Sei sulla pagina 1di 13

Chapter 11

Oriented polymers II – models and properties


11.1 Introduction
In this chapter an attempt to answer the following questions is made.

(i) What kind of theoretical models, or deformation schemes,canbe envisaged for


predicting the distribution of molecular orientations in a drawn polymer?

(ii) (ii) How do the predictions of these models agree with experimental results for the
orientation averages such as ?

(iii)(iii) Howcan experimental or theoretical values of the orientation averages be used to


predict the properties of the drawn polymer and with how much success?

These questions form the topics of discussion for sections 11.2, 11.3 and 11.4, respectively.
In extremely highly drawn material, a great deal of modification of the original structure of
the polymer must have taken place and most of the chains are essentially parallel to the draw
direction, so that dan are both close to 1. It is difficult to describe
theoretically and in detail how molecular orientation takes place under these circumstances
and thus to predict the small departures from unity for such materials. Discussion of the
prediction and use of orientation averages is therefore limited to materials of moderate draw
ratios.

Models used for interpreting the properties of highly oriented polymers are discussed
in sections 11.5 and 11.6. The deformation schemes to be described in the following section
are expected to be valid for single-phase, i.e. non-crystalline, polymers or for one of the two
phases present in semi-crystalline polymers. In such two- phase systems the prediction of the
properties involves combining those of the two oriented phases in the appropriate way. The
models considered in section 11.5 show how this can be done.

11.2 Models for molecular orientation

Two principal deformation schemes have been used in attempts to predict the distribution of
orientations:

(i) the affine rubber deformation scheme, or model; and

(ii) the aggregate or pseudo- affine deformation scheme or model.

11.2.1 The affine rubber deformation scheme

In this scheme it is assumed that, at the temperature of drawing or other deformation, the
polymer is in the rubbery state, either as a true rubber containing chemical cross-links or as a
polymer containing tangled amorphous chains, in which the entanglements can behave like
cross-link points.

Once deformation has been accomplished, the polymer is cooled down into the glassy state
before the stress is removed, so that the orientation is ‘frozen in’.

The simplest version of the rubber model makes the assumption of an affine deformation:
when the polymer is stretched the cross-link points move exactly as they would if they were
points in a completely homogeneous medium deformed to the same macroscopic deformation
(see section 6.4.4, fig. 6.12). The following additional assumptions are also made in the
simplest form of the theory.

(i) Each chain (i.e. the portion of a polymer molecule between two successive cross-link
points) consists of the same number n of freely jointed random links, each of
length l.

(ii) The length of the end-to-end vector ro of each chain in the unstrained state is equal to
the RMS vector length of the free chain, which is shown in section 3.3.3 to be n 1/2
l.

(iii)There is no change in volume on deformation, so that,

where the quantities are the extension ratios defined in section

6.3.2. (It is useful to remember that this really means that fractional changes in volume are
not actually zero but are very small compared with fractional changes in linear dimensions.)

For simplicity, only uniaxially oriented samples are considered here, which

means that and hence only , which will be called , the

draw ratio, need be specified. When the polymer is drawn, two effects contribute to the
orientation of the random links:

(i) the end-to-end vector of each chain orients towards the draw direction; and

(ii) the random links of each chain orient towards the end-to-end vector of the chain and
therefore towards the draw direction, as illustrated in fig. 11.1.

The calculation of , the distribution of orientations of the random links with respect to
the draw direction, therefore involves three steps (see fig. 11.1):

(i) the calculation of , the distribution function for the end-to-end vectors with
respect to the draw direction after drawing

(ii) the calculation of , the distribution function for the links of a chain of
stretched length r with respect to the end-to-end vector r, where r depends on ; ;and
(iii) the combining of these two functions to give .

It is a straightforward piece of geometry (see problem 11.1) to show that

The calculation of the distribution function for the random links with respect to the
end-to-end vector of the chain of stretched length r is somewhat more difficult. Provided that
n is sufficiently
large the usual statistical methods can be used, by assuming that the distribution that actually
exists is the most probable one, i.e. the one that can be obtained in the largest number of
different ways subject to the constraints that, if there are nL links inclined at L to the end-to-
end vector r, then:

This is a familiar type of problem in statistical mechanics. The solution is

Where satisfies the equation

is said to be the inverse Langevin function of r= nl . It is important to note that the


form of depends on r, which in turn depends on .

It is very difficult to combine equations (11.1) and (11.3) give ,so it is usual to
calculate only the values of and possibly in the following stages.

(a) Find the values of for for a given r, i.e. a given .

(b) Use

to calculate , the average value of for all those random links which
belong to chains of length r, for which the end-to-end vectors lie on a cone of angle .
(c) Calculate for all random links of all chains by averaging over all values of r or,
equivalently, over all values of , using the expression for from equation (11.1).

The result will be quoted only for the second-order average,

which is given by
Using only the first term is essentially equivalent to the Gaussian approximation of rubber
elasticity theory developed in section 6.4.4.

The form of versus for the rubber model is then as shown in fig. 11.2. The
relationship between and is similar but it takes much smaller values for a given
value of .

An important consequence of the affine assumption is that the rubber model in this
simple form is applicable only up to , because at this draw ratio the chains parallel
to the draw direction in the undeformed material are fully extended in the deformed material
and cannot extend further (see equation (3.5) and the sentence following it). Further
extension of the material as a whole could thus take place only non-affinely. Equation (11.6)
shows that so that the maximum value of

which the simple theory applies is approximately 0.2, as fig. 11.2 shows.
11.2.2 The aggregate or pseudo-affine deformation scheme
The basic assumptions of this scheme are the following.

(i) The polymer behaves like an aggregate of anisotropic units.

(ii) The properties of the units are those of the completely oriented polymer.

(iii)In the simplest treatment, which applies to the simplest type of uni- axially oriented
samples, each unit is also assumed to be transversely isotropic with respect to an axis
within it, often called its ‘unique axis’.

(iv) In the isotropic medium, the unique axes of the units are distributed randomly
in space.

(v) On drawing, each unit rotates in exactly the same way as would the end-to-end vector
of a chain in the affine rubber model if it were originally parallel to the axis of the
unit considered.

(vi) The units do not change in length or properties on drawing, unlike the chains
in a rubber.

The last assumption is the pseudo-affine assumption; it should be seen immediately that this
makes the model somewhat unphysical. It is there- fore sometimes called the needles in
plasticine model, in which each unit corresponds to a needle and the plasticine corresponds to
the surrounding polymer. In spite of this unreality, the model often predicts well the orien-
tation distribution of the crystallites for a drawn semicrystalline polymer.
The distribution of orientations of the unique axes of the units for this scheme is given simply
by equation (11.1) with instead of

and the corresponding expression for becomes

When is calculated from equation (11.8) it is found to depend on l as shown in fig.


11.3. The variation of with is quite similar except that it is slightly concave
upwards below and has a somewhat lower value than for a given
l,as shown in fig. 11.3. Unlike the curves for the affine rubber deformation scheme, which are
different for different values of n, these curves have no free parameters, so that they are the
same for all polymers. The shapes contrast strongly with those for the rubber model, being
concave to the abscissa, whereas the latter are convex. The curves for the affine rubber

model from fig. 11.2 are included in fig. 11.3 for comparison, from which it is clear that the
values of for a given are always very much smaller for the affine model than they
are for the pseudo-affine model.
11.3 Comparison between theory and experiment

11.3.1 Introduction

The simplest tests that can be made of the predictions of the deformation schemes considered
in the previous two sections are comparisons between the values of predicted by
the schemes and values obtained from refractive-index or birefringence measurements. In the
next two sections comparisons of this kind are made. It is clear from the proof given in the
previous chapter that the reason why the birefringence depends on but not on any
higher-order averages is that the refractive index depends on a second-rank tensor, the
polarisability tensor. Equations (A.7) show that the rotation equations for a second-rank
tensor involve products of two direction cosines and, even for general biaxial orientation,
these products can be expressed in terms of second-order generalized spherical harmonics, of
which the Legendre polynomial is the simplest and the only one required for the
simplest uniaxial orientation.

Equations (A.7) also show that, in general, the prediction of a property that depends on a
tensor of rank l will require knowledge of orientation averages of order l. The elastic
constants of a material are fourth-rank tensor properties; thus the prediction of their values for
a drawn polymer involves the use of both second- and fourth-order averages, in the simplest
case and , and thus provides a more severe test of the models for the
development of orientation. The elastic constants are considered in section 11.4.

11.3.2 The affine rubber model and ‘frozen-in’ orientation

When a polymer is spun to form a fibre the molecular network in the melt is extended and
subsequently ‘frozen in’ the extended state. In commercial fibre production, heat setting or
subsequent hot drawing above the glass-transition temperature may also be used, but for
fibres that have only been spun and drawn simultaneously the distribution of orientations in
the fibre should conform to the predictions of the rubber model.
The effective draw ratio, l, of the fibre can be determined from the shrinkage, S, when the
fibre is re-heated to a temperature above the glass-transition temperature Tg. S is defined by

Thus

It follows from equation (11.6) and the fact that the birefringence, , is approximately
proportional to that a plot of against should be a
straight line. Figure 11.4 shows data

for PET fibres of two different molar masses. The fibres were obtained by spinning only, with
different wind-up rates, giving different effective drawratios. Good straight lines are found
for each molar mass. The difference inslopes is due to differences in the degree of
entanglement, i.e. differentvalues of n, the polymer of higher molar mass having a larger
number of entanglements and thus a lower value of n.

11.3.3 The affine rubber model and the stress–opticalcoefficient

For the polymer considered in the previous section the birefringence mea-surements and the
stretching or shrinkage took place at different times; thebirefringence was measured in the
‘frozen-in’ state of orientation. It is, however, possible to measure the birefringence of a real
rubber when itis still under stress at a temperature above its glass-transition temperature.
This provides a simultaneous test of the predictions of the rubber deformation theory for both
orientation and stress.

It is useful to collect together three equations from previous sections,namely equation (10.3):

equation (10.17):

and equation (11.6), keeping only the first term on the RHS:

In the first two of these equations the subscript on the refractive index no indicates that it
refers to the mean refractive index and also distinguishes it from n in the third equation,
which denotes the number of random links per chain. If the equivalent random link is taken
as the structural unit, No is now the number of random links per unit volume and and
are the mean value and the anisotropy of the polarisability of a random link.
The second of these equations can be rewritten

Substituting the expressions for and from the other two equations leads to

where the substitution has been made, so that N is the number of chains per unit
volume. Equation (6.49) shows that the true stress on the rubber is given, in the Gaussian
approximation, by
so that

This equation shows that the ratio of the birefringence to the true stress should be
independent of stress. The expression on the RHS of equation (11.13) is known as the stress–
optical coefficient. A test of equation (11.13) can be made by plotting against , when a
straight line should be obtained. Such plots for a vulcanized natural rubber at various
temperatures are shown in fig. 11.5. The hysteresis shown in the curves for the lower
temperatures is interpreted as being due to stress crystallization, with the crystallites
produced being oriented in the stretching direction and

giving an excess contribution to the birefringence. At the higher temperatures this effect is
smaller or absent and the plot is linear to much higher stresses, confirming the applicability of
the rubber model for predicting orientation and stress.

An estimate of the number of monomer units per equivalent random link can be obtained by
dividing the value of calculated from the stress–optical coefficient by the anisotropy of the
polarisability of the monomer unit calculated from bond polarisabilities. This number can
more interestingly be expressed in terms of the number of single bonds in the equivalent
random link and is found to be about 5 for natural rubber, about 10for gutta percha and about
18 for polyethylene. (For the last two the values are extrapolated from measurements at
elevated temperature.) The number for polyethylene is considerably higher than the value of
3 suggested by the assumption of totally free rotation around the backbone bonds (see section
3.3.3 and problem 3.7).

11.3.4The pseudo-affine aggregate model

This model should apply best to glassy or semi-crystalline polymers, e.g. polyethylene.
Figure 11.6 shows the experimental birefringence plotted against for a series of drawn
samples of low-density polyethylene and calculated according to the pseudo-affine
deformation scheme. The general features of the curve are fitted, in particular the concavity
relative to the l-axis, but the fit is not quantitative.

11.4 Comparison between predicted and observed elastic properties

11.4.1 Introduction

The prediction of the elastic constants of an oriented polymer provides a further, more
stringent, test for the deformation models because these properties depend in a more
complex way on the details of the distribution of orientations. It is shown in the appendix that
the state of stress within a homogeneously stressed material can be represented by a
symmetric second-rank tensor with six independent components. For a cube with faces
perpendicular to OX1X2X3 the component represents the outwardly directed force per
unit area applied in the direction parallel to Oxi on the face perpendicular to Oxi . It thus
represents a normal stress. The component for i j represents the force per unit area applied
in the Oxj direction to the face perpendicular to Oxi , so that it represents a shear stress. It is
also shown in the appendix that the state of strain in the material can be

Fig. 11.6 Birefringence plotted against draw ratio for a series of drawn samples of low-
density polyethylene. The broken curve shows values calculated according to the pseudo-
affine deformation scheme. (Adapted by permission of I. M. Ward.)

Potrebbero piacerti anche