Sunteți pe pagina 1din 28

8 DEFORMATIONSTUDIESOF

POLYMERS USING RAMAN


SPECTROSCOPY
R. J. YOUNG
Manchester Materials Science Centre, UMIST/University of Manchester,
Manchester Ml 7HS, UK

8-1 INTRODUCTION
Developments in the area of the application of Raman spectroscopy of polymers
have been covered elsewhere in this book [1] and this chapter is concerned with
a specific application, the use of Raman spectroscopy for the characterization of
the deformation of a wide range of both polymers and polymer-based compos-
ites. It is found that the wavenumbers Av of the bands in the Raman spectra of
some of these materials change under the action of stress a or strain e. This change
in the band position is characterized by either dAv/d(T or dAv/de, which can be
related directly to the deformation of the individual bonds in the materials.
There are several reasons for the upsurge of interest in the use of Raman
spectroscopy for deformation studies. There have recently been significant
improvements in the hardware available for Raman spectroscopy. The introduc-
tion of array detectors has enabled a particular region of a spectrum to be
acquired simultaneously rather than stepping through single points. This reduces
drastically the time required to obtain a spectrum. A significant development in
this area has been the introduction of highly sensitive charge coupled device
(CCD) cameras [2], Using a typical Raman microprobe system incorporating
a CCD detector, the spectrum of a Kevlar fibre consisting of a single band in the
region of 1580-1640Cm"1 can be obtained from areas of the order of 2jim in
diameter in a few seconds using a low-power 10 mW He-Ne laser [3]. A typical
full spectrum for a fibre of Kevlar 49, made up by joining several such single-band
spectra, is shown in Figure 8.1.
More recently there have been reports [1,4, 5] of the development of Fourier
Transform (FT) Raman systems which have certain advantages over conven-
tional dispersive systems. FT Raman systems have been described in Chapter 7,
and are potentially highly efficient, with a high calibration accuracy. One
particular advantage for polymers is that the use of IR excitation wavelengths
Polymer Spectroscopy. Edited by Allan H. Fawcett
© 1996 John Wiley & Sons Ltd
Intensity (Arbitrary Units)

Wavenumber / cm"1

Figure 8.1 Raman spectrum for a single filament of Kevlar 49 in the region 1100-
1700cm"* obtained using a low-power He-Ne laser (after [38])

greatly reduces the problems of fluorescence. Unfortunately, however, the inten-


sity of the Raman scattering is reduced by an order of magnitude by the use of IR
radiation rather than excitation in the visible region because the scattering
intensity depends on the fourth power of the excitation frequency, vjj. The choice
between the use of a dispersive or a FT system depends upon the scattering
characteristics of the material under investigation and the type of information
being sought. Another significant advance is the development of the Raman
microprobe into a true Raman microscope, reported recently by Batchelder and
coworkers [6,7], whereby images of the Raman scattered light may be obtained
and spectra may be recorded under confocal conditions. There is clearly plenty of
scope for the use of these new systems in the analysis of polymers, and there is no
doubt that there will be many such reports in the years to come.

8.1.1 POLYDIACETYLENE SINGLE CRYSTALS

The first reported and one of the best examples of the use of Raman spectroscopy
to follow deformation in polymers is the case of substituted polydiacetylene single
crystals [8-12]. The macroscopic polymer crystals are produced by the solid-
state polymerization of substituted diacetylene single crystal monomers. The
reaction is a topochemical solid-state polymerization [13], and the crystals
produced have a high degree of perfection [14].
Cottle et al. [15] reported in 1978 that the imposition of hydrostatic pressure
caused an increase in frequency of the four Raman-active vibrational modes in
poly(2,4-hexadiyne-l,6-bis(p-toluenesulphonate) (polyTSHD) [14]. They related
these shifts to the decrease in the separation between the carbon atoms on the
polymer backbone. A perhaps more significant discovery had been reported
a year earlier by Mitra et al. [8], who found that there was an approximately
linear decrease in two vibrational frequencies of single crystal fibres of
poly(bisphenylurethane-2,4-hexadiyne-l,6-diol) (polyPUHD) [14] subjected to
a tensile stress. They showed that this decrease could be explained in terms of the
anharmonicity of the bonds between the carbon atoms on the chain. This study
has been followed by a series of measurements of the dependence of the vibration
frequencies upon stress and strain for several different substituted polydiacety-
lenes [9-12]. Batchelder and Bloor [9] performed an elegant series of experi-
ments upon cleaved single crystal fibres of polyTSHD fixed to the aluminium
jaws of a small micrometer-driven straining rig by an epoxy resin adhesive. They
showed that care had to be taken to obtain an accurate estimate of the strain in
relatively short fibres during deformation, and suggested that the exact values of
strain reported by Mitra et al. [8] may be in error.
The dependence upon strain of the wavenumbers for the Raman modes of
several different substituted polydiacetylene single crystal fibres has been meas-
ured by various groups of workers [10,12,14]. Most attention has been paid to
the behaviour of the V1 mode which is essentially the symmetrical stretching mode
of the C = C triple bond, as this is the most sensitive to applied strain. The
dependence of the wavenumber of this band upon applied strain for a polyTSHD
single crystal [11] is shown in Figure 8.2, and the dependence of the position of
this Raman band upon strain for four polydiacetylenes [14] with different
Wavenumber /cm*1

Strain (%)

Figure 8.2 Dependence of the wavenumber of the C = C Raman band upon strain for
a polyTSHD single crystal (after [H])
Table 8.1 Values of the strain-induced Raman band
shifts for the C = C stretching band of different sub-
stituted polydiacetylene single crystals
Polymer [14] dAv/de(cm" 7%) Reference
polyPUHD ^20 [8]
polyTSHD -20.3 ±0.5 [9]
polyDCHD -19.7±0.4 [18]
polyEUHD -18.8 ±0.4fl [10]
-19.0 ±0.3»
"Phase 1
fc
Phase 2.

side-groups is summarized in Table 8.1. It is important to note that the rate of


change of the wavenumber of the band with applied strain, dAv/de, for the C=C
stretching V1 mode is « — 20 cm " 7 % and is virtually identical for all poly-
diacetylene single crystals, even though they may have different modulus values
[14]. This is because the structure of the backbone is very similar in each case and
the shift of the Raman band wavenumber with applied strain is a function only of
the backbone structure, whereas the Young's modulus depends upon both the
properties of the backbone and the size of the side-groups [12,16, 17].

8.1.2 EXTENSIONOFTHETECHNIQUETO
OTHER MATERIALS

Over the 15 years since the original Raman deformation studies upon poly-
diacetylene single crystals, the technique has been developed and refined to
involve the study of a wide range of different high-performance polymers and
other materials. These have included rigid-rod polymer fibres [19-21], carbon
fibres [22-24] and ceramic fibres [25-27]. This present chapter will concentrate
upon recent research concerning the use of Raman spectroscopy to follow the
deformation of aramid fibres and gel-spun polyethylene fibres and the possibility
of the extension of the technique to isotropic polymers, and also the important
and developing application of the method to the study of the deformation of
fibres within composites.

8.2 HIGH-PERFORMANCEPOLYMER FIBRES


8.2.1 AROMATICPOLYAMIDEFIBRES
Aromatic polyamide (aramid) fibres are produced by spinning from liquid
crystalline solutions using solvents such as sulphuric acid [28,29]. The properties
of the fibres can be improved by a brief heat treatment under tension at elevated
temperatures [29]. Fibres of aramids with high levels of stiffness and strength
have been available commercially for several years and they are now used, under
trade names such as Kevlar (Du Pont) and Twaron (Akzo), in a variety of
applications, particularly in the areas of fibre-reinforced composites, protective
clothing, tyre cord and ropes [30]. Original materials such as Kevlar 49 were
produced with values of filament modulus of the order of 120 GPa, but recent
developments have led to fibres with higher degrees of crystallinity such as Kevlar
149, which has a modulus of 170GPa [29].
The first report of the Raman spectrum of an aramid fibre was made in 1979 by
Penn and Milanovich [31], and a typical spectrum for a single filament of Kevlar
49 was given in Figure 8.1. Aramid fibres generally give strong Raman scattering
with little fluorescence, and well-defined spectra can be obtained from single
filaments using low-power laser beams [32-38]. The structure of the most widely
used aramid, poly(p-phenylene terephthalamide) is shown below.

It has been pointed out [30] that the C-N bond has considerable double bond
character which, along with the para-substituted phenyl group, leads to a reson-
ance conjugated system giving restricted rotation about the bond. These factors
help to give rise to liquid crystalline solutions, which allows the formation of
highly oriented fibres. They also lead to the characteristic yellow coloration of
aramid fibres and probably increase the intensity of the Raman scattering due to
resonance enhancement [ H ] . Penn and Milanovich [31] were able to assign
some of the Raman bands to vibrations of structural groups in the poly(p-
phenylene terephthalamide) molecule by comparison with model compounds,
and the strong band at «1610cm" 1 has been assigned to stretching of the
p-phenylene ring.
Penn and Milanovich [31] looked at the effect of deformation upon the Raman
spectrum of Kevlar 49 and, although they found that stress caused a change in the
relative band intensities and depolarization ratios, they did not find any measur-
able shift in the band wavenumbers with stress [31]. This finding is completely at
odds with the findings of Young and coworkers [32, 33], who have shown that
significant and measurable frequency shifts of several Raman bands in Kevlar
take place on the application of stress or strain.
It is found that the position of the Kevlar 1610cm"1 Raman band shifts to
lower frequency under the application of a tensile stress, as shown in Figure 8.3.
The peak position of the band is plotted as a function of strain in Figure 8.4 for
Intensity (Arbitrary Units)

Wavenumber / cm*1

Figure 8 3 Shift in the position of the 1610 cm ~~ * Raman band for Kevlar 49 at different
levels of tensile strain (indicated) (after [38])
Peak Position, Av / cnv1

Fibre Strain, e / %

Figure 8.4 Variation of the peak position of the 1610cm"1 Raman band with tensile
fibre strain for the five different aramid fibres (after [38])

two different commercial fibres and three experimental fibres [38], and it can be
seen that there is an approximately linear shift in peak position Av with strain e up
to failure at « 3 % strain. The slope of the line is given by dAv/de, and it is
normally found [33, 37, 38] that for different aramid fibres the slope is propor-
Rate of Band Shift, dAv / de cm 1 / %

Fibre Modulus, E 1 /Gpa

Figure &5 Dependence of the rate of shift per unit strain for the 1610cm * Raman band
upon fibre modulus for the five different aramid fibres shown in Figure 8.4. The symbols
have the same meaning as in Figure 8.4 (after [38])

tional to the fibre Young's modulus £ f , i.e.


dAv/deocEf (1)
Figure 8.5 give a plot of dAv/de versus fibre modulus £ f , and it can be seen that
there is an approximately linear relationship consistent with Equation (1). This is
an indication that the Raman technique is probing molecular stretching directly
[37] and is consistent with the theoretical work of Northolt and coworkers
[39-41]. They suggested that, for the deformation of aramid fibres such as
Kevlar, the total strain is the sum of the strains due to two deformation processes,
stretching and rotation, such that

^total = ^stretch + ^rotation w

It has been shown [37] that the Raman technique follows only the crystal and
molecular stretching and, moreover, that the change in the peak position dAv is
a measure of the molecular stress rather than the molecular strain [37]. The rate
of shift per unit strain in the higher modulus fibres is a reflection of the higher
levels of molecular stress in such fibres.
Penn and Milanovich [31] expressed surprise that they did not find any
frequency shifts during the deformation of aramid fibres, bearing in mind the
earlier polymer deformation studies using infrared [42] and Raman spectroscopy
[8] and wide-angle X-ray scattering (WAXS) studies of deformed aramid fibres
[39]. The situation was confused further by the publication by Edwards and
Hadiki [34], who reported that no Raman band shifts were found when Kevlar
fibres were deformed. They claimed that their work supported the findings of
Penn and Milanovich [31] and disputed the findings of Young and coworkers
[32, 33]. The situation was eventually sorted out by a careful study by Young
etal. [35], who repeated the experiments of both groups. One significant
difference was that Penn and Milanovich [31] and Edwards and Hadiki [34] had
both used a relatively high-power argon ion laser, whereas Young and coworkers
[32,33] had used a low-power He-Ne laser. It was demonstrated that the 488 nm
line of argon ion lasers caused significant radiation damage to Kevlar fibres and
led to the fibres fracturing at very low stresses and strains [35]. Hence the groups
using argon ion lasers [31,34] had damaged their fibres, and careful inspection of
their data showed that they had not been able to apply very high strains, at least
not large enough to cause significant Raman band shifts. In contrast it was shown
that the 632 nm line of the He-Ne laser produced virtually no radiation damage,
and so high strains could be applied and significant band shifts obtained [35].
Hence the dispute was resolved.

8.2.2 POLYETHYLENE FIBRES

There has been considerable interest recently in preparing highly oriented fibres
of polyethylene with very high values of stiffness and strength [43-45]. The
techniques employed to achieve this end include ultra-drawing [43], solid-state
extrusion [44] and gel drawing or spinning [45]. It has been possible by gel
processing of ultra-high molar mass polyethylene to produce fibres with Young's
modulus values of up to 200 GPa and strengths in the range 2-5 GPa, and such
materials are now available commercially under the trade names of Spectra from
Allied Signal in the USA and as Dyneema from DSM in the Netherlands. These
values of stiffness and strength are close to the theoretical limits for polyethylene,
which are about 300GPa for the modulus [46] and 20-30GPa for the strength
[47]. Although real materials rarely have mechanical properties so close to their
theoretical limits, it is of considerable interest to know how structural features
control the mechanical properties of these high-modulus fibres and what factors
lead to this shortfall in properties. For example, it is found that fibres produced
under different conditions may have the same levels of orientation and degrees of
crystallinity but very different levels of stiffness and strength [45]. There must
clearly be structural differences in the fibres which lead to a difference in the
molecular deformation processes in the fibres.
It has been relatively difficult to devise experiments to follow the deformation
of these high modulus fibres on the molecular level. Wide-angle X-ray scattering
experiments upon fibres subjected to stress have been used extensively to monitor
strains in the crystalline regions of high-modulus polyethylene fibres [48, 49].
Such experiments can be relatively tedious, requiring long data-collection times
(unless synchrotron radiation sources are employed [50,51]), and the interpreta-
tion of the data often requires assumptions to be made about the state of stress in
the materials. It has been known for many years that the vibrational modes of
polyolefin molecules are affected by the application of stress, particularly in the
oriented state [42,52,53], and considerable effort has been put into studying the
effect of stress upon the infrared spectra of such materials. In general it has been
found that some infrared bands shift to lower frequency and increase in width,
often also accompanied by the development of an asymmetric tail. However,
difficulties are often encountered in obtaining infrared spectra from fibres, and it
has been recognized recently that the measurement of Raman spectra during the
stressing of highly oriented polyethylene fibres offers a unique opportunity to
follow the deformation of these materials on the molecular level [50-60].
It is found that well-defined Raman spectra can be obtained from highly
oriented polyethylene fibres when the polarization direction of the laser beam is
parallel to the fibre axis. Figure 8.6 shows a Raman spectrum in the region of
1080-1150 cm ~ l for a high-modulus gel-spun polyethylene fibre before and after
deformation [59]. It can be seen that there is a change in the position and shape of
the 1128 cm ~ * Raman band, which has been assigned to the symmetric stretching
of C-C single bonds [55]. It can be seen that, following deformation, the band
develops a low-frequency tail, and it is possible to fit the band to two Gaussian
curves. In fact, it has also been found that even before deformation the asymmetry
of the band can be fitted to two peaks [59,60]. The variation in the position of the
two peaks with strain is shown in Figure 8.7, and it can be seen that they both
move to lower frequency, although the smaller peak moves most rapidly. This
type of behaviour for polyethylene has been interpreted as being due to the
presence of two populations of molecules in the microstructure of the gel-spun
material. During deformation they experience different levels of stress. It is found
[59, 60] that the size of the rapidly moving peak and the rate of shift per unit
strain scale with the Young's modulus of the fibre, and hence it appears that the
changes in the Raman spectra are related to the presence of microstructural
features which give rise to the impressive mechanical properties of the fibres.
There has been some controversy in the literature concerning the exact form of
the stress-induced Raman shifts in polyethylene fibres. Some workers reported
non-linear peak shifts with significant broadening of the bands and the develop-
ment of a low-frequency tail [53, 55], whereas others apparently found linear
peak shifts with little broadening [54, 56]. There is no doubt that, in some
instances, the differences stem from differences in fibre structure due to variations
in manufacturing route and testing conditions. Band splitting is certainly more
apparent in the gel-spun polymer deformed rapidly and obtaining spectra using
a CCD camera. However, it appears that there is also a difference in the methods
of data interpretation. If the Raman band is fitted to a single function (e.g.
Gauss-Lorentz sum function [55]) then it is difficult to deal with any band
asymmetry. For example, Prasad and Grubb [55] reported that, although the
shift of the band peak fitted to a symmetrical function is not linear with applied
Intensity (Arbitrary Units)

Wavenumber/cm 1
Intensity (Arbitrary Units)

Wavenumber/cnr1
Figure 8.6 Raman spectrum in the region 1080-1150Cm"1 for an experimental high-
modulus gel-spun polyethylene fibre (after [59]). The Raman band has been fitted to two
Gaussian curves, (a) Fibre undeformed; (b) at a strain of 4.06%

stress, the shift of the centre of gravity of the full band is linear. Such problems are
overcome if the shifting Raman band is fitted to more than one function, i.e. if it is
assumed that the band splits into several components as in Figure 8.6. This
clearly implies that either the fibres must have a two-phase structure or there
must be different stresses on certain molecules in the structure. Prasad and Grubb
[55] interpreted the low-frequency tail of the 1063 cm" 1 C-C asymmetric
stretching band of their spectra as non-crystalline overstressed tie molecules, as
they did not observe the development of any tail in their WAXS peaks. However,
recent WAXS studies [57] has indicated that the Bragg peaks in gel-spun
NARROW PEAK
BROAD PEAK
Wavenumber/cm*1

Strain %

Figure 8.7 Variation in the position of the two peaks in the Raman spectra of the
high-modulus gel-spun polyethylene fibre in Figure 8.6 with strain (after [59])

polyethylene may also split into two during deformation. Clearly more work is
required to ascertain the relationship between Raman band movement and the
behaviour of WAXS peaks under stress, since this may enable a better insight to
be obtained concerning the relationship between crystalline and molecular
deformation processes.
Several models have been proposed to explain the mechanical properties of
high-modulus polymer fibres [54, 61-63], and they all assume that the fibres
contain some highly stressed molecules or crystals contributing to the high level
of modulus, with other molecules or crystals taking lower levels of stress. Wong
and Young [61] have shown that there is a bimodal distribution of stress in the
crystalline regions of gel-spun polyethylene fibres due to their two-phase micro-
structure, and have demonstrated that the molecular deformation behaviour can
be interpreted quantitatively using a parallel-series Takayanagi model. They
have shown that the Young's modulus of the crystalline regions increases with the
degree of chain extension, and that for the highest-modulus fibres may be close to
the theoretical modulus of polyethylene crystals. It appears that for the first time
Raman spectroscopy allows the deformation of the different components in the
structure to be determined directly [60, 61].
8.3 ISOTROPIC POLYMERS
Fina et al. [64] found that significant shifts in the wavenumbers of the Raman
bands in poly(ethylene terephthalate) with moderate degrees of orientation could
be obtained, and so the question arises as to whether or not measurable shifts in
Raman bands could be obtained using isotropic polymers which contain little or
no molecular orientation. In a little-quoted letter published in 1976, Evans and
Hallam [65] reported measurable shifts to lower frequency in the wavenumbers
of the bands in the Raman spectra of polymers such as polypropylene, polycar-
bonate, polystyrene and nylon 66 from samples which were presumably un-
oriented (although they did not give any details of specimen preparation). Shifts
of the order of 1 cm" * were obtained for specimens deformed up to the point of
specimen necking, but the exact values of shift were found to depend upon the
band in question [65]. They did not give any values of stress but, assuming a yield
stress of their polypropylene of % 30 MPa, then the shifts they measured
approach — 30 cm ~ VGPa, which is quite significant. However, the relatively low
yield stress of the material means that the magnitudes of the shifts will always be
relatively small. Their data for polypropylene showed that between the point of
yield and eventual specimen fracture the behaviour of the Raman bands is
relatively complex, and that the wavenumber increases, decreases or stays the
same for different bands. Unfortunately, they did not give sufficient detail of their
measurements to explain this effect. Evans and Hallam [65] pointed out the
significant advantages of the Raman technique over similar infrared measure-
ments; although a full publication of their work was promised, no record of its
publication has been found.
Over recent years a completely different approach has been adopted by Hu,
Day, Stanford and Young [66-69], who have shown that, through the syn-
thesis of specially designed copolymers, it is possible to prepare isotropic
polymers for which the deformation can be followed using Raman spectroscopy.
They have demonstrated that such materials can be used for the study of polymer
surface and interface deformation [68,69] and their work in this area is reviewed
below.

8.3.1 U R E T H A N E - D I A C E T Y L E N E COPOLYMERS

The approach that was adopted [66-69] was to prepare a series of urethane-
diacetylene copolymers in which polydiacetylene units are incorporated into
segmented copolyurethanes. It was shown in Section 8.1.1 that large strain-
induced Raman band shifts can be obtained during the deformation of poly-
diacetylene single crystal fibers [8-12]. In fact, it is found that the largest shift
measured so far ( — 20 cm"7% strain) is for the C = C triple bond stretching
band of polydiacetylenes (Table 8.1).
Polyurethanes constitute a versatile class of materials ranging from soft
elastomers to glassy resins, and are readily produced by a variety of processes in
many different forms. Fibres, films, bulk sheets and surface coatings can all be
formed either from solution or in bulk. Segmented copolyurethanes, because of
their phase-separated structures, are particularly attractive, as they enable the
combination of disparate polymer properties to be obtained within a single
material. Polydiacetylene single crystals are formed by the rapid solid-state
polymerization [13, 14] of substituted diacetylene monomer crystals on the
application of heat or radiation. It is possible to induce similar solid-state
reactions known as "cross-polymerization" in the diacetylene groups of repeat
units in certain copolyurethanes and copolyesters [70-72]. Cross-polymerization
within the crystalline diacetylene regions produces a network structure in which
the chains connecting the polydiacetylenes are analogous to the substituent
side-groups in the polydiacetylene single crystals[13,14]. In this way, linear
copolyurethanes containing phase-separated, diacetylene-containing domains
can be crosslinked in situ and transformed into insoluble and infusible materials
[73-75]. These previous studies were concerned only with elastomeric materials
formed via a two-stage solution process [73-75] at only one composition. The
more recent studies [66-69] have been concerned with the development of more
rigid copolymers formed by a relatively simple one-shot bulk polymerization
route.
The reactants used to prepare the copolyurethanes were 4,4'-methylene-
diphenylene diisocyanate (MDI), 2,4-hexadiyne-l,6-diol (HDD) and a poly-
propylene glycol (PPG400). It is possible to vary the structure and consequent
properties of the materials by varying the relative proportions of HDD and
PPG400. The exact details of the reaction conditions are give elsewhere [67],
and linear polymers can be produced with molar masses of the order of
10000 gmol" 1 . These linear segmented urethane copolymers are soluble in
a variety of solvents, and can therefore be processed into a variety of forms such as
surface coatings [68].
The structure of the material consists of diacetylene-urethane hard segments
A and polyether-urethane segments B. The development of this alternating
segmented structure results in phase separation owing to the incompatibility of
the chemically distinct segments A and B. Thermodynamic incompatibility
depends primarily on the interaction parameter between the diacetylene- and
polyether-based segments (determined by their intrinsic solubility parameters)
and their sequence lengths (degrees of polymerization). The development of
hydrogen bonding and potential crystallinity of the hard segments further
enhances the driving force for phase separation. The linear copolyurethanes thus
form as essentially a two-phase morphology consisting of rigid, highly hydrogen-
bonded hard segment domains (with a distribution of sizes) dispersed in a ductile
polyether-urethane phase. The formation of linear diacetylene-containing
copolyurethanes provides the distinct advantage, in subsequent applications, of
enabling the copolymers to be processed from solution. During or after removal
of solvent, the phase-separated copolymers can be rapidly crosslinked in situ
using heat or radiation, either of which causes cross-polymerization of diacety-
lene units within the hard segment domains.
In practice, however, the solid-state topochemical reaction involves many
chains packed within the hard segment domains, and the resulting cross-
polymerization occurs three-dimensionally as depicted in Figure 8.8. The dia-
cetylene-urethane hard segments are assumed to be crystalline and have fully
extended conformations in which the HDD unit is alUtrans, and the chains are

Figure 8.8 Schematic representation of the solid-state topochemical polymerization of


the diacetylene-urethane hard segments: (a) linear segmented block copolymer; (b) cross-
polymerized material (after [67])
staggered so that adjacent chains are linked by straight C = C • • • H — N hydrogen
bonds in both directions perpendicular to the urethane chain axes. It is found
[68] that these hard segment domains are organized in the form of spherulitic
entities, which are seen to be of the order of 1 ^m in diameter using transmission
electron microscopy.
The idealized structure in Figure 8.8, however, is unlikely to be totally
representative of the overall structure actually obtained for the hard segment
domains, although regions of such three-dimensional order must exist within
domains dispersed throughout the copolyurethanes in order to achieve overall
cross-polymerization. The formation of fully conjugated polydiacetylene (PDA)
chains within the phase-separated copolyurethanes produces dramatic colour
changes (white -• red -* deep purple) and transforms the copolymers into com-
pletely insoluble and infusible materials. The extent of cross-polymerization that
is achieved depends upon a number of factors such as the time and temperature of
heating, the concentration of hard segment domains and the degree of order
within the domains [67].
The relationship between chemical composition, structure and properties for
the copolymers has been described in detail elsewhere [68, 69]. In general it is
found that the glass transition temperature Tg and the Young's modulus E in-
crease with hard segment content and heat treatment temperature. It was found
that the material with the optimum composition and properties had a value of T%
of %80°C and a Young's modulus (isotropic) of « 1.7GPa, both of which are
typical of a conventional glassy polymer.

8.3.2 DEFORMATION STUDIES

A Raman spectrum for a sample of the cross-polymerized copolyurethane is


shown in Figure 8.9 and it can be seen that it has four main scattering bands [67,
68]. The spectrum is remarkably similar to that obtained from a polydiacetylene
single crystal fibre [ H ] . For the fibres, a strong spectrum is obtained only when
the fibre axis is parallel to the direction of polarization of the laser beam, whereas
it is found that the spectrum from the copolymer is identical for all orientations of
the sample [67,68]. This shows clearly the isotropic nature of the copolymer. It
was found [67,68] that the intensities of the bands in the Raman spectrum varied
with both the hard segment content and the heat treatment temperature and,
moreover, it was demonstrated that this variation could be used to follow the
cross-polymerization reaction. In particular, the C = C triple bond stretching
band at « 2090 cm ~ l is not present before cross-polymerization and is indicative
of the formation of the polydiacetylene chains in the structure.
Figure 8.10(a) shows the position of the C = C stretching Raman band for one
of the copolymers in the undeformed and deformed states [68]. Upon deforma-
tion there is a pronounced shift to lower frequency and a slight broadening of the
lntensity

Wavenumber

Figure 8.9 Full Raman spectrum for the cross-polymerized diacetylene-urethane


copolymer (after [67,68])

Raman band. This shift is a clear indication of stress transfer from the polyether-
urethane matrix to the diacetylene-urethane hard segments, and shows that this
stress is translated into direct deformation of the polydiacetylene chains.
Broadening of the bands suggests that a distribution of stresses is developed
within the non-uniform hard segment domains during deformation.
The dependence of the band position upon tensile strain is shown in Figure
8.10(b). It can be seen that it is approximately linear up to %1% strain with
a slope dAv/de of the order of - 5 c m " 7% strain, and is also completely
reversible on both unloading and reloading. The slope of the line in Figure 8.10(b)
is significantly lower than that for polydiacetylene single crystal fibres (Table 8.1)
but is comparable with that of high-performance fibres such as those based upon
aromatic polyamides [37, 38] or rigid-rod polymers [19-21].
It has been found that the relationship shown in Figure 8.10(b) can be used to
measure strain in a wide variety of situations. It is known that polydiacetylenes
absorb visible light very strongly, and so for the bulk copolyurethane the
spectrum is obtained only from material in the surface regions. Hence any strain
measurements will be only for surface material. Moreover, since it is possible to
focus the laser beam in the spectrometer to a spot of the order of 2 jim in diameter,
it is possible to obtain considerable spatial resolution.
Various examples of using these materials for surface strain mapping have been
presented in a recent publication [69]. The determination of stress concentra-
tions around defects such as holes or notches in a deformed plate of the
copolyurethane is shown in Figure 8.11. A circular hole and a notch of pre-
determined dimensions were accurately machined into a 3 mm thick specimen of
the copolymer. The specimen was deformed in tension in the Raman spec-
Intensity

Wavenumbor cm"1
Wavenumber/cnv1

Strain %

1
Figure 8.10 (a) Shift of the 2090 cm " Raman band with strain for the cross-polymerized
urethane-diacetylene copolymer (after [68]); (b) dependence of the peak position of the
2090 cm"1 Raman band upon strain; •,first loading; • ; unloading, O, reloading (after
[68])

trometer and the change in position of the C = C stretching band at 2090 cm" l
with copolymer strain (measured remotely with a resistance strain gauge) was
determined at different positions around the defects, as shown schematically
(inset) in Figure 8.11. The slope (dAv/de) of each line, relative to that for the
remote applied deformation data (open O) is proportional to the stress concentra-
tion at each position. For the hole, the stress concentration, as expected, is highest
at the equator (solid • ) and is essentially zero at the pole (solid • ) . For the notch,
Wavenumber / cm*1

Overall strain/%

Figure 8.11 The effects of stress concentrations on the peak position of the C = C Raman
band for a 3 mm thick cross-polymerized diacetylene-urethane plate deformed in tension.
The data were obtained from spectra obtained at the different positions indicated (after
[69])

the stress concentration (open O) increases sharply depending on the notch tip
radius and the distance from the tip. The results obtained for the various defect
geometries [69] show the stress concentration values measured by Raman
spectroscopy to be very similar to those determined from conventional stress
analyses [76].
The good solubility and adhesive characteristics of the diacetylene-containing
copolyurethanes make them attractive materials for use as surface coatings that
can be applied with controlled thickness to a variety of substrates. Subsequent
cross-polymerization, in situ, would then convert the coatings into crosslinked
materials with strain-sensitive properties that can be determined quantitatively,
in conjunction with the substrate, using Raman spectroscopy. To illustrate this
use [68], a solution of the copolyurethane was applied as a 0.05 mm coating to the
following substrates: (a) a sheet of highly crosslinked (non-diacetylene- contain-
ing) polyurethane resin, (b) an inorganic glass filament (% 25 ^m diameter) and (c)
a sheet of aluminium. The solvent was removed by evaporation and the coatings
were thermally treated at 1000C for 4Oh. The coated substrate specimens were
deformed in tension in a Raman spectrometer and the shift in the position (Av) of
the C=C triple bond stretching band was monitored as a function of the overall
specimen strain e. Excellent linearity between Av and e was obtained in each case.
The strain sensitivities of the three coated substrates determined from the slopes
of the plots are given in Table 8.2; the values near — 5 cm " 7% strain for dAv/de
Table 8.2 Strain sensitivities of the Raman
frequencies of the C = C triple bond stretch-
ing band for the diacetylene-urethane
copolymer and for the copolymer coated on
different substrates (after [68])
Substrate dAv/de (cm " 7%)
Pure copolymer 5.3 ± 0.4
Polyurethane 5.5 ± 0.4
Glass fibre 5.7 ±0.4
Aluminium 5.5 ± 0.4

are almost identical to that of the bulk sheet material. Clearly, these results
demonstrate that the copolymers can be used as coatings to monitor accurately
the deformation of a substrate, which is particularly useful if the substrate is of
complex geometrical shape or is not readily accessible for direct measurements.
As such, the polydiacetylene- containing copolyurethanes are shown to behave as
optical strain gauges.

8.4 COMPOSITES
The shifts in the peak position of the 1610cm" * aramid Raman band, shown in
Figure 8.4, can be used as calibration curves to monitor the deformation of fibres
in a composite under any state of stress or strain. Previous studies have shown
[77-81] that it is possible to map out the distribution of stress or strain along
a single short, discontinuous fibre in a low-modulus epoxy resin. This is described
in detail next.

8.4.1 SINGLE-FIBRECOMPOSITES

Figure 8.12 shows the distribution of strain along a single Kevlar 149 fibre in
a model single-fibre epoxy composite [38] calculated from the point-to-point
variation of the shift of the 1610cm" * aramid Raman band. Measurements were
taken at 20 \im intervals along the fibre for different levels of matrix strain em
ranging from 0% to 2.0% in intervals of 0.4%, and the curves drawn are best fits
to the experimental data. It can be seen that in the unstrained case (em = 0%)
there is no strain in the fibre. As em increases the strain in the fibre increases from
the fibre end up to a plateau value along the middle of the fibre. It is shown that
the strain in the central region of the fibre is approximately equal to the matrix
strain em for matrix strain levels up to 1.6%. This behaviour is qualitatively
identical to the distribution of fibre stress and strain predicted by classical
shear-lag theory [82,83], where it is found that, following the shear-lag analysis
Fibre Strain, e f / %

Distance Along Fibre, x / pm

Figure 8.12 Derived variation of fibre strain with distance along the Kevlar 149 fibre in
a single-fibre composite tensile specimen at different indicated levels of matrix strain em
(after [38])

of Cox [ 8 3 ] , the variation of tensile stress a in the fibre with distance along the
fibre x is given by

_ T cosh fli/2-x)1
1
'- ^-L cosh/?//2 J
where

where E1 is the Young's modulus of the fibre, em is the matrix strain, Gm is the
shear modulus of the matrix, A1 is the cross-sectional area of the fibre, r0 is the
fibre radius and R is the radius of the cylinder of resin around the fibre.
This behaviour is shown more clearly in Figure 8.13, in which the data points
from Figure 8.12 are fitted to theoretical curves calculated from Equation (3) at
matrix strain levels of 0.4%, 0.8% and 1.2%. The value of R was assumed to be
the half-width of the matrix resin bar, and the matrix shear modulus Gm was
calculated from the matrix tensile modulus and Poisson's ratio at the relevant
Fibre Strain, ef / %

Distance Along Fibre, x / fim

Figure 8.13 Derived variation offibrestrain with distance along the Kevlar 149fibrein
a single-fibre composite tensile specimen at different indicated levels of matrix strain em.
The curves arefittedusing Equation (3) (after [38])

matrix strains. It can be seen that, while the data points are a good fit to the
theoretical curves in the central region of the fibre, there is a slight deviation at the
fibre ends. This is due in part to the geometry of the cut fibre ends [84] and the fact
that some of the assumptions of the Cox analysis [83], such that the strain at the
end of the fibres is equal to zero, are not appropriate.
It can be seen from Figure 8.12 that at 2.0% matrix strain, when the failure
strain of the fibre is exceeded, fibre fracture occurs, with the fibre breaking into
three fragments. The strain increases from the fibre ends in each of the fragments
to a value equal to or less than the failure strain of the fibre, at which point the
fibre strain decreases. The distance from the fibre end to the point of maximum
strain defines the region IJl where the maximum length of a fibre fragment is
given by [85]:

where xy is the shear yield stress of the fibre-matrix interface, o> is the failure stress
of the fibre, /c is the critical length of the fibre, and d is the fibre diameter (2r0).
The fibre fragments initially over a large distance of between 200 and 500 ^m,
with the crack running at an angle to the fibre axis [86]. In this region the
distribution of strain is mapped along a fibrillar section of the fibre which is still
bonded to the matrix. The points of minimum strain along the fibre are recorded
where the laser is focused on the propagating crack at the surface of the fibre. The
use of Raman spectroscopy to study the fragmentation of aramid fibres in a resin
matrix has the advantage that the strain can be mapped at intervals of 20 jim or
less in order to define the fragmented regions. This may be compared with
conventional polarized light experiments [87], which do not clearly define the
fibrillar fractured regions associated with aramid fibres [86].

8.4.2 INTERFACIAL MICROMECHANICS

It will be shown finally that Raman spectroscopy may be used to compare the
interfacial properties of Kevlar fibres with different surface treatments [38].
Figures 8.14(a) and 8.14(b) show the variation of fibre strain with distance x along
the left-hand end of a sized and de-sized Kevlar 49 fibre respectively for matrix
strains ranging from 0% to 2.4%. The solid lines are a fit of the experimental data
using either asymmetric or logistic sigmoid functions with correlation coefficients
greater than 98%. The distribution of fibre strain, at matrix strain levels up to
1.2%, is similar to that shown in Figure 8.12 for the Kevlar 149 fibre, and is in
qualitative agreement with that predicted by classical shear-lag theory [83]. At
matrix strain levels greater than 1.6%, the fibre strain increases from the fibre end
at a slower rate than at lower matrix strains. This effect is clearly more
pronounced in Figure 8.14(b) for the de-sized fibre, where it is seen that the
transfer of stress from the matrix to the fibre is greatly reduced at the fibre end,
when em = 2.4%, compared with the strain in the sized fibre at the same level of
applied matrix strain.
It is possible to derive the variation of interfacial shear stress T with distance
x along the fibre from the data in Figures 8.14(a) and 8.14(b) using the relationship
[88]:

where T is the interfacial shear stress and (de/dx) is the differential of the variation
of fibre strain with position along the fibre. Figures 8.14(a) and 8.15(b) show the
derived variation of interfacial shear stress with distance x along the sized and
de-sized Kevlar 49 fibres respectively. At matrix strains up to 1.2% the interfacial
shear stress is a maximum at the fibre end (x = 0), decreasing to zero at a distance
x along the fibre equal to IJl. At matrix srains greater than 1.6%, the epoxy resin
exhibits plastic deformation, which leads to a reduction in the transfer of stress
from the matrix to the fibre. This is clearly shown in Figure 8.15(b), where the
Fibre Strain, ef / %

Distance Along Fibre, x / urn


Fibre Strain, ef / %

Distance Along Fibre, x / urn

Figure 8.14 Derived variation of fibre strain with distance along the left-hand end of
a Kevlar 49 fibre in a single-fibre composite tensile specimen at different indicated levels of
matrix strain em: (a) sized fibre; (b) de-sized fibre (after [38])
ISS1 T/MPa

Distance Along Fibre, x / ^m


ISS, x / MPa

Distance Along Fibre, x / |im

Figure 8.15 Derived variation of the interfacial shear stress (ISS) T with distance along
the left-hand end of a Kevlar 49 fibre in a single-fibre composite tensile specimen at
different indicated levels of matrix strain em using the data from Figure 8.14: (a) sized fibre;
(b) de-sized fibre (after [38])
Maximum ISS9 rmtx/MPa

Matrix Strain, em/%

Figure 8.16 Dependence of the maximum interfacial shear stress upon matrix strain for
the sized and de-sized Kevlar 49 fibres in an epoxy resin matrix (data taken from Figure
8.15). The horizontal dashed line represents the shear yield stress of the epoxy resin, along
with the scatter band of the measurements (after [38])

iterfacial shear stress at the fibre end is only of the order of 3-4MPa for an
applied matrix strain of 2.4%, compared with « 4 3 MPa for an applied matrix
strain of 1.2%.
The maximum interfacial shear stress values for the sized and de-sized Kevlar
49 fibres are shown in Figure 8.16 as a function of applied matrix strain. It is
clearly shown that the values of maximum interfacial shear stress rmax for the
de-sized fibres are less than those for the sized fibres at all levels of applied matrix
strain. It is shown that the interfacial shear stress for the de-sized fibres reaches
a maximum value of 43 MPa, which is close to the shear yield stress of the epoxy
resin matrix [77] indicated by the dashed line in Figure 8.16.

8.5 CONCLUSIONS
It has been demonstrated that Raman spectroscopy is not only a very powerful
technique for following the deformation of high-performance fibres and compos-
ites but is also of use in the study of the deformation of isotropic polymers. It has
been shown that the shifts in the Raman bands is related directly to deformation
of the bonds in the polymers, and so the technique offers a unique method of
following molecular deformation in polymers. The relationship between the shift
in a particular Raman band and stress or strain can be used to map deformation
in a wide variety of systems. For example, the point-to-point variation of strain in
high-performance fibres in a composite can be determined with a spatial resol-
ution of the order of 2 ^m, which enables the micromechanics of composite
deformation to be studied with a resolution and precision which were hitherto
unobtainable. Smart polymer coatings have been developed which allow strain
mapping in the substrates with a similar level of resolution. It is clear that the use
of Raman spectroscopy to follow deformation is in its infancy, and over years to
come there will be significant developments in instrumentation and in its
application to different materials which will allow further significant advances to
be made.

8-6 ACKNOWLEDGEMENTS

The author is grateful to a large number of colleagues and research workers in


Manchester who have helped with his work on the development of Raman
spectroscopy for the study of mechanical properties in materials. He is also
grateful to the Royal Society for support in the form of the Wolfson Research
Professorship in Materials Science.

8.7 REFERENCES
[1] P. Hendra and W.F. Maddams, Chapter 7, in "Polymer Spectroscopy", Edited by
A.H. Fawcett, John Wiley & Sons, Ltd., Chichester, 199.
[2] D.N. Batchelder, Eur. Spectrosc. News, 1988, 80, 28.
[3] RJ. Young, Chapter 6 in WJ. Feast, H.S. Monro and R.W. Richards (Eds.), Polymer
Surfaces and Interfaces II, John Wiley & Sons, Chichester, 1993.
[4] CH. Zimba, V.M. Hallmark, J.D. Swalen and J.F. Radbolt, AppL Spectrosc, 1987,
41,721.
[5] FJ. Purcell, Spectrosc. Int., 1990,1, 33; Spectroscopy, 1989,4, 24.
[6] D.N. Batchelder, C. Cheng and G.D. Pitt, Adv. Mater. 1991,3, 566.
[7] R.V. Sudiwala, C. Cheng, E.G. Wilson and D.N. Batchelder, Thin Solid Films, 1992,
210/211,452.
[8] V.K. Mitra, W.M. Risen, Jr., and R.H. Baughman, J. Chem. Phys., 1977,66, 2731.
[9] D.N. Batchelder and D. Bloor, J. Polym. ScL, Polym. Phys. Ed., 1979,17, 569.
[10] C. Galiotis, RJ. Young and D.N. Batchelder, J. Polym. ScL, Polym. Phys. Ed., 1983,
21, 2483.
[11] D.N. Batchelder and D. Bloor, Resonance Raman spectroscopy of conjugated
macromolecules, in R.J.H. Clark and R.E. Hester (Eds.), Advances in Infrared and
Raman Spectroscopy, Vol. 11, Wiley-Heyden, Chichester, 1984.
[12] G. Wu, K. Tashiro and M. Kobayashi, Macromolecules, 1989, 22, 188.
[13] G. Wegner, Pure AppL Chem., 1977,49, 443.
[14] RJ. Young, Polymer single crystal fibres, in LM. Ward (Ed.), Developments in
Oriented Polymers—2, Applied Science, London, 1987.
[15] A.C. Cottle, W.F. Lewis and D.N. Batchelder, J. Phys. C, 1978,11,605.
[16] C. Galiotis and RJ. Young, Polymer, 1983,24,1023.
[17] C. Galiotis, R.T. Read, P.H.J. Yeung, RJ. Young, LF. Chalmers and D. Bloor, J.
Polym. ScL, Polym. Phys. Ed., 1984, 22,1589.
[18] C. Galiotis, RJ. Young, P.HJ. Yeung and D.N. Batchelder, J. Mater. ScL, 1984,19,
1640.
[19] RJ. Day, LM. Robinson, M. Zakikhani and RJ. Young, Polymer, 1987, 28, 1833.
[20] RJ. Young, RJ. Day and M. Zakikhani, J. Mater. ScL, 1990, 25,127.
[21] RJ. Young and P.P. Ang, Polymer, 1992,33, 975.
[22] LM. Ribinson, M. Zakikhani, RJ. Day, RJ. Young and C. Galiotis, J. Mater. ScL
Lett., 1987, 6, 1212.
[23] C. Galiotis and D.N. Batchelder, J. Mater. ScL Lett., 1988,7, 545.
[24] Y. Huang and RJ. Young, J. Mater. ScL, 1994,29,4027.
[25] RJ. Day, V. Piddock, R. Taylor, RJ. young and M. Zakikhani, J. Mater. ScL, 1989,
24, 2998.
[26] X. Yang, X. Hu, RJ. Day and RJ. Young, J. Mater. ScL, 1992, 27,1409.
[27] X. Yang and RJ. Young, J. Mater. ScL, 1993, 28, 536.
[28] J.R. Schaefgen, Chapter 8, A.E. Zachariades and R.S. Porter (Eds.), The Strength and
Stiffness of Polymers, Marcel Dekker, New York, 1983.
[29] S.L. Kwolek, W. Memeger and J.E. Van Trump, in M. Lewin (Ed.), Polymers for
Advanced Technologies, VCH Publishers, New York, 1988, p. 421.
[30] D. Tanner, J.A. Fitzgerald, P.G. Riewald and W.F. Knoff, in M. Lewin and J. Preston
(Eds.), High Technology Fibers—Part B, Marcel Dekker, New York, 1989.
[31] L. Penn and F. Milanovich, Polymer, 1979,20, 31.
[32] C. Galiotis, LM. Ribonson, RJ. Young, B.E J. Smith and D.N. Batchelder, Polym.
Commun., 1985, 26, 354.
[33] S. Van der Zwaag, M.G. Northolt, RJ. Young, LM. Robinson, C. Galiotis and D.N.
Batchelder, Polym. Commun, 1987,28, 276.
[34] H.G.M. Edwards and S. Hadiki, Br. Polym. J., 1989,21, 505.
[35] RJ. Young, D. Lu and RJ. Day, Polym. Int., 1991,24, 71.
[36] RJ. Day, LM. Robinson, M. Zakikhani and RJ. Young, in PJ. Lemstra and L.A.
Klientjens (Eds.), Integration of Fundamental Polymer Science and Technology—2,
Elsevier Applied Science, London 1988, p. 571.
[37] RJ. Young, RJ. Day, L. Dong and W. Knoff, J. Mater. ScL, 1992,27, 5431.
[38] M.C. Andrews and RJ. Young, J. Raman. Spectrosc, 1993,24, 539.
[39] M.G. Northolt and JJ. Van Aartsen, J. Polym. ScL, Polym. Symp., 1978,58, 283.
[40] M.G. Northolt, Polymer, 1980,21,1199.
[41] M.G. Northolt and R. Van der Hout, Polymer, 1985, 26, 310.
[42] D.K. Roylance and K.L. Devries, J. Polym. ScL, Polym. Lett, 1971,9, 443.
[43] G. Capaccio, A.G. Gibson and LM. Ward, in A. Ciferri and LM. Ward (Ed.),
Ultra-High Modulus Polymers, Applied Science, London, 1979.
[44] A.E. Zachariades, WT. Mead and R.S. Porter, A. Ciferri and LM. Ward (Eds.),
Ultra-High Modulus Polymers, Ed. Applied Science, London, 1979.
[45] P. Smith and PJ. Lemstra, J. Mater. ScL, 1981,15, 505.
[46] L. Holliday and J. W. White, Pure Appl. Chem., 1971, 26, 545.
[47] AJ. Kinloch and RJ. Young, Fracture Behaviour of Polymers, Applied Science,
London, 1983.
[48] J. Clements, R. Jakeways and LM. Ward, Polymer, 1978,19, 639.
[49] K. Nakamae, T. Nishino and H. Ohkubo, J. Macromol. ScL, Phys., 1991 B30,1.
[50] DT. Grubb and JJ.-H. Liu, J. Appl. Phys., 1985,58, 2822.
[51] K. Prasad and DT. Grubb, J. Polym. ScL: Part B: Polym. Phys., 1990, 28, 2199.
[52] R.P. Wool and W.O. Station, J. Polym. ScL, Polym. Phys. Ed., 1974,12,1575.
[53] R.P. Wool, R.S. Bretzlaff, B.Y. Li, CH. Wang and R.H. Boyd, J. Polym. ScL, Polym.
Phys. Ed., 1986, 24,1039.
[54] K. Tashiro, G. Wu and M. Kobayashi, Polymer, 1988, 29,1768.
[55] K. Prasad and D.T. Grubb, J. Polym. ScL, Polym. Phys. Ed., 1989,27, 381.
[56] BJ. Kip, M.C.P. Van Eijk and RJ. Meier, J. Polym. ScL: Part B: Polym. Phys., 1991,
B29,99.
[57] J.A.H.M. Moonen, W.A.C. Roovers, RJ. Meier and BJ. Kip, J. Polym. ScL, Polym.
Phys. 1992,30,361.
[58] D.T. Grubb and Z. Li, Polymer., 1992,30, 2587.
[59] W.F. Wong, Ph.D. Thesis, Victoria University of Manchester, 1992.
[60] W.F. Wong and RJ. Young, J. Mater. ScL, 1994, 29, 510.
[61] W.F. Wong and RJ. Young, J. Mater. ScL, 1994, 29, 520.
[62] PJ. Barham and R.G.C. Arridge, J. Polym. ScL, Polym. Phys. Ed., 1977,15, 1177.
[63] A.G. Gibson, G.R. Davies and LM. Ward, Polymer, 1978,19, 683.
[64] LJ. Fina, D.I. Bower, and LM. Ward, Polymer, 1988, 29, 2146.
[65] R.A. Evans and H.E. Hallam, Polymer, 1976,17, 838.
[66] J.L. Stanford, RJ. Young and RJ. Day, Polymer, 1991,32,1713.
[67] X. Hu, J.L. Stanford, RJ. Day and RJ. Young, Macromolecules, 1992, 672.
[68] X. Hu, J.L. Stanford, RJ. Day and RJ. Young, Macromolecules, 1992, 684.
[69] X. Hu, RJ. Day, J.L. Stanford and RJ. Young, J. Mater. ScL, 1992, 27, 5958.
[70] G. Wegner, Makromol. Chem., 1970,134, 219.
[71] D. Day and J.B. Lando, J. Polym. ScL, Polym. Lett. Ed., 1981,19, 227.
[72] A.O. Patil, D.D. Deshpande, S.S. Talwar and A.B. Biswas, J. Polym. ScL, Polym.
Chem. Ed., 1981,19, 1155.
[73] M.F. Rubner, Macromolecules, 1986,19, 2114.
[74] R.S. Liang and AJ. Reiser, J. Polym. ScL, Polym. Chem. Ed., 1987, 25, 451.
[75] R.A. Nallicheri and M.F. Rubner, Macromolecules, 1991, 24, 517.
[76] J.G. Williams, Stress Analysis of Polymers, Longman, London, 1973.
[77] M.C. Andrews, RJ. Day and RJ. Young, Comp. ScL and Tech., 1993,48, 255.
[78] RJ. Young, C. Galiotis, LM. Robinson and D.N. Batchelder, J. Mater. ScL, 1987,22,
3642.
[79] RJ. Young and P.P. Ang, in I. Verpoest and F.R. Jones (Eds.) lnterfacial Phenomena
in Composite Materials '91, Butterworth-Heinemann, Oxford, 1991, pp. 45-52.
[80] H. Jahankhani and C. Galiotis, J. Compos. Mater., 1991, 25,609.
[81] M.C. Andrews, RJ. Day, X. Hu and RJ. Young, J. Mater. ScL Lett., 1992,11,1344.
[82] M.E. Cates and S.F. Edwards, Proc. R. Soc. Lond. A, 1984,395, 89.
[83] H.L. Cox, Br. J. Appl. Phys., 1952,3, 72.
[84] CF. Fan and S.L. Hsu, Macromolecules, 1989,22, 1474.
[85] A. Kelly, Strong Solids, Clarendon Press, Oxford, 1966, p. 130.
[86] A.N. Netravali, Z.-F. Li, W. Sachse and H.F. Wu, J. Mater. ScL, 1991,28,6631.
[87] L.T. Drzal, MJ. Rich and P.F. Lloyd, J. Adhes., 1982,16,1.
[88] A. Kelly and N.H. Macmillan, Strong Solids, 3rd edn., Clarendon Press, Oxford,
1986.

S-ar putea să vă placă și