NONLINEAR TEMPERATURE blSTNIBuNoNs . . . 1 _ - , DissertatiOn fo'r‘trhe Degree OfPh'y‘D.” " ’ ”'CH'GA“ STATE UNIVERSITY__ I i * ' MNSHALI—‘ch‘Am—R, e _1976* .- w LIBRARY ~"Michism§m .. ABSTRACT NONLINEAR TEMPERATURE DISTRIBUTIONS BY Marshall P. Cady, Jr. Pure, one-component fluids can exhibit large nonlinear temperature distributions in nonequilibrium situations in which nearly linear distributions are usually expected. This dissertation includes (a) a theoretical analysis of molecular energy transport mechanisms which can create nonlinear temperature distributions and (b) a Bryngdahl interferometric study of the quantitative aspects of temperature nonlinearities in liquids. The nonequilibrium thermodynamic regimes studied include time-independent and nonconvecting, time-independent and convecting, and time—dependent and convecting. The possibility that molecular vibrational degrees of freedom can create or contribute to spatial nonlinear, time-independent temperature distributions is explored with a stochastic diffusional energy transport model. Computations indicate that liquids may exhibit temperature jumps at the liquid-boundary interface but that the diffusional mechanism Marshall P. Cady, Jr. of energy tranSport cannot contribute to nonlinearities of significant spatial extent. However, the diffusional mechanism can produce gas phase nonlinearities which extend 2 0.01 cm from boundaries provided that P Q 1 x 10- atm. Computations also show the nonlinear magnitude to be 0.03 K for gases when 1 vibrational degree of freedom in 10‘1 at the boundary fails to be described by a Boltzmann distribution. A temperature jump of 0.03 K is expected for liquids when 1 vibrational degree of freedom in 102 belongs to the nonequilibrium distribution. The diffusion model has applications in describing energy transport away from catalytic surfaces, near boundaries of polyatomic gases in the "temperature jump" regime, and near boundaries of strongly emitting and absorbing liquids. The diffusional energy tranSport model is characterized by energy tranSport via diffusion of fluid molecules, the Landau-Teller transition probabilities for resonant exchange of vibrational energy during bimolecular collisions, and the assumption of non-Boltzmann distribution of molecular vibrational degrees of freedom at boundaries. The non-Boltzmann distribution is the result of boundary- Iluid interactions. Two parameters describing temperature nonlinearity associated with diffusion are deduced. They provide a measure of the spatial extent of the nonlinearity from boundaries and of the magnitude of the nonlinearity. Bryngdahl interferometry is used to examine temperature distributions in liquids bound by horizontal, Marshall P. Cady, Jr. parallel, silver plates which are maintained at different temperatures. For nonconvecting states the linear temperature distributions eXpected for pure thermal conduction are found at distances greater than 0.3 cm from boundaries for ethyl acetate, benzene, and carbon tetra- chloride. However, at shorter distances the distributions are highly nonlinear. As the metal-liquid interface is approached, the temperature gradient increases 4.5% for ethyl acetate, 12.5% for benzene, and 30% for carbon tetrachloride. Full details of the observed temperature distributions and gradient distributions are reported. In addition, we find that the ratio of integrated absolute deviation from linearity(1/l_ SEIT-T3.33ldz where T3.33 is the solution of V-kVT = 0) is 1/2.4/4.7/14.6 for ethyl acetate, benzene, carbon tetrachloride, and water, respectively. The very large integrated absolute deviation exhibited by water is caused by temperature jump phenomena. It appears that the temperature jump is a property of the boundary-liquid system and is not caused by plate temperature control and temperature measurement. Knowledge of liquid nonlinear temperature distributions makes possible an increase in the accuracy of experimentally determined Soret coefficients, thermal conductivities, and nonisothermally determined refractive index temperature derivatives. Finally, experimental studies of nonlinear temperature distributions in convecting liquids are presented. Marshall P. Cady, Jr. The Rayleigh numbers for the critical flow transitions are found to be R.I = 1667 1 137, RIII = 3.3 x 1011 4 i 12%, and R = 4.42 x 10 i 14%. At RI a transition from nonconvection IV to steady state convection occurs; at RIII there is a transition from three-dimensional steady flow to three- dimensional time-dependent flow; and at RIV there is a transition from time-dependent flow to time-dependent flow of increased frequency. Bryngdahl interferometry is uniquely suited to the study of the time-dependent phenomena. Frequencies of temperature oscillations are determined via Fourier transform of the Bryngdahl interferometric image autocorrelation function. Four major bands are observed at approximately 10, 30, 90, and 140 min-1. The band intensity is proportional to e = IV - AT and decays exponentially with angular frequency 75.8 min-1. NONLINEAR TEMPERATURE DISTRIBUTIONS BY A Marshall P2 Cady, Jr. A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1976 To Sharmaine ii ACKNOWLEDGMENTS I thank Professors F.H. Horne and R.I. Cukier for their efforts to improve the manuscript. In addition, I thank my wife, Sharmaine, whose typing, encouragement, and prodding made this dissertation possible. Finally, I express recognition of the influence Professor F.H. Horne has had upon my scientific deve10pment. The excellence of his lectures in mathematical physics, thermodynamics, and nonequilibrium thermodynamics is admired and has provided the fundamental basis of this dissertation. iii TABLE OF CONTENTS LI ST OF TABLES O O O O O O O O O O .4 O O O O O 0 LIST OF FIGURES O O O O O O O O O O O O O 0 O 0 Chapter 1. 3. INTRODUCTION . . . . . . . . . . 1.1 Apercu . . . 1.2 The Interior Conducibility. 1. 3 The Energy Source . . . . . O O O O O O O O O O O O O O O O MOLECULAR VIBRATIONAL CONTRIBUTIONS TO NONLINEAR TEMPERATURE DISTRIBUTIONS NEAR WALLS 2.1 Introduction . . . . . . 2. 2 Model 1: The Diffusion Model 2. 3 Estimates of c1 and c2 . . . 2.4 Model 2 . . . . . . . . . . . 2.5 Discussion . . . . . . . . . O O O O O O O O O O O O O O O O O O BRYNGDAHL INTERFEROMETRY . . . . . . . . . Introduction . . . Bryngdahl Interferometry: Apparatus, cell, Data 0 O O O O Refined Data; Refraction of Light . . Temperature Distribution . . . . . . 2 WWW WNW o O ISOTHERMAL, NONCONVECTING LIQUID STATES Introduction . . . . NOnlinear Temperature Distributions: o P W MHZ @Ulob “NH Index of water . . . . . . . . . . . Discussion . . . . . . . . . . . . . 41' 1!? it's]? 0 iv 0 O O O Bryngdahl Interferometry: Applications Refractive Index: Temperature Dependence Ethyl Acetate, Benzene, Carbon Tetrachloride Temperature Dependence of the Refractive O O O O O Page vi viii hnmth 19 19 25 44 53 56 59 59 61 67 78 89 9o 92 92 101 147 164 Chapter 5. NONISOTHERMAL, CONVECTING LIQUID STATES. 5.1 5.2 5.3 5.4 APPENDIX LIST OF REFERENCES . . . . Chapter 1 References Chapter 2 References Chapter 3 References Chapter 4 References Chapter 5 References IntrOduCtiono o o o o o o o o o o 0 Temperature Distributions and the First Rayleigh Transition . . . . . Fluctuations and Frequency Spectrum sumary o o o o o o o o o o o o o o O O O O O C O O O C O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O 0 0 O O O O O O O O O O O O O O O O O O O O O O O 0 Page 174 174 178 191 200 207 210. 210 213 217 223 227 LIST OF TABLES Tabl e Page 2.1 Gas phase parameters c and c of the molecular fundamental irequengy D. . . . . . . . . 48 2.2 Liquid phase parameters c and c of the molecular fundamental frequency 5.. . . . . . . . 51 pc (v)Hw(v) values for water at 300 K and vagious boundary conditions. . . . . . . . . . . . 58 3.1 Derivative estimates of M.(Equation 3.16b) for water in a 10 deg/cm temperature gradient. . . . . 85 4.1 Temperature dependence of refractive indices: n = a + bT + 0T2 at 632.8 nm, 1 atm. . . . . . . . 103 4.2 Temperature dependence of thermal conductivity: 106k=a+bT+cT2, d$T$eat1atm. . . . . . . 105 41.3 Isothermal and nonisothermal image gata. . . . . . 111 4.4 Nonisothermal fringe number: N8 = 2 a.x1 i=0 1 Where X = (2.00254 cm)/Oo254 cm. 0 o o o o o o o o 125 ‘4.5 Experimentally determined refractive index temperature derivatives: dn/dT = A + BT. A and B differ from literature values of Table 4.1 because of nonisothermal interferometric image diStortj-on O O O O O O O O O O O O O O O O O O O O 132 4.6 Temperature dependence of the refractive index of water as determined by isotherma1(I) or nonisothermal(N) methods: -(dn/dT)104 = a + bT, where céTéd. The parameters a and b are reported in the literature or evaluated with a least squares analysis of reported experimental data at 1 atm. . . . . . . . 15o vi Table 4.7 Page Nonisothermal experimental water data. N(l/2) is the fringe number at image center. 8 is the angle from vertical of fringe in deg. min and corrected for deviant verticality of the isothermal image. Angles of experiments of identical AT and T have been averaged. The The isotherm 1 paraEeters are H = 4.713 cm, 2D = 0.463 cm, and 6 = 1.239 cm with standard deviations of 0.023 cm, 0.018 cm, and 0.018 cm, respectively . . . . . . . . . . . . . . . . . . . 153 Experimental evaluations of the first critical Rayleigh number. . . . . . . . . . . . . . . . . . 188 LIST OF FIGURES Figure Page 1.1 Qualitative features of the interior conducibility, k(T), of amorphous silica . . . . . 9 1.2 Energy transport by translational degrees of freedom. The indicated positions of molecules labeled 1 and 2 are those at successive collisions. (A) TranSport far from walls. (B) Transport in the presence of a reflecting boundary. . . . . . . 15 2.1 The horizontal plate arrangement . . . . . . . . . 22 2.2 The probability distribution {xn(§)} for various g When0=1andxq(0)=1............33 2.3 Negative of H(z) for 0 1, x0(0) = 1, and various 01 . . . . . . . . . . . . . . . . . . . . 38 2.4 H(z) for 0 = 1, x1(0) 1, x4(0) 2.6 H(z) fort? = 1, w1(_l_) = 1, x0(0) = 1, and various values of 01 . . . . . . . . . . . . . . . 46 1, and various 01‘ . . . . 40 2.5 H(z) for 0 1, and various c1. . . . . 42 3.1 (A) The interferometer. (B) The cell-water jaCketassembly.......o..........69 3.2 (A) Best quality image obtainable with interfero- meter, the slit image. (B) Isothermal cell-water jacket assembly image. Z is the vertical image coordinate measured from the upper, bounding plate. 2D is the shear parameter. H is the image cell height. 8 is the distance between fringes in the isothermal cell configuration. . . . . . . . . . . 75 3.3 Refraction of light in a one dimensional gradient. z(y) is the path of a photon pencil of infinitesimal cross section which enters the liquid cell at z = 20, 2f is the exit coordinate. TU)TL. . . . 80 viii _———i Figure Page 4.1 EA) Reduced temperature gradient, dT/dz)/(dT/dz)1 2 versus z/l for carbon tetrachloride when 1 = 1.5 Cm, e = 0.05, and ATO = 2.00 c. From-G. Schodel and U. Grigull.10 (B) Deviation from linearity, T — T3 33, of CCl4 in a cell 0.810 cm high. TU = 26.4000, T = 23.6o°c. From J. D. Olson.12 . . . . . . . . 100 L 4.2 Light refraction in a nonisothermal experimental configuration; TU) TL' The pencil of light shown enters the li uid at z = 0. It leaves at coordinate zf O) and a3g1e af(0). This pencil travels through the glass cell at angle Bf(0), refracts in air, and travels at angle Yf(0) until direction is altered by the Bryngdahl optical train. 0 O O O I 0 O 0 O 0 O O O O O O O O O O O O 108 4.3 Nonisothermal Bryngdahl interferometric image of water at several temperature differences between upper and lo er silveroplates. l = 3.349 cm; ATO = 10.028 C, 13.047 C, and 16.691 C . . . . . . 113 4.4 Nonisothermal Bryngdahl interferometric image of ethyl acetate at several temperature differences between upper and lower silver pla es. 1 = 1.349 cm; ATo = 2.244 C, 6.261 C, and 10.019 C C O O O O O O O O O O O O 0 0 O O O O O l 115 4.5 Nonisothermal Bryngdahl interferometric image of benzene at several temperature differences between upper and 18wer silvgr plates.O l = 1.349 cmé ATO = 2.399 C, 7.256 C, 10.529 C, and 13.063 C . . 117 4.6 Nonisothermal Bryngdahl interferometric image of carbon tetrachloride at several temperature differences between upperoand lowea silver plates. 1 = 1.349 cm; ATO = 2.637 C, 5.976 C, 7.729 C, and 12.713 C O I O I O O O O O O O O O O 0 O O O O O O 119 4.7 Nonisothermal Bryngdahl interferometric image of methanol at several temperature differences between upper and lower sglver plates. 0 l = 1.349 8m; ATO = 4.739 C, 9.297 C, 12.125 C, and 13.488 C I I O O O O O O O O O O I O I O O O O 121 ix Figure 4.8 4.9 4.12 Page Tem erature deviation from linearity, T(z - T3 33(z), for benzene at several temperature differences between upper and lower silver plates. The refractive index temperature derivatives of Table 4.1 are used in the method of Section 3.6 to compute T(z) from the refined experimental data. 1 = 1.349 cm. 0 I O C O O O I O O I O O O 0 O O O 127 Tem erature deviation from linearity, T(z - T3 33 z , for carbon tetrachloride at several temperature differences between upper and lower silver plates. The refractive index temperature derivatives of Table 4.1 are used in the method of Section 3.6 to compute T(z) from the refined experimental data. 1 = 1.349 cm. I 0 l O O O O O O I O O O O O O I O 129 Tem erature deviation from linearity, T(z - T3 33(z), for ethyl acetate at several temperature differences between upper and lower silver plates. These curves are corrected for image distortion by using the refractive index temperature derivatives of Table 4.5 in the method of Section 3.6 to compute T(z) from the refined experimental data. 1 = 1.349 cm. . . . . 134 Tem erature deviation from linearity, T(z - T3 33(z), for benzene at several temperature differences between upper and lower silver plates. These curves are corrected for image distortion by using the refractive index temperature derivatives of Table 4.5 in the method of Section 3.6 to compute T(z) from the refined experimental data. l = 1.349 cm. . . . . 136 Tom erature deviation from linearity, T(z - T3 33 z , for carbon tetrachloride at several temperature differences between upper and lower silver plates. These curves are corrected for image distortion by using the refractive index temperature derivatives of Table 4.5 in the method of Section 3.6 to compute T(z) from the refined experimental data. A = 1.349 cm. 0 O O O O O O O I O I O O O O O 0 O 138 Figure 4.13 4.18 Page Reduced temperature gradient, (dT/dz)/(dT/dz)1/2, for ethyl acetate, benzene, and carbon tetrachloride. Corrections have been made for interferometric image distortion. 1 = 1.349 cm I o o o o o o o o o o o 143 Integrated absolute deviation from linearity, l (1/l)SOIT - T3.33ldz, as a function of ATO. Corrections have been amde for interferometric image distortion. The placement of the water curve is discussed in Section 4.3. l = 1.349 cm for ethyl acetate, benzene, and carbon tetrachloride. l = 0.474 cm for water. . . . . . 145 Per cent deviation of dn(l,T)/dT from values reported by Tilton and Taylor for water at 1 atm. 149 Temperature dependence of the refractive index. The raw experimental data is refined with the method of Section 3.4, and dn/dT mapped with the method of Section 3.5 assuming that T(O) = TU and T(l) = TL. Equation 4.38(solid line) is the best linear fit of the resulting temperature derivatives 0 O O O O O O O O O O O O O O I I O O 157 f(T) = -(dn/dT)102/(-d2n/dT2)1/2. The data are computed by refining the raw data with the method of Section 3.4, and mapping dn/dT and d2n/dT2 with the method of Section 3.5 assuming that T(O) = TU and T(l) = TL. Equation 4.40 (solid line) is the best linear fit of the above ratio. The dashed line is predicted by Equation 4. 38 I O I O O O O O O O O O O O O O O O 160 (A) Molecule 1 emits an "average" photon which is absorbed by molecule 2. This energy transport process heats the horizontal layer of molecule 2 because the photon originates within a layer of higher temperature. (B) Placement of a perfect mirror between molecules 1 and 3 alters photon energy transport processes. The photon absorbed by molecule 2 now originates from molecule 3. Since the horizontal layer of molecule 3 is of lower temperature than the layer of molecule 1, energy flux differs from the above figure . . . . 170 xi Figure Page 5.1 Nonisothermal Bryngdahl cell images. I = 1.440 cm, a = 5.961 cm, H = 3.175 cm, 2D = 0.138 cm, 8 = 0.353 cm. a a o o o o o a u o o 180 5.2 Temperature distributions for various impressed temperature gradient . . . . . . . . . . 183 5.3 Experimentally determined boundary(z = 1) temperature gradient as a function of Rayleigh number. Curve 1 is the linear least squares analysis of data(Equation 5.8); Curve 2 is the prediction of the Laplace equation description of temperature . . . . . . . . . . . . 186 5.4 Time sequential temperat e distributions for R(TM,AT) = 1037 x 10 o o o o o a o o c o o o o 190 5.5 Intenisty fluctuations at image center oi Bryngdahl image for R(TM,AT) = 8.37 x 10 . . . . . 193 5.6 Intensity fluctuations at image center of Bryngdahl image for various impressed temperature gradients. . . . . . . . . . . . . . . 195 5.7 Bryngdahl image intensity fluctuation rate at cell center versus impressed temperature gradient and versus Rayleigh number. . . . . . . . 198 5.8 Autocorrelation function of AI(t) for 5 AT = —6.849°C and R(TM,AT) = 3.71 x 10 , . . . . . 202 5.9 Fourier transform of autocorrelation function of AI(t) divided by e = IATI - 0.72700 . . . . . . 204 CHAPTER 1 INTRODUCTION 1.1 Apergu The study of nonlinear temperature distributions in pure, one-component fluids provides macroscopic, continuum data on which molecular theories of energy transport processes may be based. Clearly, all temperature distributions in nonequilibrium media are the consequence of molecular energy transport via vibrational, rotational, translational, and electronic degrees of freedom. Because of this relationship between molecular mechanism and macroscopic observation, nonequilibrium thermodynamic studies are an important 1’2 to catalog molecular complement to the modern effort energy levels and transition probabilities. Particulars of this work include both the conceptual illumination of molecular mechanisms which can cause nonlinear, Steady state temperature distributions and Bryngdahl interfer- ometric evaluation of temperature distributions in pure, one-component liquids. The possibility that vibrational degrees of freedom play a contributing role in the establish- ment of nonlinear, steady state temperature distributions in nonconvecting fluid media is explored in Chapter 2. On the IIIIIl--..____ 2 assumption that the temperature distribution far from the walls of the fluid container is linear, mechanisms are presented which invalidate the steady state Fourier-Laplace equation description of the temperature distribution when wall effects are important. Molecular diffusion and energy flux models of vibrational energy transfer are used to analyze the problem quantitatively. Fluid-dependent parameters characterizing the magnitude of the temperature nonlinearity and its extension from the container-fluid interface are deduced. Energy trans- port by diffusion is shown to contribute negligibly to non- linear temperature distributions observed within the liquid phase, but this mechanism may contribute significantly to the establishment of nonlinearities within low pressure gases. Before experimental evidence of nonlinear temperature distributions in both nonconvecting liquids(Chapter 4) and convecting liquids(Chapter 5) is presented, it is necessary to discuss details of the Bryngdahl interferometer which is used as a temperature probe. This is the purpose of Chapter 3. After a survey of the applications of Bryngdahl interferometry in Section 3.2, details of the optical apparatus, the horizontal parallel plate-cell composite, and experimental procedure are described in Section 3.3. The horizontal parallel plate arrangement heats the liquid cell at either the upper or lower plate and cools it from the opposite plate. This results in a continuous vertical temperature distribution across the liquid ‘which causes refraction of the interferometric light beam. Beam refraction must be accounted for when spatial derivatives of the refractive index are mapped from the unrefined Bryngdahl 5 interferometric data. The mapping is dw/dz ——9 dn/dz (1.1) where w, n, and z are the optical path, refractive index, and vertical coordinate, respectively. The left side of Equation 1.1 is called the unrefined data; the right is called the refined data. The mathematical mapping procedures are established in Section 3.4. Once the refined data have been mapped from the unrefined Bryngdahl interferometric data, one can either assume knowledge of the temperature gradient and compute the temperature derivative of the refractive index with the chain rule expression dn/dT = (dn/dz)/(dT/dz) (1.2a) or assume knowledge of the temperature derivative of the refractive index and compute the temperature gradient with the equation dT/dz = (dn/dz)/(dn/dT). (1.2b) The alternative mathematical prodecures are established in Sections 3.5 and 3.6, respectively. Direct interferometric observations of nonlinear, steady state temperature distributions in nonconvecting liquids are reported in Chapter 4. After an introductory discussion of linear and nonlinear distributions in Section 4.1, the qualitative and quantitative features of nonlinearities exhibited by ethyl acetate, benzene, and carbon tetrachloride are presented in Section 4.2. Besides a thorough detailing of problems inherent in nonequilibrium experimental studies, Section 4.2 includes data which lead to (1) correlation of the dependence of the nonlinearity upon the impressed temperature 4 difference between the liquid boundaries, (2) Specification of the Spatial domain of the nonlinearity, and (3) comparison of the relative magnitudes of exhibited nonlinearities between liquids. All measurements are at ambient pressure. Mean temperatures are between 24°C and 29°C, and boundary temperature differences are between 2°C and 13°C for the 1.349 cm high liquid cell. In Section 4.3 the temperature dependence of the refractive index of water between 24°C and 40°C at 632.8 nm and atmOSpheric pressure is obtained. Temperature gradients range from 4 deg cm"1 to 16 deg cm_1. Results comparable to isothermal determinations of the temperature dependence are found when the data analysis does not include the measured temperature difference between the upper and lower plates of the horizontal parallel plate system. The presence of temperature jumps would explain discrepancies which occur when the temperature difference is used in the analysis. Finally, the Bryngdahl interferometric examination of convecting thermodynamic states exhibited by water is presented in Chapter 5. There are four major features. These include the determination of the critical Rayleigh number for two types of transitions: the transitions from nonconvection to steady state convection and from steady state convection to time- dependent, turbulent convection. The nonlinear temperature distributions between these transitions are mapped. Frequencies of aperiodic motions in the turbulent states are determined. 5 1,2 The Interior Conducibility The nonequilibrium description of temperature(T) in a one-component, pure, nonreacting fluid depends upon two fundamental statements. The first is the law of conservation of energy: E + V~gi = oi (1.3) where E = E(g,t), g1 = gi(r,t), and 01 = oi(r,t) are the Specific internal energy per unit volume, the i'th form of heat flux, and the i'th form of energy source per unit volume per unit time, respectively. The heat flux and energy source have been given the superscript i because differing physical interpretations and mathematical descriptions of the energy source are associated with nonequivalent heat flux. The i'th form of the energy source is associated with the i'th form of the heat flux. The second important statement is a phenomenological relationship between heat flux and temperature gradient. For isotropic media it is J1 = -kiVT (1.4) i = k1(r) is the i'th form of the conductivity. where R For temperature fields which are uniform in the horizontal plane and for which E = 0, Equations 1.3 and 1.4 simplify to (d/dz)ki(dT/dz) = -01. (1.5) The solution of this differential equation subject to the boundary conditions T(o) = T and T(l) = TL (1.6) U is the temperature field. l,is the distance between 6 horizontal boundaries. Equation 1.5 is also used in the experimental evaluation of k1. One integration yields: 1 i -k (z) = (JT + Io (z)dz)/(dT(z)/dz) (1.7) where JT is the total heat flux in the z direction and will be called the heat flux. The heat flux, dT(z)/dz, and T(z) are experimentally determined; then, 01(2) is assumed and ki(z) computed with Equation 1.7. A final mapping ki(z) -—9 ki(T) yields the i'th form of the conductivity. A particularly important form of the conductivity is given by Fourier's3 "interior conducibility," k(T), which is defined by Jz = —k(dT/dz). (1.8) This constitutive conductivity form is recognized by many modern authors!"11 of nonequilibrium thermodynamics. There are two criteria to help with the deduction of the energy source form(hereafter called the energy source) which is conjugate to the interior conducibility. The first is due to Truesdell(as reported by Petroskia): the energy source is determined by given functions E and g and Equation 1.3. Secondly, we wish to construct the interior conducibility to be a property of the media being studied. Therefore, the interior conducibility must describe energy tranSport by all possible modes of energy transport solely characteristic of that media. Energy transport modes associated wtih container walls and the external world are not described by the interior conducibility. Very little is known about the 7 energy source,8 but with these criteria the necessary concepts will be developed in the next section. The multitude of possible energy transport mechanisms is illustrated by the interior conducibility of amorphous silica in Figure 1.1. At very low temperatures k~vT3. This dependence is the result of phonon processes for which k ~ CVL (1.9) where C is the heat capacity per unit volume, V an averaged phonon velocity, and L an averaged phonon mean path. At low temperatures L is constant because of crystal boundary and grain size restrictions; the temperature dependence of Equation 1.9 is then determined by the heat capacity. At low temperatures this dependence is the famous T3-law of the Debye theory of heat capacity of crystals. With increasing temperature, a second region of extreme complexity is found which is called "the knee." The knee is apparently caused 12 by resonance scattering of phonons by local defects. At still higher temperature a very broad region exists in which the interior conducibility is very weakly dependent upon temperature. In this region k is limited by phonon-phonon processes and by the relative inability of many of the vibrational modes, due to high spatial localization, to transmit energy. Finally, at very high temperature there is a fourth region of strong T3 dependence;13"16 energy transport by photon processes have become very important. .A photon or radiation transport process is to be understood as either the induced or spontaneous emission of a photon Figure 1.1 Qualitative features of the interior conducibility, k(T), of amorphous silica. ._. 6.29.368. :35. of... 29. MP 2 Hx >1 ‘Mugqgonpuog JO!J6.IU| 10 by a degree of freedom of the medium and the absorption of that photon by another degree of freedom. Vibrational degrees of freedom are major contributors to this mechanism via infrared photons. The second major conductivity form is called the "thermal conductivity." In this form the conjugate energy source contains all radiation tranSport processes, both those in which the photons originate within the medium and those in which thephotonsoriginate in the external world. Energy transport processes due solely to thermal motion of the elementary particles of the media are described by the 17.21 Our objection to this division thermal conductivity. of conduction and radiation is based on the presupposed ability to write mathematically and nonphenomenologically a term describing all photon processes. This term is necessary in the Equation 1.7 evaluation of the thermal conductivity. By this procedure each photon model yields a characteristic thermal conductivity and those of different laboratories are impossible to compare properly unless full computational details are published. We believe that once any model of radiation tranSport for a particular medium has been experimentally proven, then the above division of conduction and radiation becomes a valid procedure. This, luneever, is not the usual practice. For the history of the separation of conduction, radiation, and convection see 22 for the first suggestion that conductivity theory 3 Brush; 0 can also include radiation transport, see Stokes.” 11 1. The Ener Source Because the interior conducibility describes all energy transport processes characteristic of a given media, the energy source is zero far from boundaries and the external world. In this domain Equation 1.5 becomes the Fourier-Laplace equation: (d/dz)k(dT/dz) = 0. (1.10) The interior conducibilities of gases and liquids are very much like kIII(T) of Figure 1.1; they exhibit a very weak temperature dependence in a broad temperature region which may be called the "ordinary temperatures." For example, for gaseous carbon monoxide at 1 atm: 9 x 10"3 K"1 3 (1/k)(dk/dT) 2 6 x 10’“ K‘1 when 90 K s T s 1100 K.24’25 For carbon tetrachloride at 1 atm: 1.7 x 10-3K-1é -(1/k)(dk/dT)$ 2.2 x 10"3K‘1 when 255 K5 Ts 378 K.26 This demonstrates that for liquids and gases subjected to temperature differences of the order of 10 deg cm-a thermal conducibilities are essentially constant at ordinary temperatures. Thus, the temperature of fluids far from boundaries is expected to be the solution of the Laplace equation: T(z) at a + bz (1-11) where a and b are constants. This linear relationship is a consequence of the absence of an energy source term in Equation 1.10. Thus, Equation 1.11 emphasizes that the study of nonlinear temperature distributions in nonconvecting, steady state fluids must be the theoretical and experimental pursuit of the concepts behind the existence of non-zero energy source, source magnitude, and spatial domain of the source. 12 There are a number of nonlinear temperature distribution studies at ordinary temperatures and pressure (~1 atm). For a review of nonlinear observations within liquids see Section 4.1; see Figure 4.1 for an example of the nonlinearity within carbon tetrachloride. Nonlinearities at ordinary temperatures and pressure have also been found within some gases. In 1964 Gille and Goody27 used horizontal aluminum boundary plates spaced 2 cm apart and a Michelson interferometer to study temperature distributions in dry air and ammonia. Linear temperature profiles were found at all distances from boundaries for dry air. Ammonia, however, displayed nonlinear temperature profiles. The nonlinearities were found to be antisymmetric about the intermediate vertical coordinate with a negative deviation from linearity on the half closest to the heated boundary and a positive deviation on the half closest to the cooled boundary. A maximum in the deviation from linearity appeared at about 0.4 cm from the boundaries. It was 0.1800 at 0.7 atm and 0.2100 at 1.3 atm. The nonlinear temperature profiles are explained on the basis of absorption and re-emission of radiation from plates and molecular diffusion. Schimmel gt_§128 have used Mach-Zehnder interferometry to deduce boundary temperature gradients in pure gaseous 002, N20, and mixtures of CO2-CH4 and CO2-N2O in a cell bound by horizontal aluminum plates 2.55 cm apart. They report that for N20 at a temperature difference between plates of 10.700 and a mean temperature of 33°C the ratios of boundary temperature 13 gradient to the gradient predicted by Equation 1.11 are 1.09, 1.13, 1.17, 1.20, and 1.20 at pressures of 0.25 atm, 0.50 atm, 1.00 atm, 2.00 atm, and 3.00 atm, respectively. The ratios for C02 behave similarly with pressure, but are somewhat smaller. These nonlinearities are attributed to radiation transport processes. Experimental studies have shown the nonlinear phenomena to be important in vertical 29 plate systems containing pure ammonia and nitrogen- ammonia mixtures.30 The final example of nonlinear temperature distribution is provided by the observations in gases at low pressure ($10—2 atm) of what is called "temperature jump" and by the association of temperature jump with gas-wall thermal accommodation. For the history and early development of the concept of temperature jump see References 31 and 32; for the association with thermal accommodation see References 33-36; for the Boltzmann equation approach to the problem see References 37 and 38. We present a new conceptualization of the mechanism behind this nonlinearity. Consider a gas bound by two parallel plates of infinite extent. These plates are maintained at different temperatures with the "lower" wall having lower temperature, and they are far enough apart that at intermediate distances from them all energy transport is due solely to the characteristics of the gas. Figure 1.2A depicts energy tranSport by thermal motion at these intermediate distances. The indicated particle positions are those of successive 14 Figure 1.2 Energy transport by translational degrees of freedom. The indicated positions of molecules labeled 1 and 2 are those at successive collisions. (A) Transport far from walls. (B) Transport in the presence of a reflecting boundary. See text. T T, Vertical Direction G B Q Q) \/ Mirror 16 gas molecule-molecule collisions. These translational motions and collisions transport thermal energy. For example, the motion and subsequent collision of molecule 1 results in the effective heating of the horizontal layer of the collision; the motion and subsequent collision of molecule 2 effectively cools the horizontal layer of the collision. All such energy transport processes are described by the interior conducibility. Suppose a perfect mirror which reflects molecules specularly and elasticly is placed between molecules 1 and 2,andthen energytranSportprocesses are altered. Molecule 1 is now reflected by the mirror to the collision indicated in Figure 1.2B. But the interior conducibility being ignorant of the mirror's presence describes the origin of the colliding molecule as corresponding to the initial position of molecule 2. This is the mirror image of the initial position of molecule 1. The interior conducibility is assigning less energy to the colliding molecule than it actually has. Thus, the temperature distribution near the lower boundary displays a positive deviation from linearity; the energy source is positive in this domain. Now, considering the reflection of molecule 2, molecule 2 is reflected into a collision but carries less energy into that collision than predicted by the interior conducibility. The interior conducibility describes the reflected molecule as having originated at the initial position of molecule 1. This position corresponds to a horizontal layer of higher temperature and, therefore,higher 17 energy than the horizontal layer of the true initial position of the reflected molecule. Thus, the temperature distribution near the upper boundary must display a negative deviation from linearity; the energy source is negative. This conceptualization has assumed that the linear temperature distribution observed in the gas far from boundaries is maintained by some unspecified mechanism. On this basis it analyzes the effect of perfect mirrors which have a zero thermal accommodation coefficient(no energy is transferred from mirror to molecule during collision). This does not correspond to any possible laboratory experiment. However, within the framework of the model it can be easily reasoned that (1) the negative (positive) deviation from linearity prediction of this model at the upper(lower) boundary represents the maximum obtainable and (2) negative(positive) deviation from linearity at the lower(upper) boundary is not possible for any boundary-gas system which has a thermal accommodation coefficient between 0 and 1. The thermal accommodation coefficient is defined as being equal to (Ti-Tr)/(Ti-Tw) where Ti corresponds to the temperature of molecules incident upon a wall, Tr corresponds to the temperature of molecules reflected by a wall, and Tw is the temperature of the wall. Of course this conceptualization can be made phonsically realistic through the inclusion of adsorption, 18 desorption, and inelastic scattering of molecules by boundaries. The prescription for the mathematical develop— ment of the energy source is clear, however. It is the sum of energy transport processes which are occurring because of the presence of walls minus those energy transport processes which are described by the interior conducibility but are not occurring because of the presence of walls. Nonlinear temperature distributions result from radiation energy transport via mechanisms which are analogous to the above translational model. This is discussed in Section 4.4. An analogous mechanism of phonon reflection and bulk phonon-grain surface phonon interaction may also be responsible for the belief that the Fourier heat law is not valid at very low temperatures in amorphous solids.16 In this case the energy source grows with decreasing temperature. CHAPTER 2 MOLECULAR VIBRATIONAL CONTRIBUTIONS TO NONLINEAR TEMPERATURE DISTRIBUTIONS NEAR WALLS 2.1 Introduction In this chapter we examine quantitatively the possibility that vibrational degrees of freedom of a gas or a liquid contribute to spatial nonlinear temperature distri- butions within a single component fluid in the nonflowing steady state. A model in which molecular vibrational energy is transported during diffusion of molecules is considered in detail. Computations indicate that the diffusional mechanism of energy transport cannot contribute to nonlinearities of significant spatial extent within dense media. However, the diffusional mechanism can produce gas phase nonlinearities which extend 0.01 cm from boundaries provided that P s 1 x 10"2 atm. 0n the assumption of a non-Boltzmann distribution of molecular vibrational degrees of freedom at boundaries, cemputations show the nonlinear magnitude to be 0.03 K for gases when 1 vibrational degree of freedom in 1011 at the boundary fails to be described by a Boltzmann distribution. A temperature jump at a liquid-boundary interface is expected 0 When 1 vibrational degree of freedom in 10“ belongs to the 19 IIIIIIIlI-—___________________________________________nnnnnnnnnn______ 20 nonequilibrium distribution. To explore the possibility that the interaction between vibrational degrees of freedom of the fluid and the wall can result in nonlinear temperature distributions, consider Figure 2.1. The coordinate 2 measures distance from the upper plate, which is considered to extend to (i) infinity in the (x,y) plane. The distance between the two infinite plates is l. The upper plate is maintained at temperature TU while the lower is at temperature TL, with T T U L’ each layer having width z and the ith layer being associated Now imagine the fluid as being divided into layers, with the temperature T(i). Let the 1th layer be far enough away from the wall that wall effects are negligible. The transfer of energy on the molecular level can be envisioned as follows: Energy is continually being transferred from the 1th layer to the layers j+1 and j-i. For example, when a molecule goes from the 1th layer to the layer j+1, it effectively heats layer j+1 because its translational, rotational, and vibrational degrees of freedom are distributed at T(j), which is greater than T(j+1). Once in layer j+1, this excess energy is dissipated. Conversely, when a molecule undergoes the layer transition j+1-9j, the jth layer is cooled. Energy is also transferred via emission and absorption of photons 'between layers. Far from walls Fourier's interior conduci- 21 Figure 2.1 The horizontal plate arrangement. _ _ _ _ _ _ _ _ 1| 2 3 _ _ r “N... _ L_ _ _ _ /// I] l/l/l/I/ // // ////// // xxx/xx // /////// // ///// /// / ///// 2 /_.IU _ 1 1+- .1— _ _ // ///////// // TL Lower Plate 23 bility incorporates all such processes, and Equation 1.10 describes the temperature distribution. There is no experimental evidence to the contrary. Near the wall, however, the Fourier-Laplace equation may fail. For example, a molecule of the first layer is prevented from entering the "zeroth" layer by the presence of the wall. When reflected back to the first layer, it may have a translational energy corresponding to TU but have its original vibrational energy distributed at temperature T1. Since this reflected "reentry molecule" does not have the vibrational temperature TU, where TU) T1, layer 1 exhibits a temperature lower than that predicted by V2T = 0 in the steady state. Next to the lower plate, a reflected molecule may have its vibrational energy distributed at a temperature higher than TL’ As a consequence, the layers next to the lower wall will exhibit temperatures higher than that predicted by V2T = 0. This is one way in which walls might interact with vibrational degrees of freedom to produce a nonlinear temperature distribution. Analogously, reflection by the walls of photons which have been emitted by vibrational degrees of freedom of the fluid results in nonlinear temperature distributions qualitatively identical to the temperature distributions which result from molecular reflection. Also, if the upper wall emits and absorbs photons at rates and extents different from the fluid at 24 the same temperature, it is possible that there is a difference in the net number of photons absorbed by the vibrational degrees of freedom in layers near the wall beyond that described by V2T = 0 in the steady state. One final mechanistic example is that of a wall which catalyzes the production of molecules in higher vibrational states. Upon reentry into the layer next to the wall these molecules heat the layer with a resulting nonlinear temperature distribution. In the following sections, the interaction between the fluid and the wall is not directly analyzed. Rather, it is proposed that these interactions, whatever they may be, result in a non-Boltzmann distribution of vibrational energy in the layer adjoining the wall. Furthermore, it is assumed that this nonequilibrium distribution is known. Temperature distributions can then be calculated once the energy source term which causes deviation from the Fourier-Laplace temperature description has been evaluated. Specific models of vibrational energy transfer are used to evaluate the energy source. The number of nonequilibrium molecules at the boundary needed to achieve deviations from linearity of 10-2 K at ordinary temperatures and pressures is estimated, and parameters which determine the thickness of the nonlinear boundary layer are deduced. 25 The concepts used to develop the above nonequilibrium problem are stochastic: the quantities of interest are probabilities and averages rather than classical deterministic quantities. Much of the mathematics and many of the stochastic concepts developed by Rubin, Shuler, and Montroll are applied directly.1”S The models developed belong to class of models which include first-passage time problems, harmonic oscillator relaxation in a heat bath, relaxation of two interacting systems of harmonic oscillators, relaxation of Rayleigh and Lorentz gases, reaction kinetics, and nucleation theory. Reviews of these stochastic processes are available.6-13 2.2 Model 1: The Diffusion Model The time—independent temperature distribution is described by Equation 1.5: k(d2T/dz2) = -o (2.1) where k is the temperature-independent interior conducibility and o is the energy source which vanishes far from the ‘boundaries. In order to describe temperature distributions in.all domains, including that in which walls play an :hmportant role, we assert that we may divide the one 26 component fluid into two systems, the first of which is the "heat bath" (HB) system. The heat bath system of the 1th layer has all degrees of freedom, vibrational included, distributed in a Boltzmann distribution at temperature T(i). Thus, the heat bath system is associated with the measured temperature T. The second system is the system of interest(SI). The system of interest of the ith layer does not have its vibrational degrees of freedom distributed in a Boltzmann distribution at temperature T(i). Rather, the distribution is a nonequilibrium distribution which depends on the nature of the interaction between fluid and wall. The system of interest also interacts with the heat bath system via molecular collisions and perhaps radiative transfer, 1:3,, exchange of photons between HB and SI. This interaction results in a source of energy for the heat bath system, 0(2). In order to evaluate the energy source, models must be developed which enable one to describe the number density of the vibrational states of the system of interest as a function of the spatial coordinate subject to a knowledge of the vibrational energy distribution at 2:0 and 2:1, iég., sxflaject to boundary conditions on the distributions. Once the vibrational distributions have been completely determined, it is possible to write the energy source term of the bath system and solve Equation 2.1 for the temperature distribution Stflxject to the boundary conditions T(z) = TU at z = 0 and T(z) = T t z = 1. La 27 The first model, the diffusion model, has the following characteristics: (a) (b) (C) Both the system of interest(SI) and the heat bath system(HB) are composed of harmonic oscillators of frequency v. pk is the number of SI molecules per unit volume in the 5th vibrational state, and pk = pk(£,t). The density of SI molecules is such that SI molecules collide only with EB molecules and never with other SI molecules. In these SI-HB molecular collisions, resonant transfer of vibrational energy is possible. 14 Furthermore, we assume the Landau—Teller transition probabilities per collision: Pn,m = P1,0((n+1)6m,n+1 + n5m,n-1) = Pm,n’ (2'2) where Pn,m is the probability per collision that a molecule originally in the 3th harmonic state will undergo transition to the mth harmonic state, and 6m,n is the Kronecker delta, 0 if m i n 5 = 2.3) 11fm=n See References 7 and 9 for a more detailed discussion of Equation 2.2. For more recent developments, see References 15-17. SI molecules in the nth vibrational state diffuse at the rate DV2pn, where D is the diffusion coefficient and is temperature independent. 28 Because of these characteristics, the rate of change of pn at position‘s can be decomposed into the sum of a diffusion term and an SI-HB interaction term. The SI-HB interaction term is given in Reference 4. The result is (apn/at) = (apn/at)diffusion * (apn/at)SI-HB (2°43) where = DV2p (2.4a) (apn/at)diffusion n by characteristic (c), and (apn/at)SI_HB = K(1-e‘”)‘1(ne‘”pn_1 + (n+1)pn+1 - ((n+1)e-0 + n)pn) (2.4c) by characteristic (b). In the last equation, 0 = (hv/kT), where h and k are the Planck and Boltzmann constants, respectively, v is the vibrational frequency of the molecules, and T is the temperature at 5. Both the diffusion constant and the interaction constant, K, are assumed to be temperature independent. For the steady-state case, Equation 2.4 becomes 2 -0 -1 -0 DV pn + K(1-e ) (ne pn-i + (n+1)pn+1 - ((n+1)e'9 + n)pn) = 0. (2.5) Thezsolution of this equation requires two boundary conditions: 'thervibrational distribution of $1 at z = 0 and that at z ==‘l. In order to apply Montroll's and Shuler'sq Fourier series solution to the problem, SI must first be separated itho two systems. The first system is the nonequilibrium 29 distribution associated with the presence of the upper wall. Its vibration number density distribution is designated as {3.}. The second nonequilibrium distribution is that J associated solely with the presence of the lower wall. Its vibrational number density distribution is designated as )rj}. The set )sj} satisfies Equation 2.5 in the coordinate system originating from the upper wall and also satisfies the distribution boundary condition on the upper wall. Similarly, the set )rj} satisfies Equation 2.5 in the coordinate system originating from the lower wall and also satisfies the distribution boundary condition on the lower wall. When the nonequilibrium distribution associated with both walls relaxes to a Boltzmann distribution within the distance ;, SI is the sum of these two distributions. We now make the assumption that SI is this simple sum. This is not a very restrictive assumption when SI is a small fraction of HB. For instance, suppose that SI is 1% of H3 at the upper surface and that SI-HB interactions are negligibly ‘weak. In such a case the nonequilibrium distribution associated with the upper wall is altered by 1% at the lower 'wall due to the presence of the lower wall. The fraction of inolecules involved is then a mere 0.01% of HB, a number ‘which is negligible relative to the number of molecules in the nonequilibrium distribution associated with the lower 'wallx Later it will be shown that either the nonequilibrium (tistributions relax well within the distance I, or that SI i1; a small fraction of HB. With the two systems of SI 30 clearly defined, Equation 2.5 becomes DV2sn + K(1-e-0)-1(ne-0s + (n+1)s n-i n+1 - ((n+1)e'9 + n)sn) = o (2.6) which can be solved subject to the boundary condition sj(z) = Sj(0) at z = 0. The set {r3} satisfies an identical equation. To put this into dimensionless form, we make the following definitions: c1Z/l. XJ-(C) = sj(§)/§si(§). wJ-(L) = rj(§)/2iiri(c). C c1 (hf/D)“2 (2.7) Furthermore, we constrain 2si(§) and 21‘1“) to equal 1 i Es.(0) and 2r.(c ), respectively; 5x.) and {w.} are then 1 1. i. 1 1 1 1 probabilities. In accordance with the assumed form of SI, these probabilities must relax to a Boltzmann distribution far from their walls of origin. Equation 2.6 then becomes -0 2 2 -6 (1-e )(d xn/dg ) + (ne xn_1 + (n+1)xn+1 - ((n+1)e-6 + n)xn) = 0. (2.8) It is easily shown that the variation of 0 with L in Equation 2.8 is negligible to within 1% for IS 1 cm when AT/_1_.$ 5 deg/cm at ordinary temperature. Since this corresponds to the usually experimental condition, we hereafter restrict ourselves to this case. With these conditions the Fourier solution of Equation 2.8 is(see Appendix) 31 co _{1/2 xnm -—- zafiuhnew C (2.9a) =0 .1....(v) = F(—n. n+1. 1; 1-5") = e‘MflEOU-e”) (3H3) (2.91:) -0 w 0 aAU) = (1-6 )mEdlk(m)xm(O)' (“'90) Comparison of this solution with the Appendix 1 solution of Montroll and Shulerk shows that only the exponential power is different, in that the square root is now required. This difference is a consequence of the diffusion differential equation being second order. In these equations, F(a, b, c; z) is the hypergeometric function,18 and the binomial coefficients have the usual meaning, nJ/v!(n-v)! rfll>v (3) = 0 if n - (a/ay)1/2) = o. (3.10) The second term is zer0,and the first is equivalent to (d/az)(n(ay/dz>(1 + (dy/dz)2)'1/2)= 0. (3.11) Integration, evaluation of the integration constant at z = z and use of the experimental criterion dz(0)/dy = 0 0’ Yields «n(z)/neon2 - 1)(dy/dz)2 = 1; (3.12) 83 or, ay = :az/<(n)2 - 1>1/2. (3.13) Because y and z must not be imaginary, n(Z)/n(zo) > 1. (3.14) and the light beam must curve toward the higher refractive indices; that is, toward the lower temperatures. Thus, if the temperature decreases with increasing 2 in a particular experimental configuration, then the positive sign is used in Equation 3.13. If temperature increases with increasing 2, then the negative sign must be retained. We now restrictihe discussion to the case of heating from above with the z coordinate originating from the upper wall. The resulting equations apply to the case of heating from below but with the coordinate system originating from the lower wall. To facilitate the integration of Equation 3.13, expand n(z)/n(zo) in a Taylor series about z = z . Then, 0 n(z)/n(zo) = msgmmzfl (3.15a) where Nm = (1/m!n(zo))(dmn(zo)/dzom) (3.15b) and A2 = z - zo. (3.15c) Furthermore, (n(z)/n(zo))2 evsgzvmzfl (3.16a) Ivhere M =gum'NV'm. (3.16b) maO lVith Equations 3.15c, 3.16a, and the requirement y(zo)=0, 84 integration of Equation 3.13 yields Aj(v 21M Wl/gdw. (3.17) Table 3.1 shows that for water in a temperature gradient of 10 deg/cm, (Mv)v_ _3 4 are of negligible magnitude "" ’0. relative to M They can, therefore, be neglected. 2. The refractive indices of most liquids have first order temperature variations which are less dependent upon temperature than those of water; for them, (Mv)v=3 4.. are less important. Since Am ( 10m, the third order term of Equation 3.17 is less than 0.1% of the second order term, and can safely by discarded. Equation 3.17 becomes .Az y = $(M1w + M2 wz) 1/2dw (3.18a) = (-M2)'1/2cos’1(2M2M;tAz + 1). (3.18b) Inversion yields the desired equation, AZ = M1(2M2)-1(cos((-M2)1/2y) - 1). (3.19) This describes z(y) for a photon beam which enters the liquid at 20 normal to the cell face. It is now possible to write the Optical path as a function of the spatial derivatives of the refractive index and the length of the cell, a. Using the path integral expression for the Optical path (Equation 3.7). Equations 3.8, 3.12, and 3.16a, we find a W(zo) = n(zo);E0Mv éQSz)vdy. (3,20) 85 Table 3.1 Derivative estimates of Mv (Equation 3.16b) for water in a 10 deg/cm temperature gradient. v Mv de/dz dsz/dz2 1 10"3cm'1 -3x10“‘cm'2 6x10"7cm"3 2 -1O-qcm-2 --10"7cm-3 3 -1o"7cm'3 _— 4 10'8cm'“ —————- ___-— 86 The indicated integration requires the use of Equation 3.19. The integrated form is w(zo>/n(zo) = (1 + mf/(8<—M2)))a -(Mi/(16(-M2)3/2))sin(2(-M2)1/2a). (3.21) Expansian in powers of a, yields w(zo)/n(zo) = a + (Mi/12)a3 - (Mf(-M2)/6o)a5 + (Mf(—M2)2/630)a7 - (Mf(-M2)3/11340)a9 + . . . . (3.22) . m m The refined data, (d n(zo)/dzo)m=1’2. , are now mapped from (dmw(z )/dzm) by taking m derivathms with 0 O m=1,2,.. respect to zo. The result is a(dmn(zo)/dz§) = amw(zo)/az§ - iEOB§a2i+3 (3.23a) where (1/12)(dm(an)/dz§) 0335 l I {11 ll m (1/6o)(dm(anM2)/dz§) (3.23b) Bm (1/63o)(dm(anMg)/dz§) N) II C! II m (1/11340)(dm(anMg)/dz§). Additional equations are needed to associate the m) o m=1,2..‘ To flnd these unareiined data with (de(zo)/dz eQHations, let (a) Z(zo) be the image coordinate associated Witfli the cell face coordinate zo, (b) yI be the plane of fOClis within the cell, (c) zI(zo) be the image plane 87 z coordinate of the photon pencil associated with 20, and (d) note that H/l is the magnification factor of the interferometer. Then, lz(zo)/H = 21(20) - 21(0) (3.24) where 21(0) = 0 when the front face of the cell is focused. Use of Equation 3.19 in Equation 3.24 yields lZ(zo)/H = z + (M1(zo)/(2m2(zo))>(cos(<-M2(zo))1/2y1) - 1) 0 - (M1(o)/(2M2(o>))(cos((-M2(o))1/2y1) - 1). (3.25) Taking m derivatives with respect to 20, we find (l/H)(dmZ(zo)/dzm) = am,1+ (yi/4). (3.331s) T(z) is a root of Equation 3.33a (a cubic equation) and is defined to be T3.33(z). It is numerically computed from the above equations. From the refined data associated with z0 in a particular steady state experiment, Equation 3.33a is used to determine T(zo). The temperature derivatives of Equation 3.33a, evaluated at T(zo), are used in Equation 3.28 to calculate (dmn(T)/dTm)m=1’2... 3.6 Temperature Distributigg When the temperature dependence of the refractive index is known, it is possible to map the temperature distribution from the refined data. By the chain rule, dT/dz = (dn/dz)(dT/dn) d2T/dz2 = (d2n/dz2)(dT/dn) + (dn/dz)(dT2/dn2) (3.34) etc. 91 On the right hand side of these equations, the (amn(zo)/dz‘§) are known, and the (me(zo)/dnm) m=10200 m=1’2'° can be calculated from the empirical cubic equation n(T) = A + BT + 0T2 + DT3 (3.35) provided T(zo) is known. Equation 3.35 is adequate for visible wavelengths in the ordinary temperature range 20°C to 40°C. The coefficient of the cubic term is usually vanishingly small. Because T(zo) is an unknown, an iteration procedure is necessary. A suitable ith order iteration is: 1. Assume an ith order approximation to T(zo); calculate the ith order approximation to (me(zo)/dz§)m=1’2.. with Equations 3.34 and 3.35. A convenient 9th order approximation to T(zo) is (T(zo))9th = TU - (ATo/2)zo. (3.36) 2. Calculate the i+1th order approximation to T(zo) from the Taylor expansion T(zo) = §O(me(zo-€)/dzm)(ém/m!). (3.37) where E is some small distance, and TU is known. CHAPTER 4 NONISOTHERMAL, NONCONVECTING LIQUID STATES 4.1 Introduction As temperature increases, liquid density decreases. Consequently, when a horizontal parallel plate apparatus is heated from above, the liquid density increases as the lower plate is approached. This results in a liquid layer which is stable to convective motion(provided that cell end effects are unimportant). A11 liquid motion decays to stable, time-independent, nonconvecting thermodynamic states. Far from boundary or interface, measurements of liquid temperature distribution in these nonequilibrium, nonconvecting states have shown that the temperature is described by the Fourier-Laplace equation V-kVT = o. (4.1) For horizontal parallel plates in the absence of end effects, Equation 4.1 becomes (d/dz)k(dT/dz) = 0 (4.2) where the temperature distribution, T(z), is a function of the vertical coordinate, 2, which is measured from the upper plate. The parameter k of Equation 4.2 is experimentally found to be a liquid property. At ordinary temperatures 92 93 k depends to a small degree upon temperature; it is independent of temperature gradient. Highly nonlinear temperature distributions which may be found in.regions for which Equation 4.2 may not be valid(i&g,, near walls or interfaces) are described by the experimentally found temperature, T(z), minus the temperature predicted by Equation 4.2, T Equation 3.33 is the solution to 3.33’ Equation 4.2 subject to the boundary conditions which are taken to be the temperature of the upper and lower silver plates; these are TU and TL, respectively. The function T(z) - 3(z) is called the deviation from linearity T 3.3 because the temperature dependence of k is so small at ordinary temperatures that the solution of Equation 4.2 is essentially linear in the vertical coordinate. 1 called k(T) the "interior Joseph Fourier conducibility." He recognized that Equation 4.2 describes all energy flux in the absence of external effects through the Fourier heat flux relationship, Jz = -de/dz, (4.3) where Jz is the total heat flux in the z direction. We concur with Fourier but restate the external effects provision: Equation 4.2 describes T(z) in the nonflowing, steady state provided that z is far enough away from any boundary interface. The interface may be solid-liquid, solid-solid, solid-gas, liquid- gas, etc. {This statement is valid because there is no experimental evidence to the contrary. Equation 4.2 describes all 94 temperature distributions which have been observed far from boundaries. In this chapter we present quantitative, experimental temperature distributions for ethyl acetate, benzene, and carbon tetrachloride. The nonlinear temperature distributions which are found near the bounding silver walls of these pure liquids demonstrate the failure of Equation 4.2 as the liquid-solid interface is approached. The temperature distributions are determined with the Bryngdahl interferometer and mathematical analysis described in Chapter 3. Analysis of the dependence of the nonlinearity upon the temperature difference ATO = T - TL > O (4.4) U is made, and the variation of the exhibited magnitude of the nonlinearity from liquid to liquid is established. Evidence is presented which suggests that Equation 4.2 describes the temperature distribution in water at all distances from the solid-water interface but with a temperature jump or discontinuity at the interface. Such temperature discontinuity may be viewed as a degenerate nonlinearity in the sense that the nonlinearity is over a distance too small to be observed with the Bryngdahl interferometric probe. Irregular temperature distributions near liquid- solid interfaces have been reported by a number of authors. 2 In 1933 Bates found that "temperature drops" across the solid-liquid interfaces resulted in significant variations 95 in the experimentally determined thermal conductivities of water and red oil when the solid boundary was either copper or Duco lacquered copper. He reported that a surface effect exists which must be considered different from the effects of the.boundary layers associated with convecting fluids. In 1957 Longsworth3 used a nonisothermal technique and Rayleigh interferometry to study the Soret coefficients of KCl solutions. The Rayleigh cell was sandwiched between horizontal parallel p1ates(;_¢ 1 cm, ATo d 10°C). Pure water was found to exhibit highly nonlinear temperature distributions near bounding surfaces of either silver or stainless steel. Longsworth reported that the "temperature drop" in water is 2.4% to 3% of ATo with silver plates and 7% for stainless steel. He takes these temperature draps into account in the evaluation of Soret coefficients. Neglect of nonlinear temperature distribution near walls can cause systematic error in experimental thermal conductivity calculations. Recognizing this, Poltz has developed the concept of "effective thermal conductivity" for liquidsf‘"7 He defines an effective thermal conductivity,li keff’ such that km = marrow, = (gnome + Jr) (4.5a) = kc + (l/ATO)Jr (4.5h) ‘where l is the distance between the horizontal parallel ‘plates, and J2 is the total heat flux(W/cm2) in the z direction. The conductive heat flux, JC, is defined to 96 be the total heat flux, Jz, minus the radiative heat flux, Jr A11 fluxes are in the vertical, 2, direction. The parameter kc defined in Equation 4.5b is called the "thermal conductivity" by Poltz. From these definitions it is apparent that Poltz uses keff as the parameter in Equation 4.2. Thus, the validity of Equation 4.2 in describing the temperature distribution is extended to all domains whether near or far from walls. Consequently, the effective thermal conductivity depends upon (1) cell geometry and (2) the nature and magnitude of the solid-liquid interaction. Because it is not generally a property of the liquid, the effective thermal conductivity of Poltz and the interior conducibility of Fourier are not identical. Analyzing what he perceives to be the radiative heat flux, Poltz finds k = kc + (16n2aT3/3E)Y(e,t) (4.6) eff where Y(e,t) is a function such that lim Y(e,t) = 1 (4.7) £*” and where n and g are the average refractive index and average absorption coefficient(cm-1) of the liquid media, respectively. e is the emissivity of the plate surfaces, and o is the Stefan-Boltzmann constant: a = 5.66910 x 10'9 Wcm'2k‘“. (4.8) The optical density,t , is defined by r = 1g. (4.9) Taking the limit of Equation 4.6 as law , making the 97 fundamental assumption that kc is a liquid property which is independent of cell geometry, and employing Equation 4.7 we find lim k = kc + 16n20T3/3g. (4.10) l7)” eff Because the right-hand side of Equation 4.10 is the sum of two properties, the limit as l+°° of keff must be a property. Being both a liquid property and the thermal parameter of Equation 4.2, the limiting value of keff must be equivalent to Fourier's interior conducibility: 112° keff = k. (4.11) Poltz's experimental program for determination of the interior conducibility is: (1) measure keff(l) with Equation 4.5a at several cell heights and (2) extrapolate the results to l = «L This yields the interior conducibility by Equation 4.11. Using copper plates(e = 0.04), Poltz finds an increase in keff for toluene of 7% when l is increased from zero to infinity at a mean temperature of 80°C.4 When the mean temperature is 20°C, keff increases by 3.6% but the increases is 85% complete for the cell height of 3 mm. Similar increases of 4-7% are found for the weak infrared absorbing liquids: benzene, m-xylene, carbon tetrachloride, and paraffin.5’6 The effective thermal conductivity of the strong infrared absorbers n-propanol, iso-propanol, n-butanol, sec-butanol, and iso-butanol increases by less than 1% with increasing ;.7 In related work Novotny and co-workerss'9 have 98 evaluated the radiative contribution to energy transport for carbon tetrachloride using experimental, frequency dependent, absorption spectra. This seems to be the only work of its kind for a liquid. Direct interferometric observation of temperature distribution near and far from walls has been reported. In 1970 Schodel and Grigull10 used horizontal parallel plates heated from above in conjunction with a Mach- Zehnder interferometer to establish and determine temperature distribution. They found nonlinear temperature distributions for carbon tetrachloride, paraffin, and carbon disulfide, while water and methanol exhibited linear temperature distributions. Figure 4.1A reproduces their data for carbon tetrachloride. The magnitude of the temperature gradient shown in Figure 4.1A increases by 25% as the solid-liquid interface is approached from the cell center. There is also a central region(0.25 é z/l s 0.75) in which the temperature gradient is constant. The cell height is 1.5 cm; therefore, the nonlinearity lies within 0.4 cm of the liquid-silver interface(e = 0.05). Olson and Horne have used nonisothermal Bryngdahl 11 interferometry to study the refractive indices and thermal 12 of pure liquids in a horizontal parallel plate conductivities arrangement. Finding parabolic steady state fringe shapes for carbon tetrachloride, cyclohexane, and benzene,11 they suggest that the observed parabolas are caused by slight nonlinearities in the steady state temperature distribution. Further, 99 Figure 4.1 (A) Reduced temperature gradient, (dT/dz)/(dT/dz)l/2, versus z/l for carbon tetrachloride when l = 1.5 cm, e = 0.05, and ATo = 2.0000. From G. Schodel and U. Grigull.10 (B) Deviation from linearity, T - T3 33, of C014 in a cell 0.810 cm high. T = 26.40°c, TL = 23.6o°c. U From J. D. Olson.12 100 I. 2 dT/dz I. I I0 0.02 o 2. \ 8 000 *r'” ' [.— __ i I 1 1 0'020 0.2 0.4 0.6 0.8 I.0 101 they speculate that the unexpected nonlinearities are due to anomalous interactions between liquid and metal boundaries. Olson12 has computed the temperature distribution necessary to explain the parabolic fringe shape exhibited by carbon tetrachloride between silver plates(l = 0.810 cm, ATo = 2.8000C). The resulting deviation from linearity is plotted versus z/l in Figure 4.1B. Near the upper plate the temperature is a maximum of 0.0150C less than the prediction of Equation 4.2. Near the lower wall the temperature is a maximum of 0.0300C greater than the prediction of Equation 4.2. In the central region (0.2 cm s 2 $ 0.55 cm) the temperature distribution is essentially linear. Gurenkova gt all; have used a double beam diffraction interferometer to evaluate temperature distributions. They found distributions qualitatively similar to those of Figure 4.1 for toluene, hexane, and octane. The nonlinearities for water and liquid alcohols were negligibly small. 4.2 Nonlinear Temperature Distributiggg: Ethyl Acetate, Benzene, Carbon Tetrachloride All liquids studied are high purity. Methanol is "acetone free" Matheson, Coleman, and Bell A.C.S. Analyzed Reagent. Benzene is Matheson, Coleman, and Bell Spectroquality. Carbon tetrachloride is Baker Analyzed "Spectrophotometric" Reagent with 0.00 absorbance at 400 nm. Ethyl acetate is Mallinckrodt A.C.S. Analytical 102 Reagent. These organic reagents are freshly opened. Doubly distilled water with an electrical conductivity of 0.4umho/cm is degassed by boiling just prior to use. This eliminates troublesome air bubbles within the liquid cell. When the large temperature gradients needed to study nonlinear temperature distributions are applied to the pure organic liquids, strong refraction of the inter- ferometric light beam causes the beam to be refracted completely out of the optical train. This makes it impossible to count photometrically the number, N(H/2), of fringes which evolve at the image center, and the experimental procedure of Section 3.3 must be altered. All isothermal experimental aSpects remain unchanged he nonisothermal aspects are modified as follows: once the nonequilibrium, steady state temperature distribution is established, lens L1 is lowered until (1) the interferometric light beam traverses the optical train, and (2) the cell face is sharply focused. The fringe shape is measured from a photograph of the steady state image. Assuming T(H/2) is the mean of TU and TL, N(H/2) is calculated from Equation 3.6. The sum of the fringe shape and N(H/2) is the fringe number, N(Z). Table 4.1 contains the indices of refraction at 632.8 nm which are used with the methods of Section 3.6 to evaluate the temperature distribution, T(z). In computing these functions, the temperature dependent 103 maneaam new somehow auoa m o.o o.ms oamn.a mmmoooommo mamm mm neaaaem muoa x a nema.ou ones.mu ommemn.a momma mama MI emanate muoa x a sosw.ou mssw.mn admons.a saoo mama mm neamaem mnoa x a osoa.o mamm.en mmmoam.a ammo sawm mm ahead: mica x m onN.an mmsa.ou mmsmnm.a omm eonoaeeom e auoo\owOH auoo\nsoa d assess .ape a .a: w.mno he .00 aw aw B m: we czac> copsmfioo a ma zwsflcphoon: esp mH o H.¢ wands so + an + c n a "moOHUQH o>flposhmeh Ho eonounomou shapesomSoe 104 parameters A(T), B(T), and C(T) of Cauchy's formula,20 n(T,1) = A(T) + B(T)/l2 + C(T)/14, (4.12) are evaluated by the method of least squares from reported refractive indices at specific temperature and variable wavelength. A(T), B(T), and C(T) are then fitted to the best quadratic temperature equation with the least squares method. Table 4.1 lists the final results evaluated at 632.8 nm. The standard deviation of n(T,l) determined in this manner agrees with the experimental data within the reported experimental uncertainty. Table 4.2 lists thermal conductivity functions used in the calculation of T3.33. Two aspects of the data reduction must be justified: (1) neglect of the pressure dependence of refractive indices and (2) Equation 3.24. Concerning the pressure dependence, ‘1 for benzene at 24.800C, 1 (an/8P)T = 5.057 x 10'5 atm 1 atm, and 643.9 nm;14 (an/8P)T = 1.462 x 10.5 atm- for water at 23.10C, 1 atm, and 589.3 nm;21 -1 and (an/3P)T = 4.056 x 10"5 atm for methanol at 22.80C, 1 atm, and 1 589.3 nm.2 Thus a fluctuation of 0.1 atm causes a 3 refractive index fluctuation which is smaller than the experimental uncertainties reported in Table 4.1. Furthermore, since (an/3T)P = -1.071 x 10-4 o -1 C for water at 1 atm and 632.8 nm, the refractive index of water decreases by about 10-3 units during a temperature increase of 10°C, a typical temperature difference between plates; but the refractive index increases by about 5 x 10"6 units during a pressure increase of 0.1 atm. Even at these large variations of 105 ¢ maxooaavmm mm Hacked moa was 0.0 mhnmam.au mm.smoa Hoo maAmmmavmuenneaeme om ca 0.0 wao.nu mama mmmoooonmo mafimoaavmaeneoneme Ha ma o.o moo.mu mama omoo naxeomavqaasmzeuoz cos 0 ommmo.on mm.om o.omom omm oosohoHo: 00\o OO\U o a m wflswfla .86 an as a “co ao\2 an an a .Ecd H as ow.e WU . so + an + a u M OH ”%Qfl>flpo:c:oo awakens we oesowqomoc chapshomaoe N w N.J wands 106 pressure, the pressure induced refractive index variation is less than 1% of a typical temperature induced variation. These points justify the neglect of pressure variation. Equation 3.24 asserts that when the cell face is focused: Z(zo) = (H/_1_)z0 (4.13) where Z is the vertical image coordinate of a photon pencil which enters the liquid at the vertical coordinate 20; l is the cell height; and H is the isothermal image height. In general this simple relationship is incorrect. Alterations in relative lens position during the isothermal- nonisothermal experimental sequence cause changes in the magnification factor which is H/l in Equation 4.13. To prove the validity of Equation 4.13 for our experimental procedure, the nonisothermal image height predicted by Equation 4.13 is compared to the experimental, nonisothermal image height. Isothermal and nonisothermal image hieghts differ because light is refracted toward the lower temperatures, which causes a portion of the interferometric radiation to strike the lower silver plate. This portion is reflected out of the optical train, thereby reducing the image height in the nonisothermal, experimental configuration. It is easily seen from Figure 4.2 that the cross section of the interferometric beam leaving the cell-plate composite will have the effective height(l -eff) given by 1 1 (4.14) —eff = - - zend(0) 107 Figure 4.2 Light refraction in a nonisothermal experimental configuration; TU) TL. The pencil of light shown enters the liquid at z0 = 0. It leaves at coordinate zf(0) and angle af(0). This pencil travels through the glass cell at angle Bf(0), refracts in air, and travels at angle Yf(0) until direction is altered by the Bryngdahl optical train. W/m/W/fl/l \ --’ \ #2 hr —-. WWW/7, 4. ’///7/77/K/7Z 109 where zend(0) is the vertical coordinate of the light pencil, which enters the cell at 20 = 0, passing beyond the silver plate. From the definition of zend(0)’ we have the trigonometric relationship zend(0) = zf(0) + L tan flf(0) + K tan Yf(0) (4.15) where the distances L and K are defined in Figure 4.1. L and K are 0.635 cm and 2.450 cm, reSpectively. The angles af(0), Bf(0), and 7f(0) are also defined in Figure 4.2. They are related by Snell's refraction law by22 nf(0)sin af(0) = nglassSin Bf(0) = natmsin vf(o). (4.16) nf(0) is the liquid refractive index experienced by the upper most light pencil leaving the liquid. The refractive index of glass, is 1.52; the refractive index of nglass’ air, natm’ is 1.000276.23 From Equations 4.14, 4.15, and 4.16, we have 1,- leff = zf(0) + L tan sin-1((nf(0)/ng1ass(0))sin af(0)) + K tan sin-1((nf(0)/natm)sin af(0)). (4.17) To find af(0), we have tan a = dz/dy = ((n(z)/n(zo))2 - 1)1/2 (4.18) by Equation 3.12. Furthermore, combining Equations 3.16a, 3.18b, and 4.18 and evaluating the results at y = a = 5.986 cm, we have 2 af(0) = tan-1(12 MV(M1(2M2)-1(cosh(M:/2a) - 1 ))V). v=1 (4.19) 110 Equations 4.17 and 4.19 can now be used to calculate leff‘ Equation 4.13 is correct if (H/_l_)leIf = Heff equals the experimentally determined, nonisothermal, image height. Table 4.3 includes the isothermal shear parameter(2D), the interfringe distance(6), the isothermal and nonisothermal experimental image heights, and the computed effective height, H The last column is the eff“ per cent deviation of the calculated height from the experimentally determined height. In all instances, the deviation is 1% or less. Because deviations of this magnitude are within the uncertainty of the measured value of H, the validity of Equation 3.24 is established. Figures 4.3 through 4.7 are photographic prints of nonisothermal configuration images of water, ethyl acetate, benzene, carbon tetrachloride, and methanol. The fringe shapes exhibited at different ATO may be qualitatively discussed if we temporarily neglect the bending of light in a temperature gradient. Then the fringe number, N(z), is simply related to the temperature gradient by Equation 3.6: N(z) 2‘ (2Da_l_/}.H)(dn/dT)(dT/dz). (4.20) dn/dT data of Table 4.1 are linear functions of temperature in the temperature range of interest for all liquids. Thus, if dT/dz is a constant independent of z, N(z) is expected to be a linear function of z by Equation 4.20. This expectation is confirmed at all AT0 from the Figure 4.3 nonisothermal data for water. Water, 111 oa.o amn.s omn.a om¢.nm «mo.oa no.a mas.¢ mms.s mon.mm neo.na mm.o mea.a som.s mma.nm mmo.oa m mum.a asm.o eam.o o.o o m ma.o oma.s mmu.e aeo.mm mma.na am.on mom.¢ mmm.s mom.am mma.ua ha.ou wan.s a¢n.¢ emm.am uam.a an.o ems.s oom.s amm.om mmn.e n mmo.s mam.o omw.o 6.0 me no mm.ou sao.n omm.m omn.mm nas.ma ma.aa oma.n oma.m ohm.mm mme.s oa.oa mam.m mam.m >w¢.mm ona.m ma.on smn.n Ham.n mm¢.mm nmm.m a mse.n mne.o «ma.o o.o Hoo ma.ou aha.m mmm.n mmm.mm mmo.na mm.o sso.s mmo.s mas.wm mum.oa an.o umm.s oem.s coo.mm emu.n oo.o oma.a oma.s mmm.mm man.” a o oom.s amm.o oom.o o.o m o negates noises 5% .52 53a one}... so}: 33.3 mangoes Manse use Homescoa new mumposnhmm HsEHcApomw was .ae aam.a u a Q .m\AHH m n mvooa .HcOstoUfl one m deflpewbmc & .cwec owwsw Heshonvomwnon and HwanoswomH m.¢ canes 112 Figure 4.3 Nonisothermal Bryngdahl interferometric image of water at several temperature differences between upper and lower silver plates. 1 = 1.349 cm; ATO = 10.028°c, 13.047°c, and 16.69100. was .3 u [g 113 8.8.2 . ,2 8.8.9 n .2 1 lllllll‘ll 114 Figure 4.4 Nonisothermal Bryngdahl interferometric image of ethyl and lower silver plates. 1 = 1.349 cm; ATO = 2.24400, 1 acetate at several temperature differences between upper 6.261%, and 10.019°c. 1 0.73.0.3 u F4 115 ohms u [rd ousNN u F4 3138.51.41.21 .5: 116 Figure 4.5 Nonisothermal Bryngdahl interferometric image of benzene at several temperature differences between upper and lower silver plates. 1 = 1.349 cm; ATo = 2.3990C, 7.25600, 10.52900, and 15.06300. 050.2 N LLQ 117 o. wNN u FQ “sing mcmmcmm 118 Figure 4.6 Nonisothermal Bryngdahl interferometric image of carbon tetrachloride at several temperature differences between upper and lower silver plates. 1 = 1.349 cm; ATo = 2.63700, 5.976%, 7.729%, and 12.71300. 119 u. K. Na” oommx u 64 I a, o. mom u L14 u.¢m.N u L14 I ma. T A 38 ascozasfir :85 120 Figure 4.7 Nonisothermal Bryngdahl interferometric image of methanol at several temperature differences between upper and lower silver plates. I = 1.349 cm; ATo = 4.7390C, 9.297°C, 12.125°C, and 13.4880C. 121 rr< u.§.§ . .54 Egg page: 50H 122 therefore, seems to satisfy Equation 3.33; Equation 3.33 describes the temperature distribution within observable distances from the silver-water interfaces. The temperature dependence of dn/dT for ethyl acetate, benzene, carbon tetrachloride, and methanol is only 1/10th to 1/5th that of water. Because dn/dT is essentially constant for these liquids, N(z) is proportional to the temperature derivative, dT/dz. From the nonisothermal images of ethyl acetate in Figure 4.4, N(z) is nearly independent of 2 at ATo = 2.240C; therefore, dT/dz is essentially a constant. However, at ATo = 6.2600, N(z) develops curvature near both the upper and lower plates. This indicates nonlinear temperature gradients in these domains. At ATO = 6.2600 and intermediate values of the vertical coordinate, the fringe shape is seen to be a linear function of the vertical coordinate. This is the expected shape when dn/dT depends slightly upon temperature, and dT/dz is a constant. At ATo = 10.0200 very pronounced fringe curvature is seen near both upper and lower silver-ethyl acetate interfaces; the central region still displays the expected linear fringe shape. These qualitative observations indicate (1) nonlinear temperature distribution near the silver-ethyl acetate interface, (2) the magnitude of the nonlinearity increases with increasing AT and (3) there is a central region 0, in which dT/dz is constant. Qualitatively comparable results are found for benzene and carbon tetrachloride in Figures 4.5 and 4.6, 123 respectively. Only the magnitude of the nonlinearity and its extension from the metal-liquid interface vary between ethyl acetate, benzene, and carbon tetrachloride. Methanol exhibits a characteristically different fringe pattern in Figure 4.7. In the methanol nonisothermal images, the expected vertical fringe pattern has small amplitude wiggles superimposed upon it. The wiggles are time independent with amplitude which grow slightly with increasing ATO. By placing a rubber bulb over the cell fill holes and "plunging," the wiggles could be made to oscillate frantically with large amplitude at high ATO. These time dependent fringe shapes are characteristic of turbulent convection which strongly suggests that the irregular, time independent wiggles are caused by laminar convection. Observed laminar and turbulent motion experimentally proves the importance of end effects in the present apparatus, and the characteristic wiggles provide a convenient test for such liquid motion. For example, ethyl bromide is found to exhibit steady state fringe shapes identical to those of methanol, but the time dependent behavior can be induced at much lower ATO. Thus, both methanol and ethyl bromide are unsuitable for studies of temperature distribution in nonconvecting media. The characteristic fringe pattern of methanol is also observed for ethyl acetate at ATO above 1200, indicating a critical, liquid dependent ATO above which end effects become important through their production of 124 convection. To analyze quantitatively for temperature distribution, the fringe numbers are fitted in a least squares sense to spatial polynomials of degree 3 to 6. The method employed uses orthogonal Chebyshev polynomials in the discrete range.24 Results are reported in Table 4.4. The raw data are processed by methods described in Sections 3.3 and 3.4. Temperature distributions are mapped by the method of Section 3.6 with the refractive index temperature derivatives given in Table 4.1. The temperature at z = 1/2 is assumed to be given by Equation 3.33 subject to the boundary conditions T(O) = TU, T(l) = TL' T3.33(z) is computed from Equation 3.33 subject to the calculated boundary temperatures. The resulting deviation from linearity, T(z) - T3.33(z), for benzene and carbon tetrachloride are recorded in Figures 4.8 and 4.9, respectively. Certain-features of these figures are clear: (1) a negative deviation from linearity exists near the upper silver-liquid interface, (2) a positive deviation exists near the lower silver-liquid interface, (3) the deviation magnitude depends upon ATO, and (4) a central region exists in which the deviation from linearity is linear in the vertical coordinate. There are some disturbing aspects of these figures. First, because both upper and lower plates are identical, we expect a high degree of symmetry with respect to inversion through the cell center. Figures 4.8 and 4.9 have far larger negative deviations 125 0.0 0.0 0000.0 0mw.H I 000.0 00.5HI 000.0H 000.00 0H0.0H 0.0 0.0 005H.0 0000.0 I 5H0.H 00.0HI HOH.m 500.00 H00.0 0.0 0.0 0.0 0500.0 5000.0 000.0 I 000.0 000.00 000.0 ensues... H.930 0.0 0.0 000.5 H0.0HI 0H.0H 00.00I 00H.0H 005.00 0H5.0H 0.0 0.0 00.50 05.00I 00.00 00.Hml 000.0 000.00 005.5 0.0 0.0 000.0 00.0HI 00.0H H0.00I 000.5 500.00 050.0 0.0 0.0 055.H H00.0 I 00H.0 00.00I 50H.0 000.00 500.0 oUHHOHzoanoe sonhso 0.0 H5H.0I 000.0 50.0HI H0.0H 00.00I 050.0H 000.00 000.0H 00H.HI 005.0 000.0: 000.0 HHH.0 00.00I 000.0H 005.00 000.0H 0.0 0.0 0000.0 000.HI 000.0 0H.00I 000.0H 000.00 000.5 0.0 0.0 000H.0 0H00.0 I 0500.0 000.0 I 000.0 000.00 000.0 0 0 0 0 H 0 A o EO\0000H EO\ d00H EQ\ 000H EU\ 000H E0\ 000H EO\ d00H EO\ a 00\ B 00\ B4 enosnom .Eo aH 0H0 oz can 0 H onH .Eo000.0\AEo 000.0 I 00 N when: .0 N n 02 "gonad: omnwhm HashoapomHuoz .n 0.0 wmnwe 0 126 Figure 4.8 Temperature deviation from linearity, T(z) - T3.33(z), for benzene at several temperature differences between upper and lower silver plates. The refractive index temperature derivatives of Table 4.1 are used in the method of Section 3.6 to compute T(z) from the refined experimental data. I = 1.349 cm. 127 .l2 - CeHe .. G) ATO= 2.390%: .03- @ATO= 7.256°C .. ©ATO=I0529°C o .04- @ATO=I3.063°C O z .. A 5” 0 K - 0 PM I }_ -.04 -.08~ ® " 69 —.I2 ~ - @ L l l l l l 1 0 0.2 0.4 0.6 0.8 |.0 128 Figure 4.9 Temperature deviation from linearity, T(z) - T3.33(z), for carbon tetrachloride at several temperature differences between upper and lower silver plates. The refractive index temperature derivatives of Table 4.1 are used in the method of Section 3.6 to compute T(z) from the refined experimental data. 1 = 1.349 cm. T-T333 /1°C .l2 .08 129 CCI4 - CD ATO= 2.637% .. ® ATO= 5.976°C .. ® AT = 7.729°c 130 from linearity than positive; other curves are found with far larger positive deviation than negative. Second, we expect a regular increase in deviation with increasing ATO; yet, near the lower plate an irregularity is found for benzene(ATo = 7.2600). Systematic error, due to lens aberrations, is the cause of these disturbing points. Lowering lens L1 to bring the interferometric light beam back into the optical train and to refocus the cell face causes a repositioning of the light beam upon each lens. For example, in the isothermal configuration the light beam passes through the center of each lens. The need to lower L1 in the nonisothermal configuration, however, causes the light beam to traverse L1 just below center, L2 just above center, etc. This results in systematic distortion of the cell image. Correction for the distortion is made as follows. Choose two points, z and Z2, a distance m apart 1 such that 21 = (l_~ m)/2, and 22 = (l + m)/2. (4.21) Furthermore, choose the distance m such that dT(zl)/dz = dT(z2)/dz = dT(l/2)/dz, (4.22) and make the following definitions: dn(T)/dT a A + BT, (4.23) - z2 I1 = i (dn/dz)dz, and (4.24) 1 12 E f2 (d2n/dz2)dz. (4.25) Z 1 131 The integrals 11 and 12 are numerically evaluated from the refined, experimental data. Equations 4.21 to 4.25 can be combined and manipulated to show that 2 B = 12/m(dT(l/2)/dz) (4.26) where B is defined by Equation 4.23. Equation 4.26 is used to estimate B by approximately the temperature derivative as -ATo/l. Equations 4.21 to 4.25 can also be combined and manipulated to show that aA2 + bA + 0 = 0 (4.27a) where a = T(l/2)12 (4.27b) 0" fl n(dn(_l_/2)/dz)2 - 11dn(;/2)/dz + 2T(;/2)12B (4.270) (123(T(;/2))2 — Ildn(l/2)/dz)BT(l/2). (4.27s) C The final parameter of Equation 4.23, A, is the root of Equation 4.27. In this manner, the parameters A and B have been effectively adjusted so as to satisfy the condition stated in Equation 4.22; image distortions are thereby contained within these parameters. Choosing m = 0.2;, the resulting parameters are reported in Table 4.5. With the refractive index temperature derivatives of Table 4.5 and the methods of Section 3.6, the temperature distributions shown in Figure 4.10 through 4.12 are mapped. These figures are temperature deviation from linearity corrected for interferometric image distortion. They show 132 H00.0 000.0 000.H 000.0: 000.0 005.0 500.0 000.0: 000.0: 005.0 «H0.0 000.0 555.0: 505.0: 000.0 0 000.H mHH.0 000.0 00H.5: 000.H H00 000.0 H00.m 500.H: 050.0: 000.0 005.5 H00.5 000.H: 005.0: 000.5 000.0 000.0 00H.0: 000.0: 050.0 0 0 H05.H 005.H 05H.H: 000.0: 055.H m 0 0H0.5 000.5 00H.m: 000.0: 500.5 000.0 000.0 000.0: 500.0: H00.0 0 m 0 H00.H 050.H 000.0: H00.0: 000.H m 0000 00 EO\00 EO\00 EO\00 «\HAN0\000: mHem>Hh00 Mman 0>Haowhmmh 00¢ 09H: 00930300 0H0 m\HAs0\000 use 00 .aHHsoHaeeaenemeopsa sesHahepos omHs aH «\mks0\000 “A000 I M000 oesoHQHHHc ousawuonaop oosHspepmu kHHsOHHpcaouoHHowsfl one 0H e< .noHpuopch 0 00H oHHpoaouloopsH Hashospoquon no 005000: H.0 0Hnse no 005H0> auspwhopHH Bonn HeHHHc m 000 4 .90 + < u 90\:0 "00>H90>HH00 ouspmhomaop chdH 0>Hwowhnoh cmanhopoc kHpraoaHhomxm 0.0 0:39 Figure 4.10 Temperature deviation from linearity, T(z) - T3.33(z), for ethyl acetate at several temperature differences between upper and lower silver plates. These curves are corrected for image distortion by using the refractive index temperature derivatives of Table 4.5 in the method of Section 3.6 to compute T(z) from the refined experimental data. ; = 1.349 cm. 1311 CH3COOC2H5 .0 3 .. G) ATO= 2.244%: (3 ATO= 6.26l°C .0 2 _ ® ATO=IO.Ol9°C OI- 'Dl- '02P T-T3 33/1<’<: 0 e -00 3 L l J J l 135 Figure 4.11 Temperature deviation from linearity, T(z) - T3.33(z), for benzene at several temperature differences between upper and lower silver plates. These curves are corrected for image distortion by using the refractive index temperature derivatives of Table 4.5 in the method of Section 3.6 to compute T(z) from the refined experimental data. 1 = 1.349 cm. .l2 .08 136 CeHs (D ATO= 2.399% ® ATO= 7.256°C © ATO= no.529°c @ ATO= l3.063°C 0 .04 2. \ a / .3" 0 <9 / I [.— -.04 ® 0 -.08 @ -.|2 I l L l l l l O 02 0.4 0.6 0.8 |.O z/e Figure 4.12 Temperature deviation from linearity, T(z) - T3.33(z), for carbon tetrachloride at several temperature differences between upper and lower silver plates. These curves are corrected for image distortion by using the refractive index temperature derivatives of Table 4.5 in the method of Section 3.6 to compute T(z) from the refined experimental data. l = 1.349 cm. 138 .l2' .08 - T -T3.33 /1°C 0 __ CC|4 CD ATO= 2.637% © ATO= 5.976% @ ATO= 7.729% (4) ATO=12713°C d “.04 elm; -.08- ® “.I 2 " ® " . l6 " @ 1 1 1 1 1 1 1 O O 2 0.4 0.6 0.8 LO 139 (1) regular increase in negative deviation with increasing ATO near the upper plate, (2) regular increase in positive deviation with increasing ATo near the lower plate, and (3) a central region in which the deviation is a linear function of the vertical coordinate. The maximum observed deviations are -o.028°c at z = 0.243 cm and 0.025°C at l-z = 0.236 cm for ethyl acetate(AT0 = 10.019°c); -0.107°c at z = 0.216 cm and 0.079°c at $02 = 0.297 cm for benzene(AT0 = 13.063°c); -o.160°c at z = 0.233 cm and 0.12400 at l—z = 0.270 cm for carbon tetrachloride(ATO = 12.71300). Comparison of ATO and the interferometrically determined difference between the temperature of the liquid at the upper plate and the temperature of the liquid at the lower plate, AT, as reported in Table 4.5, reveals discrepancy. AT0 is the temperature difference between upper and lower silver plates as measured by thermocouples placed just outside the cell. ATo is consistently smaller than AT when Table 4.5 temperature derivatives are used to compute AT; however, ATO and AT are identical when Table 4.1 temperature derivatives are used. This latter fact is a computational artifact, just as the similarity of ATO/l and -dT(l/2)/dz(see Table 4.5) is a computational artifact in the former case. Linear least squares analysis shows AT - ATO = 1.00 x 10-2AT (4.28) O 2 0C a 0.9 x 10- for ethyl acetate; 140 AT — ATO 2.43 x 10’2ATO (4.2 ) 2 o 0:306X10- C for benzene; and for carbon tetrachloride 5.67 x 10‘2ATo (4.30) AT - ATo 2 0C a = 3.8 x 10- where a is the standard deviation. The hypothesis that these discrepancies are a true reflection of temperature control and measurement technique is tested by replacing the liquid cell with a white pine block (5.900 cm x 5.876 cm x 1.915 cm). Twenty-two nonisothermal experiments in which AT is measured by thermocouples placed at the center of the wood-silver interfaces and ATO is measured by thermocouples placed just outside the wood cell show (4.31) AT — ATO 2.37 x 10-2AT O 1.0 x 10“2 °C 0 where -4.7°04A'ros 8.1°c. ATO is smaller than AT. The discrepancy is due to the water flow in upper and lower water jackets which is effectively from a position corresponding to the metal-liquid(or wood) interface center toward the outer edges. Because heat flow is away from the upper plate, the heat bath water cools in flowing toward the outer edges. In addition, heat flow is toward the lower plate; therefore, heat bath water in the lower water jacket heats in flowing toward the outer edges. This results in a measured temperature difference, ATO, which is smaller than the temperature difference AT. 141 The reduced temperature gradient(RTG) is defined as the ratio of temperature gradient to temperature gradient at z/l = 0.5. Several important features are apparent in the RTG of Figure 4.13. First, no systematic behavior of RTG is found with increasing ATO; this ratio seems to be independent of the applied temperature gradient, ATO. Consequently, RTG's of differing AT0 are averaged for each particular liquid. A second characteristic of RTG curves is the exhibited nonlinearity near the solid- liquid interfaces. At the upper interface, the temperature gradient is 4.5% greater than (dT/dz)l/2 for ethyl acetate, 12.5% greater for benzene, and 30% greater for carbon tetrachloride. These represent extremely large increases in temperature gradient as interfaces are approached. Central regions of RTG curves demonstrate the validity of the Fourier-Laplace description of temperature far from interfaces. Centrally, the temperature gradient is constant. A final characteristic feature of RTG curves is the relative magnitude of temperature gradient increase as the lower plate is approached; the temperature gradient increases as the lower plate is approached, but the magnitude of increase is smaller than that observed as the upper plate is approached. We believe these relative differences to be due to data extrapolation problems, not to physical differences between upper and lower silver plates. At the lower plate a portion of the interferometric light beam has struck and been reflected by the plate; 142 Figure 4.13 Reduced temperature gradient, (dT/dz)/(dT/dz)l/2, for ethyl acetate, benzene, and carbon tetrachloride. Corrections have been made for interferometric image distortion. l = 1.349 cm. dT/dz (dT/dZ)f/2 L4 L3 |.2 LG 143 ® 011300002145 @ CeHe @0014 ® ..\ 0. A l l l J L l o 0.2 04 0.6 0.8 1.0 144 Figure 4.14 Integrated absolute deviation from linearity, 1 (1/1) I [T - T ldz, as a function of AT . Corrections - 0 3.33 0 have been made for interferometric image distortion. The placement of the water curve is discussed in Section 4.3. l = 1.349 cm for ethyl acetate, benzene, and carbon tetrachloride. l = 0.474 cm for water. 145 IO 9:- H20 3" 0014 '0 7- Q9 '0 “3.2 6" x c H g 51- 6 6 :3. 4" m 31—. 3- “:9 2- § . 01-13000021-15 l l I l O 2 4 6 8 Arc/1% IO l2 I4 l6 146 this portion provides no experimental data. The interval not observed is l-leff(Equation 4.17). Temperature in this interval is computed by extrapolating the observed fringe shapes of Table 4.4. This procedure underestimates the growth in N(z) as the lower plate is approached and thereby underestimates the growth in temperature gradient. The extent of the unobserved interval can be calculated from the data of Table 4.3 by taking the image height(H) at ATo = O, subtracting Heff at a particular value of AT and dividing by H/l. For benzene at ATo = 13.0630C, 0, the unobserved interval extends 0.177 cm from the lower plate. For smaller ATO, the interval is smaller. Figure 4.14 is another useful representation of the observed nonlinear temperature distributions. The curves are the integrated absolute deviation from linearity. They are functions of ATO, and the magnitude of their first derivatives orders liquids as to the magnitude of exhibited nonlinearity. Where a is the standard deviation, we find 1 (1/;) 50 IT - T3.33|dz = 1.655 x 10 3ATO (4.32) o 1.1 x 10“3 00 for ethyl acetate, N H 1 - _ -3 (1/1)‘(0 IT T3.33ld 4.059 x 10 ATO (4.33) 0 = 3.9 x 10-3 0C for benzene, and 147 7.839 x 10'3AT0 (4.34) 1 (1/1) 50 IT - T3.33ldz a = 8.8 x 10"3 00 for carbon tetrachloride. As discussed above, the temperature distribution in water is described by the Fourier-Laplace equation. Thus, neglecting the possibility of temperature jump at the silver-liquid interface, an integrated absolute deviation from linearity of zero is expected for water. Figure 4.14, however, assigns water an entirely unexpected position. This placement is discussed in Section 4.3. 4.3; Temperature Dependence of the Refractive Index of Water Interferometric determinations of the temperature derivatives of the refractive index of water seem to depend upon whether an isothermal or a nonisothermal experiment is utilized. This is illustrated in Figure 4.15 where A1 = 100((dn/dT)i - (dn/dT)TT)/(dn/dT)TT. (4.35) (dn/dT)i is the first derivative of the refractive index with reSpect to the temperature, as determined by research group i at 1 atmosphere pressure. The absolute refractive index measurements of Tilton and Taylor(TT)25 are generally considered most reliable and their thirteen parameter equation for n(A,T) is used as a standard of reference. Calculations of (dn/dT)i in Figure 4.15 are based upon the best linear least square fit of the temperature variation, 11 except for the value reported by Olson and Horne which is for a single temperature only(see Table 4.6). All .. ___—ewe-W Per by Tilton and Taylor for water at 1 atm. See Equation 4.35. I = (1) (2) (3) (4) (5) (6) (7) (8 v 148 Figure 4.15 cent deviation of dn(l,T)/dT from values reported isothermal determination, N = nonisothermal determination. I. Hawkes and Astheimer.26 l I. Eight wavelengths, Waxler gt al.14’27 I. Andreasson gt g;,28 I. Dobbins and Peck.29 N. Bryngdahl . 3O 11 N. Olson and Horne. N. Equation 4.38: Data of this work analyzed with methods of Section 3.5 and T(O) = TU’ T(l) = TL assumption. N. Equation 4.41; Implied by the ratio of Equation 4.40. 149 ///////llll[ ® 36 00 00 00.0 000.0 0.000 z 9000000 000 0o0H0 000.0 00 00 000.0 5005.0 0.000 2 00H00000000 50 00 000.0 0000.0 0.000 H 000000 000 0000000 00 00 000.0 0000.0 0.000 H 0000 00 0000000000 00.0 00 00 000.0 0000.0 0.000 H 0000 mm 00H003 m, 500.0 50 50 00H.0 0000.0 0.000 H 00000000000 000 000000 09 oo\0 oo\o 0100\9009 H100\0 as\< 00:90: @0090 nohd0m0m .0000000 00 000 000 .H u 00 0H .0 000 0:9 02 00s00HmN0 0OHAB0\Q0VI 9:090 a009aflhd> H0909 Ho G009homoha 0:9 mm 009099009nfi 00 NM .%H0>H9o0mmoh .Efi 0nd 00 :0 0M0 < 0:0 9 .590 H 90 0900 H09G0EHH0QN0 00990909 00 000%H000 m0h0300 9000H 0 0903 00903H0>0 90 0H590909HH 0:9 00 009uomoh 090 a 0nd 0 0H09080909 one .00we wo 090:3 .99 + 0 m 009Ae0\n0vl “0005905 szawah0n9omHQ0: 90 AHvHaah0s9omfi 0p 00sHah0900 00 00903 H0 x0000 0>H900HH0H 009 HO 00:00:0900 0&590A0QE09 wo “nofimm0hw0h 0.0 0H90& 151 interferometric studies which utilize isothermal states 1 - of waterlk’ 25 29 agree with TT to within 4% in the temperature range for which they are valid. The positive curvature which results in a maximum positive 4% deviation is due to the slight nonlinearity between dn/dT and T which is fully described in the values of TT. However, 11'30 both of which the two nonisothermal determinations, employ Bryngdahl interferometry, markedly differ from the results of TT. It is possible that temperature jumps in the nonisothermal, nonconvecting steady state may contribute to or be the cause of this discrepancy. The hypothesis of the existence of temperature jumps at the metal-water interface is a natural extrapolation of observed nonlinear temperature distributions(Section 4.2) in which the deviation from linearity is negative at the upper interface and positive at the lower interface. However, in the case of temperature jump, the spatial interval of nonlinearity would be smaller than that observable with the Bryngdahl interferometer(less than 0.107 cm for the present apparatus). Such physical phenomena would appear as a temperature jump: the interferometrically measured temperature difference between the liquid at the upper metal-water interface and the liquid at the lower metal- water interface, AT, would be less than the measured temperature difference between upper and lower metal plates, ATO, while the observed temperature distribution is linear. 152 To test this hypothesis we have examined water in the temperature range 24-4000 with the experimental procedure of Section 3.3. The cell height(;) and length (a) are 0.474 cm and 6.771 cm, respectively. The raw dataanx3reported in Table 4.7. The fringe number associated with the vertical coordinate l/2 times the interfringe distance,5 , has an uncertainty of $0.003 cm, and the angle of fringe from the vertica1,13, has an uncertainty of :10 min of arc. B is corrected for deviant verticality of the isothermal image by subtracting the tangent of the isothermal angle from the tangent of the nonisothermal. This equals the tangent of the corrected angle. The isothermal angle is typically 10-20 min of arc. The temperature at z/l = 0.5 is assumed to be given by Equation 3.33 subject to the boundary conditions T(O) = TU and T(l) = TL. First and second derivatives of the refractive index with respect to temperature are mapped from the refined data at z/l = 0.5 with the methods of Section 3.5. The first derivatives are then fitted by least squares to the best linear temperature equation. The second derivatives are simply averaged and standard deviation calculated. In the evaluation of the regression coefficients a and b for a function such as f(T) = a + bT (4.36) by the method of least squares, the standard error of estimate of f on T(o) is computed along with the standard error of the regression coefficient a(aa), the standard error of the regression coefficient b(ob), and the coefficient of 153 Table 4.7 Nonisothermal experimental water data. N(l/2) is the fringe number at image Center. B is the angle from vertical of fringe in deg.min and corrected for deviant verticality of the isothermal image. Angles of experiments of identical ATO and TL have been averaged. The isothermal parameters are H = 4.713 cm, 2D = 0.463 cm, and 6 = 1.239 cm with standard deviations of 0.023 cm, 0.018 cm, and 0.018 cm, respectively, for 67 isothermal experiments. l = 0.474 cm, a = 6.771 cm. Auk/00 Tl /OC N(l/2) B/deg.min 2.306 23.042 2.232 1.46 2.306 23.042 2.248 1.46 2.667 22.961 2.605 2.29 2.667 22.961 2.588 2.29 3.137 22.818 3.023 3.33 3.137 22.818 3.013 3.33 3.137 22.818 3.021 3.33 3.137 22.818 3.042 3.33 4.199 22.934 4.153 6.10 4.390 22.880 4.310 6.59 5.360 23.026 5.370 10.34 5.360 23.026 5.394 10.34 5.627 23.364 5.761 11.25 5.627 23.364 5.724 11.25 5.695 23.292 5.754 11.00 5.695 23.292 5.792 11.00 6.596 23.340 6.841 14.59 6.596 23.340 6.849 14.59 6.732 23.415 7.020 15.35 3.435 30.393 3.952 3.29 3.214 29.278 3.788 3.41 3.214 29.278 3.754 3.41 3.764 29.152 4.409 4.60 3.764 29.152 4.448 4.60 3.554 29.214 4.121 4.33 4.112 29.142 4.836 5.59 4.112 29.142 4.793 5.59 4.115 29.187 4.880 6.12 4.115 29.187 4.876 6.12 5.091 29.209 6.080 8.57 5.364 29.198 6.415 9.45 5.364 29.198 6.431 9.45 5.195 29.353 6.192 8.59 5.195 29.353 6.180 8.59 5.774 29.377 6.902 11.22 Table 4.7(cont'd.) 154 ATo/° TL/°C N(1/2) B/deg.min 5.774 29.377 6.904 11.22 6.151 29.391 7.387 12.55 6.151 29.391 7.412 12.55 4.365 31.388 5.401 6.34 4.365 31.388 5.372 6.34 4.413 31.525 5.444 6.26 4.413 51.525 5.435 6.26 4.498 31.679 5.685 7.03 4.498 31.679 5.701 7.03 5.597 31.650 7.050 10.39 5.597 31.650 7.055 10.59 4.569 32.424 5.802 7.06 4.569 32.424 5.752 7.06 5.481 32.400 6.967 10.13 5.481 32.400 6.949 10.13 5.726 32.412 7.337 10.33 5.174 50.986 6.393 9.27 5.174 30.986 6.393 9.27 5.604 31.064 6.934 10.27 5.604 31.064 6.935 10.27 5.916 31.069 7.360 11.49 7.781 31.188 9.889 20.06 7.443 31.189 9.450 18.39 6.519 31.292 8.141 14.10 5.750 32.868 7.384 11.18 6.02 32.932 7.788 12.18 6.029 32.932 7.816 12.18 6.315 32.964 8.153 13.54 6.813 33.000 8.875 16.17 6.813 33.000 8.876 16.17 5.805 34.304 7.638 10.12 5.805 34.304 7.688 10.12 5.412 34.738 7.166 9.17 5.023 35.180 6.712 8.37 4.910 35.282 6.540 8.25 4.631 35.566 6.186 7.34 4.116 36.164 5.567 5.48 4.116 36.164 5.507 5.48 4.419 36.310 5.980 6.45 4.987 36.466 6.802 8.28 5.415 36.549 7.402 9.49 determination(r2). 0, 0a, ab, and r2 are defined for m experiments by31'32 o = ((m - 2)‘1 2(11 - 1)2)1/2, (4.37a) ca = (m'1( 21‘:- - m'1(21/2 = (Di/D2)1/2N/(tans)1/2. (4.45) Because the evaluation of dn/dT via Equation 4.45 does not involve AT, results should agree wtih isothermal determinations. This explains the success of Equation 4.40. It may be that the temperature measurement technique, not a temperature jump, is responsible for the discrepancy between Equation 4.38 and 4.41; however, the "wood cell" experiments described in Section 4.2 strongly indicate that the technique is slightly compensating and not creating an apparent temperature jump. Further, the variance between Equation 4.38, Bryngdahl, and Olson and Horne is expected because the magnitude of any temperature jump will depend upon the interactions between the solid boundaries and the liquid. This interaction is a function of composition, structure, and condition of the solid surface. 163 The functional dependence of AT upon ATo can be estimated with Equation 4.42. Substituting Equation 4.38 for D1N and Equation 4.41 for dn/dT, we find T = 0.994((1 + 0.0947T)/(1 + 0.1074T))ATO (4.46) where 240C 6 T S 40°C, 2°C 5 AT0 é 8°C. Averaging this relationship over the applicable temperature range, yields AT - ATo = -9.671 x 10’2AT0. (4.47) Further, since as = (AT - ATO)(%- - Z/l). (4.48) " T133 the integrated absolute deviation is l -2 (1/1) 50 IT - T3.33ldz = 2.414 1 10 ATO. (4.49) Figure 4.14 illustrates that H20 > 001,1 > C6116 > 011300002115 (4. 50) is the observed liquid order for integrated absolute deviation. The physical literature contains several reports of unexplained water density fluctuation in the neighborhood of 35°C. Hawkes and Astheimer26’33 have encountered small, time-dependent fringe wiggle during isothermal Jamin interferometric studies of the refractive index of water. These wiggles correspond to density fluctuation of Ap/p = 6 x 10"6 and appear in the 34°C to 45°C interval. Varying the temperature during 3062 isothermal Michelson interfero- metric evaluations of the refractive index of water, 164 Dobbins and Peck29 found no evidence of any abrupt, large refractive index change. Possible irregularity is, however, indicated in their plot of mean deviation of the experimental refractive index versus temperature. The plot reveals rapid change at 34.50C. Andaloro gt $134 report a subtle transition in water between 30°C and 400C. Their study of the weakly absorbed 1.2 u combination band of water yields Arrhenius plots of integrated component intensity ratios which show neatly defined breaks occurring in the 30°C to 40°C temperature interval. Outside this interval, the experimental points are well aligned. The Bryngdahl interferometric fringe pattern of this work does not directly indicate the presence of this anomalous phenomena. Yet, the increased deviation in the ratio of (dn/dT)/(d'2n/dT2)1/2 above 35°C, shown in Figure 4.17, is further indirect evidence of its existence. 4.4 Discussion New computations, experimental observations, and experimental deficiencies have been reported. First, computed nonisothermal image heights have been confirmed experimentally, proving the validity of data analysis with respect to the experimental procedures. To our knowledge, this is the first time that the essential correctness of the mathematical analysis of nonisothermal data has been experimentally proven, not simply assumed. Characteristic fringe patterns for laminar and turbulent flows, driven by 165 end effects, have also been observed. This is important because absence of these characteristic patterns implies absence of convective liquid flow. Thus, convective flow does not contribute or cause the observed nonlinear temperature distributions. Corrections, however, are found to be necessary for nonisothermal image distortion which results from lens aberrations. Finally, the thermo- couple measurement ATO has proven to be an imperfect representation of AT. This demonstrates the need for improved temperature control and measurement. Both image distortion and temperature control problems must be considered in future nonisothermal, cell system and optical train design. After consideration of the above nonisothermal experimental problems, the temperature distribution predicted by the Fourier-Laplace temperature equation for a nonconvecting liquid is found to fail near the silver- liquid interface for ethyl acetate, benzene, and carbon tetrachloride. The observed nonlinear temperature distributions are reported in three forms: (1) temperature deviation form linearity(Figures 4.10—4.12), (2) reduced temperature gradient(RTG, Figure 4.13), and (3) integrated absolute deviation form linearity(Figure 4.14). Each form has an advantage. The temperature deviation from linearity curves directly demonstrate the magnitude of the temperature deviation from that predicted by the Fourier-Laplace equation. They are functions of both applied temperature gradient and 166 vertical coordinate. RTG curves indicate that the temperature gradient divided by the temperature gradient at cell center is independent of applied temperature gradient within experimental uncertainty. Furthermore, RTG curves reveal dramatic increases in temperature gradient as the metal-liquid interface is approached. The magnitude of the increase is liquid dependent and found to be 4.5% for ethyl acetate, 12.5% for benzene, and 30% for carbon tetrachloride when the metal is highly polished silver. Curves of integrated absolute deviation from linearity (functions of the applied temperature gradient) conveniently order liquids according to the relative magnitude of their exhibited deviation from linearity. A nonzero curve is associated with liquids which exhibit temperature jumps at the metal—liquid interface. Observed nonlinear temperature distributions are found to extend 0.2-0.3 cm from the interface for ethyl acetate, benzene, and carbon tetrachloride. Evidence strongly suggests that water exhibits a temperature jump. Knowledge of liquid nonlinear temperature distributions makes possible an increase in the accuracy of experimentally determined Soret coefficients, a = (~T/C1C2)(dCl/dz)/(dT/dz), (4.51) nonisothermally determined refractive index temperature derivatives, dn/dT = (dn/dz)/(dT/dz), (4.52) and interior conducibilities, 167 k = -JZ/(dT/dz). (4.53) Because nonlinear temperature distributions associated with interface effects have temperature gradients smaller than AT/l at cell center, use of the common gradient approximation dT(l/2)/dz fl -AT/l in Equation 4.51-4.53 yields computed values of a, dn/dT, and k which are lower than the true liquid properties. Table 4.5 shows that the linear assumption yields a 1% systematic error for ethyl acetate, a 3% error for benzene, and a 6% error for carbon tetrachloride. Figure 4.15 suggests an 8% error for water. Errors of this magnitude are much larger than typical experimental inaccuracy of 1-2% reported by independent research groups for both temperature derivatives of refractive indices and thermal conductivities of carbon tetrachloride, benzene, and water.19 The nonisothermal, Bryngdahl interferometric image also provides criteria for the establishment of liquid thermal conductivity standards. Any liquid which exhibits both linear fringe pattern and absence of a temperature jump problem similar to that observed for water will be an excellent standard. The temperature distribution in such a liquid is independent of cell size and boundary material, making the thermal conductivity measurements laboratory independent and highly reproducible. Nonlinear temperature distributions may prove very useful in testing the validity of radiation energy transport 168 models for the liquid phase. Taking the viewpoint that Fourier's interior conducibility incorporates all energy transport processes characteristic of the liquid far from walls, the time-independent, nonconvecting temperature equation becomes (d/dz)k(dT/dz) = -o (4.54) where k is the interior conducibility. 0(z) is the energy source term. It contains (1) all energy transport processes occurring within the liquid but not described by the interior conducibility and (2) all energy transport processes described by the interior conducibility but not occurring within the liquid. To conceptualize and understand, consider the averaged photon events of Figure 4.18. Photon events far from walls are depicted in Figure 4.18A. When molecule 1 emits a photon and molecule 2 absorbs that photon, the horizontal layer of molecule 2 is heated because the horizontal layer of molecule 1 has higher temperature. Likewise, the emission and absorption of a photon between molecules 3 and 4 is also a photon energy transport event. All such events are described by the interior conducibility.(0ther energy tansport processes contained within the interior conducibility include vibration-vibration, translation-translation, vibration- translation, and vibration-rotation energy transfer during molecular collisions.) Placement of a perfect mirror between molecules 1 and 3 alters the energy transport processes. The photon emitted by molecule 1 is now (A) (B) 169 Figure 4.18 Molecule 1 emits an "average" photon which is absorbed by molecule 2. This energy transport process heats the horizontal layer of molecule 2 because the photon originates within a layer of higher temperature. Placement of a perfect mirror between molecules 1 and 3 alters photon energy transport processes. The photon absorbed by molecule 2 now originates from molecule 3. Since the horizontal layer of molecule 3 is of lower temperature than the layer of molecule 1, energy flux differs from the above figure. See text. >rbei ts the TT. Vertical Direction 9 8 Mirror /’/ \ \ e ‘x \ \ \ 171 reflected and abosrbed by molecule 4. But, Fourier's heat flux law, ignorant of the mirror's presence, says that the photon absorbed by molecule 4 originates from molecule 3 whose horizontal layer is lower in temperature than that of molecule 1. The actual photon absorbed has more energy on the average. Consequently, positive deviation from linearity is found near the lower wall. Similarly, Fourier's heat flux law says that the photon absorbed by molecule 2 originates from molecule 1. This is wrong. Because of the mirror's presence, the photon has originated from molecule 3 whose horizontal layer is lower in energy. This results in negative deviation from the temperature predicted by the Fourier-Laplace temperature equation near the upper wall. By these arguments the extent of the nonlinearity from the liquid-metal interface is roughly the inverse Lambert absorption coefficient of the lowest energy vibrational mode. For water this is 7 x 10-4 cm (5 = 200 cm-l);35 for carbon tetrachloride it is 0.13 cm (5 = 320 cm-i).8 The simplest mathematical model for the energy source term: (1) assumes temperature nonlinearity is due solely to photon events, (2) treats the metal boundary as a non-emitting, non-absorbing, perfect mirror, (3) forbids multiple reflection, and (4) treats the liquid as a non- photon-scattering, isotropic media. Summing all reflection associated emission-absorption events per unit time and subtracting all emission—absorption events which due to 172 the presence of walls cannot occur yields the energy source. We find (E 6(z) = S a(z,v)dv (4.55) 0 where o(z,u)/2wa(u) = S; dz1 )0 duB(z1,v)e(a(v)/u)(z1+z) 0 -1 1 duB(zl,v)e/U> 5 +5; dz1 31 duB(zl,v)e-(a(v)/u)((l_z1)+(l-z)) 0 3. 0 -13., °d..(.1,.)./u), (.56) u = cos¢ , (4.57) a