You are on page 1of 34

Crystal Structure, Morphology, and Orientation of Polyesters 107

with the draw ratio) is attributed to the matrix surrounding the fibrils, the
crystals in the fibrils being considerably longer.
In the following, we consider characterization of the molecular confor-
mation and crystal structure of crystallizable polyesters in both the crys-
talline and amorphous solid state, and their morphology in the glassy state
and, to a lesser extent, when crystallized from solution, (unoriented) melt,
and the glass. We recently described the techniques of polymer characteriza-
tion in sonae detail, with examples of the results obtained with various types
of polymer s [6]. Here, primary attention is directed toward the appropri-
ate results for major commercial polyesters, poly(ethylene terephthalate)
(PET), poly(butylene terephthalate) (PBT), poly(ethylene naphthalate)
(PEN) and the newest member to be commercialized, poly(trimethylene
terephthalate) (PTT).
2. Conformation and crystal structure
The conformation of a polymer molecule means the relative physical ar-
rangement in space of its atoms and molecular segments as affected by both
intra- and intermolecular atomic interactions (e.g., covalent, hydrogen, and
van der Waals bonds). It is to be distinguished from the term configura-
tion (chemical or sequence structure, often used synonymously) [7]. For
the aforementioned polyesters, the chemical structure is normally known,
naost of the commercial polymer s being linear, with alter nating aliphatic
and aromatic residues. However, copolymers of PET and PEN are commer-
cialized, and for them and copolymer liquid crystalline polyesters, such as
poly(p-oxybenzoate/2,6-oxynaphthoate) (PpOBA/ONA) and PET/OBA,
the sequence distribution (randomness vs. blockiness) is of concern relative
to crystallizability, crystal structure, and properties. The effects of copoly-
merization are not discussed here.
The overall dimensions of the molecule, as characterized by the mean
square end-to-end distance (r
2
), or radius of gyration, R
9
, in either amor-
phous or crystalline regions, could be considered here, since they are de-
termined by the conformation; however, we postpone this discussion, con-
sidering first the local segmental conformation. In crystalline regions, this
is naost often determined by X-ray (or electron) diffraction, that is, the de-
termination of the shape of the molecule in the unit eeli. For studying the
molecular conformation in the amorphous regions and complementing the
diffraction studies for crystalline regions, infrared (IR) and Raman spec-
troscopy are the primary tools. They have been complemented in turn,
for characterization of conformations in amorphous regions, by techniques
such as differential scanning calorimetry (DSC), dynamic mechanical ther-
mal analysis (DMTA), nuclear magnetic resonance (NMR), and permeabil-
ity studies. For the polymer s of concern here, the naost extensive studies
have been directed at PET. Both diffraction and the spectroscopies rely
Handbook of Thermoplastic Polymers: Homopolymers, Copolymers, Elends, and Composites
Edited by Stoyko Fakirov
Copyright 2002 WILEY-VCH Verlag GmbH, Weinheim
ISBN: 3-527-30113-5
108 P. H. Geil
on traditional techniques, now supplemented by continued improvements
in Instrumentation and the application of computer-based calculations and
simulations.
The characterization of the unit eeli of a polymer crystal requires the
determination of the spatial positions of all of the atoms from one or more
molecular segments, each one physical repeat unit long, within the unit cell,
ie., the conformation of the physical repeat units and the relative positions
of the atoms. Application of standard X-ray diffraction techniques to poly-
mers is made difficult by (i) the complexity (number of atoms and low sym-
metry) of the unit cells, (ii) general lack of availability of macroscopic single
crystals permitting the application of X-ray single-crystal techniques, (iii)
disorder of various forms in the crystalline regions that can be produced,
and (iv) the C, H, N, O atomic composition (low scattering power) of most
polymer s. The application of standard X-ray crystal structure techniques
to lattice and unit cell determination of organic non-polymeric molecules
is well discussed, for instance, in [8]; although there are numerous texts
on crystal structure determinations for inorganic crystals, the problems in-
volved in the determination of the crystal structure of organic materials,
synthetic and biological, are much closer to those involved in the determi-
nation of the crystal structure of polymers.
X-ray unit cell characterization of polymers generally takes advantage
of two features relatively unique to macromolecular Systems, the helical
conformation in crystalline regions of most synthetic polymers and some
biological samples, and the ability to produce bers with an alignment of
the molecular (backbone) axes along the ber axis [9-13]. The latter iso-
lates the Bragg reflections from one of the axes of the unit cell (usually c)
along the "meridian" of the diffraction pattern, ie., parallel to the ber
axis (the "equator" is perpendicular to the ber axis), and the former some-
what simplifies the determination of the molecular conformation from that
pattern. The length of the physical repeat unit, number of chemical repeat
units per physical repeat, and thus the axial distance per chemical repeat
unit can be easily obtained from a "ber pattern". The helical structure
(number of turns of a continuous helix drawn through a given atom in a
chemical repeat unit per physical repeat unit) can be determined from the
shape of the ber pattern. Such knowledge is of major assistance in deter-
mining the conformation of the molecule in the unit cell and also allows
easy determination of the density of the perfect crystal if the unit cell di-
mensions are known. Thus, the density value needed for characterization
of the degree of crystallinity by density measurements does not require fll
knowledge of the unit cell, z.e., the location of all of the atoms within it.
However, all of the polyesters considered here are not normally considered
helical, the c-axis usually consisting of one or, at most, two chemical repeat
units.
Determining the relative positions of the atoms in the cell, along and
between the physical repeat units, is aided by knowledge of the chemical
Crystal Structure, Morphology, and Orientation of Polyesters 109
structure and typical bond lengths and angles, as determined from low
molecular weight "model compound" crystal structures. It is "confirmed",
in the end, by comparison of calculations of the structure factors or intensi-
ties of the various reflections and the experimental intensities. Details of the
determination of the helical conformation, if present, and the unit eeli are
discussed in several texts [9-14], that by Vainshtein [14] being particularly
recommended. However, as pointed out in [13], drawing, defects, crystal
thickness and, in particular, temperature can have a significant effect on
the unit eeli parameters. Increasing evidence summarized below for PET,
in particular, indicates that t her e is no unique "unit eeli" for a polymer and
that the size, shape, and molecular packing are functions of the processing
history.
IR and Raman absorption spectra result from molecular vibrations that
cause changes in the dipole moment (and polarizability) of the molecule;
these spectra are unique to each molecule. Although often used for config-
uration determinations, the spectra also depend on the conformation and
molecular packing, as these affect the intra- and inter molecular for ees, re-
spectively. For both IR and Raman, the normal mode vibrations of small
atomic groups, such as CH2 and C=O, are of concern, as affected by their
local environment. Conventional Raman spectroscopy, even though yield-
ing equivalently useful Information, has been used much less widely than
IR for the study of polymers (see [8,12]). Examples are shown below.
2.1. Poly(ethylene terephthalate)
2.1.1. Conformation in the amorphous phase
First, we consider the conformation (average) of the individual molecule in
the amorphous state and as it crystallizes during heating above the glass
transition temperature. With T
9
at ca. 65
0
C, crystallization occurs at ca.
UO
0
C when the heating rates are those used in DSC scans. In the early
196Os, based on electron microscopy and electron and X-ray diffraction
studies of ice-water-quenched thin and thick (1.6 mm) films of PET and
the effect thereon of deformation, we suggested that amorphous PET con-
tained domains in which neighboring segments have a nematic type packing
[15,16]. These results are discussed further in the morphology section. If
the domains of local order are present, one would expect a reasonable frac-
tion of the molecules to exist in a trans conformation (conformation of the
O-CH2-CH2-O moiety), similar to that in the crystal [17]. Thus, of con-
cern here are studies of the conformation of the molecules in as-prepared,
unoriented "amorphous" samples during annealing below the melting tem-
perature and near T
9
(physical aging) and during cold crystallization when
annealing well above T
9
(normally higher than ca. UO
0
C), as character-
ized by Fourier transform infrared (FT-IR) and FT-Raman spectroscopy.
A number of absorption bands have been used to determine the conforma-
tions of the ethylene glycol units, namely gauche (g) vs. trans (t) or more
110 R H. Geil
o 22O
0
C
* 20O
0
C
O 18O
0
C
16O
0
C
6 14O
0
C
100 r
120

c
10O
0
C
9O
0
C
75
0
C
80
60
40
1.0 2.0 3.0
Annealing time (log min)
4.0
Figure 1. Relative concentration of gauche conformers as a function of annealing
time at various temperatures [21]
1.43
1.39
22O
0
C
* 20O
0
C
D 18O
0
C
* 16O
0
C

14O
0
C
12O
0
C
* 10O
0
C
- c 9O
0
C
7 75
0
C
g
Q
. - -a g
0
1
0
__
0
-_ -- -
-'-" ,-UE^
1.35 f
9>
,o' ,' /
--" m ----*' *"'
... 1 1 1 1 1 1 L.
1.0 2.0 3.0
Annealing time (log min)
4.0
Figure 2. Density of PET films as a function of time at various annealing tem-
peratures [21]
Crystal Structure, Morphology, and Orientation of Polyesters 111
80
84 0

20
O
-O- - D-D1
D Gauche
Trans in amorphous
o Trans in crystalline
s
/
s
/
,/ N
0/
V.
- O -Q-O'
O 80 160 240
Annealing temperature (
0
C)
Figure 3. D istribution of rotational isomers as a function of annealing temperature
after 7 days of annealing [21]
recently, the gauche (g), trans (amorphous) (t
a
), and trans (crystalline) (t
c
).
They are listed below as used by the various authors. Bnd assignments,
in general, are described in, e.g., [18,19]. This is followed by a discussion
of the local conformation of the chains in oriented amorphous samples, the
t a content being of particular interest.
Moore et ai. [20] reported an increase in overall gauche content during
physical aging (annealing at 5O
0
C) of 0.025 mm thick samples originally
quenched from the melt to the "amorphous state", then heated to 85
0
C
and quenched at room temperature. Gauche bands used included 1042 and
896cm"
1
. In contrast to later reports, discussed below, they did not differ-
entiate between t
a
and t
c
content. On the other hnd, Lin and Koenig [21]
reported no change in isomer content when annealing glassy PET at 50 or
75
0
C for up to 38Oh; there was, however, an increase in t content when
it was annealed above T
9
. The 898 (CH
2
rock, g), 1454 (#), 973 (Q-CH
2
stretch, t), 1473 (t), and 793 (internal thickness) bands were used, with the
assignments hving been made by Miyake [22,23]. "Amorphous" films of
unstated processing history were used (p = 1.3402g/cm
3
), with subsequent
annealing of the samples between 75 and 24O
0
C for various times. The start-
ing films were said to have 14% trans isomer. Correlations with the density
permitted Separation of the three components, amorphous gauche (g
a
), t
c
,
and t
a
] their variations with annealing time and temperature are shown in
112 R H. Geil
o
3
1050. 994. 770.
Wavenumber (cm )
Figure 4. FT-Raman spectra of the surface of an injection molded, 3 mm thick
PET plaque: (a) as-molded and annealed at (b) UO
0
C, (c) 150C and (d) 205
0
C
for l h [26]
100
90
80
g 70
I
60
|50
40
S
3
O
20
10
O
_noo-a-
50 100 150 200
Annealing temperature (
0
C)
250
Figure 5. D istribution of trans and gauche conformers in PET based on Figure
4; gauche () and cry st alline trans (D) [26]
Crystal Structure, Morphology, and Orientation of Polyesters 113
Figures 1-3. Extrapolation of the density of the samples as a function of
content of the various isomers gave densities of 1.326 for all g, 1.510 for all
crystalline trans, and 1.430 for amorphous trans. The t
a
and g
a
contents
decreased rapidly above T
9
, with t
c
appearing at their expense. At 14O
0
C,
a change in slope of the t
c
and g
a
curves was observed and attributed to
an increase in mobility of the segments, permitting rearrangement to occur
by segmental motion rather than cooperatively in the vicinity of t isomers,
as at the lower temperatures. A similar change in slope, at 14O
0
C, of crys-
tallinity with annealing temperature had been shown earlier using X-ray
measurements [24,25].
Rodriguez-Cabello et ai have utilized FT-Raman and FT-IR to charac-
terize the molecular conformation of the amorpous segments in a semicrys-
talline, 3 mm thick, injection molded sample [26,27]. Figure 4 shows the
back-scattered Raman spectra in the region 750-105UCm"
1
; the peaks at
998 and 886Cm"
1
, respectively, are assigned to t (O-CH2 Stretch) and g
conformations of the ethylene glycol segments [26]. The 795cm"
1
band,
whose intensity is independent of thermal treatment, was used as an in-
ternal thickness band in order to normalize the various spectra for the
amount of material in the beam. In the IR studies, the corresponding bands
are those above, ie., 973, 898, and 793cm"
1
. From the intensity ratios of
the Raman bands, J(998)/J(795) and J(886)/J(795), the fractions of t and
g conformations of the flexible -O-CH2-CH2-O-segment can be obtained
(Figure 5). For the initial sample, the t fraction was ca. 25%. In agreement
100 200
Temperature (
0
C)
100
" I
(d)
t
200
Temperature (
0
C)
100 200 O
Temperature (
0
C)
100 200
Temperature (
0
C)
Figure
15O
0
C,
6. DSC scans of PET: (a) as-quenched and annealed at (b) 9O
0
C, (c)
and (d) 205
0
C for Ih [26]
114 P. H. Geil
50
4 5
4 0
35
O
30
25
50 100 150 200
Annealing temperature (
0
C)
250
Figure 7. Plot of t he DSC determined crystallinity of the injection molded PET
s function of annealing temperature (l h) [26]
with DSC results (Figure 6), crystallization (conversion of g to t conform-
ers) begins above T
9
for the annealing conditions at ~9O
0
C used here. The
initial bulk sample had a degree of crystallinity, s determined by DSC, of
~ 30% (Figure 7). With a portion of the t conformation known to occur in
the amorphous regions, the value of
0
is too large relative to the 25% t
conformer obtained by Raman spectroscopy. This "discrepancy", indicative
of the power of the technique, was attributed to a skin-core morphology
of the sample. The skin, being more rapidly quenched, is initially more
amorphous than the core, with DSC giving a bulk crystallinity value and
Raman smpling only, in this case, less than the 500 thickness observed
by light microscopy for the skin. Determination of the t conformer in a
glassy amorphous sample (~ 15%) permitted Separation of the distribu-
tion of g
a
, t
a
, and t
c
conformers s a function of annealing temperature
(Figure 8), with Figure 9 showing the (excellent) correlation between the
DSC-determined skin-layer crystallinity and the crystalline band content
determined by Raman spectroscopy.
As shown in Figures 5 and 8, in addition to the stepwise change at about
10O
0
C due to cold crystallization, there is a change in slope at ~ 14O
0
C,
similar to that described by Lin and Koenig [21]. Similar step changes and
changes in slope were seen, for instance, for the width of the 858 cm"
1
(com-
plex mode consisting of ring C-C and C(O)-O stretching) and 1725cm"
1
(C=O Stretch) Raman bands (see [26]). The change in slope at ~ 14O
0
C
was attributed to the development of sufficient segmental mobility to per-
mit further segmental rearrangement and an increase in crystallinity. These
Crystal Structure, Morphology, and Orientation of Polyesters 115
100
90
80
g70
l
60
|50
1
40
g30
20
10
O
_ _ O
-D Q D--

- -Cl,.

O 50 100 150 200
Annealing temperature (
0
C)
250
Figure 8. A corrected distribution of Raman-determined trans and gauche con-
formers in PET based on the trans conformer content in a "glassy amorphous"
PET sample. Amorphous trans (), crystalline trans () and gauche (D); note
that symbols are reversed in Figure 5 [26]
60
50
30
O
20
10
20 25 30 35
Trans (%)
40 45
Figure 9. Correlation of crystallinity of the skin layer as determined by DSC and
the Raman-determined crystalline trans fraction [26]
116 P. H. Geil
100
80
60
40
20
-

A
D
g.
50 100 150 200
Annealing temperature (
0
C)
250
Figure 10. Corrected distribution of IR determined trans and gauche conformers
in PET based on the trans conformer content in a "glassy amorphous" PET
sample. Amorphous trans (D ), crystalline trans (A), and gauche () [27]
3 mm thick, injection molded samples were annealed forl h in a preheated
oven and then cooled to room temperature priorto measurement.
The aut hr s of the Raman study reported similar results, using FT-IR
with photo-acoustic detection for surface characterization [27]. Figure 10
shows the fractional content of g
a
, t
a
, and t
c
on the surface of an injection-
molded sample s a function of annealing temperature, with the fraction of
t
a
being determined by comparison with total trans and the crystallinity
being determined by DSC. The 973 and 898 bands were used to characterize
the relative content of trans and gauche conformations, with 793 serving s
an internal reference band, using peak heights above baselines connecting
valleys between the peaks. The 973 band was resolved into a 975 (t
a
) and
973 (t
c
). The t
a
disappears at about 10O
0
C. There was only a mention [26] of
its possible relationship to a so-called mesomorphic or ordered amorphous
phase reported on the basis of X-ray patterns from oriented amorphous
samples, and this is discussed below. A change in slope of the trans and
gauche isomer content at 14O
0
C was again observed. The presence of a
small "fold band" (988Cm"
1
, see below) in the sample annealed at 205
0
C
was also indicated.
In a recent paper [28], Ruvolo-Filho and De Carvalho have reported
similar FT-IR studies, using the 898 (g) and 973 (t) band intensities, of
unoriented PET films annealed between 110 and 23O
0
C and correcting the
results for the absorption of acetaldehyde, which appears s a degradation
product during heating of PET above ca. 18O
0
C. This was an extension
Crystal Structure, Morphology, and Orientation of Polyesters
0.8
117
0.7-
0.6 -
0.5-
0.4-
0.3-
0.2-
100 120 140 160 180
Temperature (
0
C)
200 220 240
Figure 11. Fraction of impermeable (a
c
+ Oi
0
) (A), oriented amorphous ormeso-
morphic(a
0
) (B) and crystalline (a
c
) () as a function of annealing temperature
in samples annealed for 24 h [28]
of an earlier paper [29] characterizing the mesomorphicphase in samples
annealed between 110 and 18O
0
C by comparison of X-ray diffraction and
permeability studies. Using the technique proposed by Lin and Koenig
[21] to determine the fraction of

, i
c
, and g
a
bonds by combined FT-
IR and density measurements, they related the change in isomer content
during annealing to the amount of an "intermediate" phase, which was
determined by a correlation of X-ray and density measurements of crys-
tallinity and sorption measurements of the amount of material impermeable
to dichloromethane at low penetrant activity. The impermeable material
is considered to be composed of the crystals and a "mesomorphic" amor-
phous phase, a term originally used for amorphous, interfibrillar regions
in PET bers consisting of highly extended chains [30]. By comparing the
crystallinity of the samples with the t content, Rivolo-Filho and De Car-
valho were able to separate out from the fraction of impermeable phase, as
a function of annealing temperature (for 24 h), the fraction of mesomorphic
amorphous phase; the results are shown in Figure 11. A plot of the fractions
of the three isomer s, shown in Figure 12, suggests t hat annealing between
110 and 18O
0
C results in the conversion of g conformers to both t
a
and t
c
conformers, while heating above 18O
0
C results in crystallization of some of
the chains in the mesomorphic regions. The results are thus similar to those
of Rodriguez-Cabello et ai [26,27], except that the crystallization of the t
a
118 R H. Geil
0.9-
0.8-
0.7-
g 0.6 H

S 0.4-J
"" 0.3-
0.2-
0.1-
0.0.
J3-
D
80 100 120 140 160 180 200 220 240
Annealing temperature (
0
C)
Figure 12. Fractions of isomers in PET samples after 24 h annealing at various
temperatures. Trans amorphous (D ), trans crystalline (+), and gauche () [28]
conformers started at a highertemperature, possibly because of differences
in the initial sample preparation and method of annealing.
Shen and co-workers [31,32] have also used FT-IR to examine the con-
formational changes in PET during cold crystallization and physical aging,
using the 1340 (t) and 1370Cm"
1
(g) bands, said to show a greater change
in intensities [33] than the 973 and 898cm"
1
bands. Forthe physical ag-
ing study [31], they used solvent-cast films heated to 92
0
C and then either
quenched into ice-waterorslowly cooled to 67
0
C, after which both types of
samples were annealed at 67
0
C. The initially quenched film had a density
of 1.3356g/ cm
3
; after annealing at 67
0
C for60h, the density increased to
1.3372g/ cm
3
. The density of the slowly cooled films was initially 1.3366,
also increasing to 1.3372 after 60h at 67
0
C. The 1340 and 1370cm-
1
bands
in the quenched samples were initially of equal absorption intensity, the t
content decreasing from 9.3 to 8.8% during the physical aging (Figure 13);
changes in absorption of the 1370cm-
1
band were too small to measure.
On the other hnd, the t content in the slowly cooled sample increased
from an initial 8.2 to 8.7% with annealing. Wang et al. [31] suggest that
annealing results in closer interchain packing, with the development of new
"cohesional entanglements" [34] (attractive interactions resulting from par-
allel alignment of the chains), which act s physical crosslinks. On heating
through T
9
, gradual increases in the t content occurred for the quenched
Crystal Structure, Morphology, and Orientation of Polyesters 119
S

9.4-1
9.2 -
9.0 -
8.8 -
8.6 -
8.4 -
8.2 -
1000 2000 3000 4000
Anealing time (min)
5000 6000
Figure 13. Changes in the percentage of trans conformers as a function of sub- T
g
annealing time. (v) quenched and previously annealed at 67
0
C fortimes given,
(A) slowly cooled from 92 to 67
0
C and then annealed. Both films were initially
solution east and are said to be amorphous [31]
and slowly cooled samples, even though they originally differed consider-
ably in t content, whereas both annealed samples showed an abrupt change
at T
9
corresponding to the presence of an endothermicDSC peak (T
9
over-
shoot). In a much earlier paper [35], several of the same authors, using
the same t hermai treatment pius quenching a sample int o liquid nitrogen,
showed that the quenched sample had fewer g conformers than either the
slowly cooled or 67
0
C annealed samples. This was interpreted by the as-
sumption that the samples at 92
0
C (i. e. , at high temperatures) have a
higher t content which is frozen when quenched.
The effect of annealing quenched samples below and above T
9
on the
rate and temperature of cold crystallization, T
cc
, during isothermal and
non-isothermal crystallization, respectively, is reported in [32]. Original
films were solvent-cast, heated to 285
0
C and quenched in ice-water, then
annealed at 65, 75, and 85
0
C (T
9
= 72
0
C at 3C/ min heating rate) for
2-4 days before measuring the cold crystallization process by DSC. T
cc
decreased with increasing annealing time and temperature. The t content,
using the 1340 and 1370cm"
1
bands, increased with annealing time at all
three temperatures. The increase in t content with annealing below T
9
was
120 P. H. Geil
also reported earlier by Siegmann and co-workers [36,37]. The 1370/1340
absorption intensity ratio in the initial "amorphous" samples used by Zhang
and Shen [32] was about 2/3, while it was l/l in the "amorphous" samples
used by Wang et al [31]. Thus, the initial samples must differ, and the
effect of annealing below T
9
is in the opposite direction in the two papers
by Shen and co-workers, each hving publications by other authors agree-
ing with the results. In this paper, however, they also suggest that the
results should be interpreted in terms of an increase in parallel alignment
of the molecules, giving rise to nuclei that resulted in the decrease of T
cc
during non-isothermal crystallization and increase in crystallization rate in
isothermal crystallization [32].
Relative t o these apparently contradictory results, the papers by Hay
and co-workers [33,38] are also of interest. In the earlier paper [38], Aref-
Azar and Hay indicated that two conformations (mostly g with some ex-
tended t) of the O-CH2-CH2-O moiety can coexist in "completely" amor-
phous, unoriented PET. Using the ratio of the 896 (g) to 973 (t) absorp-
tion bands they state, but do not show, that the ratio decreased linearly
with time of annealing below T
9
, approaching the extrapolated ratio for an
"amorphous" sample measured above T
9
at a rate similar to that at which
the enthalpy decreased. Prom this, they conclude that there is "no evidence
for crystalline domains developing on physical aging". In the more recent
paper [33], it is reported, using absorption spectra and the same bands,
that their ratio increases during cooling and aging. It is suggested that the
original Interpretation was incorrect because transmission instead of ab-
sorption spectra were used; however the graph showing the change of the
893/973 band intensities in the first paper [38] is labeled in terms of the
band's absorption. These authors also confirmed the recent results, i.e., a
decrease in t relative to g during cooling through T
9
and during physical
aging, using the more sensitive 1340 and 137UCm"
1
bands; the latter were
of almost equal height in the original amorphous films. The recent results
are interpreted s suggesting that the t conformation of the glycol unit has
a higher energy and results in poorer packing in the amorphous regions.
In another fairly recent paper, Ajii et al. [39] suggested that, in samples
characterized s 5% crystalline by DSC, 10% exists in the t conformation.
They then followed the t and g content s a function of draw ratio at 8O
0
C,
using the 1340 and 1370Cm"
1
bands; s would be expected, the t content
increased s the g content decreased. With the t content always being larger
than the crystalline content, they suggested a t amorphous content.
What is the origin of the apparently contradictory results from the
various authors, including contradictions involving the same authors? In
our opinion, at least in part, none of the papers have utilized "wholly"
amorphous PET films; s will be shown later in the section on morphology
of "amorphous" PET, ice-water quenching from the melt is insufficient
to produce wholly amorphous, structureless films, "ultraquenching" into
liquid nitrogen or isopentane at their melting point being required. Ice-
Crystal Structure, Morphology, and Orientation of Polyesters
100
b
(a)
121
(b)
60
Glass spectrum
MeIt spectrum
Melt-crystallized
Solution-crystallized
4 0
20
1000 920
988 cm"
1
region in various PET samples
1000 950 900
Figure 14. (a) 988 cm"
1
IR band in various PET samples. (b) IR spectra (900-
1000 cm"
1
) of an amorphous PET sample (I) stretched 430%, then (II) annealed
taut at 14O
0
C, 30 min, then (III) restretched 50% and then (IV) reannealed at
14O
0
C, 30 min [41]
water quenching results in a domain structure proposed by us on the basis
of dark field electron micrographs, which has an aligned or nematic liquid-
crystal-like packing of the chain segments [15,4O].
With IR spectroscopy being used to characterize segmental conforma-
tion, an early paper by Koenig and Hannon [41] is also of interest. These
authors used FT-IR to characterize molecular folding in solution- and melt-
crystallized samples. They attributed the 988Cm"
1
(and 1380 cm"
1
) bnd
to the (tight) gauche conformation of an adjacent reentry fold (thus not g
a
).
With the 973 cm"
1
bnd as a measure of the degree of crystallinity (t
c
con-
tent) and the 793cm"
1
band as an internal thickness bnd, they followed
the changes in numbers of tight folds and crystallinity with thermal treat-
ment in both molded and drawn film samples. As shown in Figure 14a,b,
the number of tight folds is zero in the melt and in quenched amorphous
samples, is near zero in as-drawn films, but increases with annealing in
both types of samples. The absence of the 988cm"
1
band does not neces-
sarily indicate the absence of folds; its presence was interpreted as being
due to the particular sequence of gauche conformations in a tight, adjacent
122 P. H. Geil
(a) 50 PET
s = 0.20 A-
1
-
c- s = 0.26 A-
(b) 550
s = 0.26 A-
Figure 15. Peak resolut ion of X-ray scans of molded (unoriented) "amorphous"
PET films of two thicknesses [42]
reentry fold.
Although studies of molecular conformation by X-ray diffraction are
normally applied to the conformation in the crystal, limited application
has been made t o the conformation in the amorphous regions, with much
of the reported work being on PET. A decade ago, Murthy et al. [42] indi-
cated that the amorphous scattering, between 5 and 3520, from various
types of "amorphous" PET samples, including unoriented films, consisted
of two "peaks" at s (= 48#/ ) = 0.20 and 0.26. They suggested these
were due to two average distances between neighboring molecularsegments.
The relative height of the two peaks varied with the "amorphousness" of
the samples, s determined by the thickness of the films, the peak at smaller
s being the largest in the thinnest film (Figure 15). In the crystal structure
proposed by de Daubeny et al. [17], the 010 and 100 planes have values
of s corresponding to 0.20 and 0.29 "
1
, respectively, with cfoio being the
Crystal Structure, Morphology, and Orientation of Polyesters 123
distance between the planes of the phenylene rings (see Figures 29 and 38
below). The relative intensity depended on the thickness in quenched films,
the temperature of crystalline films as they were heated, and the azimuthal
angle in oriented fibers. The variations in relative intensity were attributed
to changes in short range ordering of the amorphous chain segment s. These
authors also suggested that the amorphous scattering in oriented samples
should be separated into isotropic and anisotropic components. As dis-
cussed below, the concept of isotropic and anisotropic amorphous compo-
nents has been extended by further studies of oriented samples. We note,
however, that, in the patterns described, there is no evidence of the 10.3
meridional reflection (8.6 20) attributed below to a mesomorphic phase of
PET.
The Suggestion due to Murthy et ai. [42] of the presence of anisotropic
amorphous scattering, mentioned above, was extended by Wu et ai [43],
who separated the amorphous scattering into isotropic and anisotropic com-
ponents, assuming the shape and intensity of the isotropic amorphous peak
to be the same as the amorphous scattering on the meridian; they used the
content of the "oriented non-crystalline" phase to explain the mechanical
properties of fibers.
These X-ray studies were extended further several years later by Wun-
derlich and co-workers [44-46]. They used a fll pattern (two-dimensional
Reitveld) wide-angle X-ray diffraction refinement [47-49], in conjunction
with DSC, SAXS, and thermal mechanical analysis, in order to character-
ize the amorphous "structure" in crystalline PET fibers prepared by high
speed spinning at 4000m/s and then annealed taut at 20O
0
C. Based on
a model of the crystallite structure, a three-dimensional diffraction pat-
tern was calculated and then "rotated" to produce a two-dimensional fiber
pattern for comparison with that observed. The model incorporated param-
eters describing the average size of the crystallites in three dimensions, the
distribution of orientation of the crystallites relative to each other, the av-
erage orientation relative to the fiber axis, the unit eeli parameters, atomic
positions (determined using a rigid body model), a temperature factor, a
paracrystallinity matrix for defects of the second kind, and instrumental
parameters.
Figure 16 shows a contour plot of one quadrant of the observed fiber
diffraction pattern, with Figure 17 being the crystalline-phase diffraction
pattern based on the model. Figure 18 shows the residual fiber scattering
after subtraction of the crystalline scattering (i.e., Figure 16 - Figure 17).
The assumption of the presence of an isotropic amorphous phase with a
truly liquid-like structure, when it is subtracted, results in the pattern
from the anisotropic, non-crystalline phase shown in Figure 19. This plot
was said to be similar to that of Murthy et ai. [42] except that it has no
peak on the meridian at s = 0.29"
1
(3.45, 003' for a 10.3 repeat).
The authors indicated [45] that they observed 003' in patterns from other
fibers and suggest that there are various types of non-crystalline order in
PET fibers.
124 R H. Geil
79800
57400
41300
29800
21400
15400
11100
7980
5740
4130
2980
2140
1540
1110
789
574
Figure 16. Contour plot of t he observed WAXS pattern for a PET fiber (one
quadrant only). The intensity values for the contours are shown at the upper left.
The pattern was corrected for absorption, polarization and Compton scattering
[45]
4 3800
29800
20300
13800
94 30
64 20
4 380
2980
2030
1380
94 3
64 2
4 38
298
203
138
0.9
0.9
Figure 17. Contour plot of the crystalline diffraction [45]
Crystal Structure, Morphology, and Orientation of Polyesters 125
22200
12500
7010
3940
2220
1250
s (A'
1
)
Figure 18. Contour plot of the non-crystalline diffraction [45]
0.9
19900 ^T 0.9
11200 '
5290
0
^
3540 > *
1990
1120
Figure 19. Contour plot of the anisotropic, non-crystalline (i.e., intermediate
phase) scattering. Obtained from Figure 18 by subtraction of an isotropic scat-
tering [45]
126 R H. Geil
Crystal phase
X-ray
Intermediate phase Amorphous phase
Crystal phase
DSC
Rigid amorphous phase
Mobile
amorphous
phase
Figure 20. Comparison of the three phases, crystalline, amorphous, and intermedi-
ate (rigid amorphous), with fractions w
c
, w
a
, and w
r
, respectively, s determined
from X-ray and DSC [45]
This amsotropic phase, defined by Wunderlich and co-workers s an
"intermediate" phase, is clearly oriented, with the broad peaks at s =
0.2 and 0.28"
1
on the equator being near the 010 and 100 reflections.
These are essentially the same positions described by Murthy et al [42],
with Murthy's s = 0.26 peak being closer to the 110 reflection. Additional
anisotropic "amorphous", relatively sharp peaks are seen near the meridian,
at about s = 0.48 (2.08) and 0.95 (1.05), corresponding to the 105
and 109 reflections; indeed, the contour plot for the anisotropic amorphous
scattering has broadened peaks at nearly all locations of strong peaks in the
crystalline phase diffraction. Wunderlich and co-workers suggest that the
initial modulus of the PET bers is primarily determined by the amount
and orientation of the intermediate phase. The tenacity also is strongly
dependent on this phase. They suggest that it is mainly present between the
fibrils, with the "truly" amorphous regions being between the crystallites
within the fibrils. The variations in relative intensity of the s = 0.20 and
s = 0.26 "peaks" described by Murthy et al. could be due to the known non-
cylindrical symmetry of PET uniaxially oriented films [50], the aromatic
rings tending t o be parallel to the film surface. The layer line spacings
above correspond to 10.4 and 9.5 physical repeat distances, the 10.4
value being close to the 10.3 distance used to define the mesomorphic
form, s discussed below.
DSC measurements of the heat of fusion (yielding w
c
), and the change
in specific heat at T
9
(yielding w
a
) allow the determination of the rigid
(rigid above T
9
) amorphous fraction, that is, w
r
= l w
c
w
a
. This rigid
amorphous fraction is presumed to be related to the intermediate phase.
Figure 20 shows a comparison of the three phases, s determined by X-ray
and DSC. Note that apportioning the w
r
to
0
orl x
c
in the two-phase
model will depend on the technique and model used.
Crystal Structure, Morphology, and Orientation of Polyesters
Tip (
1
H)
127

2
Tip (
1
^

1
10
(a)
3.3 (34%)
. Total
21 (20%)
l
"NMR -
crystalline"
"NMR-
amorphous"
o Ethylene
O Carbonyl
- Fit
10 20 30
r (ms)
40
(c)
100
25.5 (27%)
(b)
Total
"NMR -
crystalline"
NMR -
amorphous
O 10 20 30 40
r (ms)
X-R ay rry st ai 1 i n f
"NMR-rrystalliTip"
X-ray amorphous _,....
/X '
/\
Rigid Mobile
"NMR- "NMR-
amorphous" amorphous " f
<
j
| r^
i
^
/
]
-X
/-%,
\
j
N
N
Figure 21.
1
H (a) and
13
C (b) TI
p
NMR plots of PET fibers. The TI
p
are given
on the curves of intensity (M(T)) vs. relaxation tirne (r) for each slope. A model
of the structure as determined by X-ray and NMR is shown in (c) [51]
NMR spectra of PET fibers [51] have been interpreted as indicating that
the amorphous phase has two relaxation t ime constants, with the authors
suggesting the presence of three types of "structure" above T
9
: crystalline,
mobile amorphous, and rigid amorphous. Figure 21 shows the T\
p
(
13
C)
plots for a drawn, semicrystalline PET fiber of 36% crystallinity by com-
128 P. H. Geil
\
:\
Figure 22. Bonart WAXS fiber pattern of an axially stretched, paracrystalline
PET fiber. In addition to 001' and 003', 011 and 112 can be seen, indicating a
degree of 3-dimensional order, with the diffuseness of all of the reflections near
and on the equator indicating distorted lattice surfaces. A hexagonal rod packing,
in a paracrystalline monoclinic unit cell, was suggested for the disorder [53]
bined X-ray and density measurements. The total
1
^ C curve can be divided
into two components, with time constants of 3.4ms (mobile amorphous,
33%) and 54.0ms (rigid, 67%). The division into three components yields
time constants of 25.5ms (rigid amorphous, 27%), 1.3ms (mobile amor-
phous, 1.3%), and 49.7ms (crystalline, 53%). The difference between the
X-ray and NMR values for the crystallinity was attributed to smaller crys-
tals and the constrained amorphous "fringes" on the crystals, contributing
to the NMR crystalline fraction (Figure 2Ic). The TI
P
(
1
H) values are sim-
ilar. In semicry st alline bers, the rigid domain size, estimated from the
relaxation time TI
P
(
1
^ C), agreed with X-ray diffraction measurements of
the crystal size (ca. 100, along the fiber axis), while in amorphous fibers
the mobile and rigid domains were both smaller than 50 .
Bove et al [52] reported one of the first comparisons of permeability
of PET s a function of structure. Using dichloromethane s a penetrant,
they showed that the mesomorphic phase (in an amorphous film drawn 4
at 6O
0
C) was impermeable at low vaporactivity. The X-ray pattern from
the film was similarto that in Figures22 and 23; the reproduction did not
permit distinguishing between them. They reported a 55% impermeable
fraction in theirsample.
Since Bonart first described them in 1966 s "paracrystalline interme-
diate" structures [53], there have been a number of papers describing the
molecular conformation and "crystal structure" of PET resulting from de-
formation of presumably amorphous PET below or near T
9
. Using "spe-
cial conditions of drawing" (not specified), Bonart described a variety of
Crystal Structure, Morphology, and Orientation of Polyesters 129
Figure 23. WAXS fiber pattern of an
axially stretched PET fiber in which
lattice planes (OOl') Iie perpendicular
to the fiber axis, with a d spacing of
10.7 , the length of a fully extended,
trans conformation PET residue, with
all phenyl rings at the same height; thus
a statistically disordered hexagonal rod
smectic packing was proposed for the
"ordered" regions, with the disordered
regions containing oriented molecules
s well [53]
Figure 24. WAXS fiber pattern of an
axially stretched "amorphous" PET
fiber. The Suggestion was that a sta-
tistically hexagonal, nematic packing
of the molecules is present in the ori-
ented portion of the sample. The pos-
sibility of its presence in the isotropic
"amorphous" material was left open.
The method of stretching for this and
the previous two figures was not defined
[53]
paracrystalline structures produced in initially amorphous PET; the X-ray
patterns were said to be similar to those obtained by drawing an amor-
phous sample at 9O
0
C to varying degrees of deformation. In the pattern
closest to that from a normal crystalline fiber (Figure 22), 011 and 112
quadrant reflections are seen, with strong amorphous scattering and a rel-
atively sharp 010 reflection on the equator and sharp, well defined 001'
and 003' reflections on the meridian. The latter two reflections are not
present in the pattern from a crystalline fiber. In another pattern, presum-
ably corresponding to that for less draw at 9O
0
C, only the 001' and 003'
reflections are seen, along with the amorphous scatter on the equator, while
in the third pattern no sharp reflections were seen and only the amorphous
scattering on the equator was observed (Figures 23 and 24). However, in
comparison with the mesomorphic phase described later by Nicholson et ai.
[54], the 001' and 003' reflections corresponded to a physical repeat distance
of 10.7 spacing, the same s in the triclinic eeli. Bonart suggests, based
on these patterns and their stated relevance to amorphous samples drawn
at 9O
0
C, that drawing of amorphous PET results in the formation first of
130 P. H. Geil
Figure 25. Electron diffraction pattern of a thin "amorphous" film of PET cold-
drawn 6x at room temperature. The ring is from the Pt-C shadowing material
used forimaging (Figure 73); five Orders of the meridional reflection can be seen
on the original [16]
(a)
(b)
Figure 26. ED pattern of a thin PET film originally drawn 6x, annealed taut at
18O
0
C for 15 min (Figure 74a) (a) and then redrawn 2 (b). The arrow indicates
the position of the 001' reflection [16]
Crystal Structure, Morphology, and Orientation of Polyesters 131
a nematic, statistically hexagonal rod packing (Figure 24), then a smectic,
distorted hexagonal packing, with the repeat units in these regions being
all trans (Figure 23), and finally a fully extended trans conformation with a
monoclinic unit eeli and the 00/ planes normal to the fiber axis (Figure 22).
This is followed by formation of the triclinic lattice. Bonart did not discuss
the effect of thermal treatment on these samples.
Yeh and Geil, in their paper on the morphology of drawn amorphous
PET [16], also reported electron diffraction (ED) and X-ray patterns. As
shown in Figure 25, for a thin film sample drawn 6x at 25
0
C (and 65
0
C)
010 and 011 are relatively sharp, while 110 and 100 are diffuse or missing.
In contrast to the X-ray data in Figure 20, 005' (s well s the lower Orders
of 00/) can be seen on the meridian, because of the much larger sphere of
reflection for ED. Annealing of such cold drawn thin films at, e.g., 18O
0
C,
resulted in a highly crystalline, highly oriented sample, with reflections out
to the 17
th
layer line, with the tilt of the c-axes relative to the draw direc-
tion shown by the reflections on the l
st
layer line, i.e., not all lying on a
line (Figure 26a). The near meridional reflections on the 5
th
layer line are
105. It was possible to redraw these samples an additional 100%; the result-
ing ED pattern (Figure 26b) is again typical of the mesomorphic (Bonart)
structure, with 001', 010 and 011 being relatively sharp, 110 and 100 be-
ing diffuse, but an apparent 105 (or broad 005') also being present. Using
somewhat thicker films, WAXS patterns of initially amorphous samples
(solution-cast from C2CU) were taken as a function of draw temperature.
A sample drawn at room temperature showed a pattern similar to that of
Bonart in Figure 22, with samples drawn between 25 and 10O
0
C showing
necking and cold drawing, and becoming white and opaque (Figure 27).
The 5
th
layer line can be seen in the original pattern; in comparison with
M
Figure 27. WAXS pattern of a thick "amorphous" PET film drawn 4 at 25
0
C.
001' (d = 10.6 ) and 003' can be seen on the original [16]
132 P. H. Geil
the ED pattern in Figure 25, the equatorial reflection(s) is(are) much more
diffuse. Relatively sharp 011 reflections can be seen in both patterns. Sam-
ples drawn at UO
0
C and higher, however, drew uniformly, the resulting
films being transparent. This is the temperature above which cold crystal-
lization can occur. The X-ray pattern from such a sample, also drawn 4x,
shows almost no orientation, with only diffuse rings corresponding to 010,
110 and 011. Unfortunately, it is not possible to determine whether 003'
was present since its reflection would overlap the 100 reflection and 001'
would be too weak t o observe.
Asano and Seto have proposed a monoclinicunit cell forthe mesomor-
phicphase (Figure 28) [55]. Dimensions are = 4.94, b = 9.2, c = 10.5 ,
O L = 10O
0
C, based on resolving the broad equatorial peak into 4 peaks, cor-
responding t o d = 4.3 and 4.0 on the equator, and a peak with d = 4.6
slightly above and below the equator, assuming a tilt angle of the c-axis of
10 to the draw direction. Thus the meridional reflection at 10.3 would
correspond to the chain repeat of 10.5 , the reduction from the 10.7 re-
peat in the normal triclinic cell being ascribed to some undefined shrinkage
of the polymer conformation, as would occur for the conformation described
(below) by Nicholson et ai. [54].
The preceding experimental methods have been supplemented (and in
some views replaced) in recent years by semiempirical force-field calcula-
tions and molecular orbital calculations. Commercially developed programs
(e.g., [56]) allow the determination of minimum energy conformations both
for single molecules and as they are packed in a crystal. Figure 29 repre-
K)
Figure 28. The paracry st alline monoclinic unit cell proposed by Asano and Seto
to explain the mesomorphic structure of PET. The a
m
and &
m
axes represent
the monoclinic cell, while at and bt represent the triclinic cell resulting from
crystallization. The ellipse represents the plane of the phenyl ring [55]
Crystal Structure, Morphology, and Orientation of Polyesters 133
Figure 29. Simulation of the conformation of a single PET chemical repeat unit
and its packing in the unit cell. Additional chemical repeat units would be present
at each corner of the cell. The Simulation is based on the unit cell parameters of
de Daubeny et ai [ 1 7 ] and the Universal force field [6]
sents the predicted crystalline conformation of a PET molecular segment,
shown as it is arranged in the unit cell. The Cerius
2
program, from Molec-
ular Simulations, Inc., is of particular interest in that it also, based on
the crystal packing mdule, permits display of the corresponding powder,
ber, and single crystal (electron) diffraction patterns for direct compari-
son with experimental results. However, even within the same program, the
134 P. H. Geil
use of different force fields for the description of inter- and intramolecular
interactions yields different results.
Auriemma et al [57] have attempted to describe the conformation
and packing of the molecules in the mesomorphic phase in PET bers
(10.3 spacing) using X-ray diffraction, conformational energy analy-
sis, and Fourier transforms of the resulting models for comparison with
the diffraction results. They concluded that the mesomorphic phase they
described differed from the "intermediate" phase described by others
[42,44,58,59] since the intermediate phase remained after annealing above
10O
0
C while their mesomorphic phase disappeared. The mesomorphic form
is characterized by the presence of 001' and 005' reflections in the X-ray pat-
terns and by its "melting" at about 10O
0
C, whereas the intermediate phase
is an anisotopic amorphous phase, containing trans conformation chains,
that is stable above 10O
0
C and is present in oriented samples containing
the triclinic phase. These authors also conclude that the monoclinic cell
proposed by Asano and Seto [55] is incorrect since it does not explain the
absence of quadrant diffraction. Instead, they propose a model of "extended
linear chains with random sequences of monomeric units in different mini-
mum energy conformations"; there are no lng range correlations between
the planes of the phenylene rings along a chain.
Nicholson et al. also used modeling, with a forerunner of the current
Cerius
2
molecular modeling Software (Polygraf) and a molecular orbital
package (Mopac) [54]. Based on the Mopac package, they described two
stable conformations for the isolated molecule, one the normal trans, planar
form and a second one in which the CH
2
- CH
2
bond is nearly perpendicular
-C-O-CH
2
-CH
2
-
to the molecular axis. In the second conformation, the O
dihedral angles were rotated to ca. 80, i.e., as proposed by Auriemma
et ai. [56]. The single molecule repeat lengths were 10.84 and 10.61, re-
spectively. When packed into the cell, with the energy minimized using the
Polygraf package and permitting all cell parameters to vary, the repeat pe-
riods expanded to 11.01 and 10.81 , with the dihedral angles in the shorter
conformation changing from 80 to 70. Nicholson et al. [54] attributed the
larger repeat periods (than for the single molecule and for the actual crys-
tal) to the Dreiding force field and the programs' interaction energies. They
suggested that this was the same mesophase reported by Bonart [53].
In a second paper from the same laboratory, Parravicini et al. [60] reex-
amined the effect of annealing on drawn, "amorphous", melt-spun fibers,
using samples of various molecular weights. Annealing was done taut, for
15 min. The X-ray patterns were said to be independent of the molecular
weight of the samples used, although the maximum draw ratio increased
with molecular weight. The presence of a 103' reflection on the meridian
(d = 3.45 , 20 = 25.8) for the as-drawn and 6O
0
C annealed sample was
replaced by a split 103 on the 20 = 26.6 layer line, typical of the triclinic
cell, by annealing at 80 and 18O
0
C. The equatorial and quadrant reflections
Crystal Structure, Morphology, and Orientation of Polyesters 135
followed the pattern described by Bonart [53], triclinicreflections beginning
to be seen at 8O
0
C superimposed on the oriented amorphous scattering.
Asano et ai [61] have also recently reexamined the effect of annealing
cold drawn amorphous samples, using WAXS and mid-angle X-ray scat-
tering (MAXS), with the MAXS patterns being taken at the annealing
temperatures in an angular range that permitted simultaneous observa-
tion of the 001' reflection and SAXS. Using an amorphous film of unstated
processing history, they drew it 3.8x at 0.8 mm/ min, and observed slight
whitening. Annealing was done taut. The as-drawn sample showed an ori-
ented amorphous structure and broad peaks on the equatoronly. Annealing
at 5O
0
C resulted in the appearance of 001' if annealed for 100 s (but not
for 10s), while 001' appeared forall higher annealing temperatures up to
9O
0
C and was most intense forannealing at 7O
0
C. A schematicdiagram of
the WAXS, MAXS, and SAXS results of Asano et ai. is shown in Figure 30,
with an Interpretation in terms of molecularpacking in Figure 31. As in-
Unoriented
amorphous
WAXS SAXS
O None
Oriented ^*
amorphous
(room temp.)
o None
Smectic
(6O
0
C)
o None
Layer
structure
(7O
0
C)
Triclinic
structure
(>8O
0
C)
Development of
tricliniclayers
(> 10O
0
C)

' *
Figure 30. Schematicdiagrams of the effect on the WAXS and SAXS patterns of
cold-drawing and annealing "amorphous" PET films [61]
136
R H. Geil
Nematic Smectic Triclinic
Figure 31. Schematic diagrams of the molecular packing proposed by Asano et
al. for cold-drawn and annealed "amorphous" PET [61]
dicated, these authors suggest that cold drawing of an amorphous sample
results in the fornaation of an oriented nematic phase, in which the lateral
Position of the phenylene rings on neighboring molecules varies. When an-
nealed at 6O
0
C, the nematic phase is suggested to transform into a smectic
phase with a 10.7 spacing (10.3 in the earlier paper [55]), with neigh-
boring rings aligned on planes perpendicular to the draw direction and a
slight tut of the molecules. Annealing at 8O
0
C results in the triclinic phase
with the rings still on planes normal to the draw direction and a larger
molecular tut.
In the most recent studies of the mesomorphic structure, Mahen-
drasingam et al. have utilized time-resolved Synchrotron radiation to fol-
low the development of crystallinity during strain-induced crystallization
of amorphous PET films [62-64]. Figure 32a,b shows patterns taken during
stretching at 9O
0
C at a rate of 13 s"
1
(72 000%/min), each with an exposure
time of 40ms. In frame4 (Figure 32a, pattern (b)), there is an increase of
intensity on the equator, indicating the development of segmental orienta-
tion. In frame5 (Figure32a, pattern (c)), recorded close to the end of the
deformation, there is a high degree of "equatorial orientation", no sign of
crystalline reflections, and a weak meridional 001' reflection at a spacing
of 0.098-
1
(10.2), s well s 003' and 005' reflections (the latter are
not commented on). From frame5 to frame7, it increases in intensity, with
signs of crystalline reflections first showing in frame 7 (Figure 32a, pattern
(d)). For the next 5 frames, all following the end of deformation, there was
a gradual decrease in the intensity of the 001' reflection and an increase in
the crystalline reflections.
Mahendrasingam et al. [63,64] followed the development of crystallinity
s a function of draw temperature and draw rate, and defined three regimes,
depending on the relationship between the draw rate and the relaxation
Crystal Structure, Morphology, and Orientation of Polyesters 137
(a)
(b)
Figure 32. Synchrotron X-ray diffraction patterns taken during stretching an
"amorphous" 500 thick PET film at 9O
0
C, draw rate 72 000%/ min, to an
overall draw ratio of 3.7. Shown in the top set of patterns are frames l, 4, 5, 7,
12, and 24, each with a 40 ms exposure, from a total of 124 frames. The bottom
set of patterns show the change in intensity of the 001' reflection in the central
portion of the patterns forframes 5-10 of the same set of data [62]
138 R H. Geil
0.4 0
0.35

s " o .3o
Sc
~Q^0.25
^
1
T
1
O 0.20
o ^
^ T*| 0.15
" ^
I
10
Q
0.05
0.00
D raw ratio +
Crystallinity x
Half-width of (010) reflection *
************* x****m ********
0.0 0.5 1.0
Time (s)
1.5 2.0
Figure 33. Development of the (010) peak area (crystallinity) and its half-width
during draw at a rate of 3.5 s"
1
to a final draw ratio of 3.2 [64]
times T
B
r
c
, the first being related to retraction of the molecules within
their deformed "reptation" tubes and the second to reptation out of the
deformed tubes t o a new, isotropicconformation [63]. At draw rates faster
than I/T
B
, crystallization does not commence until after deformation has
stopped. At draw rates slower than l/ r
c
, the molecules relax faster than
they are oriented by the draw and no oriented crystallization occurs. At
intermediate rates, oriented crystallization can occur during deformation.
The mesomorphicphase developed only forthe highest draw rates at the
lowest temperature used (9O
0
C), developing during deformation and dis-
appearing following cessation of draw and initiation of crystallization, i. e. ,
in the first regime. Figure 33 shows the Variation in crystallinity and the
half-width of the 010 reflection as a function of time, in comparison with
the draw ratio, fora sample drawn at a rate of 3.5s"
1
at 9O
0
C [64].
The 001' reflection was not observed in this sample. The half-width is es-
sentially independent of the time while the crystallinity is increasing (after
cessation of deformation, i. e., the first regime). This was interpreted as indi-
cating that the primary crystallization was related to an increase in number
or amount of crystals rather than to an evolution of density fluctuations
(i. e. , spinodal decomposition) as proposed by Strobl for crystallization of
a pre-oriented amorphous PET sample when heated just above T
9
[65].
Mahendrasingam et ai suggest, however, that the sporadically nucleated
Crystal Structure, Morphology, and Orientation of Polyesters 139
crystals in these samples may originate from the "oriented density fluctu-
ations" described by Imai et ai [66] as developing during the induction
period for isothermal cold crystallization of isotropic PET, with the paral-
lel alignment of the segments being aided by the orientation; these results
are discussed further below.
Based on the above, it is reasonable to conclude that "amorphous"
PET, in the form usually studied, obtained, e.g., by quenching a molten
film into ice-water or by injection molding, contains segments in which the
-C-O-CH
2
-CH
2
-
conformation of the O unit is Irans rather than gauche.
There was little direct evidence above, however, that these Irans units are
collected to form domains in either the mesomorphic or the intermediate
or rigid amorphous phases in the unoriented "amorphous" samples. On
the other hnd, there is such evidence in oriented amorphous samples be-
fore and after annealing. The presence of X-ray reflections on layer lines
in the cold drawn, low temperature annealed samples could be interpreted
in terms of the molecular transform, but the presence of the broad reflec-
tions on and near the equator implies a nematic or smectic type alignment
and thus domains, as pictured in FigureSl for the mesomorphic form. The
NMR results suggested that the domains were less than 50 in size and,
if the breadth of the several "amorphous" X-ray peaks is interpreted in
terms of crystal size, a similar value would be obtained. The permeability
results require impermeable domains, but do not yield a size. IR or Ra-
man spectoscopy, on the other hnd, which was the basis for evidence of
Irans conformations in amorphous PET, samples only local conformations
and interprets most of the bands involved as unaffected by nearest neighbor
units on the same or adjacent chains, and thus not requiring domains. How-
ever, electron micrographs of quenched PET were interpreted in the 196Os
as indicating the presence of nodules, domains of local order (described
as a type of liquid crystalline ordering) [15], and thus it is reasonable to
suggest that the Irans amorphous segments are collected in the nodules.
The domains, when present, probably differ in their molecular packing and
morphology in various types of samples as a function of sample preparation,
and in semicrystalline vs. non-crystalline samples.
This stiil leaves the question of the relationship, if any, between the
mesomorphic phase and the intermediate or anisotropic amorphous phase.
According to Auriemma et ai [57], e.g., the mesomorphic phase, found in
cold-drawn amorphous samples annealed below 80-10O
0
C, is defined by the
presence of a 10.3physical repeat distance based on the presence of 001',
003', and 005' meridional reflections. We have been unable to resolve the
problem that in some papers the layer line spacings are said to correspond
to a 10.3 repeat and in others to a 10.7 repeat, even in some cases
for papers from the same laboratory; since these are layer line spacings,
the corresponding physical repeat distances are independent of molecular
140 P. H. Geil
tut. This phase transforms into triclinic when annealed at 80 or 10O
0
C,
depending on the aut hr. Auriemma et al [57] suggest the mesomorphic
phase differing from the intermediate phase in that the latter is present in
semicry st alline (triclinic) samples that have been annealed above 10O
0
C,
with the 10.3spacing being absent. However, Wunderlich and co-workers
[44-46] report the presence of a 10.4 distance (fifth layer line) in their
intermediate-phase, partially cry st alline samples. Since in both types of
samples broad equatorial reflections are present, domains of liquid crystal
order are suggested. We thus conclude that the differences are more a result
of the processing conditions, leading to formation of such domains and their
resultant morphology. Clearly, there is need for further research (see also
Chapter 4).
2.1.2. Conformation in the cry stalline phase; crystal structure
Since PET is the first and still most commercially important polyester, its
crystal structure has been intensively studied for more than 50 years (e.g.,
[17,24,44,54,55,67-86]). The earliest report on the unit cell dimensions of
PET was in 1946 by Astbury and Brown [67], measured from X-ray ber
diffraction photographs. Later, in 1954, using better oriented specimens,
Daubeny, Bunn, and Brown [17] observed additional reflections which did
not fit Astbury and Brown's cell. Another unit cell, now considered the
conventional unit cell for PET and often described s Bunn's cell, was sug-
gested, hving = 4.56, b = 5.94, c = 10.75, a = 98.2, = 118,
and 7 = 112, with a calculated crystal density of p
c
= 1.455g/ cm
3
. The
chain conformation and atomicpositions in the crystals were also deduced
from the relative intensities of the X-ray reflections. Ten years later, in
1964, Tomashpolskii and Markova [70] used electron diffraction to obtain
patterns from solution-cast PET films that had been stretched 700% and
subsequently heated at 18O
0
C, and proposed similar, but slightly different
cell parameters (see Table 1). At about the same time, Yamashita reported
the morphology of PET folded-chain, single, solution-grown crystals, but
only obtained [101]*ED patterns [71]. In 1975, Fakirov et al [72] rechecked
the unit cell dimensions using X-ray diffraction from 5.5x drawn samples of
l mm thick "amorphous" bristles annealed taut for6 h at various tempera-
tures between 100 and 26O
0
C; these authors suggested anotherset of unit
cell parameters with a higher calculated density of the crystalline phase
(pc = 1.515 g/ cm
3
) forsamples annealed above 14O
0
C and a slightly larger
cell forlower annealing temperatures (p
c
= 1.484 g/ cm
3
at 10O
0
C). They
also suggested the presence of "highly distorted paracrystalline nuclei" dur-
ing low temperature stretching based on an X-ray pattern similarto that
for the mesomorphicphase.
*Use is made of italics, i. e. , [hkl] in this chapter to represent zones in ED patterns
t o prevent confusion with the use of [xxx] t o represent references.

You might also like