You are on page 1of 24

Laser -Vibrational Scattering by Polymers

R. F. SCHAUFELE*
Esso Research and Engineering Co.

I. 11. 1 1 1 . IV.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Scattering Sources ...................... Laser Excitation and Physical Form.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . Polymethylene Chains.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A. Solid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . B. Liquid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . V. Future Progress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

67 69 71 77 77 87 89 90

I. INTRODUCTION
Vibrational characterization of polymers has been utilized extensively as a n approach towards the elucidation of their molecular geometry. Most experiments, however, have involved the absorption of infrared energy, owing to difficulties which, in the past, have plagued workers attempting Raman scattering studies (1). Although other complementary, direct, structural approaches, such as magnetic resonance and X-ray diffraction may be utilized, reliance solely upon the absorption of infrared energy for structural information limits the number and types of vibrational modes observable to those which are infrared active. A number of intrinsic, complementary differences exist between Raman scattering and infrared absorption spectroscopy, arising from basic dissimilarities in the participating physical processes. The absorption of infrared energy by a molecule requires a change in its electric dipole moment to occur during the transition if this process is to take place. Quantum mechanically this requires that the dipole transition moment integral M = $\k, < re > \kl dv # 0, where \ k 2 and \El are the upper and lower vibrational state functions and < re > = p d i p is the dipole moment operator a t any instant of time, i.e.,
* Present address: Owens-Illinois Inc., Red Cedar Research Park, Okemos, Michigan. 67

68
pdip

R . F. SCHAUFELE

PO-I-

(dp/@k)oQk,

Qk

being a normal molecular coordinate

and the subscript referring to values for nuclei in their equilibrium positions. The time-dependent differential term determines whether or not the transition occurs, its probability (intensity) being proportional to MZ. In contrast a vibrational transition via Raman scattering depends upon the dipole moment induced in a molecule,pind~dip upon interaction with the oscillating electric field vector, E, of incident light. In the first order pind-dip = (Y El where the polarizability tensor (Y N (YO ( d ( ~ / d Q k ) ~ Q k , the subscript refers to values when

the nuclei are at their equilibrium positions, and Q k is a normal molecular coordinate. Again the time-dependent differential term determines whether or not the Raman transition moment integral M = $9, <pind-dip> 9 1 dv # 0. Since the operators < p j n d - d i p > and < p d i p > are respectively symmetric and antisymmetric, they connect the zeroth vibrational state to upper states which are also respectively symmetric and antisymmetric. I n molecules with high effective symmetry, such as a center of inversion, this selectivity is such that ACTIVITY
INFRARED ACTIVITY

MOLECULAR SYMMETRY AND VIBRATIONAL

& IW

v , CT

>

2 w
3 0
I

INTERMEDIATE

2
U

5 v ,

a W
CT

SHADED AREAS DENOTE MUTUAL OVERLAP OF RAMAN AND INFRARED ACTIVITY LOWEST

0 g v w 0

Fig. 1.

Moleciilar symmetry and vibrational activit.y.

LASER VIBRATIONAL SCATTERING

69

infrared and Raman transitions are mutually exclusive. Relating this result to hydrocarbon polymers, infrared-active modes are primarily C-H motions, whereas Raman-active modes include the skeletal motions of the C-C backbone. With a progressive decrease of molecular symmetry, however, increasing mutual activity occurs, as shown schematically in Figure 1. Experimentally, Raman scattering occurs as emission sidebands of the incident frequency, having a typical intensity - 1 O V that of the incident light and with a displacement frequency equal to the interval between participating states.
11. SCATTERING SOURCES

Before the advent of the laser, observation of Raman scattering by polymers was hampered, if not prevented, by a combination of factors including low source intensity, interference by other line and continuum source frequencies, and competition by various processes for the incident excitation energy. The latter include absorption by the scattering specimen, with subsequent radiationless dissipation or its emission as fluorescence a t other frequencies, photodecomposition of the molecule into other species, and elastic (Rayleigh or Tyndall) scattering. These disadvantages are particularly pronounced for the low preossure mercury (Toronto) arc lamp, generally utilized for its 4358 A emission line. Although the electronic absorption bands of pure synthetic polymers generally lie above 25,000 cm-', impurities and degradation products may absorb a t -23,000 cm-l (4358 A) or lower, hampering or preventing scattering studies in many instances (1). Although some improvement has been made towards increasing the intensity of mercury arcs by utilizing a geometry in which the luminus flux is localized in a small volume, increasing its directionality and facilitating its focusing to a small spot (2,3), the other aforementioned disadvantages associated with this frequency remain. Studies have been reported (4), which utilized lower excitation frequencies with sources such as cesium, rubidium, and helium, but lacked the intensity and simplicity of mercury arc emission. The laser was conceived in the mid-1950's (5,6) and first reduced to practice with the stimulated emission by ruby (7) (6943 A) which was soon followed by the He-Ne gas laser (8) (6328 A) and has since been extended to hundreds of other lasering frequencies in other systems.

70

R. F. SCHAUFELE

TABLE I Selected Laser Lines Suitable for Raman Excitation Powers (watts)
1 0.8 0.1 0.1 5 5

GQ
Ar+ Kr+ Xe+ He-Ne GaAs YAG: bNd3+
4880, 5145 5208, 5309, 5682, 6471, 6764 5419, 5971, 6271 6328 8400-9000 (depends upon temperature) 10,650

8 Output power refers to the strongest individual line which, for multiple lines, is underlined.

The laser's monochromaticity, intensity, directionality, and linear polarization provide an ideal exciting source for Raman scattering spectroscopy. Also, the large range of laser frequencies now available permit considerable flexibility in the selection of a laser frequency to eliminate the disadvantages of mercury arcs. A partial list of laser frequencies particularly suitable for continuous Raman scattering are given in Table I. These have been chosen on the basis of their intensity, their reliability as practical devices, and also for spanning a frequency range in which the probabilities for absorption, fluorescence, and photodecomposition are reduced, while remaining in the spectral region for which the photocathodes of available photoelectric detectors retain adequate response. Together, S-20 and S-1 photocathodes encompass this range of frequencies. The laser's high degree of directionality permits several orders of magnitude enhancement of the exciting intensity by focusing or reduction of the beam diameter. However, since the intensity of a spontaneously scattered Raman band, IR, is proportional to (G4/A'i;) l o , where 'i; is the exciting source frequency of intensity I 0 and A'i; is the mode frequency, the scattered intensity of a given band is reduced by aofactor of -23 in changing the exciiation frequency from Ar+ (4880 A), -20,500 cm-', to Nd3+ (10,650 A), -9,400 cm-'. Detector response, as well as grating and reflector efficiencies, introduce additional factors to the overall response so that the selection of a particular excitation frequency is a balance of all these considerations.

LASER VIBRATIONAL SCATTERING

71

LASER-EXCITED RAMAN SPECTROMETER

RECORDER

POLAR

LOC K- IN AMPLIFIER

6 3 2 8 i He-Ne LASER

BAND- PA SS FILTER

5-20 PHOTOMULTIPLIER

7
SCATTERING C E L L

GRATING DOUBLE

MONOCHROMATOR

ANALYZING

PRISM

Fig. 2. Schematic diagram of her-excited Raman spectrometer system.

The first successful application of a laser to induce Raman scattering by a polymer was reported by Schaufele (9) for isotactic polyproplene, utilizing a continuous He-Ne laser together with photoelectric detection. An excellent Raman-Rayleigh intensity ratio was observed with little or no fluorescence or photodecomposition, permitting the definitive characterization of even the lowest frequency Raman vibrational modes to within -50 ern-' of the laser frequency. Information of similar high quality has been observed and reported (10) in a survey of polymers having various physical forms. The instrumental arrangement is represented schematically in Figure 2. Specific experimental details have been described elsewhere (11,12).

111. LASER EXCITATION AND PHYSICAL FORM


Several different experimental arrangements are required to maximize the Raman-Rayleigh scattering intensity ratio, depending upon the physical form of the polymer. For a translucent solid this is best accomplished by focusing the laser beam into a conical cavity cut either into a cast piece or a pellet formed by compression of

LASER EXCITATION ARRANGEMENTS FOR VARIOUS PHYSICAL STATES

+
LASER BEAM

LASER BEAM

LENS
MIRROR

1
L ASER BEAM

LASER BEAM

P
q

LENS

LIQUID. OBSERVATION
1

OBSERVATION

f
CAPILLARY TUBE

MULTIPLEPASS CELL

WEDGED MIRROR

. TRANSLUCENT SOLID
b . TRANSPARENT SOLID
c.

I. SOLUTION LIQUID

Fig. 3. Experimental arrangements for laser excitation of specimens in various physical forms.

LASER VIBRATIONAL SCATTERING

73

powder (9), as shown in Figure 3a. The cavity functions as a light trap, producing multiple scattering of the incident photons via the intrinsic reflectivity and transmittance of the specimen to the laser frequency. By suitably adjusting the thickness from the cavity wall to the front surface (generally 1 mm or less) an optimum RamanRayleigh intensity ratio is reached which is superior to either front or back surface illumination. Schrader and Bergmann (13) have since independently verified this theoretically. The resulting Raman bands are depolarized by subsequent Rayleigh scattering. An illustration of a result utilizing this approach is shown in Figure 4 for poly(viny1 chloride). X-ray diffraction studies (14) have shown the unit cell to contain 2 helical chains, each of which has 4 skeletal atoms and 1 turn per repeat period. A normal coordinate analysis

c
N 0

3000

2800

1 4 0 0

1200

loo0

800

600

400

200

A'v ( cm-'1

Fig. 4. Laser-excited Raman scattering by translucent, helical

CH3 [CHzCHCI], CH3 (300K).

74

R. F. SCHAUFELE

of the vibrational modes of this sytem has been reported (15), and further analysis is presently in progress (16). When the specimen is a transparent, solid polymer, directionality of the laser beam is preserved in passing through, producing a scattering image which is colinear with the monochromator slit aperature for scattering observed at 90" to the direction of incidence, as shown in Figure 3b. Reduction of the incident laser beam diameter, together with collimation over the active specimen volume, results in either a higher spectroscopic resolution or a higher signal/noise ratio than that occurring with a larger beam diameter. This technique also permits small spatial regions within the scattering specimen to be selectively studied. Alternatively, a second or multiple passes may be produced through the scattering specimen by suitable reflection of the exit beam. The study of polarization characteristics of polymer Raman bands is facilitated by the utilization of laser excitation together with a transparent specimen. This is attributable to the high degree of linear polarization of the former and the low probability of depolarization of Raman light via Rayleigh scattering in the latter. An illustration of results from this approach is shown in Figure 5 for solid, transparent polyisobutylene. Linear band polarizations parallel and perpendicular to the experimental scattering plane (defined by the laser and observation directions) have been separated into their components by means of an analyzing prism. Largest degrees of polarization occur for bands a t -720 cm-' and in the -2900 cm-l C-H stretching region, reflecting their totally symmetric type displacements (factors of 0.56 and 1.6 are necessary to correct polarization intensity ratios, It// II,for instrumental interactions with linearly polarized light in these respective frequency regions). Normal coordinate analysis for this polymer has not, as yet, been reported. Band polarization characteristics of polymers nominally translucent may be studied either in the molten state of the infinite chain, or correlated with bands of the low molecular weight liquids (oligomers). One of the differences between transparent liquids and solids is that the former generally requires a scattering cell to prevent flow during measurement. By a suitable reduction of the laser beam diameter, thereby decreasing the active scattering volume, specimen size can be reduced to 5 p1 or less as, for example by utilizing a capillary cell, as shown in Figure 3c. The liquid state often has a greater number of active degrees of motion than the solid, which may be

LASER VIBRATIONAL SCATTERING

75

indicated by a broadening in the half-widths of Raman scattered bands of liquids, above values for the solid. An illustration of this CH3, an oligoeffect, utilizing a 5 pl specimen of CH30(CHzCHz0)4 mer of helical polyoxyethylene, is given in Figure 6. Although a preliminary study has been reported (17) for scattering by this

!
<

XI

X25

POLARIZATION 1 TO SCATTERING P L A N E

f #
n

POLARIZATION I1 T O SCATTERING PLANE

~lsbo '

l2bo

'

ea,

40

A= lcm-'l '

Fig. 5. Polarized, laser-excited Raman scattering by solid, transparent CHa [CH&(CH,),], CHa (300K).

76

R . F. SCHAUFELE

RAMAN W T l E R I f f i BI c LIOUID 3JO.K

np(whCy

FUARIZATION I I TO StATTERING R A N E

LIWID 300.K POLARIZATION I TO SCATTERING PLANE

W I D 77%

POLARIZATION I TO SCeiTTERlffiR A N E

-1'

3000 2000

1600

'

1200 Ai;(cm-')

'

'

000

'

400

'

Fig. 6. Polarized, laser-excited Raman scattering by liquid and solid CH,O(CH&HzO)rCH, (300 and 77K).

LASER VIBRATIONAL SCATTERING

77

material a t room temperature, the present study provides information regarding polarized scattering by the liquid a t 300K and depolarized scattering by the solid a t 77K. The helix of the infinite chain is known (18) to contain '/* chemical units (CH2CH20) per turn. The 300K polarized study shows that bands a t 279 cm-' and in the 2800-2900 cm-' C-H stretching region have a high degree of linear polarization. At 77K sharpening of bands and the disappearance of diffuse structure occurs for the solid, probably indicating a reduction in number of those rotational isomers present in the liquid. For example, the localized CH2 twisting mode a t 1283 cm-' (300K) undergoes a reduction in half-width from 39 to 12 cm-'. More striking, however, is the apparent disappearance of the broad band a t 279 cm-' (300"K),whose half-width at 300K is 85 cm-', revealing a number of sharp bands for the 77K solid above and below this frequency, with band half-widths of 9-13 cm-' (experimental resolution 8 cm-l). The latter series of bands are skeletal deformation modes for which further analysis is presently in progress (19). The intensity of Raman scattering by polymers in solution is attenuated by a factor proportional t o the solute volume fraction. This reduction can be offset by the higher intensity and directionality of lasers over their mercury arc counterpart. Exploitation of the latter characteristic is illustrated in Figure 3 4 in which the laser beam is repeatedly reflected within a scattering cell to multiply the Raman intensity. The polarized scattering by sodium polyphosphate (Z 'v 200; 5 w t %) in water solution, utilizing this technique, is shown in Figure 7. The sharp, polarized band a t 1156 cm-' is characteristic of a localized vibration, probably a symmetric P-0 stretching mode. Certain of the broader bands arise from more highly delocalized modes, whose half-widths reflect a large number of conformational configurations. This information, which is unattainable via infrared absorption experiments due to solvent absorption, suggests many areas for further development.

IV. POLYMETHYLENE CHAINS

A. Solid
A series of bands in the Raman scattering spectra of short-chain polymethylenes is observed to occur below 600 cm-' (see Fig. 8 ) , their frequency varying inversely with the chain length. These bands

78

R . F. SCHAUFELE

Polarization II to scattering plone

'
1600

_L-i

1 - _ 1 - _ 1

2000

1000

1400

1200

A;

1000 (cm- )

800

600

400

200

Fig. 7. Polarized, laser-excited Raman scattering by HO [Na+P03-],H in water (300K).

were found t o persist in the monoclinic modification, which contains two chains per unit cell, characteristic of pure even-numbered polymethylenes above C26H64 a t room temperature, a s well as for the triclinic form, containing only one chain for even-numbered members below C28H68 (20,21). This latter observation precluded their assignment to intermolecular modes and indicated that they result primarily from intramolecular transitions. Early Raman studies by several workers (22) reported a number of low frequency bands for short chain polymethylenes in the liquid state, i.e., below C18H38. At low temperatures where these materials solidify only one sharp band was reported below 800 cm-l, its frequency being inversely proportional

LASER VIBRATIONAL SCATTERING

79
67 4

2883
8X
>

1132

303

I
2000

3000

I 1400

I
200

1200 1000 AZ (crn)

800

600

400

, Fig. 8. Raman scattering by crystalline C W H ~(300K).

to the chain length. Mizushima and Shimanouchi (22) showed that this band could be simply explained by assignment to the symmetric, longitudinal, accordionlike motion characteristic of an extended, planar zig-zag carbon backbone, as shown schematically in Figure 9. The additional bands of the liquid were attributed to rotational isomers within the chain. Using a semiemperical approach they showed that the low temperature frequencies could be fitted to the linear approximation for the longitudinal fundamental frequency of a linear chain of identical harmonic oscillators
A;
N

(E/~)l~ cd)-l(m/n), (2 for m > >n

(1)

where E is Youngs elastic modulus, p is the density, c is the speed of light, d refers to the distance between chain units, m is the vibration order, and n is the number of chain units. This approximation

80

R. F. S C H A U F E L E

Fig. 9. Longitudinal accordion mode of an ext,ended normal hydrocarbon molecule.

results from the basic relationship derived by Born and von Karman (23) for the finite chain given by
A; = ( E / P ) ' / ~ ( cd ~ r)-l sin (rmln)

(2)

Utilizing E = 3.4 x 10l2dynes/cm2 as a first approximation (22), together with the X-ray density of the polyethylene crystal (24), p = 0.9834 g/cm3, and fully extended intermethylene distance, d = 1.275 x lo-* cm, longitudinal fundamental frequencies (m = 1) were calculated for Eq. (2). Excellent agreement occurs with the lowest frequency polymethylene bands observed here (as),as shown in Table 11. Perdeuteration of C36H74 shifts the observed fundamental longitudinal frequency to values predicted by the ratio of densities, as shown in Table 111. The observed high intensity of this fundamental results from a large change of polarizability for this mode, together with its low Av. This suggested that overtones of this vibration were, perhaps, also observed. Longitudinal, fundamental, and overtone displacements are shown in Figure 10 as a function of chain length. In this instance even values of m give no change of polarizability so that only odd values are predicted (25). Utilizing Eq. (2) as an approximation to verify specific assignments for m, band frequencies were plotted versus m/n, as shown in Figure 11. The resulting function differs somewhat from Eq. (a), as might be expected for the actual planar zig-zag geometry of the polymethylene chain, with methyl end groups. Allowance for these per-

LASER VIBRATIONAL SCATTERING

81

TABLE I1 Longitudinal Accordion Mode Frequencies of Polymethylene Chains Utilizing Eq. (1) and (2) Number of carbon atoms
~~ ~

Order m

Calc Eq. (1) A? (cm-l)


135 122 101 87 76 68 55 26

Calc Eq. (2) A9 (cm-')


134 121 101 87 76 68 55 26

Obs
A; (cm-I)

18 20 24 28 32 36 44 94

132.5 120.0 97.8 84.7 75.8 67.4 56.5 26

m=l
(accordion I

m=2
u)

rn33

v,
0

m=4
0

s
m=5

Fig. 10. Longitudinal displacements of a linear chain.

82

R. F. SCHAUFELE

TABLE 111 Longitudinal Acoustical Frequencies of Polymethylene Chains Utilizing Eq. (3)
n m
A$

Calc (cm-I) CnHzn+z


128.3 358 501 116.0 327 473 97.4 278 423 84.0 241 378 73.8 212 339 65.9 189 306 403 473 54.3 156 254 26.0 75 123 170 216 261 305 345 382 414 443 467 489 510 531 557

A? Obs (crn-')

18

20

24

28

32

36

44

94

1 3 5 1 3 5 1 3 5 1 3 5 1 3 5 1 3 5 7 9 1 3 5 1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31

132.5 355 493 113.6 324 475 97.8 279 422 84.7 241 377 75.8 211 337 67.4 189 303 403 475 56.5 155 259 26 71 121 168 216 265 305 346 386 417 444 467 49 1 512 536 556

LASER VIBRATIONAL SCATTERING

83

TABLE 1 1 1 (continued)
n

36

1 3 5

63 .Oa 177" 283*

62.0 175 281

Frequencies were calculated utilizing Eq. ( l ) , i.e.,


A: (CnH2n+d A i (CnDzn+z)
= (PD/PH)~"

turbations was taken by means of a least-squares fit of the observed frequencies to a (m/n)series expansion of Eq. (2). Best agreement occurred when even powers of (m/n)were also included, as given by Eq. (3).

where

A B C D E F

= = = = =
=

2495 f 86 cm-1 2.855) x lo3 cm-' -(5.867 (6.253 f 3.537) x lo4 cm-' -(3.485 f 2.058) x lo5 cm-' (7.329 f 5.676) x lo5 cm-l -(4.724 f 5.964) x lo5 cm-'

Frequencies calculated from Eq. (3) are compared with the observed values in Table 111. For the infinite chain Eq. (3) reduces to Eq. (1). From the value of A in Eq. (3) the elastic modulus of the fully extended polyethylene chain was calculated, showing that E = (3.58 f 0.25) x 10l2 dyn/cm2. This is comparable with that obtained experimentally by X-ray diffraction techniques in the direction of the polyethylene fiber axis (as), although the value obtained from dynamical mechanical measurements is an order of magnitude lower (27). Young's modulus, estimated theoretically from the bond stretching and bending force constants, is also in agreement with this value, if the molecular force field is adequate (28).

84

R. F. SCHAUFELE

I
0.1
m/n

I
0.2

0 . 3

Fig. 11.

Low frequency Raman bands plotted as a function of m/n and their assignments n-m.

LASER VIBRATIONAL SCATTERING

85

It is also instructive to relate the longitudinal frequencies observed for finite chains to those calculated by means of normal coordinate analysis of the infinite chain. Such calculations have been carried out by Tasumi and Shimanouchi (29), and by Tasumi and Krimm (30), for vibrational branch frequencies as a function of 6 , , the phase difference between neighboring CH2 groups, taking into account interactions between internal and lattice vibrations of polyethylene. From the Born-von Karman theory of a linear chain with n identical units
6,
=

7r(m/n)

(4)

Comparing Tasumi and Krimms theoretical values with the longitudinal frequencies reported here (25), excellent agreement occurs between infinite chain theory and the finite chain observables, as shown in Figure 12. For an isolated polymethylene chain, v 5 and v g correspond to the CCC skeletal deformation and the CC torsional vibrational modes respectively. In the polyethylene unit cell, which contains two chains, a and b, these modes undergo admixture. For the infinite chain the longitudinal vibration becomes the longitudinal acoustical branch. It is interesting that the long chain, C94H190, takes the extended planar zig-zag form in crystals, since the 26 cm- accordion vibration would be frequency shifted if the chain were folded. Accordion-type vibrations are also predicted to occur for folded polyethylene. For a clamped-clamped chain, vibrational frequencies and mode symmetries match those for the free-free configuration. Whether the longitudinal mode frequency is characteristic of the folded segment length or of the total chain length, however, depends upon how well the node a t the fold decouples motions between adjacent fold segments. This question has been partially resolved by utilizing a prototype molecule for a folded polymethylene chain, namely the (C&)34 ring. Its crystal and molecular structure have been determined by Newman and Kay (31,32), employing X-ray diffraction. Its triclinic unit cell contains one molecule which consists, approximately, of two parallel, planar zig-zag chains of 15 methylenes each, linked on each end by two closure methylenes. A schematic representation of the actual structure is shown in Figure 13. A series of Raman-scattered longitudinal mode bands were observed for (CH2)34 from whose frequencies its chain segments are observed to move individually, but

86

R. F. SCHAUFELE

40C

300
1

A I

5 2
W 3

U w
IA

200

100

0 0

0.1
m/n

0.2

0.3

Fig. 12. Low frequency Raman bands and theoretical (7) dispersion curves of v6 the CCC bending, and vg, the CC torsional, vibrations for each of the two polymethylene chains, a and b, in the polyethylene unit cell.

LASER VIBRATIONAL SCATTERING

87

GTTTTTTTTTTTTG' G G' T T G G' GTTTTTTTTTTTTG'

Fig. 13. Schematic representation of the molecular structure for the (CHd34 ring.

with a degree of coupling interaction. These results and a study of folded polyethylene will be reported elsewhere. Longitudinal fundamental modes of decoupled polyethylene fold segments, -100 methylene units or less, or overtones, for chains several times longer, may be observed via Raman scattering. However, the mode fundamental of an extended chain or strongly coupled segments, lo5methylene units long, for example, occurs at 10-2cm-1, in the region accessible to Fabry-Perot interferometers and optical heterodyning experiments. A distribution of effective chain lengths for a particular specimen results in a longtidudinal mode distribution function characteristic of each sequence length and its abundance. The utilization of Raman scattering for the study of short or folded chains and the latter approaches for extended infinite chains are also applicable to the longitudinal branches of polymers in helical or coil configurations. B. Liquid

It is interesting to compare the behavior of polymethylene chains in the liquid and solid states by means of the Raman scattered longitudinal vibrational mode. Such a comparison is shown in Figure 14 for CH3(CH2)12CH3, for example, an oligomer of poly-

88

R. F. SCHAUFELE
RAMAN SCATTERING BY C H ~ I C H $ I ~ C H ~

1
X 1/3

LIOUID - 3OO-K POLARIZATION I I TO SCATTERINGPLANE

rm

LlOUlD - X 0 - K
POLARIZATION 1 TO SCATTERING PLANE 10 lD

f
X 1/3
Lu

Ic

jL
rP-

SOLID - 77OK
POCARIZATION I TO SCATTERING

L m u

Ic

-3000

2600
Ai;(cm-')

Fig. 14. Polarized, laser-excited liamari scattering by liquid and solid CH3(CH2)&H, (300 and 77K).

LASER VIBRATIONAL SCATTERING

89

ethylene. At 300"K, this mode occurs a t 219 cm-I with a band half-width of 50 cm-'. For the solid a t 77K this mode frequency has shifted to 159 cm-' with a half-width of 11 cm-' (experimental resolution 9 cm-I). Utilizing Eq. 3, its calculated value for the fully extended chain a t 300K is 164 cm-'. The 60 cm-I upward frequency shift and 5-fold band half-width increase in going from the solid to the liquid indicate an increased distribution of effective chain lengths for the liquid, with values considerably less than the fully extended configuration. Considering the various parts of the chain to be planar zig-zag segments, as a first approximation, utilizing Eq. (3), a 219 cm-' longitudinal mode frequency occurs for a segment of 10 methylene units. The longitudinal band for the shorter (4) methylene segment is predicted to occur in the vicinity of 400 cm-l, where the observed broad structure may also include the first overtone of the 219 cm-' fundamental. This effect is consistent with a type of chain folding in which the segments are vibrationally decoupled by a gauche bond and have a distribution of lengths, the most likely segment pair being -10 and 4 methylene units. For band half-width frequencies of 194 and 244 cm-I this corresponds to -(ll;3) and -(9.5;4.5) methylene segment pairs, respectively. A more detailed study of this effect will be reported elsewhere.

V. FUTURE PROGRESS
Current areas of study in the vibrational analysis of polymers reflect the progress made during recent years. Theoretical studies have begun on the interactions (29,30) between polymer chains and also for specific nonregular structures (16,33). These advances have been facilitated by improved theoretical techniques leading to accurate values for hydrocarbon polymer force fields, the advent of large, high speed computers, and the regenerative interaction between experiment and theory. Future developments will probably include improved force fields for heteronuclear polymers (eventually extending to biopolymers) relating band intensities to specific structures, and improved predictability of physical properties. Structural studies of polymer liquids and solutions will receive increased emphasis. Certain of the observables should be related to statitical configurational distributions for which theoretical studies show considerable progress, (34,35,36).

90

R . F. SCHAUFELE

References
1. 2. 3. 4.

J. R. Nielsen, J . Polymer Sci., 7C, 19 (1964) and references therein. H. Moser and D. Tieler, 2. Angew. Phys., 12, 280 (1960). C. E. Hathaway and J. R. Nielsen, Spectrochim. Acta, 23A, 881 (1967). E. R. Lippincott, F. X. Powell, J. A. Creighton, and D. G. Johnes, Develop.
Appl. Sprectr., 3, 106 (1963) and references therein. J. P. Gordon, H. J. Zieger, and C. H. Townes, Phys. Rev., 95, 282 (1954). A. L. Schawlow and C. H. Townes, Phys. Rev., 112, 1940 (1958). T. H. Maiman, Nature, 187, 493 (1960). A. Javan, W. B. Bennett, Jr., and D. R. Herriott, Phys. Rev. Letters, 6, 106 (1961). R. F. Schaufele, J . Opt. Soc. Am., 57, 105 (1967). R. F. Schaufele, paper presented at the Pittsburgh Conference for Applied Spectroscopy, March 5-10, 1967 and included at the Polymer Gordon Research Conference, July 3-7, 1967. R. F. Schaufele and M. J. Weber, Phys. Rev., 152, 705 (1966). R. F. Schaufele, Trans. N . Y . Acad. Sci., 30, 69 (1967). B. Schrader and G. Bergmann, Z. Anal. Chem., 2, 230 (1967). G. Natta and P. Corradini, J . Polymer Sc., 20, 251 (1956). T. Schimanouchi and M. Tasumi, Bull. Chem. SOC. Japan, 34,359 (1961). M. Tasumi, private communication. K. Machida and T. Miyazawa, Spectrochim. Acta., 20. 1865 (1964). H. Tadokoro, Y. Chatani, M. Kobayaski, T. Yoshihara, and S. Murahashi, Rept. Prog. Phys., 6, 303 (1063). T. Miyazawa, private communication. A. E. Smith, J . Chem. Phys., 21, 2229 (1953). H. M. M. S. Shearer and V. Vand, A d a . Cryst., 9, 379 (1956). San-Ichiro Mizushima and Takehiko Shimanouchi, J . Am. Chem. SOC., 71, 1320 (1949) and references therein. M. Born and T. von Karman, 2.Physik., 13, 297 (1912). C. W. Bunn, Trans. Faraday Soc., 35,482 (1939). R. F. Schaufele and T. Shimanouchi, J . Chem. Phys., 47, 3605 (1967). I. Sakurada, T. Ito, and K. Nakamae, J . Polymer Sci., C15, 75 (1966). J. M. Crissman, J. A. Sauer, and A. E. Woodward, J . Polymer Sci., A2,5075 (1964). T. Schimanouchi, RI. Asahina, and S. Enomoto, J . Polymer Sci., 59, 93 (1962). M. Tasumi and T. Shimanouchi, J . Chem. Phys., 43, 1245 (1965). M. Tasumi and S. Krimm, J . Chem. Phys., 46, 755 (1967). B. A. Newman and H. F. Kay, J . Appl. Phys., 38, 4105 (1967). H. F. Kay and B. A. Newman, Acta. Cryst., in press. R. G. Snyder, J . Chem. Phys., 47, 1316 (1967). A. Abe, R. L. Jernigan, and P. J. Flory, J . Am. Chem. SOC.,88, 631 (1966). R. G. Gordon, J . Chem. Phys., 42, 3658 (1965). R. N. Work and S. Fujita, J . Chem. Phys., 45, 3779 (1966).

5. 6. 7. 8. 9. 10.

11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36.

You might also like