in the capillary on the side of the solution in
response to osmotic pressure, as is done in the static method, a
dynamic equilibrium
method can be used. Here a counter pressure is applied to maintain
equal levels of the liquid in both capillaries and prevents flow of
the solvent. Different types of dynamic membrane osmometers are
available commercially. The principle is illustrated in
Fig. 2.19.
Fig.
2.18
Membrane osmometer
Fig.
2.19
Schematic of a high speed osmometer (from
ref [68])
The results obtained from either method must
still be extrapolated to zero concentration for van’t Hoff’s law to
apply. Such extrapolation is illustrated in Fig. 2.20. Because all
molecules contribute equally to the total pressure, osmotic
pressure measurements yield a number average molecular weight.
Fig.
2.20
Extrapolation to zero concentration
Light
scattering measurement is a technique for determining the
weight average molecular weight. When light passes through a
solvent, a part of the energy of that light is lost due to
absorption, conversion to heat, and scattering. The scattering in
pure liquids is attributable to differences in densities that
result from finite nonhomogenuities in the distribution of
molecules within adjacent areas. Additional scattering results from
a presence of a solute in the liquid. The intensity or amplitude of
that additional scattering depends upon concentration, the size,
and the polarizability of the solute plus some other factors. The
refractive index of pure solvent and a solution is also dependent
upon the amplitude of vibration. The turbidity that arises from
scattering is related to concentration:
where n o is the
refractive index of the solvent, n is the refractive index of the
solution, λ is the
wavelength of the incident light, N o is Avogadro’s number,
and c is the concentration.
The dn/dc relationship is obtained by measuring
the slope of the refractive index as a function of concentration.
It is constant for a given polymer, solvent, and temperature and is
called the specific
refractive increment.
Because scattering varies with different angles
from the main beam of light, the results must be extrapolated to
zero concentration and zero angle of scattering. This is done
simultaneously by a method developed by Zimm. A typical Zimm plot
is illustrated in Fig. 2.21.
Fig.
2.21
A typical Zimm plot
A very popular technique for determining
molecular weights and molecular weight distributions is
ge l permeation chromatography. It is
also called size exclusion chromatography [69, 70]. The
procedure allows one to determine M w and M n, and the molecular
weight distribution in one operation. The procedure resembles HPLC.
It separates molecules according to their hydrodynamic volumes or
their effective sizes in solutions. The separation takes place on
one or more columns packed with small porous particles. As the
solution travels down the columns, there is retention of the
polymer molecules by the pores of the packing. It was postulated in
the past that the separation that takes place by molecular sizes is
due to smaller molecules diffusing into all the pores while the
larger ones only into some of the pores. The largest molecules were
thought to diffuse into none of the pores and pass only through the
interstitial volumes. As a result, polymer molecules of different
sizes travel different distances down the column. This means that
the molecules of the largest size (highest molecular weight) are
eluted first because they fit into the least number of pores. The
smallest molecules, on the other hand, are eluted last because they
enter the greatest number of pores and travel the longest path. The
rest fall in between. The process, however, is more complex than
the above postulated picture. It has not yet been fully explained.
It was found, for instance, that different gels display an almost
identical course in the relation of dependence of V R (retention volume) to
the molecular weight. Yet study of the pores of different gels show
varying cumulative distributions of the inner volumes. This means
that there is no simple function correlating the volume and/or the
size of the separated molecules with the size and distribution of
the pores [69]. Also, the shape of
the pores that can be inferred from the ratio of the area and
volume of the inner pores is very important [70]. Different models were proposed to explain
the separation phenomenon. These were reviewed thoroughly in the
literature. They are beyond the scope of this book.
As indicated above, the volume of the liquid that
corresponds to a solute eluting from the columns is called the
retention volume or elution volume (V R). It is related to the
physical parameters of the column as follows:
where, V o = the
interstitial volume of the column(s)
K = the
distribution coefficient
V
1 = the internal solvent volume inside the pores
The total volume of the columns is V T that is equal to the sum
of V O and
V 1. The
retention volume can then be expressed as follows:
From the earlier statement it should be clear
that polymer fractionation by gel permeation chromatography depends
upon the spaces the polymer molecules occupy in solution. By
measuring, experimentally, the molecular weights of polymer
molecules as they are being eluted one obtains the molecular weight
distribution. To accomplish this, however, one must have a
chromatograph equipped with dual detectors. One must detect the
presence of polymer molecules in the effluent. The other one must
measure their molecular weights. Such detectors might be, for
instance, a refractive index monitor and a low angle laser light
scattering photogoniometer to find the absolute value of
M.
Many molecular weight measurements, however, are
done on chromatographs equipped with only one detector that
monitors the presence of the solute in the effluent. The equipment
must, therefore, be calibrated prior to use. The relationship of
the ordinate of the chromatogram, commonly represented by
F(V), must be related to the molecular
weight. This relationship varies with the polymer type and
structure. There are three methods for calibrating the
chromatograph. The first, and most popular one, makes use of narrow
molecular weight distribution reference standards. The second one
is based upon a polydisperse reference material. The third one
assumes that the separation is determined by molecular size. All
three methods require that an experimentally established
calibration curve of the relationship between the molecular size of
the polymer in solution and the molecular weight be developed. A
chromatogram is obtained first from every standard sample. A plot
is then prepared from the logarithms of the average weights against
the peak retention volumes (V R). The values of
V R are measured
from the points of injection to the appearances of the maximum
values of the chromatograms. Above M 1 and below M 4 there is no effective
fractionation because of total exclusion in the first place and
total permeation in the second case. These are the limits of
separation by the packing material.
To date the standard samples of narrow molecular
weight distribution polymers that are available commercially are
mainly polystyrenes. These samples have polydispersity indexes that
are close to unity and are available over a wide range of molecular
weights. For determining molecular weights of polymers other than
polystyrene, however, the molecular weights obtained from these
samples would be only approximations. Sometimes they could be in
error. To overcome this difficulty a universal calibration method is used. The basis for
universal calibration is the observation [51] that the multiplication products of
intrinsic viscosities and molecular weights are independent of the
polymer types. Thus, [η]M is the universal calibration parameter. As a result, a plot of
log ([η]M) vs. elution volume yields a curve
that is applicable for many polymers. The log ([η]M) for a given column (or columns), temperature, and elution volume is assumed to be a constant
for all polymers. This is illustrated in Fig. 2.22.
Fig.
2.22
Molecular weight calibration curve for gel
permeation chromatography
Numerous materials have been used for packing the
columns. Semi rigid crosslinked polystyrene beads are available
commercially. They are used quite frequently. Porous beads of glass
or silica are also available. In addition, commercial gel
permeation equipment is usually provided with automatic sample
injection and fraction collection features. The favorite detectors
are refractive index and ultraviolet light spectroscopic detectors.
Some infrared spectroscopic detectors are also in use. Commercially
available instruments also contain pumps for high-pressure rapid
flow and are usually equipped with a microcomputer to assist in
data treatment. Also, they come with a plotter in the equipment to
plot the detector response as the samples are eluted through the
column or columns. A typical chromatogram is illustrated in
Fig. 2.23 and a schematic for the basic equipment is
shown in Fig. 2.24. When polydisperse samples are analyzed,
quantitative procedures consist of digitized chromatograms with
indication of equally spaced retention volumes. These can be every
2.5 or 5.0 mL of volumes. The resultant artificial fractions
are characterized by their heights h i , their solute concentrations
C i , and by the area they occupy
within the curve A
i . The
cumulative polymer weight values is calculated according to:
After conversion of the retention volumes
V i into molecular weights (using
the calibration curve), the molecular weights, M w, M n, and M z can be calculated:
If the chromatogram is not equipped with a
microcomputer for data treatment, one can easily determine results
on any available PC. Programs for data treatment have been written
in various computer languages. They are available from many
sources.
Fig.
2.23
A typical digitized gel permeation
chromatogram
Fig.
2.24
Schematic illustration of gel permeation
equipment (the illustration only shows one sample column. Several
sample columns are often used) (from ref [68])
Recently, there were several reports in the
literature on combining size exclusion with high pressure liquid
chromatography for more comprehensive characterization of polymers.
Thus, Gray et al. reported that a combination of high pressure
liquid chromatography with size exclusion chromatography allows
comprehensive structural characterization of macromolecules
[71].
On the other hand, Chang et al. reported on using
a modified form of high pressure liquid chromatography analysis,
referred to as interaction
chromatography for polymer characterization. The process
utilizes enthalpic interaction of polymeric solutes with the
stationary phase. Such interaction depends on both, the chemical
composition and on molecular weight. It is claimed to be less
sensitive to chain architecture and to offer superior resolution to
SEC. The typical HPLC instrument is modified to precisely control
the temperature of the column. The temperature of the column and
the mobile phase is controlled by circulating a fluid through the
column jacket from a programmable bath/circulator. The mobile and
stationary phases require careful choices to adjust the interaction
strength of the polymer solutes with the stationary phase so that
the polymer solutes elute out in a reasonable elution time. The
process depends upon variations of the column temperature for
precise control of the solute retention in the isocratic elution
mode. Mixed solvent system of a polar and a less polar solvent are
often employed to adjust the interaction strength [72].
2.8 Optical Activity in Polymers
Optical activity in biopolymers has been known
and studied well before this phenomenon was observed in synthetic
polymers. Homopolymerization of vinyl monomers does not result in
structures with asymmetric centers (The role of the end groups is
generally negligible). Polymers can be formed and will exhibit
optical activity, however, that will contain centers of asymmetry
in the backbones [73]. This can be
a result of optical activity in the monomers. This activity becomes
incorporated into the polymer backbone in the process of chain
growth. It can also be a result of polymerization that involves
asymmetric induction [74,
75]. These processes in polymer
formation are explained in subsequent chapters. An example of
inclusion of an optically active monomer into the polymer chain is
the polymerization of optically active propylene oxide. (See
Chap.
5 for additional discussion). The process of
chain growth is such that the monomer addition is sterically
controlled by the asymmetric portion of the monomer. Several
factors appear important in order to produce measurable optical
activity in copolymers [76]. These
are: (1) Selection of comonomer must be such that the induced
asymmetric center in the polymer backbone remains a center of
asymmetry. (2) The four substituents on the originally inducing
center on the center portion must differ considerably in size. (3)
The location of the inducing center must be close to the polymer
backbone. (4) The polymerization reaction must be conducted at
sufficiently low temperature to insure stereo chemical selectivity.
An example is a copolymerization of maleic anhydride with optically
active l-α-methylbenzyl vinyl ether. The copolymer exhibits optical
activity after the removal of the original center of asymmetry
[77].
An example of an asymmetric induction from
optically inactive monomers is an anionic polymerization of esters
of butadiene carboxylic acids with (R)-2-methylbutyllithium or with
butyllithium complexed with (−)methyl ethyl ether as the catalyst.
(This type of polymerization reaction is described in Chap.
4) The products, tritactic polymers exhibit
small, but measurable optical rotations [78]. Also, when benzofuran, that exhibits no
optical activity, is polymerized by cationic catalysts like
aluminum chloride complexed with an optically active co catalyst,
like phenylalanine, an optically active polymer is obtained
[77].
By contrast, an example of formation of
enantioselective polymer from achiral monomers, where the chirality
is inherent in the main chain is polymerization of 1,5-hexadiene
with an optically active catalyst [77]. The catalyst precursors are (R,R) or (S,S)-[ethylene-1,2-bis(η5-4,5,6,7-tetrahydro-1-indenyl)zirconium
(1,1′-by-1-naphtholate). The product is an optically active version
of poly(methylene-1,3-cyclopentane):
This polymer is highly isotactic and contains 72%
trans rings.
It was also observed that conjugated polymers
that are also electrical conductors (see Chap.
10) exhibit optical activity that depends
critically on their structural organization [78]. Thus, strong chiroptical properties can be
obtained from substituted polythiophene [79] (Chap.
10) with optically active side chains, especially
when the monomers are coupled within the polymer in a regioregular
head-to-tail fashion. Actually, optical activity of these materials
is only found when the polymers are aggregated at low temperature,
in poor solvent, or in solution cast films. This contrasts with
other optically active polymers, like polypeptides, poly(l-alkynes)
and polyisocyanates that show an optically active conformation of
the main chain in the absence of supramolecular association.
In addition, it was shown that a repetitive
inversion of optical activity in films can be obtained by warming
and cooling cycles, where the cooling rate determines the
handedness of the associates [78].
A similar result concerning inversion of chirality has been found
in solution, depending on the composition of a binary solvent
mixture.
2.9 Review Questions
2.9.1 Section 2.1
1.
What are the secondary bond forces that influence
the physical properties of macromolecules?
2.9.1.1 Section 2.1.1
1.
Explain and illustrate dipole–dipole interactions
in polymers and how do they affect the properties of polymeric
materials. Can you give other examples?
2.9.1.2 Section 2.1.2
1.
What are the induction forces?
2.
Explain what type of chemical structures and
chemical bonds in the backbones of the polymeric chains stiffen
them and what type flexibilizes (or “softens”) them.
3.
How do the pendant groups affect the melting
points of polymers?
4.
Does copolymerization raise or lower the melting
point of a polymer?
5.
Which polymer, A or B, would have a higher
melting point? Explain
6.
Explain what the induction forces of polymers and
how they affect the physical properties of these materials?
7.
Illustrate and explain why nylon 11 melts
approximately 50°C higher than linear polyethylene
8.
Why does poly(ethylene adipate) melt at a
considerably lower temperature than does poly(ethylene
terephthalate)?
9.
What type of bonds in the polymeric chain tend to
soften them?
10.
Does copolymerization lower or raise the melting
temperature of polymers? Explain
2.9.2 Section 2.2
1.
Describe an amorphous state of polymers
2.9.2.1 Section 2.2.1
1.
Describe the glassy state of polymers
2.
What is the second order transition temperature?
Are the send order transition temperatures of polymers absolute
values or do they vary depending upon various conditions?
Explain.
3.
How can the transition to a glassy state be
observed?
4.
Explain structural recovery and the asymmetry of
structural recovery.
5.
What equation describes the asymmetry of
structural recovery?
6.
What is the Flory–Fox equation and what is it
limited to?
2.9.2.2 Section 2.2.2
1.
Explain the phenomenon of elasticity.
2.
What natural structural arrangement of polymeric
chains leads to rubber elasticity.
3.
What is meant by the negative coefficient of
expansion?
2.9.2.2.1 Section 2.2.2.1
1.
Write the equation for the reactive force of an
elastomer and explain what each term represents and define an ideal
elastomer.
2.
What is the entropy of elasticity? What is it
proportional to?
3.
What is the relationship between the retractive
force of an elastomer and its temperature?
4.
Write the equation for the work done in
stretching a chain of an ideal elastomer and explain what each term
represents.
5.
Write the equation for a free energy change of a
stretched elastomer.
6.
Write the equation for the average energy per
(stretched) chain
7.
What is the free energy change for a stretched
elastomer? Explain.
2.9.2.2.2 Section 2.2.2.2
1.
Define a Newtonian liquid and a Hookian ideal
elastic solid, and an ideal elastomer.
2.
What is meant by viscoelasticity?
3.
Explain what is meant by a yield point and a
Bingham Newtonian fluid.
4.
How is the viscosity of a molten polymer related
to its temperature?
5.
How is shear stress defined mathematically?
6.
What is a thixotropic liquid?
7.
What is the relationship Newtonian viscosity of
amorphous polymers and their chain length?
8.
What is shear creep compliance equal to ?
9.
Why are shear thinning and thixotropy two
different phenomena? Explain
10.
What relationship describes flow behavior of
liquids that deviate from Newtonian flow?
11.
Write the equation for shear-stress-relaxation
modulus for viscous fluids. Explain
12.
What are the two important quantities in behavior
of polymeric liquids?
13.
Describe two techniques for measuring the
viscosity of molten polymers.
2.9.3 Section 2.3
1.
How does crystallinity of polymers that are
formed from the melt differ from that of small molecules?
2.
What are the two ways that crystal growth can
take place in polymeric materials?
3.
What type of polymers tend to crystallize?
4.
What is the typical size of polymeric crystals
formed from the melt?
5.
What is a fringed micelle or a fringed
crystallite model?
6.
What is the folded chain lamella? An
adjacent-reentry model? A switchboard or a nonadjacent-reentry
model?
7.
What is drawn fibrilar morphology? What is
vitrification?
8.
What are spherulites? How are they
observed?
9.
Explain the difference between a crystallite
formed from the melt and a polymer crystal formed from a dilute
solution
2.9.3.1 Section 2.3.1
1.
What are the three different interfacial free
energies that are characteristics of crystals?
2.
What equation described the dependence of the
melting temperature of a polymer crystal upon its chain length?
Explain and write the equation.
3.
What is the first order transition
temperature?
2.9.3.2 Section 2.3.2
1.
How can one obtain information on the kinetics of
crystallization
2.
What is the Avrami equation and what are its
drawbacks?
2.9.4 Section 2.4
1.
What are mesophases ?
2.
What is anisotropic behavior?
3.
What are lyotropic and thermotropic liquid
crystals?
4.
Explain nematic, smectic, and cholesteric liquid
crystal arrangements.
5.
Which of the two polymers would you expect to
exhibit liquid crystalline behavior? Explain.
6.
What is the difference between the main chain and
side chain liquid crystals?
2.9.5 Section 2.5
1.
What happens to the arrangement of the polymeric
chains in films and fibers upon application of an external
stress?
2.
What is orientation and how does that benefit the
properties of polymeric materials?
2.9.6 Section 2.6
1.
What are the two stages of dissolution when a
polymer dissolves in a solvent?
2.
What is the heat of mixing of binary systems?
Write equation.
3.
What is the radius of gyration? Write
equation
4.
Define unperturbed dimensions and expansion
factor
2.9.6.1 Section 2.6.1
1.
What is the change in entropy of mixing,
according to the Flory–Huggins theory? Explain and write the
equation.
2.
What is the change in the free energy of mixing
according to the Flory–Huggings theory? Write equation.
3.
What is the Flory temperature? write equation and
explain what happens to a solution of a polymer below this
temperature
2.9.7 Section 2.7
1.
Define the degree of polymerization
2.
What is the DP of polystyrene with molecular
weight of 104,000 and poly(vinyl chloride) with molecular weight of
63,000?
3.
What are the important features of chain-growth
and step-growth polymerizations? Can you explain the difference
between the two? Can you suggest an analytical procedure to
determine by what mechanism a particular polymerization reaction
takes place?
4.
What is the DP of polystyrene with molecular
weight of 104,000 and poly(vinyl chloride) with molecular weight of
630,000?
5.
Explain the differences between thermosetting and
thermoplastic polymers and define gel point.
6.
Give the definitions of oligomer, telomer, and
telechelic polymers
7.
Why must statistical averages be used to express
molecular weights of polymers?
8.
What is number average molecular weight? What is
the equation for number average molecular weight?
9.
What is a weight average molecular weight? What
is the equation for the weight average molecular weight?
10.
In a mixture of two kinds of molecules there are
ten of each in kind. The molecular weight of molecules A is 10,000
and the molecular weight of molecules B is 100,000. What is the
number average molecular weight of the mixture and the weight
average molecular weight?
11.
What is the viscosity average molecular weight
and how does it differ from the weight average molecular
weight?
12.
What is a molecular weight distribution? What is
a monodisperse polymer and a polydisperse polymer?
13.
What is the Mark–Houwink–Sakurada equation? Can
you suggest a way to determine K and a constants experimentally for a given
polymer?
14.
Give the definitions and formulas for the
relative viscosity, specific viscosity, reduced viscosity, inherent
viscosity, and intrinsic viscosity.
15.
Why is it necessary to extrapolate to zero
(explain how this is done) in order to obtain intrinsic
viscosity.
16.
Discuss the various methods of molecular weight
determination explain why a particular method yields a number of a
weight average molecular weight, or, as in case of GPC, both.
2.9.8 Section 2.8
1.
Discuss optical activity in polymers.
2.9.9 Recommended Reading
-
H. Sperling, Introduction to Physical Polymer Science, 2nd ed., Wiley, New York, 2006
References
1.
S. Mizushima T. Shimanouchi,
J. Am. Chem. Soc., 1964, 85, 3521; L.H. Sperling, Introduction to Physical Polymer
Science, 3rd. Ed., J. Wiley, New York, 2006; G. Strobel, The Physics of Polymers, II edition,
Springer, Berlin, 1997
2.
J. W. le Noble, Highlights of Organic Chemistry,
Dekker, New York, N.Y., 1974
3.
M. Avrami, J. Chem. Phys., 1939, 7, 103; ibid. 1940 8, 212
4.
F. P. Redding, J. Polymer Sci., 1956 21, 547
5.
L. Mandelkern, Am. Chem. Soc. Polymer Preprints,
1979, 20 (1), 267; D.C. Basset, Principles of Polymer Morphology,
Cambridge University Press, Cambridge, 1981; A. Sharples, “Crystallinity”,
Chapt. 4 in Polymer
Science, Vol.I, A. Jenkins, ed., North Holland, Amsterdam,
1972; L. Mandelkern. “The
Crystalline State”, Chapter 4 in J.Mark, K.Ngal, W. Graessley, L.
Mandelkern, E. Samulsky, J. Koenig, and G. Wignall, Physical Properties of Polymers, 3rd
ed.,Cambridge, 2004
6.
J. Brandrup and E.H.
Immergut., eds.,Polymer
Handbook, Wiley and Sons, 2nd and 3rd Ed.; E.A. Turi,
ed.,Thermal Characterization of
Polymeric Materials,,Academic Press, New York, 1981; W.W. Wedlandt, Thermal Analysis, 3-rd ed., Wiley, New
York, 1986; T. Nakaoki, R.
Kitamaru, R.G, Alamo, W.T. Huang, and L. Mandelkern, Poymer J., 2000, 32, 876
7.
F. W. Billmeyer, Jr.,
Textbook of Polymer Science Wiley-Interscience,
3rd ed., New York, 1984
8.
L. Marker, R. Early, and S.
L. Agrarwall, J. Polymer
Sci., 1959, 38, 369
9.
T. W. Campbell and A. C.
Haven, Jr., J. Appl
. Polymer Sci ., 1959, 1, 73
10.
G. Natta, F. Danusso, and G.
Moraglis, J. Polymer
Sci., 1957, 25, 119
11.
F. P. Redding, J. Polymer Sci ., 1956, 21, 547
12.
A.V. Tobolsky, Properties and Structure of Polymers,
J. Wiley, New York, 1960
13.
G.I. Wilks, J. Chem Ed., 1981, 58, 880
14.
D.J. Williams, Polymer Science and Engineering,
Prentice-Hall, Englewood Cliffs, N.J., (1971)
15.
M. G. Zachman, Angew . Chem ., Intern . Edit ., 1974, 13, (4), 244
16.
F. Bueche, Physical Properties of Polymers,
Interscience, New York, 1962
17.
D.J. Plazec and K.L. Ngai,
in Physical Properties of Polymers, Handbook, J.E. Mark, ed., AIP
Press, Woodbury, New York, 1996; also see. K.L.Ngai, Chapter 2 in
J.Mark, K.Negai, W. Grassley. L. Mandelkern, E. Samulski. J.
Koenig, and G. Wignall, Physical
Properties of Polymers, 3rd ed., Cambridge, 2004
18.
T.G.Fox, Jr. and P.J. Floty,
J. Appl. Phys.,
1964, 21, 581
19.
Fang Liu, William L. Jarrett
and Marek W. Urban, Shelby F. Thames, Macromolecules, 2010, 43, 5330
20.
R. D. Deanin, Polymer Structure, Properties and
Application, Cahmers Publishing Co., Inc., Boston, Mass.,
1972; S.-Q. Wang, S.
Ravindranath, and P.E. Boukany, Macromolecules, 2011, 44, 183
21.
P. J. Flory, Principles of Polymer Chemistry,
Cornell Univ. Press, Ithaca, N.Y., 1951
22.
L.R.G. Treloar, The Physics of Rubber Elasticity,
3rd ed., Clarendon Press, oxford, 1975; A.N.Gent, Rubber Elasticity: Basic Concepts and
Behavior, Chapter 1 in F.R. Eirich, (ed.) Science and Technology of Rubber,
Academic Press, New York, 1978
23.
T. Koga and F. Tanaka,
Macromolecules,
2010, 43, 3052.
24.
W.W. Graessley, Chapter 3 in
J.Mark, K.Negai, W. Grassley. L. Mandelkern, E. Samulski. J.
Koenig, and G.Wignall, Physical
Properties of Polymers, 3rd ed., Cambridge, 2004
25.
J. Mark, Chapter 1 in
J.Mark, K.Negai, W. Grassley. L. Mandelkern, E. Samulski. J.
Koenig, and G. Wignall, Physical
Properties of Polymers, 3rd ed., Cambridge, 2004
26.
H.F. Mark, J. Polymer Sci ., 1965, C9, 1 A. Peterlin, J. Polymer Sci ., 1965, C9, 61
27.
E. P. Chan, K.A. Page, and
C.A. Stafford, Am. Chem. Soc.
Polymer Preprints, 2010, 51
(1),109
28.
Huang, R. J. Mech. Phys. Solids 2005, 53, 63–89
29.
J. R. Collier, Ind . and
Eng . Chem., 1969, 61, (10), 72
30.
G.Natta and P.Carradini,
J. Polymer Sci,
1959, 39,29 (1959)
31.
P.H. Geil, Polymer Single Crystals, Wiley-Interscience, New York,
1963
32.
P.H. Geil, N.K.J. Symons,
and R.G. Scott, J. Appl
. Phys ., 30, 1516
33.
R.H. Geil, J. Polymer Sci ., 1960, 44, 449
34.
B.H. Ranby, F.F. Morehead,
and N.M. Walter, J. Polymer
Sci ., 1960, 44, 349
35.
A. Keller, “Growth and
Perfection of Crystals,” Wiley, New York, 1958: A. Keller and A.
O’Connor, Polymer,
1960, 1, 163
36.
W.M.D, Bryant, J. Polymer Sci., 1947, 2, 547
37.
R.St.J. Mauley, Nature, 1961, 189, 390; S. Hosoda, Y. Nozue,
Y.Kawashima, K. Suita, S. Seno, and T.Nagamatsu, Macromolecules, 2011, 44, 313
38.
M. Farina, M. Peraldo, and
G.Natta, Angew . Chem ., Intern . Ed ., 1965, 4, 107
39.
D.C. Basselt, F.C. Frank and
A. Keller, Nature,
1959, 184, 810; W.D. Niegisch and P.R. Swan,
J. Appl Phys ., 1960, 31, 1906; D.H. Reneker and P. H. Geil,
J. Appl Phys ., 1960, 31, 1916
40.
M. Goodman, A. Abe and Y,L.
Fan, Macromol. Rev.,
1967, 1, 8
41.
N.G. Gaylord and H.Mark,
Linear and Stereoregular Addition
Polymers, Interscience, New York, 1959
42.
L. Mandelkern, Chapter 4 in
J.Mark, K.Negai, W. Grassley. L. Mandelkern, E. Samulski. J.
Koenig, and G. Wignall, Physical
Properties of Polymers, 3rd ed., Cambridge, 2004
43.
Z. Wang, Y. Li, J. Yang, Q.
Gou, Y. Wu, X. Wu, P. Liu, and Q. Gu, Macromolecules, 2010, 43, 4441
44.
E.B. Sirota, Macromolecules, 2007, 40, 1045
45.
W.Z. Oswald, Z. Phys. Chem., 1897, 22, 286
46.
G. Stroble, Eur. Phys, J., 2000, 3, 165
47.
G. Stroble, Eur. Phys, J. 2005, 18, 295
48.
G. Stroble, Prog. Polym. Sci., 2006, 31, 398
49.
B. Heck,T. Hugel, M. Iijima,
and G. Strobl, Polymer,
2000, 41, 8839
50.
K. Hayashida, A. Takano, and
Y. Matsushita, Phys. Rev.
Lett., 2007,
98, 195502
51.
S. Chandrasekhar,
Liquid Crystals, Cambridge
University Press, London, 1977
52.
A. Ciferri, W.R. Krigbaum,
and R.B. Meyers, eds., Polymer
Liquid Crystals, Academic Press, New York, 1982
53.
Tokita, M.; Kim, K.W.; Kang,
S.; Watanabe, J. Macromolecules 2006, 39, 2021; Tokita, M.; Funaoka, S,,
Watanabe, J. Macromolecules, 2004, 37, 9916
54.
Keith, H. O. and Chen, W.
Y., Polymer, 2002, 43, 6263- 6272.; Lots, B.and Cheng, S.
2. D. Polymer, 2005, 46, 577- 610
55.
Y. Gnanou and M. Fontanille,
Organic and Physical Chemistry of
Polymers, Wiley,New York, 2008
56.
N.M. Abukhdeir and A.D. Rey,
Macromolecules,
2009, 42, 3841
57.
E.T. Samulski, Chapter 5 in
J.Mark, K.Negai, W. Grassley. L. Mandelkern, E. Samulski. J.
Koenig, and G. Wignall, “Physical Properties of Polymers,” 3rd ed.,
Cambridge, 2004
58.
T.-H. Tong, A. Primak, S.
Kumar, and L.-C. Chien, Macromolecules 1998, 31, 3537
59.
R. Deschenaux, F. Turpin,
and D. Guillon, Macromolecules, 1997, 30, 3759
60.
A. Sanchez-Ferrer and H.
Finkelmann, Macromolecules,
2008, Web
article10.1021/ma7025644S0024- 9207(70)0264-0
61.
S. Rendon, W.R. Burghardt,
M.L. Auad, and J.A. Kornfield, Macromolecules, 2007, ASAP Article 101021/ma062912c; Web
Release Date August 14, 2007
62.
Ahn, Deshmukh, and Kosi,
Am. Chem. Soc. Polymer
Preprints, 2010,
51, (2) 430
63.
M.L. Huggins, Physical Chemistry of High Polymers,
Wiley, New York, 1958
64.
N.C. Billingham,
Molar Mass Measurements in Polymer
Science, Halsted Press, New York, 1977. J.F. Rabeck, Experimental methods in Polymer
Chemistry, Wiley, New York. 1980
65.
J. R. Collier, Ind . and
Eng . Chem ., 1969, 61, (10), 72
66.
J. Brandrup and E.H.
Immergut., eds.,Polymer
Handbook, Wiley and Sons, 2nd and 3rd Ed.; E.A. Turi, ed.,
“Thermal Characterization
of Polymeric Materials”, Academic Press, New York,
1981;
67.
F. P. Redding, J. Polymer Sci ., 21, 547 (1956)
68.
A. Ravve, Organic Chemistry of Macromolecules,
Delkker, New York, 1967
69.
J. Janca, ed., “Steric Exclusion Liquid Chromatography of Polymers”, Dekker, New York,
1984
70.
T. Provder, ed.,
“Detection and Data Analysis in Size Exclusion Chromatography”, Am. Chem.,
Soc.,Washington, 1987
71.
M. Gray, C. Silverman, and
P. Kile, Am. Chem. Soc. Polymer
Preprints, 2010,
51(2), 162
72.
T. Chang, S. Ahn, and H.
Lee, Am. Chem. Soc. Polymer
Preprints, 2010,
51 (2), 160.
73.
N. Beredjick and C.
Schuerch, J. Am
. Chem . Soc ., 1956,78, 2646; ibid ., 1958, 80, 1933
74.
G. Natta, M. Farina, M.
Peraldo, and G. Bressan, Makromol . Chem ., 1961, 43, 68
75.
M. Farina and G. Bressan,
Makromol . Chem ., 1963, 61, 79
76.
P. Pino, F. Ciardelli, and
G. P. Lorenzi, J. Am
. Chem . Soc ., 1963, 85, 3883
77.
P. Pino, F. Ciardelli, and
G. P. Lorenzi, Makromol
. Chem ., 1964, 70, 182
78.
M. Goodman, A. Abe and Y,L.
Fan, Macromol. Rev.,
1967, 1, 8
79.
B. Goderis, H. Reynaers, H.
Scharrengerg, V. Mathot, and M. Koch, Macromolecules, 2001, 34, 1770