.. € -5 A IL K 0490!. it ..I ...»r\‘b‘0» »\r ‘3‘; .3....oV"Ym mIK.v¢ofll.l3vle s I'D? . Ovvz UN \4.|.4.v00$n00. . I... . .“l‘txl.ih‘lnl.lv .9 l n x. I .. .'|&. v . . Hum . with... My? !.fcI..J km). I. h 6‘ in!“ c . .A :4 "13;..- .u. t.“ 0 It 'I l D 6an VIII 1!: (Irv 0?. a“: U .x u. u .. . v1 a’!!!i|ll.l[n|.£{dl . .IIAIIIJ‘}. .4». it!“ .II‘. I.’ (It. ‘ . .r... '9‘. v . X >1....JI;. Kl . .4 . ‘8’. .A.»O.v1t . . nu, . 1".VD..’II4'A-uu .‘ V V gilttfil $11.; t £4. ».H :I 1" \cn _!A {Info-la It If. ...... . \u. y ([3 I4 . .’.|.|.llnu \n 111‘"... .s i . ~ ~ . h vfltlu‘lnyll . o 2. {p . . . ‘ 4 94:43,... use b.” , a .s t 12), its). 3|... sl:¥.‘2§§.§~th§l , 3151. - v|1l 31V .\‘\ If, I)! 1.: 3:31:03 31% was A .l g? ah v . v ‘ . n y c. I"! 4 .lulhllvot‘. 3 Lanii‘xivftutiollt. .1. I) .0 afioll . -.l I ‘l)~ilv?1-qull\'1uti.lllr 1 .v a \)v(1vl.1\\)..1\s \\.A‘:IV‘:II\(0I£|£V" §.|.|11l1 .8 . {VS 3.; :3. $.lgll‘} 1:1 . Pt“is\|\(.v‘»uug 2’1 , . valavaviavsl. I: ~ .. ‘ . . .1 . .y .7333... ; 11.1.3.1. . , .11.! Pianr‘ . ‘32]; . \‘ulliix‘huvuruézfi . .zx.3..xzfif..:I-aif ill \. gk . . 113.313.1115}: .5 titan} 3. .313. 1!. .. 1 ‘ é? {$.13,§§1§u113§£ 1...... it‘l} « ‘ I. I it: . , .11.). f iii?! A (ails t aliquifzdizlii. ‘ u I .1.le étif if; \KIIHI D“ £‘.‘fr‘(l§.t fulfil-7... Dix .gf‘tlr’gff I‘EI‘.‘ ' { Ill‘ ££»£§v‘§~§ u u... l .) lillr villi; THESlS This is to certify that the thesis entitled Thermal Conductivity of Solid Argon presented by Juan Javier Bautista has been accepted towards fulfillment of the requirements for Ph.D. _ Physics degree 1n g Major professor A ril 4, 1980 Date p 0-7 639 THERMAL CONDUCTIVITY OF SOLID ARGON BY Juan Javier Bautista A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics 1980 ABSTRACT THERMAL CONDUCTIVITY OF SOLID ARGON BY Juan Javier Bautista An apparatus was designed and constructed to measure the thermal conductivity of rapidly cooled samples of solid argon grown from the melt. A linear heat flow method was used to obtain the in situ measurements while the samples were under their own equilibrium vapor pressure in the temperature range of 2.2 K to 83 K. The measurements obtained indicated the presence of a thermal contact resistance which was quantita- tively taken into account. The corrected data are in reasonable agreement with the constant and equilibrium volume data of previous experimen- ters. In addition, the present data are of sufficient density to compare to previous theoretical calculations. To Simona, Carlye and Geraldo ii ACKNOWLEDGMENTS I would like to express my extreme gratitude and appreciation to Professor G. L. Pollack for his unyielding support and encouragement throughout the course of this research and my graduate career at Michigan State University. I would also like to thank Professors J. Kovacs, S. D. Mahanti and T. Kaplan for their support and especially Professor J. Cowen who provided suggestions and guidance at crucial moments. I am also greatly indebted to my wife, Simona, for her patience, love and support throughout our stay at Michigan State University. I would also like to thank her for her dedication and persistence toward the typing and completion of this manuscript. I would also like to thank Rob Godbey for his assistance in preparing some of the figures. Finally, I would like to thank the Michigan State University Physics Department, the Department of Energy (formerly E. R. D. A.) and the Ford Foundation for con— tributing to the financial support of this work. TABLE OF CONTENTS Chapter Page I. INTRODUCTION....... .......... ................ l A. Discovery and General Properties of the Rare—Gas Solids...... ........... ........ l B. Thermal Conductivity of an Insulator.... 2 C. Thermal Conductivity of Solid Argon- Experimental Background ................. 11 D. Purpose. ................................ 25 II. THE EXPERIMENT ............................... 27 A. Cryogenic Apparatus ..................... 27 1. Description of Apparatus............ 27 2. Temperature Control ................. 33 3. Temperature Measurement ............. 40 B. Sample Preparation ...................... 45 1. Sample Growth ....................... 45 2. Sample Manipulation ................. 52 3. Thermal Conductance Measurements.... 57 4. Thermal Conductivity Measurements... 60 III. EXPERIMENTAL RESULTS ......................... 63 A. Results of the Force Experiments ........ 64 1. Measured Thermal Conductance versus Force ............................... 64 iv Chapter Page 2. Remarks on the Observed Effects of the Applied Force on the Measured Thermal Conductance ....... .......... 65 3. Force Experiment Errors..... ..... ... 73 B. Results of the Thermal Conductance Meas- urements........... ..... ..... ......... .. 74 1. Thermal Resistance versus Sample Length............. ................ . 75 2. Calculation of the Thermal Conduc- tivity ....... .. ...... ... ........... . 75 3. Remarks on the Thermal Conductivity Measurements ...... ...... ........... . 90 4. Thermal Conductivity Errors......... 94 IV. DISCUSSION AND CONCLUSION ................... . 99 LIST OF APPENDICES Appendix Page A. Effective Thermal Conductance at Constant Force .............. .. ....... ... ......... .... 109 B. Effective Thermal Conductance Data .......... 111 C. Defect Scattering — Classical Analogs ....... 120 D. Error Associated with the Magnitude of AT... 122 LIST OF REFERENCES .......................... 124 vi LIST OF TABLES Table Page 1. Summary of Sample Preparation............ ..... 64 2. Effective Thermal Conductance at Constant Temperature.... ....... ................ ........ 66 3. Effective Thermal Conductance at Constant Force............... ..................... ..... 109 4. Effective Thermal Conductance Data. ...... ..... 111 5. Smoothed KEff Values..... ..................... 79 6. Calculated Thermal Conductivity and Thermal Contact Conductance ....... ........ ............ 81 7. Experimental High Temperature Thermal Conduc— tivity Data.... ................. ....... ....... 88 8. Thermal Conductivity Data of Short and Long Samples ........ . ........... . ............. . ... 92 Figure l. 2. 10. LIST OF FIGURES A normal process and an Umklapp process in a one dimensional crystal of lattice constant a Typical behavior of thermal conductivity for a finite crystal with no defects and a crys- tal with dislocations ........ . ............... A cross section of the thermal conductivity apparatus .............. .... ...... . ........... A block diagram of the Artronix temperature controller ...... . ........................... . A schematic diagram of thermometer and lower Al block heater circuits... ........ ... ....... A plot of the effective thermal resistance of sample 6 as a function of the reciprocal of the applied force ............................ A plot of the effective thermal resistance of sample 7 as a function of the reciprocal of the applied force ................... . ........ A plot of the effective thermal conductance for different methods of making mechanical contact..... ....... ........ ....... . ..... ..... A plot of the effective thermal conductance versus sample length for several isotherms between 2.25 and 5.00 K ...................... A plot of the effective thermal conductance versus sample length for several isotherms between 8 and 26 K.... ....................... viii Page 29 35 43 68 69 72 76 77 Figure Page 11. A plot of the thermal conductivity versus temperature. Included for comparison are the data of Clayton and Batchelder and Krupskii and Manzhelii ........... . .......... 82 12. A plot of the effective thermal conductance versus temperature for samples of different 1engths........ ........................ . ..... 84 13. A plot of the thermal contact conductance versus temperature. .......................... 85 14. A plot of the effective thermal conductivity versus temperature for samples of different lengths.. .............. . ........ ... .......... 86 15. A plot of the high temperature thermal con— ductivity data and the three—phonon thermal conductivity calculated by Christen and Pollack... ................................... 87 16. A plot of the calculated thermal conductivity versus temperature for short and long samples 93 17. A plot of the effective thermal conductance for two samples of approximately the same length displaying different values of RC0nt" 18. A plot of the low temperature (solid curves) and high temperature (dashed curve) theoreti- cal calculations of Christen and Pollack. Included for comparison are the present data. 101 96 19. A plot of the thermal conductivity data of several workers for Ar under its own equilibrium vapor pressure....... ...... . ..... 104 20. A plot of the thermal conductivity data of Christen and Pollack from runs 8 and 10. In— cluded for comparison is a plot of the "effective thermal conductivity" of the pre- sent data, sample 12 run 1 ................... 105 ix Figure Page 21. A plot of the thermal conductivity versus temperature of the present data. Included for comparison are the data of Christen and Pollack, runs 7,8 and 10 ......... ............ 107 I. INTRODUCTION A. Discovery and General Properties of the Rare—Gas Solids Historically, the first of the rare-gases to be identi- fied was argon (Ar). In 1785 Henry Cavendish was the first to observe argon while attempting to identify the constitu- ents of atmospheric air.1 However, it wasn't until approxi- mately a century later that the "lazy one", argon, was identified as a new element by Lord Rayleigh and Sir William Ramsay. Subsequently, Ramsay within a short span of four years went on to discover the rest of the elements (He, Ne, Xe and Kr) comprising the eighth column of the periodic table with the exception of radon (Rn) which was discovered two years later by Ernest Rutherford.2 Since the introduction of quantum physics the scientific community has devoted considerable attention to the problem of understanding the properties of the rare-gas solids. Although the bulk of the work has been devoted to lattice dynamics,3 work has also been carried out over a wider range. For example, there have been recent studies of the inter- ’5 and action of gaseous xenon with biological materials,4 also recently, solid xenon has been converted from an insulator to a conductor by the application of very high 2 pressures.6 Of the solid state properties of rare-gas solids (RGS) investigated to date the relatively least understood is thermal conductivity.7 Although the rare-gas solids are easily exploited as theoretical models, they are less amenable to experimental handling. The rare-gas solids are well suited for theoretical thermal conductivity studies since they form simple close packed structures containing one atom per unit cell and are, ordinarily, electrical insulators due to their closed shell electronic structure.8 Hence, when thermal conductivity is to be investigated it is only necessary to consider the heat transported by lattice vibrational waves. The complicating effects due to the heat transport of conduction electrons present in metals and semiconductors or due to optic modes occurring in solids with more than one atom per unit cell can safely be ignored. Furthermore, the intermolecular forces are essentially central in nature and the lattice dynamics are relatively well understood.7 B. Thermal Conductivity of an Insulator The specific question that currently arises is the following: What can one learn by investigating the thermal conductivity of an insulator? The answer is that since heat is transported by atomic vibrations in insulators one can gain valuable information about their nature. The thermal conductivity for these materials is entirely determined by the way in which lattice vibrational waves (phonons) interact with one another and how they interact with defects, 3 impurities and the sample boundaries.9 For the purposes of calculation the total lattice poten— tial may be expanded in a Taylor series in terms of the displacements of the atoms from their equilibrium sites. For small oscillations the harmonic term is taken to repre- sent the ideal crystal, while the higher order terms are then treated as perturbations.lO The energy of these lattice vibrations or phonons is quantized with one phonon having energy E = hm and crystal momentum E = ha (where w is the phonon frequency and q is the phonon wave vector). The average number of phonons having this energy is given by the Planck distribution mm) = l/(exp(hw/kBT)-l) (l) where kB = Boltzmannconstant. Hence one can View a finite insulating solid as a rigid- walled box containing a phonon gas. By heating one end of the box we locally increase the number of phonons having energy hm. These will then tend to diffuse toward the cooler end and a net flow of energy will result. A useful expression for the thermal conductivity of an insulator is given by K = (l/3)CVU£ (2) where Cv is the specific heat per unit volume at constant volume, v the velocity of sound and 2 the phonon mean free path.ll It should be pointed out that although Equation (2) also gives the thermal conductivity for an ideal classical gaS, the number of phonons (or particles) does not remain constant as in the case for a classical gas. — 4 Although Equation (2) is not accurate in general, it can be used to gain valid physical insights. Since the specific heat CV is a well known function of temperature and the velocity of sound U is not strongly temperature dependent, it is then only necessary to define the mean free path 2. In general R is a very complicated function of temperature and frequency. However, in order to gain a good qualitative understanding it is sufficient to consider the mean free path characteristic of the dominant scattering mechanism. For temperatures above the Debye temperature (0), Cv is constant and the mean free path is proportional to the inverse of the temperature (T_l). Thus K is proportional to T_l. Since phonons cannot interact (scatter from one another) in the harmonic approximation this behavior is entirely explained by the anharmonic terms in the Taylor series which couples the phonons. That is, a phonon may now combine with a second to give a third or it may break up to give two phonons. Although it may be obvious that £«T_l (since lal/n and n+kBT/hm as T+W), not all of the phonon—phonon inter— actions contribute to the thermal resistivity. Only the so called Umklapp processes contribute directly to the thermal resistivity (l/K) while the normal processes only redistri— bute the phonon energies. Although normal processes do not contribute directly to the thermal resistivity, they do contribute indirectly by keeping those states occupied which scatter by Umklapp processes. 5 This can be understood by considering the conservation rules for combining two phonons to produce a third inside a crystal. These are the energy conservation fiwl+hw2 = hw3 (3) and wave vector conservation given by + —) —> ql+q2 _ q3ifi . (4) Here R is a reciprocal lattice vector of the crystal. When R = 0 in Equation (4) the scattering processes are called normal processes and when R # 0 Umklapp processes.12 To simplify the argument we will consider the inter- action of two phonons in a one dimensional solid of lattice spacing a. Figure 1(a) illustrates two phonons combining to yield a third that is within the first Brillouin zone. Normal processes do not change the direction of energy flow, hence cannot directly contribute to the thermal resistivity. In Figure 1(b) the phonons combine to yield a third that lies outside the first Brillouin zone. This phonon is equivalent to one lying within the first zone but pointing in the opposite direction. For this case there has been a reversal of energy flow, hence only Umklapp processes contribute directly to the thermal resistance. As the temperature is lowered toward the Debye temperature the number of phonons present that can partici— pate in Umklapp scattering decreases exponentially, i.e. naexp—(O/bT). Thus the thermal conductivity will increase exponentially as the temperature is lowered further.7 For a perfect infinite solid the mean free path would (a) (b) I ' ' t 1 C -K ‘ I I I < I 4' ‘ a I q ,q l q]. qz I l l 2 ————+»———r——+ I | I 0 q3 q3 ' I ‘ ’ | | <.-. ' l I I l | 1 J - n/a 0 n/a —n/a 0 n/a 93 = ql + q2 q3 = q1 + q2 - K, K = 2n/a Figure l. (a) A normal process and (b) an Umklapp process in a one dimensional crystal of lattice constant a. an(T) Figure 2. Typical behavior of thermal conductivity for (a) a finite crystal with no defects and (b) a crystal with dislocations. 7 continue to rise exponentially without bound as the tempera- ture is lowered significantly past the Debye temperature. However for a finite but otherwise perfect solid the mean free path will become constant and of the order of the dimensions of the solid. That is, phonons propagate without scattering to the surface of the solid where they are absorbed and reemitted. Thus at very low temperatures the thermal conductivity will be proportional to the specific heat, i.e. KmT3 for a perfect but finite solid. For defected solids the mean free path is determined by the type of defects and their concentration. The kinds of defects that concern us are point defects (vacancies, impurities, isotopic impurities and interstitials) and line defects (dislocations). At low temperatures the mean free path due to phonon scattering from imperfections is not intrinsically temperature dependent. It is strongly dependent on the phonon frequency w. Hence in order to establish the temperature dependence of the thermal conductivity associated with a particular type of defect a slightly more complex expression than Equation (2) is needed. It can be shown that in the Debye approximation (m=Uq) for a cubic and isotropic crystal the thermal conductivity is given by 2 339/114 -2 K(T) = kBu/Zn (kB/hu) T g x expx(expx—l) £(q(x))dx (5) where O = Debye temperature and x = hm/kBT, a dimensionless 8 variable. In the special case where only one type of defect is present in the crystal £(q(x)) can be adequately repre- sented by a power law of the form £(q(x)) = Aq'n = A(KBT/hu)‘nx'“ . (6) If we now take the low temperature limit (T<\>>Vl/3 where V = volume of the sphere, the intensity of the scattered sound wave is proportional to q4 (see Appendix C).15 Phonons of wavelengths greater than the dimensions of the defect (a few lattice spacings) will 9 then display a mean free path that is inversely proportional to q4 (since £(q)¢l/O(q)). Thus at low temperatures a solid with only point defects present will have KmT-l. The scattering of phonons from static dislocations is characterized by two main features. These are the core of the dislocation consisting of a narrow region along its axis and the surrounding strain field (whose radial exten- sion is of the order of several wavelengths). The strain of the crystal about the dislocation varies as b/r where b = Burgers vector of the dislocation and r = radial distance from the core. To incident phonons the core can be regarded as a long and narrow cylindrical obstruction with cross sectional area A = flaz. Rayleigh also showed that in the long wave- 1/2 length limit (A>>A ) the scattering cross—section per unit length of dislocation varies as a(a/)\)3 (see Appendix C).15 3 Thus 2core is proportional to q . For the surrounding strain field the Rayleigh theory is not applicable, since lgAl/Z. The effect of the surrounding strain field can be estimated by considering an analogy with geometrical optics. As a phonon passes through the strained region the phonon's velocity will be altered due to the anharmonicity in real crystals. The phonon velocity in this region is given by u = u0(ltyb/r) (8) where U0 = velocity in the unstrained region, and Y is the Grueneisen constant (a measure of the anharmonicity). To 10 first order for small scattering angles the incident phonon will be deviated from its original direction by an angle ¢~yb/p where p is approximately the closest distance that the phonon will approach the dislocation core. For a small scattering angle ¢ the scattering cross-section per unit length of dislocation will be proportional to YZbZ/pO where pO is the smallest allowed value for p. Our geometrical optics analogy breaks down unless p031. Hence, the scattering cross—section is proportional to q and thus 2 a —1 16,17 strain q ‘ The ratio of the mean free paths is thus Q . /2 strain core . . 3 2 2 . which varies as a(a/A) /(Y b /l). Since Y_l'£strain/2core thus varies as (a4/y2b2)/12=(a2/b)2/A2. In the low tempera- ture region the wavelength of a typical phonon is much greater than both the dislocation core radius a and the <2 Burgers vector b, hence £ Thus a defective < strain core’ solid dominated by dislocations will display a thermal conductivity that is proportional to T2.14 Therefore K(T) is seen to be a sensitive indicator of crystalline quality at low temperatures, while at the higher temperatures it is a sensitive indicator of the interatomic potential anharmonicity. Figure 2 contains a graph of the thermal conductivity of an insulator displaying the main features we have just discussed. 11 C. Thermal Conductivity of Solid Argon- Experimental Background Since the intermolecular forces are weak and short ranged the RGS are characterized by low melting temperatures, high vapor and sublimation pressures and a relatively large ratio of heat of fusion to heat of vaporization, it is necessary to carry out experiments at cryogenic temperatures and/or high pressures.8’54 In addition, since the proba- bility of stray nucleation is relatively high, large grained single crystals are difficult to obtain. The difficulty of obtaining a defect free sample is compounded by the small activation energies necessary for inducing various types of crystal defects and an unusually large thermal expansivity which may well be 100 times larger than that of the container it is grown in. Hence, experimenters must be content to work with plastic and easily deformed solids at low tem- peratures.8’18 The thermal conductivity for an isotropic solid is defined by the relation 3 = —K§T (9) where h is the heat flux, ET is the temperature gradient and the constant of proportionality K is the thermal conductivity. Hence in order to measure the thermal conductivity in principle it is only necessary to know the heat flux and the temperature gradient associated with it. To date six groups have measured the thermal conducti- 7,19,20,21,22,23,24,25 Vity of solid Ar. These experiments 12 can be classified as either constant volume or equilibrium volume measurements. There are of course inherent trade— offs when one choses to perform one type of experiment over another. The major disadvantage of performing constant volume measurements is that high pressures must be applied. Hence chambers containing the samples are necessarily opaque and usually have thermal conductivities comparable to the sample itself. This latter problem may be partially obviated in one of two ways depending upon the geometry employed for the measurements. First, the standard, direct linear heat flow method may be employed for constant volume measurements. In this method heat is conducted parallel to the sample walls. Now the thermal conductivity of solid Ar at high temperatures is comparable to that of glass. This is usually much smaller than the thermal conductivity of a typical metal sample chamber. Thus, this method has the disadvantage of tending to make the high temperature region very difficult to measure. A second method that may be chosen is the radial heat flow method. In this method heat is conducted radially outward toward and perpendicular to the sample chamber walls. Since the sample chamber can be chosen to have a large heat capacity and thermal conductivity, it can be used as a heat sink that defines a thermal equipotential surface.23 This method has the advantage over the first 13 that most of the heat is carried by the sample itself, hence measurements can be carried out over the entire temperature range without difficulty.7 Once the difficulties associated with the application of high pressures to RGS are overcome, the thermal conducti- vity as a function of T at constant volume can be studied as well as the thermal conductivity as a function of density at constant T.7 Furthermore, since the volume is kept con- stant, the sample is usually in good thermal contact with the heat sinks and the temperature probes throughout. The major drawback associated with equilibrium volume experiments is the unusually large thermal expansivity of solid Ar. In cooling down from its triple point (83.8 K) to liquid helium temperatures the volume of the sample of Ar will contract approximately 9%, hence the temperature probes as well as the thermal heat sinks will tend to pull away from the sample. The thermal contact problem is also aggravated by the vapor phase transport of material from the heat sinks and probes in the presence of a thermal gradient. This causes the formation of voids about the heat sinks and probes.8 Furthermore, a small yield strength near the triple point and relatively high brittleness at low temperatures make the introduction of defects particularly easy whenever large thermal gradients are introduced.26 Hence in spite of the difficulties mentioned so far, the major advantage of the equilibrium volume method is that the application of high pressures are not required so 14 that transparent, thin-walled and low thermal conductivity sample tubes can be employed. The advantages of using this type of sample tube are that most of the heat will be con- ducted by the sample throughout the entire temperature range of interest; and one can visually inspect the quality of the sample at any time during an experiment. Another advantage is that it has been demonstrated that growing Ar crystals from the melt at equilibrium volume is the best method so far for growing single crystals.27 To date all of the thermal conductivity experiments performed at equilibrium volume have employed a linear steady—state heat flow method. Of the samples examined most were grown from the melt using a Bridgman technique. That is, the samples were directionally grown solidifying from bottom to tOp as heat was extracted from the bottom of the sample tube. Although the sizes of the samples varied all were cylindrical. As was mentioned earlier, since these samples are allowed to contract the thermal boundary re- sistance at the heatLSinksq;probes, and other interfaces may not have been properly taken into account. White and Woods19 were the first to measure the thermal conductivity of solid neon, argon and xenon. Their samples were grown from the melt in thin-walled Inconel tubes of 1.3 cm diameter by 7.6 cm length. The temperature difference was measured by two He gas thermometers connected to two copper wires approximately 5 cm apart.28 These wires were in turn stuck through the tube perpendicular to the axis 15 of the sample tube and soldered in place. Although White and Woods were not able to visually examine their samples, they were able to estimate the quality of their samples from trial experiments performed using a glass sample tube. From these they estimated that the growth rates varied from 1 to 2 mm/min. For these growth rates one might expect grain sizes between 0.1 and 1 mm.26 The trial samples were initially transparent and were found to become opaque and cloudy when rapidly cooled from 77 K to 4 K. The results of their thermal conductivity measurements supported their preliminary observations. The six samples they studied displayed very low thermal conductivity values at low temperatures, indicating that they were highly de- fective due to the constraints imposed by the sample tube during cooling. It was further noted that the fifth sample yielded the lowest values even though greater care was exercised to minimize the thermal strains.19 This result may have been caused by the loss of good thermal contact. Since this sample was cooled at a slower rate than previous samples a significant amount of mass migration through the vapor phase may have occurred at the thermal probes. Berne, Boato and DePaz22 cognizant of the difficulties encountered by White and Woods19 succeeded in eliminating some of the earlier experimental difficulties. Berne at 31. grew over 50 samples of solid argon from the melt. Only 12 of these yielded data. These specimens were grown in pyrex l6 tubes (0.55 cm inner diameter by 6 cm long) at rates of l to 3 mm/hr (1/60 of the rates used by White and Woods). The average grain sizes were observed to be between 1 and 4 mm. In addition, the thermal strains induced by the con— tainer were completely removed. After the sample was grown, near the triple point temperature, it was cooled down to a uniform temperature of about 75 K. The sample was then separated from the container walls by subliming a small amount of material at the sample chamber walls. This was accomplished by pumping gently on the vapor above the sample.27 Once the sample was completely separated, it was lifted out of its container by means of a pyrex rod which was embedded in the top of the sample during crystal growth. The sample was then positioned between four copper spring clamps that were anchored to the conductivity measuring stage of the appara- tus (a heat source, two helium gas thermometers and a heat sink). To prevent the sample from subliming further and to reduce the thermal gradients during cooldown the sample was surrounded by helium exchange gas at a pressure of 200 Torr. The samples were then slowly cooled to 4 K. Although the sample was slightly reduced in size (since some sublimation nonetheless occured), the samples remained optically transparent. With the sample at 4 K the clamps were then closed, gripping the sample, and the measurements were performed. 17 Spring clamps of moderate strength were used, since it was found that clamps which were too strong would break the sample and clamps which were too weak would achieve very poor thermal contact. Since the vapor pressure of argon increases quite rapidly as the temperature is increased Berne, 3: 31. confined their measurements to low temperatures between 3 K and 15 K. Although their values were much higher than those obtained by White and Woods (indicating better quality samples), the data were not reproducible from sample to sample as well as for a single sample. This was attributed to the different ways in which the samples were grown and to thermal contact which may have varied from sample to sample. Krupskii and Manzhelii20 unlike Berne, at 31. concen- trated their efforts on making precise measurements at high temperatures. Three specimens were grown from the vapor at 70 K in glass sample tubes (1.9 cm inner diameter by 5 cm length) at rates less than 5 mm/hr and the expected grain sizes are of l to 4 mm.27 To measure the thermal gradient the sample was conveniently grown about a differential copper-constantan thermocouple. Since their work covered the temperature range of 24 K to 73 K, where the thermal conductivity is defect indepen— dent, Krupskii and Manzhelii made no attempt to remove the 20 strains induced by the sample chamber. By cooling their samples slowly at rates of 10 K/hr to the desired temperature, 18 Krupskii and Manzhelii were able to keep all but one of the samples free of any visible defects. That is, the first two samples remained transparent while the third became translu- cent after cooling. The data of Krupskii and Manzhelii were nonetheless reproducible from sample to sample. As with other workers, Krupskii and Manzhelii also experienced trouble with loss of thermal contact at the heat sink and source as the sample was cooled. This pro- blem was partially overcome by the introduction of He gas into the sample chamber to reestablish contact. Since it was established by Peterson, 33 31.27 that crystals grown from the vapor are much more defective than those grown from the melt, it is mildly surprising that at about 25 K Krupskii and Manzhelii obtained higher values of thermal conductivity than White and Woods. Krupskii and Manzhelii attributed this difference to having obtained better quality crystals than White and Woods. However, a more likely explanation is that Krupskii and Manzhelii obtained better thermal contact than White and Woods, since at 25 K the thermal conductivity is essentially defect independent. Christen and Pollack24 unlike previous workers attempted to remove both the thermal boundary resistance (at the tem- perature probes, heat source and heat sink) and the thermal strains. The samples tested were grown from a "seed" in a thin, transparent Mylar tube (1 cm diameter by 3 cm length) at the rate of 0.7 mm/hr. The seed was prepared by locally l9 maintaining the bottom the liquid filled sample tube slightly below the Ar triple point temperature until a thin wafer of solid about 0.5 mm thick appeared. It was then allowed to anneal for a period of 12 to 24 hours before the remainder of the sample was grown. It was observed that samples pre- pared in this way had grains which were 5 mm to 10 mm in size.25 The principal difference between the measurements per- formed by Christen and Pollack and those of other workers was the number and the location of the thermometers used to measure the temperature gradient. Since argon tends upon cooling to separate from the temperature probes, a single germanium thermometer was embedded in the heat source located at the bottom of the sample tube. The temperature difference between the top and the bottom was measured in the following way. First, the entire sample was allowed to reach a uniform temperature. In this case the temperature was that of the heat sink situated at the tOp of the sample. This was accomplished by electronically controlling the heat needed to keep the heat sink at a constant temperature. This initial temperature was noted and then heat was applied to the bottom of the sample via the heat source located there. Once a steady—state condition was reached, the temperature at the heat source was once again noted. Hence, assuming a thermal contact resistance was not present at the interfaces the temperature gradient along the sample is the difference between the final and initial temperatures at the 20 heat source divided by the length of the sample. To reduce thermal strains Christen and Pollack first partly detached the sample from the walls of the sample chamber in essentially the same manner employed by Berne, gt 31. Once separated the samples were cooled to liquid helium temperatures at the rate of 1 K/hr. To prevent the sample from separating from the heat source and sink the sample tube was lifted toward the heat sink during cooling. This was accomplished by the use of a metal bellows attached to the top of the sample tube. Due to the high vapor pressure of argon near 83.8 K an over- pressure of over half an atmosphere keeps the bellows ex— panded. As the temperature is lowered the vapor pressure drops quite rapidly8 and hence nearly a constant positive stress is exerted on the sample as the temperature is lowered further. In order to prevent atoms from migrating away from the interface between the heat source and the bottom of the sample during the cooling process Christen and Pollack kept the bottom of the sample slightly colder than the tOp. This was accomplished by the introduction of a small amount of helium exchange gas around the sample tube. It should be noted that the presence of this thermal gradient tends to encourage migration of atoms toward the bottom of the sample. Thus possibly reducing the contact area at the interface between the heat sink and the top of the sample. Of the 10 samples grown by Christen and Pollack only 21 three yielded low temperature data. The data were found to be reproducible for a given sample but not from sample to sample. Most of the samples were found to turn cloudy and all suffered at least some surface defects on cooling. This was attributed to the sample bridging to and subsequently separating from the walls of the sample chamber. To check for a thermal boundary resistance Christen and Pollack measured two samples of different lengths. If the thermal contact resistance (l/K ) and the thermal resis- Cont tivity (l/K) is the same for both samples, then the effective thermal resistance (l/KEff) should be a linear function of the lengths with a positive non zero y-intercept (l/K ). Cont Quantitatively expressed we have l/KEff = L/KA + l/KCont where KEff = Q/AT . (11) In Equations (10) and (11) K is the effective (or meas- Eff ured) thermal conductance, L is the length of the sample and A is its cross sectional area, K is the thermal conduc- tivity of the argon crystal, K is the thermal contact Cont conductance, O is the power supplied to the heat source and AT is the temperature difference between the heat source and sink associated with the application of O. Since their two samples yielded results that implied that 1/KCont was a negative quantity Christen and Pollack assumed that a ther- mal boundary resistance was either not present or negli- gible.25 22 In view of the great care exercised in handling the samples it is surprising the data at low temperatures in- dicated that these samples were only slightly better than 19 those tested by White and Woods. It should be pointed out that in Equation (10) K and KCont could have also changed for the two samples tested by Christen and Pollack. Ideally, the same sample of two different lengths would have been preferred. The possibility of the existence of a boundary effect should not have been so easily dismissed, since it may have been, in fact probably was, present. Clayton and Batchelder23 were the first workers to perform constant volume thermal conductivity measurements. Their work has been recognized as the first direct veri- fication of one of the oldest predictions of solid state physics. In 1914 Debye9 showed that at high temperatures the thermal conductivity at constant volume should be inversely proportional to the temperature. Clayton and Batchelder's experiments showed that this prediction is valid. For this extremely difficult experiment,23 5 samples of various molar volumes were grown from the melt at essentially constant pressure. The pressures used to grow the samples were in the 1 to 5 kilobar range. A high thermally conducting copper tube (7.0 cm long and 1.5 cm i.d.) contained within a high pressure cell served as the sample chamber. The thermal conductivity measuring part of the appa- ratus was located within the sample chamber tube concentric 23 with the tube's axis. The heat source was constructed of a long thin steel rod wound with heater wire which lay along the tube axis. Two concentric copper rings suspended with nylon string served to define two thermal equipotentials. A difference thermocouple attached to the two rings measured an average radial temperature difference. These samples were grown from the melt at constant pressure in a manner similar to that of Crawford and 30 and Leake, gt gt.3l Before a Daniels,29 Daniels, gt gt. sample was grown it was first determined what the melting temperature and pressure should be such that at a given temperature the sample experienced no external pressure. The sample chamber is filled with liquid at the temperature Tm and a pressure slightly less than Pm and left to equili- brate for several hours. Next, in the presence of a slight thermal gradient along the length of the sample chamber the pressure was gradually increased to Pm and crystallization was marked by a momentary rise in temperature. As solid continued to fill the sample chamber more argon was intro- duced to maintain a constant pressure.26 The samples were then left to anneal for 16 to 39 hours near their respective melting temperatures. Although the samples could not be visually examined, the low-temperature thermal conductivity data obtained by Clayton and Batchelder23 indicated that the samples were of the same quality as those grown by Berne, gt g1. Although Clayton and Batchelder had no direct way to test for 24 a thermal contact resistance, it is unlikely that one was present, since the external pressure on the samples is not expected to pass through zero. The nylon string used to support the concentric COpper rings aided Clayton and Batchelder in determining the strained state of their samples. That is, if the strings were found broken after performing an experiment, it was assumed that the samples had been severely strained and the data were discarded. Clayton and Batchelder found that for a given sample the data were reproducible for the entire temperature range investigated, except in a small region slightly below the melting temperature. Between the melting temperature and just 20 K below it their data fluctuated randomly. The in— stability in this region was attributed to significant re- crystallization of the sample during the measurements. It has been demonstrated that samples grown at high pressures consist of a large single crystal surrounded by many small ones26 having grains of l to 10 mm in size. This is a consequence of the high thermal conductivity of the sample chamber walls. One of the experimental difficulties encountered by Clayton and Batchelder was that near the thermal conducti- vity maximum the error in measuring the temperature differ- ence was largest. Near the maximum they found that the relatively larger amount of power required to produce a measurable temperature difference caused an appreciable 25 warming of the sample. It was later suggested by Clayton and Batchelder that this difficulty could be overcome by employing a linear heat flow method. Weston and Daniels7 subsequently pursued this suggestion and performed measurements at constant volume near the con- ductivity maximum. The temperature range covered was bet- ween 5 K and 40 K. Weston and Daniels' sample preparation techniques were similar to Clayton and Batchelder's. How- ever, the results of their work indicated that their samples were of a significantly better quality, than those studied by previous workers. D. Purpose In View of the difficulties experienced by previous experimenters with the thermal contact resistance, one of our chief interests was to determine what role, if any, the thermal contact resistance plays in the measurement of the thermal conductivity of solid Ar at equilibrium pressures. Once this was determined our next goal was to provide reliable data so that the various theoretical models cur- rently available can be tested. Natural Ar was chosen because of its value as a re- search material. Its low triple point temperature (83.8 K) allows the use of inexpensive liquid nitrogen as the cryogen during crystal growth. The interatomic potential is the best known of all the rare gases. Natural argon is rela- tively inexpensive and is available in very pure form (less then 10 ppm total impurity content) and is nearly isotOpically 26 pure. The argon content of atmospheric air is 99.6% Ar40, 36 and 0.063% Ar38.2 0.337% Ar For the present work we grew several crystalline sam- ples of Ar under controlled conditions. The measurements of the effective thermal conductance were all performed 22 gttg. We first measured the effective thermal conductance at constant temperature as a function of the force applied to the ends of the crystal at the heat source and sink in- terfaces. We then measured the effective thermal conduct~ ance for various sample lengths while we varied the tem- perature. The data obtained from these experiments will be com- pared to those of previous experimental and theoretical workers. II . THE EXPERIMENT A. Cryogenic Apparatus To perform the current experiments an apparatus was designed and constructed to measure thermal conductivity at low temperatures. The apparatus was mounted within a double dewar system to allow the use of liquid He or liquid N as the cryogen. 2 1. Description of Apparatus The design of the apparatus is similar in principle to the steady-state linear heat flow systems used by previous 19'20'22’24 In particular, it is patterned on the system employed by Christen and Pollack.24 Although the workers. present system incorporates many of the same features as theirs, the use and construction of the sample chamber is considerably different. A sketch of the cryostat is shown in Figure 3. The cylindrical sample chamber was constructed from a sheet of transparent Mylar 0.005" in thickness. A tube 3 to 4 cm long was formed by gently heating the Mylar sheet with a heat gun while it was wrapped tightly around a Teflon mandrel until it conformed to the mandrel. The seam was then secured with ordinary quick-setting epoxy glue. 27 C—HICW'nmmDnCD) I'— 7\"‘ N Figure 3. ‘ Ar Gas Line He Gas Line Brass Gears Upper Cu (Block) Heat Sink Pt Sensor Ge Sensor Ar Needle Valve Threaded Brass Rods Large Cu Bellows Vacuum Chamber Short Cu Bellows Sample Tube Cu Flange Mylar Sample Chamber Al Lower Block Heater Ge Thermometer apparatus. 28 ~ -<>Ti. Now after the Al lower block heater is turned on in the final steady state 59 the net rate of heat conducted from the lower block to the upper block is given by Q QNet = QApp- i = KEff(Tf_Tu)' that is, QApp = KEff(Tf_Tu)+Qi Substituting for Oi in the above expression we see that QApp = KEff(Tf-Ti)' that is K = K = Eff meas QApp/(Tf-Ti)' It can also be easily seen that the above result is also true for T ST.. u i To test for a possible dependence of KEff on AT we measured K for several values of AT. We observed no sys- tematic dependence on AT. In general the AT's used were 0.02 K at the lowest temperatures near 2 K, 0.05 K near the thermal conductance maxima, 4-10 K, and 0.25 K near 83 K. These values were large enough to provide a small error in the measurement of AT, yet small enough so that the fractional difference between KEff(T) and K(T), the true thermaJ.conductance,.is much less than 1% (See Appen- dix D). The latter criterion is fulfilled when ATEO.7 at the lowest temperatures and AT528 K at the highest.l4’25 To investigate what effect pressing on the sample ends would have on the values measured for the thermal conduct- ance, the short bellows was soldered in place between the sample tube flange and the Cu lifting flange (See Figure 3). To perform these measurements we started by first making 60 very light contact at the Cu-Ar interface. The thermal con- ductance was then measured as the applied force was in- creased while the temperature was kept constant. The applied force was determined by measuring the relative ex- pansion of the bellows With a Wild cathetometer. The pre- viously measured force versus bellows expansion was then used to determine the applied force. A variation of the above experiments was conducted to determine the effect on the temperature dependence of the thermal conductance for a constant applied force at two different values. For this study and the subsequent ones the short copper bellows was removed. The first set of measurements was carried out after thermal contact was established at approximately 5 K. We then maximized the thermal conductance at this temperature by squeezing on the ends of the sample until we reached the maximum force that could be applied. Measurements of KEff were carried out for temperatures between 3-20K. A The temperature was then lowered back down to 5 K where the thermal contact was first broken and then re- established. However, in this case the force applied was such that K was approximately half the thermal conduct- Eff ance of the previous experiment at 5 K and measurements were then conducted over the same temperature range. 4. Thermal Conductivity Measurements The results of the previous experiments indicated the presence of a thermal contact resistance that would have to 61 be determined in order to determine the thermal conductivty of the sample itself with the present apparatus. It was thus necessary to measure the thermal conductance as a function of temperature for samples of various lengths at the same thermal contact resistance. Of course the ideal experiment would have been to measure K for various lengths of the same sample. We did attempt to follow this route. However, vacuum leaks in the sample tube limited the number of different lengths we could examine for the same sample. In spite of the diffi- culties we were able to measure KEff for two different lengths of the same sample for at least three different samples. The procedure was to first measure K as a function Eff of temperature for the longest sample. Next, the sample's temperature was raised to slightly above 83.8 K and the top of the sample was melted until the sample was the desired length. The excess Ar was then pumped out in the usual manner and the sample was allowed to anneal at 82 K until it regained its optical clarity. We then repeated the ex- periment for this next length. The lengths of the samples were measured with the Wild cathetometer and their cross-sectional area was taken to be the same as the inner cross—sectional area of the Mylar sample tubes corrected for thermal expansion. The thermal conductivity as well as the thermal bound- ary conductance was determined by fitting data to a straight 62 line of the form given by Equation (10) using the method of least squares. III . EXPERIMENTAL RESULTS A total of sixteen samples were grown to carry out our investigations. Table 1 contains a summary of the growth conditions for the samples that gave meaningful data. The average growth rate in Table l is defined as the length of the sample divided by the total time required to grow that length. For each of the first runs the annealing period is the time interval between the end of the growth period to the actual beginning of the first measurement of thermal conductance. For subsequent runs it is the time between the end of the previous run to the beginning of the next measurement of thermal conductance. Samples 6, 7, and 10 were used to study the dependence of the thermal conductance on the force applied to the sam- ple ends. Specifically, samples 6 and 7 were used to in— vestigate the thermal conductance as a function of the applied force from 0.0 to 18.0 nt, while sample 10 was used to study the thermal conductance as a function of tempera- ture for the two forces 15 nt and 30 nt. Samples 10, ll, 12, 13, 14, 15, and 16 were used to determine the thermal conductivity of solid Ar as a function of temperature from 2.2 to 83.0 K. We should further note that sample 15 was 63 64 used to carry out extensive thermal conductivity measure- ments in the high temperature range of 48 to 83 K. Table 1 Summary of Sample Preparation Sample Run No. Length Ave. Growth Annealing No. (cm) Rate (mm/hr) Time (Days) 6 2 0.86 0.20 5 6 3 0.67 0.20 2 7 2 1.72 0.34 2 7 3 0.83 0.34 l 10 l 1.57 0.34 1/2 10 2 1.17 0.34 2 11 3 1.97 0.26 1/2 11 5 1.34 0.26 6 12 1 2.84 0.31 4 13 l 0.33 0.36 0 13 3 0.95 0.28 2 l4 1 3.00 0.25 4 l4 2 2.98 0.25 4 15 l 1.28 0.19 1 l6 1 0.63 0.25 2 l6 2 2.41 0.37 1 A. Results of the Force Experiments Measured Thermal Conductance versus Force Table 2 contains the results of samples 6 and 7. Sam- ple 6 run 4 consists of two sets of data for different but fixed temperatures one for 5.2 K and the other for 34 K. Sample 7 also consists of two sets ofcknxh that is, and 4. runs 3 These are both for two different lengths approxi- mately two to one in ratio taken at 9 K (See Table l for sample lengths). The results of sample 10 run 1 are contained in Table 3 (See Appendix A). Included in this table are two sets of 65 data for the measured thermal conductance as a function of temperature for two different applied forces. Although the short Cu bellows (that allowed us to measure the applied force) had been removed for these and subsequent experiments, we can nonetheless estimate the two forces. The applied force for the first set of data of Table 3 was approximately the maximum force we could apply 30 nt. The force for the second set of data of Table 3 was adjusted so that the thermal conductance at 5.0 K was approximately equal to one-half the thermal conductance at 5.0 K for the maximum applied force. To a first approximation the rela- tionship between the thermal conductance and the applied force can be assumed to be linear, as we shall later show, so that the second force is approximately 15 nt. 2. Remarks on the Observed Effects of the Applied Force on the Measured Thermal Conductance The data in Table 2 indicate that the measured thermal conductance increases as the applied force increases. This is expected since the area in contact at the Cu—Ar and Al-Ar interfaces increases, since the Ar will yield plas- tically at these interfaces as the force is increased. Since the yield stress of Cu and A1 is 102-103 times greater than the yield stress of Ar throughout the temperature range studied, it can be safely assumed that Ar is the only mate- rial to plastically deform during these experiments. Ber- man41 measured the yield stress of Cu at 4.2 K and found it to be 3.8 kbar, while Leonteva, gt gt.42 found the yield 66 Table 2 Effective Thermal Conductance at Constant Temperature Sample No. 6 Run No. 4 T = 5.2 K T = 34.0 K Force(nt) KEff(mw/K) Force(nt) KEff(mw/K) 0.0 0.12 i 4% 0.0 1.62 $ 6% 3.6 $ 6% 0.75 $ 4% 3.6 $ 6% 4.60 $ 6% 7.9 i 6% 1.70 i 4% 7.9 $ 6% 5.18 i 6% 17.5 i 3% 2.29 $ 4% 13.3 $ 4% 5.85 $ 6% Sample No. T = 9.0 K Run No. 3 Run No. 4 Force(nt) KEff(mw/K) Force(nt) KEff(mw/K) 0.0 0.12 $ 4% 0.0 0.43 $ 4% 0.8 i 12% 0.33 $ 4% 0.6 t 17% 3.75 $ % 1.7 $ 12% 7.56 $ 4% 1.4 t 14% 5.13 $ % 3.6 i 6% 9.89 $ 4% 2.1 $ 10% 6.79 $ 4% 5.6 $ 9% 13.07 i 4% 2.9 $ 7% 7.37 $ 4% 7.9 i 6% 13.84 $ 4% 6.2 $ 8% 11.12 $ 4% 13.1 i 4% 14.92 i 4% 7.2 $ 7% 12.10 $ 4% 7.2 $ 7% 11.25 $ 4% 67 stress of polycrystalline samples of Ar to be 14.0 bar at 4.2 K. That the Ar does indeed deform plastically at the sur- faces can be made clearer by plotting the effective thermal resistance (REff = 1/KEff) as a function of the reciprocal of the applied force (F_l) for reasons which will be de- scribed below (See Equation (15)). The plots of these variables calculated from Table 2 are displayed on Figures 6 and 7. Although the data are sparse for sample 6 run 4, the 4 sets of data on Figures 6 and 7 clearly fit a straight line. This behavior can be understood by noting that REff consists of the thermal resistance of the crystal (R = L/KA) in series with the thermal contact resistance (RCont = l/KCont) at the two interfaces (Al-Ar and Cu-Ar). That is, REff = R + RCOnt (14) but REff = R + CF_1 (15) where C is just some proportionality constant. Since on a microscopic level the ends of the sample are quite rough, we expect that the contact area will increase since the surface asperities deform readily while the force is increased. We can thus conclude that the thermal contact conductance at the interfaces is proportional to the applied force, that is,K F. 0: Cont 68 1.4- c5 c3 t3 FL 01 00 C: no O .2 EFFECTIVE THERMAL RESISTANCE R5,,(K/Mw) 0.2 Figure 6. SAMPLE N0. 6 ' RUN NO. 4 IS.2 K A3“) K 1 1 1 0.1 0.2 0.3 INVERSE FORCE F'1(NT)'1 A plot of the effective thermal resistance of sample 6 as a function of the reciprocal of the applied force. EFFECTIVE THERMAL RESISTANCE R5,,(K/Mw) 69 .20_ SAMPLE N0. 7 A RUN N0. 3 O RUN N0. 4 T=9.0 K .15 0,0 1 1 1 1 1 1 1 0.0 0.2 0.4 0.6 INVERSE FORCE F'1(N'I’)'1 Figure 7. A plot of the effective thermal resistance of sample 7 as a function of the reciprocal of the applied force. 70 This result is in agreement with the experiments of 41 43 44 Berman, Berman and Mate, and Mate who measured the thermal contact conductance between pressed surfaces of various materials (Cu—Cu, Cu-diamond, Au-Au, sapphire- sapphire) as a function of the applied force at liquid He temperatures and observed an almost linear dependence. An approximate relationship between K and F can be Cont obtained by considering the following simplified picture of the surfaces. Assume that the average sized asperity is contained within a cube of side d, then the total boundary conductance of n asperities in contact (which can be con- sidered to be in parallel) is given by KCont = an (16) where K is the thermal conductivity of the material of the individual asperities. The total surface area A in contact is determined by yield stress PC of the material and the force applied F. That is, as force is applied the asperi- ties will continue to deform plastically until the ratio F/A = Pc the yield stress. That is A = F/PC = nd2 so that d = (F/nPC)% . (17) Substituting this result in Equation (16) yields that KCont = (nF/Pc)%K. (18) Tabor has shown that for two steel surfaces in contact nng, hence KContaF3/4’ which is in approximate agreement with our results. Tabor has also shown that K FO‘6 CI Cont for deformations that take place elastically.45 71 In Figure 8 we show the data plotted from Table 3 (open circles, filled circles and filled triangles) including the data from sample 10 run 2 (filled squares). The two sets of data from Table 3 (open and filled circles) diSplay essen- tially the same qualitative behavior, except that the maxi- mum of KEff(T) has been shifted. The shift of the peak toward a higher temperature for the smaller applied force (filled circles) is probably due to the increased dominance of RCont over R. Curiously, below the peak the two lower sets of data (Open and filled circles) in Figure 8 have a definite T2 dependence, while the data of sample 10 run 2 (filled squares) have a temperature dependence somewhat less than T2. It should be recalled that for sample 10 run 2 we attempted to maintain thermal contact with the crystal ends during the cooldown process. Although better thermal contact is achieved as is evidenced by the almost two-fold increase in the effective thermal conductance near the peak for run 2 (filled squares) over run 1 (open circles), it is unfortunately accomplished at the expense of the quality of the sample. Although some deformation of the entire sample is expected, severe deformation is probably confined to the sample ends. However this may be an irrelevant point since the samples are thermally strained 1% or more during cooldown. Even a strain this small is large enough to induce plastic deformation above 65 K.42 TOO tn c: I I I II Ln I EFFECTIVE THERMAL CONDUCTANCE K5,,(Mw/K) E3 II .I I I I I I I I 000 o I o 0 0C I o f 0 Q 93 04 o - O ‘ o a... o 0 o . O I A o o '0 o n . O ICONTACT ACHIEVED AT 80 K ‘INCREASING FORCE CMAXIMUM FORCE AT 5 K OHALF-MAXIMUM FORCE AT 5 K 1 I I I I I | I I I I I 1 Jill I I I l l 50 TEMPERATURE T(K) Figure 8. A plot of the effective thermal conductance for different methods of making mechanical contact. 100 73 3. Force Experiment Errors The error in measuring the applied force F is due principally to the hysteresis associated with the non- elastic behavior of the small Cu bellows when large forces are applied. In a separate experiment this bellows was calibrated at 77 K by pressurizing its interior with He gas. The applied force F was determined from the applied over pressure and the cross-sectional area of the bellows (1.13 cm2). Its extension was then measured as the force (pressure) was increased from zero to a maximum of 21 nt (1390 Torr) and then decreased to zero. Since these meas- urements yielded two sets of curves, the calibration curve was taken to be the mean of these two measured curves. Thus the maximum error in measuring the applied force F due to the inelastic behavior of the bellows was determined to be $0.1 to $0.2 nt for applied forces between 0.0 and 5.0 mt and $0.5 nt for forces between 5.0 nt and above. The percentage error varied from 17% for the smallest applied force to 3% for the largest applied force. The error in measuring the absolute temperature T is due principally to the limits on our ability to keep the sample temperature constant (See Chapter II, Section 2). We estimate that the maximum error in measuring T (ST) is then $0.001 K for 2 KST<10 K, $0.005 K for 10 KST<4O K, and $0.010 K for 40 KETE83 K. Since we were able to measure O to better than 0.1% the largest source of error introduced in determining KEff 74 is in measuring AT. The maximum error that AT will in- troduce is twice the maximum error in T (26T). The AT's used varied from 0.05 K near 4 K to 0.25K near 80 K. We estimate from these values that the maximum percent error in KEff varies from 4% at 4 K to 8% at 80 K. B. Results of the Thermal Conductance Measurements Table 4 in Appendix B contains all the measured thermal conductance values and their corresponding temperatures for samples 10-16. These samples were used to determine K. The data presented in this table satisfy two experimental criteria. First, the sample tube remained reasonably leak tight during the experimental run. That is, the sample did not gradually leak out of the sample tube into the surrounding vacuum region over a 24 hour period. A gross leak of this size could be easily verified by the inability of the vacuum pumping system to reduce the vacuum chamber pressure below 10 mTorr. Second, mechanical contact with the sample was main— tained throughout the cooldown process. That is, it should have been possible to continuously remove the slack on the lifting screw due to the contraction and sublimation of the sample. If the sample tube and the Cu piston remained frozen together for more than 15 minutes, we assumed that the mechanical contact had been lost. Any data that did not meet these two criteria were discarded. 75 1. Thermal Resistance versus Sample Length As suggested earlier, in the Introduction, the simplest model for the effective thermal resistance is two thermal resistances in series. One of these is due to the imperfect 1 ont' The contact achieved at the sample ends RCont = K; other thermal resistance is due to the presence of the Ar sample which may be written, by definition, R = L/KA (L = length of the sample and A = cross—sectional area of the sample). Recalling Equation (10), this is expressed as 1/KEff = L/KA + l/K (10) Cont ‘ In order to determine whether the above relation is reasonably represented by our data, we must know KEff for each L at exactly the same temperature for the temperature range investigated. Since it is not feasible to measure KEff at exactly the same temperature for each L, we drew the smoothest curve through the raw K (T) data for each Eff length to interpolate and extrapolate for KEff at conven- ient temperatures. The values of KEff(L) for these temper- atures were determined graphically and are contained in Table 5 along with the corresponding temperature. 2. Calculation of the Thermal Conductivity Figures 9 and 10 are plots of REff as a function of L (length) at constant temperature for various isotherms from 2.25 K to 26 K. It is apparent from Figure 9 that be— low 3.5 K the longer samples show a greater scatter than the shorter samples. However as the temperature increases above 3.5 K the scatter diminishes and the data approach EFFECTIVE THERMAL RESISTANCE Raff(Mw/K) 76 .20 — o 2.25K O _ 0 2.50K _ 0 3.00K o — I 3.50K _ 4A 5.00K D o D .15 __ D I c o _ o D 3 o .10 __ O C I o o D I — C] D O I D D . I ; — o o o . , I A i I I ‘ | I “ I .05- ' I A I — A | A A A A i ” | .00 l I I I L I 0.0 1.0 2.0 3.0 LENGTH L(CM) Figure 9. A plot of the effective thermal conductance versus sample length for several isotherms between 2.25 and 5.00 K. EFFECTIVE THERMAL RESISTANCE REff(K/Mw) 77 .20 __ A A1 26K P II 22K I 0 18K L C1 12K I r CD 8K I- A . .15 __ L o | o I- I I o I— A .1oI_ I A A . L I _ D I ' D _ O D _. O .A. . D .05.__ I I D P D ; A. ° 0 1 . a 1. “ I I. ,OOI I I I I 1 I 0.0 1.0 2.0 3.0 LENGTH L(CM) Figure 10. A plot of the effective thermal conductance versus sample length for several isotherms between 8 and 26 K. 78 very nice straight lines. Since the samples were grown and manipulated in essen- tially the same manner, we assume that K (thermal conducti- vity) and K (thermal contact conductance) are the same Cont for each sample at the same force. Thus, in order to deter- mine these two quantities we fit the data of Table 5 using the linear regression method to a straight line of the form REff = mL + RCont (19) where K - (mA) and KCont — RCont . calculations along with the correlation coefficients are The results of these contained in Table 6. A plot of the thermal conductivity K(T) is contained in Figure 11 along with the data of Clayton and Batchelder23 and Krupskii and Manzheliizo’21 for comparison. Above 65 K our plotted thermal conductivity values in Figure 11 are those obtained from Table 5 for sample 15 run 1. Before we present the high temperature data in row form, we mention three interesting effects displayed by the raw data contained in Table 4. The three effects we are considering are as follows: (a) the dominance of RCont(T) for the shorter samples (L S 1.34 cm) at low temperatures (T E 6 K), (b) the broadening and shifting of the maximum (T) and (c) the of (T) due again to presence of K KEff Cont dominance of K at the higher temperature (T Z 38 K). It is clear by comparing columns 2-5 in Table 5 with column 3 in Table 6, for temperatures below 6 K, that the KEff is very nearly equal to the thermal contact conductance. 79 on.m om.m ov.w 00.x N.HH m.mH m.ma v.mm N.mm oo.mm om.m ma.m v.5 H.m h.NH o.¢a o.mH m.vm o.mm oo.om om.h om.h n.m o.oa m.¢a m.ma m.na v.hm o.mm oo.ma oa.m om.m H.OH m.mH m.ma m.mH 0.0m m.om N.H¢ oo.oa m.oa ~.oa o.mH o.mH N.om o.mm e.mm o.vm v.66 oo.va m.ma m.ma o.va m.na m.vm «.mm m.>m H.mm m.ov oo.ma b.¢a m.¢a m.mH 0.0m m.mm m.>m n.mm m.mm m.>v oo.HH n.mH m.oH m.ha N.mm m.mm m.mm o.mm w.ov 0.56 oo.oa 5.5H m.ha m.ma e.mm 0.0m v.om «.mm 6.Hv o.>v om.m m.ma m.ma m.ma v.vm m.Hm o.Hm m.vm o.m¢ m.ov oo.m o.om m.ma m.om m.mm m.mm m.mm m.mm N.m¢ m.wv om.m o.HN N.HN >.HN m.mm m.mm o.mm o.om o.mv o.mv oo.m v.am o.mm m.mm H.nm m.vm ~.mm m.mm o.H¢ m.mv om.n m.Hm v.mm v.mm 6.5N m.vm o.mm m.mm «.mm S.H¢ oo.w m.om m.mm m.HN 0.5m m.mm m.mm o.vm m.>m w.mm om.o m.ma h.am m.am m.mm o.mm o.Hm m.mm o.mm o.mm oo.m m.ma m.om 0.0m N.mm 0.0m m.mm m.om «.mm m.mm om.m o.wa «.ma m.nH m.mm m.hm e.mm w.nm o.mm 0.0m oo.m h.ma 0.0H m.mH h.ma o.vm H.mm o.vm v.mm N.mN om.v >.HH w.ma w.mH m.ma b.om H.ma h.om m.am m.mm oo.v om.m v.HH v.oa o.vH m.ha o.mH m.>a v.mH m.ma om.m oo.m ov.m om.m v.HH m.¢H +m.ma m.va o.mH «.ma oo.m om.h cv.m om.> +H.0H o.ma +m.HH o.ma m.ma n.ma mn.m om.m ov.h om.m +m.m +m.HH +m.oa m.HH m.HH H.ma om.m on.m ov.m +mm.m +m.h +0.0H +N.m o.oH +v.oa +m.oa mN.N I s ~ ~ I s s s s m mmmm m mmmm n mmmx m mwmm m mwmx v wwmx m mmmm m mmmm H wmmm Amve mmsam> AM\3EvmmmM ConuOOEm m mHQmB 80 mm.H 0H mh.a vm.a NH.N 0¢.N Hm.m ¢H.m mm.m mm.v u 0H mmmm wm.m m 00.N Hbm.m +0N.m +00.m +Nv.v 00.m m wmmm «m.m m mo.m mN.N Hm.m mh.m 00.m mm.m mm.m ma.v mm.v \ m mwmx HV.N h mm.m mm.m mh.m 00.m mm.m mw.m 00.v 0m.v 00.m 00.m 0 mmmx >0.H 0 0N.m H00.m +0m.0 00.0 s w wmmm wm.a m mm.m ow.m NN.m wo.m ma.v 00.0 m0.m mm.m om.m 0m.0 00.0 0m.h 0N.m 00.0 00.0 m mmmm A©.pcoov 0H.H v 00.0 +0m.0 v.0H m.HH w mmmm m magma mm.0 m vu.m HmH.0 +mm.0 +0H.n 00.0 0m.m 00.0 00.0 0.0H 0.HH m mmmm m0.0 N m0.n mm.m N0.m mv.0H I0.HH mm.0a m.aa m.NH 0.0H m.ma m.0H N.mH 0.0a N mmmx mm.o H ba.a Ho.HH mm.NH ma.ma m.mH 1m.mH Mm.aa ma.ma 0.0m m.am 6.mm m.am m.mm o.mm 6.am H mmmm mmsHm> pmumHommnuxm + 15004 “C s C mwmx 00.00 00.mm 00.00 00.m0 00.00 00.mm 00.0m 00.0w 00.Nv 00.00 00.mm 00.0m 00.vm 00.Nm 00.0m 00.0w 00.0N 00.0w Axes 81 Table 6 Calculated Thermal Conductivity and Thermal Contact Conductance T(K) K(mw/ch) KCOnt(mw/K) Correlation KCont/K(Cm) CoeffiCient 2.25 33.4 $ 22% 13.9 $ 20% 0.900 0.42 2.50 38.9 $ 23% 15.9 $ 21% 0.912 0.41 2.75 44.4 $ 23% 18.0 $ 20% 0.922 0.41 3.00 49.7 $ 21% 20.2 $ 18% 0.931 0.41 3.50 61.9 $ 20% 24.4 $ 17% 0.941 0.39 4.00 75.6 $ 17% 28.8 $ 14% 0.948 0.38 4.50 90.8 $ 17% 33.1 $ 14% 0.951 0.36 5.00 107.9 $ 17% 37.5 $ 13% 0.954 0.35 5.50 122.1 $ 16% 41.3 $ 11% 0.966 0.34 6.00 129.4 $ 12% 44.8 $ 9% 0.977 0.35 6.50 128.9 $ 12% 48.1 $ 10% 0.977 0.37 7.00 126.3 $ 12% 51.0 $ 10% 0.984 0.40 7.50 119.8 $ 12% 53.7 $ 11% 0.988 0.45 8.00 110.6 $ 11% 56.3 $ 12% 0.989 0.51 8.50 99.2 $ 10% 59.1 $ 12% 0.992 0.60 9.00 89.0 $ 9% 61.5 $ 13% 0.993 0.69 9.50 80.8 $ 8% 63.1 $ 13% 0.994 0.78 10.00 73.1 $ 7% 64.8 $ 13% 0.993 0.89 11.00 61.0 $ 6% 70.3 $ 14% 0.995 1.15 12.00 52.3 $ 5% 73.8 $ 15% 0.995 1.41 14.00 40.6 i 4% 78.7 $ 16% 0.994 1.94 16.00 33.2 $ 4% 76.3 $ 16% 0.994 2.30 18.00 27.4 $ 4% 79.1 $ 22% 0.995 2.88 20.00 23.0 $ 4% 82.1 $ 22% 0.995 3.58 22.00 19.7 $ 3% 84.8 $ 23% 0.994 4.31 24.00 17.1 $ 3% 81.0 $ 22% 0.994 4.73 26.00 15.1 $ 3% 96.0 $ 26% 0.994 6.40 28.00 12.7 $ 3% 211.3 $ 57% 0.997 16.69 30.00 11.3 $ 3% 353.3 $ 96% 0.997 31.21 32.00 10.2 $ 2% 221.6 $ 48% 0.999 21.68 34.00 9.27 $ 2% 356.9 f 77% 0.999 38.50 36.00 8.39 $ 2% 1504.7 >$ 100% 0.998 179.30 38.00 7.58 $ 2% 666.6 >$ 100% 0.998 87.942 40.00 6.86 f 2% -1l4.0 >$ 100% 0.997 >102 42.00 6.65 $ 2% -236.5 >$ 100% 0.996 >102 46.00 6.38 $ 2% ~469.7 >$ 100% 0.996 >102 50.00 5.65 $ 2% -275.3 >$ 100% 0.997 >102 55.00 4.67 $ 2% +276.8 >$ 100% 0.992 >102 60.00 4.33 $ 2% 274.2 >$ 100% 0.997 >102 65.00 3.75 $ 2% 1213.9 >$ 100% 0.992 >10 82 200 A A A a°°°a A n o 100.. A A o _. £5 0 o C A ° :0 501 A a A 0 AN 0 1A _. 0 0 g A a P D "‘ D \ 0 £1 C; D X __ £5 0 E: A‘ D > s in §10__ '0 c.) _ m :s; -— “3 E — Chev FE _' II ‘7 St A CLAYTON a BATCHELDER DATA g V _ v SMOOTHED CLAYTON a BATCHELDER DATA '3 vV _ . SMOOTHED KRUPSKII a MANZHELII DATA ‘8 v o PRESENT DATA in O l I 11L11111 I 11111111 1 5 10 50 100 TEMPERATURE T(K) Figure 11. A plot of the thermal conductivity versus temperature. Included for comparison are the data of Clayton and Batchelder and Krupskii and Manzhelii. 83 For example, at 2.25 K,(RCOnt/REff)(100%) varies from 77% to 42% for the shortest to the longest sample. To illus- trate this effect more clearly we show in Figure 12 plots (T) of the effective (or measured) thermal conductance KEff for samples of various lengths. Below 6 K the data converge to KCont as the samples become smaller in length. Figure 12 also illustrates the second feature, that ff lS broadened and shifted toward higher temperatures. We expect is, as the samples become shorter the maximum of KE that as L+O, K That this is approximately correct Eff+KCont' can be seen by comparing Figure 12 with Figure 13 which con- tains a plot of K (T) taken from the data of Table 6. Cont Between 2 and 20 K it can be seen that KEff(T) does indeed approach KCont(T) as L becomes shorter. (Above 20 K the behavior of KCont(T) is slightly more complicated. We shall discuss this later.) The third feature is that above 38 K the conductance of the sample is now dominant. For example at 38 K (RCont/REff) (100%) varies from 2.8% to 0.3% from the smallest to the largest sample. To further illustrate this effect in Figure 14 we plot KEff(L/A), the effective thermal conductivity, as a function of temperature for several sample lengths. In view of this last effect we can safely assume that above 38 K the measurements we have carried out have yielded K(T) directly. In Figure 15 we plot the raw thermal con- ductivity data (corrected for KSp the spurious heat conduc— tance due to the mylar tube, wire leads, etc.) of Table 7. 84 100 L. I— 50.. I.‘. . _. moo .0 :2 EB. '0. U \\ _. are . g .000000 V C : Ido AA‘ C 01 D I “ A o :4 P— l. o ‘ . g ..00. A - E I C. O L) o. :3 Q . OA €3.10L_ o L) __ 4 . o __J A < )— E A m )— LLJ A g; .— LU 5.. E: I 0.33 CM *— tfl 7' D 0.63 CM ft uJ __ 0 0.95 CM 0 1.97 CM 1. ‘ 2.84 CM 1L 1 1 1 11111 1 11 1 S 10 50 TEMPERATURE T(K) Figure 12. A plot of the effective thermal conductance versus temperature for samples of different lengths. 85 1000 1.. F r— 500% /\ __ .\_4 E r F— V w c O U __ :2 uJ L) 2: < p.— L) Fa” z 100_ C) U )- .._ I—— L) < r— P— 2: __ 8 , __, :0. <2 E 33 F— I—- IIIIII I I IIIIII l 5 10 50 100 TEMPERATURE T(K) Figure 13. A plot of the thermal contact conductance versus temperature. ...... C) I— 100 /CM K) KMW I K: ,, x LENGTH/AREA U1 h 50 p._1 CD 86 L. _. - II 1... I I I- . _\41CO(‘_ I -c * __ J L I . C A“A CC I . Q ‘A‘ AA e— cC A‘ 0 ‘ ' . C A C] DOD. o ‘ D C) C] . I C ‘ D A C D D - I— A I ‘ 00 ..° 00/ A 0 ,°'° . . I A D 0 . 0 ° ° DC ’ U ... o .0 A D o ._ ‘ [P . ' 9:- y—_ C] O I I— ° '7 o ‘0 F— 0 ° +— 0 $7. I- . I +— . | I 2.98 CM I I. 0 1.34 CM ' I . 0.95 CM I L— l 1 D 0.63 CM . i - 0.33 CM 1 I 1 I I I I I II 1 I 1 : I l 5 10 SO TEMPERATURE T(K) Figure 14. A plot of the effective thermal conductivity versus temperature for samples of different lengths. THERMAL CONDUCTIVITY K(MN/CM K) 87 15.0%. 10.0L 8.0u 6.0 _ A.0_ A 0.33 CM 0 0.63 CM . 1.28 cm _ A 1.34 CM 0 SMOOTHED DATA CORRECTED FOR THERMAL BOUNDARY ° CONDUCTANCE 2.0—- ——— THREE PHONON THEORETICAL THERMAL CONDUCTIVITY (CHRISTEN a POLLACK) 1.5 l 1 l l L a a L 15 20 A0 60 80 100 TEMPERATURE T(K) Figure 15. A plot of the high temperature thermal conductivity data and the three-phonon thermal conductivity calculated by Christen and Pollack. 88 Table 7 Experimental High Temperature Thermal Conductivity Data T(K) KEff(mw/K) Ksp(mw/K) K(mw/cm K) A. Sample No. 11 Run No. 5 37.83 5.51 0.05 8.38 45.40 4.28 0.07 6.47 63.45 2.63 0.10 3.89 B. Sample No. 13 Run No. l 39.51 19.26 0.05 7.31 49.53 16.43 0.07 6.23 60.87 10.49 0.09 3.96 C. Sample No. 15 Run No. 1 83.08 1.82 0.12 2.51 81.70 1.77 0.12 2.44 79.90 1.90 0.12 2.63 78.27 2.00 0.12 2.78 76.00 1.97 0.11 2.74 72.85 2.14 0.11 3.00 68.72 2.37 0.10 3.35 65.75 2.55 0.10 3.62 62.92 2.86 0.10 4.08 59.65 3.04 0.09 4.35 55.93 3.29 0.09 4.73 52.85 3.83 0.08 5.54 51.34 3.92 0.08 5.68 48.29 4.36 0.07 6.33 45.99 4.61 0.07 6.71 52.50 3.91 0.08 5.66 89 Table 7 (cont'd) T(K) KEff(mw/K) Ksp(mw/K) K(mw/cm K) D. Sample No. 17 Run No. 1 41.38 11.11 0.06 8.03 42.25 10.07 0.06 7.28 44.61 9.34 0.06 6.74 47.61 7.87 0.07 5.60 50.39 8.25 0.07 5.94 53.81 6.72 0.08 4.82 E. Sample No. Run No. 39.31 2.64 0.05 7.19 9O Along with these data we also present for comparison the theoretical three phonon K(T) calculation of Christen and Pollack.24 We should note that between 30 and 80 K the experimental data and the theory are in good agreement. We shall discuss this later. 3. Remarks on the Thermal Conductivity Measurements The three main features we have just discussed can be understood by recalling some of the unusual properties of solid Ar: (i) the large coefficient of thermal expansion and (ii) the low yield stress. These two properties are clearly manifested by the behavior of K (T) (Figure 13). Cont Between 4 K and 10 K the relative change in the lattice constant of Ar is negligible (approximately 0.01%)46 so that for a constant applied force (as is the case here) d and n in Equation (16) should be essentially constant in this temperature range. In the last column of Table 6 we show K = nd which does remain essentially constant Cont/K in this temperature range. Thus KCont(T) is proportional to K. This temperature behavior (Figure 13) below 10 K is in quantitative agreement with the results of M05547 who measured K(T) of plastically deformed crystals of CaF2 and found that K varied somewhat less than T2. Above 10 K, n and d are expected to increase with an increase in temperature. This is because the sample as well as the asperities will begin to experience a signifi- cant increase in size. In going from 4 K to 20 K the rela- tive increase in the lattice constant is 0.13%, while in 91 going from 4 K to 40 K it is 0.69%.46 As the sample in- creases in length relative to the stationary sample tube ends, the applied force to the sample ends is effectively increased. The increasing force coupled with a diminishing yield stress results in a rapid increase in the contact area above 20 K. Hence, we expect that above 20 K,K (T) Cont and K /K will both rapidly rise. A quick glance at Cont Figure 13 and at the last column of Table 6 show convinc- ingly that our results support this conclusion. The first and second effects, (a) and (b), mentioned in the previous section are due mainly to a reduction in the contribution of R (thermal resistance of the sample) because of a reduction in L and also possibly an increase in the thermal conductivity K for the shorter samples. This latter possibility can be investigated by calculating K(T) separately for the short samples (0.33 cm 5 L S 1.34 cm) and long samples(l.97cm1$ L S 2.98 cm). In Figure 16 we have plotted only the meaningful data contained in Table 8. The plots in Figure 16 indicate that the shorter sam- ples are apparently less damaged (defected) than the larger ones. That is, the thermal conductivity peak of the shorter samples is higher (in fact a factor of 2 higher) and occurs at a lower temperature than for the longer samples. This is consistent with both the experimental observations and the fact that the shorter samples will experience smaller thermal gradients during the cooldown process. This result supports our belief that the thermally induced strains cause 92 hmm.o om.bm ma.mn Hmm.o mH.Hm H.m> oo.oa wvm.o m~.mm Hm.mm mmm.o mm.mm m.mm om.m mvm.o wo.vm ma.vm .mmm.o mo.>m m.HoH oo.m mmm.o mm.om >.moa nnm.o nm.vm m.mHH om.m Nam.o om.>v m.mma mom.o mm.am m.mma oo.m ham.o ~¢.m¢ m.mma mmm.o am.mv v.nma om.> oam.o mm.om H.omH mmm.o mm.mv m.mma 00.5 nbm.o No.mv m.¢ma mmm.o oa.mw H.mam om.o mom.o mv.mm H.NHH vvm.o Ho.mm m.¢mm oo.w vom.o mm.am mm.am Hom.o mm.mm m.m¢m om.m 5mm.o va.mm mm.mn mmn.o mm.Hm m.oom oo.m vmm.o mw.vm Ho.om mom.o mm.>~ m.mam om.v vvm.o mm.mv vm.mm mm>.o ¢¢.mm N.oma oo.v mmh.o Hm.mm hm.mv mmh.o Ho.aa m.m>a om.m vm>.o mm.vm mh.mv mmn.o Hm.mH m.nma oo.m omm.o mm.na vm.mv mmm.o mm.ma o.mmm m>.m wmm.ol Hm.¢ vw.h¢1 omm.o mm.ma h.mvm om.m www.cl wm.m vo.mm| omm.o mm.oa c.0HN mm.m ucwflowmmmoo ucmHOmemOU coflumamuuoo AM\3Evucoox AM EO\BEVy cowumamuuoo AM\BEVuCOUM Ax EO\BEVy AMVB EU mm.m w d w E0 hm.H EU vm.H w a w EU mm.o mmHmEmm mcoq paw uuonm mo mumo wpfl>flpospcoo Hmanmce m mHQMB 1000 93 500 100 50 THERMAL CONDUCTIVITY K(Mw/CM K) T E T 10 ° SHORT SAMPLES 0.33 - 1.34 CM 0 LONG SAMPLES 1.97 - 2.98 CM I ALL SAMPLES 0.33 - 2.98 CM 0 o c. . . . o o .... w.- e 9 r '9 P— M? __ I __ I _. I I P. I I r- I I I I L, T l T 1 l 111 l l | l l l 5 10 50 TEMPERATURE T(K) Figure 16. A plot of the calculated thermal conductivity versus temperature for short and long samples. 94 more damage to the entire sample than pressing on the sam- ples ends. The third effect (c) we mentioned in the previous sec- tion is simply due to the rapid increase in contact area resulting from the combination of the increasing applied force and the reduction of the yield stress of solid Ar. This results in the significant reduction of RCont relative to R near 38 K for all samples. 4. Thermal Conductivity Errors In determining K for Ar from 2.25 K to 65 K we recall that the data were fit to Equation (10) using from 9 to 3 different lengths (See Table 5) for a given isotherm using the method of least squares. For clarity we recall Equa- tion (19) REff : mL + RCont (19) l = 1/K where REff = KEff’ m = l/KA and R Since we Cont Cont' were able to measure L directly within i 0.3%, we will assume that L is known exactly to determine the uncertainty in K and K The quantities contributing to the uncer- Cont' tainty are therefore REff and A the cross sectional area of the samples. 'In this case if the error in REff is the same for each length Li’l°e'6REff,i = SREff,then the variance of m is . 48 given by v = (a )2/(2 L 2 - (2 L )Z/N) <20) REff ii ii and the variance of R is given by Cont 95 V( = <(5 )2> RCont 2 RCont) _ 2 2 _ 2 — (aREff) (i Li )/(N§ Li T; Li) ). (21> l Ordinarily, if R were exactly reproducible for each run Cont at a given length Li' then GREff would be proportional to 6(AT). Because of the way in which we smoothed the data, we expect the errors associated with AT to average to zero. Hence, the major contribution to the uncertainty in the determination of REff is the uncertainty of reproducing the contact resistance SRCont for each run. To estimate the maximum uncertainty associated with RCont we compared the data of sample 14 runs 1 and 2 shown in Figure 17. These two runs we carried out on the same sample for approximately the same length 3.00 cm and 2.98 cm, respectively. We note that for run 1 mechanical contact was lost and then reestablished at 25 K, thus these data were not used in determining K. For the expressed purpose of estimating 6R we Cont' will assume that K remained unchanged from run 1 to run 2 and that L and A are known exactly. This is a valid as- sumption, since the sample was manipulated in essentially the same manner for both runs and annealing between runs is expected to restore the sample to its original unstrained 49,50 state. If we take run 1 to be representative of the maximum value of RCont and run 2 to be representative of the minimum, then the error in RCont (or REff) is given by, 6Reont = 6REff = i 1/2 (REff,l ' REff,2) (22) 100 EFFECTIVE THERMAL CONDUCTANCE K5,,(Mw/K) 96 SO._ 100 _ ltggoou. .30 0' ID - .C] D. ID - 10.. . o D I — D _ I a DI _ I n u I o D - __ I D S_ C] DI _ D _ E] SAMPLE N0. 14 a RUN N0. 1 L = 3.00 CM I RUN NO. 2 L = 2.98 CM 1 TIIIIII I IILIlll‘ l 5 10 50 Figure 17. TEMPERATURE T(K) A plot of the effective thermal conductance for two samples of apprOXimately the same length displaying different values Of RCont- 97 where REff,l and REff,2 are the effective thermal reSist- ances for runs 1 and 2, respectively, at a given tempera- ture. The values of SREff determined in this way were found to vary from 0.021 K/mw at 2.25 K to 0.001 K/mw at 40 K. The fractional error in KCont is then determined from Equation (19) and the relation SKCont/KCont = GRCont/RCont (23) The value of (SKCOnt/KCont)(100%) was determined to vary from a minimum of 9% at 6 K to a maximum of over 100% above 36 K. This large error near 36 K is just the consequence of RCont approaching zero as T approaches 38 K. In fact the appearance of the physically unreal, negative values of in Table 6 in the vicinity of 40 K is just the KCont result of the straight line to which we fit our data, statistically missing zero as the y-intercept. Since K = l/mA, the fractional error for K is then given by (SK/K = i (<6m2>/m2 + (6A/A)2);i. (24) Recalling that (6A/A)(100%) = 2.0%, the values of (6K/K)(100%) were determined to vary from 22% at 2.25 K to i 2.0% at 65 K. All of the percentage errors for KCont and K are shown in Table 6. For the raw thermal conductivity data, corrected for the spurious heat conductance (Ksp), shown in Table 7 and plotted in Figure 15, we compute the percent error in a straightforward manner. Since for these data 98 K = (KEff - Ksp)L/A, the fractional uncertainty in K is given by _ _ _ 2 éK/K — : {(6(KEff Ksp)/(KEff Ksp)) L + (6L/L)2 + (6A/A)2}2 . (25) In a separate experiment with the sample tube and vacuum chamber evacuated we measured KSp directly and found it to vary from 1.11 x 10.1 mw/K at 83 K to 1.10 x 10_2 mw/K at 10 K. The uncertainty in this measurement was less than 0.7% throughout that temperature range. The magnitude of KSp represents a correction of 7% to KEff at 83 K and a correction of less than 1% at 38 K for a sample of intermediate length, say 1.3 cm. (Since KSp represents less than 1% of the total thermal conductance below 38 K it was unnecessary to correct KEff data for KSp at tempera- tures below 38 K.) Thus the evaluation of Equation (21) gives the result thatthe percenterror for the raw data representingkcis approximately i 8% for the temperatures between 38 K and 83 K. IV. DISCUSSION AND CONCLUSION In the paragraphs that follow we shall briefly discuss how our data compare with the results of previous workers and with the theoretical calculations of Christen25 and Christen and Pollack.24 In our concluding remarks we will cite the work of Kimber and RogersSI, who carried out meas- urements of K for constant volume samples of Ne, to aid us in suggesting possibly a better approach to measure the thermal conductivity of Ar. At high temperatures (T20) K is expected to be proportional to l/T for a constant volume sample. Clayton and Batchelder who measured K of isochoric samples of Ar having approximately the same molar volume as ours did in- deed find that K c l/T above 20 K. Our data, on the other hand decrease faster than l/T above 20 K and are in good agreement with the results of Krupskii and Manzhelii21 (see Figure 11) who found that l/K = AT + BT2 (A,B > 0). The high temperature, theoretical, K calculation of Christen and Pollack appears to reconcile the deviation of our data and the data of Krupskii and Manzhelii from the l/T behavior. In this complicated computer calculation Christen and Pollack24 used the best known interatomic 99 100 potential (Barker and Pompesz) to calculate the contribution of the anharmonic components of the crystal potential (three— phonon contribution) to the thermal resistivity. In their calculation, in order to take into account the effects of the lattice expansion, the lattice frequencies were eval- uated for the equilibrium density at a given temperature. The results of their calculation are plotted in Figures 15 and 18 along with our data for comparison. In the temperature range of 30-80 K our data are in reasonable agreement with the three-phonon calculation of Christen and Pollack.24 However, as the temperature falls below 30 K the theoretical curve (dashed curve in Figure 18) falls significantly below our data. This discrepancy as suggested by Christen and Pollack is probably due to the model used. That is, it was assumed, for this calcu- lation, that the collision frequency for normal processes was much larger than that for Umklapp processes. This has the effect of pOpulating those states which scatter strongly by Umklapp processes, depressing the value of the theoretical K between 6 K and 30 K. In the low temperature region for a defective solid dominated by dislocations the expected temperature depen- dence of K is T2. In general below 10 K our data are in qualitative agreement with the constant volume data of Clayton and Batchelder23 (see Figure 11). However, unlike the T2 behavior displayed by the data of Clayton and Batchelder, our data show a behavior that is somewhat slower 101 200 \ D‘IUOD 100 L. o a E. 0 0 i \ ”a u. D 50 __ 0 __ 0 Q \ 0 § ~ \ . é \ a X i - \\ D 0 ES \\\o L; o g 10 \ C 8 : \30 I 2;; VERA 85 I‘ 0 55 SE—' /// \9 1 ‘— - o PRESENT EXPERIMENTAL DATA ‘10 I _ / -— RELAXATION TIME THEORY \) ' _/ — — THREE PHONON THEORETICAL R I THERMAL CONDUCTIVITY “\o lI— I ll 1 I lllllll I ITIIJILI l 5 10 50 ‘ TEMPERATURE T(K) Figure 18. A plot of the low temperature (solid curves) and high temperature (dashed curve) theoretical calculations of Christen and Pollack. Included for comparison are the present data. 102 than T2 as the temperature decreases below 5 K. This is the same temperature dependence that was observed by Christen and Pollack. This observed temperature dependence may well be due to the deformed state of the samples we studied. However, since our error in this range is so large this question can only be properly answered after further investigation. In Figure 18 we have also included the relaxation time theory calculation of Christen, although this theory of low temperature K is still quite phenomenological. The three solid curves shown in this figure were computed for three different dislocation densities (0) having values of 2.5 x 109 cm_2, 5.2 x 109 cm_2 and 1.5 x 1010 cm"2, respectively, from the uppermost curve to the lowest. These calculations were performed considering only the conven- tional scattering mechanisms (dislocations, isotopic impu- rities, sample boundaries and 3 phonon, normal and Umklapp, scattering). The calculations also parametrically take into account the importance that normal processes, relative to the other scattering processes, have in redistributing phonon states among the various scattering mechanisms. Below 3 K the lst curve in Figure 18 (o = 2.5 x 109 cm ) is in agreement with the data of Clayton and Batchelder, while below 5 K the 2nd curve (0 = 5.2 x 109 cm_ ) turns out to be in rough agreement (within experi- mental uncertainty) with our data. Unfortunately, these values are 3 orders of magnitude higher than the values 103 27 determined by Peterson, 32 31. from x-ray analysis of sam- ples of Ar who set a lower limit of o = 106 cm-z. To end this part of the discussion we will present the results of previous workers who also carried out equilibrium volume measurements. In Figure 19 we show a plot of our results along with the data of Berne, gt £1.22 and White and Woods.19 Between 3 K and 13 K our data do appear to be in reasonable agree- ment with the data for one of the samples of Berne, gt 31. (filled circles). In general, near 20 K,K is expected to be independent of scattering from defects, however, our data appear to be significantly higher than the results of White and Woods.19 This difference can probably be attributed to a presence of a thermal contact resistance at the tempera- ture probes, in the data of White and Woods, that was not properly taken into account. To conclude this discussion we show in Figure 20 the data of Christen and Pollack for their runs 8 and 10 along with the "effective thermal conductivity" = K x L/A) (KEff Eff of our sample 12 run 1. If there were no thermal contact resistance, from the apparent agreement of the data we would probably, conclude that the thermal conductivity is inde- pendent of the cooling rate. This is unlikely since our cooling rates were approximately 70 times those of Christen and Pollack. The more likely explanation is that the ther- mal contact resistance in the experiments of Christen and Pollack probably changed from run to run. 200 THERMAL CONDUCTIVITY K(Mw/CM K) 100 50 5.; O 104 I I I I l I dime II ' o "0 r— 0. . O _. D _. CL. . 0 O __ U. I - U r— C] C _8l3 . D _. o A o ‘r I" o __ .AP 0 D I U .A A. ”' o _ ODD ‘—' O P - . o DERNE, BOATO a DEPAz DATA __ A WHITE 3 NOODS DATA DA 0 PRESENT DATA “’0 I I I l i ' i I L ' 1 S 10 50 TEMPERATURE T(K) Figure 19. A plot of the thermal conductivity data of several workers for Ar under its own equilibrium vapor pressure. 100 THERMAL CONDUCTIVITY K(MW/CM K) 105 100 — o oeoonl0 — . 3"”‘4‘. 50 ,_ 0 .% Q?:.A O 0? 'kA . L. ch.' (“‘5‘ 2 1. _ (3“ 2. 10__ L: a E- o RUN N0. 10 CHRISTEN & POLLACK DATA SP- - RUN N0. 8 CHRISTEN a POLLACK DATA r- . EFFECTIVE THERMAL CONDUCTIVITY __ SAMPLE N0. 12 RUN N0. 1 PRESENT DATA _f.___'e 1, l Llllllll 4 1111111 5 10 50 100 TEMPERATURE T(K) Figure 20. A plot of the thermal conductivity data of Christen and Pollack from runs 8 and 10. Included for comparison is a plot of the "effective thermal conductivity" from the present data, sample 12 run 1. pa 106 Figure 21 contains a plot of our data corrected for RCont including the data of Christen and Pollack for runs 7,8 and 10. The data of run 7 seem to be in agreement with our data and suggest that possibly RCont for this run was smaller than for the other two. It is unfortunate that the presence of RCont masks the true thermal conductivity in measurements of this type, since the quality of the samples studied by Christen and Pollack is probably greater than what the data would imply. In summary, we have demonstrated that above 38 K the linear heat flow method we have employed yields K of Ar directly. Unfortunately, at low temperatures (2 to 15 K) RCont becomes significant and must be reproduced reliably in order to determine K. In addition, this method yielded only an average value of K for all of the samples studied. Thus, unless one is willing to work with infinitely large samples, our method will not yield K directly. If one is to avoid the complicating effects of a thermal contact resistance a truly potentiometeric method must be used. That is, the probes used to measure AT must be in good thermal contact with the sample and must conduct a negligible amount of heat current during the K measure- ments. The experiments carried out on equilibrium volume samples of Ne by Kimber and Rogers51 employed just such a technique. Kimber and Rogers grew cylindrical samples of Ne in a glass tube (in a manner similar to ours) around THERMAL CONDUCTIVITY K(MW/CM K) 107 150 00000 D D 100% O a L. . 0 t 0 008.0000 .00 L— [3 (>03’ru‘.': . 0 50E_ 0 O 2‘? q o D 3 o _ C) 0 ° 0 D 0.! g o I—- 00 D O o L. 00 o D °o D D lOI_ C] _ D O r“ o _ Ebb 5‘ - - c CHRISTEN a POLLACK DATA 0 I RUNS 7, 88.10 ”D. j“ o PRESENT DATA 0 IH- P - DD 1 I IiIIIILL I IIllLll l 5 10 50 100 TEMPERATURE T(K) Figure 21. A plot of the thermal conductivity versus temperature of the present data. Included for comparison are the data of Christen and Pollack, runs 7, 8 and 10. 108 looped platinum wire 0.5 mm in diameter. Before measure- ments were commenced the samples were first separated from the glass tube and then cooled in a slow and even manner to the temperature range of interest (from 24 K to 4 K in 3 hours). The results for isotopically pure 20Ne samples were in excellent agreement with the results of Clemans53 who carried out measurements of K for constant volume sam- ples of 20Ne of the same purity. In this cleverly designed experiment Clemans had the capability of measuring the K of the sample in segments and was thus able to test whether the sample was of a uniform quality throughout. The sample chamber used by Clemans consisted of a thick-walled stainless tube with three annu- lar heaters equally spaced one cm apart and a single ther- mometer located at the bottom was used to measure AT. In this way Clemans was able to measure K for the portion of the sample between the upper heat sink and any one of the three annular heaters during a single run. The excellent agreement between these two methods, un— doubtedly suggests that the successful method used by Kimber and Rogers be employed in future experiments to obtain direct measurements of the thermal conductivity of equilib- rium volume samples of Ar. APPENDICES APPENDIX A Table 3 Effective Thermal Conductance at Constant Force Force Gradually Increased to Fmax = 30 nt T(K) KEff(mw/K) T(K) KEff(mw/K) 5.01 8.72 5.12 12.17 5.07 8.74 5.12 12.77 5.06 10.57 5.13 13.16 5.05 11.65 Force = 30 nt 5.13 13.47 6.52 16.05 4.86 12.41 6.80 16.22 4.49 10.64 7.04 16.29 4.14 9.89 7.30 16.45 3.86 8.19 7.57 15.97 3.59 6.87 7.78 16.25 3.94 8.25 8.01 16.37 3.58 7.10 8.27 15.62 3.28 5.62 8.52 15.70 3.27 5.60 8.74 15.73 4.59 9.42 9.00 15.27 4.60 10.27 9.23 15.15 4.61 10.82 9.50 14.94 4.61 10.79 9.76 14.12 4.83 11.66 9.99 14.12 4.99 12.39 10.46 13.46 5.25 13.94 11.00 12.59 5.50 15.05 11.46 12.25 5.79 14.08 12.02 9.63 5.82 14.47 12.96 11.60 6.02 14.53 14.00 10.40 6.26 15.04 15.00 8.59 109 T(K) 15.01 15.00 17.79 20.04 20.02 20.03 20.03 12.54 12.98 12.01 5.00 5.00 4.51 4.05 4.05 5.52 6.01 6.48 7.01 7.50 KEff(mw/K) 9.58 9.23 8.01 8.29 6.95 6.53 6.94 11.75 11.01 11.68 Applied Force 5.88 5.88 4.72 3.66 3.66 6.77 8.09 9.13 9.68 10.36 110 Table 3 (cont'd) T(K) 9.99 7.96 6.06 6.95 6.95 5.01 5.01 5.01 5.00 15 8.06 8.52 9.08 9.58 10.01 11.04 12.02 12.99 13.98 14.98 nt KEff 14 10. 10. 10. 10. .00 .39 .34 .14 .15 .39 oxxiooooxo (mW/K) .57 17. 17. 18. 18. 14. 14. 14. 14. 15 01 68 18 20 34 75 27 71 36 35 46 T(K) 5.05 5.05 5.15 5.21 5.50 5.99 6.02 6.26 6.51 6.78 7.00 7.23 7.42 7.44 7.43 7.42 7.41 7.74 8.02 8.20 8.51 9.03 9.52 9.96 11.08 12.00 13.07 14.03 15.30 16.18 17.24 KEff Sample No. 29.03 27.66 28.84 28.00 26.70 29.53 31.83 32.56 32.62 33.00 33.10 33.83 33.74 34.32 33.33 33.38 34.23 32.95 32.77 32.80 32.11 31.73 29.91 28.93 26.89 25.44 23.35 21.11 19.40 18.64 18.33 (mW/K) Table 111 APPENDIX B 4 Effective Thermal Conductance Data 10 Run No. T(K) 19.87 21.44 23.32 24.79 26.75 28.61 28.59 28.61 28.61 26.09 24.23 23.14 21.78 20.05 17.88 16.07 14.82 13.78 11.95 11.08 9.98 9.52 9.03 8.52 8.03 7.49 7.03 6.51 6.06 5.73 5.52 KEff 14.70 14.23 11.39 11.94 10.44 9.75 8.57 8.80 9.34 10.72 10.41 10.80 13.06 14.52 17.64 20.81 19.84 21.57 23.92 27.32 29.10 30.01 31.51 32.03 32.40 31.57 33.57 32.08 31.02 28.72 28.37 (mW/K) T(K) 5.25 5.01 4.76 4.51 4.52 4.24 4.00 4.91 5.00 5.12 5.30 5.45 5.64 4.81 4.67 4.49 4.35 4.21 2.91 2.94 2.98 3.01 3.06 3.16 3.26 3.35 3.47 3.52 3.47 3.53 3.64 3.77 3.87 4.02 4.14 4.27 4.38 4.52 4.64 4.59 KEff(mw/K) 27.58 24.71 24.37 22.09 23.05 21.07 20.49 Sample No. 23.84 23.89 24.09 24.56 24.32 24.73 21.21 20.96 19.94 18.72 18.85 11.03 11.97 10.81 11.20 11.66 11.93 12.89 13.42 13.88 13.55 13.81 14.06 14.97 15.61 16.08 16.82 17.77 18.12 18.80 19.40 20.12 20.51 Table 4 (cont'd) T(K) 3.77 3.49 3.27 3.22 3.27 3.27 11 Run No. 4.79 4.92 5.01 5.08 4.99 5.13 5.22 5.29 5.41 5.52 5.65 5.81 5.89 5.98 6.09 6.09 6.21 6.43 6.60 6.84 7.03 7.18 7.40 7.56 7.78 8.05 8.24 8.38 8.01 7.38 8.38 8.69 8.86 3 18.00 15.95 14.64 15.02 15.15 14.16 21.58 23.24 23.10 24.06 23.94 23.65 24.30 25.41 25.95 25.10 25.34 25.90 26.13 26.25 27.56 27.16 27.04 26.90 27.20 26.27 26.37 27.34 28.72 27.19 26.33 27.09 24.80 25.76 27.06 25.90 23.54 22.98 25.12 KEff(mw/K) T(K) 8.84 9.06 9.23 9.56 9.79 9.97 10.55 10.99 2.92 2.96 3.00 3.04 3.07 3.16 3.24 3.33 3.40 3.51 3.61 3.73 3.80 3.90 4.03 4.17 4.28 4.42 4.55 4.75 4.95 4.85 5.09 5.26 5.45 5.68 5.64 5.86 5.98 6.12 6.22 KEff)mw/K) 26.94 25.98 24.93 22.82 21.60 20.57 21.21 19.65 Sample No. 14.15 14.27 14.47 14.71 14.95 15.30 15.55 16.41 16.95 17.37 18.56 19.52 19.86 20.50 21.37 22.21 22.99 23.90 24.52 25.51 28.12 27.49 29.01 29.10 30.13 30.08 30.03 29.61 30.26 33.05 32.94 Table 4 (cont'd) T(K) 11.94 13.08 14.17 15.89 18.36 20.78 24.66 11 Run No. 6.47 6.51 6.73 6.99 7.22 7.47 7.77 8.00 8.23 8.55 8.78 8.32 8.65 8.96 9.20 9.37 9.57 9.78 9.96 10.56 11.07 12.29 14.19 16.19 21.15 25.25 31.16 37.83 45.40 63.45 KEff(mw/K) 17.91 15.47 15.08 13.10 10.09 8.77 6.54 33.57 34.15 34.38 34.78 34.00 34.20 33.68 33.43 33.26 33.28 32.15 33.11 32.05 32.00 30.85 30.48 29.55 28.76 28.76 27.32 26.00 23.41 19.67 16.91 11.71 9.46 6.82 5.51 4.28 2.63 114 Table 4 (cont'd) Sample No. 12 Run No. l T(K) KEff(mw/K) T(K) KEff(mw/K) 4.67 16.79 4.23 15.00 4.67 16.97 4.35 15.37 4.74 17.22 4.48 15.92 4.89 18.08 4.59 16.63 4.99 18.37 4.69 16.72 5.10 18.83 6.02 21.82 5.20 19.54 6.19 22.03 5.21 19.34 6.47 22.14 5.33 19.83 6.62 22.31 5.47 20.24 6.87 22.14 5.58 20.96 7.00 22.31 5.66 21.18 7.27 22.13 5.76 20.96 7.46 21.93 5.85 21.38 7.66 21.88 5.94 21.73 7.85 21.75 6.02 21.77 8.01 21.24 6.08 21.96 8.25 20.42 6.17 22.09 8.48 19.95 6.27 22.44 8.71 19.42 6.37 22.20 8.90 18.86 2.30 6.70 9.04 18.62 2.35 6.74 9.24 18.06 2.40 6.99 9.49 17.35 2.47 7.27 9.70 17.03 2.54 7.60 10.04 16.07 2.61 7.87 10.63 14.93 2.68 8.19 11.09 14.17 2.77 8.52 11.59 13.29 2.85 8.80 12.07 12.47 2.93 9.10 12.69 11.52 2.99 9.25 13.11 11.15 3.08 9.71 14.25 9.91 3.13 9.88 15.10 9.14 3.21 10.05 10.60 15.30 3.29 10.34 11.04 14.28 3.38 11.03 12.57 11.91 3.50 11.43 13.71 10.67 3.63 12.05 14.73 9.63 3.75 12.52 15.65 8.75 3.86 13.07 16.41 8.32 3.97 13.58 17.65 7.53 4.10 14.21 18.71 6.63 T(K) 17.70 18.79 19.95 21.52 4.59 4.69 .87 .02 .18 .32 .97 70 .48 .23 .98 .77 .57 40 .76 .98 .55 .06 \IONU'IU'IU'INNNWWUJWbUIUIub .42 .61 .79 .14 90 .68 .41 .18 .90 .75 .28 .39 .49 .62 NNNNNNwwwwebpp KEff 7.38 6.99 6.73 5.54 Sample No. 27.01 27.51 29.06 30.11 31.22 24.98 22.41 20.61 18.98 17.10 15.23 13.95 12.63 33.17 35.55 36.06 38.81 41.32 Sample No. 23.92 24.96 26.07 22.00 20.38 18.89 17.28 15.56 13.61 13.00 10.14 10.86 11.45 12.33 (mW/K) Table 4 (cont'd) T(K) 24.19 26.93 32.00 37.87 13 Run No. 7.55 6.02 8.03 8.51 8.93 9.51 10.01 11.07 12.51 14.94 17.16 20.31 24.85 31.36 39.51 49.53 60.87 13 Run No. 5.00 5.20 5.52 5.82 5.56 5.98 6.27 6.52 6.73 7.03 7.49 7.96 8.49 9.05 KEff 4.61 4.00 3.02 2.33 44.29 36.55 45.66 45.61 46.64 47.06 46.90 49.09 46.23 43.47 39.60 37.13 29.17 25.23 19.26 16.43 10.49 27.65 28.65 29.21 32.50 30.95 32.26 33.07 33.74 34.26 37.26 35.55 36.82 35.46 34.28 (mW/K) T(K) 9.51 10.01 11.08 12.56 4.61 4.76 4.96 5.20 4.29 4.42 4.54 3.99 4.08 4.19 2.39 2.49 2.59 2.73 2.88 2.99 3.11 3.22 3.34 3.48 3.70 3.80 3.94 5.16 5.36 5.48 5.66 5.84 6.03 6.18 6.36 6.57 KEff 116 Table 4 (cont'd) (mW/K) 32.75 32.19 29.69 26.48 Sample No. 12.35 13.04 13.88 14.88 11.17 11.67 12.25 10.01 10.40 10.75 4.92 5.22 5.54 5.94 6.34 6.57 6.93 7.18 7.70 8.16 8.95 9.31 9.87 14.52 15.49 15.92 16.75 17.12 17.74 17.85 18.32 18.88 14 Run No. T(K) 15.01 23.84 28.16 35.20 6.71 6.87 7.02 7.20 7.41 7.65 7.87 8.03 8.28 8.56 8.80 9.03 9.28 9.50 9.72 9.74 9.99 10.37 10.53 11.09 11.85 11.46 13.21 12.39 14.20 15.23 15.91 18.80 21.57 25.90 33.26 KEff (mW/K) 21.56 11.85 10.22 7.32 18.85 18.84 19.05 19.13 19.13 19.22 19.15 19.05 18.55 17.94 17.50 17.00 16.53 16.22 15.61 15.79 15.42 14.58 14.25 13.75 12.39 12.82 10.43 11.51 9.81 9.13 8.31 6.75 5.44 4.30 3.06 117 Table 4 (cont'd) Sample No. 14 Run No. 2 T(K) KEff(mw/K) T(K) KEff(mw/K) 4.60 14.34 5.92 19.79 4.60 14.35 6.05 20.05 4.70 14.79 6.23 20.41 4.85 15.41 6.40 20.61 4.96 15.73 6.61 20.89 5.07 16.46 6.77 21.02 5.16 16.83 7.02 21.25 5.28 16.44 7.23 21.20 2.28 5.76 7.54 21.43 2.37 6.08 7.75 21.43 2.50 6.55 8.03 21.01 2.62 6.90 8.30 20.27 2.75 7.30 8.74 19.36 2.90 7.73 8.97 18.67 3.01 8.01 9.35 17.81 3.09 8.24 9.67 17.32 3.22 8.78 9.90 16.85 3.29 9.03 10.30 15.86 3.38 9.34 6.03 20.01 3.50 9.81 6.50 20.79 3.55 9.98 7.02 21.27 3.63 10.29 7.45 21.23 3.76 10.69 7.97 21.16 3.90 11.38 8.99 18.64 4.03 11.82 10.08 16.58 4.14 12.39 10.58 15.69 4.23 12.63 11.01 14.90 2.36 13.26 11.95 13.24 4.55 14.09 12.56 12.42 5.40 17.81 13.62 11.14 5.52 18.27 15.01 9.90 5.64 18.95 16.69 8.39 5.73 19.18 18.53 7.46 5.83 19.79 20.92 6.36 5.93 19.83 23.88 4.98 6.06 20.10 118 Table 4 (cont'd) Sample No. 15 Run No. 1 (T) KEff)mw/K) T(K) KEff(mw/K) 83.08 1.82 62.92 2.86 81.70 1.77 59.65 3.04 79.90 1.90 55.93 3.29 78.27 2.00 52.85 3.83 76.00 1.97 51.34 3.92 72.85 2.14 48.29 4.36 68.72 2.37 45.99 4.61 65.75 2.55 52.50 3.91 Sample No. 16 Run No. 4.66 26.86 9.00 42.23 4.88 27.60 9.46 41.47 5.12 29.74 9.98 40.77 5.44 32.91 10.48 40.66 2.59 12.50 11.04 39.18 2.65 12.89 11.97 37.48 2.87 14.20 13.16 36.20 3.00 14.93 13.99 33.65 3.27 16.83 14.94 32.80 3.50 18.43 16.14 30.16 3.77 20.39 18.02 27.46 4.02 22.25 20.08 25.44 4.26 23.91 23.40 19.86 4.42 24.98 25.11 17.77 4.54 26.02 26.87 18.09 4.75 27.45 28.20 17.30 5.00 29.03 30.21 15.32 5.48 32.41 32.18 13.94 5.69 33.94 34.42 12.56 5.99 35.43 36.23 11.84 6.28 36.44 38.14 10.97 6.66 37.89 41.38 11.11 6.99 39.31 42.25 10.07 7.39 40.54 44.61 9.34 7.81 42.54 47.61 7.87 8.01 42.00 50.23 7.37 8.50 41.96 53.81 6.72 119 Table 4 (cont'd) Sample No. 16 Run No. 2 T(K) KEff(mw/K) T(K) KEff(mw/K) 4.69 16.39 6.67 22.36 4.82 17.82 6.82 22.31 4.99 17.98 7.01 22.21 5.23 19.07 7.22 22.39 5.40 19.86 7.41 22.30 2.47 6.09 7.72 22.26 2.62 6.76 7.94 22.12 2.79 7.48 8.21 21.56 3.00 8.24 8.46 20.87 3.25 9.17 8.71 20.14 3.46 10.42 8.97 19.41 3.76 11.81 9.14 19.23 4.01 13.18 9.39 18.78 4.23 14.20 9.70 18.29 4.50 15.48 10.00 17.73 4.78 16.77 10.35 17.53 4.32 14.51 10.78 16.39 4.62 16.15 11.06 16.02 4.82 17.09 12.02 14.12 5.12 18.47 12.82 12.92 5.29 19.06 14.16 11.69 5.51 20.09 15.11 10.60 5.66 20.70 16.09 9.71 5.75 21.28 18.42 8.69 6.00 21.66 20.66 7.18 6.15 21.55 22.63 6.16 5.99 21.15 24.27 5.36 5.83 21.04 27.15 4.82 5.65 20.43 29.65 4.09 6.00 21.13 34.43 3.23 6.32 21.69 39.31 2.64 6.47 21.88 APPENDIX C Defect Scattering - Classical Analogs 1. Rayleigh Scattering from a Spherical Object - Dimensional Analysis The Rayleigh scattering law for a spherical obstruction can be verified by energy conservation and dimensional anal- ysis considerations.15 Consider a plane wave of unit ampli- tude wi = exp(iqz) incident upon a spherical object of volume V such that 1>>Vl/3. The amplitude of the scattered wave ws is necessarily proportional to 1/r since the scattering intensity is proportional to lwslz |2 a 1/r2. This is because flwslzrzdfl must be equal to a constant. Now the Q and st only other guantities that ws can depend on are u (velocity of sound), A (wavelength) given by 2'rrq_l and V (volume). Since U is the only quantity that involves time as one of its dimensions, it is discarded. Hence, the simplest prod- uct that we can form from 1, r and V such that 05 remains dimensionless is V/Azr = L3/L2L = 1. Thus, $5 a V/Azr and the scattering cross section is proportional to V2/14. That is, o is proportional to q4 120 121 2. Scattering from a Cylindrical Object - Dimensional Analysis The Rayleigh scattering law foreulinfinite cylindrical obstruction can be verified as in the preceding case.15 In this instance since we have cylindrical symmetry, the pro- blem can be reduced to 2 dimensions. Scattering intensity considerations yield that the amplitude of $5 a 1/03, since glwslzpde must be a constant. The other quantities on which ws can depend are A and A = naz the cross sectional area of the cylinder. Thus the simplest product we can form is A/A3/Zpg, i.e., $3 « A/A3/20;5 = Lz/LB/ZL15 = 1. Therefore, the core scattering cross section per unit length of disloca- '2 tion is o a [05 ~ AZ/A3 = a(a/A)3. core 3. Scattering from the Strain Field Surrounding the Cylindrical Core - Geometrical "Optics" Limit For a phonon that is deviated from its original direc- tion by an angle 0 4 yb/p on passing through the strain field, the change in momentum along the original direction is proportional to (1-cos¢).32 Since the change in heat current will be reduced by an amount which is proportional to (1-cos¢), the scatteringcmoss secthmnper unit length of dislocation is thus proportional to p 2 2 2 f¢m(l-cos¢>)d¢> = meWb/p) dp = Y b /pO 0 where p0 is the least value of p (p0 ~ 1). Thus 0 is proportional to yzbZ/X. strain APPENDIX D Error Associated with the Magnitude of AT In our measurements the expression we use to evaluate the effective thermal conductance is given by KEff (T) = Q/AT (01) where T = TO + AT/2 is the average temperature. For small AT Equation (D1) is a good approximation of the following equation . TO+AT K = Q/AT = (f K(T)dT)/AT (D2) Eff T0 . . 14,25 where K(T) is the true effective thermal conductance. At low temperatures K(T) 2 ATZ. On substituting this expression into Equation (D2) and evaluating the integral we obtain for the effective thermal conductance K = ((T 3 3 Eff + AT) - To )A/3AT . (D3) 0 The true effective thermal conductance at T is given by K(T) = A(TO + AT/2)2. (D4) The relative difference (K(T) - KEff)/K(T) to lowest order in AT/T0 is given by AK/K : (AT/T0)2/12 . (05) Hence, if we want Equation (D1) to stay within 1% of the 122 123 true effective thermal conductance, then the largest AT that we should not exceed at the lowest temperature (2 K) is given by AT : 2(o.12);5 = 0.7 K . (D6) Similarly at the highest temperatures, where K 2 BT-1 after performing the appropriate expansions, the error will also be given to lowest order in AT/TO by Equation (D5). Hence, if at 80 K we wish to stay within a 1% error, then the largest AT that we should not exceed is given by L T S 80(0.12)2 = 28 K. LI ST OF REFE RENCES 10. 11. 12. 13. 14. 15. 16. 17. 18. LIST OF REFERENCES H. Cavendish, Phil. Trans. R. Soc., 13, 372 (1785). G. A. Cook in Argon, Helium, and the Rare Gases I, G. A. Cook ed., (Interscience Publishers, New York, 1961). G K. Horton in Rare Gas Solids l, J. A. Venables and . L. Smith eds., (Academic Press, London, 1977). L. Pollack, Biophysical J. 33, 49 (1977). . Unsworth and F.C. Gillespie in Diffusion Processes , J. N. Sherwood, g; 33. eds., (Gordon & Breach, cience Publishers, Inc., New York, 1971). B G J 2 S D. A. Nelson and A. L. Ruoff, Phys. Rev. Lett. 33, 383 (1979). D. N. Batchelder in Rare Gas Solids 2, J. A. Venables and B. L. Smith eds., (Academic Press, London, 1977). G. L. Pollack, Rev. Mod. Phys. 33, 48 (1969). R. E. Peierls, Quantum Theory 9£ Solids (Oxford Univer- sity Press, New York, 1964). P. Carruthers, Rev. Mod. Phys. 33, 92 (1961). C. Kittel, Introduction 39 Solid State Physics (John Wiley and Sons, Inc., New York, 1971). R. E. Peierls, Ann. Phys. Leipzig 3, 1055, 1101 (1929). P. G. Klemens in Thermal Conductivity 1, R. P. Tye ed., (Academic Press, New York, 1969). R. Berman, Thermal Conduction ig_Solids (Oxford Univer- sity Press, Oxford, 1976). Lord Rayleigh, Theory 9f Sound 2 (Dover Publications, New York, 1945). F. R. N. Nabarro, Proc. Roy. Soc. A209, 278 (1951). J. M. Ziman, Nuovo Cimento Z, 353 (1958). B. L. Smith, Contemp. Phys. 33, 125 (1970). 124 19. 20. 21 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 125 G. K. White and S. B. Woods, Phil. Mag. 3, 785 (1958). I. N. Krupskii, V. G. Manzhelii, and L. A. Koloskova, Phys. Stat. Sol. 33, 263 (1968). I. N. Krupskii and V. G. Manzhelii, Phys. Stat. Sol. 33, K53 (1967) and Soviet Phys. JETP, 33, 1097 (1969). A. Berne, G. Boato and M. DePaz, Nuovo Cimento 33, 182 (1969). F. Clayton and D. N. Batchelder, J. Phys. 93, 1213 (1973). D. K. Christen and G. L. Pollack, Phys. Rev. 812, 338 (1975). D. K. Christen, Thermal Conductivity 9: Solid Argon, Ph.D. Thesis, Michigan State University (1974). J. A. Venables and B. L. Smith in Rare Gas Solids 2, J. A. Venables and B. L. Smith eds., (Academic Press, London, (1977). O. G. Peterson, D. N. Batchelder and R. 0. Simmons, J. Appl. Phys. 33, 2682 (1965). G. K. White and S. B. Woods, Canad. J. Phys. 33, 58 (1955) and Nature, London 177, 851 (1956). R. K. Crawford and W. B. Daniels, J. Chem. Phys. 32, 3171—83 (1969). W. B. Daniels, 25 33., Phys. Rev. Lett. 33, 548-50 (1967). J. A. Leake, g3 33., Phys. Rev. 181, 1331—60 (1969). J. M. Ziman, Electrons and Phonons (Clarendon Press, Oxford, 1960). G. K. White, Experimental Techniques 32_Low—Temperature Physics (Clarendon Press, Oxford, England, 1968). A. C. Rose-Innes, Low Temperature Techniques, (English University Press, Ltd., London, 1964). V. I. Grushko, D. N. Bol'shutkin, G. N. Scherbakov and N. F. Shevchenko, Soviet Phys.-Crystallography 33, 536 (1970). 36. 37. 38. 39. 40. 41 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 126 M. Gsanger, H. Egger, G. Fritsch and E. Lfischer, Z. Angew. Phys. 33, 334 (1969). G. L. Pollack and E. N. Farabaugh, J. Appl. Phys. 33, 513 (1965). I. Lefkowitz, K. Kramer, M. A. Shields and G. L. Pollack, J. Appl. Phys. 8, 4867 (1967). G. L. Pollack and H. P. Broida, J. Chem. Phys. 38, 968 (1963). _—' C. R. Tilford and C. A. Swenson, Phys. Rev. 33, 719 (1772). R. Berman, J. Appl. Phys., 33, 318 (1956). A. V. Leonteva, Yu. S. Stroilov, E. E. Lakin and D. N. Bol'shutkin, Phys. Stat. Sol. 33, 543 (1970). R. Berman and C. F. Mate, Nature, Lond., 182, 1661 (1958). ‘— C. F. Mate in "Heat Flow below 100 K", p. 101, Annexe. 1965-2 Suppl. Bull. Inst. Refrig. (1965). D. Tabor, The Hardness 93 Metals (Oxford University Press, 1951). O. G. Peterson, D. N. Batchelder and R. 0. Simmons, Phys. Rev. 150, 703 (1966). M. Moss, J. Appl. Phys. 33, 3308 (1965). S. L. Meyer, Data Analysis for Scientist and Engineers, (John Wiley and Sons, Inc., New York, 1975). W. R. G. Kemp, P. G. Klemens and R. J. Tanish, Phil. Mag. 3, 845 (1959). J. N. Lomer and H. M. Rosenberg, Phil. Mag. 3, 467 (1959). R. M. Kimber and S. J. Rogers, J. Phys. 33, 2279 (1973). J. A. Barker and A. Pompe, Australian J. Chem. 33, 1683 (1968). J. E. Clemans, Phys. Rev. B 33, 1072 (1975). G. Boato, Cryogenics 3, 65 (1964).