You are on page 1of 30

7 FOURIER TRANSFORM

INFRARED AND RAMAN


SPECTROSCOPIES IN THE
STUDY OF POLYMER
ORIENTATION
P. J. HENDRA and W. F. MADDAMS
Department of Chemistry, University of Southampton, Highfield, Southampton,
SOU IBJ, UK

7.1 INTRODUCTION
Infrared and Raman spectroscopy can be used to detect the presence of orienta-
tion in polymer specimens and also to quantify it. The use of infrared methods for
films has a long history, but more recently the technique has been extended to
bulky specimens. Raman scattering can in principle provide a more detailed
insight into molecular anisotropy, but is dogged by experimental difficulties. To
review the field in its entirety would be quite impossible, but it is feasible to
introduce the subject and through examples to explain its scope. Orientation
measurements based on infrared or Raman methods can be applied to static
samples or more recently, to those exposed to sinusoidally varying stress. We will
consider first the principles governing the effect that orientation has on the
observed vibrational spectra, review some of the applications and then move on
to the dynamic studies, and conclude with an account of the effect of large strain
on elastomers.
Any molecular property which is anisotropic, i.e., that shows a direcional
dependence, is per se capable of providing information on orientation both in
solids and in appropriate melts or liquids. Such properties include optical
birefringence, infrared polarization, anisotropic Raman scattering, broad line
NMR and X-ray diffraction, each of which has been exploited in the study of
polymers. These various approaches yield differing amounts of information, for
fundamental reasons; in order to appreciate why this is so it is necessary to
consider their theoretical basis, and also to have a convenient mathematical
framework by which to quantify degrees of orientation. These two topics will now
Polymer Spectroscopy. Edited by Allan H. Fawcett
© 1996 John Wiley & Sons Ltd
be considered, first in the case of infrared spectroscopy, which provides a com-
paratively simple introduction, and then for Raman spectroscopy which, in
principle, provides more detailed information at the expense of complexity of
interpretation of the experimental measurements.

7.1.1 THE BASIS O F ORIENTATION MEASUREMENTS


BY INFRARED SPECTROSCOPY

The absorption of infrared radiation occurs to a maximum degree when the


direction of the electric vector of the radiation is parallel to the direction of the
dipole moment changes involved in the various vibrational modes of the absorb-
ing molecule. If the direction of the electric vector lies at an angle to the direction
of the dipole movement change, the component of the former resolved along the
latter direction is involved in the absorption process, which thus occurs less
strongly. This is the directional property that makes infrared spectroscopy useful
for orientation studies.
This directional property is not usually a pertinent factor in determining the
intensities of the bands in an infrared spectrum, because the radiation is not
polarized and the direction of the electric vector is random in the plane perpen-
dicular to the propagation direction. Furthermore, in solutions, and in many
solid samples, the absorbing molecules are randomly oriented.
In the situation where the incident light is plane polarized, so that the electric
vector lies in a fixed direction, and the sample is partially or completely
orientated, as is often the case with polymer specimens, changes in peak intensity
will usually occur. This is readily understood by reference to Figures 7.1 (a), (b)
and (c). The first shows the simple situation of a group of molecules, which are
conveniently taken to be polymer chains in the present context, lying parallel to
each other. This type of orientation is known as uniaxial, and it often occurs as the
result of stretching or rolling in the direction along which the chains are lined up.
Consider a molecular vibration in which the direction of the dipole movement
change lies along the chain, or the principal orientation direction. Consider
further the situation when the electric vector is plane polarized parallel to this
direction. In these circumstances maximum absorption will occur, and its
magnitude will be determined by the rate at which the dipole moment changes
with changes in bond length during the vibration.
Complete uniaxial orientation seldom occurs in practice, and the typical
situation is that of a range of orientations about the direction of complete
orientation. One such chain is shown in Figure 7.1(b). The component of the
electric vector resolved along this chain, lying at an angle 9 to the complete
orientation direction, is cos 6. Then, since the absorption intensity is proportional
to the square of the dipole moment change, it follows that, if the intensity of
absorption is / along the direction of complete orientation, it will be / cos 2 0 for
Direction of dipole
moment change
1a 1b 1c

Figure 7.1 The relationship between the directions of incidence and the vibrator in
a polymer

the chain at angle 0. It is evident that for 0 = 90°, zero intensity will be observed.
The converse is that in the situation shown in Figure 7.1(a), with complete
uniaxial orientation, if the direction of the electric vector is perpendicular to the
chain direction, / will be zero. If these two intensities are denoted by J1 and J1,
their ratio, which is known as the dichroic ratio Z), will tend to infinity. If the
dipole moment change lies perpendicular to the chain direction it follows that for
complete uniaxial orientation the dichroic ratio will be zero. In both situations
the dichroic ratio will be unity for random orientation, so measurements on the
way in which it changes during a process such as stretching or rolling, which
frequently induce orientation, will yield useful information.
Figure l(c) shows the situation which is common in practice, a range of
orientations which, in the uniaxial situation, will be symmetrical about the
direction of maximum orientation. The average value of cos 2 0, denoted by
<cos20>, will then define the overall orientation of this system. It may be shown
that <cos 2 0> = /„//, + 2I1 and, therefore, that <cos20> = D/(D + 2). Thus,
<cos 2 0> may be determined by measuring D. The situation considered in
Figures l(a)-(c) is simplistic in that, in general, the direction of dipole moment
change in the polymer molecule will not lie precisely along the chain axis, but at
some angle <j>. If </> is known it then becomes possible to determine <cos20> in the
general situation. In practice <t> is often hard to locate for two reasons. Although
detailed crystallographic studies and molecular vibrational calculations have
been undertaken for most of the better known addition polymers, this is not the
case for some newer materials, and <j> for some vibrational modes is not known.
Secondly, for amorphous polymers, and indeed for the amorphous component of
partially crystalline polymers such as polyethylene, the chains may occur in more
than one conformational form, so that perpendicularity becomes rather indeter-
minate. Nevertheless, the examination of D as a function of processing operations
will often yield very useful comparative information which is unobtainable
from X-ray diffraction studies, which are specific for crystalline regions only. In
particular, it is worth noting that X-ray diffraction applies only to crystalline
fragments, but infrared dichroism is indicative of orientation throughout the
system.

7.1.2 THE BASIS O F ORIENTATION MEASUREMENTS


BY RAMAN SPECTROSCOPY

The information provided by the Raman spectrum of an oriented polymer differs


from its infrared counterpart because of the fundamentally different processes
involved in the generation of the spectra. In the infrared absorption process, as
already noted, the absorption intensity is dependent on the angle between the
electric vector and the direction of the dipole moment change. The Raman
spectrum results from inelastic photon scattering, details of which are determined
by changes in the polarizability of the chemical bonds involved. Polarizability is
a tensor quantity, which results in complications but, in principle, provides
additional information. As we have seen, infrared spectroscopy involves only one
beam of polarized radiation, and the fraction of the radiation absorbed by
a molecule depends only on the orientation of the molecule with respect to the
polarisation vector of the radiation. However, Raman scattering involves two
beams of radiation, those of illumination and collection, and the scattered
intensity depends on the orientation of the molecule with respect to the polarisa-
tion vectors of both beams, which may, of course, be different. This necessitates
more detailed measurements in order to obtain the relevant information.
It may be shown that the Raman measurements are capable of yielding
information on both < cos 2 0 > and < cos 4 6 >. The availability of < cos 4 6 > data can
be valuable is distinguishing between the differing types of stress deformation
mechanisms that have been proposed. However, an interpretation of the band
intensities in terms of <cos 2 0> and <cos 4 0> is possible only when the principal
components of the derived polarisability tensor are known. This information is
often not available and assumptions must then be made; these then render the
method non-absolute. Examples of this approach will be considered briefly below.
The interpretation of detailed orientation measurements by Raman spectro-
scopy has led to the use of various abbreviated procedures. However, unlike the
use of infrared dichroic ratios for comparative purposes, simplified Raman
procedures present pitfalls for the unwary and must be used with due care and
attention.
It is useful at this juncture to note the capabilities of the other major methods
for assessing orientation in polymers. Birefringence measurements yield values
for <cos 2 0> and broad line NMR provides <cos 2 0> and <cos 4 0>, together with
<cos60> in certain favourable circumstances. X-ray diffraction measurements
define orientation uniquely, and so give values for <cosn0>, where n takes on all
even values. Infrared spectroscopy and birefringence measurements yield the
least detailed information, but are by far the simplest methods from the experi-
mental point of view, and this is often the decisive factor.
There is no difficulty in making infrared measurements on polymer specimens
having some degree of biaxial orientation, but the interpretation of the results
may not be straightforward. It is possible in principle to express the results in
terms of a more generalized type of mathematical orientation function. However,
the approach that has tended to be used in practice, so far as is possible, is to
utilize bands arising from vibrations for which the direction of the dipole moment
change is well established, and is presumably either parallel or perpendicular to
the chain axis. The complexity of interpretation of the Raman spectra of biaxially
oriented specimens has discouraged work in this area.

7.2

7.2.1 EXPERIMENTAL T E C H N I Q U E S O N STATIC SAMPLES

In the infrared, plane polarized radiation is required, and is provided by passing


the radiation through a polarizer transparent in the wavelength domain of
interest.
A variety of these devices has been used over the years, but today the use of
a fine wire grid on a KRS-5 (thallium bromoiodide) support is pre-eminent. The
device is usually in the form of a disc 25 mm in diameter mounted in a metal
holder, but the devices are very fragile because only the slightest accidental
contact with the face coated with the gold grid is fatal to its efficiency. It is normal
to study polymers as films in transmission, having identified a reference direction
in the specimen. To remove polarization effects arising from the interferometer,
the polarizer should lie in a fixed orientation and the sample should be rotated,
rather than the reverse. Where orientation has been introduced by processing
operations such as drawing or rolling, visual inspection of the specimen usually
reveals the processing directions, widely known as the 'machine' directions, and
can be uniaxial or biaxial. The major experimental limitation in infrared trans-
mission measurements is the requirement for films to be thin enough to obtain
adequate results. In practice, this requires film thicknesses in the range 25-100/mi
depending on the type of polymer involved. See Figure 7.2.
Measurements have been successfully made using polarized infrared radiation
in attenuated total reflectance (ATR) and specular reflectance. In the ATR experi-
ment the conventional prism is replaced by a 'pent roof device and the polymer is
clamped to its surfaces with the machine direction parallel to one of the sides. The
experiment is then carried out through two adjacent faces, i.e. with the polarized
radiation passing along or normal to the machine direction. See Figure 7.3.
Comparison of the two spectra is as for transmission. Specular reflection has
a much more complex theoretical background and is not considered here.
Absoitoance

Wavenumber v (cm~1)

Figure 7.2 Dichroic infrared measurements on a film of biaxially oriented polyethylene


terephthalate

Sample

AT. R. Crystal

Sample oriented

Pent roof A.T.R. prism


Il Position
Rotation about axis G
produces 1 Position
Beam enters face A

Figure 7.3 ATR carried out with polarised light. The method was developed by the late
Micrometer

Figure 7.4 The optical arrangements typical of a FT Raman spectrometer operating


anisotropically. The laser is turned by prism P1 to the left and brought to a focus in plane S.
Samples in this plane scatter light collected by lenses L2 and L3 and pass it to the
interferometer, which lies to the right

Confining our attention to FT Raman experiments and to polymers makes the


experimental arrangements involved far simpler than they might otherwise be. In
FT instruments backscattering is almost invariably used. See Figure 7.4. The
radiation, which is from a laser source and almost always plane polarised,
irradiates the sample. The scattered light leaving the sample in the reverse
direction to the incident beam is collected with a large lens or mirror and passed
to the interferometer and detector. The collected radiation is analysed with an
analyser in the optical train (frequently between the interferometer and the
detector). Several experiments can be attempted by rotating the sample as
required and also by rotating the polarization analyser. In theory, correction
should be made for the polarization effect of the interferometer itself, but this is
usually ignored. The various experiments are too disparate to describe as || or
1 and so a nomenclature originally devised by the late S.P.S. Porto is normally
involved. It is defined in Figure 7.5. In principle, samples of any reasonable
thickness can be studied with FT Raman spectroscopy and the sampled volume
can be located from the surface back into the bulk at will. However, a high degree
of optical clarity is essential if the polarization sense is not to be scrambled.
With care, convincing measurements can be made on opalescent materials,
e.g., ultrahigh-modulus polyethylene, which is milky and non transparent. See
Figure 7.4. If study in the bulk phase of an inhomogeneous specimen, e.g.
a transparent moulding in, say, polymethyl methacrylate or poly-4-methyl-l-
pentene is planned, a word of caution is pertinent. As the laser passes through the
Crystal axes

Direction of View Analyze


Direction of Illumination Laser Polarisation

Figure 7.5 The experimental nomenclature defined by the late Dr S.P.S. Porto

surface of the specimen and then the anisotropic bulk on its way to the volume
viewed by the spectrometer, the plane of polarization will be rotated. The
scattered light will also be affected, so the conclusions reached about the
orientation within the specimen may be unreliable.

7.2.2 INFRARED SPECTROSCOPIC STUDIES ON


ORIENTED POLYMERS
The purpose of this section is to give readers an indication of the type of
information that may be obtained, not to provide a detailed literature survey. It
covers both the use of dichroic ratio measurements, the more common approach,
and the formal use of orientation functions.
The value of dichroic ratio measurements, taken in conjunction with polymer
processing variables, is demonstrated excellently by a study on cold drawn linear
polyethylene, made a quarter of a century ago by Glenz and Peterlin [ I ] . The
results are shown in Figure 7.6. Consider first the results for the band at
1894 cm " l . The dichroic ratio decreases rapidly with increasing draw ratio and
asymptotically approaches zero as the draw ratio exceeds « 5 . The 1894 cm" 1
band is a combination mode of the Raman active methylene rocking mode at
1170 cm ~ * and the methylene rocking mode at 720 cm " *, and is characteristic for
crystalline regions of the polymer. Its dipole moment is perpendicular to the
Draw ratio
Figure 7.6 Dichroic ratio D vs. draw ratio for cold drawn linear polyethylene [1]

c axis, i.e., the long chain direction of the polymer. Hence, the early decrease in the
dichroic ratio with increasing draw ratio shows conclusively that there is rapid
orientation of the c axes of the crystallites parallel to the draw direction as the
sample is deformed.
The peak at 1368cm" x arises from the symmetric CH 2 wagging mode for the
CH 2 —CH 2 group, where the C—C band is at the centre of a gauche-trans-
gauche sequence. It differs from the behaviour of the band at 1894 cm" x in two
obvious respects: namely, that the approach to the limiting value with increasing
draw ratio occurs decidedly more slowly as deformation increases, and the
ultimate value indicates that there is only partial orientation of this structural
unit. This was interpreted as evidence for a clear segregation of the crystalline and
amorphous phases. The corresponding asymmetric wagging vibration gives rise
to the band at 1303 cm" 1 , and the polarized intensity of this is virtually
dependent on draw ratio, showing that substantial orientation is not occurring,
although a knowledge of the direction of dipole moment change is required in
order to interpret the results more fully. The peak at 1078 cm" 1 comes from
a gauche C—C stretching and some CH 2 wagging. Its behaviour is intermediate
in that it reaches its ultimate dichroic ratio of «0.7 at relatively low draw ratios,
but the final value is a reasonable indication of partial orientation only, as would
be expected for methylene units in amorphous regions.
The value of the approach using dichroic ratios coupled with a knowledge of
the assignments of peaks is further illustrated by measurements on plain poly-
ethylene films [2]. The production process leads to a rather complex orientation
pattern. The molten polymer is extruded as a thin, hollow cylinder, in what may
be termed the machine direction. It is simultaneously expanded in a plane
perpendicular to the machine direction by the application of internal pressure.
Additional variables are the extrusion temperature and the rate at which the
blown film has been cooled. The resulting orientation behaviour is best studied
by x-ray diffraction pole figure measurements [3,4,5] but the infrared approach
provides a relatively simple means for obtaining a useful amount of information,
particularly for the behaviour of chains in amorphous regions.
Dichroic ratios were measured for the bands at 1080, 1303, 1352 and
1368 cm" 1 . The first of these is associated with tie chains between crystalline
lamellae in amorphous regions, and the remaining three involve methylene group
vibrations of loose chain folds in amorphous domains. It was possible to correlate
the results with the occurrence of two types of orientation that had earlier been
characterized by X-ray diffraction measurements. They are termed high- and low-
stress orientation, whose occurrence depends on the blowing conditions and those
during film production. The first type of orientation is analogous to that found in
cold drawn polyethylene, discussed above, in having the c axis distribution of the
crystalline regions substantially along the machine direction. The low-stress
orientation, which occurs the more frequently, is the result of the type of crystal-
lization process described by Keller and Machin [6]. The a and c axes are inclined
at an angle to the plane of the film, with a strong tendency for the greater concen-
tration of a axes to lie near to the machine direction. The three peaks at 1303,1352
and 1368 cm" * show appreciable orientation in the high-stress type of films but very
little with the low-stress materials. Conversely, the extended tie chains, characterized
through the 1080 cm" i band, are appreciably oriented in the low-stress films but
not so in their high-stress counterparts. These results, together with those from
X-ray diffraction measurements, proved to be of appreciable value in selecting the
best blowing conditions to manufacture films having optimum tear strengths.
The more formal approach, involving values of <cos 2 0>, has been used by
Purvis et al. [7] to study uniaxially oriented specimens of polyethylene tereph-
thalate). These measurements were part of a wider study involving birefringence
and Raman studies, and the results are more conveniently considered in the
following section.

7.2.3 RAMAN SPECTROSCOPIC STUDIES O N


ORIENTED POLYMERS

The ability of Raman spectroscopic studies on oriented polymers to yield values


for both <cos 2 0> and <cos 4 0> has led to its use with polyethylene, atactic
poly(methyl methacrylate), poly(ethylene terephthalate) and poly(vinyl chloride).
Some results that have been obtained will be considered briefly.
Despite the problems of Raman spectroscopic studies when working with
polyethylene specimens, Maxfield et al. [8] obtained spectra of adequate quality
by immersing thin drawn films of low density polyethylene in silicone oil to
minimize surface scattering. They used the 1170cm" 1 band, characteristic for
chains in crystalline regions, and one at 1081 cm" \ specific for the amorphous
ones, to obtain values for <cos 2 0> and <cos 4 0>. The former agreed well with
those deduced from infrared measurements. Both sets of values are reasonably in
line with theoretical estimates obtained from the Roe and Krigbaum [9] cross-
linked rubber network model, and the crystalline orientation factors are as
expected on the basis of the phenomenological theory of Yoon et al. [10].
The major workers in the field have been Professor LM. Ward and his
colleagues at the University of Leeds. The technique was first applied to
poly(ethylene terephthalate) [T]9 abbreviated to PET for convenience, using
uniaxially oriented tape specimens. Measurements were confined to the benzene
ring mode band at 1616cm" x and only three independent intensities, in terms of
possible orientations, were measured. Nevertheless, some interesting results were
forthcoming. Values of <cos 2 0> were plotted as a function of birefringence, and
a straight line was obtained. Extrapolation to <cos 2 0> = 1 permitted a value for
the maximum birefringence to be obtained, and this agreed well with that
suggested by Kashiwagiet al. [ H ] . Plots of <cos 2 0> and <cos 4 0> as a function
of draw ratio were compared with the predictions from the affine rubber elasticity
model and the pseudo-affine aggregate deformation model. There is moderately
good agreement with the rubber model for draw ratios less than « 3 , and some
indication that for those in excess of 4.5 the pseudo-affine aggregate model
becomes more appropriate.
Purvis and Bower [12] subsequently extended the work, using four additional
Raman peaks, all of which predominantly involve motion of the terephthalyl
moiety, to a first approximation. The results suggested that there is no preferred
orientation of the plane of the ester group with respect to the plane of the benzene
ring in the amorphous phase. They also concluded that the drawing of PET
occurs in a similar manner to the extension of a rubber-like network.
Purvis and Bower [13] also examined drawn specimens of amorphous
poly(methyl methacrylate), using four peaks in the Raman spectrum. They were
not able to distinguish between two plausible structural models, one essentially
linear and the other helical, but they were able to show that their results are
consistent with the affine deformational model.
The success of these measurements prompted a study on drawn specimens of
plasticized PVC [14]. This poses problems because of the complexity of the
spectrum, which shows overlapping peaks arising from configurational and
conformational isomers. In this first study, attention was confined to peaks at 608
and 638 cm" 1 specific for long planar syndiotactic sequences, the type of unit
INTENSITY

W A V E N U M B E R ( C m-1)

Figure 7.7 FT Raman measurements on ultra high-modulus polyethylene. Spectra


recorded in two ways-analyzer vertical or horizontal polarised vertically.

involved in the crystalline regions of the polymer. All possible orientation and
polarization combinations were used, giving a total of eight intensity measure-
ments per peak. The results, including a comparison of < cos 2 6 > values with those
from birefringence measurements, showed that the crystalline regions are more
highly oriented than the non-crystalline ones in samples containing the larger
amounts of plasticizer and drawn at the higher temperatures. In a continuation of
this work, Bower et al. [15] used the intensity of the 616 cm" * band, specific for
short syndiotactic sequences probably present in amorphous regions. The results
support the earlier tentative conclusion that amorphous chains behave in
a rubber-like way during orientation.
The results described above all refer to Raman measurements made prior to
the introduction of FT methods and near-infrared sources. More recent work
shows that anisotropic measurements are far easier than they were, and can be
made at room temperature on heated or cooled specimens with consummate
ease. The measurements on highly oriented polyethylene shown in Figure 7.7 are
simple to produce both at room temperature and at — 1800C. An analysis is
available [16].
Several very preliminary reports have appeared in the recent literature, where
a 'dichroic' measurement has been attempted, in that spectra have been recorded
with no polarization analyser and with the machine direction of the sample set
parallel or perpendicular to the electric vector of the laser. They are, of course, in
each case the sum of several of the spectra shown in Figure 7.7. Further, the
polarization of ingoing and outgoing radiation may be rotated as described in
Section 7.2.1. The results indicate the orientation and variations thereof but
cannot be used to give quantitative data. We find the approach useful but it has to
be used with caution.

7.3 TIME RESOLVED MEASUREMENTS


As we have seen, the changes to a vibrational spectrum as orientation is induced
are ones of intensity. Application of stress is well known to induce frequency
shifts. However, both effects only subtly change a well developed spectrum
characteristic of the specimen itself. One obvious way to study these subtle
changes is to apply force sinusoidally and to discriminate electronically between
the DC component of the signal (the invariant) and the AC (that of interest).
Further, in several practical situations polymers are regularly subjected to
variable loads, and their behaviour under these situations is critical. In addition,
their deformations under quasi-static stresses are very different from those under
alternating ones.
Polymers behave to varying degrees as viscoelastic materials, and this has
considerable consequence for their response to loading at moderately high
frequencies. Their behaviour under such conditions has been studied by a variety
of methods that come under the general heading of dynamic mechanical testing.
However, until comparatively recently, spectroscopic methods have not been
applied to the problem. The work in this area will be considered and, in order to
provide the foundation for this discussion, the basic theory for the response of
a viscoelastic system to sinusoidal stress will be given first.

7.3.1 THE RESPONSE O F A VISCOELASTIC SYSTEM TO


S I N U S O I D A L STRESS

Consider an applied stress varying as a function of time according to the rela-


tionship (T = (T0 sin co, where co is the stress frequency. If the material were wholly
elastic, and obeyed Hooke's law, the strain would then vary as e = e0 sin cot.
However, for a viscoelastic material the strain lags behind the stress. Let this
lag be denoted by S9 which may be called the phase angle or the phase lag, and is
the relative angular displacement of the stress and strain.
The appropriate equations then become
G = a0 sin cot
and
e = e0 sin ((Dt — S)
Hence
e = e0 sin cot cos 3 — e0coscot sin S (1)
The strain can therefore be considered in terms of two components, one of
which, e0 sin cot cos <5, is in phase with the strain, and the other, e0 cos cot sin <5, is
out of phase with the strain by n/2. It is therefore possible to define two dynamic
moduli, E1 in phase with the stress and E2, which is n/2 out of phase with the
stress. E1 = (cr Je0) cos d and E2 = ((T0Ze0) sin S.

E1=(V0Ze0)COsS + E2 = {<70/e0)sin8
Substitution into (1) then gives
e= E^I/GQ sin cot — E2e\/a0cos cot (2)
and
tan S = E2/EX (3)
tan S is frequently used as a measure of viscoelastic character. For a given
temperature, E1^E2 and tan S are functions of the frequency of the applied stress.
In general, tan 5 and E2 are usually small at very low and very high frequencies
and their values pass through maxima at some intermediate frequency. On the
other hand, E1 is high at high frequencies in the case of glassy polymers and low at
low frequencies for rubbery polymers.
One of the main reasons for studying the frequency dependence of the dynamic
mechanical properties is that it is often possible to relate peaks in E2 and tan S to
particular types of molecular motion in the polymer, via what may be regarded as
a 'resonance' effect. For example, there are particularly strong 'resonances' at the
glass transition. Peaks in E2 are also often found at the melting temperature in
semi-crystalline polymers, a consequence of the greater freedom of molecular
motion that is possible when the molecules are no longer arranged into a regular
crystalline structure. Other types of molecular motion, such as those involving
the rotation of branches, often give detectable but smaller peaks in E2 and tan <5,
and these are usually referred to as secondary transitions. If a series of such
transitions occurs over a range of temperature, the peak which occurs at the
highest temperature is termed a, and the subsequent ones are called /?, etc.
Hence, dynamic mechanical measurements are usually made over a range of
temperature and frequencies in order to cover the various types of molecular
motion that may occur. The oscillatory strain amplitudes used are very small,
typically below 1.0% of the total sample dimension, to ensure a linear viscoelastic
response.
Molecular motions in polymers, particularly those types that involve some
reorganization of functional groups such as branches, should be amenable to
study by vibrational spectroscopy. The spatial movement of functional groups
involves a change in the directions of dipole moment and polarizability changes
during molecular vibrations. Hence, the measurement of linear dichroism using
polarized radiation should prove useful as a complement to dynamic mechanical
measurements, or as a characterizational technique in its own right.
Because the oscillatory strain amplitudes involved are very small, it is easier to
measure the difference between A1 and A19 the absorbances measured for
a particular peak for light polarized in planes parallel and perpendicular to
a fixed reference direction of the sample, than to measure their ratio A^fA19 as is
commonly done in making orientation measurements on appreciably deformed
polymers. A1 — A1 is termed the dichroic difference and is usually denoted by AA,
or AA(t) in the case of a time dependent signal. It is then easy to show, using the
type of reasoning involved in obtaining Equation (2), that the dynamic dichroism
signal AA(t) can be separated into two orthogonal components given by the
equation
AA(t) = AA' sin ot + AA" cos cot (4)
11 1
and tanS = AA JAA , where tan<5 is termed the dichroic dissipation factor,
analogous to the mechanical dissipation factor considered above. The terms AA'
and AA" are known as the in-phase spectrum and the quadrature spectrum
respectively of the dynamic infrared linear dichroism. They represent compo-
nents of dynamic optical anisotropy caused by the re-orientation of electric
dipole transition moments. The in-phase spectrum is a measure of the instan-
taneous strain and the quadrature spectrum characterizes the component of
re-orientation proportional to the rate of strain which is n/2 out of phase with the
stress. The applied oscillatory stress provides a mechanism for perturbing the
system and stimulating individual functional groups into the specific reorienta-
tional responses, which are then characterized by the resulting dichroic measure-
ments.

7.3.2 EXPERIMENTAL

The equipment required to measure in-phase and quadrature linear dichroic


spectra consists of two components, the transducer necessary to provide the small
amplitude oscillatory strain and a suitably modified infrared spectrometer. So far
as the former is concerned, it is desirable that the transducer system is capable of
operating over a range of frequencies, in order to provide the flexibility that may
be required in probing a change of re-orientational responses. The transducer
may either form part of a dynamic mechanical analyzer, as described by Noda
et al. [17] or may be a simple unit used solely as part of an infrared spectrometer
system.
The majority of the measurements reported hitherto have been made on
modified dispersive spectrometers, and the system of Noda et al. [17] is typical.
This involves three successive modulations of the infrared beam as it passes
through the spectrometer, which is custom designed and built. The first stage of
modulation is via a mechanical light chopper, in order to eliminate background
radiation and sample emission. The beam is then polarization modulated using
a combination of a wire grid polarizer and a photoelastic modulator, whereby the
plane of polarization of the light alternates rapidly between directions parallel
and perpendicular to a fixed reference axis, conveniently the direction along
which the strain occurs. Two photoelastic modulators were used, having modula-
tion frequencies of 37 and 84 kHz, with corresponding time resolutions of 14 and
6 ^s. The dynamic dichroism of the sample leads to the third modulation of the
beam.
In view of this triple modulation process, a rather complicated demodulation
scheme has been employed in order to extract the required information from the
signal from the detector. This involved a set of five lock-in amplifiers. A full
mathematical analysis of this system has been provided by Noda et al. [17]. In
addition to the fine time resolution performance already noted, the sensitivity of
the instrument is such that it will detect a signal as small as 10"4 absorbence unit
at a resolution of 4cm" *. Chase and Ikeda [18] have also described a dispersive
instrument of this type.
More recently, and predictably, Fourier transform infrared spectrometers have
been used to measure dynamic linear dichroic spectra. However, substantial
modifications to the conventional type of FTIR spectrometer have been necess-
ary in order to overcome a basic problem. This arises from the fact that it is
necessary to separate the time dependence of the sample response from that of the
spectral multiplexing, because the interferogram itself is a cosine function of time.
For example, if the moving mirror has a velocity of 0.5 mm per second at
3000 cm" 1 the radiation is frequehcy modulated at 300Hz.
Until recently the almost universal approach to the problem has been the use of
step-scanning FT interferometers. Such instruments avoid the problem of the
separation of two time dependent variables by creating the interferogram point
by point. At each retardation position of the interferometer mirrors, data are
collected for as long as is required to obtain the desired signal-to-noise ratio, and
a single interferogram is recorded for Fourier transformation. By this means the
spectral multiplexing is uncoupled from the time domain.
The essential difference between conventional FT instruments and the step-
scan devices is that, for successful operation, it is necessary to control the
retardation (mirror) velocity in the case of the former and the retardation (mirror)
position for the latter. In both cases, the method used to control the retardation
involves a collinear or parallel helium/neon laser interferometer. In continuous
scan operations the laser interference fringes are used to generate feedback signals
to maintain constant mirror velocity, and in the step-scan mode the laser
interferogram provides the means for the control of the mirror position via
a feedback signal.
Several methods are available for setting the retardation and the stepping. An
early approach was that of Manning et al. [19,20] for use with an IBM IR-44
spectrometer. The control signal is generated by path difference modulation or
'dithering' of the moving mirror via an AC voltage applied to the drive mechan-
ism. The resulting 'phase modulation' in the control laser detector signal is used in
a lock-in feedback circuit to the drive mechanism system. This step-scan instru-
ment has a retardation position uncertainty of +15 nm and a minimum stepping
rate adjustable over the range 0-100 Hz, although data are normally collected
near to the bottom of this range.
Gregoriov et al. [21], in a subsequent development, used either complete
digital control of the position and stepping or a combination of analogue and
digital control, with phase modulation available as an optional extra, in convert-
ing a Nicolet System 800 into a step-scanning instrument. They were able to
achieve a positional uncertainty of « ± 1 nm. This type of control has been used
by Bruker in their IFS 88 instrument and the more recent Bio-Rad Digilab FTS
70A. Fuller details of step-scan instruments are given in the excellent review
article by Palmer et al. [22] Marcott et al. [23] have made comparative dynamic
measurements on a thin film of atactic polystyrene, using their dispersive
spectrometer described above in comparison with a Bio-Rad FTS 60A instru-
ment operating in the step-scanning mode. They concluded that, although the
dispersive approach produces higher signal-to-noise ratios over small spectral
regions, the multiplex advantage makes the FT approach attractive when
broader spectral coverages are required.

Block Diagram of Dynamic Infrared System


Based on Rapid Scanning FTIR Spectrometer

to 90'Phase S/H
ADC
Shifter
ADC trigger

Control

25Hz Reference
Osc S/H S/H

Low
Amp Pass S/H Multiplex
SYS2000 Filter
FTIR
Polariseir Sample MCT Infrared beam
Stretcher Detector S/H Sample & Hold

Figure 7.8 Block diagram of the Perkin-Elmer 2000 system for studying time resolved
phenomena.
MODULATION
RAYLEIGH LINE FREQUENCY
ARB. UNITS

ALIAS OF 2ND HARMONIC


RAMAN SPECTRUM
WATER

ARTEFACTS

UPPER SIDE-BAND
D.C. SPECTRUM
SHIFTED BY 5260 CM-1

WAVENUMBER/cm1

Figure 7.9 Bennett's experiment on heated polystyrene. He applies a modulation to the


laser source at 540 Hz. This shifts the Fourier transformed spectrum by ± 5260 cm " ! . The
shifted spectrum labelled 'upper side-band' is that of the polymer, the unshifted one, the
DC spectrum, that of polystyrene over the black body emission

Turner and Hoult [24] have demonstrated recently that very satisfactory
results may be obtained from a conventional scanning FTIR system, using
synchronous lock-in detection. Their system has a photoelastic polarization
modulator operating at 74 kHz. A reference signal is taken from this modulator
and fed to the lock-in amplifier and sampled at the same time as the infrared
signal at the detector. There are, therefore, two associated data points for each
optical path difference of the interferometer, namely the modulated interfero-
gram signal and the photoelastic modulation signal. The data are demodulated
and then accumulated with an appropriate software routine. See Figure 7.8. This
approach has also been successfully exploited by Bennett [25] for the removal of
thermal backgrounds from Raman spectra using a modulated laser source in
conjunction with a conventional scanning FT Raman spectrometer; see Fig-
ure 7.9. There are therefore now proven methods for obtaining dynamic linear
dichroic infrared spectra. This has already prompted a range of studies and more
will doubtless follow.
In the early studies the in-phase and quadrature spectra were examined by
conventional interpretational techniques, and useful information was forthcom-
ing. However, the value of correlation analysis quickly became evident, as the
result of the pioneering activities of Noda [26, 27], and this had led to the use of
so-called two dimensional infrared (2D IR) spectroscopy to display graphically
the results of correlation analysis.
As noted above, the in-phase and quadrature spectra represent components of
dynamic optical anisotropy caused by the re-orientational behaviour character-
istic of the type and local environment of each group. Reorientation processes
tend to synchronize if there is a specific chemical interaction or connectivity
between them, and herein lies the value of correlation analysis, in that it provides
a valuable method for studying the time dependent variation of infrared dichro-
ism signals.
If dichroic differences are measured at two wavenumbers V1 and v2, two
orthogonal correlation spectra may be defined as follows:

0(V1, v2) = IAAXv1)AAXv2) + A^l"(v1)A^'/(v1)A^//(v2)]/2 (5)


*(vi, v2) = IAA-(V1)AAXv2) - AAXv1)AAXv1)AA-(V2W (6)
They are respectively referred to as the synchronous and asynchronous 2D
infrared spectra. The synchronous spectrum characterizes the degree of coher-
ence between the dynamic fluctuations of signals measured at two wavenumbers,
and the correlation intensity becomes significant only if the reorientation rates of
dipole transition moments are similar to each other. The asynchronous spectrum,
however, characterizes the independent, uncoordinated out-of-phase fluctu-
ations of the signals. Hence the asynchronous correlation intensity becomes
non-vanishing only if the signals vary at different rates.
Peaks along the diagonal position of a synchronous 2D spectrum are referred
to as autopeaks. They indirectly represent the local mobility of chemical groups
contributing to the molecular vibrations at that wavenumber. Peaks located at
off-diagonal positions of a 2D spectrum are known as cross peaks. They appear
when the dynamic vibration of infrared dichroism at two different wavenumbers
are correlated with each other because the two signals are fluctuating more or less
in phase with each other. As long as the normal modes of vibrations correspond
to reasonably pure group frequencies, the cross peaks in a synchronous 2D
spectrum may be used to map out the degree of intra- and inter-molecular
interaction of various functional groups.
Cross peaks in an asynchronous 2D spectrum provide complementary infor-
mation. They appear if the signals are out of phase with each other. Even if the
characteristic band frequencies are similar and absorption peaks in the conven-
tional infrared spectrum overlap an asynchronous 2D spectrum can differentiate
them, because their time dependent intensity fluctuations differ slightly. Spatial
and temporal information about the re-orientation processes of transition mo-
ments and their associated chemical groups can be obtained from the sign of cross
peaks. If the sign of a synchronous cross peak is positive, the corresponding pair
of dipole transition moments reorients in the same relative direction. If the sign is
negative the re-orientation directions are perpendicular to each other. A positive
Wavenumber, V2
Wavenumber, V1

Figure 7.10 A schematic representation of synchronous 2D correlations. For details see


text. Diagram due to Noda et al. [44]

peak in an asynchronous spectrum indicates that the transition moment with


vibrational frequency V1 re-orients before v2.
This may be appreciated more readily by reference to Figure 7.10, which shows
schematic synchronous 2D correlations for two pairs of peaks A and C, and B and
D. The shaded areas represent negative intensity correlations. A cross peak is
negative if the changes are in opposite directions, as in the case with bands A and
C. To summarize, 2D spectra are capable of revealing rather detailed informa-
tion, as will emerge clearly below when typical examples are considered.

7.3.3 SOME EXAMPLES OF DYNAMIC LINEAR


DICHROIC INFRARED STUDIES

A detailed review of the published work is not possible in the available space.
Selected examples will therefore be used to illustrate the information that may be
obtained from the inspection of in-phase and quadrature spectra, and from the
use of correlation analysis. Studies on polystyrene provide a very convenient
introduction.
Noda et al. [17] have measured the in-phase and quadrature spectra of thin
films of polystyrene supported on a Teflon film, concentrating on the spectral
regions 1425-1525 cm~* and 2800-3200cm" 1 . The former contains two peaks of
particular interest, at «1450 and 1490 cm" 1 . The first of these is made up of
overlapping bands from two uncoupled vibrations: a CH 2 scissoring motion in
the polymer backbone and an aromatic ring stretching vibration that is locally
polarized along the bond between the phenyl group and the polymer backbone.
The 1490 cm" l band is assigned to the coupling of an aromatic CH deformation
with another aromatic ring stretching mode, locally polarized in the plane of the
ring, perpendicular to the bond between the phenyl group and the backbone
aliphatic chain. The 1490 cm" 1 band has a significant signal in the quadrative
component, which is shifted to higher wavenumbers, whereas the 1450 cm" 1
peak is closer to being in phase with the applied stress. This difference suggests
that there may be some fraction of the aromatic side chains which is responding
to the applied stress at a rate different from that of the polymer backbone.
Unlike the situation at 1450 cm " x , where there is clear separation between the
aromatic and aliphatic C-H stretching bands, the specific band assignments in
the aromatic C-H stretching region are not wholly certain. There are two clear,
positive peaks at 3028 and 3058 cm" 1 in the in-phase spectrum, and this could
indicate that the relevant transition moments are locally polarized in the plane of
the phenyl ring, perpendicular to the band between the phenyl group and the
backbone aliphatic chain. However, this does not prove that this is universally
the case; in a non-crystalline polymer such as atactic polystyrene, individual
functional groups can be oriented in a variety of directions relative to the
reference strain axis. The aliphatic CH 2 symmetric stretching mode giving the
band at 2854 cm" 1 is also interesting. This band is quite strongly negative in

CH2-stretchtng

Asymmetric Symmetric
Wavenumber, v2

Cross peak
Autopeak

Wavenumber, V1

Figure 7.11 The infrared spectrum of polystyrene shown in a 2D presentation. For


details see text. Diagram due to Noda et al. [17]
the in-phase spectrum and is consistent with the polymer chains tending to align
in the direction of the applied stress.
Noda et al. [28] have interpreted measurements on atactic polystyrene using
the correlation technique, and the two groups of C-H stretching bands are again
of interest. Figure 7.11 shows the synchronous 2D correlation spectrum over the
aliphatic C-H stretching region. The large autopeak on the diagonal represents
the re-orientational motions of dipole transition moments assignable to the
symmetric and antisymmetric methylene CH 2 stretching vibrations. The appear-
ance of synchronous cross peaks at the corresponding spectral coordinates
indicates that the dipole transition moments for these two bands re-orient at
a similar rate, as might be expected. The sign of the synchronous cross peaks is
positive, suggesting that the transition moments of the two vibrations are both
realigning in the same direction, namely perpendicular to the direction of applied
stress. As both dipole transition moments are known to lie perpendicular to the
polymer backbone, it is reasonable to conclude that the chain of polystyrene must
be aligning in the direction of applied stress, even in the glassy state.
The re-orientation dynamics of the side group phenyl rings is more complex.
Figure 7.12 shows the asynchronous spectrum involving the phenyl side groups
and the backbone methylene groups, and strong cross peaks are present. This
suggests that the re-orientational motion of the phenyl groups under strain is
quite different from that of the backbone. This, in turn, requires that there must be

Side-group phenyl
Backbone methylene

Wavenumber, v2

Wavenumber, V1

Figure 7.12 The infrared spectrum of polystyrene shown in an asynchronous form.


Diagram due to Noda et al. [17]
substantial freedom for the side groups to realign independently of the main
polymer chain. Consequently, it is impossible to characterize the main chain
dynamics of polystyrene simply by studying the re-orientation of phenyl groups.
However, 2D correlation spectra in the ring stretching modes region are more
productive, because the dipole transition moment directions are well established.
Such spectra show that, in the glassy state, side group phenyls tend to re-orient
along the direction of the main chain backbone. This is somewhat unexpected, as
such re-orientational motions require highly distorted local conformations of
chain segments. As the temperature is raised well above the glass-to-rubber
transition point, the asynchronous relationship between side groups and the
main chain of polystyrene decreases considerably. It is therefore reasonable to
conclude that the anomalous re-orientation behaviour of phenyl groups of glassy
polystyrene results from highly localized and constrained motions of functional
groups in a molecular environment with limited free volume at temperatures
below Tr
Polymer blends have proved another fruitful field for study, with considerable
technological implications. Particularly simple, but commercially important, is
the behaviour of a blend of low density polyethylene (LDPE) and high density
polyethylene (HPDE); a 50:50 blend of these two materials is semi-crystalline. In
Wavenumber

Wavenumber

Figure 7.13 The asynchronous correlation map for low and high density polyethylenes.
Diagram due to Gregoriou Noda et al. [29]
view of the spectral similarity of these two materials, Gregoriou et al. [29] used
a blend of perdeuterated HDPE and conventional LDPE. A portion of the
asynchronous correlation map is shown in Figure 7.13. There are clearly two
components within the 1088 cm" 1 band, at 1085 and 1091cm" 1 , as is evident
from the appearance of the corresponding cross peak. Additionally, a negative
cross peak exists between the transition dipole moments at 730 and 1091 cm" \
suggesting that the crystalline component of rf-HDPE of the blend re-orients
faster than the crystalline LDPE portion under a positive cross peak between the
dipoles at 721 and 1085 cm ~ l and also that, in the corresponding synchronous
plot, the corresponding peak is also positive.
It has been suggested [30] that, when a melted blend of HDPE and LDPE
cools, HDPE crystallizes first because of its higher melting temperature, resulting
in a volume-filling superstructure of HDPE crystals forming a skeletal network.
When the LDPE begins to crystallize it forms disjointed crystallites filling the
interstitial space of the HDPE network. If such a system is deformed, the initial
observable response will result from the deformation of the HDPE crystalline
network. The stress will then transfer to the interstitial spaces, and there will be
a secondary orientation of LDPE crystallites. The interpretation of the 2D
spectra of the polyethylene blend is in good agreement with this model.
The miscibility of some polymer blends is of considerable technological
importance although, fundamentally, the reasons for the miscibility are not
completely understood. Polystyrene (PS) and poly(2,6-dimethyl-l,4-phenylene
oxide) (PPO) is one such system. 2D correlation studies have been made on
a blend of 80% of the former and 20% of the latter by Palmer et al. [31]. The
results suggest a different dynamic behaviour for the PS and PPO portions of the
blend, depsite their compatibility, with the PS chains responding to the pertur-
bing force faster than those of PPO. Some asynchronous cross peaks develop
between the constituents, indicating the possible existence of submolecular level
microheterogeneity.
Atactic polystyrene and poly(vinyl methyl ether) provide another case of
miscibility of two very different polymers, the latter being water soluble. Al-
though the conventional infrared spectrum shows a single peak for the C-H
stretching mode of the methoxyl group, the asynchronous 2D spectrum of the
blend reveals two separate peaks assignable to this mode. Strong synchronous
cross peaks exist between one of these two methoxyl peaks and some phenyl
group modes of the polystyrene, indicating the possible existence of a specific
intermolecular interaction between the phenyl and methoxyl groups.
The tri-block polymer styrene-butadiene-styrene, with a weight ratio of 78/28
in favour of butadiene, has also been studied [32]. The in-phase and quadrature
spectra over the region 2700-3100Cm"1 show that, at room temperature, the
styrene portion of the copolymer displays negligible dynamic response when
butadiene forms the continuous matrix. This is not unexpected, as PS is in the
glassy form whereas PB is a rubbery phase. The interesting feature of the two
spectra is the appearance of incipient fine structure centred at « 2920 cm l. The
2D correlation spectra prove revealing. The synchronous map indicates the
existence of peaks at 2933 and 2915Cm - Mn the asynchronous correlation map
there are several peaks in the vicinity of the asymmetric C-H stretch at
2920cm" 1 . There are a positive cross peak between 2936 and 1915cm" 1 ,
a positive cross peak between 2936 and 2930 cm " \ a negative one between 2930
and 2922 cm" 1 and a positive cross peak between 2922 and 2915 cm" *. Interest-
ingly, Fourier self-deconvolution failed to resolve these features under the broad
asymmetric C-H stretching peak.
This ability to enhance spectral resolution has also been demonstrated in the
case of atactic poly(methyl methacrylate). This has three very overlapped peaks in
the C-H stretching region, whose presence has been revealed by studies on
polymers with varying degrees of deuteration [33], a useful but not particularly
convenient approach. The three peaks are specific for the ester methyl groups, the
a-methyl groups and the backbone methylene groups. The 2D correlation
approach yields equally specific information without resort to deuterated poly-
mers. Strong synchronous cross peaks occur at spectral coordinates specific for
ester methyl groups, and the a-methyl group is clearly differentiated from the
ester methyl group on the basis of cross peaks appearing in the asynchronous
spectrum. Furthermore, analysis based on the signs of cross peaks provides
detailed information on submolecular reorientation mechanisms that occur with
small strains.
The technique has also been used to study the dynamic behaviour of a poly-
mer-dispersed liquid crystal subjected to an electric field [18]. The liquid polymer
used was the commercially available nematic liquid crystal mixture E7, which
contains four nitrile and ethyl substituted bi- and tri-phenyls. It was blended with
a polymer precursor consisting of a mixture of an acrylate monomer, an acrylate
oligomer and a UV curing agent. The 2D correlation analysis showed that the
rigid core of the liquid crystal molecules re-orients as a unit, and suggests that the
polymer side chains existing in the interface between the polymer and the liquid
crystals may re-orient in phase with the liquid crystal re-orientation by interac-
tion with the liquid crystal molecules.
The work discussed hitherto has been concerned solely with polymers subjec-
ted to sinusoidally varying stress. The use of stress with this simple wave form is
not a necessary condition for the production and use of 2D infrared correlation
spectroscopy. Noda [35] has shown that signals fluctuating as an arbitrary
function of time may be dealt with and, in some circumstances, offer advantages
over sinusoidal signals. He has provided the necessary mathematical framework
for this more general approach, and the method has been used to study the
photopolymerization of acrylic and epoxy monomers [36]. By this approach,
features associated with spectral intensity changes and peak shifts arising from
the polymerization reactions were clearly observed. It is reasonable to predict
that the method will find further applications in this field.
7.4 ELASTOMERS UNDER STRESS
Unlike thermoplastics, elastomers are capable of supporting massive reversible
strains and yet recovering their original dimensions, i.e., they behave as 'classic'
springs recovering their dimensions elastically and reversibly after deforming to
three and even six times their original length. Examination of the stress: strain
curve shows, however, that many rubbers do not obey Hooke's law, their
modulus rising with strain. This comes about because they partially crystallize at
higher strains, the micelles of crystalline order lying parallel to the stress
direction. Since crystalline polymers show molecular vibrational characteristics
different from the non-crystalline materials (because the vibrations are sensitive
to the rotational isomerism of the backbone and the intermodular interactions),
the spectrum of the highly strained polymer will differ from that of the relaxed
state. In addition, of course, the spectrum will reflect the onset of orientation in
the otherwise random matrix and in the crystallites as they form. To demonstrate
these points, we offer in Figure 7.14 spectra of vulcanized natural rubber, both
stressed and relaxed [37].
The vibrational changes that occur on orientation and crystallization have
been used to research the origin of the residual orientation frequently found in
blown or extruded film. These materials frequently show quite well developed
orientation, and hence are useful as shrink wrapping. As the flowing melts from
which they are formed are optically dichroic, it seems reasonable to propose
a model involving flow-orientation-crystallization and solidfication in an
oriented manner. It has been shown, however, that the orientation of a flowing
polyethylene melt (as measured by infrared, Raman diffraction and X-ray
diffraction) is very small [38].
This can only mean that the few longer chains present are oriented under flow,
and these nucleate oriented crystallization. The thesis has been confirmed by
examining the virbational spectra of polyethylene rubber [39] (linear polyethy-
lene cross-linked and kept above its melting point). The material shows hardly
any orientation when highly extended, but cooling the film without relaxation
produces highly oriented and crystalline material.
Although most of the investigation of the effect of strain on the vibrational
spectrum of elastomers has been confined to the infrared spectrum of natural
rubber films, more recently FT Raman results have appeared. Again see Fig-
ure 7.14. An analysis of the bands which alter under strain is in hand [40].
Similarly, work on butyl rubber (polyisobutylene) containing a small concentra-
tion of a diene and cross-lined shows that the spectrum changes dramatically as
the rubber is strained. See Figure 7.15. There is, however, a persistent experimen-
tal problem in this type of investigation. FT Raman study is far more convenient
and relevant than infrared because the sample does not have to be restricted to
a thin film, and any stretching rig can simply and easily be mounted in the sample
area of the instrument. Unfortunately, however, the near-infrared laser radiation
Stretched

Relaxed
intensity /arbitrary units

Frequency /wavenumber

Stretched

Relaxed
Intensity/arbitrary units

Frequency /Wavenumber

Figure 7.14 The FT Raman measurements on crosslinked natural rubber stretched and
relaxed and recorded in two ways as illustrated. Laser polarised vertically.
(i) Butyl Rubber, unstretched
Intensity

(ii) Butyl Rubber, stretched

Raman shift / wavenumbers

Figure 7.15 Raman spectra of butyl rubber

focused into the specimen causes heating and hence atypical stress: strain
patterns around the sampled point. The problem is simply illustrated in that
changes to the spectrum induced by strain may be apparent only if low laser
powers are used. Although experimentally most restricting, one excellent method
of avoiding the problem is to stretch and then cool in a cold cell. Clearly, novel
cold cell/stretching facilities need developing, but fortunately pioneering work in
this field was completed many years ago by Downes [42], who recorded
conventional Raman spectra on elastomers strained near their Tg. The pro-
ceedures used have been refined and are reported in Ref. 40.

7.5 CONCLUSION
Quite clearly, the advent of Fourier transform infrared and Raman methods and
the extension to dynamic or time resolved processes have already produced
a whole raft of new results and will certainly continue to do so. The measurement
of orientation and the structural changes that occur to specimens when stressed
in a variety of ways are bound to be studied in detail, if for no other reason than
that both FTIR and Raman spectroscopies are versatile, simple to apply and
rapid.
At a conference held recently, Everall [43] described a newly developed Raman
system involving fibre optical coupling between the optical system and the
spectrometer, and showed how it could be used to monitor and control on-line
the commercial production of polyethylene terephthalate film. Thus, the exten-
sion of these methods into commercial control is upon us.

7.6 REFERENCES
[1] W. Glenz and A. Peterlin, J. MacromoL ScL, Phys., 1970, B4,473.
[2] M.A. McRae and W.F. Maddams, J. Appl. Polym. ScL, 1978, 22, 2761.
[3] W.F. Maddams and J.E. Preedy, J. Appl. Polym. ScL, 1978, 22, 2721.
[4] W.F. Maddams and J.E. Preedy, J. Appl Polym. ScL, 1978,22, 2739.
[5] W.F. Maddams and J.E. Preedy, J. Appl. Polym. ScL91978, 22, 2751.
[6] A. Keller and MJ. Machin, J. MacromoL ScL Phys., 1967, Bl, 41.
[7] J. Purvis, D.I. Bower and LM. Ward, Polymer, 1973,14, 398.
[8] J. Maxfield, R.S. Stein and M.C. Chen, J. Polym. ScL, Polym. Phys. Ed., 1978,16,37.
[9] RJ. Roe and W.R. Krigbaum, J. Appl. Phys., 1964,35, 2215.
[10] D.Y. Yoon, C. Chang and R.S. Stein, J. Polym. ScL, Polym. Phys. Ed., 1974,12,209.
[11] M. Kashiwagi, A. Cunningham, AJ. Manuel and LM. Ward, Polymer, 1973,14,111.
[12] J. Purvis and D.I. Bower, J. Poly. ScL, Polym. Phys. Ed., 1976,14, 1461.
[13] J. Purvis and D.I. Bower, Polymer, 1974,15, 645.
[14] M.E.R. Robinson, D.I. Bower and W.F. Maddams, J. Polym. ScL, Polym. Phys. Ed.,
1978,16,2115.
[15] D.I. Bower, J. King and W.F. Maddams, J. MacromoL ScL Phys., 1981, B20, 305.
[16] PJ. Hendra and P. Bentley, Spectrochim. Ada, Part A, 1995, in press.
[17] I. Noda, A.E. Dowrey and C. Marcott, Appl. Spectrosc, 1988, 42, 203.
[18] B. Chase and R. Ikeda, Appl. Spectrosc, 1993, 47, 1350.
[19] MJ. Smith, CJ. Manning, R.A. Palmer and J.L. Chao, Appl. Spectrosc, 1988,42,546.
[20] CJ. Manning, R.A. Palmer and J.L. Chao, Rev. ScL Instrum., 1991,62,1219.
[21] V.G. Gregoriou, M. Dawn, M.W. Schauer, J.L. Chao and R.A. Palmer, Appl.
Spectrosc, 1993,47,1311.
[22] R.A. Palmer, J.L. Chao, R.M. Dittmar, V.G. Gregoriov and S.E. Plunkett, Appl.
Spectrosc, 1993,47,1297.
[23] C. Marcott, E.A. Dowrey and I. Noda, Appl. Spectrosc, 1993, 47,1324.
[24] AJ. Turner and R.A. Hoult, Poster presented at the 9th International Conference on
Fourier Transform Spectroscopy, Calgary, August 1993.
[25] R. Bennett, Spectrochim. Ada, Part A, 1994,5OA, 1813.
[26] I. Noda, J. Am. Chem. Soc, 1989, 111, 8116.
[27] I. Noda, Appl. Spectrosc, 1990, 44, 550.
[28] I. Noda, A.E. Dowrey and C. Marcott, Makromol. Chem., Makromol. Symp., 1993,
72,121.
[29] V.G. Gregoriou, I. Noda, A.E. Dowrey, C. Marcott, J.L. Chao an R.A. Palmer, J.
Polym. ScL, 1993, B31.
[30] H.H. Song, D.Q.Wu, B. Chu, M. Satkouski, M. Ree, R.S. Stein, and J.C. Phillips,
Macromolecules, 1990,23, 2380.
[31] R.A. Palmer, V.G. Gregoriou and J.L. Chao, Polym. Prepr., 1992, 33(1), 1222.
[32] K. Saijo, S. Suehiro, T. Hashimoto and I. Noda, Polym. Prepr. Jpn., 1989,38,4212.
[33] S.K. Dirtikov and J.L. Koenig, Appl. Spectrosc, 1979, 33, 555.
[34] R. Hasegawa, M. Sakamoto and H. Sasorki, Appl. Spectrosc, 1993, 47,1386.
[35] I. Noda, Appl Spectrosc, 1993,47,1329.
[36] T. Nakano, S. Shimada, R. Saitoh and I. Noda, Appl Spectrosc, 1993,47,1337.
[37] C. Jones, Spectrochim. Ada, Part A, 1991, 47A, 1313.
[38] PJ. Hendra, M.A. Taylor and H.A. Willis, J. Polym. ScL, 1986, 24, 83.
[39] PJ. Hendra, T.H. Stevenson, W.F. Maddams, V. Zichy and M.E.A. Cudby, Plast.
Rubber, Process. Appl, 1990,14, 7.
[40] A.M. Healey, PJ. Hendra and Y.D. West, Spectrochim Ada, 1995, 51A in press
[41] PJ. Hendra and P. Bentley, Spectrochim. Acta, in press.
[42] J.B. Downes, MPhil Thesis, University of Southampton 1972.
[43] N. Everall, ICI Wilton Research Centre, Personal communication; J. Andrews,
Spectrosc Eur., 1995,7(8), 8.
[44] I. Noda, A.E. Dowrey and C. Marcott, Appl Spectrosc 1993,47, 1317.

You might also like