LIGHT SCATTERING SPECTROSCOPIC STUDlES OF SILICONE FLUIDS Dissertation for the Degree of Ph. D. MICHIGAN STATE UNIVERSITY KATHERINE EMILY REED 1977 This is to certify that the thesis entitled Light Scattering Spectroscopic Studies Of Silicone Fluids presented by Katherine Emily Reed has been accepted towards fulfillment of the requirements for Ph . D . degree in Chemistry Major professorK/ Date W7 0—7639 ABSTRACT LIGHT SCATTERING SPECTROSCOPIC STUDIES OF SILICONE FLUIDS BY Katherine Emily Reed A Brillouin light scattering spectroscopic study of both linear and cyclic polydimethylsiloxane liquids was undertaken to examine the effect of polymer chain length on various spectral parameters. Eight fluids were investigated at seven temperatures between 20°C and 85°C. The refractive index, Brillouin shift frequency and full Brillouin linewidth at half-height were obtained experimentally. These properties were found to be linearly decreasing functions of temperature for all fluids. The sonic velocity, adiabatic compressibility and sound absorption coeffcient were calculated and also found to vary linearly with temperature for each silicone. At low molecular weights the molecular configuration, linear or cyclic, did not affect the acoustical properties. However, as the chain length or ring size increased the cyclic liquids exhibited larger shift frequencies and sonic velocities and lower compressibilities than the linear liquids. At a single temperature the sonic velocity was found Katherine Emily Reed to increase rapidly to a molecular weight of approximately 1000 where it continued to rise very slowly with a further increase of molecular weight. Although experimental data supported Rao's Rule, a simpler empirical equation, relating the velocity linearly to reciprocal molecular weight, was found to adequately represent the data. In addition, the Brillouin shift frequency, full Brillouin linewidth and sound absorption coefficient varied with molecular weight in the same manner. Landau-Placzek ratios were found to be relatively independent of temperature. Results were in good agreement with heat capacity data for silicones and other liquid polymers. The acoustical parameters have been extended more than an order of magnitude in frequency in this study and are found to be in qualitative agreement with data obtained by ultrasonics. LIGHT SCATTERING SPECTROSCOPIC STUDIES OF SILICONE FLUIDS BY Katherine Emily Reed A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1977 \J TO MY HUSBAND DR. PAUL G. GERTENBACH ii ACKNOWLEDGEMENTS The author wishes to thank Professor Jack B. Kinsinger who served as research preceptor for this work, Professor Richard H. Schwendeman who served as second reader, and the Department of Chemistry which provided financial support. The generous gifts of silicone fluids from the Dow Corning Corporation are gratefully acknowledged. iii Chapter I. II. III. IV. V. TABLE HISTORICAL LIGHT SCATTERING SILICONE FLUIDS THEORY C O O O O O APPARATUS OF CONTENTS EXPERIMENTAL INFORMATION A. The Brillouin Spectrophotometer . . . B. Control and Measurement of Temperature SAMPLE PREPARATION . . . . . . . . . . . . REFRACTIVE INDEX REFRACTIVE INDEX VARIATION OF THE WITH TEMPERATURE VARIATION OF THE WITH TEMPERATURE VARIATION OF THE WITH TEMPERATURE VARIATION OF THE WITH TEMPERATURE VARIATION OF THE WITH TEMPERATURE VARIATION OF THE WITH TEMPERATURE CONCLUSIONS . . LITERATURE CITED . RESULTS AND DISCUSSION BRILLOUIN FREQUENCY SHIFT AND MOLECULAR WEIGHT . . VELOCITY OF SOUND AND MOLECULAR WEIGHT . . ADIABATIC COMPRESSIBILITY AND MOLECULAR WEIGHT . . BRILLOUIN LINEWIDTH AND MOLECULAR WEIGHT . . ABSORPTION COEFFICIENT AND MOLECULAR WEIGHT . . LANDAU-PLACZEK RATIO AND MOLECULAR WEIGHT . . iv Page 26 35 40 40 44 52 71 86 89 104 112 115 118 LIST OF TABLES Table Page 1 Measured Temperatures of the COpper Block and Sample Liquid at Various Controller Settings . 38 2 Measured and Corrected Temperatures Obtained During Refractive Index Measurements of Silicone Fluids . . . . . . .‘. . . . . . . . . 42 3 Refractive Index As a Function of Temperature for MDM Observed at 5890 A and Calculated at 5145 A . 45 4 Refractive Index As a Function of Temperature for MDZM Observed at 5890 A and Calculated at 5145 A . 45 5 Refractive Index As a Function of Temperature for MDSM Observed at 5890 A and Calculated at 5145 A . 46 6 Refractive Index As a Function of Temperature for MD9M Observed at 5890 A and Calculated at 5145 A . 46 7 Refractive Index As a Function of Temperature for MDllM Observed at 5890 A and Calculated at 5145 A 47 8 Refractive Index As a Function of Temperature for MD13M Observed at 5890 A and Calculated at 5145 A 47 9 Refractive Index As a Function of Temperature for D. C. 200 Fluid, Viscosity Grade 12, 500 cts, Observed at 5890 A and Calculated at 5145 A . . . 48 10 Refractive Index As a Function of Temperature for D4 Observed at 5890 A and Calculated at 5145 A . . 48 ll Refractive Index As a Function of Temperatureofor D6 Observed at 5890 A and Calculated at 5145 A . . 49 12 Least—Squares SlOpes, Intercepts at 0°C, and Correlation Coefficients for Refractive Index versus Temperature Plots . . . . . . . . . . . . . 51 13 Brillouin Shift Frequencies of Some Substances at 30°C . . . . . . . . . . . . . . . . 53 V 14 15 16 17 18 19 20 21 22 23 24 25 Observed Brillouin Frequency Shifts, Velocities of Sound, Full Brillouin Linewidths at Half—Height, Sound Absorption Coefficients, and Refractive Indices of MDM as a Function of Temperature . . . Observed Brillouin Frequency Shifts, Velocities of Sound, Full Brillouin Linewidths at Half-Height, Sound Absorption Coefficients, and Refractive Indices of MDZM as a Function of Temperature . . . Observed Brillouin Frequency Shifts, Velocities of Sound, Full Brillouin Linewidths at Half-Height, Sound Absorption Coefficients, and Refractive Indices of MDSM as a Function of Temperature . . . Observed Brillouin Frequency Shifts, Velocities of Sound, Full Brillouin Linewidths at Half-Height, Sound Absorption Coefficients, and Refractive Indices of MD9M as a Function of Temperature . . . Observed Brillouin Frequency Shifts, Velocities of Sound, Full Brillouin Linewidths at Half-Height, Sound Absorption Coefficients, and Refractive Indices of MD11M as a Function of Temperature . . Observed Brillouin Frequency Shifts, Velocities of Sound, Full Brillouin Linewidths at Half-Height, Sound Absorption Coefficients, and Refractive Indices of MD13M as a Function of Temperature . . Observed Brillouin Frequency Shifts, Velocities of Sound, Full Brillouin Linewidths at Half-Height, Sound Absorption Coefficients, and Refractive Indices of D4 as a Function of Temperature . . . . Observed Brillouin Frequency Shifts, Velocities of Sound, Full Brillouin Linewidths at Half-Height, Sound Absorption Coefficients, and Refractive Indices of D6 as a Function of Temperature . . . . Linear Least-Squares Slopes, Intercepts at 0°C, and Correlation Coefficients of the Brillouin Shift vs. Temperature Plots for All Silicones . . Velocity of Sound as a Function of Frequency For Several Silicones at 25°C. . . . . . . . . . . Linear Least-Squares Slopes, Intercepts at 0°C, and Correlation Coefficients of the Velocity of Sound vs. Temperature Plots for All Silicones . . Ultrasonic and Brillouin Velocities of Sound for Linear Polydimethylsiloxanes at 30°C . . . . . vi 56 57 58 59 6O 61 62 63 68 72 77 78 26 27 28 29 30 31 32 33 34 35 36 37 38 39 Brillouin Velocity of Sound Data for Selected Polymers and Other Liquids at 20°C . . . Rao Numbers, Molar Volumes and Related Quantities for All Silicones at 25°C . . . . . . . Velocities, Densities and Adiabatic Compressibilities for Siloxanes at 25°C . . . . . Adiabatic Compressibilities for Some Organic Liquids at 25°C . . . . . . . . . . . . . Adiabatic Compressibility as a Function of Temperature for MDM . . . . . . . . . . . . . . . Adiabatic Compressibility as a Function of Temperature for MDZM . . . . . . . . . . . . . . Adiabatic Compressibility as a Function of Temperature for MDSM O O O O O O O O O O I O O O Adiabatic Compressibility as a Function of Temperature for MD9M . . . . . . . . . . . . . . Linear Least-Squares Slopes, Intercepts at 0°C, and Correlation Coefficients of the Full Brillouin Linewidth vs. Temperature Plots for All Silicones O O O O O O C O I O O C O O O O O 0 Relaxation Times of the Silicone Fluids as a Function of Temperatire . . . . . . . . . . . . Activation Energies for Relaxational Processes in Silicone Fluids . . . . . . . . . . . Linear Least-Squares Slopes, Intercepts at 0°C, and Correlation Coefficients of the Sound Absorption vs. Temperature Plots for All Silicones Absorption Coefficient Ratios for Silicones at 25°C . . . . . . . . . . . . . . . . Landau-Placzek Ratio as a Function of Temperature for the Silicones . . . . . . . . . . vii 79 83 87 86 90 90 91 91 97 101 102 109 112 114 Figure \JO‘WDWN m 10 11 12 13 14 15 16 17 LIST OF FIGURES Block Diagram of the Brillouin Spectrophotometer Three Rayleigh-Brillouin Spectral Orders . . . . Cross-Sectional View of the Thermostatted Jacket Calibration Plot for the Thermostatted Jacket . The Light Scattering Sample Cell . . . . . . . . Calibration Plot for the Thermocouple . . . . . Refractive Index at 5145Atas a Function of Temperature for All Silicone Fluids . . . . . . Brillouin Spectra of MDZM and MD13M at 85°C . . Brillouin Spectra of MDM at 25°C and 85°C . . . Brillouin Shift Frequency as a Function of Temperature for MDM, MDZM and MDSM . . . . . . Brillouin Shift Frequency as a Function of Temperature for MD9M, MDIIM and MD13M . . . . . Brillouin Shift Frequency as a Function of Temperature for D4 and D6 0 I o o o o o o o o o Brillouin Shift Frequency as a Function of Molecular Weight for Linear and Cyclic Silicones at 80°C . . . . . . . . . . . . . . . Brillouin Shift Frequency as a Function of Reciprocal Molecular Weight for Linear Silicones at 80°C . . . . . . . . . . . . . . . Velocity of Sound as a Function of Temperature for MDM, MDZM and MD5M . . . . . . . Velocity of Sound as a Function of Temperature for MDQM, MD11M and MD M . . . . . 13 Velocity of Sound as a Function of Temperature for D4 and D6 . . . . . . . . . . . viii Page 27 34 36 39 41 43 50 54 55 64 65 66 69 70 74 75 76 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 Velocity of Sound as a Function of Molecular Weight for Linear and Cyclic Silicones at 80°C Rao Number as a Function of Molecular Weight for Linear Silicones Velocity of Sound as a Function of Reciprocal Molecular Weight for Linear Silicones at 80°C Adiabatic Compressibility as a Function of Temperature for MDM, MD Full Brillouin Linewidth as a Function 2 M, MD 5 M and Temperature for MDM, MDZM and MDSM . . Full Brillouin Linewidth as a Function M and MD M Temperature for MD Full Brillouin Linewidth as a Function Molecular Weight for Linear and Cyclic Silicones at 80°C Full Brillouin Linewidth as a Function Reciprocal Molecular weight for Linear Silicones at 80°C Temperature Dependence of Relaxation Times for MDM and MD Sound Absorption Coefficient as a Function M and MD M 9 M, MD 13 M Temperature for MDM, MD Sound Absorption Coefficient as a Function M and MD13M Temperature for MD 9 M, MD 6 2 Temperature for D4 and D6 Sound Absorption Coefficient as a Function Molecular Weight for Linear and Cyclic Silicones at 80°C Sound Absorption Coefficient as a Function Reciprocal Molecular Weight for Linear Silicones at 80°C ix 11 Full Brillouin Linewidth as a Function Temperature for D4 and D 11 Sound Absorption Coefficient as a Function 5 13 MD M of 9 80 84 85 92 94 95 96 99 100 103 106 107 108 110 111 CHAPTER I HISTORICAL LIGHT SCATTERING The phenomenon of light scattering has been of interest to laymen and scientists for hundreds of years. Tyndall's success in simulating "blue sky" atmospheric scattering in the laboratory by passing white light through a gold sol and observing blue light perpendicular to the incident beam led to concentrated efforts to devise a workable theory of light scattering (1). He erroneously attempted to explain scattering using the laws of reflection, but it was Lord Rayleigh (2,3,4) who correctly identified it as a diffraction process and developed the foundation of elastic scattering of light for transparent, optically isotropic, non-interacting particles which were small compared to the wavelength of light. Later Smoluchowski (5) suggested that due to thermal motions of molecules in a liquid, the local density fluctuated through a region on the order of the wavelength of light. Using statistical thermodynamic fluctuation theory Einstein (6) derived a mathematical relationship between the total intensity of scattered light and density fluctuations. Mie (7) in 1908 develOped a general electromagnetic theory of light scattering for non-interacting, spherical particles of arbitrary size which could be applied to systems in which the scatterers were large compared to 10’ the wavelength of incident light. The frequency distribution of the scattered light was first examined by Brillouin (8) in 1922. He treated the density fluctuations (also called sound waves, acoustical waves or phonons) which propagated throughout the medium with no directional bias. He maintained that the light scattered by these travelling fluctuations would undergo frequency shifts of equal magnitude analogous to the Doppler shift of classical physics and consequently derived the following equation: Aw = i 2w°(Vs/c)n sin 8/2 [1] where wois the incident frequency, Aw is the Brillouin frequency shift, c is the velocity of light, 8 is the angle of observation, V8 is the velocity of sound in the medium, and n is its refractive index. These Brillouin modes were observed first by Gross (9) in 1930. In Gross' work and all subsequent experiments an anomalous unshifted peak was also observed and was attributed for several years to a technical error such as dust. This central component was explained by Landau and Placzek (10) in 1934. By a fortuitous change of the two independent variables from.density and temperature (used by Einstein) to entropy and pressure, they were able to effect a separation of the total intensity into two contributions: entropy fluctuations at constant pressure and pressure fluctuations at constant entropy. The entropy fluctuations were very slow, non-propagating thermal fluctuations, dissipated by thermal diffusion and, therefore, had lifetimes dependent on the thermal diffusivity of the scattering medium. The unshifted intensity, commonly called the Rayleigh peak, was due to these isobaric entropy fluctuations. The pressure fluctuations, on the other hand, were propagating and gave rise to the Brillouin lines, one on either side of the Rayleigh component. The lifetime of the propagating acoustic wave was determined by the rate at which it was damped by losing energy to other modes of motion and to overcoming the intermolecular frictional forces and was in turn dependent upon the sound absorption coefficient. Debye was responsible for extending the theory of light scattering to dilute solutions of macromolecules (11,12). After his initial suggestion many others made contributions to the thermodynamic theory of Rayleigh light scattering by small Optically isotropic molecules (13,14,15). In the 1960's equivalent but more successful approaches were developed including an extrapolation of Van Hove's (16) theory of neutron scattering to light scattering (l7) and the introduction of the phenomenological method of linearized hydrodynamics (18-24) in order to implicitly consider molecular interactions. The development of intense, monochromatic radiation sources in the form of gas lasers in the early 1960's led to a surge in experimentation in light scattering spectroscopy. In addition, development of high resolution interferometers contributed to the increased accuracy of the experimental data and facilitated the observation of yet another intensity component called the "relaxation mode" or "Mountain line". When it exists, it is evidenced by a low intensity background upon which the Rayleigh and Brillouin peaks are superimposed. The source of this added intensity is theorized to be coupling between the external translational modes of the scatterer and its internal vibrational modes. Energy in the form of acoustical waves in the medium is absorbed as vibrational energy and is later released thermally as the molecules relax back to their ground states. 8 ILICONE FLUIDS Since the silicones became industrially important in the 1940's, a wealth of information pertaining to their chemical and physical properties has been obtained and published. The silicones are organosilicon polymers in which the silicon atoms are bound through oxygen atoms to form chains or rings and have the following general form: RRR I I I -Si-O-Si-O-Si-O- (32510)n III. The usage of the term silicone became common when it was believed that there would be far reaching analogies between the chemical behaviors of carbon and silicon. The repeating unit R2810 appeared to correspond to a ketone RZCO. Later it was found that the 81:0 bond was unstable compared to the =0 bond and instead silicon formed single bonds to oxygen resulting in polymeric compounds. However the term."si1icone" remained firmly implanted. There are many different silicones but our attention in this work is restricted to the polydimethylsiloxane (PDMS) fluids and the following notation is used to designate specific molecules: Cyclic: D Linear: MD M y x CH 3 (‘2H 3 ('IH ('21-! 3 Si-O H C-Si- O-Si O-Si-CH 3 3 I Y I I I CH3 CH3 CH3 CH3 A number of both cyclic and linear molecules have been synthesized and fractionated to high purity from a polymer mixture of low molecular weights. In addition liquids containing a distribution of molecular weights are designated by their viscosities in centistokes. Several authors have compiled the information published in the literature concerning the silicones (25,26,27). The structure of PDMS in absence of solvents or diluents is determined by the Si-O and Si-C bond distances and Si-O-Si and C-Si-C bond angles. It has long been known that the Si-O interatomic distance is considerably'smaller than the sum of the atomic and ionic radii. The relatively large 'valence angle of the siloxane oxygen coupled with the large size of the Si atoms compared to those of C is thought to result in a long Si-CH3 bond distance whose effect is to push the methyl groups so far from the Si-O-Si axis that 'virtually unrestricted rotation about the siloxane bond is ;possible. The small Si-O bond distance and large Si-O-Si lyand angle have been explained using a resonance description of the bond including covalent, polar and double bond contributions (28). The presence of double bond character would explain the short Si-O bond and the large Si-O-Si angle. The partial donation of a lone pair of electrons from oxygen to the unfilled 3d levels of silicon (dn+pn) is thought to be responsible for the double bond character. The Si-C bond distance, on the other hand, has been found to be almost equal to the sum of the covalent radii and is probably basically a covalent single bond. It is believed that the pure liquid molecular conformation is partially helical with siloxane bonds oriented toward the screw axis and methyl group pointed outward (29,30). This stable helix is believed to be due to the internal dipole compensation caused by the polar component of the siloxane bond giving rise to weak dipoles which align in the helix in such a way as to have Opposing dipoles facing one another. Further evidence consists of entropy of dilution of the siloxanes in benzene and heat of mixing data (31). The high values of the entropy and lack of chain dependence on the heat of mixing both indicate that the liquid structure of a linear PDMS is destroyed upon solvation, so is considerably more ordered than many pure liquids. PDMS rings with more than three -D- units are puckered. D2 is an unstable ring due to strain. Isolation of rings containing as many as thirty atoms has been accomplished. The physical behavior of the cyclic molecules is very different from that of their linear counterparts (25,26). Several cyclic compounds are crystalline near room temperature. The temperature coefficient of viscosity is not as low for the cyclic as it is for the linear conformations (32). Also the viscosity of the cyclic molecules increases more rapidly with molecular weight than in the linear series (33). The intermolecular forces in the polydimethylsiloxanes are unusually weak and lead to high compressibilities as well as other anomalous physical properties, including a low temperature coefficient of viscosity, low freezing and pour points, high vapor pressures and many others (34-38). The lack of change of viscosity with temperature has been interpreted in terms of the chemical structure Of the linear molecules (29). Two Opposing effects, the expansion Of the helical coils and the increase in volume, which occur upon heating, are thought to somewhat compensate for one another resulting in very little net change in the mean intermolecular distances. However, this reasoning is speculative only and the physical factors responsible for the unusual behavior of the siloxanes remain unclear. Other investigators (39,40,41) have shown that the methyl groups also demonstrate virtually unhindered rotation and this is hypothesized to cause large space requirements for the methyl groups which in turn cause molecular volumes to be large (32) and intermolecular forces to be small. Ultrasonic and viscoelastic techniques have been applied to the silicones to determine their acoustical properties. Weissler (42,43) was the first to investigate the relationship between the velocity of sound and chain length and reported that at a molecular weight of 1000 the velocity reached an asymptotic value. Other workers have pursued this further (44,45,46,47) and have attempted to correlate their results with relaxation theories. Also Brillouin measurements were made recently on selected linear and cyclic polydimethylsiloxanes in order to elucidate the temperature and molecular weight dependence of their thermodynamic and acoustical properties (48). The intent of this Brillouin scattering study was to examine spectroscopically several linear and cyclic polydimethylsiloxane liquids in order to determine the effects of chain length and molecular configuration on various acoustical properties including sonic velocity and sound absorption coefficient. It was hoped that this information would provide insights into the factors responsible for the weak intermolecular forces and unique liquid characteristics of this series. Further, it was hoped some information regarding the molecular motion responsible for phonon prOpagation and attentuation could be obtained. CHAPTER II THEORY THEORY Light is scattered by a medium because it contains optical inhomogeneities. Inhomogeneities in pure materials are known to be caused by fluctuations in the Optical dielectric constant. Static nonuniformities would not produce a frequency shift in the scattered light. However, time dependent thermal motions of molecules in a dense fluid create a frequency spectrum in the scattered light which is characteristic of the resulting time-dependent fluctuations. Light scattering theory has developed from hydrodynamic, quantum mechanical and statistical mechanical treatments. When an electromagnetic wave strikes a scattering particle, it interacts with the polarizability and induces an oscillating dipole in the direction perpendicular to that of the incident electric vector. The magnitude of the induced dipole moment is directly proportional to the particle polarizability and the magnitude Of the incident electric vector is_given by + "x = axy By“ [2] The elctromagnetic wave, propagating in the z direction with its electric vector in the y direction, induces a dipole in the x direction proportional to the deformability 10 of the particle electron distribution. In general, the polarizability is a second order tensor but only the diagonal elements contribute when scattered light is polarized. The intensity of the scattered light, L), for an isotropic sample is given by the Rayleigh equation (2,3) 16w“|5|2 Ie= IO 2 u (l + cosze) [3] r A where I0 is the incident intensity, r is the distance from the scatterer to the point Of Observation, A is the wavelength of light in the medium, 8 is the scattering angle measured from the direction of propagation Of the incident radiation, and U’is the average isotropic polarizability given by l OI= §(°‘xx+°‘yy+ (122)- [4] A Clausius-Mossotti type equation was used by Bhagavantam (49) to express E'in terms Of the mean fluctuation Of the dielectric constant about its average value _- V<|Ae|> a = ---- [5] 4n where V is the small spherical scattering volume. Substitution from Equation [5] into [3] gives n2V2<(A€ 13> I = I ’1 + cos2 6 8 o r21“ \ 0): [ ] which is the initial equation used by Einstein and by Landau and Placzek in their statistical thermodynamic 11 treatments Of the intensity Of scattered light. For simplicity Equation [6] is usually written for 9 = 90°. Clearly the mean square fluctuation of the dielectric constant is the only unknown and cannot be directly evaluated. Instead it is evaluated in terms of two other linearly independent thermodynamic properties. Einstein (6) expressed the fluctuation Of the dielectric constant as fluctuations in the density and temperature: Be 86 48 = (-) AD + (-) AT [7] 8p T 3T p ' He then assumed that the contribution made by (as/8T)p was negligibly small compared to (as/8p)T. By eliminating the AT term, Einstein developed the following equation for the intensity Of light scattered at 90°. nzv 38 2 2 I = I —- (—) 9 km . [8] 90 O r21“ 8p T T In this equation k is the Boltzmann constant and ST is the isothermal compressibility. The assumption that the change in dielectric constant with temperature at constant density is negligible has been experimentally determined to be valid for some substances but nOt for others (50). The experimental verification of the existence Of a frequency spectrum of scattered light led Landau and Placzek (10) to attempt a separation of the intensity into terms which corresponded to the lines of the Rayleigh-Brillouin triplet. By changing independent variables from density 12 and temperature to entropy and pressure, as as A6 = (--) m + (—) AP, [9] as P 39 S and proceding in the manner of Einstein, they succeeded in making the separation. The fluctuations in entropy and pressure have been shown to be statistically independent of one another (51), so the mean square dielectric constant fluctuation is given by as as <(Aef> = I-—fP + I-—?S<(Apf> [10] as 8P where the angular brackets indicate an ensemble average over the initial states of the system. Each term in [10] has been evaluated by statistical thermodynamics (51) and the results are as as 8T 88 T (SE11, = (5;)P(;;)P = (3;)1, EVE; ’ [111 <(Asf> = kpvcP , [12] as as 8p 38’ (3;)5 = (3;)SISEIS = (3;)5 p85 [13] and ' <(Api> = kT/VBS [14] where CP is the heat capacity at constant pressure and BS is the adiabatic compressibility. By substituting for the appropriate terms into [10] we obtain as kTZ as p28 kT 2 S <(A6) > = (-Y$-*- + (-¥g --- l15] By using Equation [7] and the fact that (ET/3p)S = 13 (aP/aT) T/Cvpz, it is found that p as as 85 SP T (-) = (--)T + (-) (-) -—- - [15] 3p 3 3p 3T 9 3T9 cvp2 Then assuming that (as/8T)p<< (ac/8p)T (53), the same assumption made by Einstein, results in elimination of the second term in [16], so that the final form Of <(Ae)2> can be written as kT’ ‘ae pzeskr <(A€)2> = (-—I; + (-I; . [17] 3T pVCP 8p V Now, substituting this expression into [6] we Obtain an intensity equation containing two terms I =I nzv as M2 90 O IZAH as (3T);-C- + (;-); pszfig] . [18] D P D Rayleigh Brillouin The first term involves fluctuations with temperature at constant pressure and represents a nonpropagating mode. This component of the scattered light was thought to correspond to the unshifted Rayleigh line. The Rayleigh linewidth is proportional to the rate of decay of the fluctuations and broadened somewhat by the damping thermal dissipative processes. The second term involves fluctuations with density at constant temperature or entropy which are propagated through the medium. They can be viewed as successive compressions and rarefactions which propagate or travel in a wavelike manner. This component of the 14 scattered intensity was thought to correspond to the mode which is shifted in frequency by an amount proportional to the velocity of the thermal waves. Since scattering can occur from waves travelling in opposite directions, this intensity component is actually divided between two lines which together are called the Brillouin doublet and are equally shifted in frequency from the Rayleigh line. The lifetimes of these propagating thermal sound waves are determined by the rate at which they are damped by losing energy to other modes of motion and intermolecular frictional forces. The remarkable similarity between the Einstein total intensity equation and the Brillouin intensity term in their equation led Landau and Placzek to propose the intensity ratio which bears their names: Iggggfeln IR+ZIB I (n2V/r21“)[(ae/ap);] pszBT O Igangigfigiaczek 21B Io(w2V/r21“)[(86/8p);] pszBS This intensity ratio is usually rearranged to the following form and is called the Landau-Placzek ratio. H II IA? I I-' II IS“ I H II .< I I-‘ [20] a: H 'm U) 0 < Thus absolute scattering intensities are not necessary to obtain heat capacity ratios of the medium. 15 The statistical thermodynamic method as delineated above provides only a qualitative treatment Of light scattering. It is based on equilibrium thermodynamics and as such its conclusions are not completely accurate. For example, this approach does not predict or take into account any frequency dependence or dispersion of any of the parameters in the equations. Cummins and Gammon (53) have added correction terms for dispersion to [18] and [20] but we must turn to the linearized hydrodynamic theory of irreversible thermodynamics for a detailed development of the time and frequency dependences Of the thermodynamic and hydrodynamic variables. The hydrodynamic theory of fluctuations differs from the thermodynamic approach because it examines the behavior in time of density fluctuations and takes into account the effect of density fluctuations in adjacent volume elements. The function which contains information about the fluctuations is the generalized structure factor, S(£,w), and is the space and time Fourier transform of the density-density correlation function. Here A is the change in the wave vector and m is the shift in the angular frequency of the scattered light. The intensity of the scattered light is directly proportional to S(E,w) as given by + + I(R,w) = Io(a29“N/2nc“R2) sin2¢ S(k,w) [21] ‘where N spherically symmetric molecules of polarizability a scatter light at an angle ¢ measured from.the electric 16 + vector of the incident wave to R, the point of Observation. The angular frequency of the scattered light is 9, the shift in the angular frequency is w, and the change in the wave vector is A which is defined by + + - k = k12 sin (6/2) [22] where 8 is the scattering angle measured from the direction of propagation of the incident light and R1 is the wave vector of the incident light. In order to Obtain the correlation function from which the generalized structure factor is deduced, it is necessary to solve the linearized hydrodynamic equations for the time-dependent, kth Fourier component Of the density. The following develOpment is essentially that of Mountain (18-23) and details of the calculations are omitted in order to more clearly demonstrate the method. The linearized hydrodynamic equations are the continuity equation, the longitudinal part of the Stokes-Navier equation, and the energy transport equation. The continuity equation assumes that matter appearing in a volume element does so by flowing through a boundary of that element and is given by 8p1 -— + podiv V = 0 [23] at where p, is the equilibrium value of the density, pl is the time-dependent density increment, and V is the flow velocity. The Navier-Stokes equation is 17 av c3 C3800 4 Do(-) + (--)gradp1 + grad T1 - (~ns+n ) grad div V = 0 at 'Y Y 3 B [24] and the energy transport equation is 3T CV(Y‘1) 301 2 pocVI—I - —— (-) - Av T1= o [25] at B at where T and T1 are the equilibrium and time-dependent o values of the temperature, 8 is the thermal expansion coefficient, A is the thermal conductivity,n and nB are S the shear and bulk viscosities, y is the heat capacity ratio, and Co is the low frequency limit of the sound velocity. The first step in Mountain's adaption of Van Hove's neutron scattering method is to obtain the Fourier and Laplace transforms of the linearized hydrodynamic equations of motion and next to solve the resulting equations for the Fourier-Laplace transform Of the density, designated as n(k,s) and given by 32+ (a+b)kzs + abk“ + c§(1-1/y)k2 n(k,s) = n(k) 53+ (a+b)kzsz+ (Cfik2+abk“)s + aCfik“/y [26] where a = A/poCV, b = (408/3 + nB)/po and s is a dummy variable used to evaluate the Fourier and Laplace transforms of the density fluctuations. The inverse Laplace transform of n(k,s) is n(k,t), the time-dependent kth Fourier component of the density. In order to obtain the inverse Laplace transform of [26], it is necessary to find the roots to the 18 cubic equation in the denominator, called the dispersion equation: s3 + (a+b)k2s2 + (C3k2+abk“)s' + aCfik“/y = 0 . [27] An approximate solution to [27] consisting of a power series of coefficients is preferred to the exact algebraic one because it is more easily evaluated and interpreted in terms of the physical parameters characterizing the system. The lowest order solutions to the dispersion equation are 4n /3+n 1 l A s=riCok-!5 S B+—(--—---—)k2 [28] p, ‘3va CP 3 = - lkz/pOCP . [29] The real component of the first solution is the halfwidth at half-height of a Brillouin line 4nS/3+nB l A A F k2 = 8 ---- + -(- - -—) k2 [30] DO 90 CV CP and P is the temporal absorption coefficient or lifetime of the fluctuation. It is now possible to determine n(k,t), the inverse Laplace transform of n(k,s) and the lowest order terms are c -c c n(k,t) = n(k)[-L¥exp(-Ak2t/pocp) + Jlexp(-rk“I-.)cos:c,,kt] . c c p p [31] 19 The kth component of the density-density correlation function can now be constructed from n(k,t) as indicated by F(k,t) = n(-k)n(k,t) [32] C -C C = [fE—-yexp(-lk2t/pOCP) + -yexp(-Fk2t)cosCokt] . C C P P [33] It is now possible to write the generalized structure factor, which is directly proportional to the scattered light intensity and related to F(k,t) as follows S(k,w) = 2Re£mdt exp(iwt) F(k,t) [34] Since S(k,w) = S(k)O(k,w) [35] where S(k) = . [36] the function O(k,w), the ordinary structure factor, contains the frequency distribution of the scattered light and is c -c 2Ak2/p c 0(k,w) = P V ° P 2 2 2 cP (1k /poCP) + w CVI: rk2 I‘kz __ + ] [37] cP (Pk2)2+(w+Cok)2 (Pk2)2 +(w-Cok)2 All terms are easily recognized as Lorentzian in shape since -1 Iterm°=(A+Bw2) . The first term corresponds to a density fluctuation relieved by a thermal diffusion process where A/poCV is the thermal diffusivity. This is a nonpropagating 20 mode; therefore, there is no shift from the incident frequency, and information about the Rayleigh line is contained in this term. The second term represents a prOpagating decay in the fluctuation and results in two Brillouin components which are equally shifted from the incident frequency. The frequency spectrum of light scattered by most pure substances cannot be accurately predicted by this analysis. With the exception of liquid metals, most materials have been found to exhibit experimental sound absorptions which are always larger than classically predicted. Two general processes are thought to be responsible for the excess absorption and are dispersive relaxation phenomena. The first is a thermal relaxation in which there is a weak coupling between the internal vibrational and external translational degrees of freedom of the molecule. The equilibrium distribution of energy is disturbed by the periodic compressing and decompressing of volume elements by the travelling sound wave. The exchange of energy between the internal and external modes which occurs while the system is returning to equilibrium may be expressed by a simple rate equation dE(Ti)dt = -[E(Ti) - E(tfl‘/T' [38] where T' is the relaxation time and Ti and T are the tenperatures of the internal and external degrees of freedom. In light scattering theory the relaxation is described by 21 introducing a frequency dependent bulk viscosity = 0 I ° I ”B nB + nB/(l + le ) [39] or by adding an additional hydrodynamic equation for a new state variable, the internal degree of freedom. The second type of relaxation process assumes that the passage of a sound wave actually induces a structural change in the molecule and the relaxation time is determined by the rate at which the molecule returns to its initial structure. Mountain (19) has extended his hydrodynamic treatment of the spectral distribution of scattered light to include a thermal relaxation process which can be described by a single relaxation time. He introduced a bulk viscosity consisting of two terms, one dependent and the other independent of frequency. The modification, therefore, was incorporated into the Navier-Stokes equation which became 3V C2 CfiBpo Do(-) + (-1-) grad 91+ ( ) grad T1 - (4nS/3+nB) grad div V at Y Y t - J nWt-t') grad div V(t')dt = 0 [40] where n3 is the frequency independent part and n' is the Fourier transform of the frequency dependent part. The procedure to obtain the generalized structure factor was identical to that previously outlined so only the significant results are presented here. The most striking difference appears in the ordinary structure factor as an additional 22 intensity mode. [ 2Ak2/pOCP ] kAkZ/poCP)2+w2J 0(k.w) = (1-l/Y) + [ICi-C§)k2-(V2/C§-l)(C3/V2T2+C§k2(l-l/Y)'] C3/V“T2+ Vzk2 [[1-c3/v2(1-1/Y)][v2k2+c§/vzri - (Ci-C§)k2] + C3/V“tz+ Vzk2 PB F x + B [41] §+(w- Vk)2 P;+(w+Vk)2 As in [37] all terms are Lorenzian in character. The first term is identical to the Rayleigh mode obtained in the absence of relaxation. The second term is also a nonpropagating mode of decay and is related to the coupling between the internal and external degrees Of freedom of the molecules. The last term corresponds to the phonon modes or Brillouin lines and contains terms in T, the relaxation time of the new thermal process. Not only the intensity but also the linewidth of a Brillouin line is altered by the presence of relaxation. The linewidth is not a simple additive property as can be seen by comparing the following expression 23 c3 ak2 blkz 2r = ak2 + bokz - (——)(-—-) + (-——-———-)(l-ak21) [42] B v2 y 1+V2k2T2 with equation [30]. In light scattering experiments the unshifted relaxational line is a very broad low intensity line upon which the Rayleigh-Brillouin triplet is superimposed and is recognized by a low intensity between the Rayleigh line and its associated Brillouin lines. The intensity ratio in the case of relaxation is a much more complicated expression than.the classical Landau-Placzek result because the unshifted intensity of the new thermal mode is included when evaluating Ic' and the Brillouin intensity is also affected: I (l-l/Y)(Ci-C§)k2-(V2/C§-l)[(CZ/V212)+C§k2(l-l/Y)] 21 [1-cfi/VZII-1/Y)][v2k2+c§/vztz]- (cfo-cfim2 . [43] Equation [43] reduces to the classical Landau-Placzek value when phonon frequencies are small, that is, Vkr <>l, the result is IC 7 C2-C3 -- = (Y'l) 1 + (‘-')( m ) v-1 c: [44] In actual practice it is difficult to compare the theoretical prediction with experiment because there is usually large error involved in determining the intensity Of the second nonpropagating mode because it is so broad. 24 Mountain (20) has also deduced the ordinary structure factor for a thermal relaxation by introducing an equation- of motion for the thermal property. A comparison of the two approaches gave identical results for a structural relaxation but slightly different ones in the higher order corrections for the case of coupling between the internal and external degrees of freedom. Litovitz and Herzfeld (54) have develOped a classification scheme for liquids based on the behavior of their sound absorption coefficients, a, with temperature. All fluids except liquid metals have been shown to exhibit sound absorption coefficients significantly higher than predicted by the Stokes equation of classical absorption oz 3 - -1 Ac v3p0n+4w )I/P) OLclass [45] 2 3 where the first term represents absorption due to viscosity and the second due to heat conduction. Liquid metals are alone in that they are the only fluids for which heat conduction is an important source of absorption. Liquids are divided into the following classifications: Group I : Normal Liquids - Monatomic and Diatomic Liquids, Liquid Metals da/dT = 0 Group II: Kneser Liquids - Most Organic and Some Inorganic Liquids da/dT > 0 Group III: Associated Liquids - Water, Alcohols da/dT < 0 25 The normal liquids have absorptions only slightly greater than the classical value. Kneser liquids have a positive temperature coefficient and have a ratio a/aclass which generally falls between 3 and 400. It is thought that the excess absorption is due to a thermal type of relaxation. Associated liquids, on the other hand, are characterized by a negative temperature coefficient, have a ratio a/aclass between 3 and unity, and probably owe their excess absorption to a structural relaxation process. CHAPTER III EXPERIMENTAL INFORMATION APPARATUS A. THE BRILLOUIN SPECTROPHOTOMETER The Brillouin spectrOphotometer used for all light scattering measurements was designed and characterized by S. Gaumer (55). However, in order to elucidate the instrumentation necessary to obtain Brillouin spectra and because modifications have been incorporated recently, a block diagram, Figure l, and brief description of the instrument follow. Further details as well as design considerations are available in the literature (55). All Optical components Of the system were mounted on a vibrationally insulated aluminum slab which served as an Optical bench. The light source was a frequency stabilized Spectra Physics argon ion laser. The most intense single mode of the 5145 A line was chosen as the incident frequency. The mode was selected with the aid of a Spectra Physics Optical Spectrum Analyzer and an external etalon. The laser cavity itself contained a solid etalon which, when tuned, scanned the modes Of the laser lines under examination. The analyzer output was displayed on a Tektronix Storage Oscilliscope which allowed the selection of a single mode. Typical laser output powers were in the 500 mw range. The exact polarization of a polarization rotator attached to the laser was calibrated by a piece Of glass with known refractive index mounted at its Brewster angle.(55) 26 27 Hmuofiouonmouuommm cflsoaawnm map mo EMHOMHQ xooam .H wusmwm 30:2... .1 _ 528: «9.52-2 - , . I I m2: 1 7522,32,. . fiwfiwfiflwT 313. -m I , I . ~95:sz - I k , 12$. «32: — 5&3, fl 5&3 _ , #2:? .10.: J x $25.. $26.. , L 3: a a a PI :1 _\1 _ _ I use]? > m: > LI ezwanUvMI \.. _ __. 0:5. £33.. -3325 «.61: u «R is_emmufize 2:25.: a: n..— m.fl so ezEOEmOO zozaaom: «$5...pr ua>u m> u>§> 2255 lo :59; «35... £3.33," a no? If» 259...: 1ru\u "a u02<¢ ..<¢.—Uumm mum“— 3.31 34 35 In order to ensure that the Brillouin lines were associated with the Rayleigh lines nearest them, the initial mirror separation was taken to be very small so that spectral orders were separated by a large distance. Since a plot of vB/frs versus d is linear and passes through the origin, the SIOpe was determined from the results of the preliminary spectrum. Then the mirror separation corresponding to one third of the free spectral range was calculated. All mirror separations used were approximately 1.40 centimeters. The depolarized component Hv of the scattered light was found to be negligible in all fluids. The pV depolarization ratios of other fluids in the series have been reported (48) to be on the order of 10.2 indicating very low HV intensities. B. CONTROL AND MEASUREMENT OF TEMPERATURE A cross-sectional view of the thermostatted jacket used to thermally insulate and heat the scattering cell is shOwn in Figure 3. It was designed primarily by Gaumer (55) and modified as follows. An additional hole was drilled through the epoxy shell into the copper block to accomodate a thermometer. Asbestos tape served to insulate the copper block from a five foot strand of 28 gauge nichrome resistance wire wound around it. Another layer of asbestos tape was wrapped around the wire and secured by silicone tape. A voltage was supplied to the nichrome wire by a YSI Model 72 Temperature Controller. A thermistor inserted into the copper block and coated with Wakefield Thermal Compound provided temperature information 36 Mercury Thermometer Hole for Thermometer \\\\\ \ gi‘éfiix I :ffl Hole for ’////”Thermistor Pflv ~\‘ rd ‘ ______ _. Sample / Copper Cell \\\\\\\ Jacket \ I N ' '——_‘_] Cooling G1Ycerine L“‘F ”’T’COIls ‘\\\\\\ r,,..Heating 180° Wires Viewing -~ Port \ \ U 1 _‘,,..Sample ___j L.— Cell ‘ ‘ ‘ Holder Insulating I Base ‘ Plate 7“~\. E] /J | Aluminum Bar on 1"::EJ\\\F‘Centering Pin Rotating Table TOp Figure 3. Cross-Sectional View of the Thermostatted Jacket 37 to the controller. It was found that the voltage supplied to the wire was excessively high and melted the wire so a Variac was used as a voltage divider. With the controller bandwidth set at 0.1°C and a Variac setting of 20, the temperature of the copper block was found to vary less than 0.1°C. Due to the viewing slot in the thermostatted jacket, the sample was not completely insulated and it was therefore suspected that at higher temperatures the sample temperature would not equilibrate at the same temperature as the c0pper block. For this reason the thermostatted jacket was calibrated in terms of both the block and sample temperature. Two calibrated, accurate etched-stem mercury thermometers with divisions of 0.0l°C were used as temperature probes. A Fisher NBS calibrated thermometer and an Owens-Illinois thermometer were inserted to the proper depth into the glycerine sample and the copper block respectively. The calibration data are given in Table l. Plotting the copper block temperature versus the glycerine temperature yielded the calibration curve shown in Figure 4. The least-squares slope and intercept of the line were used to calculate the actual temperature from the observed block temperature in all Brillouin studies. 38 Table 1. Measured Temperatures of the Copper Block and Sample Liquid at Various Controller Settings Controller Block Liquid Setting Temperature Temperature 24.0°C 24.05°C 24.10°C 26.0 25.90 25.90 28.0 27.70 27.75 30.0 29.50 29.55 32.0 31.55 31.60 34.0 33.65 33.55 36.0 35.65 35.45 38.0 37.55 37.25 40.0 39.50 39.05 50.0 49.50 48.50 60.0 59.50 58.20 70.0 69.60 68.00 80.0 79.55 77.80 90.0 89.55 87.25 39 70 40 20 90 " 80 ‘ 0 0 0 0 30 T 20 (3.) ilfllVUidWil X3018 Figure 4. Calibration Plot for the Thermostatted Jacket LIQUID TEMPERATURE (°c) 40 SAMPLE PREPARATION The following fluids were the generous gift of the Dow Corning Corporation: MDSM, MD9M, MDllM' MD13M, D4 and D6. The remaining fluids, MDM and MDZM were obtained from Ohio Valley Speciality Chemical Company. All samples had been analyzed by the respective companies to be greater than 99.5% pure and were used as obtained. Since dust particles add intensity to the Rayleigh line thereby invalidating the Landau-Placzek ratio, it was necessary to clean the samples scrupulously. Each fluid was forced under pressure through a 0.25 micron pore size Millipore filter into the 1.0 inch 0.D. scattering cell which had been attached to a 15 millimeter Fischer-Porter joint as shown in Figure 5. The filtering process was repeated until the liquid in the cell appeared totally free of dust. The fluid was then degassed by the standard freeze- pump-thaw method and then sealed under vacuum. REFRACTIVE INDEX Refractive index measurements of all silicone fluids were made at five different temperatures with a Bausch and Lomb Abbe 3-L Refractometer. The temperature of the prisms was maintained by a Haake circulating bath temperature control unit. A Doric Thermocouple Indicator Type T DS-350 with digital temperature readout was used to determine the temperature of the liquid between the prisms. The digital voltmeter was calibrated by immersing the thermocouple in an 41 TR 1 v.” 272” F1” I (J Fisher-Porter Joint / Sample Cell A /Viewing / Window 1v.” 7/ Figure 5. The Light Scattering Sample Cell 42 ice bath and five additional constant temperature baths whose temperatures were determined by the Fisher calibrated NBS thermometer. The calibration data are tabulated in Table 2 and plotted in Figure 6. The refractive index data were obtained at 5890 A, the sodium D line. The dispersion tables supplied by Bausch and Lomb allowed calculation of the refractive index at 5145 A, the wavelength of the incident radiation. Table 2. Measured and Corrected Temperatures Obtained During Refractive Index Measurements of Silicone Fluids Measured Corrected -5.70°C 00.00°C 18.1 23.45 37.2 41.70 59.3 62.35 78.5 81.60 94.8 98.20 43 100— THERMOCOUPLE READING (°C) .5 O 1 ° I I I F I 20 40 so no 100 mu: TEMPERATURE (’c) Figure 6. Calibration Plot for the Thermocouple CHAPTER IV RESULTS AND DISCUSSION REFRACT IVE INDEX In order to calculate the velocity of the acoustic wave at various temperatures, it was necessary to determine the refractive index of the scattering fluid in the temperature range of interest for the Brillouin study. All measurements were made at 5890 A, the sodium D line of an Abbe refractometer and since refractive index is a frequency dependent property, dispersion corrections were made to obtain the value at 5145 A, the Brillouin excitation wavelength. Measured and calculated values of the refractive index at 5890 A and 5145 A respectively are presented as functions of temperature in Tables 3 - 11. Details of the correction procedure are given in the instrument manual. All of the silicone fluids exhibited a linear decrease of refractive index with increasing temperature as shown in Figure 7. The least-squares slopes, intercepts and correlation coefficients are listed in Table 12 for the refractive index-temperature data at both wavelengths. The linearity indicates that there are no optical anomalies occuring in this temperature region. It is interesting to note that the molecules of cyclic conformation have refractive indicies significantly higher than their linear counterparts. Also, within the series of linear fluids, the refractive index at a single temperature increased with 44 45 Table 3. Refractive Index As a Function of Temperature for MDM Observed at 5890 A and Calculated at 5145 A Temperature n5890 n5145 34.1°c 1.3790 1 0.0005 1.3824 1 0.0010 44.9 1.3732 1.3762 52.6 1.3689 1.3720 63.2 1.3630 1.3659 72.5 1.3582 1.3614 Table 4. Refractive Index As a Function of Temperature for MD M Observed at 5890 A and Calculated at 5145 A2 Temperature n5890 n5145 34.1°C 1.3836 1 0.0005 1.3870 1 0.0010 44.9 1.3791 1.3823 52.6 1.3735 1.3769 63.2 1.3692 1.3723 72.5 1.3645 1.3677 46 Table 5. Refractive Index As a Function of Temperature for MD M Observed at 5890 A and Calculated at -5 5145 A Temperature n5890 n5145 37.3°C 1.3885 1 0.0005 1.3921 1 0.0010 44.4 1.3841 1.3874 56.3 1.3795 1.3829 64.7 1.3762 1.3795 73.8 1.3712 1.3745 84.7 1.3636 1.3669 Table 6. Refractive Index As a Function of Temperature for MD9M Observed at 5890 A and Calculated at 5145 31 Temperature n5890 n5145 34.0°C 1.3944 0.0005 1.3979 0.0010 43.3 1.3895 1.3925 51.1 1.3858 1.3892 63.2 1.3818 1.3848 72.5 1.3775 1.3808 47 Table 7. Refractive Index As a Function of Temperature for MD M Observed at 5890 A and Calculated at 5145 A11 Temperature 115890 115145 34.0°C 1.3950 1 0.0005 1.3985 1 0.0010 44.9 1.3904 1.3939 51.1 1.3866 1.3900 63.2 1.3829 1.3859 72.5 1.3791 1.3821 Table 8. Refractive Index As a Function of Temperature for MD13M Observed at 5890 A and Calculated at 5145 Temperature ”5890 n5145 34.1°C 1.3690 1 0.0005 1.3995 1 0.0010 44.9 1.3916 1.3952 52.0 1.3877 1.3913 63.2 1.3829 1.3862 72.5 1.3796 1.3827 48 Table 9. Refractive Index As a Function of Temperature for D. C. 200 Fluid, Viscosity Grade 12,509 cts, Observed at 5890 A and Calculated at 5145 A Temperature n5890 n5145 34.0°C 1.4008 1 0.0005 1.4042 1 0.0010 45.7 1.3960 1.3996 52.5 1.3926 1.3962 62.5 1.3881 1.3917 72.5 1.3854 1.3885 Table 10. Refractive Index As a Function of Temperature for D4°Observed at 5890 A and Calculated at 5145 A Temperature n5890 n5145 34.0°C 1.3906 1 0.0005 1.3938 1 0.0010 44.6 1.3854 1.3889 52.0 1.3799 1.3834 63.7 1.3753 1.3784 72.5 1.3706 1.3739 78.1 1.3666 1.3695 49 Table 11. Refractive Index As a Function of Temperature for D Observed at 5890 A and Calculated at 5145 51 Temperature n5890 n5145 34.0°C 1.3965 1 0.0005 1.3998 1 0.0010 44.0 1.3920 1.3951 52.1 1.3876 1.3910 63.2 1.3830 1.3863 72.5 1.3790 1.3822 78.1 1.3750 1.3779 50 3 El MD9M O MDHM D MD'I3M . ll [as \ 0 012,500 cts 09.4 1.3900— A if, u\\ § 0 E .. \. us > 1.3800- : A U < .. a: u. m M 1.3700— . V MDM - ll “V‘EAZIV‘ I MDSM 1.3600— ‘ D4 l W I l f I I x 1 40 50 60 7O 80 TEMPERATURE (°c) Figure 7. Refractive Index at 5145 A as a Function of Temperature for All Silicone Fluids 51 eeemm.o meH6.H meem.e- memmm.o mmee.e 6065.6- 6 Hmemm.o mmae.a eemm.m- mmemm.o emee.a mamm.ms 46 WHO oaeam.o «mae.e mmea.eu emmmm.o eeae.a eeae.eu eem.me Hemmm.e evee.e meee.e- «Nemm.e meHe.H eemm.e- amen: ememm.o emae.a mvm~.e- Hemmm.o emee.a mmao.eu seam: mommm.o mHH6.H heem.eu mmemm.o mmee.a mme~.eu zen: memmm.o moae.e eeme.m- emmmm.o Neoe.e oemm.eu zmez Heemm.o meoe.a ~mmo.m- meemm.e moee.a mane.mu zmez mmmmm.o oaee.a meom.m- ~mmmm.e eemm.e move.m- 2e: nemeoememou Huxu.c eHx unmeoemmmoo Huxo.c eax cowumamuuou ummonmucH m on :oHumHmHHou ummoumuoH mmoam m meam m emmm muoam wusumummEma .m> xmccH m>fluomummm How mucmHOAMMwoo cowumamuuou com .060 um mummoumuqH .mmmoam mmumsqmuummmq .NH wanna 52 increasing molecular weight to a chain length of about thirteen silicon atoms where it then leveled off in an asymptotic manner as indicated by the small spread in refractive index over a large molecular weight range. VARIATION OF THE BRILLOUIN FREQUENCY SHIFT WITH TEMPERATURE AND MOLECULAR WEIGHT When linearly polarized light is scattered by a substance, the scattered light exhibits a frequency distribution about the incident frequency which is character- istic of the time-dependent fluctuations in the medium. Entropy fluctuations and static inhomogeneties are responsible for the scattered component centered around the incident frequency. Light scattered by pressure fluctuations gives rise to three pairs of symmetrically displaced Brillouin lines corresponding to scattering from the longitudinal and shear waves (56). Only the longitudinal doublet is regularly observed in liquids. The magnitude of the Brillouin displacement, VB, is directly proportional to V the velocity SI with which the presSure or sound wave travels through the liquid, and is given by v3 = 1 2vo(n/c)VS sin 0/2 [46] where v0 is the incident frequency, n is the refractive index of the medium, c is the speed of light in_vacuo, and 0 is the angle of observation. Values of the frequency shift are 53 usually a few gigahertz (GHz). The Brillouin shift is measured experimentally as shown in Figures 8 and 9. For a_given liquid the frequency shift changes with liquid molecular weight. The values of the shifts as well as sound velocities and other parameters are listed for various temperatures for each liquid in Tables 14 through 21. Values of the Brillouin shifts at 30°C for the silicones and several other materials are tabulated below. The low frequency shifts of the silicone Table 13. Brillouin Shift Frequencies of Some Substances at 30°C Material vB(GHz) Reference Pyridine 6.3 57 Dimethylsulfoxide 6.0 57 PolyprOpylene Glycol 4.0 58 Polybutadiene 6.0 59 Polymethylmethacrylate 9.5 60 (amorphous) n-Alkanes 5.0 56 Silicones 3.3-3.7 This Work fluids compared to common solvents and polymers are paralleled by low sound velocities as well and will be discussed in the next section. The temperature dependences of the Brillouin shifts for the liquid silicones are illustrated in Figures 10 - 12. Over the temperature range examined each fluid exhibited a linear decrease in shift frequency with increasing temperature. This type of behavior is well established for pure liquids (61-64). The least-squares slopes, intercepts and correlation coefficients are listed for each silicone 54 MD13M MDZM Figure 8. Brillouin Spectra of MDZM and MD13M at 85°C 55 25°C 85°C Figure 9. Brillouin Spectra of MDM at 25°C and 85°C 56 Table 14. Observed Brillouin Frequency Shifts (v ), T(‘C) 24.6 34.4 44.0 53.6 63.2 72.8 82.4 Velocities of Sound (VS), Full Brillou n Linewidths at Half-Height (F ), Sound Absorption Coefficients (a), and Refracgive Indicies (n) of MDM as a Function of Temperature n5145 vB(GHz) VS(m/s) Ih(GHz) 6(cm-1) 1.3875 3.61 946 0.254 1340 1.3821 3.48 916 0.238 1300 1.3768 3.27 864 0.212 1230 1.3715 3.11 825 0.199 1210 1.3662 2.99 796 0.191 1200 1.3610 2.84 759 0.180 1190 1.3557 2.86 719 0.170 1180 Uncertainties in Measured and Calculated Quantities 10.1 10.001 10.03 110 10.005 120 57 Table 15. Observed Brillouin Frequency Shifts (v ), T(°C) 23.6 34.4 44.0 53.6 63.2 72.8 82.4 Velocities of Sound (V ), Full Brillouin Linewidths at Half-Height (PB ), Sound Absorption Coefficients (a), and Refractive Indicies (n) of MDZM as a Function of Temperature n5145 1.3924 1.3870 1.3821 1.3772 1.3723 1.3674 1.3625 vB(GHz) 3.48 3.33 3.19 3.07 2.96 2.84 2.78 VS (Tn/S) 909 873 840 811 785 756 742 FB(GHz) 0.297 0.286 0.270 0.244 0.233 0.228 0.201 a(cm 1630 1640 1610 1500 1480 1500 1350 Uncertainties in Measured and Calculated Quantities 10.1 10.001 10.03 110 10.005 120 _1) 58 Table 16. Observed Brillouin Frequency Shifts (v ), Velocities of Sound (V ), Full Brillou n Linewidths at Half-Height (PB), Sound Absorption Coefficients (a), and Refractive Indicies (n) of MDSM as a Function of Temperature T(°C) n5145 vB(GHz) VS(m/s) PB(GHz) a(cm'1) 24.4 1.3986 3.62 942 0.376 2000 34.4 1.3935 3.46 903 0.345 1910 44.0 1.3887 3.33 872 0.323 1850 53.6 1.3839 3.22 846 0.281 1660 63.2 1.3790 3.15 831 0.265 1590 72.8 1.3742 3.06 810 0.254 1570 82.4 1.3694 2.89 768 0.249 1620 Uncertainties in Measured and Calculated Quantities 10.1 10.001 10.03 110 10.005 120 59 Table 17. Observed Brillouin Frequency Shifts (VB), Velocities of Sound (V ), Full Brillouin Linewidths at Half-Height (P ), Sound Absorption Coefficients (a), and Refracgive Indicies (n) of MD9M as a Function of Temperature T(°C) n5145 ' VB(GHz) VS(m/s) FB(GHz) 6(cm-1) 23.4 1.4017 3.67 952 0.438 2300 34.4 1.3970 3.59 935 0.416 2220 44.0 1.3928 3.46 904 0.406 2240 53.6 1.3887 3.38 885 0.378 2130 63.2 1.3846 3.26 857 0.362 2110 72.8 1.3804 3.21 846 0.334 1970 82.4 1.3763 3.10 819 0.312 1900 Uncertainties in Measured and Calculated Quantities 10.1 10.001 10.03 110 10.005 120 60 Table 18. Observed Brillouin Frequency Shifts (VB), Velocities of Sound (VS), Full Brillouin Linewidths at Half-Height (P ), Sound Absorption Coefficients (a), and Refracgive Indicies (n) of MDllM as a Function of Temperature T(°C) n5145 vB(GHz) VS(m/s) FB(GHz) a .mH ousmam 3.1 $35.12.: om om oh om om ow on om r . . r r L . . L _ . I oo» "IIIIIII. II nXWN l oom 5A one u e: / c. F com T 0mm 75 mos How ousumnomama mo cowuoosm o no ossom mo wuwooam> .oa ousmwm 3.. 5.35.1123 23oz was 2:9,. .2 ON 00 0m O? on ON P . — I» P P b . _ F b I. 0mm 1. 00m SA T 0mm \I w l C. e)§mwrnu')‘.‘V I. nXUAZF II\ 2.52 . son: I T. OMOP I 00:. 76 on can eo How musumnmmfioa mo coauocsm 6 mm ocsom mo muflooHo> .ha musmwm 3.. 5.35.22... cm on on cm on ow on ON FI......_..._IP.. Iomh Ioow sA ) Iommm I c. 125 w on Iomo mwAH a. #009 77 Table 24. Linear Least-Squares Slapes, Intercepts at 0°C and Correlation Coefficients of the Velocity of Sound vs. Temperature Plots for All Silicones Fluid Slope Intercept Correlation (m/sec°C) (m/sec) Coefficient MDM -3.948 1044 0.9980 MDZM -2.909 972 0.9952 MDSM -2.776 1002 0.9922 MD9M -2.286 1008 0.9959 MDllM -2.902 1064 0.9961 MD13M -2.684 1030 0.9994 D4 -2.157 926 0.9606 D -3.565 1069 0.9909 78 Table 25. Ultrasonic and Brillouin Velocities of Sound for Linear Polydimethylsiloxanes at 30°C Liquid MW Ultrasonic V Brillouin V (m/sec) S (m/sec) S MM 162.2 873 1 10 873 1 10 MDM 236.3 901 925 MDZM 310.4 919 884 MD3M 384.5 931 929 MDSM 533.1 942 918 MD7M 720 953 983 MD9M 829 939 MDllM 978 976 MD 3M 1126 966 949 100 cts 6400 985 975 In order to place in perspective the magnitudes of the velocities of sound for the silicones, the Brillouin velocity results for other polymers and a few organic liquids are listed in Table 26. The siloxane sound velocities, ranging from 875 to 975 meters per second, are significantly lower than those tabulated. High sound velocities have been empirically correlated with strong intermolecular forces and low compressibilities which are thought to facilitate passage of the sound wave. For this reason associated liquids such as dimethylsulfoxide have higher velocities than the nonpolar normal alkanes. In this respect the velocity of sound is a qualitative probe of the extent of inter- molecular interaction. This interpretation leads to the supposition that the intermolecular forces in the siloxanes are weak in comparison to many liquids and are weaker than in the alkanes of corresponding chain length. Also, the increase of the sound velocity with molecular weight as shown in Figure 18 may be an effect of the increasing Table 26. Brillouin Velocity of Sound Data for 7 9 Selected Polymers and Other Liquids at 20°C Compound Polymethylmethacrylate (amorphous) Benzene Water Polybutadiene C14H30 Toluene C12H26 Carbon disulfide Polypropylene glycol 425 Acetone Acetic acid Ethyl alcohol n-Hexane Methyl alcohol Carbon tetrachloride Ethyl ether Silicones V S (In/see) 2870 1500 1480 1400 1350 1350 1310 1250 1200 1190 1160 1160 1110 1110 1000 1000 900 - 1000 Reference 60 53 53 56 56 53 56 53 58 53 53 53 53 53 53 53 This Work 80 o.om no mocooeaem oaaoso com Hooded How panama Hmaoomaoz mo cowuocsm o no ocdom mo huwooam> .mH musmflm b.1033 ¢<._:Uw._O<< 00g OONP 000—. Dow 80 00V CON . 1\)/(\w . . . ., c _ Auan 1..ummw _. I 84. 25.6 4 .. sA “(NZ-.— 0 T Omh m} \ / c. m l 00w m I omm I com 81 interaction between siloxane functional groups. In long chain molecules there are a greater number of covalent bonds between the siloxane units than in short chain molecules. The decrease of the velocity of sound with increasing temperature can also be attributed to the increasing randomness of the liquid structure as the liquid is heated. A Rayleigh-Brillouin light scattering study of polyprOpylene glycol 425, a polymer somewhat simdlar in structure to the siloxanes, has been reported by Wang and Huang over a broad range of temperatures (58). They examined the behavior of the velocity of sound with temperature as the polymer underwent transitions from the glass to rubber and finally to the liquid state. The behavior of the siloxanes very closely paralleled that of polypropylene glycol in the liquid region. From 40°C to 90°C the velocity of sound in polyproPylene glycol varied from 1000 to 750 meters per second and the siloxane velocities fell in this same range. Also the temperature coefficients of velocity are virtually identical for the two types of polymer liquids. The velocity of sound was found to depend not only on molecular weight but also on molecular structure. Figure 18 shows the smooth curve describing the relation- ship between the velocity and molecular weight for both linear and cyclic conformations. Since the velocity is calculated from the Brillouin shift frequency, the same 82 functional form is expected for both, as is indeed the case. The sloPe is higher at low molecular weights and the velocities of the cyclic fluids exhibit similar behavior but are somewhat higher than those of the corresponding linear fluids. A recent Brillouin study of normal alkanes, C12 to C36' reported by Champion and Jackson (56) revealed that the sound velocity increased with increasing chain length, although no attempt was made to delineate the mathematical relationship. On the other hand, Wang and Huang (74) report no molecular weight dependence on the velocity for polypropylene glycol ranging from weight average molecular weight 425 to 1000. Although Brillouin and ultrasonic theories do not contain any information pertaining to molecular weight, a semi-empirical approach was developed by Rao. He observed that for many unassociated organic liquids (67) 1 8V 1 3V --<-—S) /-——<—)P [49] vS 3T P v 3T where V is the molar volume. In integral form this equation is R = v v [50] where R is the Rao number. The validity of equation [49] was checked for several silicone fluids and ratios ranging from 2.7 to 2.9 were obtained. Rao predicted that R would be a linear function of molecular weight (67). 83 Table 27 contains the molar volumes and Rao numbers for all the silicones studied. Table 27. Rao Numbers, Molar VOlumes and Related Quantities for All Silicones at 25°C Fluid VS(m/sec) p(g/cc) V(cc/mole) R MM 894 0.7619 212.9 2052 MDM 945 0.8200 288.4 2830 MDZM 899 0.8536 364.0 3513 MD M 945 0.8755 439.3 4312 MD3M 932 0.9110 585.2 5717 MDSM 995 0.9134 745.6 7445 MD7M 950 0.9223 898.7 8837 MDilM 991 0.9305 1051 10,480 MD13M 962 0.9335 1206 11,900 The linear relationship between the Rao number and molecular weight, MW, was verified for the silicones as shown in Figure 19. Therefore, according to Rao, the velocity of sound is given by the following third order equation V S 03[b(1/MW) + a]3 [51] pa[b3(l/MW)3+3ab2(l/MW)2+3a2b(l/MW)+a3] where p is the density and a and b are the sloPe and intercept of the Rao number versus molecular weight plot. However, for the silicone fluids, a completely empirical first order equation -4 vS = (-3.10 x 10 )(1/MW) + 852 [52] also fits the data as demonstrated in Figure 20. The contribution of the second and third order terms in Rao's equation give numbers on the order of 10 and 0.1 m/sec which are within the limits of experimental error for this 84 12.001 10.00— 8.00- 6.00“ Rx10'3 4.00“ 2.004 I I I I I j 200 400 600 800 1000 1200 MOLECULAR WEIGHT Figure 19. Rao Number as a Function of Molecular Weight for Linear Silicones 85 Uoow um mmGOOAHHm Hmmcfiq Mom pnmflmz Hmasomaoz Hmooumflomm mo cofluocom 6 mm endow mo huflooam> .om musmfim o— X .2103; ¢<._3Um._0<‘\— n v a N w —[ _ P _ ovo _ loam Ions loos W‘") ‘4 loom Io; loan 86 study. In order to determine which equation better describes the behavior of these liquids, a much larger data set would be necessary. VARIATION OF THE ADIABATIC COMPRESSIBILITY WITH TEMPERATURE AND MOLECULAR WEIGHT The compressibility is inversely related to the strength of the intermolecular forces or degree of association so it is a probe of liquid structure. It is possible to calculate the adiabatic compressibility of a liquid using the velocity of sound and density from 2 1 v 5 /ps . B [53] Table 28 lists the compressibilities of all siloxanes for which velocities of sound have been determined. The compressibilities of the silicones are unusually high. Values for some other liquids are given in Table 29 for comparison. Table 29. Adiabatic Compressibilities for Some Organic Liquids at 25°C Liquid BS x 1010 (cmz/dyne) Reference Pyridine 0.438 57 Dimethylsulfoxide 0.406 57 Decane 0.935 56 Dodecane 0.847 56 The associated liquids, pyridine and dimethylsulfoxide, are much less compressible than the silicones and even 87 Table 28. Velocities, Densities and Adiabatic Compressibilities for Siloxanes at 25°C Liquid p(g/cm3) VS(m/sec) BS X 1010 (cmZ/dyne) MM 0.7619 895 1.64 MDM 0.8200 945 1.37 MDZM 0.8536 399 1.45 MD3M 0.8755 946 1.28 MDSM 0.9110 933 1.26 MD7M 0.9134 996 1.10 MDgM 0.9233 951 1.20 MDllM 0.9305 992 1.09 MD13M 0.9335 963 1.16 100 cts 0.9579 990 1.06 D4 0.9500 872 1.38 D5 0.9528 933 1.20 D6 0.9621 980 1.08 D9 0.9756 1010 1.00 D 0.9736 1050 0.94 15 88 the normal alkanes are for molecules of corresponding chain length. A possible explanation for these observations lies in the large space requirement of the methyl groups. Neutron scattering and nuclear magnetic resonance studies have demonstrated the high mobility of the methyl group hydrogen atoms which is attributed to rotation about the Si-C and Si-O bonds (39,71,72,73). In other words the molar volumes are large and the intermolecular forces are weak. Also, nearly free rotation about the silicon oxygen bond contributes to the overall mobility of the chain. This freedom of motion is considerably more restricted for cyclic molecules than linear ones as evidenced by the smaller molar volumes (32) and compressibilities of the cyclic liquids. The compressibilities of the cyclic molecules fall more rapidly with increasing molecular weight than do those of the linear molecules. The smallest ring containing three silicon atoms is planar so there is little rotation about the Si-O bond. As the ring increases in size, it puckers and there is increased floppiness or freedom of motion. Rings and corresponding chains containing at least nine siloxane units have equal compressibilities. The compressibility of the linear series initially decreases with increasing molecular weight then reaches a limiting value. The statistical model for polymers considers the chain to be composed of molecular segments whose motions are virtually independent of one another (76). The asymptotic behavior of the compressibility may be related 89 to these segmental motions. The compressibilities of four silicones were calculated for several temperatures and the results are tabulated in Tables 30 - 33. The listed densities have been calculated from the reported temperature coefficients of density for these liquids (26,68). Each fluid exhibited a linear increase in compressibility with increasing temperature as is shown in Figure 21. In general, the compressibility and its temperature coefficient decreased with increasing chain length. VARAIATION OF THE BRILLOUIN LINEWIDTH WITH TEMPERATURE AND MOLECULAR WEIGHT The measured Brillouin linewidths were determined directly from the experimental Brillouin lines. An observed Brillouin line is actually a convolution of the true Brillouin line and the instrumental response function, the central of Rayleigh peak. Immediately prior to each series of Brillouin measurements, the linewidth at half-height of the instrumental response function was determined with the use of a strong Rayleigh scatterer. The central line of a Brillouin spectrum.was considered the instrument response function because both had the same linewidth. Since the instrument function and the observed Brillouin line have been 90 Table 30. Adiabatic Compressibility as a Function of Temperature for MDM Table T(°C) p(g/cm3> BS x101°(cm2/dyne) 24.6 10.1 0.8140 10.001 1.37 $0.05 34.4 0.8039 1.48 44.0 0.7939 1.69 53.6 0.7839 1.87 63.2 0.7739 2.04 72.8 0.7639 2.27 82.4 0.7540 2.56 31. Adiabatic Compressibility as a Function of Temperature for MD M 2 T(°C) p(g/cmg) BS x1010(cm2/dyne) 23.6 10.1 0.8521 10.001 1.42 10.05 34.4 0.8405 1.56 44.0 0.8301 1.71 53.6 0.8197 1.85 63.2 0.8093 2.01 72.8 0.7989 2.19 82.4 0.7885 2.30 91 Table 32. Adiabatic Compressibility as a Function of Temperature for MDSM 3 T(°C) p(g/cm ) BS x1010(cm2/dyne) 24.4 10.1 0.8975 10.001 1.26 $0.05 34.4 0.8873 1.38 44.0 0.8774 1.50 53.6 0.8676 1.61 63.2 0.8578 1.69 72.8 0.8480 1.80 82.4 0.8381 2.02 Table 33. Adiabatic Compressibility as a Function of Temperature for MD9M . 3 M C) p(g/cm ) as x101°(cm2/dyne> 23.4 $0.1 0.9233 30.001 1.19 10.05 34.4 0.9128 1.25 44.0 0.9037 1.35 53.6 0.8945 1.43 63.2 0.8854 1.54 72.8 0.8763 1.59 82.4 0.8671 1.72 92 00 r‘ N San: cam zmoz 2 as .202 How wnoumummawa mo coauocom m mm huflafinfimmmumfiou oflumnmwod .HN wusmflm 3.. #35222 06 cm ow on on ow on _ _ _ low; lot— mm x ’9’ Ice; snow.— (3NAG an") Ioo& IONN Iota 93 successfully approximated by Lorentzian line shapes (48,57), the observed Brillouin linewidth is the sum of the linewidth of the true Brillouin line and the central line: exp = true + 54 PB PB PC [ ] All linewidths reported in Tables 14 through 21 have been corrected for the contribution of the instrumental response function. The behavior of the linewidth with temperature is shown for each silicone fluid in Figure 22 - 24. The linewidth was found to be a linearly decreasing function of temperature. The least-squares slopes, intercepts and correlation coefficients are_given in Table 34. Although there is little information available in the literature for comparison, the same linear decrease of Brillouin linewidth with increasing temperature has been reported for pyridine and dimethylsulfoxide (57). Hydrodynamic theory predicts that the linewidth should be proportional to the viscosity of a simple non-relaxing fluid: 4n /3+n 1(Y-1) r = [-—§-———§ + ] E2 [55] s p pCP where "S and "B are the shear and bulk viscosities, A is the thermal conductivity and :2 is the square of the wave vector. For most liquids the second term in Equation [55] is very small compared to the first. For the silicones the maximum contribution of the thermal conductivity term 94 Ema! can ENDS .zoz How musumummEmB no cowuocsm o no nuow3mcwq cfisoaaflum Adam 3; 2:25.22 ow ON 00 cm 0% on .— — p — r P F _ L — P — bl .zmos. nines. 4 .222 .mn «Haven 8 t 83 T 80.0 ... coed 95 m EMHQE Ucm SHADE .2 a: Mom muoumummame mo cofiuocom m mm nuofl3mcflq aflooaafium Hasm .mm muomflm Au; 5.25.253... om ow o» 00 cm oe om om P: P p _ r — . Pl . _ _ _ . P OKNNAU I oomd III I... a .. ommd M 1 1 9 T oovd H I I owed I oomd 96 m a was we now mnsumnomfioe mo coauocdm o no suoflzoCHq :flsoHafiHm Adam .qm muomflm 3.. 335.22.: om o» 8 cm 9. on 8 b P . P t r . b . P Ll P t _ omwuo room... .u 8 10mm... M 1 1 I . 9 coco H z m 18.3 .5 o .. 8m... 97 Table 34. Linear Least-Squares SlOpes, Intercepts at 0°C, and Correlation Coefficients of the Full Brillouin Linewidth vs. Temperature Plots for All Silicones Fluid Slope Intercept Correlation (GHz/°C) (GHz) Coefficient MDM -0.001444 0.2837 0.9919 MDZM 20.001619 0.3378 0.9883 MDSM -0.002302 0.4222 0.9711 MD9M -0.002144 0.4925 0.9946 MDllM -0.002214 0.4872 0.9840 MD13M ~0.001605 0.4420 0.9396 D4 -0.003432 0.5567 0.9864 D -0.003946 0.6549 0.9889 98 was calculated to be 0.001 GHz, too small to be detected in this study. Since the viscosity is known to decrease slowly with temperature for the silicones (32), the experimental results are in qualitative agreement with theory. The Brillouin linewidth was also found to vary with molecular weight as shown in Figure 25. Since the viscosity increases as the chain length is increased, this is not an unexpected result. Figure 26 demonstrates that the linewidth is a linear function of reciprocal molecular weight. The reciprocal of linewidth has the dimension of time and may be described as a relaxation time or distribution of relaxation times. Although the molecular relaxation process occurring has not been interpreted for polarized Brillouin scattering, it is possible to estimate the activation energy for the process by assuming its temperature dependence is described by the Arrhenius equation l/1rI‘B = A exp(E/kT) [56] where l/flI‘B is the relaxation time, A is a constant, B is the activation energy, k is the Boltzmann constant and T is absolute temperature. Table 35 contains the values of the relaxation times at various temperatures for all the silicones. The activation energy can be calculated from the slope of the line obtained by plotting ln(l/nFB) versus l/T. The linearity of such a plot is established in Figure 99 oovo Uoom um mmcooaaam Uflaoau com Hmmcwq mom pcmflmg Hmaoomaoz mo coauocsm m we nuofl3oswq swooaaflum Hash .mm musmflm 5022, «53.62 comp 006—. com 000 00* CON. \\)/(\1 . F pr . . _ u 83 95.6 4 w ¢o one Hmmcwq How pnmflmz Hmaoomaoz mo oowuocom m we ucm«0flmmmou cowumHOmnm venom .am ousmflm 210.25 5333.52 coco com. 08. com com 8.. com _ . _ _ 2 r _ \/\ I 8.. I 00: 2.6.6 4 I 82 5.5.2: . m I 00:. I OOON I oomN a)’><) ( 111 o.om um mmcooflasm Mmmcfla you unmflmz Hmaoooaoz HMUOHmHomm mo cofluocsm o no ucmfiowmmoou :oflumuomnfi ocsom .Nm madman «op x $.03; «(ADUmaOE\— v n N F F _ _ _ 0:2. [coup 189 (run) )0 Face. looou [Gown 112 Hunter also pointed out that absorptions for the silicones were less than the classical Stokes values and that the deviation was more pronounced at higher frequencies. The Brillouin absorptions are also lower than the Stokes values as indicated by the ratio a/aclass listed in Table 38. Table 38. Absorption Coefficient Ratios for Silicones at 25°C Liquid aIcm'l) a/a . class MDM 1300 0.317 MDZM 1650 0.251 MD M 1950 0.145 MngM 2330 0.0954 MDllM 2250 0.0758 MD13M 2200 0.0586 n 2700 0.263 02 2850 0.108 Hunter conjectures that the Stokes prediction is too large because the viscosity undergoes relaxation. Evidence for his position was provided by the d/a class ratio varying from 2 at very low frequencies to 0.25 at high frequencies, thereby indicating a frequency dependent viscosity. The even smaller value of a/a at the hypersonic Brillouin class frequencies provides further substantiation for his hypothesis. VARIATION OF THE LANDAU-PLACZEK RATIO WITH TEMPERATURE AND MOLECULAR WEIGHT The Landau-Placzek ratio is the ratio of the integrated intensity of the central peak to the sum of the Brillouin intensities and is denoted JV. It was determined for each 113 liquid as a function of temperature and the values are listed in Table 40. This ratio was found to be relatively insensitive to both temperature and molecular weight. The large experimental error associated with the measurement, in conjunction with the very small spread of the data, prohibits more than a qualitative interpretation. For a simple liquid the Landau-Placzek ratio is related to the heat capacity ratio by Jv = cP/cV - 1 [59] The heat capacity ratios for several silicones were determined by Weissler (42) and ranged from 1.31 for MDM to 1.22 for MD11M' These values are in good agreement with the Brillouin results. In addition, Wang and Huang found similar Jv values for liquid polyprOpylene glycol and noted the very slight decrease of J with temperature. v The high Landau-Placzek ratios of MD M and D are not 9 6 consistent with the values for the other liquids in the series. Sample contamination by a small amount of high molecular weight polymer would account for this observation.- The other spectral parameters would be influened only slightly since the principle effect would be to increase the intensity of the Rayleigh line and therefore, Jv 114 Table 39. Landau-Placzek Ratio as a Function of Temperature for the Silicones T(°C) MDM MDZM MDSM MDgM 25 0.30 0.26 0.31 1.11 35 0.28 0.27 0.29 0.93 45 0.26 0.26 0.27 0.93 55 0.25 0.25 0.24 0.86 65 0.24 0.23 0.23 0.75 75 0.23 0.23 0.22 0.57 85 0.22 0.21 0.22 0.49 T(°C) MDll MD13M D4 D6 25 0.27 0.27 0.35 2.2 35 0.27 0.23 0.35 2.2 45 0.25 0.23 0.33 2.1 55 0.24 0.21 0.31 2.0 65 0.23 0.18 0.30 1.9 75 0.24 0.18 0.28 1.9 85 0.22 0.17 0.26 1.8 The relative uncertainty in the Landau-Placzek Ratios is 15%. 115 CONCLUSIONS Brillouin light scattering has been used to study the effect of molecular weight on spectral parameters in a homologous series of polydimethylsiloxane fluids. Each liquid behaved in the expected manner exhibiting linear relationships of refractive index, Brillouin shift, velocity of sound, adiabatic compressibility, Brillouin linewidth, and absorption coefficient with temperature. The high compressibilities and low sound velocities were evidence of the weak intermolecular forces in the silicones. Sound velocities agreed with those determined by ultrasonic methods. The Brillouin shift, velocity of sound, Brillouin linewidth, and absorption coefficient were all found to vary in the same way with molecular weight. Each parameter increased rapidly to a molecular weight of about 1000 but" increased only slightly as the molecular weight increased further. An empirical equation was found which described the given parameter as a linear function of reciprocal molecular weight. It has been pointed out that a chain length of eleven silicon atoms, corresponding to a molecular weight of approximately 1000, is just long enough to form a planar, unstrained ring (68). Since long polymer chains are known to be coiled, these steric considerations suggest that this does not occur below a molecular weight of 1000. 116 In summary: 1. The first comprehensive Brillouin study of a homologous series of silicone fluids has been completed and the results are in agreement with reported ultrasonic data. The acoustic properties of the cyclic fluids were distinctly different from those of the linear fluids. Sonic velocities and absorption coefficients in cyclic liquids were higher than in linear ones. This behavior was interpreted in terms of the more restricted motions of cyclic compared to linear molecules. Adiabatic compressibilities for all the silicones were found to be unusually high, indicative of their weak intermolecular forces. The small ring liquids were less compressible than the corresponding linear ones but this difference disappeared for the higher molecular weight liquids. Estimated activation energies for relaxation processes were found to be numerically the same as calculated barriers to rotation about the siloxane bonds. The temperature coefficient of the sonic absorption coefficient was negative for all fluids, placing them in the category of associated and viscous liquids. The classical prediction for absorption due to viscosity alone was found to be too large for all the silicones in contrast to most other liquids which 117 exhibit excess absorptions. This behavior has been attributed to a frequency-dependent viscosity. The lack of velocity dispersion, absence of a Mountain line and low sonic absorptions indicate that the silicones behave as simple fluids in which there are no thermal or structural relaxations. CHAPTER V LITERATURE CITED (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (ll) (12) (13) (14) (15) (16) (17) (18) (19) (20) (21) (22) J. J. unt‘ :9 b w: a» :z Q ‘ w "0"!) L. 12. LITERATURE CITED Tyndall, Phil. Magi, 31, 384 (1869). Rayleigh, Phil. Mag), 41, 447 (1871). Rayleigh, Phil. Mag., 42, 81 (1881). Rayleigh, Phil. Mag,, 41, 375 (1889). Smoluchowski, Ann. Physik, 25, 205 (1908). Einstein, Ann. Physik, 33, 1275 (1910). Mie, Ann. Phys. Lpz., 25, 771 (1908). Brillouin, Ann. de Physique (Paris), 11, 88 (1922). Gross, Nature, 126, 201 (1930). Landau and G. Placzek, Physik. Zeits. Sowjetunion, 172 (1934). Debye, J. Appl. Phys., 15, 338 (1944). Debye, J. Phys. Chem., 51, 18 (1947). Zimm, J. Chem. Phys., 13, 141 (1945). Fixman, J. Chem. Phys., 22, 2074 (1955). Benoit and W. Stockmayer, J. Phys. Rad., 11, 21 (1956). VanHove, Phys. Rev., 35, 249 (1954). Komarov and I. F. Fisher, Soviet Physics JETP, 1358 (1963). Mountain, J. Research NBS, 69A, 523 (1965). Mountain, J. Research NBS, 70A, 207 (1966). Mountain, J. Research NBS, 72A, 95 (1968). Mountain, Rev. Mod. Phys., 38, 205 (1966). Mountain, J. Acous. Soc. Amer., 42, 516 (1967). 118 (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) (35) (36) (37) (38) (39) (40) (41) (42) (43) (44) (45) 119 . Mountain, J. Chem. Phys., 33, 832 (1966). . Bhatia and E. Tong, Phys. Rev., 173, 231 (1968). Bondi, J. Phys. Colloid Chem., 33, 1355 (1951). . Smith, ”Analysis of Silicones", Wiley, New York, . Y., 1974. 255 3’ (P w W. Noll, "Chemistry and Technology of Silicones", Academic Press, New York, N. Y., 1968. Y. Kurita and M. Kondo, Bull. Chem. Soc. Japan, 31, 160 (1954). W. Fox, P. Taylor and W. Zisman, Ind. Engng. Chem., 33, 1401 (1947). W. Noll, Naturwissenschaften, 33, 505 (1962). . Newing, Trans. Faraday Soc., 33, 613 (1950). . Hurd, J. Amer. Chem. Soc., 33, 364 (1946). M C D. Wilcock, J. Amer. Chem. Soc., 33, 691 (1946). D. Bradley, J. Acoust. Soc. Am., 33, 1565 (1963). J . Boggs and W. Sibbitt, Ind. Eng3 Chem., 31, 289 (1955). R. Sauer and D. Mead, J. Amer. Chem. Soc., 33, 1794 (1946). J. Dale, Pure Appl. Chem., 33, 469 (1971). M. Hunter, J. Hyde, E. Warrick and H. Fletcher, J. Amer. Chem. Soc., 33, 667 (1946). E. Rochow and H. LeClair, J. Inorg. Nucl. Chem., 3, 92 (1955). w. Roth, J. Amer. Chem. Soc., 6_9, 474 (1947). W. Roth and D. Harker, Acta. Crystallogr. (London), ,1. 34 (1948). A. Weissler, J. Amer. Chem. Soc., 13, 93 (1949). A. Weissler, J. Amer. Chem. Soc., 13, 419 (1949). J. Hunter, T. Lewis and C. Montrose, J. Acous. Soc. Am., 33, 481 (1963). - J. Hunter, C. Montrose and J. Shively, J. Acous. Soc. Am., 33, 953 (1964). (46) (47) (48) (49) (50) (51) (52) (53) (54) (55) (56) (57) (58) (59) (60) (61) (62) (63) (64) (65) 120 J. Hunter and P. Derdul, J. Acous. Soc. Am., 33, 1041 (1967). J. Lamb and P. Lindon, J. Acous. Soc. Am., 33, 1032 (1967). A. Kumar, Ph.D. Thesis, Michigan State University, East Lansing, Michigan (1973). S. Bhagavantam, "Scattering of Light and the Raman Effect", Ahdhra University, Waltair, India, 1940. I. Fabelinskii, "Molecular Scattering of Light", Plenum Press, New York, N. Y., 1968. L. Landau and E. Lifshitz, "Statistical Physics", Pergamon Press, New York, N. Y., 1958. D. McQuarrie, "Statistical Thermodynamics", Harper and Row, New York, N. Y., 1973. H. Cummins and R. Gammon, J. Chem. Phys., 33, 2785 (1966). K. Herzfeld and T. Litovitz, "Absorption and Dispersion of Ultrasonic Waves", Academic Press, New York, N. Y., 1959. S. Gaumer, Ph.D. Thesis, Michigan State University, East Lansing, Michigan (1972). J. Lascombe (ed.), "Molecular Motions in Liquids", Reidel, Dordrecht, Holland, 1974. M. Tannahill, Ph.D. Thesis, Michigan State University, East Lansing, Michigan (1973). Y. Huang and C. Wang, J. Chem. Phys., 33, 120 (1975). Y. Huang and C. Wang, J. Chem. Phys., 33, 1869 (1974). E. Friedman, A. Ritger and R. Andrews, J. Appl. Phy§,, 22, 4243 (1969). C. O'Connor and J. Schlupf, J. Acous. Soc. Am., 33, 663 (1966). D. Rank, E. Kiess and U. Fink, J. Opt. Soc. Am., 33, 163, (1966). D. Eastman, A. Hollinger, J. Kenemuth and D. Rank, J. Chem. Phys., 33, 1567 (1969). L. Wong and A. Anderson, J. Opt. Soc. Am., 33, 1112 (1972). G. Benedek and T. Greytak, Proc. IEEE, 33, 1623 (1965). (66) (67) (68) (69) (70) (71) (72) (73) (74) (75) (76) 121 C. Friedlander and J. Kinsinger, unpublished data. R. Beyer and S. Letcher, "Physical Ultrasonics", Academic Press, New York, N. Y., 1969. C. Sutton and J. Mark, J. Chem. Phys., 33, 5011 (1971). P. Flory, V. Crescenzi and J. Mark, J. Am. Chem. Soc., 33, 146 (1964). L. Scales and J. Semlyen, Polyger, 31, 601 (1976). M. Froix, C. Beatty, J. Pochan and D. Hinman, J. Polym. Sci., 33, 1269 (1975). C. Huggins, L. St. Pierre and A. Bueche, J. Phys. Chem., 33, 1304 (1960). G. Allen, P. Brier, G. Goodyear and J. Higgins, Faraday Symp. Chem. Soc., No. 3, 169 (1972). C. Wang and Y. Huang, J. Chem. Phys., 33, 4847 (1976). J. Blitz, "Fundamentals of Ultrasonics", Butterworth and Co., London, 1967. M. Volkenstein, "Configuration Statistics of Polymeric Chains", Interscience, London, 1963. 178 3552 1293 03 "II I. R H all): III"... I“ u". all". “ A I“ ll In" I" I'll. "(Ill 3