This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
—NCO,
chain-extended with 1,4-butanediol, HO(CH 2 ) 4 OH, to give the structure C H 2 — < < 3 ^ N H . CO . O(CH 2 ) 4 O . OC .
for the hard segment di-isocyanate. The degree of hydrogen bonding of the NH groups was studied by examining the v(NH) region of the spectrum (see fig. 6.12) and looking for the presence of a shoulder in the vicinity of 3450 cm" 1 , indicative of free NH groups. This appears on the much stronger peak due to bonded NH groups at 3320 cm" 1 and its intensity is difficult to measure with precision, but it cannot amount to more than 15 % of the total, so that at least 85% of the NH groups are hydrogen bonded. The peak arising from the carbonyl stretching mode is split into two components, one at 1733cm" 1 and the other at 1703cm" 1 . Studies on model compounds such as N-phenylurethane, \ O ) — N H . CO . OC 2 H 5 , show that the two maxima are assignable to free and bonded carbonyl groups, respectively. In order to determine the degree to which the carbonyl groups are bonded it is necessary to know the relative molar absorptivities of the 1733 and 1703 cm" 1 modes. Studies on a range of model compounds suggest that there is some
260
The microstructure of polymers
increase in molar absorptivity when hydrogen bonding occurs, but generally not greater than 20%. It is therefore useful to consider two extreme possibilities: that there is no change in absorbance or that it increases by a factor of 1.2. As the ratio of the absorbances of bonded and free carbonyl groups is 1.55, it follows that 6 1 % or 56% of the urethane carbonyl groups are involved in hydrogen bonding for the two assumptions, respectively. This then leads to the conclusion that the remaining bonded NH groups, at least 25% of the total, must be involved with a second acceptor group, and this can only be the ether
Fig. 6.12. F r e e ^ N H absorption at 2.90 ^01(3450 cm" 1 ) for the polyurethane -f-(CH 2 ) 8 OCONH(CH 2 ) 6 NHCOO^; at (a) 22°C and (b) 100°C. (From D. S. Trifan and J. F. Terenzi, Journal of Polymer Science 28 443 (1958). Copyright © (1958) John Wiley & Sons, Inc. Reprinted by permission of John Wiley & Sons, Inc.)
s
2.9 3.0 Wavelength (jum)
6.6 Hydrogen bonding
261
oxygen linkage. The complexity of the peaks in the v(C=O) region for polyester urethanes prevents the straightforward use of this approach. The second example covers a more quantitative interpretation of the infrared spectra and their temperature dependence for two types of polyurethane. For one type, a polyether polyurethane (ET), the soft segment diol is again polytetrahydrofuran, and for the other, a polyester polyurethane (ES), it is poly(tetramethylene adipate), H-^O(CH 2 ) 4 . O . CO . (CH 2 ) 4 . CO^O(CH 2 ) 4 OH. For both types the hard block is MDI extended with butanediol, as for the first example. The v(NH) region of the spectrum was computer fitted with three Gaussian peaks, centred at 3420,3320 and 3185 cm" 1 . The first of these is due to free NH groups, the second to bonded NH groups and the third, which is rather weak, is attributed to bonded NH groups in the conformation in which the NH group is cis to the C = O group rather than, as is more usual, trans to it. The area of this peak was therefore added to that of the peak at 3320 cm" 1 in calculating the concentration of bonded NH groups. The ratio of the molar absorptivities at 3320 and 3420 cm" 1 was assumed to be 3.46, a value based on studies on model compounds. The v(C=O) profile for the ET series was fitted with two Gaussian peaks at 1733 and 1703cm" 1 , characteristic for free and bonded carbonyl groups, respectively, whose molar absorptivities were taken to be equal. It was not possible to fit the v(C=O) profile in the ES series, because of the ester groups present in the soft segments. The results for the ET series, covering a range of compositions in terms of MDI hard segments, were, with one exception, very uniform, showing 82 ± 2 % of bonded NH with 66 ± 2 % of bonding to the urethane carbonyl group. The amount of bonding to the urethane ether varied somewhat more, with upper and lower bounds of 20 % and 12 %. The exception was a material in which the soft segment was of appreciably higher molecular weight, and the percentage of hard segments was high. This material, which was semi-crystalline, had 78 % of the NH groups bonded, 73 % to the carbonyl group and only 5 % to the urethane ether. This suggested a greater than average phase separation, in line with other properties of this material. The polymers of the ES series showed a comparable degree of NH bonding, ranging between 75% and 84%. Studies of the temperature dependence of the degree of hydrogen bonding in the ET series showed, as expected, that it decreased with increasing temperature and the results were used, together with other information, to provide evidence about the influence of hard and soft
262
The micro structure of polymers
segment lengths on the perfection and purity of the hard and soft segment phases.
6.6.4
Biopolymers Hydrogen bonds are important in stabilizing biopolymers in their native conformations. In proteins, for example, many hydrogen bonds are formed between t h e ^ N H and C = O groups of the backbone peptide groups and these tend to stabilize the backbone into one or other of a variety of helical conformations in which the hydrogen bonds are aligned in a cooperative fashion. Additional hydrogen bonds are formed from N—H, C = O and O—H groups of the protein side chains and, in general, all potential hydrogen-bond forming groups are in fact involved in hydrogen bonds either to other groups on the protein or to the water molecules that are normally present in biological systems. Despite the stabilizing effects of hydrogen bonds and other interactions such as van der Waals interactions, the energy difference between the native state of a protein and the unfolded, or 'denatured', state is quite low, commonly of the order of 100 kJ mole" 1 ; in an aqueous environment, denaturation can be promoted at temperatures not very greatly above ambient, the favourable interactions within the protein molecule that are disrupted being replaced by only slightly less favourable interactions with water molecules surrounding the unfolded state. In fact, the stability of proteins is determined not so much by the enthalpic interactions as by the higher entropy of the bulk water when the protein is completely folded. The phenomenon of denaturation causes changes in the infrared spectrum because of the greater strength of the cooperative arrangement of hydrogen bonds in the native protein. A particular and widely studied example of denaturation is that of the fibrous protein collagen, the material of tendon. This molecule consists of three helical chains super-coiled around each other in the form of a rope, stabilized by van der Waals interactions and hydrogen bonds between the three chains (see fig. 6.13). Denaturation, in which the fibre collapses into a gelatinous state, involves a 'melting' of the triple helical structure with a concomitant change in v(NH) from 3330 to 3300 cm" 1
Fig. 6.13. Schematic diagram of part of a collagen molecule, showing the regular pattern of hydrogen bonds that link the three helical chains of the molecule so that they super-coil round each other. (From J. R. Holum, Principles of Physical, Organic and Biological Chemistry. Copyright © (1969) John Wiley & Sons, Inc. Reprinted by permission of John Wiley & Sons, Inc. After an original drawing by A. Rich.)
6.6 Hydrogen bonding
263
Nl
often hydroxyproline often pro line glycine
264
The microstructure of polymers
as the regular pattern of inter-chain hydrogen bonds is replaced by an irregular pattern of hydrogen bonds to neighbouring water molecules. Raman spectroscopy is often the favoured approach with biopolymers because of the ease of sampling and because of the generally high water content. This technique has been used to study the 'melting' of DNA, the genetic material, whose double helical conformation is stabilized by hydrogen bonds that provide specific interactions between complementary planar nitrogenous bases attached to the two sugar-phosphate backbone chains. In a synthetic analogue formed by one chain of polyribo-adenylic acid and one chain of polyribo-uridylic acid the Raman spectrum shows a clear peak at 1681 cm" 1 at low temperatures. This is one of the v(C=O) vibrations from the uracil base when it is hydrogen bonded to adenine. At higher temperatures, the double helix 'melts' as the hydrogen bonds between pairs of bases are disrupted and at 59°C the 1681 cm" 1 peak is replaced by two at 1660 and 1698 cm" 1 characteristic of the free uracil moiety. 6.7 6.7.1
Chain order and crystallinity Introduction The way in which the vibrational modes of polymers are modified when the chains are present in crystalline regions has been discussed in chapter 3, primarily from the fundamental angle. The consequences will now be considered in more detail, by reference to four polymers, polyethylene, polyoxmethylene, poly (vinyl chloride) and polytetrafluoroethylene, as will the practical implications. These include the determination of crystallinity, the testing of postulated intermolecular force fields, the provision of evidence for a third phase, neither crystalline nor amorphous, in polyethylenes and the provision of evidence for or against particular crystal structures. In addition, it is now known that certain vibrations that were previously assumed to be specific for chains in crystalline regions are actually characteristic for no more than one-dimensional order along the chain length. They can, nevertheless, sometimes still prove useful for obtaining an approximate measure of crystallinity, and the general question of chain order and regularity peaks in vibrational spectra will be discussed after some clear examples of the effects of crystallinity have been given. By way of introduction, it will be useful to summarize the principles set out in more detail in chapter 3. In essence, if there is a single chain passing through the crystallographic unit cell the spectrum will differ only marginally from that of a single, isolated chain or a group of chains with the same conformation in an amorphous region. Modes that were degenerate for a single chain
6.7 Chain order and crystallinity
265
may have different frequencies, though in general the splitting will not be very great, and modes inactive for the single chain may be activated. On the other hand, by comparison with a group of chains in the amorphous regions the peaks will tend to be somewhat sharper because of the removal of the spatial disorder which causes each vibrational frequency to be spread over a small range. If there are m chains per unit cell, correlation splitting of the modes occurs for m > 1. Any mode of the single chain is split into m modes which differ, to a first approximation, only in the relative phases of the motions of the m chains in each unit cell; the motions are correlated. In addition, there are 4m — 3 low frequency lattice modes for all values of m, including m = 1. Two assumptions are involved in this description of the effects of crystallinity. They are that the intermolecular forces are small compared with the intramolecular forces, so that the perturbation of the single chain vibrations is small, and that the splitting is not obscured in practice because the corresponding peaks in the spectrum are too broad or are overlaid by peaks due to non-crystalline material. 6.7.2
Polyethylene The formal theoretical treatment of the factor group modes of crystalline polyethylene has already been given in subsection 3.2.2. The nature of the modes resulting from correlation splitting is very easily illustrated and a simple vectorial approach allows an easy deduction of the directions of the infrared transition moment dipoles. Consider the CH 2 rocking vibration of species B2u for the single chain, shown in fig. 6.14a as it would appear if viewed along the length of the chain. The dipole moment changes consequent upon the movements of the four hydrogen atoms are shown resolved along axes parallel and perpendicular to the projected C—C bond direction in fig. 6.14b. It is clear that the net dipole moment change is perpendicular to the C—C bond direction, as indicated by the broad arrow in fig. 6.14a. The unit cell of polyethylene is orthorhombic and contains two molecules: one molecule passes through the centre of the cell and one molecule at each corner of the cell is shared with three other unit cells. The symmetry of the factor group allows only two different phase relationships between the vibrations of the two chains in the unit cell and these are shown in fig. 6.14c. When the net dipole moments of the chains are resolved parallel to the a and b axes of the unit cell and summed, bearing in mind that the contribution per unit cell from the central chain will be four times that from each of the corner chains, the overall dipole moment is seen to lie parallel to the b axis in the upper diagram and parallel to the a axis in the lower one.
266
The micro structure of polymers
Measurements on the 720/731 cm~ 1 infrared doublet, assigned to the rocking mode, using plane polarized radiation and a biaxially oriented sample, show that the peak at 720 cm" 1 originates from the dipole moment change parallel to the b axis (B2u species for the crystal) and that the peak at 731 cm" 1 originates from the dipole moment change parallel to the a axis (Blu species). A similar conclusion has been obtained from studies on single crystals of high molecular weight normal paraffins. A similar splitting occurs for the Blu species bending mode, (3(CH2), giving absorption peaks at 1462 cm" 1 (B2u) and 1473 cm" 1 (Blu). The splitting of the other infrared-active modes is too small to be observed for two reasons. First, the size of the splitting
Fig. 6.14. The nature of the Biu/B2u rocking doublet modes in polyethylene: (a) the rocking vibration of a single chain viewed along the chain axis. For simplicity the motions of the carbon atoms are not shown. The broad arrow indicates the direction of the instantaneous dipole moment, which is the vector sum of those due to the four C—H bonds; (b) the bond dipole changes resolved parallel and perpendicular to the projected C—C bond direction; (c) the two allowed phase relationships of the vibrations of the two chains in the unit cell for factor group modes and the corresponding net dipole changes.
(ft)
w (b)
6.7 Chain order and crystallinity
267
depends on the type of vibration involved and is greater for those modes that involve motions predominantly perpendicular to the chain direction because the inter-chain interaction is stronger. Secondly, the width of the peaks is a relevant factor, as two peaks separated by less than their width will not be seen as resolved. One method for reducing this limitation on the observation of splitting is to cool the sample. This sharpens the peaks (see section 4.6) and also leads to a thermal contraction of the crystal lattice, bringing the chains closer together and so increasing the inter-chain interaction and the splitting. This method has proved of considerable value for resolving correlation splitting doublets for several modes in the Raman spectrum of crystalline polyethylene. The B2g skeletal stretching mode, a single peak at 1066 cm" 1 at ambient temperature, appears as a doublet at 1065 cm" 1 (Blg) and 1068 cm" 1 (B2g) at - 196°C and at this temperature there are peaks at 1295 cm" 1 (B2g) and 1297 cm" 1 (Blg) corresponding to the twisting mode yt(CH2) of species Blg for the single chain, which is observed as a single peak at 1296 cm" 1 at ambient temperature, as shown in fig. 6.15. Furthermore, splitting of the B2g species wagging mode, yw(CH2), of perdeutero-polyethylene gives components at 827cm" 1 (Blg) and 830cm" 1 (B2g) at - 160°C. It is possible to calculate values for the splittings by assuming values for the intermolecular and intramolecular force constants and the values so obtained tend to be larger than those observed, particularly for small splittings. No doublets have been observed corresponding to calculated splittings less than 5 cm" 1 and the failure to observe them is simply a consequence of the small splittings and not a failure of the predictions of group theory. Although correlation splitting of the <5(CH2) mode was established very early in the study of the infrared spectrum of polyethylene, an understanding of the corresponding region of the Raman spectrum has been more difficult to obtain. As shown in fig. 5.1, three peaks are observed at 1416, 1440 and 1464 cm" 1 . Polarization studies show that they are of symmetry species Ag, B3g and jB3g, respectively. In one of the earliest studies of correlation splitting in the Raman spectrum it was suggested that the peaks at 1416 and 1440 cm" 1 were the correlation split doublet corresponding to the <5(CH2) mode of symmetry Ag for the single chain. Other authors have, however, suggested that the three observed peaks are simply the result of the superposition of a large group of peaks resulting from the Fermi resonance of the components of the correlation split doublet due to the S(CH2) mode with the overtones and combinations of the correlation split doublet due to the yr(CH2) modes at 7207731cm"1.
268
The microstructure of polymers
The fact that the intensity of the peak at 1416 cm " 1 , relative to the rest of the spectrum, is insensitive to temperature change between — 180°C and +100°C even though its frequency changes by about 4 c m " 1 suggests that it cannot be involved in Fermi resonance, since the intensities of the components of such resonance split peaks show a very strong dependence on the degree of interaction, which in turn depends on the precise frequencies of the interacting vibrations. This peak is thus assignable to the Ag component of the correlation doublet. The assignment of the other two peaks is less simple and may involve Fermi Fig. 6.15. The effect of temperature on the splitting of the B lg /B 2g species twisting doublet observed at 1296 cm" 1 in the Raman spectrum of polyethylene: {a) sample at room temperature; (b) sample at — 196°C; (c) sample at — 196°C with increased spectral resolution and expanded wavenumber scale (•—• represents 5cm" 1 in each spectrum); {d) forms of the vibrations. The + and — signs represent displacements up and down with respect to the plane of the diagram. Only two differently oriented chains are shown for simplicity.
(a)
6.7 Chain order and crystallinity
269
resonance. In addition, as discussed below, the peak at 1440 cm" 1 is partially due to non-crystalline chains. Although the occurrence of correlation doublets would appear to offer a method for the determination of crystallinity, quantitative work is not straightforward. This is true even for the 720/731 cm" 1 doublet, which corresponds to the single peak at 720 cm" 1 in the amorphous material, so that measurement of the ratio of the absorbances in the two peaks would in principle lead to a determination of crystallinity. The reason is that it is difficult to produce totally unoriented samples, and since the transition moments for the two components are at 90° to each other, correction for the polarization introduced by the spectrometer must be made. Although this is not difficult in principle, other methods have now come to be preferred. The first of these utilizes the absorption peak at 1894 cm" 1 , which arises from the combination vibration of the Raman-active CH 2 rocking fundamental at 1170 cm" 1 and the infrared-active CH 2 rocking fundamental at 731cm" 1 . This combination vibration is specific for chains in crystalline regions and is less sensitive to sample orientation than the CH 2 rocking doublet. It is often convenient to determine the ratio of the intensity of the peak at 1894 cm" 1 to that of a convenient internal standard peak, such as the peak at 910 cm" 1 due to the out-ofplane deformation vibration of the chain-terminating vinyl group, as this eliminates the need for the measurement of sample thickness. The method requires calibration and this may be done in terms of crystallinity values measured by X-ray diffraction or density determination. The alternative approach is to determine the proportion of chains in amorphous regions, using a peak at 1303 c m ' 1 , due to CH 2 wagging in amorphous regions. This has two significant advantages. Direct calibration is possible using a paraffinic hydrocarbon that is liquid at ambient temperature, such as normal hexadecane, CH 3 (CH 2 ) 1 4 CH 3 , or an amorphous polyethylene obtained by light cross-linking by yirradiation in the molten state. Secondly, although it is difficult to avoid some orientation in amorphous regions, the intensity of the peak at 1303 cm" 1 is substantially unaffected. The method is well proven, as plots of amorphous content determined by X-ray diffraction against amorphous content based on the absorbance at 1303 cm" 1 give straight lines over the approximate concentration ratio 20% to 100%, with a slope of unity, as shown in fig. 6.16. Both of these approaches to the measurement of crystallinity tacitly assume that the polymer is a two phase system, consisting of wholly crystalline and wholly amorphous material. As explained in subsection
270
The microstructure of polymers
1.2 A, the crystalline regions consist of lamellae, typically a few hundred Angstroms in thickness. Originally it was assumed that a sharp change occurs at the boundary between a lamellar crystallite and the adjacent amorphous region. More recently, however, several studies have suggested that there is a layer of intermediate order at the interface. Some evidence for this model has come from the Raman spectra of various polyethylene samples. The spectrum in the CH 2 bending region consists of three peaks, superimposed upon a much broader peak characteristic for chains in amorphous regions. The peaks at 1416 and 1440cm" 1 , both assigned in these studies to the correlation doublet, are of approximately equal intensity for highly crystalline samples. For polymers of lower crystallinity, however, their intensities differ markedly and it is not possible to account for the observed spectrum in terms of partially crystalline and partially amorphous material. The intensity at
Fig. 6.16. Correlation of infrared and X-ray diffraction estimates of amorphous content in high molecular weight hydrocarbons. (From R.G.J. Miller and H.A.Willis, Journal of Polymer Science 19 485 (1956). Copyright © (1956) John Wiley & Sons, Inc. Reprinted by permission of John Wiley & Sons, Inc.) 110
10
20
30
40 50 60 70 Infrared % amorphous
80
90
100
6.7 Chain order and crystallinity
271
1440 cm" 1 is too high, suggesting the presence of an additional peak, which has been assigned to regions, presumed to be at the crystal/ amorphous interface, where the chains are in the all-trans conformation but do not show regular packing in the lateral direction. This assignment is based on the Raman spectra of normal alkane crystals near to, but below, their melting points, which show a single peak at 1438 cm" 1 . Estimates of the concentration of the interfacial material have been made by using the areas under the peaks at 1416 and 1303 cm" 1 as measures of the concentration of truly crystalline and truly amorphous material, respectively. The area under the peak at 1296 cm" 1 , due to CH 2 twisting and insensitive to crystallinity, was used as an internal standard, with calibration factors for crystalline or amorphous material being determined from measurements on extremely highly crystalline or molten polymer, respectively. 6.7.3
Polyoxymethylene Polyoxymethylene, -(-CH2—O-fe, prepared by conventional routes crystallizes in a hexagonal unit cell with only one chain passing through it. The conformation of the chain is helical, with nine monomer units in five turns, a 9 5 helix (see fig. 6.17b and d). The vibrational spectrum of crystalline samples differs from that of almost amorphous samples mainly in having rather sharper peaks. Slow polymerization in alkaline conditions yields a polymer which crystallizes to give an orthorhombic unit cell with two molecules passing through it. The conformation of each of these chains is a 21 helix, illustrated in fig. 6.17 a and c, in which the bond angles and the angles of rotation around bonds are not very different from those in the 9 5 helix; a 2l helix may equivalently be thought of as a 105 helix. It is therefore not surprising that the vibrational spectrum of the orthorhombic form is quite similar to that of the hexagonal form. The most important difference is that the spectrum of the orthorhombic form shows correlation splitting of some of the modes. The factor group of the line group for the 9 5 helix is isomorphous with the group D9 and the spectroscopically active modes are distributed among the symmetry species as follows: 5 ^ + 5A2 + 11E1 + 12£2- The factor group for the 2X helix is isomorphous with the group D2 and the 20 true vibrational normal modes are distributed equally among the four symmetry species A, B^, B2 and B3. The spectroscopic activities and the correspondence between the symmetry species for the two helices are shown in table 6.3. The A modes of the 21 helix correspond to the A1 modes for the 9 5 helix, the B1 modes correspond to the A2 modes (taking the helix axis as Oz) and the B2 and B3 modes correspond to the El
272
The microstructure of polymers
modes, which lose their degeneracy. The E2 modes of the 9 5 helix also lose their degeneracy and become identical to A or Bx modes of the 2X helix. The factor group of the space group for the orthorhombic structure is also isomorphous with the group D2 and the relationship between the factor groups of the site and space groups is such that each of the line group modes of species A or Bl splits into two components of species A and B1 and each of the modes of species B2 or B3 splits into two components of species B2 and B3. The relationships between the symmetry species and spectral activities of the single 9 5 and 21 helices and the orthorhombic crystal are also shown in table 6.3. The final predictions for the infrared spectrum are that active A2 modes of the 9 5 helix should give rise to active Bx modes of the Fig. 6.17. The chain conformations of polyoxymethylene: {a) and (b) show the projections of the carbon and oxygen atoms of two chemical repeat units for the 2j and 9 5 helices, respectively, onto planes normal to the helix axes; (c) and (d) show projections which are normal to the chain axes and to the lines AB for the 21 and 9 5 helices, respectively.
(c)
6.7 Chain order and crystallinity
273
Table 6.3. Relationships between symmetry species and spectroscopic activities for the vibrational modes of different forms of polyoxymethylene 95 helix species
Single 21 helix R
IR
species
R
Orthorhombic crystal IR
species
R
IR
P d P d
d d
o
B2,B3
d
B2,B3
d
Notation as explained in section 5.1.
orthorhombic crystal, active Ex modes should give rise to active B2, B3 doublets and inactive Ax or E2 modes should give rise to active Bl species modes. Some of the correlation split doublets, those corresponding to Ex modes of the 9 5 helix, should thus be observable directly in the infrared spectrum, whereas others, corresponding to Al9 A2 or E2 modes, require the observation of Raman spectra. The task of assigning the modes in the infrared spectrum and of seeking the B2 and B3 doublets, assuming that the intermolecular coupling constants are large enough, is complicated by the fact that the orthorhombic form is unstable and reverts to the hexagonal form at 70°C or on stretching. It is therefore necessary to examine the sample in powder form and polarization information cannot be obtained. Such information can, however, be obtained for the hexagonal form, for which A2 species should be n modes and Et species should be a modes. A comparison of the infrared spectra of the hexagonal and orthorhombic forms then suffices to establish which modes are doubled in the orthorhombic form and to identify which line group modes they correspond to. The results are shown in table 6.4 for those modes which can be clearly established as doubled. The labelling in terms of group modes is, as usual, only approximate.
274
The microstructure of polymers
Table 6.4. Vibrational modes of polyoxymethylene which show correlation splitting Frequency (cm" 1 ) Mode
Hexagonal form
Symmetry species Orthorhombic form for single D2 chain
C—H stretching 2984
3000 2980 2965
Bl9B3
1471
1488 1466
B3
1238
1237 1220
B2
1120
1120 1110
B3
456
434 428
B2
CH 2 bending CH 2 rocking + COC bending C—O stretching COC bending
Adapted from Table III of V. Zamboni and G. Zerbi, Journal of Polymer Science C7 153 (1964). Copyright © (1964) John Wiley & Sons, Inc. Reprinted with permission of John Wiley & Sons, Inc.
Some of the splittings are quite appreciable, which shows that the intermolecular forces are comparatively large. As expected, the C—H stretching region is rather complex. Theory predicts one peak from the Bx mode and doublets from the B2 and B3 modes, but the fact that only four peaks are clearly resolved is not surprising, as this spectral region is likely to be overlaid by first harmonics of CH 2 bending modes, which will blurr some of the detail. The doublet at 428/434 cm " * is particularly clear and may be suitable for the determination of crystallinity.
6.7.4
Poly(vinyl chloride) Correlation splitting in poly(vinyl chloride) has been discussed briefly and formally in subsection 3.2.2 and it will now be considered in rather more detail because both the theoretical and experimental aspects of the topic present interesting features. In the earlier discussion it was shown that the presence of two chains in the orthorhombic unit cell of the syndiotactic polymer leads to correlation splitting but that,
6.7 Chain order and crystallinity
275
unlike the examples considered so far, only one component of each doublet is infrared active while the other is only Raman active. Hence the proof for the existence of the splitting requires the careful comparison of peak positions in the two types of spectrum. Furthermore, because the commercial type of polymer is only marginally syndiotactic and contains only about 10 % of crystalline material, the spectrum is complicated by the presence of peaks due to many different conformational and configurational isomers, as explained in subsection 6.2.3. This latter problem can be overcome by the use of the highly syndiotactic polymer made by the urea clathrate method. Nevertheless, only one measurement has been reported in which it was clearly established that the peak positions were determined with sufficient accuracy to prove the existence of the splitting. This study gave frequencies of 601 ± 1 and 608 + 1 cm" 1 for the antisymmetric C—Cl stretching modes in the infrared and Raman spectra, respectively, and the frequency 639 + 1 cm " x for the symmetric stretching mode for both spectra. The C—Cl stretching regions of the infrared and Raman spectra are shown in fig. 3.7. The corresponding calculated values for the antisymmetric stretching mode are 604 and 602 cm" 1 , and for the symmetric stretching mode they are 637 and 642 cm" 1 . The differences between the observed and calculated values and the fact that the values are reversed in order for the antisymmetric modes shows that the interchain force field used for the calculations is incorrect. This is confirmed by the fact that the observed splittings for the C—C—Cl bending vibrations of Ax and B2 species, at about 358 and 315 cm" 1 , respectively, are only about 3 cm" 1 or less, whereas the corresponding calculated splittings are 16 and 6 cm" 1 . Despite the difficulties involved in working with commercial polymers, which have a low degree of syndiotacticity, the values of 608 and 601 cm" 1 obtained for the antisymmetric C—Cl stretching mode from spectra of the urea clathrate polymer have been confirmed. It has been possible to fit the observed Raman spectrum of the commercial polymer with nine peaks over the interval 600-700 cm" 1 , including peaks at 608 and 638 cm" 1 (see fig. 6.7). Second and fourth derivative infrared spectra, which have sharper peaks than the normal spectrum (see section 1.7), also reveal peaks at about 601 and 638 cm" 1 . Furthermore, it has proved possible to use the sum of the areas of the peaks at 608 and 638 cm" 1 in the Raman spectrum of a powder sample, expressed as a fraction of the total area under all the v(CCl) peaks, as an approximate measure of crystallinity, although the complexity of this approach precludes it as a routine practical method.
276 6.7.5
The micro structure of polymers
Lattice modes in polyethylene and polytetrafluoroethylene As explained in subsection 3.2.1, the three-dimensional order present in polymer crystallites leads to the occurrence of lattice modes. These modes are vibrations in which the distortion of the crystal corresponds, to a first approximation, to the bodily movement of each chain with respect to its neighbours. Vibrations of this type have been most extensively studied for polymers containing long methylene sequences, and for polyethylene in particular. The far infrared spectrum of polyethylene at low temperatures has two absorption peaks at about 80 and 108 cm" 1 . The assignments of these two peaks, based partly on the results of normal coordinate calculations, are well established. The one at 80 cm" 1 results from a translational vibration in which the two chains in the unit cell move in opposite directions parallel to the b axis of the cell, whereas that at 108 cm ~* has its origin in a similar translational vibration in which the chains move parallel to the a axis. The interpretation of the low frequency spectrum of polytetrafluoroethylene (PTFE) has proved more challenging and the problem has been attacked by a combination of normal coordinate calculations and careful experimental measurements, including studies of the polarized infrared spectra of oriented samples. The early experimental studies showed that several new peaks appear in the far infrared region when the temperature of the sample is below the transition temperature of 19°C (see subsection 5.2.4) and that splittings also appear in several peaks in the Raman spectrum. These observations were difficult to reconcile with the assumption that the change in structure at the transition temperature is purely intramolecular and involves only a transformation from a 157 to a 136 helix, since the most natural explanation would be that the low temperature form has at least two chains per unit cell. Early X-ray diffraction studies suggested, however, that both forms have a unit cell containing only one molecule. In an attempt to resolve this paradox it was suggested that the changes in the vibrational spectrum for samples at low temperature were a consequence of a relaxation of the helical selection rules, due to the effect of neighbouring chains, or to conformational disorder. The two possible interpretations of the spectroscopic data stimulated further low temperature studies. The original interpretation of the splitting observed in the Raman spectrum, as correlation splitting, led to the conclusion that the crystal structure was monoclinic, with space group P2X. This structure has two chains passing through the unit cell in such a way that they are related by a two-fold screw axis perpendicular to the chain axes and the factor group of the space group is isomorphous with the point group C 2 . For
6.7 Chain order and crystallinity
277
this factor group all five lattice modes are infrared active and it follows that three of them are A modes polarized parallel to the C2 axis, i.e. perpendicular to the chain axis, and as a consequence they should be observed as o modes in oriented samples. These are the libration of the two chains in the same direction around the chain axis, the translational vibration of the two chains with respect to each other parallel to the chain axis and the translational vibration of the two chains with respect to each other perpendicular to the C2 axis. The other two modes are B species modes and may have their oscillating dipoles anywhere in the plane perpendicular to the C2 axis, so that they are neither o nor n modes. These correspond to the translational vibration parallel to the C 2 axis and to the librational motion of the two chains in opposite directions. The experimental results show that there are at least seven absorption peaks in the region 30-90 cm" 1 , four of which, at 45, 54, 58 and 86 cm" 1 , are a polarized and the other three of which, at 32, 71 and 76 cm" 1 , are without strong dichroism. Several slightly different interpretations of these peaks have been given, but it is generally accepted that the peak at 32 cm ~x is probably to be assigned to the internal E1 mode on the dispersion curve labelled v8 in fig. 5.8 and that those at 45, 54, 58 and 71 cm" 1 are probably to be assigned to lattice modes. The latter four show appropriate polarization for lattice modes associated with C 2 symmetry and are found to decrease in intensity when the samples are irradiated with y-rays, which leads to a lowering of the crystallinity. If they were due to defects they would be expected to increase in intensity, and if they were due to changes in the selection rules for the low temperature phase they would be expected to maintain essentially constant intensities, since the crystal structure is not further changed by irradiation. In addition, lowering the temperature causes the peaks at 45, 54 and 58 cm" 1 to move towards higher frequencies, as would be expected for lattice modes, since thermal contraction of the lattice leads to increased interaction between the molecules. The observations on the low frequency vibrational spectrum thus clearly support the idea that the low temperature form of PTFE has a structure with two chains per unit cell. The results of recent X-ray and electron diffraction studies of the low temperature form have been interpreted in terms of a unit cell with two chains passing through it. The two chains have helical conformations of opposite chirality which correspond closely, but not exactly, to 136 helices. No definite space group symmetry could be assigned to the structure. The gradual elucidation of the structure and spectral assignments of the low temperature form of PTFE, which is not yet
278
The microstructure of polymers
complete, provides a good example of the complementary role that vibrational spectroscopy can play to conventional crystallographic studies of polymers. 6.7.6
Differentiation between crystallinity and chain order It is clear from the results considered in the present chapter that the major factor which differentiates the vibrational spectrum of polymer chains in crystalline regions from the spectrum of those in amorphous regions is the presence of order, i.e. translational symmetry, along the chain direction. The additional changes that result from lateral order, due to crystalline packing of the chains, are small. Not surprisingly, therefore, a careful distinction has not always been drawn between the two types of order and values of crystallinity have been reported which were actually based on peaks characteristic for order along the chain direction only. It will therefore be useful to summarize how the two types of order manifest themselves and to consider to what extent peaks characteristic for the latter type of order may be used, albeit empirically, as a measure of crystallinity. Over 20 years ago Zerbi, Ciampelli and Zamboni divided so-called regularity peaks into three types and as their system is logical, and has stood the test of time, it will be followed here. The first group of peaks characterize conformational regularity or irregularity, of which the most obvious example is trans/gauche rotational isomerism. For example, the peaks at 1450 and 1435 cm" 1 in the spectrum of trans-1,4polybutadiene are characteristic for the trans and gauche conformers, respectively, of the CH 2 —CH 2 bond. Such peaks may occur both in the solid state and in the melt or solution, although the ratio of the concentrations of the two conformers may be different for the two states. The trans isomer is often of lower energy than the gauche isomer and is then found in the crystalline state. It is unwise, however, to assume that the intensities of peaks characteristic for trans conformers are a direct measure of crystallinity. This has often been done for poly(ethylene terephthalate), for which peaks at 975, 1340 and 1470 cm" 1 in the infrared spectrum are characteristic for the trans conformer. It has, however, been shown that trans units occur in amorphous regions in concentrations which may depend on annealing temperature and time or on the degree of molecular orientation. Crystallinity values based on the intensity of a peak due to trans conformers may therefore be too high. The second class of regularity peaks, often erroneously called 'crystallinity peaks', consists of peaks associated with long stereoregular chain segments. Some examples have been discussed in section 6.2, of
6.7 Chain order and crystallinity
279
which the best known is isotactic polypropylene. This forms a regular helix with three monomer units per turn and it was noted that peaks at 805, 840, 898, 972, 995 and 1100 cm" 1 are characteristic for the helical structure. This type of chain regularity peak is usually absent from the spectrum of the molten polymer, but occasionally it is replaced by a much broader peak. Of the six peaks that are characteristic for the polypropylene 3X helix, all are absent from the spectrum of the molten polymer except the one at 972 cm" 1 , which is considerably broadened. A detailed analysis of the profile suggests that short helical units are still present in the melt. For this reason, although regularity of individual chains is a necessary condition for the occurrence of crystallinity, the intensity of a particular peak characteristic for such regularity may not be a direct measure of the crystallinity. The third class of regularity peaks comprises the true crystallinity peaks. Their occurrence is a consequence of inter-chain interaction and is predictable by group theory although, as is evident from the examples considered in subsections 6.7.2 to 6.7.4, the magnitude of the correlation splitting cannot, at present, be calculated with any degree of confidence. These true crystallinity peaks disappear when the sample is melted, but this may not prove a useful diagnostic test because of the instability of a number of polymers, e.g. polyoxymethylene and poly(vinyl chloride), at temperatures well below their melting points. Hence, there are often real practical difficulties, particularly with new or unusual polymers for which detailed vibrational analyses are not available, in diagnosing with certainty which, if any, of the observed peaks are specific for crystallinity and whose intensities may therefore be used as a measure of crystallinity, provided that allowance is made for complications such as those arising from preferred orientation. In these circumstances it is necessary to look for alternative approaches to the measurement of crystallinity. As noted above, modes specific for chains in amorphous regions are sometimes more suitable for quantitative work although here also it may not always be easy to identify such peaks with certainty. In addition, it must be recognized that the definition of the degree of crystallinity for a polymer is not straightforward. States of order intermediate between the truly crystalline state, with three-dimensional order over regions large compared with a monomer unit, and the truly amorphous state, with no order extending for distances more than a few monomer units, may exist. Each different method of measuring crystallinity may then lead to a different result. If, for a given polymer, samples are available covering a range of crystallinities, as determined by a technique such as X-ray diffraction or
280
The microstructure of polymers
density measurement, it may be possible to deduce an empirical correlation between this measure of crystallinity and the intensity of one or more peaks in the vibrational spectrum. Such correlations can be useful, but should be treated with caution if the reason for the correlation is not understood. For example, although the peak at 1141 cm ~1 in the infrared spectrum of poly (vinyl alcohol) has not been shown to be a crystallinity peak its intensity correlates with values of crystallinity obtained by X-ray diffraction methods. Similarly, for those ethylene/propylene/dicyclopentadiene terpolymers which contain more than 65 % of ethylene and are partially crystalline, correlations have been found between the intensity of the C—C stretching mode, which gives a peak at 1065 cm" 1 in the Raman spectrum, and the degree of crystallinity determined by X-ray diffraction. The virtue of this approach is that once a calibration in terms of a second technique has been established, vib rational spectroscopy usually provides a simple and rapid method for the measurement of crystallinity. 6.8
Chain folding in polyethylene In the preceding section, lateral order between polymer chains was considered in the important but specific context of interactions between chains in the crystalline unit cell and the dimensions of the crystallites were merely assumed to be much greater than those of the unit cell, so that surface effects could be neglected. As already pointed out in subsection 1.2.4, the crystals of polyethylene and various other polymers obtained from solution consist of lamellae a few hundred Angstroms in thickness, but with considerably larger lateral dimensions, and it is now well established by a variety of methods that the polymer chains fold back and forth so that their axes are approximately normal to the lamellar plane. This may be represented as shown schematically in fig. 1.7. Two assumptions have been made in drawing this figure: that the majority of the chains re-enter the lamella in a position adjacent to where they emerge, by tight chain folds, and that the folding takes place along a particular crystallographic plane. An alternative extreme assumption would be that the chains re-enter in positions randomly related to those from which they emerge, the so-called 'switchboard' model. The identification by electron microscopy of the growth face of the diamond-shaped lamellar crystal of polyethylene (see fig. 6.18a) as a (110) crystallographic plane has led to the conclusion that folding occurs in this plane, but it is possible to grow truncated crystals (see fig. 6.18fo) using different conditions, suggesting that chain folding may occur in the (200) plane. The morphological characteristics of polyethylene
6.8 Chain folding in polyethylene
281
crystallized from the melt are more difficult to observe, but the presence of lamellae has been established. Several experimental techniques have been used in attempts to decide whether or not adjacent re-entry predominates in chain-folded crystals and, if so, which are the predominant fold planes. Among these methods, infrared spectroscopy has been, and continues to be, important. Its contribution originated in 1968 with the work of Krimm and his colleagues. Their method is based on infrared studies of co-crystallized mixtures of polyethylene (PEH) and perdeutero-polyethylene (PED), and the principles may be understood by reference to fig. 6.19. Fig. 6.19a shows the arrangement, in a cross-section perpendicular to the chain direction, for a regular 1:1 mixed crystal of PEH and PED in which folding is assumed to occur in the (110) planes. It is evident that the unit cell contains two chain segments of PEH and two of PED, and there will therefore be correlation splitting of the various modes of both species. Since each chain now interacts with only two chains of the same kind but different orientation instead of the four such chains that it interacts with in the normal crystal, the splittings will be halved. They will, nevertheless, still be readily observable for some modes. Additionally, it should be possible to distinguish between adjacent and random re-entry, because if there is a significant amount of the latter there will be an appreciable number of isolated chains of either species, giving peaks at frequencies between the two components of the
Fig. 6.18. Electron micrographs of single crystals of polyethylene crystallized from dilute solution in xylene: (a) diamond-shaped crystals; {b) truncated crystals. (From (a) V. F. Holland and P. H. Lindenmyer, Journal of Polymer Science 57 589 (1962); (b) D. J. Blundell et al., Journal of Polymer Science B 4 481 (1966). Copyright © (a) (1962) {b) (1966) John Wiley & Sons, Inc. Reprinted by permission of John Wiley & Sons, Inc.)
^
(a)
•
^
(b)
282
The microstructure of polymers Fig. 6.19. Possible regular structures for 1:1 mixed crystals of normal and fully deuterated polyethylene: (a) regular folding in the (110) planes; (b) regular folding in the (200) planes. O indicates (CH2)n and # indicates (CD2)M. (From M. Tasumi and S. Krimm, Journal of Polymer Science Al 995 (1968). Copyright © (1968) John Wiley & Sons, Inc. Reprinted by permission of John Wiley & Sons, Inc.)
(a)
°T, (b)
st °t
±ZJ")_
X* X. °l X. °r ^ • t ' - .
r^
X.
T"
°i
&J
—°r
6.8 Chain folding in polyethylene
283
correlation split doublets. Reference to fig. 6.1% shows that for regular (200) folding the unit cell contains only one chain segment of each species and correlation splitting will therefore not occur. The absence of correlation splitting cannot, however, be regarded on its own as concrete proof for (200) folding because if there is extensive random reentry with (110) folding there will also be very little splitting. While these conclusions have been derived for regular alternating arrangements of PEH and PED chain segments, which lead to specific lattice structures, they should be valid for any crystal in which a (110) plane containing one species has nearest neighbour planes containing the other species, by the following line of reasoning. Consider a mixed crystal in which the concentration of PEH chains is very high, thus ensuring that the planes of PED chain segments are isolated from each other. A plane of PED chain segments may then be considered as a onedimensional lattice which, for (110) folding, has two equivalent but differently oriented oscillators per repeat period, whereas for (200) folding it has only one oscillator per repeat period. The former will give split modes, whereas the latter will not. Three modes of vibration were used for the experimental studies: the CH 2 bending and rocking modes observed at 1463/1473 and 720/731 cm" 1 , respectively, in pure PEH and the CD 2 bending mode observed at 1084/1091 cm" 1 in pure PED. The results of the original measurements of the splittings on a series of solution crystallized and melt crystallized (cast film) mixtures of PEH and PED of various compositions are given in table 6.5, which also shows calculated values of the splittings for pure crystals of either type. The calculated values for the pure PEH and PED are in good agreement with the experimental values. The limited number of calculated values for mixtures were somewhat lower than those observed, but the difference is not unexpected, as any aggregation of chains will increase the observed splitting. The fact that for cast films splitting of the modes of the minor component is observed even when this forms only 10 % of the mixture shows that chain folding occurs predominantly in the (110) plane with some adjacent re-entry. Furthermore, the absence of an obvious third peak between the components of the doublet suggests that the folding in fact occurs mainly by adjacent re-entry. The spectra of melt-crystallized samples showed a broad single peak for the species at low concentration, with some residual doublet structure occurring as shoulders, and this suggested that either (200) folding or random re-entry predominates for samples of this type. Mixed crystals of PEH and PED with random re-entry were simulated by preparing and examining mixtures, of varying composition, of the
Table 6.5. Splittings of bending and rocking modes in mixed crystals of PEH and PED Avb(CH2) (cm-
Av r (CH 2 )(cm"
Avb(CD2) (cm- l) Melt crystallized
Cast film
Melt crystallized
Cast film
10.3 9.8 10.1 9.8
10.5 10.4 9.9 9.6
— 5.5 5.9 6.6
— 0 0 0
10.8 10.7 10.4 10.3
11.3 11.0 10.5 9.9
1:1
9.3
8.8
6.7
4.8
9.5
9.0
1:2 1:4 0:1 0:1 (calc.)
8.3 7.4
0 0 —
7.1 7.5 7.7
6.8 7.3 8.0
8.7 8.4 —
0 0
PEH:PED
Cast film
l:O(calc.) 1:0 10:1 4:1 2:1
Melt crystallized 10.5
9.9
8.1
Adapted from Table IV of M. I. Bank and S. Krimm, Journal of Polymer Science A2 71785 (1969). Copyright © (1969) John Wiley & Sons, Inc. Reprinted by permission of John Wiley & Sons, Inc.
6.8 Chain folding in polyethylene
285
normal paraffin C 3 6 H 7 4 and its fully deuterated analogue. Since there is no chain folding in paraffin crystals, the distribution of hydrogenated and deuterated chains in these mixed crystals will be random. The observed splittings for the species at high concentration in the paraffin mixtures were found to be appreciably less than for the corresponding melt crystallized polymer mixture. It was therefore deduced that there must be a substantial amount of adjacent re-entry in the polymer mixes and, as a corollary, that the chain folding is largely in the (200) plane. It must be noted, however, that the lattice dimensions of melt-crystallized polyethylene and normal paraffins are different at room temperature, and this conclusion is not supported by more recent work. The interpretation of the results in the way indicated depends upon the assumption that there is no segregation of the PEH and PED chains, since otherwise it does not follow that adjacent chain segments of the species with low concentration necessarily belong to the same molecule. It has been shown that although segregation can take place, because of a difference in the melting points of the two species, the conditions used in the original work were close to those required to eliminate this effect. Fig. 6.20. The CD 2 bending region of the infrared spectrum of a PED/PEH mixed crystal grown from dilute solution in xylene and containing 0.5% PED of MW 216 000. The spectrum was obtained with the sample at the temperature of liquid nitrogen and is shown fitted with a singlet and two doublets; the sum of these is shown by the dashed line. Further increasing the number of doublets improves the fit. (Reproduced from S. J. Spells et al., Polymer 25 749 (1984), by permission of the publishers, Butterworth & Co. (Publishers) Ltd. ©)
1070
1080 1090 Wavenumber, v (cm"1)
1100
286
The microstructure of polymers
More recent studies have been made with samples cooled to liquid nitrogen temperature and it has become clear that the structures of the CH 2 and CD 2 bending regions of the spectrum are more complex than originally thought. A detailed analysis of the regions into sets of overlapping peaks (see fig. 6.20) has shown that for solution crystallized samples they can be understood in terms of a model in which a molecule folds in the (110) plane to give a sheet containing 10 to 20 parallel chain segments. The folding is largely, but not completely, accompanied by adjacent re-entry, so that the chain segments occur within the sheet as blocks separated by blocks of chain segments from another molecule. For high molecular weights the molecule 'superfolds' to give multilayer sheets of this type. The model is also consistent with the results of neutron scattering experiments and was in fact originally developed to explain those results. For melt-grown crystals the more recent infrared studies, in which the broadened single peak observed for the CD 2 bending mode has been analysed into several components, suggest that random re-entry predominates but that there is possibly a small amount of (110), rather than (200), adjacent re-entry. The results of neutron scattering studies suggest a similar structure, in which groups of a few chain segments related by adjacent or nearly adjacent re-entry occur among chain segments for which re-entry is largely random.
6.9
Longitudinal acoustic modes As long ago as 1949, in a general study of the Raman spectra of solid normal paraffins (n-paraffins), it was noted that there was a peak in the low frequency region whose frequency appeared to be inversely proportional to the chain length. This observation was confirmed and extended nearly 20 years later, as was the original suggestion that this peak has its origin in an accordian-like motion of the all-trans chain present in the crystals. There is maximum displacement of the atoms at the surface of the crystal and a node of zero displacement at the centre of the crystal, as shown in fig. 6.21b. This motion will change the polarizability of the crystal and the vibration will thus be Raman active. It is also evident that this vibration is the fundamental of a series of longitudinal acoustic modes (LA modes or LAMs). The higher members of the series will have two or more nodes within the crystal, as shown in fig. 6.21c for three nodes. It is clear that those with an even number of nodes will not produce a net change of polarizability for the crystal as a whole and so will not be Raman active, whereas those with an odd number of nodes will give peaks in the Raman spectrum at frequencies which, to a first approximation, will be the appropriate multiple of the
287
6.9 Longitudinal acoustic modes
fundamental frequency. Such higher order LAMs are readily observable in the spectra of the n-paraffins. To first approximation, a LAM of any order should correspond to a point on the dispersion curve v5 for polyethylene (see fig. 4.13 and section 4.5), but if the corresponding wavelength is large enough it is simpler to consider the vibrating chain as an elastic rod. Elementary theory shows that the vibrational frequency, v, of such a rod vibrating in a mode with m nodes is given by the equation: m
(6.3)
where / is the length of the rod, E is the Young's modulus and p is the density. This relation fits the experimental results for solid n-paraffins well in the limit where l/m is much greater than the translational repeat unit and it is possible to obtain a 'spectroscopic' value for the Young's
Fig. 6.21. The displacements of the repeat units and the forms taken by the backbone of the molecule in the longitudinal acoustic modes (LAMs) of polyethylene (schematic); (a) shows the equilibrium state of the chain backbone and (b) and (c) show the first and third order LAMs. In {b) and (c) the upper and lower diagrams show the forms taken by the backbone in opposite phases of the motion and the central diagram shows the corresponding displacements, with displacement to the right shown above the straight line representing the undisplaced positions. The full curves refer to the upper diagrams and the dashed curves to the lower diagrams. The magnitudes of the distortions of the backbone are greatly exaggerated.
<«> wwwwwwwwwwwwv vwwwwvwwwvwvwwv (b)
vwwwwwvwwwwwwv WWWWWV\AAAA/WWWWV (c)
V\MA/VWWVWWWWWWW
288
The microstructure of polymers
modulus of polyethylene from the limiting slope, at v = 0, of a plot of v against m//, since the density of crystalline polyethylene is known. At the time that the later work on n-paraffins was in progress the lamellar nature of the single crystals of polyethylene obtained from solution (see subsection 1.2.4) was well established, as was the fact that the lamellar thickness is, typically, 200 A. It was therefore possible to deduce that the lowest LAM frequency for the chain segments in the lamellae occurs in the range 10-20cm" 1 , and the corresponding peak will be difficult to observe as it will lie on the 'wing' of the intense line due to Rayleigh scattering. For this reason, the LA mode was not observed in polyethylene until 1971. In this initial study it was clearly established, as anticipated, that the frequency decreased as single crystals were annealed and so became thicker. The change in lamellar thickness with annealing temperature, as assessed by this approach using the value of Young's modulus obtained from the measurements on n-paraffins, proved to be very similar to that deduced from small angle X-ray diffraction measurements. Fig. 6.22 shows the peaks due to the LA mode in the Raman spectra of two polyethylene samples prepared in different ways. In the last decade, with the emergence of the holographic diffraction grating which gives a high discrimination against the Rayleigh line, it has proved comparatively easy to work within about 6 cm " * of this line, and there have been numerous studies on the LA mode of polyethylene and, to a lesser degree, on that of other polymers which crystallize in the lamellar form. Higher order LA modes have been observed, but most effort has been concentrated on the lowest order LAM because of its greater intensity. There have also been a number of theoretical treatments of the LA mode, prompted largely by anomalies which appeared in the experimental results. This work has proved interesting and has provoked several questions. For example, is it possible to determine lamellar thicknesses with reasonable accuracy from the positions of the corresponding LAM peaks by using the value of Young's modulus derived from measurements on solid n-paraffins? In view of the all-trans conformation required for the LAM vibration, what will be the effect of an occasional defect structure in the crystalline region, such as a gauche unit or a chain branch? Is it correct to assume that the LAM frequency is determined only by the length of the straight chain segments ('stems') in the crystal lamellae so that it is unaffected by the presence of disordered 'amorphous' material between lamellae or chain folds at their surfaces? Since a range of lamellar thicknesses will occur in a given sample, the observed LAM peaks will be the concentration-weighted sum of the peaks for each particular stem
289
6.9 Longitudinal acoustic modes
length, and so a broader composite peak will be obtained; is it possible to obtain information on the distribution of lamellar thicknesses from the shape of this peak? These topics will now be considered briefly. The early work on n-paraffins gave a value of 3.6 x 1011 N m~ 2 for Young's modulus. This value was derived by assuming that the Fig. 6.22. The low wavenumber regions of the Raman spectra of two samples of linear polyethylene, showing the fundamental LAM. Each sample was compression moulded at 160°C. Spectrum (a) was obtained from a sample quenched directly into cold water from 160° and spectrum (b) was obtained from a sample which was slowly cooled to 110°C before being quenched.
(b)
r/ > yj-
Jh 1
Observed v / intensity j
/
— Fitted exciting-line tail
/
\
\
- ^
\
-****
i
(a)
//—-Fitted / exciting-line / tail
Observed __ intensity
j
X. \\
y/_ i
30
y Fitted LAM
1
20 10 Wavenumber, Ai> (cm"1)
290
The microstructure of polymers
frequency depends only on the length of the sequence of CH 2 groups in an individual stem but later work indicated that weak interlamellar forces affect the peak position and an estimate of the magnitude of this effect led to a new value of 2.9 x 1011 N m " 2 . This is in quite good agreement with the value of 3.3 x 1011 N m~ 2 obtained from inelastic neutron scattering data for the v5 dispersion curve of polyethylene. It is also very close to the value obtained by direct stress/strain measurements on highly oriented samples of polyethylene and since these are not perfect single crystals the true crystal modulus must be somewhat higher, probably close to that obtained from the neutron scattering data. Even if the correct crystal modulus and the observed LAM peak frequency are substituted into equation (6.3) the value of / obtained will not be the mean thickness of the lamellae in the sample, as might be expected, since it has been shown that the peak position depends on the temperature of the sample and the half-width of the peak. These factors will not be discussed in detail and it will suffice to note that even in the favourable circumstances of a narrow peak with a half-width of about 6 cm ~ 1 , corresponding to a narrow distribution of sequence lengths, the correction is about 1 cm ~x for a peak at 20 cm ~*; for a half-width three times larger the corresponding correction is about 7 cm ~ x . The correction, which reduces the observed frequency and so leads to a higher value for the mean lamellar thickness, rises rapidly with decreasing wavenumber of the peak position and may be comparable with the measured value. In these circumstances it is not surprising that the absolute measurement of lamellar thicknesses from observed LAM frequencies poses considerable problems and the method is of much greater value for following the changes that occur during treatments such as annealing. It should also be noted that if the chains are inclined at an angle to the normal to the lamellar surface, as is often observed, the angle of tilt must also be established. Nevertheless, by combining the results of Raman measurements with those from small angle X-ray diffraction studies, which give the repeat distance of the 'sandwich' structure of lamellar crystalline and amorphous layers which usually occurs, it has been possible to deduce reasonably reliable values for the thickness of the amorphous layers. The evidence for the effect of defect structures comes primarily from normal mode calculations for models in which one or more successive gauche units are introduced near to the centre or to the end of a long straight chain segment. They show that such a structure near to the chain end has little influence on the LAM vibration but when it is near to the chain centre it reduces the intensity of the peak considerably without
6.9 Longitudinal acoustic modes
291
affecting its position perceptibly. Several pieces of evidence suggest that the presence of chain folds and amorphous material between lamellae does not affect the LAM frequency corresponding to a given stem length. One important piece of experimental evidence is provided by the results of measurements following treatment with nitric acid. This material selectively attacks the amorphous material and the chain folds, leaving isolated stems in the lamellae, but the LAM frequency does not change significantly after this reaction. This evidence is supported by the results of calculations based on an extension of the elastic rod model in which a composite rod is considered, which has one length, density and modulus for that part of the molecule contained within a lamella, and two ends of different lengths, densities and moduli for the parts of a particular chain lying within amorphous regions. Using this model it is possible to obtain a modified expression for the LAM frequency which, when applied to the specific case of polyethylene, shows that the frequency is only modified slightly by the fact that the stems traversing the lamellae are parts of chains that also cross the interlamellar noncrystalline regions. A model of a similar nature, but which treats the chain more correctly at the microscopic level, shows similarly that the dispersion curves obtained by neutron scattering should be largely unaffected by the interlamellar disordered material. These conclusions probably apply only for melt-crystallized samples, since there is some evidence that the LAM frequencies of 'mats' of sedimented solutiongrown single crystals may be influenced by interactions between the lamellar surfaces, which consist of more regular chain folds than do those of melt-grown crystals. As noted above, if there is a range of lamellar thicknesses the LAM peak will be broadened because of the superposition of the peaks, somewhat displaced with respect to each other, characteristic for each stem length. Not surprisingly, there has been considerable interest in the possibility of deducing information about the distribution of lengths from the width of the LAM peak. Unfortunately, the half-width of the chain-length distribution function is not obtained from the observed intensity distribution simply by converting the frequencies at the halfintensity points to chain lengths using equation (6.3). Just as the peak position depends on temperature and the distribution of stem lengths, as discussed above, so does the shape of the peak. The following equation relating the distribution of stem lengths, /(/), and the line shape, J(v), has been derived: f(l) = Cv2B(v)I(v)
(6.4) c/fcT
where C is a constant, B(v) is the factor 1 — e~^
and / is the stem
292
The microstructure of polymers
length which gives rise to the LAM frequency v. An additional fact which must be taken into account is that the LAM peak has a finite width even for a sample consisting of stems of identical length; measurements on crystalline n-paraffins lead to half-width values of 2.5-3.5 cm "* for peak positions ranging from 6 to 30 cm" 1 . The effect of these corrections, which have not been applied in several published studies, is to reduce the spread of the calculated stem lengths. For a particular sample of solution-crystallized polyethylene, the corrected spread was found to be as low as 15 A, in line with the results of small angle X-ray measurements, whereas an earlier study in which the appropriate corrections were not applied gave a spread of 40 A. Although LAM studies on polyethylene have attracted the most attention, a range of other polymers has been examined. These fall into two types: those that have (CH2)n sequences as part of the overall structure and those that do not. The former category includes materials such as poly(decamethylene sebacate), poly(alkylene oxide)s and poly(alkylene sulphide)s. The second category comprises a range of ordered polymers, such as polytetrahydrofuran, cis-l,4-polybutadiene, polytetrafluoroethylene, polyoxymethylene, poly(ethylene oxide) and isotactic polystyrene. Suitable short-chain molecules related to these polymers as the n-paraffins are to polyethylene are not available, nor generally are neutron scattering data for constructing dispersion curves, so that spectroscopic values of Young's modulus cannot always be obtained. It has also been shown that values determined in other ways are usually less reliable. The use of LA mode frequencies in morphological studies of these polymers is therefore more difficult than it is for polyethylene and will not be discussed here. 6.10 6.10.1
Molecular orientation Introduction In earlier sections it has been explained how infrared spectra obtained from samples in which the molecular chains have been preferentially oriented, by drawing or stretching the sample, can be used as an aid to the assignment of vibrational modes. For this purpose it is usually desirable for the chains to be as highly oriented as possible, and in the ideal sample all chain axes would be parallel to the draw or stretching direction. Many polymers in practical use are oriented to some degree, either intentionally, to improve the mechanical properties, or incidentally, during injection moulding, cold forming or other processes. This latter type of orientation may lead to undesirable distortions of the products when they are heated, because of relaxation of the orientation. Both intentional and incidental orientation are often
6.10 Molecular orientation
293
very far from complete and it is important to be able to characterize the degree of orientation so that the properties of the products may be more fully understood and usefully modified. A number of methods have been devised for studying orientation in polymers. X-ray diffraction is often the preferred method for studying the orientations of crystallites, particularly if the degree of crystallinity is high, but it is quite difficult to obtain reliable information about the orientation of the non-crystalline chains from X-ray diffraction studies. Among the other available methods, infrared and Raman spectroscopies using polarized radiation are particularly important because they can potentially give a great deal of information, about both crystalline and non-crystalline chains, since measurements on any group mode can give information specifically about the distribution of orientations of the corresponding groups in the polymer. Infrared spectroscopy provides the simpler method, but it cannot give as much information about the distribution of orientations as Raman spectroscopy because it involves only one beam of polarized radiation and the fraction of the radiation absorbed by a single molecule depends only on the orientation of the molecule with respect to the polarization vector of the radiation. Raman scattering, in contrast, involves two beams of radiation and the intensity scattered by a molecule depends on the orientation of the molecule with respect to the polarization vectors of both beams, which may be different. Suitable measurements of scattered intensities can thus give more information about molecular orientation. In addition, infrared absorption measurements are usually restricted to samples of thickness less than about 50 /im, because thicker samples are opaque to the infrared radiation, whereas Raman scattering can sometimes give good results on samples as thick as 1 cm. In the next subsection the theoretical background to the methods will be given and in subsection 6.10.3 some examples of their use. 6.10.2
Theoretical background We shall assume that the symmetry of the distribution of orientations in the sample is such that a set of three mutually perpendicular axes OXYZ may be chosen within it so that rotating it through 180° about any one of them does not change the distribution of orientations with respect to a fixed observer. If the distribution is unchanged by rotation through any arbitrary angle around one of these axes, the sample is said to be uniaxially oriented, or transversely isotropic, with respect to that axis; otherwise it is said to be biaxially oriented. These terms originate from crystal optics and are used because the optical properties of the samples are similar to those of uniaxial and
294
The microstructure of polymers
biaxial crystals, respectively. It is usual to label the unique direction of a uniaxial sample OZ. We shall concentrate in detail on the simplest type of uniaxial orientation and the information that can be obtained by means of infrared spectroscopy. A brief indication will then be given of the additional information which can be obtained if the samples are biaxial or if Raman spectroscopy is used, or both. We assume that the sample is made up of only one type, or possibly a small number of different types, of structural unit and that the units of any one type, such as a crystallite or a non-crystalline chain segment, are identical. Each unit is imagined to have embedded within it a set of axes Oxyz. In the simplest form of uniaxial orientation each type of unit has a special axis with the property that if this axis lies in a particular direction for a typical unit, all orientations of units obtained by rotation around that axis are equally represented in the distribution. If Oz is chosen to coincide with this special axis we can then say that there is no preferred orientation of the structural units around their Oz axes. For this type of uniaxial sample the distribution of orientations of a particular type of unit can be described fully by a function N(6) such that N(6)d(o measures the fraction of units for which the Oz axes lie within any infinitesimal solid angle dco at angle 6 to the OZ axis and 2n JQ N(9) sin 6 dO = 1. The aim of orientation studies is to determine as much as possible about the form of N(6). It is often convenient to expand the function N(6) in terms of Legendre polynomials, as follows:
N{6)=
h
where, because of the orthogonality of the Legendre functions, it follows that the coefficients ax are the averages,, of the corresponding polynomials P^cos 6) taken over the distribution of orientations. For distributions in which equal numbers of chains point in opposite directions only even values of at are non-zero. The lowest values of ax are then a0 = 1
(6.6a) 2
a2 == ^(3
(6.6b)
a4 == i(3 - 30
(6.6c)
Consider an infrared-active vibrational mode for which the oscillating dipole moment /i of a particular unit makes the angle 0o with Oz and 6^ with OZ, as shown in fig. 6.23. If infrared radiation, of the appropriate frequency, polarized parallel to OZ falls normally on a sample in the
295
6.10 Molecular orientation
form of a thin sheet with its plane containing the OZ axis, the intensity / | transmitted will be given by (see subsection 5.4.1)
*•({;)-*.« and similarly the intensity transmitted perpendicular to OZ will be
(6.7a) if the polarization
is
(6.7b) where / 0 is the incident intensity, fc| and kL are the mean absorptivities for a single structural unit for radiation of the corresponding polarization, n is the number of units per unit volume and t is the thickness of the sample. Since the energy absorbed by an individual unit is proportional to the square of the component of the electric vector Fig. 6.23. The orientation of a structural unit and its attached infrared dipole moment \i with respect to axes OXYZ fixed in a uniaxially oriented sample, with OZ parallel to the symmetry axis (draw direction) of the sample. Oz is the special axis of the unit around which it has no preferred orientation. The points P and Q lie in the OXY plane.
296
The microstructure of polymers
parallel to the oscillating dipole moment, it follows that fcn = 3/c2
(6.8a) 2
kL = 3/c<sin 0M cos >M>
(6.8b)
where k is the mean absorptivity per structural unit for a randomly oriented sample, for which < c o s 2 ^ > = 3 , and depends on the vibrational mode but not on the distribution of orientations. Because of the uniaxial nature of the sample, 0M takes random values, independently of 0M, and thus /c± = |/c<sin20/x>
(6.9)
Equations (6.8a) and (6.9) show immediately that 3/c = /c | +2/c 1
(6.10)
and
It follows from equations (6.7) that logl f I + 2 log! -± I = A,, + 2A± = nf(fc, + 2/c±) = 3ntk = 3A
(6.12)
lo=
-?—logl — I H- 2 log! — 1
Ay+2AL
D+2
(6.13)
where A^ and A± are the absorbances of the sample for radiation polarized parallel and perpendicular, respectively, to the draw direction, OZ, and D is the dichroic ratio defined in section 4.2.1. The quantity A is the absorbance of a randomly oriented sample with the same value of nt. The average value of cos 2 0^ or, equivalently,*s ^ e o n ' v quantity related to molecular orientation which may be determined directly by means of infrared spectroscopy. It may be shown, however, that
= Pf(cos 0o)
(6.14)
6.10 Molecular orientation
297
for all /, so that if the angle 90 is known, a2 or, \j/, where the angle \j/ measures rotation around the Oz axis of the unit. This function may be expanded in terms of a set of generalized spherical harmonics, of which the Legendre functions are the simplest type. Equations (6.8) are then replaced by three equations which relate the absorptivities for radiation polarized parallel to OX, OY or OZ linearly to the averages of four second order generalized spherical harmonics. The coefficients in these equations depend on the angles 90 and 0 which specify the orientation of the oscillatory dipole with respect to Oxyz. Equations (6.8) could have been written in a similar way by the use of equations (6.6) and (6.14). It is therefore possible in principle, by making absorption measurements on more than one absorption peak for which 90 and >0 are known, to solve the resulting equations for the four orientation averages. Similarly, the intensities of Raman scattering from a sample with the most general type of biaxial orientation usually considered may be expressed, for all possible combinations of polarization vectors of incident and scattered radiation, as linear functions of the averages over the distribution of orientations of the four second order and an additional nine fourth order generalized spherical harmonic functions. The coefficients in these expressions are second order functions of the Raman tensor components referred to the symmetry axes of the unit. For the simplest type of uniaxial orientation the only non-zero averages are and and if the ratios of the principal Raman tensor components and the orientation of the principal axes within the structural unit are known it is possible to determine these two averages, or . One of the difficulties of the Raman method is that this information about the Raman tensor is not always available. In addition, for biaxially oriented samples it is not possible to determine all the relevant averages from measurements on one mode and although, in principle, measurements on several Raman-active modes could be combined, this has not yet been satisfactorily done.
298
The micro structure of polymers
In the treatment given above, corrections for reflection at the surfaces of the sample and for certain 'internal field' effects have been neglected, but they must be taken into account in order to obtain accurate results from either the infrared or Raman method. Studies of orientation using vibrational spectroscopy have been made on a number of polymers and some of the more detailed of these will now be briefly described.
6.10.3
Orientation in polyethylene and polyethylene terephthalate) Infrared studies of uniaxially oriented polyethylene have been made by a number of authors. In one study of linear polyethylene two sets of samples of very different molecular weight distribution were drawn at 60°C to draw ratios between 5 and 25. Some of the samples were prepared before drawing by slow cooling from the melt and some were prepared by quenching from the melt. Dichroic ratios were determined for the 1894 cm " x crystal mode and the 1078,1303 and 1368 cm " x amorphous modes. The 1894 cm" 1 mode is assigned to a combination of the Raman active Ag methylene rocking mode at 1170 cm" 1 and the infrared-active B2u/Blu methylene rocking modes at 720/731 cm" 1 and its dipole moment is perpendicular to the crystal c-axis. The amorphous mode at 1078 cm" 1 is assigned to gauche C—C stretching with some CH 2 wagging and the 1303 and 1368 cm" 1 modes are assigned to asymmetric and symmetric CH 2 wagging modes for the CH 2 —CH 2 group where the C—C bond is at the centre of a gauche-trans-gauche sequence. Fig. 6.24 shows a plot of dichroic ratio, D, against draw ratio for the four modes. D 1 8 9 4 rapidly approaches zero with increasing draw ratio, which indicates full orientation of the c-axes of the crystallites parallel to the draw direction. The D values for the three amorphous modes approach saturation values which do not indicate full alignment either parallel or perpendicular to the draw direction and this is clearly in accord with their assignments, since the corresponding groups are unlikely to become very highly oriented. In addition, the approach to saturation, particularly for the 1368 cm" 1 mode, is much more gradual than that for the crystal mode and this has been interpreted as evidence for a clear segregation of the crystalline and amorphous phases. An extensive series of studies has been performed on oriented samples of poly(ethylene terephthalate) by several spectroscopic techniques, including infrared and Raman. The first quantitative Raman studies were performed on a series of tapes drawn at 80°C to draw ratios up to 5.9. They were assumed to have the simplest type of uniaxial orientation and the Raman-active mode used was that at 1616 cm " x , assigned to a
6.10 Molecular orientation
299
good group vibration of the benzene ring with the symmetry species Ag with respect to the D 2h symmetry of the paradisubstituted ring (see fig. 6.25). The Raman tensor was assumed to have cylindrical symmetry around Oz0 and the ratio a x /a z was determined from the linear depolarization ratio for a random sample. This assumption has been shown to be reasonable by a study of the Raman scattering from single crystals of the model compound bis(2-hydroxy ethyl) terephthalate, HO(CH 2 ) 2 O . O C — < Q ) - C O . O(CH 2 ) 2 OH. The values offor the chain axes were calculated from the observed scattering intensities by assuming that the molecular conformation of all
Fig. 6.24. Dichroic ratio, D, plotted against draw ratio for various infrared absorption peaks of samples of linear polyethylene drawn at 60°C. The wavenumbers of the peaks are indicated for each curve and the assignments are given in the text. O» slowly cooled Marlex 6050; # quenched Marlex 6050; A slowly cooled Marlex 6002; A quenched Marlex 6002. (Reproduced by permission from B. E. Read in Structure and Properties of Oriented Polymers, ed. I. M. Ward, Applied Science Publishers (1975), adapted from W. Glenz and A. Peterlin, Journal of Macromolecular Science Physics B4 473 (1970), by courtesy of Marcel Dekker, Inc.)
10
15 Draw ratio
300
The microstructure of polymers Fig. 6.25. Approximate form of the vibration assigned to the peak observed at 1616 cm" 1 in the Raman spectrum of poly(ethylene terephthalate). The set of axes used in discussing the orientation of the ring is shown.
Fig. 6.26. A plot of
0.4 0.00
0.04
0.08
0.12 0.16 Birefringence, An
0.20
0.24
6.11 Further reading
301
chains in the polymer was similar to that in the crystalline regions except for the possibility of gauche bonds in the —CH 2 —CH 2 — groups. The results were compared with the birefringences of the samples, i.e. the differences between the refractive indices for light polarized parallel and perpendicular to the draw direction, which may be shown to be approximately equal toArcmax, where Anmax is the birefringence of a fully oriented sample. A straight line resulted (see fig. 6.26) which extrapolated to the value An = 0.24 at
=l, in good agreement with the value for Arcmax calculated for a fully oriented sample from bond polarizability data. The points representing pairs of values of
6.11
Further reading
The following books and reviews, in addition to those referred to in section 5.7, should be found particularly useful for further reading on the topics of this chapter: The Chemical Microstructure of Polymer Chains by J. L. Koenig, Wiley, 1980; Infrared and Raman Spectroscopy of Polymers by H. W. Siesler and K. Holland-Moritz, Dekker, 1980; 'Polymer Characterisation by Raman Spectroscopy' by D. L. Gerrard and W . F . Maddams in Applied Spectroscopy Reviews, 22(2-3) 251 (1986); Biological Spectroscopy by I.D. Campbell and R. A. Dwek, Benjamin-Cummings, 1984; 'Determination of Molecular Orientation
302
The micro structure of polymers
by Spectroscopic Techniques' by I.M. Ward in Advances in Polymer Science 66 81 (1985). A list of references to some of the more important papers on the topics covered in this chapter and in chapter 5 is given on the following pages. References given in figure captions or in connection with tables may also be found useful.
References for chapters 5 and 6
The following are some specific references to topics in chapters 5 and 6. References given in figure captions or in connection with tables will also be found useful but they are not repeated here.
5.2 J. J. Fox and A. E. Martin, 'Investigations of infrared spectra. Determination of C—H frequencies (~3000cm" 1 ) in paraffins and olefins, with some observations on "polythenes"', Proc. Roy. Soc. A175 208-33 (1940). M. C. Tobin, 'Selection rules for normal modes of chain molecules', J. Chem. Phys. 23 891-6 (1955). J. R. Nielsen and A. H. Woollett, 'Vibrational spectra of polyethylene and related substances', J. Chem. Phys. 26 1391-400 (1957). C. Y. Liang, G. B. B. M. Sutherland and S. Krimm, 'Selection rules and frequencies of the skeletal vibrations of long chain polymers in the crystalline state', J. Chem. Phys. 22 1468-9 (1954). J. R. Nielsen and R. F. Holland, 'Dichroism and interpretation of the infrared bands of oriented crystalline polyethylene', J. Mol. Spectroscopy 6 394-418 (1961). R. G. Snyder and J. H. Schachtschneider, 'Vibrational analysis of the n-paraffins - 1 . Assignments of infrared bands in the spectra of C 3 H 8 through n-C 1 9 H 4 0 '. '.. .-II. Normal coordinate calculations', Spectrochim. Acta 19 85-116; 117-68 (1963). R. G. Snyder, 'A revised assignment of the B2g methylene wagging fundamental of the planar polyethylene chain', J. Mol. Spectroscopy 23 224-8 (1967). J. Barnes and B. Fanconi, 'Critical review of vibrational data and force field constants for polyethylene', J. Phys. Chem. Ref. Data 1 1309-21 (1978). S. Krimm and C. Y. Liang, 'Infrared spectra of high polymers. IV. Polyvinyl chloride, polyvinylidene chloride, and copolymers', J. Polymer Sci. 22 95-112 (1956). W. H. Moore and W. Krimm, 'The vibrational spectrum of crystalline syndiotactic poly(vinyl chloride), Makromol. Chemie Suppl. 1 491-506 (1975). J. L. Koenig and F. J. Boerio, 'Raman scattering and band assignments in polytetrafluoroethylene', J. Chem. Phys. 50 2823-9 (1969); 'Refinement of a valence force field for PTFE by a damped least squares method', J. Chem. Phys. 52 4826-9 (1970). C. Y. Liang and S. Krimm, 'Infrared spectra of high polymers. VI. Polystyrene', J. Polymer Sci. 27 241-54 (1958). R. W. Snyder and P. C. Painter/Normal coordinate analysis of isotactic polystyrene: I. A force field for monosubstituted alkyl benzenes'; ' . . . II. The normal modes and dispersion curves of the polymer', Polymer 22 1629-32, 1633^1 (1981). F. J. Boerio, S. K. Bahl and G. E. McGraw, 'Vibrational analysis of polyethylene terephthalate and its deuterated derivatives', J. Polym. Sci.: Polym. Phys. 14 109246 (1976).
303
304
References for chapters 5 and 6
5.4 J. G. Grasselli, M. A. S. Hazle, J. R. Mooney and M. Mehicic, 'Raman spectroscopy: an update on industrial applications', Proc. 21st Colloq. Spectrosc. Int., Heyden, London, 1979, pp. 86-105. R. S. Silas, J. Yates and V. Thornton, 'Determination of unsaturation distribution in polybutadienes by infrared spectrometry', Anal. Chem. 31 529-32 (1959). 5.6 G. Martinez, C. Mijangos, J. L. Millan, D. L. Gerrard and W. F. Maddams, 'Polyene sequence distribution in degraded poly(vinyl chloride) as a function of the tacticity', Makromol. Chemie 180 2937^5 (1979). D. L. Gerrard, W. F. Maddams and J. S. Shapiro, 'Influence of annealing on the formation of long polyenes during PVC degradation', Makromol. Chemie 185 184354 (1984). 6.1 M. M. Coleman and J. Zarian, 'Fourier-transform infrared studies of polymer blends. II. Poly(£-caprolactone)-poly(vinyl chloride) system', J. Polym. Sci.: Polym. Phys. 17 837-50 (1979). C. Tosi, P. Corradini, A. Valvassori and F. Ciampelli, 'Infrared determination of sequence distributions and randomness in copolymers', J. Polym,. Sci. C 22 108592 (1969). I. V. Kumpanenko and K. S. Kazanskii, 'Analysis of IR-band shapes in studies on the monomer sequence distribution in macromolecules', J. Polym. Sci. Symp. 42 973-80 (1973). M. Meeks and J. L. Koenig, 'Laser-Raman spectra of vinyl chloride-vinylidene chloride copolymers', J. Polym. Sci. A2 9 717-29 (1971). 6.2 G. Natta, 'Precisely constructed polymers', Scientific American 205 No. 2 33—41 (1961). R. G. Snyder and J. H. Schactschneider, 'Valence force calculations of the vibrational spectra of crystalline isotactic polypropylene and some deuterated polypropylenes', Spectrochim. Acta 20 853-69 (1964). M. Peraldo and M. Cambini, 'Infrared spectra of syndiotactic polypropylene', Spectrochim. Acta 21 1509-25 (1965). R. G. Snyder and J. H. Schactschneider, 'Valence force calculations of the vibrational frequencies of two forms of crystalline syndiotactic polypropylene', Spectrochim. Acta 21 1527-42 (1965). 6.3 C. Baker and W. F. Maddams, 'Infrared spectroscopic studies on polyethylene. 1. The measurement of low levels of chain branching', Makromol. Chemie 111 437-48 (1976). M. A. McRae and W. F. Maddams, 'Infrared spectroscopic studies on polyethylene. 2. The characterisation of specific types of alkyl branches in low-branched ethylene polymer and copolymers', Makromol. Chemie 111 449-59 (1976). G. S. Park, 'Short chain branching in poly(vinyl chloride)', J. Vinyl Technol. (1985).
References for chapters 5 and 6
305
6.4 S. Crawley and I. McNeill, 'Preparation and degradation of head-to-head PVC\ J. Polym. Sci.: Polym., Chem. 16 2593-^06 (1978). 6.5 C. Baker and W. F. Maddams, 'Infrared spectroscopic studies on polyethylene. 1. The measurement of low levels of chain branching', Makromol. Chemie 111 437-48 (1976). M. A. McRae and W. F. Maddams, 'Infrared spectroscopic studies on polyethylene. 3. The bromination of unsaturated groups', Makromol. Chemie 111 461-71 (1976). W. F. Maddams, 'Spectroscopic studies on the characterisation of defects in the molecular structure of poly(vinyl chloride)', J. Vinyl TechnoL, 1 65-73 (1985). 6.6 L. R. Schroeder and S. L. Cooper, 'Hydrogen bonding in polyamides', J. Appl. Phys. 47 4310-17 (1976). V. W. Srichatrapimuk and S. L. Cooper, 'Infrared thermal analysis of polyurethane block polymers', J. Macromol. Sci. Phys. B 15 267-311 (1978). 6.7 F. J. Boerio and J. L. Koenig, 'Raman scattering in crystalline polyethylene', J. Chem. Phys. 52 3425-31(1970). M. Glotin and L. Mandelkern, 'A Raman spectroscopic study of the morphological structure of the polyethylenes', Colloid and Polymer Sci. 260 182-92 (1982). M. E. R. Robinson, D. I. Bower and W. F. Maddams, 'Intermolecular forces and the Raman spectrum of syndiotactic P V C , Polymer 17 355-7 (1976). W. Frank, H. Schmidt and W. Wulff, 'Complete temperature dependence of the far infrared absorption spectrum of linear polyethylene from 14°K to the melting point', J. Polym. Sci. Polym. Symp. 61 317-26 (1977). J. F. Rabolt, 'Characterization of lattice disorder in the low temperature phase of irradiated PTFE by vibrational spectroscopy', J. Polym. Sci.: Polym. Phys. 21 1797-805 (1983). G. Zerbi, F. Ciampelli and V. Zamboni, 'Classification of crystallinity bands in the infrared spectra of polymers', J. Polym. Sci. C 1 141-51 (1964). 6.9 R. F. Schaufele and T. Shimanouchi, 'Longitudinal acoustic vibrations of finite polymethylene chains', J. Chem. Phys. 47 3605-10 (1967). W. L. Peticolas, G. W. Hibler, J. L. Lippert, A. Peterlin and H. Olf, 'Raman scattering from longitudinal acoustical vibrations of single crystals of polyethylene', Appl. Phys. Lett. 18 87-9 (1971). B. Fanconi and J. F. Rabolt, 'The determination of longitudinal crystal moduli in polymers by spectroscopic methods', J. Polym. Sci.: Polym. Phys. 23 1201-15 (1985). 6.10 A. Cunningham, I. M. Ward, H. A. Willis and V. Zichy, 'An infrared spectroscopic study of molecular orientation and conformational changes in poly(ethylene terephthalate), Polymer 15 749-56 (1974).
306
References for chapters 5 and 6
D. I. Bower, 'Investigation of molecular orientation distributions by polarized Raman scattering and polarized fluorescence', J. Polym. Sci.: Polym. Phys. 10 2135-53 (1972). D. I. Bower, D. A. Jarvis and I. M. Ward, 'Molecular orientation in biaxially oriented sheets of poly(ethylene terephthalate) I. Characterization of orientation and comparison with models', J. Polym. Sci. B 24 1459-79 (1986).
A note on the use of the indexes
It is not possible to cite in the main index every occurrence of the use of many of the basic concepts introduced in chapters 1-4, and in general only selected examples have been given there. It is hoped that judicious use of the specialised indexes, in conjunction with the main index, will enable others to be readily found. The following abbreviations have been used in the indexes, in addition to more obvious ones: ATR f.f. FTIR H-bonding h-h h.p. IR LA(M) l.d. PBD PE PED PEH PIB PMMA POM PP PS PTFE PVC PVDC R res. R
307
attenuated total reflection spectroscopy force field Fourier transform infrared spectroscopy hydrogen bonding head-to-head high pressure infrared spectroscopy or spectrum longitudinal acoustic mode low density polybutadiene polyethylene deuterated polyethylene; potential energy distribution ordinary hydrogenated polyethylene polyisobutene poly(methyl methacrylate) polyoxymethylene polypropylene polystyrene polytetrafluoroethylene poly(vinyl chloride) poly(vinylidene chloride) Raman spectroscopy or spectrum resonance Raman spectroscopy or spectrum
Index of spectra illustrated
copolymer of ethylene and propylene IR fig 6.2 232; various compositions, CH 2 rocking region, 660-900cm" 1 ethylene and vinyl acetate IR fig 6.1 230; v(C=O) region, 1690-1770cm"1, resolved into triads normal paraffin C 1 2 H 2 6 IR fig 5.2 170; 7OO-15OOcm"1, sample at - 1 9 6 ° C Novolak resin IM184 IR fig 5.17 212; cast film, 400-3800cm" 1 Nylon 6,6 IR fig 5.18 213; cast film, 400-3700cm" 1 polybutadiene IR fig 1.8 14; 1,2-polybutadiene, 65O-35OOcm~1 R fig 5.19c 220; 1600-1700 c m 1 polyethylene IR and R fig 5.1 164; 400-4000cm" 1 IR fig 5.13a 207; h.p. type, 400-4000cm" 1 , compared with PIB and PP IR fig 6.8 248; branched PE, methyl absorption region, 1200-1700cm"1 IR fig 6.10 253; Phillips, l.d. and Ziegler-Natta polymers, 800-1100cm" 1 R fig 6.15 268; y,(CH2) doublet, 1290-1300cm" 1 , effect of temperature IR fig 6.20 285; <5(CD2) region of mixed PED/PEH crystal, 1067-1104cm"1 R fig 6.22 289; LAM region, 8-30cm" 1 , effect of thermal history poly(ethylene terephthalate) IR fig 4.5 121; oriented sample, polarized spectra, 770-1070cm" 1 IR and R fig 5.11 201; 200-3200cm" 1 , showing interference fringes in IR polyisobutene IR fig 5.13c 207; compared with PE and PP spectra, 400-4000cm" 1 poly(methyl methacrylate) IR fig 5.15 209; cast film, 400-3700cm" 1 IR fig 5.21 222; overtone region showing 6135cm" 1 monomer peak polypropylene IR fig 5.13b 207; atactic, 400-4000cm" 1 , compared with h.p. PE and PIB IR fig 6.3 236; iso-, syndio- and atactic, polarized spectra, 700-1600cm" 1 polystyrene IR and R fig 5.9 195; atactic sample, 400-4000cm" 1 R fig 5.19a 220; 1500-1700cm" 1 , compared with sample toughened with PBD toughened with polybutadiene R figs 5.19b 220; 1500-1700 cm"1, compared with pure PS spectrum polytetrafluoroethylene IR and R fig 5.7 189; 7O-15OOcm"1, polarized IR in region 35O-15OOcm~1 poly(vinyl acetate) IR fig 5.14 209; cast film, 400-3800cm" 1 poly(vinyl chloride) IR and R fig 3.7 102; v(CCl) region, 540-720cm" 1 , correlation splitting IR and R fig 5.3 174; 400-3700cm" 1 , R spectrum shows 'fluorescence' res. R fig. 5.22 225; thermally degraded sample, 50-4000cm" 1 IR fig 6.5 241; v(CCl) region, 590-720cm" 1 , samples polymerized at +50, - 7 8 ° C R fig 6.7 243; v(CCl) region, 500-800cm" 1 , showing fitted Lorentzian peaks
308
309
Index of group modes
poly(vinylidene chloride) IR and R fig 5.6 183; single crystals and quenched sample, 5O-33OOcm~1 polyurethane IR fig 6.12 260; v(NH) region, 3200-3600 cm" \ for £(CH2)8OCONH(CH2)6NHCOOJn
Index of point groups Cj Cs Cn Cx C2 C3 C 13
54, 66, 88, 118, 128-30, 201 54, 66, 88, 90-3, 118, 127, 178, 181-5 54, 62-3, 65-6, 88, 95-7, 127, 151 54, 88, 90, 93, 118 88, 93, 95, 118, 182, 194, 276-7 127, 199,235 187-8
C(12TC/13)
187-9
Cnh 54,62-3 C 2h 88, 95, 103, 105, 118, 128-30, 169 Cnv 54, 65 C2v 72-3, 88, 90-3, 95, 100-1, 107, 110, 118, 127, 169, 174-9, 181-2, 195-8, 237 C 3v 55-6 C4v 59-60,64,118-19 C^ 54 Dn 54, 63, 65, 88, 95-7, 127 D2 54,66,88,95-7,107,118,127,181-4, 235,271^
D 3 127 D6 96-7 D9 271-3 D 1 3 187-94 D(12TC/13) 187-94 Dnd 54 D2d 54, 107 D 3d
120
Dnh 54,65 D2h 54,66,77,84-8,90-3,95,100-6,107, 118, 128-30, 154-6, 163-73, 265-7, 274-5, 276, 298-9 D 4h 116-17 D6h 128-30,197-8 D^ 54,114 Sn 54,62-3,65-6 S2 54 S6 54 Td 55
Index of group modes See, in addition, fig 4.9, p 133 and fig 5.12, p 204 amide I, II, III modes, 204,212^, 223,258 backbone see skeletal below benzene for modes of benzene or modes involving a substituted benzene ring, see main index: benzene C—C stretching in a terpolymer, 280 in polyethylene, 298 in polypropylene, 208, 235-8 in poly(vinyl chloride), 226 see also skeletal modes below C = C stretching in allyl alcohol, 218 in ethylene, 77-8 in polybutadiene, 218-9
C = C stretching (continued) in poly(vinyl chloride), 226, 255 in unsaturated hydrocarbons, 206 in unsaturated polyethylene, 254 C—C—C bending in polypropylene, 237 see also skeletal modes below C—C—Cl bending in poly(vinyl chloride), 275 C—(CH3) stretching in polypropylene, 208, 237 C—C—O bending in polyethylene terephthalate), 202-3 C—Cl bending of —CC12— groups in poly(vinylidene chloride), 186 of lone CC1 groups in poly(vinyl chloride), 176, 179, 240 C—Cl2 rocking in poly(vinylidene chloride), 186
310
Index of group modes
C—Cl stretching in carbon tetrachloride, 149 of —CC12— groups in poly(vinylidene chloride), 181-5 of lone CC1 groups in poly(vinyl chloride), 102-3, 176, 178-9, 240-5, 275 C—Cl2 twisting in poly(vinylidene chloride), 186 C—Cl wagging of —CC12— groups in poly(vinylidene chloride), 186 of lone CC1 groups in poly (vinyl chloride), 176, 179 CD 2 bending in deuterated polyethylene, 283-6 CF 2 bending in polytetrafluoroethylene, 192-4 CF 2 rocking in polytetrafluoroethylene, 194 CF 2 stretching in polytetrafluoroethylene, 192-4 CF 2 twisting in polytetrafluoroethylene, 192-4 CF 2 wagging in polytetrafluoroethylene, 194 C—H bending of —CH 2 — groups in polyethylene, 85-7, 154-5, 165-9, 173, 199, 266-8, 270-1, 283-6 in poly(ethylene terephthalate), 239 in polyoxymethylene, 274 in polystyrene, 199 in poly(vinyl chloride), 175, 177-8, 217 240 in poly(vinylidene chloride), 186 of = C H 2 groups in ethylene, 77-8 of CH 3 groups in polyethylene, 247-50 in polyisobutene, 208 in polypropylene, 207 of lone C—H groups in polypropylene, 208, 235-8 in polystyrene, 199 in poly (vinyl chloride), 175, 177-8 = C H 2 out of plane deformation of vinylidene end group, 252-4 CH 2 rocking in adipic and sebacic acids, 214 in ethylene/propylene copolymers, 231-3 in polyethylene, 85-7,155-8,165-9,171, 199, 206, 265-7, 269, 283-5, 298 in polyoxymethylene, 274
CH 2 rocking (continued) in polypropylene, 237 in polystyrene, 199 in poly(vinyl chloride), 175, 178 in poly (vinylidene chloride), 186 of ethyl branches in polyethylene, 250 (CH 2 ) 2 modes and trans/gauche isomerism in polyethylene terephthalate), 239^0,278 trans 1,4 polybutadiene, 278 in h-h poly(vinyl chloride) and polypropylene, 251 CH 3 rocking in polyethylene branches, 249, 252 in polyisobutene, 208 in polypropylene, 208, 237 C—H stretching and early IR instruments, 23 influence of nearby oxygen atom, 203 in methane, 135, 149 of —CH2— groups in polyethylene, 85-7, 154-5, 165-7, 173, 176, 199 in poly(ethylene terephthalate), 203 in polyoxymethylene, 274 in polystyrene, 199 in poly (vinyl chloride), 175-7 in poly (vinylidene chloride), 184-6 of = C H 2 groups in ethylene, 77-8, 131-2 of CH 3 groups in polypropylene, 206-7 of lone CH groups in poly (vinyl chloride), 175-7 CH2 twisting in polyethylene, 85-7,106,136-7,155-8, 165, 167-71, 206, 267, 271 in polystyrene, 199 in poly (vinyl chloride), 175, 178 in poly (vinylidene chloride), 186 CH wagging of —CH2— groups in polyethylene, 85-7,136-7,155,164-5, 167-9, 171-3, 199, 247, 267, 269, 271, 298 polyethylene terephthalate), 239 polystyrene, 199 poly(vinyl chloride), 175, 178 poly (vinylidene chloride), 186 of lone CH groups in poly(vinyl chloride), 175, 177-8 C—N stretching in poly amides, 213 C = N stretching in acrylonitrile copolymers, 218
Index of polymers C—O stretching in acetates, 208-9 in acrylates, 210 in aliphatic carbonates, 208 in cellulose acetate, 208 in poly(ethylene terephthalate), 202-3,210 in polymethacrylates, 209-10 in polyoxymethylene, 274 in polyurethanes, 214 in saturated esters/polyesters, 208-10 in terephthalates, 202-3, 210 C = O rocking in poly(ethylene terephthalate), 202 C = O stretching and hydrogen bonding, 259-64 in aromatic polyesters, 210 in biopolymers, 264 in carbonyl compounds, 223 in carboxylic acids, 223-4 in cellulose acetate, 208 in COO" ion 223-4 in dilauryl thiodipropionate, 223 in oxidised polyalkenes, 223-4 in poly amides, 213 in poly(e-caprolactone) blends 230-1 in polyethylene terephthalate), 202-3, 301 in poly(methyl methacrylate), 209 in polyurethanes, 214, 259-61 in poly (vinyl acetate), 229-30 in saturated aldehydes, 224 in saturated esters/polyesters, 202-3, 208-10, 301 in saturated ketones, 224 in vinyl acetate copolymers, 217, 229-30 of acetate groups in polyoxymethylene, 255
311 C = O stretching (continued) variation of freq. with environment, 203^ C = O wagging in PET 202 C—O—C bending in polyethylene terephthalate), 202-3 in polyoxymethylene, 274 C—O—C stretching in an aliphatic ether, 212 in polyurethanes, 214 in resole resins, 212 N—H in-plane deformation in polyamides, 213 N—H stretching and early IR instruments, 23 and hydrogen bonding, 257-64 in biopolymers, 262 in Nylon 6, 6 and Nylon 10, 258 in polyamides, 213, 257-8 in polyurethanes, 258-61 O—H stretching and early IR instruments, 23 and hydrogen bonding, 257 in novolak resins, 211-12 in phenolic antioxidants, 222 in resole resins, 212 of end groups in polyoxymethylene, 255 out-of-plane deformation of unsaturated groups in polyethylene, 252^, 269 vibrations in polybutadiene, 218-19 skeletal modes in polyethylene, 85-7, 155, 165, 167-71, 173, 267 polyisobutene, 208 polypropylene, 208, 235-8 poly(vinyl chloride), 176-9, 226 YX 2 stretching vibrations, 166, 184
Index of polymers See in addition table 5.10, p. 205, for a polymer classification scheme by characteristic infrared frequencies. amino resins, 7 biopolymers, 3, 32, 213, 256, 262-4 blends of polymers, 15-6, 214, 216 polybutadienes, 218-19 poly(e-caprolactone) and PVC, 230-1 cellulose, 256
cellulose acetate, 208 collagen, 262-4 copolymers of ethylene and 1-alkenes, 248 ethylene and propylene, 231-3 ethylene and vinyl acetate, 217-18, 229-30 ethylene, propylene and dicyclopentadiene, 280 Nylons, 214 styrene and acrylonitrile, 218
312
Index of polymers
copolymers of (continued) styrene and butadiene, 219 vinyl chloride and vinyl acetate, 217 vinyl chloride and vinylidene chloride, 233-4 DNA, 264 elastomers, 31 epoxy resins, 255-6 gutta percha (trans 1,4-polyisoprene), 10 novolak resins, 210-12 Nylons, 5-6, 212-14, 220-1, 257-8 perspex: see poly (methyl methacrylate) phenolic resins, 7, 210-12 plexiglass; see poly (methyl methacrylate) polyacetates, 208-9 polyacetylene, 4, 226 polyacrylates, 208-10 polyacrylonitrile, 3-4 see also copolymers poly(alkylene oxide)s, LA modes, 292 poly(alkylene sulphide)s, LA modes, 292 polyamides, 5-6,204,212-14,220-1,257-8 polybutadiene, 4, 218-19, 278, 292 see also copolymers poly(butyl methacrylate), 210 polycaprolactam (Nylon 6), 6, 214, 220-1 poly(e-caprolactone), 230-1 poly(decamethylene sebacate), 292 polydicyclopentadiene, 280 polydienes, 4, 10 see also individual polymers poly(diethylene glycol bisallyl carbonate), 208 polyenes, low MW, 226 polyester polyurethanes, 6, 214-15, 259, 261 polyesters, 5-6, 208-10, 214 see also individual polymers polyether polyurethanes, 6,214-15,259-61 polyethers, 4 see also individual polymers polyethylene additives, defects, end groups and degradation, 11-12, 163, 167-8, 172, 222-4, 246-50, 252-4, 269 chain dimensions, 187 copolymers; see copolymers cross linking, 254, 269 crystal structure and symmetry, 103-6, 166, 265 crystallinity, morphology and chain folding, 13, 156, 163, 246, 280-92 deuterated, 136-7, 166, 267, 281-6
polyethylene (continued) discovery of, 166 dispersion curves; finite chains, 154-8, 168-72, 192, 287, 290 effects of crystallinity in, 103-6, 156-7, 163, 166-8, 171, 173, 231, 247, 265-71, 276, 281-6, 298 longitudinal acoustic mode, 34, 286-92 model compounds, 157, 166-73, 247-8, 269 molten, 166, 271 non-crystalline chains, 166, 168, 172, 247, 269, 271 orientation, 166-7, 172, 266, 269, 290, 298-9 overtones and combinations, 163,167-8, 170, 222, 267 perfect single chain, 81-91, 154-6, 163-73 polymerization and types, 3,172, 246-7, 249, 250, 252-4 sample thickness for IR, 29 single crystal texture, 172-3 single crystals, 13, 280-6, 288, 291-2 vibrational calculations, 154, 166-8, 170-2, 267, 276 poly(ethyl methacrylate), 210 poly(ethylene adipate), 214 polyethylene oxide), 214, 292 poly(ethylene terephthalate), 7 assignments and recognition, 200-3,210 model compounds, 134, 202-3, 240, 299 orientation, 121, 240, 298-301 polymerization; structure; symmetry, 6, 128, 200-1 rotational isomerism and amorphous regions, 11,202-3,239-40,278,299 vibrational calculations, 203 polyisobutene, 207-8 polyisoprene, 4, 10 polymethacrylates, 208-10 poly(methyl acrylate), 210 polymethylene, 247 see also polyethylene poly(methyl methacrylate), 4, 13, 29, 208-10,221-2 poly(a-methyl styrene), 4 polyoxymethylene, 4,252,255,271^, 279, 292 polypeptides, group frequencies, 213 polypropylene, 3-4 additives, defects, end groups and degradation, 223-4, 251, 254 conflgurational isomerism and spectrum, 234-8 copolymers; see copolymers
Main index polypropylene (continued) group frequency approach, 206-8 polymerization, 3-4, 234 structure and symmetry, 3-4,231,234-5, 279 vibrational calculations, 208, 235, 237 poly(propylene oxide), 214, 233 polyribo-adenylic acid, 264 polyribo-uridylic acid, 264 polystyrene, 3-4,31,194-200,218-20,251, 292 see also copolymers polyterephthalates, 210 see also poly(ethylene terephthalate) polytetrafluoroethylene, 4, 185-94, 276-8, 292 polytetrahydrofuran, 259, 261, 292 poly(tetramethylene adipate), 261 polythene, see polyethylene polyurethanes, 5-6, 206, 214-15, 258-62 poly(vinyl acetate), 3 ^ , 208-9, 230 see also copolymers poly(vinyl alcohol), crystallinity, 280 poly(vinyl chloride) blend with poly(6-caprolactone), 230-1 copolymers; see copolymers defects, end groups and degradation, 224-6, 246, 250-1, 255, 279 deuterated, 173, 177-8, 240
313 poly(vinyl chloride) (continued) effects of crystallinity, 100-3, 274-5 isomerism, amorphous regions and spectrum, 10, 134, 179, 239, 240-5, 275 models for, 134, 240-2 spectral assignment of, 173-80 structure/symmetry, 3 ^ , 100-3, 128, 174 syndiotactic, 100-3, 274-5 factor group modes of cry St., 100-3 from urea clathrate, 179, 240, 275 vibrational calculations, 173,177-9,240, 275 poly(vinyl fluoride), 8 poly(vinylidene chloride), 4, 180—6 see also copolymers poly(vinylidene fluoride), 8 resole resins, 210-12 Rigidex9, 247 rubber, natural (cis 1,4-polyisoprene), 10 vinyl polymers, 3-4, 8-10, 16 symmetry and vibrational modes, 89-93 symmetry and selection rules, 126-8 see also individual polymers vinylidene polymers, 3-4, 8-9 see also individual polymers
Main index absorbance, infrared definition, 20 determination, 23, 26 for oriented samples, 296 ratios in analysis, 217-18, 247, 256, 269 absorption coefficient, 20,215,218-19,248 absorptivity, molar, 215, 259-60 acceptor group, in H-bonding, 256,259-60 acetate end groups in POM, 255 acetylacetone, 202 acetylene, 4, 227 acoustic modes: see modes, acoustic activated species, 3 activity, spectral, 2, 65, 8 1 ^ , 100-1,105-6 and crystal field, 98-100, 105-6, 265 for model 'polymer' chain, 154 see also selection rules addition polymerization,3-4, 8 additives, 219-20, 222-3 adipic acid, 5, 214
adjacent re-entry, 13, 280-6 aldehydes as oxidation products, 224 aliphatic alcohols, 214 aliphatic amides, 223 aliphatic carbonates, 208 aliphatic polyesters, 214 alkanes; see paraffin 1-alkenes, 246, 248, 252 alkyl benzenes, 200 alkyl chlorides; see chlorides allyl alcohol, 218 allyl chloride, 227 allylic chlorine group, 224,255 amide group/linkage, 5-6, 204, 212-14, 223, 256-8 amino resins, 7 ammonia molecule, symmetry of, 55, 65 amorphous regions; see non-crystalline regions amplitude of vibration, 85, 139, 221
314
Main index
analysis chemical, 218 spectroscopic, 15-16, 133 qualitative, 162, 203-15 quantitative, 215-26, 242, 247, 252-6 angular momentum, 74-5 anharmonic vibrations, 136, 158-61, 221 anisotropy of tensor, 1 2 3 ^ see also orientation annealing and lamellar thickness, 288,290 antioxidants, 220, 222 antisymmetric species, 52, 65-6, 74 antisymmetric stretching modes, 46, 74-5, 275 see also index of group modes application of known assignments, 162 various applications, 215, 301 areas under peaks, 229-30, 242, 271, 275 aromatic isocyanates, 214-15 see also benzene assignment, 162 for various polymers, 163-203 meaning and methods, 122,125-38,149, 292 associative law and group elements, 53 asymmetric carbon atom, 233 asymmetric peaks, 161, 229 atactic polymer, definition, 8 atomic radius, 8, 181 atomic structure and nucleus, 42-4 ATR, attenuated total reflection, 20, 30-1 axes choice of orientation and symmetry species designation, 66, 74,81,85-6,92-3,100,103,237,271 within structural unit, 294 symmetry, definitions and types, 47-52, 65, 81, 93 axial glide; see glide, axial axial point group, 54, 65, 118-20 back-biting, 12, 254 basis for representation, 58-63, 67, 70, 77, 117-19 Beer's law; see Lambert-Beer law bending; see modes, bending; index of group modes benzene, 128 fundamental and symmetry modes, 195-8 ring monosubstituted, 195-200 paradisubstituted, 128, 201-2 see also: phenolic resins; polyesters; polystyrene; polyurethanes in index of polymers
biaxial orientation, 266, 293, 301 binding energy of molecule, 42 biopolymers, 3, 32, 213, 256, 262-^ birefringence, 300-1 bis(2-hydroxy ethyl) terephthalate, 299 blends (of polymers), 15-16, 216, 219, 230-1 blowing, 223 Boltzmann factor, 113 bond angle, 1, 75-7, 147-8, 181-2 chemical, nature of, 1, 42 dipole, 109-11, 265-6 double and cis/trans isomerism, 10 and end groups or defects, 227, 252-5 and resonance Raman, 224-6 in monomers, 4, 221 electrostatic; see hydrogen bond(ing) hydrogen; see hydrogen bond(ing) length, 76, 147-8 rotation around and trans/gauche, 9-10 bonding, tetrahedral, 8, 47 Born's cyclic boundary conditions, 150 borrowing of intensity, 197 branch/branching and degradation in PVC, 224, 246 as defect structure, 8, 11 effect on crystallinity, 246 longitudinal acoustic mode, 288 mechanism for, 11-12 studied by spectroscopy, 246-50 see also methyl group(s) breathing mode of methane, 135 bridged structures, 7, 210 broadening of peaks asymmetric, 161, 229 in copolymer, 229, 233 in the melt, 279 involving non-factor group modes, 161 see also width of peaks bromination of polymers, 252-5 1,4-butanediol in polyurethanes, 259 calculation of frequencies; see vibrational calculations calibration for analysis, 247, 252, 255 for crystallinity determination, 280 caprolactam, 6, 220-1 carbamate groups, 214 carbon dioxide, normal modes, 45-6 disulphide, 218, 223 tetrachloride, 47-8, 149, 221
Main index carbonyl compounds/group, 202-3, 223 and hydrogen bonding, 256, 259-64 see also C = O in index of group modes carboxylic acids, 223-4 casting of samples, 29 catalyst, 233, 234, 246, 254 cellulose, 256 cellulose acetate, 208 central force field; see force field centre of inversion, 49 centrosymmetric molecule, 81-9, 114-16, 200-1 chain folding, 13, 280-6, 288, 291 transfer during polymerization, 12 terminating groups; see end groups number per unit cell, 98-106, 264-5; see also splitting, correlation regular, 7, 15, 80-96; see also: order; isolated regular chain character of a representation matrix, 63-5 contribution of each atom, 71-2, 84 table, 64^7, 72-3,79,84-5,109,118,168 chemical analysis, 218 chemical repeat unit, 2, 15, 81, 95 chirality, 94^5 chlorides, secondary as models for PVC, 134, 240-2 as models for PVDC, 185 force constants, 173^, 185 chlorine atom, diameter, 181 chloromethyl branches in PVC, 227, 250 chloroparaffins 2-chlorobutane, 174, 242 2-chloropropane, 241 see also chlorides chromatography, 30, 220 CH 3 group; see methyl group CH 4 (methane), 135, 149 cis configuration/conformation, 10-12, class of operation, 64^5, 71, 72, 95 coatings, 20, 31, 36 collagen and hydrogen bonding, 262-4 coloured samples, 33 column matrix; see matrix, column comb, in dispersive, IR, 23 combination; see overtone/combination frequencies combination rule in group theory, 53-6, 57 commutation of elements/operators, 53,56 complex eigenvectors, 63, 143, 150-1, 154 complex matrices, 62-3, 119, 143 composite material, 258 computers and IR, 21, 26-7, 247, 255
315 concentration; see analysis condensation reactions, 5-7, 214 configuration; see isomer/isomerism conformation; see isomer/isomerism conformer; see isomer/isomerism, conformational conjugated double bonds, 224-6 convolution, 40-1 cooling; see temperature coordinates Cartesian, 67-8, 138-9, 152 complex, 63, 143 for helical molecule, 150-1 generalized, 139-41 internal, 76-7, 139, 146-8 internal symmetry, 77, 130-1, 146 mass weighted, 69, 141 normal, 69-70, 75, 111-14, 142-3, 145 rotational and translational, 146 symmetry, 74^8, 139, 145-6, 178 copolymer analysis, 16, 216-19, 227-31 and crystallization, 227 and polymerization mechanisms, 227-8 block, 217, 228, 258 definition, 2, 8 distribution of units, 227, 229-34 interactions in, 217, 228-31 random, 217, 228-9, 231 segmented block, 258 see also index of polymers correlation splitting; see splitting tables for descent in symmetry, 92 coupled oscillators, 1, 156, 168, 192 coupling of group vibrations, 133, 154 see also mixing cross-linking, 7, 210-12, 218,254,256, 269 crystal field; see field, crystal size and spectrum, 156-7 structure hexagonal, 271 monoclinic, 276 orthorhombic, 100, 103, 265, 271 spectroscopic study of, 99, 264, 276-8 crystallinity and isomerism, 234, 240, 250, 278 and order, 278-80 degree of, 163, 246, 275 determination of degree of by density measurement, 269, 279-80 spectroscopy, 264, 269-71, 275, 278-80 X-ray diffraction, 269, 279-80
316
Main index
crystallinity (continued) effects on properties, 13, 234 spectrum, 97-106, 126, 166-7, 264-80; see also splitting, correlation; field, crystal peaks, 278-9 crystallization and polymer morphology, 12-13 effect of branching on, 246 in copolymers, 227, 231 crystallites dimensions of, 13-14, 270, 280, 288 lamellar, 13-14, 270, 280-6, 288-92 crystals and regular chains, 15, 80 single, 13-14, 266, 280-1, 288, 291 vibrational modes and activity, 17, 97-106, 113, 156-7,264-5 cubic groups and symmetry, 55,66, 125 cubic terms in potential energy, 159, 161 curing of resins, 210, 256 curve fitting, 39-40, 215, 229-30, 240-3, 261, 275, 285-6, 289 cyclic boundary conditions, 150 cyclohexane, 257 damping, 225-6 data processing and IR, 26-7, 247, 255 deconvolution, 40-1 defect structure(s), 8 and copolymerization, 227 effect on LAM, 288, 290-1 quantitative information on, 217 see also: branch/branching; end groups; isomer/isomerism degeneracy accidental, 58, 161 lifting/loss of, 265,272; see also splitting, crystal field degenerate modes; see modes, degenerate degradation, 223-7, 246, 251, 252, 255, 279 degrees of freedom, 44 denaturation of proteins, 262—4 density and crystallinity, 269, 279-80 depolarized Raman peaks, 125, 180 depolarization ratio, 123-5, 299 derivative spectra, 40 descent in symmetry, 89-93, 99, 105 detectors for IR spectroscopy, 27-8 for Raman spectroscopy, 35-6, 167 deuterated polymers polyethylene, 136-7, 166, 267, 281-6 other polymers, 173,177-8,199,202,240
deuterium substitution, 135-7 di-phenyl methane-4,4'-di-isocyanate, 214 diagonal glide, 104 diagonal matrix, 61, 70; see also matrix diagonal plane, 49 diatomic molecule, 42-4, 49 dibutyl tin dilaurate, 224 a,co-dichloroalkanes, spectra, 172 dichroism/dichroic ratio, 120-2, 296, 298-9 see also polarization difference frequency, 161; see also overtone/combination frequencies spectroscopy, 39,221,223,224,247,253, 255 diffraction, electron, 13, 277 diffraction grating, 17, 24, 34-5, 288 dihedral axis/plane, definitions, 49 dihydroxyethylene terephthalate, 202 dilauryl thiodipropionate, 223 dimethyl terephthalate, 6, 202 dimethylhexatetracontane, 247 diphenylmethane di-isocyanate, 214, 259 dipole (moment), electric, 107-12 addition and IR, 110-11, 175, 265-6 and IR selection rule, 109-11 and orientation studies, 294 of bond or segment, 109-11, 121, 265-6 direct product group, 66 direction cosines, 108, 123 disorder, spatial, 265 disordered structures and group modes, 15, 206 dispersion curves, 150-8 for model 'polymer' system, 152-4 for PE,154-8,168-72,191-2,287,290-1 for PTFE, 192-4, 277 non-crossing and repulsion, 154-6 dispersive spectrometers, 21-4, 34-6 distearone, 224 distortion of oriented samples, 292 distribution of lamellar thicknesses, 289, 291-2 of orientations of units, 293-301 DNA, 264 donor group, in H-bonding, 258 double bond; see bond, double double helix of DNA, 264 drawing (stretching), 120, 240, 242, 273, 292, 298, 301 dynamic studies, 27, 36 eigenvector (s) and symmetry, 47, 52-3, 70, 74 and vibrational calculations, 138-45
Main index eigenvectors (continued) as sum over symmetry mode vectors, 74-5, 145 assignment to, 125, 130, 134 approximate, symbols for, 85, 163 complex, 63, 143, 150-1, 154 definition, 45-7 for degenerate modes, 46, 57-63, 70 for modes in crystals, 98-9, 103, 265-6, 268 for PE 171 normalization, 69, 142-3 orthogonality of, 69, 141-3 elastic rod model and LAM, 287, 291 elastomers and ATR, 31 electric dipole; see dipole electric field; see field electromagnetic spectrum, 17-18 electron diffraction, 13, 277 electron microscopy, 280-1 electron volt, definition, 18 electronegative atom, 256 electronic states/transitions, 43, 44, 224-5 electrostatic bond; see hydrogen bond(ing) element of group, 53 elevated temperatures, 160-1 see also temperature end groups as defect structures, 8 concentration by spectroscopy, 251-6 effects on modes and activity, 156-7 in various polymers, 224, 251-6, 269 see also methyl group(s) energy kinetic, 42, 140-3 levels, 17-18, 112-13, 159-60 of isomeric forms, 10, 13, 239, 278 potential, 42-5, 52-3, 140, 142-4, 158-9 distribution (PED), 144, 175-9 thermal, 18 enthalpy and protein stability, 262 entropy and protein stability, 262 environment and group frequency, 203 environmental health, 219 epoxy resins, 255-6 equations of motion, 45, 58, 74, 138, 140, 152 equilibrium, thermal, 113 esterification reaction, 5-6 esters, saturated, 208-10; see also polyesters in index of polymers ethanol, 257 ether bridges in phenolic resins, 7, 210 ether linkage in polyurethanes, 214 ethers, cyclic, polymerization of, 4
317 ethylene, vibrational modes, 77-8, 131-2 ethylene dibenzoate, 240 ethylene glycol, 6, 128 excited states, 113 external modes; see mode(s), external extinction coefficient, definition, 215 extrusion, 223 F matrix, 139-48 factor group; see: group, factor; mode(s), factor group analysis for various polymers, 81-93, 100-6, 163-203, 235-8, 271-3 far infrared region, definition, 17 Fellgett advantage in FTIR, 26 Fermi resonance, 161, 167-8, 267-9 fibres and Raman spectroscopy, 37 field crystal, 53, 98-9, 100, 105-6 electric, 295-6 and induced dipoles, 107-12 internal, 298 films as samples for IR, 28-30 'finger print', spectrum as, 162 finite chains, vibrations of, 156-8 fitting of spectra; see curve fitting 'fluorescence' and Raman spectra, 32, 37, 174 (fig 5.3) fluorine atom, radius, 8 folding of chains, 13-14, 280-6, 288, 291 force constant(s), 44-5, 76-8, 134, 135, 138-50, 152-4 derived from small molecules, 148-9, 154, 167, 171-2, 173-4, 184-5, 200 for non-bonded atom interactions, 148 interaction force constants, 148 no. of independent, 146, 148 refinement of, 149-50 transferabilty of, 148-9 force field, 146-50 central and general central, 147 for hydrocarbons, 171, 200 for monosubstituted alkyl benzenes, 200 intermolecular, 264, 275 Urey-Bradley-Shimanouchi, 148, 154 valence/simple valence, 148-9, 171, 184-5 forces interatomic, 42-5, 52, 76-8, 80, 98-9, 136, 148, 267; see also force field interlamellar, 290 intermolecular, 80, 98-9, 265, 267, 275; see also splitting, correlation van der Waals, 8, 148, 262 formic acid as solvent for Nylon 6, 221
318
Main index
Fourier series, 159 transform, 26 transform IR, 24-7, 255 free energy, 10, 13, 239 freedom, degrees of, 44 frequency angular, 62-3, 138, 152-4 of electromagnetic wave, 17 of modulation in FTIR spectroscopy, 26, 28 range for normal modes, 17 frequency-phase-difference curves; see dispersion curves frequency/structure correlations, 203 see also mode(s), group fringes, interference, 201 (fig 5.11) FTIR; see Fourier transform IR fundamental frequencies of unit cell, 170 fundamental frequencies/modes for various polymers, 163-203 fundamental frequency, 136, 158-61 fundamental LAM, 286-8 g (gerade) species defined, 66 G matrix, 139-46 gamma-irradiation, 254, 269, 277 gauche; see isomer/isomerism Gaussian line shape, 39, 261 generalized coordinates, 139-41 generalized position, 58, 67-76 generating operations, 55 genuine normal mode, 44,74,77,84-5,146 geometry, molecular, 58, 134, 135 Gibbs free energy, 10, 13, 239 glasses, polymeric, 13 glide plane, 81, 100, 152 axial, 81, 104 diagonal, 104 glycols, 5-6, 128 ground state, 113, 159-61 group axial point, 54, 65 crystal space, 97-9, 100, 103, 166, 272, 276 direct product, 66 factor and local symmetry, 126-31 for helix, 95, 151 of line group, 80, 84, 91, 98-9, 151 of site group, 103, 105 of space group, 97-9 see also factor group frequency; see mode(s), group homomorphous, 56-7
group (continued) isomorphous; see isomorphous groups line, 83, 151, 167; see also factor above mathematical, definition, 53 mode; see mode(s), group; for specific types see index of group modes multiplication table, 55-6, 60, 84 point, 53-6, 84, 85, 89 pseudo-factor, 151 representation; see representation site, 103, 105 space, 97-9, 100, 103, 166, 272, 276 theory, 52, 53-74, 79 vibration; see mode(s), group gutta percha, 10 halo-alkanes, 172, 239 see also chlorides hard segment, 258-62 harmonic approximation, 44-5, 158 harmonic functions, 159 harmonic oscillator, 1, 43-5, 112-13 harmonic vibrations, 136 harmonic wave, 17 head-to-head/tail addition, 8,231-3,250-1 health, environmental, 219 heat treatment, 126 see also annealing and lamellar thickness helical molecules biopolymers, 262-4 CF 3 (CF 2 ) 12 CF 3> 194 symmetry and normal modes, 93-7, 150-2 various polymers, 10, 154, 185-94, 199, 234, 235, 271, 276-7, 279 hexadecane, 269 hexagonal unit cell, POM, 271 hexamethylene diamine, 6 hexamine, in novolak resins, 211 1-hexene, 246 highly symmetric molecules, 206 hindrance, steric, 4, 8 homomorphous groups, 56-7 homopolymer, definition, 2 horizontal mirror plane, definition, 49 hot bands, 160-1 hydrocarbons, force constants, 154, 200 see also paraffin hydrochlorides, 214 hydrogen bond(ing), 222, 256-64 in biopolymers, 262-4 in polyamides, 257-8 in polyurethanes, 258-62 temperature dependence, 257-8, 261 hydrogen chloride, 224 hydrolysis of Nylons, 214
Main index hydroperoxides, 223 hydroxyl group, 222, 256-7, 262 see also index of group modes, OH icosahedral group/symmetry, 55, 66, 125 identification; see analysis identity element of group, 53-5 operation, definition, 49 period, 185-7; see also translational period representation, 57, 65, 117-18 impact strength, 227 improper rotation, definition, 48 inclusions, study by Raman, 32, 36, 254 infinite chain, modes of, 15, 80-97 infrared absorption mechanism, 109 frequency range, 17 selection rules, 107, 109-11 inhomogeneity, study by Raman, 254 initiation site for degradation, 224 initiation step in polymerization, 3 instrumental line shape, 40 intensity ratios in Raman, 32,216,219,268 interaction force constants; see force constants interactions between repeat units, 15, 199, see also coupled oscillators see also forces interatomic distances, non-bonded atoms, 148 interatomic forces; see forces, interatomic interchain interaction; see intermodular forces interference fringes, 201 (fig 5.11) interferogram, 26 interferometer, Michelson, 24-7 intermolecular forces; see forces in various polymers, 163, 239, 274, 277, 281 see also hydrogen bond(ing) internal coordinates; see coordinates internal field, 298 internal mode, 98-9, 131 internal standard, 32, 256, 269, 271 interpretation of polymer spectra, 162 for various polymers, 163-215 intramolecular bonding, 256 invariant, of tensor, 124 inverse of group element or operator, 55 of block matrix, 146 inversion, centre of, definition, 49 irradiation wth X- or y-rays, 254, 269, 277
319 isobutane, 171 isocyanurate ring, 206 isolated regular chain, 80-99, 150-6 examples, 163-203, 234-8, 271-4 isomer/isomerism cis/trans, definition, 10 configurational/conformational, 8-11 and determ. of crystallinity, 278 and group vibrations, 133 characterization of, 234-45, 297 config., examples, 179, 218-19, 275 conform., examples, 179, 200-3, 275, 278, 298-9 trans/gauche, definition, 10; see also conformational above sequence; see head-to-head/tail addition isomorphous groups, 56-7 examples, 84, 89, 91-2, 95, 97, 99, 100, 103, 127, 271, 272; see also section 5.2 isotactic chain, definition, 8-9 isotope substitution, 134-7 see also deuterated polymers iteration, 149-50 Jacquinot advantage in FTIR, 27 ketones, 202, 224 kinetic energy, 42, 140-3 KRS5, 29 LAM, 34, 286-92 / and d sequence distribution, 233 Lagrange's equation of motion, 140 Lambert-Beer law/equation, 215, 217 lamellar crystals, 13, 270, 280-6, 288-92 laser for IR spectroscopy, 21 for Raman spectroscopy, 18, 33-4, 36-7, 124 for resonance Raman, 226 lattice modes, 17, 98-9, 113, 131, 265, 276-7 least squares fit, 171 Legendre polynomials, 294-7, 301 librational modes, 98-9, 277 light, velocity of, 17 light scattering; see scattering line group; see group, line linear molecule, 44, 46, 114-15 linear momentum, 74-5 line shape, 39-41 see also shape of peak
320
Main index
methylol group, 7, 212 Michelson interferometer, 24-7 microconformations and morphology, 13 microscope and Raman spectroscopy, 36 microstructure of copolymers, 227-34 microtome, for preparing IR samples, 30 mirror plane, definition, 47-9 influence on dispersion curves, 154-6 mixed crystals of PED and PEH, 281-6 mixing or coupling of symmetry or group modes in various polymers, 133, 154, 156, 168, 171, 175, 177, 185, 192,194, 197, 200, 202-3, 206, 208, 212-13 macroconformations and morphology, 12 mixtures of polymers; see blends (of polymers) maleic anhydride, 6 mode(s) mass {for specific group modes, see index of reduced, 135, 137 spectrometry, 220 group modes) weighted coordinates, 69, 141 acoustic, 17, 153-4, 157, 161 mathematical techniques, 38-41 longitudinal acoustic, 34, 286-92 matrix effects on vibrations, 229-30 bending, 74^8, 110, 115 matrix (mathematical), 79 breathing, methane, 135 conformationally-sensitive, 200,239-45, character of; see character column, 58-62,67-9,108,139-45,150-1 297 complex, 62-3, 119, 143, 150-1 degenerate, 57-65, 69, 70, 98 accidental degeneracy, 58 diagonal, 61 defn. of symmetry determined, 45, 58 diagonal block, 70, 145-6 eigenvectors for, 46-7, 58-63 of direction cosines, 123 for helical molecules, 96-7 of force constants, 139-50 overtones and combinations of, 160 representation Raman tensors for, 119-20 of group elements, 57 with 1-dim. representations, 62-3 of operator, 57-63 see also splitting, crystal field of tensor, 108 external, 77, 98-9 Wilson G and F, 139^18 factor group mats of single crystals, 291 for helix, 95-7, 151 mechanical properties, 227,234,246,292,301 for model 'polymer' system, 153 melamine and amino resins, 7 of crystal space group, 97-9,100,103, melt crystallization, 13, 280-6, 291 member of group, 53, 55 105 methane, 135, 149 of line group of isolated regular chain, methanol, 223 80-9 methyl acetate, 202 see also main heading 'factor group' methyl group(s), 172, 246 genuine normal, 44, 74, 77, 84-5, 146 group, 15, 131-3, 137, 203-15 as end groups and orientation, 133, 293 in paraffins, 156, 172 interaction in copolymers, 229 in PE, 167, 168, 250, 252 symbols for simple named types, 85-7 in/as branches, 167, 207-8, 246-50, 252 see also mixing methyl laurate, spectrum, 172 internal, 98-9, 131 methyl methacrylate in PMMA, 221-2 lattice, 17, 98-9, 113, 131, 161, 265, methylene 276-7 bridges in phenolic resins, 210 librational, 98-9, 277 ether group in phenolic resins, 210-12 low frequency, 17, 160, 276-7 group, definition, 81 non-factor group, 150-8, 161, 206; see sequences, as branches in PE, 246 also dispersion curves see also index of group frequencies
linkages, amide, ester, urethane, 5, 258 see also: amide group; esters; urethane group liquid cell, 217 liquid nitrogen, 157, 286 liquid paraffinic hydrocabon, 269 local symmetry, 126-31, 194-200, 200-3 lone CH group, 179, 199 see also index of group modes longitudinal acoustic modes, 34, 286-92 Lorentzian line shape, 39-41, 243 low frequency modes, 17, 160, 276-7 low temperature, 276-7
Main index mode(s) (continued) non-genuine, 136; see also mode(s), genuine normal normal, 1, 14-5, 44-7, 52-3, 57-64, 66-78 for helical molecule, 96-7, 150-1 for model 'polymer' system, 152-4 no. of for each symmetry species, 70-4 simple description of, 78 see also: factor group above; eigenvectors; vibrational calculations optical, 1 5 3 ^ , 161 skeletal, 85; see also index ofgroup modes stretching, 74-8, 110-11 symmetry, 74-8, 85-9, 131, 137 for helical molecule, 96-7 mixing of; see mixing torsional, frequency range, 17, 113 true vibrational; see mode(s), genuine normal twisting, 85, 87 zero frequency, 69, 73; see also: mode(s), non-genuine; rotational modes; translation model compounds, 130, 134 for various polymers, 157, 166-72, 185, 189, 194, 202-3, 240-2, 248, 259, 266, 269, 299 model 'polymer' system, vibrations of, 152^ models for molecular orientation, 301 modulation frequency in FTIR, 26, 28 molar absorption coefficient, 215, 248, 259-60 molecular weight, 217, 252, 254, 255 molten polymers, 239, 271, 278, 279 moment of inertia, 98, 136 momentum, angular and linear, 74-5 monochromators, 21-4, 33, 34-6 monoclinic crystal, 276 monomer,3, 12 determination of residual, 219-21 monomeric OH vibration, 257 monosubstituted benzene; see benzene morphology, 12-13 see also lamellar crystals motion, equation(s) of, 45, 58, 74,138,140, 152 mulls for IR spectroscopy, 28-30 multicomponent mixtures, analysis, 215-16 examples, 217-26 multidetector arrays and Raman, 36 multiplex advantage in FTIR, 26 multiplication (combination) in group theory, 53-6, 57 table, 55-6, 60, 84
321 mutual exclusion rule, 113-16, 201 N atom, electronegative, 256 N atom in hydrogen bonding, 256 near infrared, 17, 221 neopentane, 171 neutron scattering, 1, 151, 158, 286, 290-2 Newman projection, 242 (fig 6.6) NH group and hydrogen bonding, 256-64 nitric acid degradation and LAM, 291 nitrogen-containing polymers, 212-15 see also index of polymers nitrogen, liquid, 157, 286 NMR spectroscopy, 211, 212, 217, 218, 219, 229-30, 248, 250, 251, 255 node in LAM, 286 noise, 23, 27, 41 non-crystalline region, 13, 80 between lamellae, 288, 290-1 concentration of; see crystallinity conformational content, 238-45 effect on spectrum, 100, 126, 265 of PE, 168, 247, 269-71 of various polymers, 179, 194, 202 non-factor group modes, 150-8, 161, 206 non-harmonic vibration, 136, 158-61, 221 normal coordinate calculations; see vibrational calculations normal coordinates, 69-70, 75, 111-14, 142-3, 145 normal mode; see mode(s), normal notation for symmetry operations, 47-52, 81-3, 93 species, 65-6 used in tables, 162-3 see also symbols novolak resins, 210-12 nucleus of atom, 42-4 null balance in dispersive IR, 23, 27 number of normal modes for species, 70-4, 84-5 Nylons 5-6, 212-14, 220-1, 257-8 O atom, electronegative, 256 O atom in hydrogen bonding, 256 o-o and o-p linkages in novolaks, 210-11 oleamide in polyethylene, 223 opaque samples, 36 operations, 47-52, 55, 81-3, 91-3, 104 optical modes, 153-4, 161 order of a group, 53,71,98-9 of a symmetry element, 52, 54 order (regularity) chain order and crystallinity, 264-80
322
Main index
order (regularity) (continued) intermediate, 269-71, 279 loss of, 231, 235 peaks characteristic for, 229,235-8, 264, 278-80 orientation averages, 294-301 molecular and assignment, 120-2, 126-30, 276, 292; see also section 5.2 (163-203) for various examples and conformational content, 240, 297 and IR spectroscopy, 28, 120-2, 127 and polymer processing, 16-7 and Raman spectroscopy, 34, 125 and Young's modulus, 290 characterization of, 133, 242, 292-301 corrections for, 240, 269, 279, 296 of axes; see axes, choice of orientation orthogonality of eigenvectors, 69, 141-3 of Legendre polynomials, 294 orthorhombic crystal, 100, 103, 265, 271 oscillator, harmonic; see harmonic oscillator overlapping peaks, 126, 247 see also curve fitting overtone/combination frequencies, 17, 158-61 for PE, 163, 167, 168, 170, 222 for various polymers, 191, 197, 213, 274 for vibrations involving H atoms, 221 oxidation, degree of, 223-4 oxirane ring, 255 oxygen-containing groups, 200-3, 210-12 see also polyesters in index of polymers oxygen, direct determination of, 217 paradisubstituted benzene; see benzene paraffin branched, 171, 247 normal crystal structure, 173 definition, 156, 166 deuterated, 171, 283-5 force constants, 171-2 in liquid state, 269 LAM, 286-90, 292 models for PE, 156-8, 166-72, 266, 269 selection rules and spectra, 168-71 single crystals, 172, 266 parallel (n) mode, definition, 121 partially polarized Raman peaks, 125 peak area, 229-30, 242, 271, 275
peak (continued) broadening; see width of peaks shapes, 3 9 ^ 1 , 229, 233, 241, 289, 291-2 shift; see shift of peak splitting; see splitting width; see width of peaks PED (potential energy distribution), 144, 175-9 1-pentadecene, 252 perfect isolated chain; see isolated regular chain perpendicular (or) mode, definition, 121-2 perspex; see index of polymers phase (of vibration) relative of atoms in normal modes, 45-7 of chains in correlation splitting, 99, 100, 102-3 106, 265-6 in lattice modes, 99 of groups within repeat unit, 103, 154, 174-6 of masses for model 'polymer', 153-4 of methylene groups in normal paraffins, 168-9 of units in helical molecules, 96-7, 151, 188, 192, 194 phase (type of structure) in polyethylene, 264, 269-71, 298 in polytetrafluoroethylene, 187, 276-7 in polyurethanes, 258-62 phenolic antioxidants, 222 phenols and phenolic resins, 7, 210-12 phenyl group; see benzene N-phenylurethane, 259 Phillips process, 246, 252, 254 photochemical decomposition, 223-6 photographic recording for Raman, 167 photon, 112 phthalic anhydride, 6 physical properties/structure, 227 and spectroscopy, 16-17 see also mechanical properties pi (n) mode, definition, 121-2 planar square molecule, 115-16 planar zig-zag, 8-10, 13, 82, 89-92, 100, 127, 173, 177-80, 185, 199, 235, 237, 240 plane, see mirror plane plasticizers, 226 plexiglass; see index of polymers point group and symmetry, 47-56 polar groups, 163; see also hydrogen bond(ing) polarizability in longitudinal acoustic mode, 286 derivative, 113; see also tensor, Raman
Main index polarizability (continued) molecular, 107-9, 111-15, 160, 225 see also tensor polarization and infrared spectra, 28, 67, 109-11, 120-2, 127; see also 266, 276-7 and section 5.2 (163-203) for various examples and orientation, 292-301; see also orientation, molecular and Raman spectroscopy, 33-5, 37, 67, 122-5,127; see also 267 and section 5.2 (163-203) for various examples instrumental, 28, 34, 122, 269 polyene low MW, 226 sequences in PVC, 224-6 polymer, definition of, 2 polymerization and polymerization mechanisms, 3-7, 227-8, 233, 234, 240, 246, 271 position, generalized, 58, 67-76 potassium bromide disks, 29, 30 potential energy; see energy powder samples, 28-30, 36, 37, 275 prediction of frequencies, 149-50, 172 preferred orientation around axis, 294 pressing, hot for IR samples, 29 principal axes of a tensor, 108 principal axis of a molecule, 49, 65 principal polarizabilities, 107-8 prisms in IR spectrometers, 23 processing of polymers, 16, 223, 292 product rule, Teller-Redlich, 136-7 propagation step in polymerization, 3 propylene glycol, 6 propylene oxide, polymerization, 233 proteins, hydrogen bonding in, 262-4 pseudo-baseline technique, 38-9, 215 pyramid, square, molecule, 59 pyrolysis of novolak resin, 211 qualitative analysis; see analysis quantitative analysis; see analysis quantum mechanics, 17-18, 43, 112-13, 159-61 quartic terms in potential energy, 159 Raman; see scattering, Raman random re-entry, 280, 283, 286 randomly coiled chains, 13 Rayleigh scattering, 2, 32, 33, 112, 288 ratio recording IR spectrometer, 23 ratios of absorbances, 217-18, 247, 256, 269 of Raman intensities, 32, 216, 219, 268 reactivity ratios, 228
323 reduced mass, 135, 137 reflection loss, correction for, 20, 298 operation, 47-9 spectroscopy, IR, 20, 30-1 refractive index, 301 regular polymer chain; see chain regularity peaks, 229, 235-8, 264, 278-80 see also order (regularity) relaxation of orientation, 292 repeat unit(s) chemical, 2, 15, 81, 95, 180, 235 dimensions, 15, 180, 187 interactions between 15, 199; see also: coupled oscillators; phase physical, 15, 95, 97, 150-2, 235 translational, 81-3, 90-1, 103, 151-2, 185-7, 238 representation of a group based on 2nd order functions, 117-20 by matrices, 57-63 reducible, irred. & equiv. reps, 60-3 3n-dimensional, 84 3N-dimensional, 67-70 repulsive forces, 148 resole resins, 210-12 resolution, spectral, 18, 21, 27 resonance, Fermi, 161, 167-8, 267-8 resonance Raman, 18, 43, 224-6 rigidity, 234 rotation around bonds, 9-10; see also isomer/isomerism axes/operations, definitions, 47-9, 65, 81,93 rotational isomerism; see isomer/isomerism, conformational rotational modes/motion, 42,44,69, 73-5, 77,84-7,98, 110, 146, 192 rotatory mode (libration), 98-9, 277 rubber, natural, 10 rule of mutual exclusion, 113-16, 201 SHH, SHH and SCH structures, 241-5 sample preparation, 28-30, 36-7 Santonox R, 220-1 scattering neutron, 1, 151, 158, 286, 290-2 of light, 33 by gratings and mirrors, 34 by spherulites, 13 Raman, 2, 18 anti-Stokes, 112-13 ease of sampling, 32, 36, 254, 264 efficiencies, 242 first polymer spectrum, 31
324
Main index
scattering (continued) Raman (continued) for model 'polymer' chain, 154 intensities, 112-13, 120 origin of, 111-13 selection rules, 107, 113-20 Stokes, 2, 18, 112-13 Rayleigh, 2, 32,33, 112,288 resonance Raman, 18, 43, 224-6 scission of chains, 224 screw axis/operation, 81, 93, 150, 154 see also helical molecules sebacic acid, 6, 214 second order functions, 117-20 secondary amide group, 213; see also amide group/linkage secondary chlorides; see chlorides secular equation, 139-40, 152 segregation in mixed crystals, 285 selection rules, 81, 101, 107, 113-20, 129 applications; see section 5.2 (163-203) infrared, 109-11 for overtones and combinations, 160 Raman, 107, 113-20 sequence length, copolymers 227-8, 231-4 shape of peak, 39-41, 229, 233, 241, 289, 291-2 sharpening of peak by cooling, 157, 267-8 by crystallization, 265 shift of peak by isotope substitution, 134-7 in copolymers, 229, 230, 233 with temperature, 277; see also temperature side groups; see branch/branching sigma (ar) mode, definition, 121-2 signal-to-noise ratio, 23, 27, 41 similarity transformation, 62 simple harmonic; see under harmonic single crystal texture PE, 172-3 single crystals, 13, 266, 280-6, 288, 291-2 site group, 103, 105 skeletal modes, 85 see also index of group modes slip agents, 223 small samples and Raman, 32, 254 soft segment, 258-62 solid angle, 294, 297 solubility of polymers, 254, 256 solute/solvent interactions, 229-30 solution crystals from, 13, 280-6, 288, 291-2 of polymer, 29, 221,278 solvents for polymers, 29,217,218,221,223 sound waves; see mode(s), acoustic
sources for IR/Raman spectroscopy, 27, 33 see also laser space group; see group, space spectrometer(s), 18-28, 31-6, 113 spherical harmonics, 118-19, 297 spherical part of tensor, 123 spherulites, 13 splitting of modes/peaks, 16 correlation, 99-100, 265, 279 in PE, 105-6,167,173, 265-70, 281-6 in various polymers, 100-3, 184, 272-6 crystal field, 98-100, 265 due to isomerism, 239-45, 278 obscured by amorphous peaks, 100, 265 of group modes by order, 206 spring, 42, 152, 154 square molecule, planar, 115-17 square pyramidal molecule, 59 stability; see degradation stabilizers, 224, 226 standard, internal, 32, 256, 269, 271 stereoregular polymerization), 8, 233, 234 steric hindrance, 4, 8 stress crack resistance, 227, 246 stress/strain measurements, 290 stretching, 273; see also drawing modes, 46, 74-8, 275; see also: mode(s), stretching; index of group modes structural defect; see defect structure sub-units and assignment, 128-30, 200-3 subspectrum, 126 substitution of isotopes, 134-7 see also deuterated polymers subtraction of spectra; see difference sulphur-containing materials, 220-1 summation frequency, 161 see also overtone/combination frequencies superfolding in PE crystals, 286 superposition of spectra, 126, 228-9 surface coatings, 20, 31, 36 switchboard model of chain folding, 280 symbols and electromagnetic waves, 17-18 for types of group vibrations, 85-7 see also notation symmetric species, 52, 65-6 symmetric stretching modes, 46, 74-8, 275 see also index of group modes symmetry and nature of normal modes, 47, 52-3 and spectral activity, 15, 67, 81, 180-5; see also selection rules coordinates, 74-8, 139, 145-6, 178 descent in, 89-93, 99, 105
Main index symmetry {continued) elements and operations definitions, 47-52, 55, 81-3, 91, 93, 104 local, 126-31, 200-3 lowering of, 98, 172 mode; see mode(s), symmetry species, 53 definitions and symbols, 53, 64-6 no. of normal modes for each, 70-4, 84-5 of dispersion curve, 154-6 translation^, 81, 90, 278 symmetry-related set of atoms, 136 syndiotactic chain, definition, 8-9 tacticity, definition, 8-9 tail-to-tail addition, 8, 231-3, 250-1 Taylor series, 111 Teller-Redlich product rule, 136-7 temperature and intermolecular interactions, 267, 277, 286 and structural transitions, 185-7, 273, 276-7 effect on H-bonding, 257-8, 261, 2 6 2 ^ LAM peak, 290-2 peak position/width, 157, 267-8, 277 high and degradation, 246 excited states, 113, 160-1 of polymerization and tacticity, 240-1 tensile strength, 227 tensor, 161 anisotropy of, 123-4 derivative polarizability, 113; see also Raman below invariant of, 124 molecular polarizability, 107-9, 111-15 160, 225 principal axes and components, 108 Raman, 113, 116-20, 122-5, 225, 297, 299 second rank, 108 spherical part of, 123 spherically symmetric, 124-5 symmetric, 108 terephthalates, 202, 210 see also polyterephthalates in index of polymers terephthaloyl residue or group, 128, 134 terminating group; see end groups termination of polymerization, 3 tetrahedral angle/bonding, 8, 47, 181 tetrahydrofuran, as solvent, 29-30, 217
325 thermal contraction, 267, 277 thermal decomposition/degradation, 223-6, 251, 262-4 thermal equilibrium, 113 thermal excitation, 160-1 thermodynamic equilibrium, 239 throughput advantage in FTIR, 26-7 tilt of chains in lamellar crystals, 290 toluene as solvent for PE, 223 torsional vibrations, 17, 113 totally polarized Raman modes, 125 totally symmetric mode/species, 65, 124-5, 135, 149, 226 trans; see isomer/isomerism trans-in-chain groups, 253-4 transformation, similarity, 62 transition moment, 265, 269; see also dipole structural, 185-7, 273, 276-7 translation/translational modes/motion, 42,44,69,73-5,77,84-6, 109-10, 146, 153, 192 period, SI, 95; see also repeat unit(s) symmetry operation/element, 81,91,278 transmittance, 20, 23, 26 transparency and polymer morphology, 13 transverse isotropy, 293 triad units in copolymers, 229-30 triatomic molecule bent, 67-8, 72-7, 110-11 linear, 46, 114-15 triple helix of collagen, 262-3 true vibrations; see mode(s), genuine normal turbidity, of PTFE, 189 twisting modes, 85, 87 see also index of group modes u (ungerade) species defined, 66 UBS force field, 148, 154 ultraviolet/visible region, 43, 224, 225 uniaxial orientation, 293 unit monomer, 3 repeat; see repeat unit structural, 7, 294 unit cell definition, 97 examples; see crystal structure fundamental modes of, 170; see also mode(s), factor group no. of chains through, 98-9, 264-5; see also splitting, correlation units and electromagnetic waves, 17-18 unsaturated groups, 206, 224, 253-5; see also vinyl and vinylidene end group
326
Main index
'unzipping' in PVC, 224 uracil base, 264 urea and amino resins, 7 urea clathrate, 179, 240, 275 urethane group/linkage 5-6, 214 see also polyurethanes in index of polymers Urey-Bradley-Shimanouchi f.f., 148, 154 valence force field (VFF), 148-9, 171, 184-5 van der Waals forces/radius, 8, 148, 262 vector basis, 58-62 displacement, 58 velocity of light, 17 of mirror in FTIR, 26 vertical axis/mirror plane defined, 49 vibrational calculations, 135-56 and assignment, 126, 137-8, 149 and 'prediction' of frequencies, 149-50, 172 arbitrariness, 149-50 for mixed PED/PEH, 283 for polymers, 150-6 for PE, 154, 166-8, 170-2, 267, 276 for PVC, 173, 177-9, 240, 275 for various polymers, 183-5,193-4, 200, 203, 208 procedure, 149 use of computers, 171 vibrations/vibrational frequencies coupled; see mixing see also mode(s) vinyl chloride, 227 vinyl end group in PE, 224,250,252^, 269 vinyl fluoride, 8 vinyl polymers, 3-4 and head-to-head addition, 8 conformational/configurational isomerism, 8-10, 234-45 examples atactic, 194-200; isotactic, 199-200, 234-6 syndiotactic, 100-3, 173-80, 235-8
vinyl polymers (continued) symmetry of chains, 89-93, 126-7 tacticity, definition, 8-9 see also index of polymers vinylidene end groups, 252-4 vinylidene fluoride, 8 vinylidene polymers, 3, 8 see also index of polymers visible region of spectrum, 18, 43, 224, 225 water, 67 and hydrogen bonding, 256, 262-4 and Raman spectroscopy, 32, 264 in poly (methyl methacrylate), 221-2 wave-function, 43, 160-1 wavelength of electromagnetic wave, 17 typical for IR or Raman, 15, 17, 18 wavenumber of electromagnetic wave, 17 waves, certain vibrational modes as, 63 width of peaks, 15,16,18,157,267,270,271 and mathematical techniques, 38-41 effect on resolution, 100, 265, 267 for LA mode, 289, 291-2 influence of crystallinity on, 271 Wilson G and F matrices, 139^48 X-rays and crystal structure, 173, 179, 185-7, 200-1, 234, 239, 240, 276-7 and crystallinity determination, 269, 279-80 and orientation, 293, 301 and Young's modulus, 288 irradiation of PE with, 254 small angle diffraction of, 288, 290, 292 Young's modulus and LAM, 287-92 and X-ray diffraction, 288 zero-frequency modes, 69, 73 see also: mode(s), non-genuine; rotational modes; translational modes Ziegler-Natta catalysts, 234, 246, 254 zig-zag, planar; see planar zig-zag