M LIBRARY Michigan State University THESIS ,l i This is to certify that the dissertation entitled The Efiecznicaz Rebiazivity and the Thenmoe£ectnic Ratio 06 PotaAAium, Sodium, Lithium, Rubidium, and PotaAAium— Rubidium AZZoyA 610m 0.07K to 4.2K presented by Zhao-Zhi Va has been accepted towards fulfillment of the requirements for P11- 1). degree in 1714ij ms ’é» jégLvfihr 1/ Major professor MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 Date 10/5/84 )V‘ESI.) RETURNING MATERIALS: Place in book drop to LIBRARIES remove this checkout from AmnfizyuunL. your record. flfl§§ will be charged if book is returned after the date stamped below. THE ELECTRICAL RESISTIVITY AND THE THERMOELECTRIC RATIO OF POTASSIUM, SODIUM, LITHIUM, RUBIDIUM, AND POTASSIUM- RUBIDIUM ALLOYS FROM 0.07K TO 4.2K BY Zhao-Zhi Yu A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 1984 © 1985 ZHAO-ZHI YU All Rights Reserved THE ELECTRICAL RESISTIVITY AND THE THERMOELECTRIC RATIO OF POTASSIUM, SODIUM, LITHIUM, RUBIDIUM, AND POTASSIUM- RUBIDIUM ALLOYS FROM 0.07K TO 4.2K BY Zhao-Zhi Yu AN ABSTRACT OF A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 1984 ABSTRACT THE ELECTRICAL RESISTIVITY AND THE THERMOELECTRIC RATIO OF POTASSIUM, SODIUM, LITHIUM, RUBIDIUM, AND POTASSIUM- RUBIDIUM ALLOYS FROM 0.07K TO 4.2K BY Zhao-Zhi Yu The electrical resistivities (P) of thick (diam. = 1.0 - 3.0 mm) samples of K, Na, Li, and Rb; of thin (diam. a 0.09 - 0.5 mm) samples of K and Na; of samples of K and Na encased in polyethylene and teflon tubes; and of samples of K-Rb alloys were measured from 4.2K down to 0.07K with a precision of 0JL4L01 ppm. The thermoelectric ratios (G) of the same samples were measured over the same temperature range. In free-hanging, bare, high-purity thick samples, we found a T2 variation of the electrical resistivity for K, from about 1.1K down to about 0.3K; for Li, from 4.2K down to about 1.6K; and, perhaps, for Na from 1.9K down to 1.2K. No temperature range over which a T2 variation was dominant was found for thick Rb samples. Clear deviations from a T2 variation for K, Na, and Li below the temperature ranges just listed were also found. In thin samples of K cooled in He gas, we found a size effect pattern in P leading to a negative dP/dT when the Zhao-Zhi Yu diameter became smaller than electron mean-fee-path. The only model we know which might explain the negative dP/dT we observed is an interaction between surface scattering and normal electron-electron scattering first prOposed by Gurzhi and then calculated by a Monte Carlo method by Black. A Kondo-type anomaly was observed in the electrical resistivity below 1K for K samples encased in polyethylene tubes. The data can be well fitted by p= 90 + AT2 - BlnT, where -BlnT is a typical Kondo effect term. The resistivity of 9.4 at.% Rb alloy samples evidenced a strong deviation from the expected T2 behavior for TglK. The reasons for this deviation are not yet clear. ‘All G data of thick samples of pure K, Na, Li, and Rb and K-Rb dilute alloy could be well fitted by c a G0 + 13*1'2 + (C*/T)Exp(-6*/T), where the third term is negligible for Na and Li. Both the normal and Umklapp phonon drag terms were quenched by increased surface and impurity scattering. G data of K samples encased in polyethylene tubes were well . fitted by G = G0 + B*T2 + D*/T, (D*<0) where the last term is attributed to the Kondo effect. TO MY WI FE ii ACKNOWLEDGEMENTS It is a great pleasure to thank my thesis advisor, Professor Jack Bass, for his supervision, support, criti- cism, and invaluable aid in all aspects of this research. I would also like to thank Professors Pratt and Schroeder for their precious advice, discussions, and help at many stages of this study. Thanks also are due to Professor Spence for his help in testing the single-crystal potassium, and to Professor J. C. Garland of Ohio State University for providing high-purity sodium. I wold like to extend thanks to Dr. Mark Haerle and Dr. Vernon Heinen for their unselfish aid in various stages of this research, to Mr. John Zwart for his help in design and construction of the sample can, and to Zhao Jing and Qian Yao-Jin for their help in taking some data. I wish also to thank the good humored fellows in the machine shop for their help in constructing the sample can, the presses and other apparatus. Finally, the financial support of the National Science Foundation is gratefully acknowledged. iii LIST OF LIST OF CHAPTER I. II. TABLE OF CONTENTS TABLES O O O O O O O O O O O O O O O O O O FIGURES O O O I O O O O O O O O O O O O 0 INTRODUCTION 0 O O O O O O O O O O O O O O 1.1 Alkali Metals and Basic Electronic Transport PrOperties . . . . . . . 1.2 PreVious work O O O O O O O O O O I I 1.2.1 Previous Work on Resistivities 1.2.2 Previous Work on ThermOpower S and Thermoelectronic Ratio G 1.3 Present Thesis . . . . . . . . . . . 1.3.1 Electrical Resistivity . . . . 1.3.2 Thermoelectric Ratio G and Thermopower S . . . . . . . THEORY O O O O O O O O O O O O O O O O O O 2.1 Standard Theory of Electrical Resistivity for Alkali Metals . . . . . . . . . 2.1.1 Electron-Phonon Scattering . . 2.1.2 Electron-Electron Scattering . 2.1.3 The Residual Resistivity . . . 2.2 Beyond the Standard Model of Electrical Resistivity for Alkali Metals . . . 2.2.1 Electron-Electron Scattering in the Presence of Anisotropic Scat- terers such as Dislocations 2.2.2 Inelastic Scattering of Electrons Due to Low Energy Excitations Associated with Dislocations 2.2.3 CDW Theory and Electron-Phason Scattering . . . . . . . . . iv Page vii ix 13 l4 l4 17 18 19 19 24 25 25 29 32 III. IV. 2.3 The Knudsen Flow Model . . Size-Dependent DMR . . . . The Gurzhi Theory . . . . The Kondo Effect . . . . . Inelastic Scattering of Electrons by Impurities and Defects . . Electron-Electron Interaction Effect and Localization Effect . o o e o o N NNNNN o o o o o \D (DNO‘UIIh N NNNNN Thermoelectric Power S and Thermoelectric Ratio G I O O O O I O O O O O O O I O 0 2.3.1 Thermoelectric Power S 2.3.2 Thermoelectric Ratio G . . . . . . EXPERIMENTAL TECHNIQUES . . . . . . . . . . . 3.1 3.6 Introduction . . . . . . . . . . . . . . 3.1.1 The Main Equipment for Measurements. 3.1.2 Thermometers O O O O O O O O O O 0 3.1.3 The Main Equipment for Sample Preparation . . . . . . . . . . The Glove Boxes and Presses . . . . . . . Sample Can . . . . . . . . . . . . . . . Sample Preparation . . . . . . . . . . . Measurement Method . . . . . . . . . . . 3.5.1 Resistivity . . . . . . . . . . . 3.5.2 Thermoelectric Ratio G and Thermo- power S O O O O O O O O O O O O Uncertainties . . . . . . . . . . . . . . 3.6.1 Uncertainties in G and S Measurements . . . . . . . . . . 3.6.2 Uncertainties in P(4.2K), Po, an dP/dT Measurements . . . . . . . EXPERIMENTAL RESULTS AND ANALYSIS . . . . . . 4.1 Free Hanging, Bare, Thick (d=1.5-2.0 mm) High-Purity K Samples 0 o o o o o o o 0 4.1.1 The Resistivity . . . . . . . . 4.1.2 The Thermoelectric Ratio G . . . . V 51 51 54 55 55 56 57 59 S9 62 68 87 87 97 99 99 100 103 103 104 107 V. 4.2 4.3 4.4 4.5 4.6 4.7 Free Hanging, Bare, Thin (d=0.09-0.5 mm) 4.2.1 The 4.2.2 The 4.2.3 The Resistivity . . . . Thermoelectric Ratio G Thermoelectric Power 8 Nominal Single-Crystal K Sample . Resistivity . . . . Thermoelectric Ratio G High-Purity K Samples in Contact with 4.3.1 The 4.3.2 The Plastics 4.4.1 The 4.4.2 The K-Rb Alloy 4.5.1 The 4.5.2 The Free Hanging, Bare, Thick, High-Purity Na, or Oil 0 O O O O O O O Resistivity . . . Thermoelectric Ratio G Samples . . . . . . . Resistivity . . . . Thermoelectric Ratio G Li, and Rb Samples . . . . . . 4.6.1 The Resistivity . . . . 4.6.2 The Thermoelectric Ratio G . Free Hanging, Bare, Thin, High-Purity Na Samples . 4.7.1 The 4.7.2 The Na Samples 4.8.1 The 4.8.2 The Resistivity . . . . Thermoelectric Ratio G Encased in Polyethylene Tubes Resistivity . . . . Thermoelectric Ratio G SUMMARY AND CONCLUSIONS . . . . . . . LIST OF REFERENCES . . vi Page 113 113 132 139 145 145 148 148 148 168 175 175 185 187 188 201 206 206 209 211 211 213 215 222 Table 1.. 1 LIST OF TABLES Previous work on potassium . . . . . . . . . . . Characteristics of samples . . . . . Characteristics of samples . . . . . Characteristics of samples . . . . . Characteristics of samples . . . . . Characteristics of or plastics . . . Characteristics of samples . . . . . Characteristics of ethylene tubes . . bare, free-hanging, pure K bare, free-hanging, pure Rb bare, free-hanging, pure Li bare, free-hanging, pure Na K samples in contact with oil bare, free-hanging, K-Rb Na samples encased in poly- The electrical resistivities of some alkali metals Coefficient A and other parameters of thick K samples . . . . . L for isotrOpic ($3?. 38)) . . . . scattering (after Black Coefficient A' deduced from the fitting of Eqn. (4'2*) 0 o o o e e The coefficients from fits to the G data of thin samples cooled in He gas . . . . . . . . The coefficients from fits to the G data of thin samples cooled in Ar gas . . . . . . . . . . . L(T)/Lo values calculated from G*/G and from Wiedemann-Franz Law for sample K-1/2H . . . . . vii Page 10 70 73 73 74 81 85 86 89 106 123 125 135 138 146 4-10' Page The coefficient B from fits to the P data of K samples encased in polyethylene tubes . . . . . . 165 The coefficients from fits to the G data of K samples encased in polyethylene tubes . . . . . . 174 The coefficients from the local phonon mode fits to P data of Li samples . . . . . . . . . . . . . 200 The coefficients from the bound electron level fits to 0 data of Li samples . . . . . . . . . . . 200 viii \IO‘U'IuwaH LIST OF FIGURES Page Normal and Umklapp process of e-ph scattering (A) Phonon annihilation . . . . . . . . . . . . . 21 (B) Phonon creation 0 o o o o o o o o o o o o o o 21 The effect of a CDW perturbation on the Fermi surface. The sphere is distored to a lemon shape (AFter Overhauser (ref. 1) . . . . . . . . . 34 The Umklapp scattering for a distored Fermi surface (After Overhauser (ref. 1) . . . . . . . . 34 Plot of resistivity vs T for the data of sample K2C of Rowlands et a1. (ref. 16). The fitting curve with J4 is shown. In the inset, extrapola- tions of Tl's, J2, J , J5 fits to lower tempera- tures are shown. (After BishOp and Overhauser (ref. 17)) O O O O O O O O O O O O O O O I O I O O 37 Two kinds of stainless steel presses . . . . . . . 61 sample hOIder 1 O I O O O O O I O O O O O O O O O 63 Cover of the sample holder . . . . . . . . . . . . 64 Three different ways of mounting samples . . . . . 65 The way of sucking potassium into a tube . . . . . 75 The system for purifying paraffin oil . . . . . . 77 The system for pulling single crystal potassium sample 0 O O O O O O O O O O O O O O O O O I O O O 7 9 The low temperature circuit. The components inside the broken line are inside the sample can . 90 (P(4.2K)/PT) (AP/AT) vs T for free hanging, bare, high-purity, thick K samples . . . . . . . . . . . 105 A vs PO for 5 thick K samples cooled in He . . . . 108 G vs T for free hanging, bare, high-purity, thick K samples 0 O O O O O O O O O O O O O O O O I O O 109 ix 4-10 4-11 4-12 4-13 4-14 4-15 4-16 4-17 4-18 4-19 4-20 P(4.2K) (AlnP/PT) vs T for thin wires of K cooled in a He atmosphere . . . . . . . . . . . . . . . P(4.2K) (AlnP/AT) vs T for two K samples of Rowlands et a1. (ref. 13) with d=0.08 mm cooled in He atmosphere. For comparison, the solid lines represent data from Fig. 4-4 for K samples having the diameters indicated . . . . . . . . . P(4.2K) (AlnP/AT) vs T for two K-0.08 at.% Rb alloy samples with d=0.25 mm . . . . . . . . . . P(4.2K) (AlnP/AT) vs T for samples indicated in Fig. 4-4 With fitting Curves Of Eqno 402 o o o o P(4.2K) (AlnP/AT) vs T for thin wires of K cooled in an Ar atmosphere or a partial vacuum. For comparison, the solid lines represent data from Fig. 4-4 for K samples having the diameters indicated . . . . . . . . . . . . . . . . . . . 90 vs 1/d for thin K samples . . . . . . . . . . G vs T for thin wires of K cooled in a atmosphere 0 O O O O O O O O O O O O O O O O O O G vs T for thin wires of K cooled in a atmosphere . . . . . . . . . . . . . . . . . . . cooled in a partial G vs T for thin wires of K vacuum 0 O O O O O I O O O S vs T for sample K-1/2H . . . . . . . . . . . . G and G* (=S/LOT) vs T for sample K-l/ZH . . . . L(T)/Lo vs T for sample K-l/ZH . . . . . . . . . (P(4.2K)/P) (APAAT) vs T for the nominal single crystal K sample . . . . . . . . . . . . . . . . (P(4.2K)/PT) (AP/AT) vs T for the nominal single crystal K sample C O O O O O O O O O O O O O O O G vs T for the nominal single crystal K sample . (P(4.2K)/T) (AlnP/AT) vs T for sample K-S and five runs of sample K-PHl . . . . . . . . . . . A version of Fig. 4-19 to a larger scale . . . . Page 115 117 119 126 127 131 133 136 140 142 143 144 147 149 150 153 154 Figure 4-21 4-22 4-23 4-24 4-25 4-26 4-27 4-28 4-29 4-30 4-31 4-32 4-33 4-34 4-35 4-36 4-37 4-38 4-39 4-40 (P(4.2K)/T) (AlnP/AT) vs T for four runs of sample K-PHZ. For comparison, the data of samples 2a, 2b, and 2c of van Kempen et al. are indicated 0 O O O O O O O O O O O O O I O O O O O A version of Fig. 4-21 to a larger scale . . . . . (9(4.2K)/T) (AlnP/AT) vs T for five runs of sample K-PAl O O O O I O O O O O O O O O O O O O O A version of Fig. 4-23 to a larger scale . . . . . (P(4.2K)/T) (AlnP/AT) vs T for five runs of sample K-PAZ o o o o o o o o o o o o o o o o o o o A version of Fig. 4-25 to a larger scale . . . . . (P(4.2K)/T) (AlnP/AT) vs T"2 for samples K-PHl and K-PHZ o o o o o o o o o o o o o o o o o o o o (P(4.2K)/T) (AlnP/AT) vs T"2 for sample K-PAl . . (P(4.2K)/T) (AlnP/AT) vs T‘2 for sample K-PAZ . . (P(4.2K)/T) (AlnP/AT) vs T for K samples in contact with other plastics . . . . . . . . . . . G VS T for sample K-PHl o o o o o o o o o o o o o G vs T for sample K-PHZ . . . . . . . . . . . . . G vs T for samples K-PHl, K-PHZ, K-PAl, and K-PA2 with fitting curves at temperature below 1K . . . p0 VS C for K-Rb samples 0 o o o o o o o o o o o o (P(4.2K)/T) (AlnP/AT) vs T for dilute K—Rb alloy samples 0 O C O O O I O O I O O O O O O O O O O 0 Normalized AP/AT vs T for K-Rb alloy samples . . . Three different trial fittings for (P(4.2K)/T) (AlnP/AT) of two K-9.4 at.% Rb alloy samples . . . C vs T for K-Rb samples 0 o o o o o o o o o o o o (P /T) (AlnP/AT) vs T for free hanging, bare, thick, high-purity Na samples . . . . . . . . . . xi Page 155 156 157 158 159 160 162 163 164 167 169 170 171 173 176 178 180 182 186 189 Figure Page 4-41 (P(4.2K)/T) (AlnP/AT) vs T for Rb, Li, and K samples 0 O O O O I O C C C O C O O Q C C O O O O 191 4-42 P(4.2K) (AlnR/AT) vs T for Rb, Li, K, and Na samples from 0.07K to 1.4K. The data are the same as in Fig. 3-40 and Fig. 4-41. The dashed lines indicate T resistivity variations inferred from all of the available data . . . . . 194 4-43 Trial fittings for P(4.2K)A1nP/AT of Li samples . 196 4-44 Trial fittings for (9(4.2K)/T) (AlnP/AT) of Li samples 0 O O O O O O O O O I O I '0 O I O O O O O 197 4-45 G vs T for thick Na samples . . . . . . . . . . . 202 4-46 G vs T for Li samples . . . . . . . . . . . . . . 204 4-47 G vs T for Rb samples . . . . . . . . . . . . . . 205 4-48 (Po/T) (AlnP/AT) vs T for thin Na samples . . . . 207 4-49 (ea/P) (AP/AT) vs T for selected thin Na samples 0 O O O C C I O O O O I O O O O O O C O O 208 4-50 G VS T for thin Na samples 0 o o o o o o o o o o o 210 4-51 (t%/T) (AlnP/AT) vs T for Na samples encased in polyethylene tubes . . . . . . . . . . . . . . 212 4-52 pbAlnP/AT vs T for Na samples encased in polyethylene tubes . . . . . . . . . . . . . . . . 214 xii CHAPTER 1 Introduction This dissertation is a report of experimental studies of electrical transport properties such as electrical resistivity and thermo-electric ratio in the alkali metals K, Li, Na, Rb, and in K—Rb alloys. One of the main purposes of this study is to try to find the concealed reasons for disagreements between experimental results from different groups (see section 1.2.1). Particular attention is concentrated on: 1) a size effect in pure K samples at temperatures below 1K; 2) a contact effect of polyethylene and oil on K at temperatures from above 1K down to 0.1K; 3) anomalous deviations from the simple T2 variation expected for electron-electron scattering in thick, pure, free hanging K, Li, Na, and Rb samples when the temperature is lower than a certain limit for each; and 4) a deviation from the expected T2 behavior in the resistivity of K-Rb alloys at temperatures below 1K. In this introduction, we first explain why people are interested in measurements on the alkali metals, especially potassium, and then review previous work on K, Li, Na, Rb, ‘ and K-Rb and discuss some disagreements among work done previously by different groups. In Chapter 2 we introduce 2 all related theories. In Chapter 3, the experimental techniques of the present study are described. The experimental results and interpretations are presented in Chapter 4. Summary and conclusions are given in Chapter 5. 1.1 ;_1; V a, a P =‘ 3.8 _-0 RA _'0:. "0’ i .E. The alkali metals have been attracting renewed theoretical and experimental interest in the past few decades. These metals are simple monovalent.metals which have bec. lattice structures at room temperature, so that they have nearly spherical Fermi surfaces (to within 0.1% for K according to deA measurements). They have no unfilled d- or f- shells to complicate calculations. No superconducting phenomena and Kondo effects at low tempera- tures have been previously seen in the alkali metals. Potassium has been studied most often among the alkali metals. The reasons are: 1) Unlike lithium and sodium, potassium does not undergo a martensitic phase transformation at low temperatures. It thus keeps its bxxc. structure and spherical Fermi surface. 2) Potassium is less reactive than rubidium and cesium and thus easier to work with. 3) Potassium is softer than lithium and sodium, and can be easily extruded through small dies into thin wires. 3 4) Overhauser has claimed that potassium is the best material for testing the existence of charge density waves (CDW) (ref. 1) in simple metals. If found to exist in K, CDW theory would change our basic understanding of electronic transport in metals. In all studies of electronic transport, the fundamental bases are the microscopic transport equations, 3.. =’- Ell-E. + L12VT (1‘1) 2'5 = RAE + 322% (1-2) where 3 is the electrical current density, 3 is the heat flow current density, E is the electric field, and‘aT is the temperature gradient. In general, all four 3 coefficients are tensors, but in cubic metals, such as alkali metals under zero magnetic field, the L's are scalars. Then we can rewrite (l-l) and (1-2) as E = p§+s‘v"r (1-1') 5 = 3p-S-E-K3T <1-2') The electrical resistivity is defined as a (1‘3) ' 'VT = 0 'b n 'ullml where E and 3 are in the same direction. Considering electron-phonon, electron-electron, and electron-defect scattering, the electrical resistivity at 4 temperatures well below the Debye temperature is expected to have the form (ref. 2) P(T) = Po + Pe_e +"§-ph +’Pg-ph (1-4) Here!% is the residual resistivity caused by elastic scat- tering of electrons by impurities or lattice defects in the metal and is temperature independent in the first approxima- tion; ‘E-e is attributed to electron-electron scattering and is expected to have a temperature dependence of the form ATZ, where A is basically independent of 05; p§_ph is due to electron-phonon scattering normal processes and, according to Bloch-Gruneisen theory, is expected to have the form p§_ph = BTS; and the last term pg-ph' comes from electron- phonon Umklapp scattering processes which, in a metal with a spherical Fermi surface completely within the first Brillioun zone, are expected to lead to the form (ref. 3,4) U n -6*/T pe_ph = CT e (1-5) The thermal conductivity K is defined as s K = - ——- (1-6) VT E-O. We have the Wiedemann-Franz law (ref. 2) K9 _ _ T — L(T) (l 7) When elastic scattering of electrons is dominant, n2 k2 0 3 e2 L(T) = L 2.45 x 10'3v2K'2 (1-8) where L0 is the ideal Lorentz number. The thermopower S is defined as E S = -—— a (1-9) The thermoelectric ratio G is defined as G = % a (1-10) q B=0 In general, , S S = GLT 1.e., G = LT = xp (1-11) Where L is usually temperature dependent. When elastic scattering is dominant, L(T) = Lo and S = GLOT (1—12) In pure potassium this is true only when the temperature is below 1K (according to our present data). At low temperatures, when elastic scattering is dominant, G is expected to have the form (ref. 5,6,7,8) Be"e1"/T - 2 G - G0 + AT + T (1-13) Where Go is attributed to electron diffusion, the second term is due to normal phonon drag, and the last term is an 6 Umklapp phonon drag term appropriate to the alkali metals with b.c.c. lattice structure. 1.2 EBEEIQH$_EQBK 1.2.1 WW Earlier resistivity measurements on the alkali metals have made the electron-phonon scattering contribution to the electrical resistivities of the alkali metals, including the contribution of normal and Umklapp processes, well understood (ref. 9,10,11,12,13). Since van Kempen et al. developed high-precision (1 ppm) measurements of the electrical resistivity at low temperatures (1K and above) in 1976, (ref. 14) the electron- electron scattering contribution to the electrical resistivity of simple metals, and deviations from this contribution under various conditions, have been a subject of interest. Four groups in the world have made high precision measurements of the electrical resistivities of alkali metals in the last few years. They are the Wyder, van Kempen group in the Netherlands, the Greenfield group in Israel, the Woods and Rowlands group in Canada, and our group at Michigan State University (MSU). This previous work is briefly described as follows: 7 1.2.1.1 W W In 1976 van Kempen et al. measured the resistivity of high-purity potassium (purity 99.97%) samples each clad in a l-m-long polyethylene tube with an inner diameter of 0.9 mm. They reported that from 1.7SK down to 1.1K, their data could be fit to the formula p a po + ATS + BTn e-9*/T (1-14) with n a l, 9* = 19.9 i 0.2K, and s = between 1 and 2. Using 3 a 2, as expected for electron-electron scattering, their coefficients, A varied from sample to sample by a factor of as much as 3.6, in contradiction to the standard theory. They also reported that the residual resistance ratio [RRR = R(300K)/R(OK)] of the samples changed after annealing the samples at room temperature. One of their samples originally had RRR a 3000, then changed to 6300 after two days annealing at room temperature in a He atmosphere, and finally changed to 8100 after 80 days annealing at room temperature in vacuum. The coefficient A decreased as the RRR increased (ref. 14). Following on the heels of the results of van Kempen et al., a theory of electron-electron scattering in the presence of anisotropic scatterers such as dislocations was prOposed by M. Kaveh and N. Wiser (ref. 15) to explain the sample-dependence of A in P(T) = ATZ. In 1978, Rowlands et al. (ref. 16) measured the resistivity of bare, high-purity potassium samples, 1.8 m 8 long with diameter of 0.79 mm, which were wrapped into a helical groove on a 3-cm-diam Teflon cylinder under a helium atmosphere. They reported that the RRR of their samples increased upon annealing at room temperature (e.g., from 1300 to 4800 in 3 weeks), and that the best fit of their data from 0.5K to 1.3K was pm = 3133/2 with B = (86 1». 10) x 10'6 Po K'3/2. However, P(T)op us on 0» csop xmm.ove how 0 o nnnnnnnnnnnnnnn 3:0 to?“ an angle: .6qu nuclei ”muoz Swans): 3:50 9333. 215a rm; mmsmma: 2 icing... a Gaufieixmes manages can; u: m s 2: vs; 2 u at ,2 n a: on mm.oua now 0» xm.ou9 how xe.H o Mane how xv.H om Mane how panama u< can mm mm c« when was» mama one» mama oocmum a“ when Imnuoaaom ca Iwnuowaom c“ Issoufiu “soc: o.~ . v.Hv soc: m.m u m.H soc: e.m u m.o soc: m.mH . mm.o xx~.¢av 3a OOHm I comm come I coma ooomfla comm I can mam 5.: Com HO Mod fl 0 55 Mac n U SE CoH n U SE moo fl. my “@UQEMfiU onEmm m: can sacs m: can xm.o ma o>onm ad m>obm cowmou 9 5mm H.o 5mm o.H 5mm o.H 5mm H.c :ofimaooum Aasouo amzc mucuazom .Hm um .Hm um .Hm no mom can mnooz uaoqmcoouu comsom cm> .Eaammcuom so xuos msofi>oum .HIH manna 11 cause inelastic scattering of electrons. His deformed K samples showed complete recovery upon annealing at tempera- tures above 165K. 1.2.1.2 W 111.915. In 1980, C. W. Lee et a1. (ref. 23) reported resistivity measurements of bare, free-hanging K-Rb alloy samples with nominal concentrations of 2.24, 0.83, 0.32, 0.13, and 0.05 At.% Rb from 180mK to 4K. The samples were about 4cm long with d = 3.0 mm. They claimed that below 1K, dP/PdT = (mo/(Do + 2Ai)T, (i.e., p= p0 + (A0 + Aipo)T2, with A1 = (8.5 i 0.3) x 10"5/K2 and A0 = (2.2 i 0.31) x 10'13 52cm/K2. The AC was consistent with the A they found in pure K, and the Ai was comparable to the theoretical values of 13.7 x 10"5/x2 from P. L. Taylor (ref. 24) and 12.5 x 10"5/K2 from Kus and D. W. Taylor (ref. 25). 1.2.1.3 W In 1979, Greenfield et a1. (ref. 18) reported resistivity measurements of high-purity Na samples clad in polyethylene tubes with d = 1.0 mm. As they did for their K samples, they plotted their Na data as (P-Po)/'r2 versus T with an adjustable parameter Po, and claimed (KT) = AT2 was found from 1.1K to 2.1K. The RRR of their samples increased slightly with sample annealing time at room temperature; in 9 days Po changed from 0.916 n cm to 0.787 chm, while the coefficient A changed form 0.195 to 0.180 chm/Kz. They claimed that A was a linear function of (Pd/Po)2, in accord 12 with the theory of Kaveh and Wiser (ref. 15). The lowest temperature they reached was 1.1K; the sensitivity they achieved was one to few ppm. 1.2.1.4 Ere1i9us_work_on_resistixitx.9f_hi In 1971, G. Krill (ref. 26) reported electrical resistivity measurements of a high-purity Li sample with RRR a 7000 and Po = 7.30 chm from 40K down to 1.3K. His sample was 50 cm long with d = 0.5 mm. He found PCT) « T4 at 10K0. But below 0.2K, their data start turning down from the fit, and can be fitted better by G = G0 + 8*T2 + A*T, with Go<0, B*<0, and A*>0. 1.3 EBESENI_IHEELE 1.3.1 W As noted in the introduction above, at low enough temperatures, the electrical resistivity P(T) in metals is predicted to vary as T2 due to electron-electron scattering. It is predicted (ref. 21) that such behavior would be seen in alkali metals such as K when the temperature is lower than about l-2K. However, experimental results on potassium below 1.5K by various groups showed discrepancies with this simple theory, as well as disagreements with each other, as described above in section 1.2.1. In order to find the reasons for these discrepancies and disagreements, in this thesis more than 100 measurements of dP/dT were made on alkali metals such as K, Na, Li, Rb, and on K-Rb alloys, from 4.2K to about 0.07K with 0.1 ppm - 0.01 ppm precision. 15 1. To study the deviation from T2 behavior found by Lee et a1. below 0.35K for pure, thick K samples cooled in Ar gas, as described above in section 1.2.1, we used im- proved techniques to measure 14 free-hanging, bare, pure K samples, with d = 1.5 and 2.0 mm, under Ar gas, He gas, or partial vacuum, respectively. The improved techniques in- cluded a better glove box, an improved sample can with an external thermal radiation shield, and a computer averaging technique for data taking. 2. To investigate the approximately T3/2 behavior reported by Rowlands et al. for pure K samples of d = 0.79 mm cooled in the gas and measured from 1.4K to 0.5K, as described in section 1.2.1, sets of pure samples with d ranging from 1.5 down to 0.09 mm were cooled in He gas, Ar gas, or partial vacuum and measured from 1.8K down to 0.07K. An unusual size-effect was found, which lead to a negative dP/dT in very thin samples. To test whether the electron mean-free-path was an important parameter in this size- effect, measurements of dR/dT were made on two K-0.08% Rb samples with d a 0.25 mm, which had a much smaller electron mean-free-path (A.z 0.04 mm) than the pure bulk K samples (A.= 0.2 mm). In an attempt to get a longer electron mean- free-path, so as to see whether the size-effect could be observed in a thicker sample, we tried to make and measure a single crystal K sample with d = 1.5 mm. To test for effects of surface corrosion, two different thicknesses of 16 samples were allowed to thin by means of such corrosion, which occurs naturally inside the sample can. 3. To investigate further the variations in magnitude of the T2 coefficient reported by van Kempen et a1. and Levy et al. for pure thick K samples clad in polyethylene tubes and measured down to about 1K, and described in section 1.2.1., we measured 4 pure K samples clad in polyethylene tubes with d a 1.6 or 0.9 mm down to 0.07K. To test for annealing effects, after each measurement, each sample was allowed to anneal for varying periods of time at room tem- perature under Ar gas, He gas, or partial vacuum, respec- tively. Samples were remeasured after being annealed for a few days, a few weeks, and a few months. Anomalous behavior in dP/dT which might be due to a Kondo effect was seen. To test whether the contact with polyethylene is an important condition for this anomalous behavior, we measured pure K samples in contact with other materials such as Teflon, Kel- F, and paraffin oil. 4. To investigate further the resistivity P of K-Rb alloys, with improved techniques, ‘we made and measured K - 0.08 at.% Rb and K - 9.4 at.% Rb samples from 4.2K down to about 0.07K. 5. Previous work on Li and Na at temperatures down to about 1K by Greenfield group, as described in section 1.2.1, showed T2 behavior in P(T) from 4.2K down to 1.1K for Li, and from 2.1K down to 1.1K for Na. No previous work on p(T) of Rb has been reported. In order to see whether there are, 17 as in K, anomalous deviations from T2 behavior of P at lower temperatures in Li, Na, and Rb, we extended resistivity measurements of all these three pure metals down to about 0.07K. We measured thin high-purity Na samples and high- purity Na samples encased in polyethylene tubes to test whether the size-effects and polyethylene effects seen in pure K can also be seen in pure Na. 1.3.2 W About twenty years ago, MacDonald et a1. (ref. 5,6,7) measured the thermal e.m.f. of pure K from 3K down to 0.1K and deduced the thermopower S from their data. More recently, Lee et a1. made G measurements on pure K and K-Rb alloys from 4.2K down to 0.07K, and M. Haerle et a1. made G measurement on deformed pure K samples from 4.2K down to 0.07K, as described in section 1.2.2. We measured G for most of the samples mentioned above concurrently with dP/dT; (i.e., we extended G measurements to thin samples of K, Na, and K-Rb, thick K samples in contact with plastic and oil, and thick samples of Li, Rb, and Na from 4.2K down to 0.07K). In order to find the low temperature limit of L(T) a Lo (equation 1-8) and thus 8 = GLOT (equation 1-12), and to find the form of L(T) in K, the thermOpower S was concurrently measured with G in one pure K sample with d = 0.5 mm. CHAPTER 2 Theory In this chapter the theory of electronic transport, especially for alkali metals, is reviewed. First, the basic "standard model" of the electrical resistivity for alkali metals is described. Second, since many measurements on many different topics have been done, and various deviations from the standard model have been seen, it is necessary to review quite a few models which might be candidates for explaining our data. Some of these models were proposed to explain previous work on the resistivities of alkali metals. Some were previously used for other metals. Some are candidates for only one topic, some for more than one. The candidates for explaining deviations from the standard theory in our bulk pure samples are: 1) anisotropic electron-electron scattering; 2) electron-phason scattering based on CDW theory; 3) inelastic scattering of electrons due to low energy excitations associated with dislocations; and 4) inelastic scattering of electrons by impurities. The candidates for explaining deviations in our thin samples are: 1) the Knudsen flow model; 2) electron-phason 18 l9 scattering; 3) size-dependent DMR; and 4) the Gurzhi theory. The candidates for explaining deviations in alloys and in pure samples in contact with hydrocarbons are: l) the Kondo effect; 2) inelastic scattering of electrons by impurities; 3) localization effect; and 4) electron-electron interaction effects. These models are introduced and discussed separately. Third, the theory of thermopower S and thermoelectric ratio G is introduced and discussed. 2.1 40.41. ..o: 0 METALS ’0 I! p. i ‘ O .00 4 p )0 When an electric potential is applied across a piece of metal, the conduction electrons in the metal will drift in a definite direction to form a current. Any mechanism of electron scattering that reduces the total crystal momentum of the electrons will cause resistance. The standard theory considers electron and phonon systems inside a crystal with impurities and defects. So the mechanisms contributing to the resistivity include: electron-phonon scattering, electron-electron scattering, electron-impurity scattering, and electron-defect (including electron-surface) scattering. 2.1.1 W (ref. 2,28) There are two different kinds of collisions between electrons and phonons: normal and Umklapp. It is helpful to understand these two processes first. When an electron is in the periodic potential of a lattice, the free electron 20 model can no longer be used; hk, where k is the wave vector nhvu) of the electron wave function, is not a real momentum. hk acts, however, rather like a real momentum, and is thus called the crystal momentum of the electron. Strictly speaking, phonons do not carry momenta, because they are associated with the vibrations of whole lattice. But when a phonon interacts with another kind of particle, such as a photon, a neutron, or an electron, it acts like it possesses a momentum hq, called the crystal momentum of the phonon. Here, a is the w.v. of the phonon. When an electron is scattered from a phonon, the total crystal momentum is conserved to within a reciprocal lattice vector G, (i.e., hk+hq+h5=hk' or k'=k+§+§wherekis the wave vector of the incoming electron and k“ the wwv. of the scattered electronh This process is called phonon anni- hilation. In phonon creation we have bi? + hG = hk' + THE or k + G = k" + q. For a normal process, 5 is zero, and there is no crystal momentum exchange between the lattice and the electron. When G is not zero, it is called an Umklapp process, and there is crystal momentum exchange between the electron and the lattice. The alkali metals in the bcc structure have nearly spherical Fermi surfaces, completely inside the first Brillouin zone. Figure 2.1 shows schemati- cally the normal and Umklapp processes of phonon annihila- tion and creation in the alkali metals with bcc structure, where the Brillouin zone has been simplified. Clearly, for Fermi . . surface\‘* Simplified M Brillouin zone k" k (A) Phonon Annihilation kF/OM =O.877 for the alkali metals in bcc structure Normal process: 'k + a = k' Umklapp process: a. k + a +'§ =‘F' Normal process: 72 1.12.4.3 37L Q ll .. k . 1;::::::::::> (B) Phonon Creation Umklapp process: A A k + G =‘k' +‘a Fig.12-l. Normal and Umklapp processes of e-ph scattering. 22 an Umklapp process to take place, a minimum value of phonon w.v. (qmin) is required. For electron-phonon scattering, assuming: 1) thermal equilibrium phonon distribution; 2) the relaxation time approximation; 3) only normal processes; and 4) the Debye model of phonons, Bloch (ref. 28) predicted that, pe-ph = const (T/ea)5 j5(6d/T), (2-1) where: 99/4 jn =J£ zn (ez-l)'1 (1-e'z)’1dz (2-2) and.6D is the Debye temperature. At high temperatures, T>>Gb, j5~(6b/T)4, so Pe_ph«(T/eb); at low temperatures, T< Pg-e(T) = AUT2 isotropic limit . (2-10) -e(T) __>‘,N,ani(T) +'PUe- e(T) anisotropic limit ~ pNLani(T) = ANTZ (2-11) Finally, they predicted the coefficient of the e-e T2 term to be A = AU + AN with A“ a: (pod/po)2 = (1 + Poi/Paal'2 (2-12) where P61 and pod are the residual resistivity contributions due to impurities and dislocations, respectively. This model was able to describe the sample dependence of the coefficient A for K data of van Kempen et al. of Levy et al., and of Rowlands et al. The main difficulty with this model is that the re- quired number of dislocations in the K samples are at least two orders of magnitude more than is believed could easily exist in strain free K samples. Moreover, there was no way to determine poi' pod' and their ratio, so that one 29 adjustable parameter was needed to make each set of data fit the theory. 2.2.2 W E E 'II' E 'Iill'IlE'] I‘ .As mentioned in section lu2.1, M. Haerle (ref. 22) used theories of inelastic scattering of electrons due to low energy excitations associated with dislocations to explain anomalous deviations at the low T end in bulk potassium samples. In this section, this model is introduced. Two kinds of low energy excitations are discussed; one is asso- ciated with local electronic states, the other with local phonon modes. Theoretical and experimental work on understanding the contribution of dislocations and grain boundaries to the residual resistivity in metals was summarized by Brown in 1981 (ref. 29). He suggested that additional bound states for electrons with energy slightly larger than the Fermi energy exist near the cores of dislocations. He estimated the energy of these additional levels to be above the Fermi level by e which is a few meV for common metals such as c0pper, gold, aluminum, silver, molybdenum, tungsten and zinc, and about 0.1 meV for potassium. Earlier, Gantmakher and Kulesko (ref. 30) had derived an equation for the addi- tional resistivity due to changes in the elastic scattering cross-section because of filling of electron levels local- ized at dislocations as pe-dis = all + flexp(e/kT)1'1 (2-13) 30 where p is the spin degeneracy of the additional level andc: is a pr0portionality constant. Fulde and Peschel (ref. 31) calculated the resistivity contribution of inelastic scat- tering of electrons off of a localized energy level. Their theory could be applied to scattering off local electronic states caused by dislocations such as the ones predicted by Brown. They obtained yzé/T Sinh(6/T)] (2'14) 0(T) = ao[1 - They compared the above result with the corresponding expression one would obtain by applying Matthiessen's rule, 2 aM(T) = ao[l + 7 1'1 (2-15) 1 + 2/3 Sinh2 6/2T where do is the residual conductivity, 6 is the energy separation between the associated two states, and )’is a constant. These two functions have the same limiting values for T<<6 and T>>6. In another mechanism, inelastic scattering of electrons off of dislocations is associated with local phonon modes. Anderson and O'Hara worked on the lattice thermal conductivity in undeformed superconducting Al and deformed superconducting Pb and Ta, and claimed that their data showed resonant phonon absorption at certain energies. Their results were in good agreement with calculations based on the elastic-string model of localized phonon modes associated with dislocations (ref. 32). The elastic-string 31 model of dislocations considers a dislocation which is anchored at both ends. Since there is an elastic energy associated with unit length of dislocation, there will always be a tendency for a dislocation to make itself as short as possible so as to minimize this energy, like an elastic string does. In this model, additional local phonon modes can come from the oscillation of the elastic disloca- tion string; the longest wave length, and thus the lowest energy state, are associated with the length of the disloca- tion. The anchoring sources can be impurities, nodes in a dislocation network, or some other crystal defect. Adding more impurities and dislocations will shorten the length of dislocations, and thus increase the.lowest energy of this system. ‘The random placement of the dislocations and im- purities would give a random distribution of lengths of the anchored dislocations. The scattering of electrons from these local modes should be inelastic. Approximating the local modes by a single frequency oscillator, Gantmakher and Kulesko (ref. 30) calculated the contribution of such inelastic scattering to the resistivity as of the form Pe-dis = (C/4T) Sinh’2(hw/2kT) (2-16) where a2is the frequency of the ground state and C is a constant. 32 2.2.3 Wm An electron-phason scattering mechanism from CDW theory, first proposed by Boriack and Overhauser (ref. 33) and later improved by Bishop and Overhauser (ref. 34), was proposed to explain the T3/2 variation in the resistivity of potassium, reported by Rowlands et a1. (ref 16). According to Overhauser (ref. 1), in a metal for which the positive-ion lattice closely approximates the deformable jellium model, the electronic ground state is a charge density wave (CDW) state. The spatial density of conductive electrons is not uniform throughout the crystal, but is of the form P(F) = -en[1-P cos(5°r +¢)] (2-17) where n is the density of the electron gas, and P, 6, and.¢ are the amplitude, wave vector, and phase of the electronic CDW. In potassium Q is claimed to be slightly greater than 2kF (ref. 1). In order to neutralize this charge modulation, the lattice tends to be deformed. Potassium might undergo this deformation easily because of its softness. This deforma- tion of the ionic lattice is given by -. 6(E) = A sin(6°fi +95) (2-18) with parameters defined above, and [AI = P/IGI. In general, the CDW wave vector 6 is incommensurate with the reciprocal lattice vector G. Thus the lattice 33 should have an extra one dimensional periodicity along the direction of 6. This extra periodicity introduces addi- tional Bragg planes, and thus, extra Brillouin zone bounda- ries. In potassium, Q is presumed to be slightly greater than 2kF, so the first extra zone boundary is very close to the Fermi surface, and the spherical Fermi surface of potas- sium is changed to a lemon shaped surface that just touches the extra zone boundary (Figure 2-2). One of the consequences of this modification is that in the region of contact points between two Fermi surfaces in neighboring new Brillouin zones, the minimum wave vector of phonons joining electron-phonon Umklapp scattering is reduced (Figure 2-3), thus the scattering is enhanced, and its contribution to the electrical resistivity increases. In addition, electron-electron Umklapp scattering can occur not only through E1 + k2 = k} + k3 + 5, but also through k1 + k2 = ki + k} + 5, (i.e., the CDW enhances this scattering and its contribution to the resistivity). According to Overhauser, a CDW crystal is typically divided into 5 domains. The direction of 6, along with the domain boundaries, is very unstable at high temperature, but is frozen at low temperatures. Overhauser predicted that the enhanced electron-phonon Umklapp scattering along the direction of 5 would cause an anisotropy of the electrical resistivity of potassium at low temperature as large as four to one (ref. 34). For wire samples, the orientation of the 5 domains is sample-dependent, so the electron-electron Fig: 2-2. ii " 0 ll .—-.——.-’-- New Fermi ;//surface \ due to CDW z' perturbation 7\_.- _ -._._._ ————-—-— / New zone boundaries ”/r due to CDW perturbation The effect of a CDW perturbation on the Fermi surface. The Sphere is distored to a lemon shape. (After Overhauser (ref.l)) The Umklapp scattering for a distored Fermi surface. (After Overhauser (ref.l)) 35 scattering contribution to the resistivity should also be sample dependent. This can be used to explain the sample dependencies of Po and of A, the coefficient of the T2 term in the resistivity of potassium, reported by van Kempen et al., Levy et a1. and Rowlands et al. 0 Another consequence of a CDW, is phonon-like excitations, called phasons. Without a CDW, phonons arise from thermal vibration of the lattice ions because of the lattice periodicity; Similarly, phasons arise from the new ionic periodicity of CDW, since 6 is incommensurate with G. Boriack and Overhauser (ref. 33) suggested an electron- phason interaction of the form Ve¢ = 1/2 G§¢q {cos[Q + q)°r - wqt] - cos[(5 - ii)? - <0th (2-19) where G cos(6€§ is the total self-consistent potential, and ¢H*‘“§* a are magnitude, frequency, and wave vector of the phason, respectively. With this potential, and assuming the Fermi surface to be rigidly displaced in k-space by the external electric field, Bishop and Overhauser (ref. 34) obtained the electrical resistivity by using the variational method as Pe¢(T) = A(\/‘2' T/6¢)5 J5(6¢/\/‘2' T) + 3(T/9¢)4 J4(eh/T) + C(T/6¢)2 J2(e5/T) (2-20) 36 where x n Jnix) 8] 2 dz 0 (ez-l)(l-e“z) is the Bloch-Gruneisen function. Each one of the three terms given above can fit well to Rowlands' data with different parameters 6¢, but their extensions below 0.5K differ from each other (Figure 2-4). It is important to note that this model cannot lead to a negative dP/dT, since all three terms given above increase monotonically with temperature. This means that this model of electron-phason scattering cannot explain the appearance of a negative dP/dT in our thinnest K samples, which will be analyzed in Chapter 4. 2-2-4 Ih£_KnndS£n_£lQE_MQdfil By analogy with the Knudsen flow (ref. 35) of rarefied gases in a cylindrical tube, Rowlands et al. (ref. 16) -proposed a size effect model associated with the Knudsen flow of electrons to explain their experimental results for K (section 1.2.1). They noted that the relaxation time (T) for electron- electron normal scattering in potassium at 1K from Lawrence and Wilkins' calculation (ref. 21) is about 1.09 x 10-7 sec, and the Fermi velocity (VP) of potassium is about 8.6 x 105 m/sec. Thus, the mean free path (A) of this scattering is about 9 cm. Comparing A with the radius (r) of their samples (r = 0.4 mm), they argued that most collisions of 250 l zoo—r RESISTIVITY (|O"50hm-cm) Fig. Z-Q, (50- '00" 0.2 0.4 37 \T|.5 I i l I l l 0.0 0.5 LO (.5 TEMPERATURE Plot of resistivity vs T for the data of sample KZC of Rowlands et al. (ref.l3). The fitting curve with Jh is shown. In the inset, extrapolations of TI'S, J2, J4, J5 fits to lower temperatures are shown. (After BishOp and Overhauser (ref.l7)) 38 electrons below 1K were with boundaries, and thus the Knudsen flow model can be used. Assuming diffusely scattering walls they proposed an appropriate formula for the resistance to the electron "Knudsen flow" as p = po (1 + kr/A) (2-21) where k is about 1 for gasses (found by Dushman) (ref. 36), for phonons in liquid helium (found by Whitworth) (ref. 37), and for electrons (found later by Black in a Monte Carlo calculation (ref. 38)), For electron-electron scattering, AaT'Z, so the temperature-dependent resistivity should be _ _ 2 - with Borpor Black (ref. 38) pointed out later that in the Knudsen case, Po is proportional to 1/r, and hence, the coefficient B is independent of the residual resistivity Po. Black also argued that if P0 is something other than the size—limited resistivity, as suggested by Rowlands et al., then B a pot could be valid. We noticed that r/A a 0.005 at 1K in Rowlands case, and if k ~ 1, as calculated by Black, the PT(1K)/Po is about 5 x 10'3. But the data of van Kempen et al., of C. W. Lee et ale, and even of Rowlands et al. themselves, showed 39 P'T(1K)/Po to be about 2 x 10'4. Even considering a possible smaller effective radius due to additional scattering from other scatterers, as suggested by Rowlands et al. for their data, the predicted ratio PT(1K)/Po is still about an order of magnitude too big. Alternatively, according to Kukkonen and Smith (ref. 39) 7'could be an order of magnitude smaller. In such a case, Rowlands et al. proposed that a term in (r/A)2 ought to be taken into account to give 9T = BT2 - CT4 (2-23) In this case, the ratio PT(1K)/Po is even two orders of magnitude too big compared to experimental results- Starting from a force-force correlation function formula for the residual resistivity, March and Woods provided a possible alternative justification for equation (2-21) above (ref. 40). In their form p.-.p (1+£+ ) (2-24) 0 A C O C where A? is found to reflect long-range correlations in the electronic motions and not necessarily to be size limited. A.is electron-electron scattering mean-free-path in low temperature resistivity study of potassium. In explaining data of the potassium wire samples from Rowlands et al., they suggested A" ~ 1/2 mm. So, the PT(1K)/PO in their model is still an order of magnitude too big. 40 As mentioned above, we need a model to explain the negative dP/dT in our thinnest K samples. In the two models described in this section, only the second term in equation (2-23) seems to be able to generate a negative dP/dT. But when sample radius r is smaller, huQJz is less important, so -CT4 term is less important. So if for bigger r we do not see a negative dP/dT, (idh, with our thick samples), then for thinner samples this equation cannot lead to a negative dP/dT. 2.2.5 W In this section we first describe two size-dependent DMRLmodels, which were previously used for explaining size- effects in other metals. The possibility for using each of these models to explain the negative dP/dT found in our thinnest K samples is discussed. The well known Matthiessenfs rule (MR) can be written in the form Pa(c.T) = Pp(T) + 90(c) (2-25) where Pa(c,T) is the resistivity of a dilute alloy contain- ing a concentration c of impurity, Pp(T) is the resistivity of an ideally pure metal, and Po(c) is the impurity produced resistivity at zero K. In fact, there are always deviations from MR. Pa(c,T) can be written exactly as Pa(C.T) = Pp(T) + 90(0) + A(C,T) (2-26) 41 For 'normal' size dependent DMR, induced by a spatial variation in the distribution function fk(f) due to the existence of boundaries, Dingle (1950) (ref. 41) developed a theory for metal wires of circular cross section. Dingle assumed that the electrons in the metal were free-electron- like with a spherical Fermi surface and that the surface scattering might be characterized by a single specularity parameter, p (defined as the probability that an electron incident upon the surface is reflected specularly as though from a mirror) which was assumed to be independent of the incident angle. A few years later, Ziman (ref. 2) (1960) formulated an angle-dependent specularity parameter Ps(0) = expl-(4nh/Ae)2coszel, by assuming that all electrons have the same wavelength Xe and that the surface can be charac- terized by a root mean square surface roughness h. More recently this 95(0) was incorporated into Dinglefls model by Sambles et a1. (ref. 42), obtaining an expression for the ratio of bulk to thin wire resistivity, pa, 12 ”/2 2 2 ”72 - l — -—' cos 95in OdOJf sin¢d¢ P flk o o (l—p)[1-exp(-ksinw/sin9)] 1-p.exp(-ksinw/sin0) (2-27) where k d/Awand p = expl-(llrrh/Ae)2 sin20 sinzw]. Since Rn/P is a function of only k, Sambles defined f(k) = (fa—poo )k/Poo = (p/poo - l)k. Thus P-Poo= f(k)P°° /k = f(k)° {LAW/d. To determine the sign of 6P/6T, Sambles (ref. 43) 42 argued that 23:3. - 6p 6T 6T (’0 pm) +5? BmQ-f-é—kl- + % (since Mis a constant) P A 6f(k) 6k 6p . . . _emaa_____.. __ m d 6k 8T + 6T (Since f(k) is a function of only k) - if. he, - , 9. = an... — (l + 6k) 6T (Since k A“, Ankw thus Bk/d‘l‘ = (ti/Pack...) (GAO/6T) (2-28) Finally, Sambles argued that in their model, 6f/ak cannot be smaller than -1, dP/OT cannot be negative unless ago/am is negative (ref. 43). Boughton and Neighbor (1972) (ref. 44) calculated size- dependent DMR under several different situations. For wires of circular cross section, their calculation led to two different results: a) For metals with cylindrical Fermi surface, they obtained the ratio of thin wire to bulk resistivity .312 [ 16k P .. _ iii-1 _ P..' 0.590 ln3fl] (2 29) as a function of only k, with k = d/Aq” This gave f(k) - (——p 1) k - 16 [0.590 1n3 ] k (2-30) and 43 6f ar’l with the same argument for equation (2-28), one can conclude that aP/BT cannot be negative unless dam/6T is negative. in ,For metals with spherical Fermi surface, they obtained p 1 3 3n -—-= —-+ — -—-- - - 1 D; it 8 [1n 8 lnk IE] (2 3 ) where I; is a constant, for wires of circular cross-section I; < -1 can be obtained from an equation given by them (ref. 44). This gave 3k 3n f(k) = 1 +-§- [ln§—-- lnk - [g] - k (2-32) Thus, 62:“ > and, g; > 0 (as long as ggé > 0). 2.2.6 W (ref. 45) None of the theories associated with size-effect discussed above can provide an explanation of negative dP/dT in our data of thinnest potassium. In this section, the only published candidate which might provide such an expla- nation is introduced. 44 In 1962, Gurzhi predicted that normal electron-electron scattering (NEES) should affect the path length between two sequential electron-surface collisions in impurity-free conductors and thus cause a negative temperature derivative of resistivity (dP/dT < 0). Gurzhi argued that a) The frequency of electron- electron scattering decreases with temperature more slowly than the frequency of the electron-phonon scattering and will be dominant at low temperatures. b) In a bulk conductor, electron-electron scattering gives rise by itself to electrical resistivity only through Umklapp processes, NEES makes no contribution. c) When 1V>>d, where 1V is the mean free path connected with the scattering by the inhomogeneities such as phonons, impurities and internal defects, and d is the sample diameter, the effective mean free path with diffusely scattering walls should be of the form leff ~ d if the effect of NEES is not taken into account. But, when d>>1e-e (le-e = the mean free path of NEES), before an electron situated deep in the sample reaches the walls, it would be scattered by other electrons many times and thus would move like a Brownian particle. According to the formula of Brownian motion, the length of the path length between two collisions with the wall is of order d2/1e_e, i.e., leff ~ d2/1e_e. But 1/1 or l/‘re_eat e-e T2, so Pal/leffale_e/d2aT'2. This gives us dP/ch1-T‘3, a negative derivative! According to the Gurzhi theory, a negative temperature 45 derivative of resistivity would be expected in a high-purity thin sample with le_e<>6D and Z is the excess charge of the impurity ion. Considering that the displacement of the impurity potential induced by the thermal motion of the impurity ion leads to a reduction of elastic as well as inelastic scattering of electrons, Taylor (ref. 24) pointed out that this reduced the additional resistivity due to inelastic scattering. In 1963, Koshino (ref. 48) argued that this reduction cancels out the additional resistivity due to inelastic scattering at high temperatures, but brings no essential change in the additional resistivity at low temperatures. Taylor (1964) (ref. 24) recalculated the 47 additional resistivity at low temperatures and obtained an expression without z as ”ZhZKZ _ av 2 _ -5 2 - ‘Hnel - -——-§-POT - 1.37 x 10 PCT (2 34) 2MkBGD where Kav is an averaged wave vector close to 2kF (kg = Fermi wave vector), M the host ionic mass, 6D the Debye temperature, and k8 the Boltzmann's constant. Kus and Taylor (ref. 25) obtained for K-Rb _ -5 2 _ pinel - 1.25 x 10 POT (2 35) Frobose and Jackle (ref. 49) derived the same formula as equation (2-34) for structure defect inelastic scattering, and calculated pinel for K and obtained p. - 1 8 x 10"5 p T2 (2-36) inel ‘ ' o C. W. Lee et al. (ref. 23) reported that in K-Rb alloy resistivity measurements they obtained p = p0 + (8.510.3l x 10‘6pr2 + (2-2i0-3) x 10’13T2: and claimed the second term, (8510.3) x 10‘5po'r2, as an unambiguous Koshino—Taylor term. 2.2.9 WWW Effects In this section, two different theories of the temperature dependence of the resistivity in impure metals with nonmagnetic impurities are given. 48 The first is the theory of electron-electron interac- tion effects. A square-root temperature dependence of resistivity due to the electron-electron interaction effects was first ob- tained by Altshuler et a1. (ref 50). They argued that electron-electron interaction accompanied by the impurity scattering leads to a singularity in the density of states near the Fermi surface, and thus leads to a zero bias anoma- ly in the tunneling current-voltage characteristic. They showed that at low temperatures, the resistivity decreases as -x/"l" with increasing T. In 1980, Fleurov et a1. (ref. 51) obtained a more clear picture. They argued that when T150 pacm), for which 16 may be as small as an atomic spacing, the second term in Eqn. 50 (2-38) can be neglected, then, _3 1 1 0(T) :- ._.._.. n2 h Li(T) At low temperature, 1101-) or T‘2 T<6D (2-39) for both electron-electron and electron-phonon inelastic scattering. At high temperature, the electron-phonon scattering leads to -1 11(T) at T T>6D but 2 - Li(T) - 1/2 leli(T) Thus, Howson obtained 0(T) « T T<6D and 0(T) «VT T>9D (2-40) when ab>>a«T), where do is temperature-independent part of conductivity, then one has p = p0 "' CT T<6D and p 00 - CJT T>9D in first approximations, where C ang. (2-41) 51 2.3 W As described in section 1.1, generally there are three terms in S (or G) at low temperatures. The first is the diffusion term. The other two terms are due to normal and Umklapp phonon drag, respectively. In normal phonon drag (described in section 2Jul), electrons tend to flow along with phonons from the hot to the cold end, and thus contribute a negative term in S (or G). In Umklapp phonon drag, electrons tend to flow opposite to the phonon flow, and thus contribute a positive term to S and G. Details of the theory of S and G for Alkali metals have been described by C. W. Lee (ref. 8) and M. Haerle (ref. 22) in their theses. Only the main points of this theory are briefly reviewed in this section. 2.3.1 W At low temperatures, the thermopower S of alkali metals with bec. lattice structure, can be written in the general form 8 = A'T + B'T3 + C'(T) exp(-9*/T). (2-42) The first term (A'T) is the diffusion term, where A' can be written as the generally used Mottls expression (ref. 2): _ fizkz 61na(E) 3e 88 E=Ef . A' = (2-43) 52 Here Ef = the Fermi energy, 0(E) = energy dependent conductivity. Simplifying this general form, Guenault and MacDonald (ref. 5) obtained . nzkz aln[n(E)v2(E)l aln‘r(E) A = " 3 i + -- (2-44) er alnE alnE E=Ef where 7(8) = the relaxation time, n(E) = the density of states, and V(E) = the average electron velocity at energy B. When temperature independent mechanisms such as impurity scattering, dislocation scattering, etc. dominate, a temperature-independent A' is expected from either equation (2-43) or (2-44). And for spherical Fermi surface, as in alkali metals with bxnc. structure, we expect to have a negative A' (ref. 8). Applying Matthiessen's rule P= Pi + Pd to Mott's formula, one gets the Gorter-Nordheim relation where Pi and Si are the electrical resistivity and thermopower, respectively, due to impurity scattering, and Pd and 3d are the electrical resistivity and thermopower, respectively, due to another temperature independent scattering (such as dislocation or surface scattering). A slight change in the concentration of impurities, or in the density of dislocations, can easily change the dominant scattering mechanism and thus change the magnitude of A'. 53 The second term B'T3 is the normal phonon drag term, where B' is a simple constant. At very low temperatures in a sufficiently pure metal, the boundary scattering of phonons will dominate,the contribution of normal phonon drag to S will then go roughly as the lattice specific heat, which has T3 temperature dependence (ref. 8). The third term (C'(T)exp(-6*/T) is the Umklapp phonon drag term. For Umklapp scattering, as described in section 2.1.1, a minimum phonon wave vector (qmin) is needed, and thus the number of phonons available for such events should be pr0portional to exp (-6*/T). Guenault and MacDonald (ref. 7) obtained an approximation for C'(T) at low temperatures as l/r .. c'(T) = [ 9“ e 1 (0.200) [5) [9:1 1/"'ph--e + 1/ Tph-i e an 9* T [5" + 3 + 65;] (2-46) where Tph-e = phonon-electron scattering relaxation time, Tph-i = phonon-impurity scattering relaxatin time, 9D = Debye temperature, and EN = a constant. Ziman (ref. 57) and Bailyn (ref. 58) developed more detailed theories, but so far only the simplest forms for the Umklapp scattering have been used for fitting experimental data. Guenault and MacDonald (ref. 7) claimed that their potassium data could be fitted by 54 s = A'T + B'T3 + c'exp(-e*/T) with C' = a constant and 6*~21K. They found, as expected, that A' and B' were negative while C' was positive. 2.3.2 Wm ‘Very little theory has been worked out for G. ‘But, as described in section 1.1, G may be related to S by S = GL(T)T (1-11) and when elastic scattering of electrons is dominant L(T) = L = 2.45 x 10‘8 V2 K'2 (1-8) 0 According to our present data, L/Lo ~ 0.96 in K at 1K, the ratio L/Lo becoming closer to l at lower temperatures. At very low T, G is thus expected to be I G = G0 + B'T2 + $~exp(-6*/T) (1-13) or even G = G0 + B'T2 (2-47) where G0 = A} is the diffusion term, B'T2 is the normal phonon drag term, and C'/T exp(-6*/T) is the Umklapp phonon drag term. CHAPTER 3 Experimental Techniques 3.1 W In this chapter the basic experimental techniques and equipment, and our samples, are briefly described. Since most of the techniques and equipment have already been described elsewhere (ref. 8,22), only some modifications are given in detail. As mentioned above, one of the main purposes of this study is to try to find the concealed reasons for disagree- ments between experimental results from different groups (section 1.2JJ, and to study various deviations from electron-electron scattering when according to standard theory electron-electron scattering should dominate over electron-phonon scattering at low enough temperature. Thus, it was necessary to get the temperature below 1K. Second, there are two ways to study the temperature dependence of resistivity at low temperature» One is to measure P'direct- 1y. But P = P0 +P (T), and at 1K, P(T)/Po is about 200 ppm for pure potassium of typical RRR~5000 (ref. 14,59). Since Po cannot be directly measured, it must then be left as an adjustable parameter, which produces an inevitable uncer- tainty in P(T). The other way, used in this study, is to 55 56 measure dP/dT. Since PO is a constant, this must be exactly equal to dP(T)/dT with no adjustable parameters. To deter- mine the measured quantity AP/PAT with a precision of 1%, extremely high precision (g 0.1 ppm) of AC/C (=AP/P) was needed, because for our highest purity potassium AP/P z 10"5 with a AT = 0.1K near 1K. 3.1.1 W 1) WWW with good vibration isolation, giving a temperature range from 4.2K to 60 mK. The whole system could be cooled down to liquid nitrogen temperature (~77.4K) in about 12 hours by adding liquid nitrogen to the outer dewar with appropriate exchange gasses, and then cooling further to liquid helium temperature UL2K) by transferring liquid helium into the inner dewar. This transfer took about 2 hours. 1.3K could be obtained at the 1K pot by pumping on the liquid helium. The lowest temperature of 70 mK was reached at the mixing chamber and the attached sample can about 3~5 hours after circulation of the dilution refrigerator was started. That means that all of our samples were cooled slowly; See C.lL Lee's thesis (ref. 8) for details. 2) WWW consisting of a commercial direct-current comparator modified by D. Edmunds et al. (ref. 60) and a very sensitive SQUID (Superconducting Quantum Interference Device) null- detector, capable of detecting voltages less than 10’15V, or current less than 10‘9A. The ratio of two currents (slave 57 current to master current) could be read with a precision of 0.1 ppm using a set of eight decade-dials. Nearly 0.01 ppm could be achieved by interpolating beyond the last dial using a computer averaging technique. 3) A_screened_rggm for screening out radio frequency noise that affects the operation of the SQUID. l) and 2) were enclosed in the screen room, and all the pumps were kept outside (ref. 8). 3.1.2 Thermometers Temperature measurements were very important in this study. Three germanium resistance thermometers GRT2, GRT4, and GRTS were used. The calibration of these three thermometers was done by C. W. Lee and M. L. Haerle (ref. 8,22). GRT4 and GRTS were calibrated simply by carefully comparing them to GRT2. GRTZ itself was calibrated in three steps. First, it was carefully compared to a Cryocal CRSO thermometer, calibrated by J. L. Imes (ref. 61) and G. L. Neiheisel (ref. 62) from 0.065K to 4.2K by using the He3 vapor pressure and a susceptibility thermometer. A fit of the data with the equation, 9 . 1n R = 2 Ai (1nT)1 (3-1) i=0 was used for calculating the temperature from 1.3K to 4.2K. Below 1.3K, this fit was not accurate enough for the present study. Second, using a set of Superconducting Fixed Point Devices SRM 767 and SRM 768 from the National Bureau 58 of Standards (ref. 63) to get 6 absolute temperature points from 0.099K to 1.17K and using a susceptibility thermometer to interpolate between these six points. The fit of this set of data, from 0.059K ot 1.24K, with the above equation (3-1) has been used for measurements below 0.5K. Third, for (LSK to 1.3K calibration, a large powdered sample of cerous magnesium nitrate and four of the fixed points between 0.519K to 3.414K were used to again fit Eqn. (3-1). The accuracy of these calibrations was tested by using the Wiedemann-Franz law and a Ag—0.1 at% Au Alloy. One end of this alloy was attached to the mixing chamber at a fixed temperature, and GRT2 was mounted to the other end for measuring the temperature there. The calibration gave the product TAT to within an average value of 0.47% with a standard deviation of 1.2%. The maximum deviation from the Wiedemann-Franz law was 2.6% (ref. 22). To ensure that there were no problems with temperature measurements while taking the data, the Wiedemann Franz law was also used to double check the thermometry for every run (see section 3.4). The resistances of GRT2 and GRT4 were measured with two SHE conductance bridges in the 4-terminal mode. The bridges, using low excitation voltage (10 to 100 uV), were accurate to better than 0.5%. They were self-balancing and had a differential output, which was also used for regulat- ing the mixing chamber temperature (section 3.5.1). 59 3.1.3 IhWWW The alkali metals are all very reactive. They react rapidly with air and water vapor. Secondly, in this study, samples were measured under different circumstances: under Ar, He, or vacuum, and either bare, or encased in, or touched by, plastic. Therefore, some special equipment was needed for sample preparation. The main equipment included: a) two glove boxes sup- plied by Vacuum Atmospheres Company; b) two different kinds of stainless steel presses used for making wire samples and potential leads; c) two sample cans, used to protect the samples from reacting with air; d) a high vacuum system, with all metal parts, capable of getting a 10'9 torr vacuum, used to pump out the sample can and clean some plastic tubings and pieces, glass tubings and molecular sieves, all needed for sample preparation. The details of the glove boxes, presses, sample cans, sample preparation, measurement method, and uncertainty analysis will be given in what follows. 3.2 IHE_§LQ¥E_BQKES_AND_RBESE§S Two glove boxes supplied by Vacuum Atmospheres Company were used for sample preparation. One had a built-in Pedatrol pressure control, Dri-train Mo 40-1 inert gas purifier with circulation rate of 50 cfm for He, and A0 316- C Oxygen analyzer. It was used under He atmosphere with an oxygen contamination level less than 0.5 ppm. Fresh 60 potassium exposed inside this glove box stayed shiny for at least 2 hours. The other glove box was evacuable and had only a built-in Pedatrol pressure control and locally built inert gas purifier with circulation rate of about 1 cfm. It was used under Ar atmosphere. Fresh potassium exposed inside of it stayed shiny for at least 2 hours also. All the samples were made inside these two glove boxes. Two kinds of stainless steel presses were used. Figure 3-1 shows the structure of these presses. The big one, which is threaded, was much more powerful than the other and was used for making samples which could not easily be extruded, such as Li or thin Na samples. When the big one was used, it was stabilized in a big vice, and a long wrench was used to screw the piston in. There were two ways to put the material into the press. For K, Rb, or Na, which were originally in ampoules, the material was melted in the glove box with a hot-plate and then poured into the press. For Li or Na, which were not originally in ampoules, the surface was first cleaned with a stainless steel knife, and the metal was then cut to reasonable size and put into the press. After the material was poured or put into the press, it was melted and allowed to solidify into one piece. Different sample thicknesses were produced by using stainless steel dies with different hole diameters. 61 _ <—- Piston / GI ‘10. \Q-"\\’I-.§’(\’O-C \Il'l ’0 0|. w>>>~If .u""ilv O ...---..- .. f (o) \It I ‘ \ .- J. s. \c.... o s .. ..I J- \._ I ........ JL. '0” w r . . . . o . . . o 4 0 Q xx presses. Two kinds of stainless steel Fig. 3-1. 62 3.3 SAMELE.CAN Two sample cans were used. Figures 3-2 and 3—3 show the structure of sample can number 1, used for K, Rb, and Na samples. Its sample holder has two separated sides, each with two c0pper supports for current leads, and two more for voltage leads. Sample can number 2, used for Li, is a modified version of sample can number 1. The main differ- ence is that sample can number 2 has 8 copper clamps instead of 8 copper supports, because Li is not soft and thus not easily cold welded to copper. All eight copper supports or clamps were carefully cleaned with alcohol and filed just before being brought into the glove box, and then filed again inside the glove box. Before the samples were made, a little material was extruded out to clean the stainless steel die. Then more material was squeezed out and smeared (cold welded) onto all copper supports or clamps. After a sample was made, it could be mounted onto either side of the sample holder in three different ways (Figure 3-4). In method (a), the sample was first cold welded onto the current supports. Then two potential probes of the same material with the same diameter were extruded and cold welded first to the sample, typically 5 cm apart, and then onto the voltage supports. In method (b), used for mounting thin samples, two thick (d=l.0 or 1.5 mm) leads were made from the same material, and each was cold welded onto a current support and the neighboring voltage support. Then a thin sample (d30.25 mm), 63 .mtouoeoetozu may mc_v_o; not mean—u tmaaou .m. .cmsotcuuoou .ueto;u_toaa0u .s. .mvmo_ m:_uu=tcootoa:m .m— .toumo; u .N. .maota toaa0u .._ .maota co_>z .o. .mtmo. _m_ucouoa ecu mutoaazm chance .m .mpMp. ucottau to» mutoaaam toaa0u .w .omcm_m whatm .m .monau .ooum mmo_c_mum .o .mmnau toaaou .m .x:__ _mEtocu m< .: .mtouoeoetoch .m .consmcu mc_x_E och ea Damagomuum rem mc_t co_mmu tam __oz toaaou .N._ u. ton—o; o_aEmm .NA .3“. 64 .onau Landau o>o_m cm_:oo_oe :u .m .m .oanu mouth po___m e=_pc_ .m .mc_c mmmcm .N .cmo toaaou .mc_mmo oa_a mmmtm .: .consmcu "top—o: o_aEmm as» $0 tm>ou oc__ mc_ae:a 0h a. .mum cm— a 65 ucoctsu Am ®\ll.li .: .moZEmm ucottau tom ass—o to utoaasm toaaou ._ .mm_aEmm .4 com Adv .mppd. _m_ucouoa .m .momo_ _o_ucouoa com asp—o to utoaaam toeaou "mo—aeom mc_uc:oe mo m>m3 ucocomm_p motes .oz oeo .x «o mo_anm c_;u tom Anv om humm— .~ .mdhd. .62 can A .a-m .m_a .em .x mo mo~aEmm xo_;u com ov fillllll'! erliv L 66 typically 0.5-1.0 cm long, was cold welded onto those two thick leads. As described later, for 0.25 mm K samples, the data were same by methods (a) and (b). In method (c), used for Li, the sample and potential leads were first clamped by the copper clamps. The sample was then cold welded to the potential leads by pinching them together with a stainless steel tweezer. Our samples were cooled in one of three different atmospheres: Ar atm., He atm., and partial vacuum (~10qu of He). Under Ar atmosphere, the sample can was simply closed with an indium "O” ring. Under He atmosphere, some cleaned Linde Molecular Sieve Type 13X pellets were left inside the can to adsorb the He gas at low temperature, so as to prevent heat exchange between the samples and the can. The molecular sieve was cleaned by pumping it down to a vacuum of less than 10'6 torr in an Erlenmeyer flask with a stopcock, and then heated to about 200°C under vacuum for two days. The flask was then allowed to cool to room temperature, sealed by closing the stopcock, moved into the glove box, and opened just before closing the sample can. After the pellets were poured into the sieve chamber, a piece of stainless steel screen was pressed in to hold the pellets in place. To evacuate the sample can down to 10 ”Hg, and then seal it properly, a piece of indium-filled brass tube was needed. In making these tubes, first the brass tube was 67 filled with molten indium. After the indium solidified, a hole was drilled through it and the tube was soldered onto the sample can as shown in Figure 3-3. [See M. L. HaerleT; thesis for details (ref. 22).] An extension tube with a ‘valve was soldered to the indium tube before the can was moved into the glove box. The sample can was sealed at room temperature by closing the valve, and then was taken out of the glove box and hooked onto a high vacuum system. In order to prevent the diffusion of the air through the pump to the can, the vacuum system was first completely evacuated, then the input valve of the pump was closed and the valve on the can was opened, this procedure was repeated until the sample can was evacuated properly; typically, about 10 u-Hg helium gas at room temperature was left inside the can as exchange gas for the cooling process. The indium tube was then pinched to seal the can. The indium cold- welded to itself, providing a leak-tight seal at the lowest temperature in this study. .Finally, the valve was removed from the sample can, and the can was attached to the mixing chamber of the dilution refrigerator through a copper well and teflon ring (Fig. 3-2). As noted above, temperature measurements were critical to the results of this study. The two thermometers, GRT2 and GRT4, used for measuring the temperatures of the samples, were clamped with copper clamps onto two c0pper thermal feedthroughs outside the sample can. To ensure that each of the thermometers was at the same temperature as the 68 sample it was measuring: 1) The thermal resistance between the sample and the thermometer was minimized by using a well annealed copper rod, 8 cm long and 3 mm in diameter, as the thermal feedthrough; 2) Each feedthrough was sealed onto a thin wall stainless steel tube using stycast 3850 GT epoxy, and the stainless steel tube was then soldered onto the sample can (Fig. 3-2). The samples were thus thermally isolated from the sample can, because both epoxy and stainless steel are very effective thermal insulators at very low temperatures; 3) The sample can and dilution refrigerator were both thermally isolated from liquid He by enclosing them inside a vacuum can. To cut down thermal radiation from the vacuum can to the thermometer sitting on the sample can, the sample can was shielded by a thin copper can, which was attached to the mixing chamber. 3.4 SAMELE.£BEEABAILQN All of our high-purity potassium (99.95%), rubidium (99.95%) and some of the high-purity sodium (99.95%) were obtained from the Gallery Chemical Company, sealed in glass ampoules under Ar gas. Some high-purity sodium was kindly supplied by Prof. J. C. Garland of Ohio State University, packed under paraffin oil in a plastic bottle. Our high- purity lithium (99.99%) was obtained from Atomergic Chemetals Corp. in a low sodium dry pack. In order to keep the glove box clean, the oil bottle of Na was first moved into a glove bag filled with Ar gas, the 69 Na pieces were taken out of the oil with a clean stainless steel tweezer, soaked in a beaker of petroleum ether for a few minutes to dissolve the oil, dried, sealed in a glass container, and then transferred into the glove box. The common way of preparing bare pure K, Rb, Li, and Na samples was already described in section 3.1. In this section, the details of each sample are given in Table 3-1 (for K), Table 3-2 (for Rb), Table 3-3 (for Li), and Table 3-4 (for Na). The details of making pure K samples encased in plastic tubings, coated by oil, or touched by various kinds of plastic pieces, and the details of attempting to grow a single crystal K sample and of making some K-Rb alloy samples will be given next. In order to make a K sample inside a polyethylene, teflon, or glass tube, an ampoule of K was opened inside a glove box, heated with a hot plate until the potassium melted, and then the molten potassium was sucked up into the tube using a syringe (Figure 3-5). The tubes had been freshly cleaned by placing them in a flask, pumping the flask down to less than 10'”6 torr, and heating the tubes with a heat lamp to above 100°C for two days. In order to determine effects of different plastics, each ampoule of potassium was allowed to come into contact with only one kind of plastic. And, after a tool touched one kind of plastic, it was cleaned by alcohol before being used for dealing with another kind of plastic. The surface of a K sample touching polyethylene was always shiny during all the Table 3-1 70 Characteristices of bare free hanging pure K samples. (RRR : R(295K) K(u.2K)) p(u.2x) Ampoule Circumstance d [Pb] Mount Samples No. at R.T. (mm) (nflcm) RRR method Other K-7 # 1 Ar atm. 1.50 1.41 5100 (a) [1.13] K-8 4 1 Ar atm. 1.50 1.50 4790 (8) [1.22]“ K-9 # 2 Ar atm. 1.50 1.72 4180 (a) [1.441‘ K-lO # 2 Ar atm. 1.50 1.47 4890 (a) [1.1913 x-11 #11 Ar atm. 2.00 1.26 5710 (a) [1.01] K-Hla # 3 He atm. 1.50 1.70 4230 (a) (1.421“ K-Hlb # 3 He atm. 1.50 1.65 4360 (a) Second run [1.371“ of K-Hla K-HZ # A He atm. 1.50 1.27 5660 (a) [0.991“ K-H3 # 5 He atm. 1.50 1.97 3650 (a) [1.591“ K-Hu # 5 He atm. 1.50 1.22 5890 (a) Annealed [0.9“]‘ at 50°C for 30 min. K-H5a # 8 10 qu He 1.50 1.27 5660 (a) [1.02] K-H5b # 8 10 )1Hg He 1.50 1.39 5170 (a) Second run of K-HSa K-H6a # 8 He atm. 1.50 1.23 5850 (a) (0.951“ Table 3-1. (Continued) 71 P(4.2K) Ampoule Circumstance d [PB] Mount Samples No. at R.‘1‘. (mm) (nSZcm) RRR method Other K-H6b # 8 10 )1Hg He 1.50 1.32 5406 (a) Second run [1.05:1'~ of K-H6a K-H6C # 8 100 )ng He 1.50 1.35 5330 (a) Third run [1.10] of K-H6a K-0.9H6a # 5 He atm. 0.90 1.29 5570 (a) Done by M. [1.01]“ Haerle K-O.9H6b # 4 He atm. 0.90 1.37 5250 (a) Second run [1.051“ of K-O.9H6a K-0.5H # 5 He atm 0.50 1.79 4020 (a) [1.511“ K-.25H1a # 6 He atm. 0.25- 2.37 3030 (a) (2.051“ K-.25H2a # 6 He atm. 0.25 2.36 3050 (a) [2.081‘ K-.25H1b # 6 He atm. 0.16“ 3.13 2300 (a) Second run K-.25H2b # 6 He atm. 0.16' 3.27 2200 (a) Second run of K-.25H2a K-.25H3 # 6 He atm. 0.25 2.25 3200 (b) [1.98] K-.25H4 # 6 He atm. 0.25 2.23 3220 (b) [1.96] K-.1OHla # 9 He atm. 0.10 11.60 620 (b) [11.30] K-.10HZa # 9 He atm. 0.10 13.80 520 (b) [13.40] K-.10H1b # 9 He atm. 0.09' 11.60 620 (b) Second run [11.30] of K-.10H1a K-.10H2b # 9 He atm. 0.09. 13.00 550 (b) Second run [12.70] of K-.10H2a 72 Table 3-1. (Continued) P(4.2K) Ampoule Circumstance d [Po] Mount Samples No. at R.T. (mm) (anom) RRR method Other K-.25V1 # 7 ~10 )ng He 0.25 2.11 3410 (b) . [1.86] x-.25v2 # 7 ~10 pHg He 0.25 2.25 3200 (b) [1.99] ' K-.10V1 # 7 -10 qu He 0.10 7.70 930 (b) (7.421“ K-.1OV2 # 7 ~10 pHg He 0.10 7.16 1000 (b) (6.881“ K-.1OV3 # 7 ~10 )1Hg He 0.10 8.83 810 (b) [8.49] K-.10V4 # 7 ~10 pHg He 0.10 7.83 920 (b) [7.49] K-.10A1 #10 Ar atm. 0.10 6.22 1160 (b) Without [5.91] molecular sieve K-.10A2 #10 Ar atm. 0.10 8.31 870 (b) Without [7.99] molecular sieve K-.10A3 #11 Ar atm. 0.10 3.67 1960 (b) With [3.37] molecular sieve K-.10A4 #11 Ar atm. 0.10 5.00 1440 (b) with [4.59] molecular sieve 'Diameters of corroded versions were estimated by the change of room temperature resistances. “Values are determined by p are determined by P0 z P(~ =p(4.2K) - 0.28 nil-cm. all other values 1K). 73 Table 3-2. Characteristics of bare free hanging pure Rb samples (from one ampoule) 9(4.2K) Ampoule Circumstance d [Pb] Mount Samples No. at Rm. Temp. (mm) (nflcm) RRR method Rb-Hl #1 He atm. 1.50 41.8 300 (a) [35.61“ Rb-HZ #1 He atm. 1.50 44.8 280 (a) [38.61“ Rb-V1 #1 ~10 pHg He 1.50 31.0 403 (a) [24.7] Rb-VZ #1 -10 (ng He 1.50 31.0 403 (a) [24.9] “Values determined by P =P(4.2K) - 6.252n em, all others determined by PC as MAR). Table 3-3. Characteristics of bare free hanging pure Li samples (from same batch) Circumstance d P(4.2K) Mount Samples at Rm. Temp. (mm) (nflcm) RRR method Other Li-Hla He atm. 3.0 9.0 1030 (c) Li-HZa He atm. 3.0 10.8 860 (c) Li-H1b He atm. 3.0 8.3 1120 (c) Second run of Li-la Li-H2b He atm. 3.0 10.4 900 (0) Second run of Li-2a 74 Table 3-4. Characteristics of bare free hanging pure Na samples. P(4.2K) Circumstance d [Po] Mount Samples Source at Rm. Temp. (mm) (nflcm) RRR method Na-Hl Gallery He atm. 1.0 12.40 380 (a) Chemical Company Na-HZ Callery He atm. 1.0 13.00 360 (a) Chemical Company Na-H3 J. C. He atm. 1.0 1.02 4660 (a) Garland [1.01] Na-H4 J. C. He atm. 1.0 1.00 4720 (a) Garland [1.00] Na-.25A1 J. 0. Ar atm. 0.25 1.18 4010 (b) Garland [1.18] Na—.25A2 J. C. Ar atm. 0.25 1.14 4160 (b) Garland [1.14] Na—.25H1 J. C. He atm. 0.25 1.42 3350 (b) Garland [1.42] Garland [1.49] Na-.10H1 J. c. He atm. 0.10 2.08 .2270 (0) Garland [2.08] Na-.10H2 J. C. He atm. 0.10 2.57 1850 (0) Garland [2.57] 75 N!-OQ .1909 9! '0 Fig. 3-5. The way of sucking potassium into a tube: l. Melting potassium, 2. Tube, 3. Syringe. 76 experiments. The surface of K samples touching teflon turned black right away, but the rest of the sample below the very thin surface remained shiny during all the experiments. Two Na samples encased in polyethylene tubes were made in the way described above. One K sample was coated by cleaned paraffin oil and several were measured while in contact with plastic. To try to grow a single crystal K sample, a beaker of 0-122 White Heavy Paraffin Oil from Fisher Scientific Company was carefully cleaned in the system shown in Fig. 3-6. The oil was heated to about 150°C while it was being pumped by a mechanical pump for about 3 days until no more bubbles came out. Then the input of the mechanical pump was closed and liquid nitrogen was poured into the thermos and the input valve of the cold trap was opened to better evaluate the system. Some crumbs of clean potassium were put into the oil to clean out left over dissolved air or other impurities. This cleaned paraffin oil was used sepa- rately also for coating a K sample, (sample K-0). A piece of precision bore pyrex capillary, with internal diameter d=l.5 mm from Wilmad Glass Co., Inc., was carefully cleaned in the same manner as the plastic tubing described above. The cleaned oil and pyrex capillary were moved into the Ar filled glove box. First, a little cleaned paraffin oil was sucked into and pushed out of the capillary with a syringe to lubricate the internal wall of the capillary. Second, 77 .muo__oa xm. oa>h o>o_m cm_soo_oz otc_m cacao—o ea_z to___c auto e_oo .e .toaoto_e e_ae_s .m .o>_o> .5 .esoa so. .m .oum_auuo: .N .__o e.mmmtmm ._ "__o e.muatma mc_>m_tsa ecu Ecum>m one .mum .m_m 1.1. L .l .. N u. u. 4.. . .. .m i m m ”m F a. m . Doooo‘. . .J I' l J i l l l aEsc .mo_cmzooe oh 78 molten potassium was sucked into the capillary and allowed to solidify. Third, the cleaned paraffin oil was poured into a clean hot-beaker, the part of the capillary with potassium in it was submerged in the oil, and the oil was heated to about 70°C, a few degrees above the melting point of potassium. The capillary was then slowly pulled out of the hot oil at the rate of 0.5 cm per half an hour (see Fig. 3-7). After growth, the sample was pushed out of the capil- lary with a syringe. Its ends were carefully scraped with a cleaned surgical knife to clean off the oil and were cold welded to the current and potential leads as in method (b) of Fig. 2-4. After the measurements were completed, a 1 cm long piece of this sample was cut out and put into a thin wall glass tube filled with cleaned paraffin oil to protect the sample from reacting with air. Then it was tested on GE. XRD-S x-ray machine with X= 0.7107 A and beam diameter d = 0.4 mm. The result showed a sharp peak at 29 = 15.6°. From Bragg equation nk = 2dsin0, d = 5.2 A was obtained for n =- 2, thus, at least, within the spot of x-ray beam, a single crystal part with d a 0.4 mm was found. Unfortunate- ly, since only one spot of the sample was tested, the veri- fication of the whole piece as a single crystal K sample was not established, but Dr. P. A. Schroeder (ref. 64) and his collaborators did grow some single crystal K samples in the Netherlands using the same technique with even faster pull- ing speeds. Their samples were tested with x-ray diffrac- tion at three different spots along each sample. 79 :IEZEE '_1C) in. 1 :: ‘l. .. -Lz- --. 2__k \ 34""‘*'::::E:::::: Fig. 3-7. The system for pulling single crystal potassium sample: 1. Precision bored PX capillary, 2. Cleaned paraffin oil, 3. Hot-beaker. 4. Manipulator with adjustment in two perpendicular directions. 80 The K(0.077 at% Rb) alloy was obtained by first making a.KK1.3 at% Rb) mixture and then adding some of this mixture to pure potassium to form a.KK0.077 at% Rb) alloy. The K19.40 at% Rb) alloy was obtained by adding Rb directly to pure potassium. The percentages above were calculated from the weights of each material, which were weighed on a preci- sion scale inside the glove box (ref. 22). The samples of the alloys were extruded through a die and mounted in the same way as the pure K samples described earlier. For the KK9.40 at% Rb) alloy, the resistivity was high. Since a high resistance sample would cause big heat- ing effects at low temperature when a current passes through it, the samples and potential leads were thick (d a 3.0 mm) and short (potential leads were ~l.8 cm apart, each sample was ~4 cm longJ to keep the resistance low. The sample holders were shortened for holding those short samples. The characteristics of the possible single crystal K sample, of the K samples coated by clean paraffin oil, and those enclosed in polyethylene or teflon tubes, are given in Table 3-5. The characteristics of the K-Rb alloy samples are given in Table 3-6. The characteristics of the Na samples encased in polyethylene tubes are given in Table 3-7. 81 Table 3-5. Characteristics of K samples in contact with oil or plastics P(4.2K) Days Ampoule Circumstance d Mount 9 Rm. Samples No. at Rm. Temp. (mm) [ P ] (nflgm) RRR method Temp. Other K-S K-O K-Phla K-Phlb K-Phlc K-Phld K-Phle K-PhZa #12 #15 #8 #8 #8 #8 #8 #15 Try to be grown as a single- crystal sample Ar atm. coated by cleaned paraffin oil Ar atm. in a polyethy- lene tube 10 pHg He in a polyethy- lene tube 10 pHg He in a polyethy- lene tube He atm. in a polyethy- lene tube 10 pHg He in a polyethy- lene tube 100 qu He in a polyethy- lene tube He atm. 1.5 1.5 1.6 1.6 1.6 1.6 1.6 .9 1.17 [0.91] 1.32 [1.07] 2.97 [2.70] 1.68 [1.42] 1.18 1.20 1.20 [0.95] 3.30 [3.02] 5150 5450 2420 4280 6090 5990 5990 2180 (b) (a) (a) (a) (a) (a) (a) (b) 13.0 0.5 11.0 24.5 26.0 27.0 0.5 Second run of K-Phla Third run of K-Phla Fourth run of K-Phla Fifth run of K-Phla Table 3-5. (Continued) 82 P(H.2K) Days Ampoule Circumstance d [ Po] Mount 9 Rm. Samples No. at Rm. Temp. (mm) (nflcm) RRR method Temp. Other K-Ph2b #15 a. in a .9 1.35 5330 (b) 2.5 Second polyethy- [3.02] run of lene tube K-PHZa b. He atm. K-Ph2c #15 a. in a .9 1.32 5060 (b) 13.0 Third polyethy- [1.17] run of lene tube K-PHZa b. He atm. K-Ph2d #15 a. in a .9 1.28 5620 (b) 73.5 Fourth polyethy- [1.03] run of lene tube K-PHZa b. He atm. K-TH1a #16 a. in teflon 1.5 8.31 855 (b) 0.5 tube [8.10] b. He atm. K-TH1b #16 a. in teflon 1.5 5.17 1390 (b) 2.5 Second tube [n.87] run of b. He atm. K-TH1a K-TH1c #16 a. in teflon 1.5 1.62 5370 (b) 13.0 Third tube [1.33] run of b. He atm. K-TH1a K-TH1d #16 a. in teflon 1.5 1.20 5990 (b) 73.5 Fourth tube [0.99] run of b. He atm. K-TH1a K-PA1a #1“ a. in a .9 5.u6 1320 (b) 0.5 polyethy- [3.02] lene tube b. Ar atm. K-PA1b #1” a. in a .9 1.u5 H960 (b) 9.5 Second polyethy- [1.20] run of lene tube K-PA1a b. Ar atm. Table 3-5. (Continued) 83 P(“.2K) Days ' Ampoule Circumstance d [ 90] Mount @ Rm. Samples No. at Rm. Temp. (mm) (chm) RRR method Temp. Other K-PA1c #1“ a. in a .9 1.39 5170 (b) 16.0 Third polyethy- [1.1“] run of lene tube K-PA1a b. Ar atm. K-PA1d #1“ a. in a .9 1.57 “580 (b) 19.0 Fourth polyethy- run of lene tube K-PA1a b. Ar atm. K-PA1e #1“ a. in a .9 1.36 5290 (b) 97.0 Fifth polyethy- . [1.12] run of lene tube K-PA1a b. Ar atm. K-PAZa #1“ a. in a 1.6 5.“8 1310 (a) 0.5 polyethy- [5.20] lene tube b. Ar atm. K-PAZb #1“ a. in a 1.6 1.22 5890 (a) “.5 Second polyethy- [0.97] run of lene tube K-PAZa b. Ar atm. K-PAZc #1“ a. in a 1.6 1.38 5210 (a) 16.0 Third polyetny- [1.12] run of lene tube K-PA1a b. Ar atm. K-PAZd #1“ a. in a 1.6 1.23 5850 (a) 19.0 Fourth polyethy- run of lene tube K-PA1a b. Ar atm. K-PA2e #1“ a. in a 1.6 0.95 7570 (a) 97.0 Fifth polyethy- [0.73] run of lene tube K-PA1a Ar atm. 84 Table 3-5. (Continued) p(“.2K) Days Ampoule Circumstance d [ Po] Mount @ Rm. Samples No. at Rm. Temp. (mm) (nflcm) RRR method Temp. Other K-KA #17 a. In con- 1.5 1.28 5620 (a) tact with [1.03] two pieces of Kel-F b. Ar atm. K-PPA a. Bare sam- 1.5 1.22 5890 (a) ple with [0.96] potential leads in polyenty- lene tube b. Ar atm. 85 Table 3-6. Characteristics of bare free hanging K-Rb samples . P(“.2K) Ampoule Circumstance At % d [ no] RRR(“.2)Mount Samples No. at Rm. Temp. Rb (mm) (nflcm) [RRRO] method K—Rb1 K: #58 #6 He atm. 0.077% 0.25 -- --- (b) : #2 [11.2] [6“0] K-Rb2 K: #58 #6 He atm. 0.077% 0.25 --- --- (b) Rb: #2 [11.6] [620] K-Rb3 K: #13 Ar atm. 9.“OO% 3.20 1010 8.30 (a) Rb: #3 [1010] [8.30] K-Rb“ K: #13 Ar atm. 9.“005 3.20 10“0 8.10 (a) Kb: #3 [10“0] [8.10] 86 Table 3-7. Characteristics of Na samples encased in polyethylene tubes P(“.2K) Days Circumstance d [ 90] Mount 6 Rm. Samples Source at Rm. Temp. (mm) (nflcm) RRR method Temp. Other Na-PA1a J. 0. Ar gas 1.6 2.27 (a) 12.0 Garland Na-PAZa J. C. Ar gas 1.6 2.52 (a) 12.0 Garland Na-PA1b J. 0. Ar gas 1.6 2.29 (a) 33.0 Second Garland [2.29] run of Na-PA1a Na-PAZb J. C. Ar gas 1.6 2.“O (a) 33.0 Second Garland [2.39] run of Na-PAZa 87 3.5 MEASQBEMENI_MEIEQQ 3.5.1 W More than fifty runs were made in this study. In each, we measured PULZK), Po (usually), and both dP/dT and G from 0.07K to 4.2K. In each run, two twin samples or similar samples were measured. 3.5.1.1 R§295K) First, the room temperature resistance of sample was measured using a constant current supply and a Keithley digital nanovoltmeter. The room temperature T was also measured. Then the room temperature resistance was converted to 295K resistance R(29SK) by assuming a linear temperature dependence, i.e., P(295K) x L(295K) 31.22252. . “295‘“ 0.2220. (3-2) R(T) _ MT) P(T) x L(T) A(T) where L and A are the length and cross-section area of sample respectively, and L4295k)/A(295K) = LATO/A(T) to a precision better than 0.1% at room temperature. Hence, _ 9(295K) R(295K) - R(T) -_3(TT_ = R(T) 1 + 21.80 (3_3) l + a(T - 273.2) assuming 88 P('I‘) = P(273.2K) [1 +oz(T - 273.2)] (3-4) Actually, we happened to use the approximation R(295K) = R(T) x 295/T for some measurements of potassium, this made R(295K) not more than 0.5% too big in our room temperature range. The a and 9(295K) used for K, Li, Rb, and Na were those as listed in Table 3-8 (from Landolt-Bornstein Numerical Data and Functional Relationships in Science and technology, Group III, Volume 15). 3.5.1.2 Circuit for Measuring P(4.2K), Po, and dP/dT. When the samples were cooled to 4.2K or lower, the circuit in Fig. 3-8 was used. This circuit consisted of a SQUID (Superconducting Quantum Interference Device) null detector and three resistors wired in series. These three resistors were the samples, R1 and R2, and a standard resistor RSt =.LJ34 #9 made of a tin-indium alloy, chosen to be superconducting at about 3.8K. For thin pure samples and thick alloys with high resistances at low temperature, an inductor (Lt z 50 pH) was inserted in series with the three resistors to keep the relaxation time of the circuit long enough to have the SQUID lock properly. The inductor and all the wires in this circuit were made of Niomax CN, a multifilament Nb-Ti superconducting (Tc > 4.2K) wire with Cu-Ni cladding made by Imperial Metal Industries (ref. 22). To keep the noise introduced by stray magnetic fields as low as possible: 1) Rst and most of the wiring except 89 Table 3-8. The electrical resistivities of some alkali metals Metals (273.2K) a P(295K) (calculated) K 6.45 .0050 7.19* Li 8.50 .00445 9.32 Rb 11.25 .0051 12.50 Na 4.29 .00485 4.74 * This number should be 7.15 from the calculation, but we used 0(295) = 7.19 for K to be consistent with van Kempen et al. (ref. 14). 90 Mixing Chamber Thermal Link with Electrical Isolation -—=) EKJLHE) U1 Controller U2 3::::;:3 ‘ ‘ [:;:::: % RL1 81.2% Q 50010 :23 a: 1: G] Sample Can G2 I _.__-__-__--..___...__...._.---J Fig, 3-8. The low temperature circuit. The components inside the broken line are inside the sample can. 91 that near the samples were shielded in superconducting lead (Pb)('Tc = 7K) tubing. 2) All the wires leading to the samples were carefully tied or varnished down to reduce vibration and thus reduce magnetically induced currents. 3) The vacuum can, with the sample can inside, was shielded by one Nb can, and the whole refrigerator was surrounded with another u-metal can. Both the Nb can and the u-metal can had openings only at the tops. The wires were drawn out of the can through a special electrical feedthrough. This feedthrough was made by running several pieces of insulated Niomax wire through a clean 1/8” diameter stainless steel tube, which was then inverted in a small cup of liquid Stycast 1266 epoxy that was then allowed to harden. The ratio of the resistances for any two of the three resistors in this low temperature circuit could be measured by using the current comparator (mentioned in section 3.1). Let Rm and RS represent the resistances of these two resistors respectively; When the current comparator was used to compare them, the master current Im was passed through Rm while the slave current IS was passed through RS. Here, IS = CIm, where C was the switch setting of the cur- rent comparator. A standard current reversal technique was used to eliminate thermal EMFs generated in this circuit as follows. For currents going one way (the directions of I currents through two resistors in this circuit should be opposite to each other), the SQUID voltage is 92 v+ = I Rm - I R + v m s s T ImRm - cxmas + vT (3-5) where VT represents the stray voltage due to any thermal EMF. After the currents on the two sides are reversed, the SQUID voltage is V -ImRm + ISRS + VT —ImRm + CImRS + VT (3-6) C was adjusted to make V+ = V‘. Then, ImRm - CImRs = -ImRm + CImRs (3-7) and hence, Rm/Rs = C independent of VT' At 4.2K, the resistances of two samples, R1 and R2 in Figure 2-8, were measured by using the current comparator to compare each of them with Rst' Then the ratio of the 4.2K resistances to the 295K resistances were calculated. neglecting the change of L/A between 4.2K and 295K (the error caused by this neglecting will be discussed in section 3-5), one has P(4.2K)/P(295K) = R(4.2K)/R(295K) = l/RRR. For pure samples, P(295K) Ppure (295K),1and (”4.2K) P(295K)/RRR. For K-Rb alloys, assuming Matthiessen's Rule (MR) as a first approximation, 93 P(295K) P(4.2K) + PpureK(295K). Therefore, P(4.2K) a PpureK(295K)/(RRR-l). The impurity-dependent deviation from MR might cause an error as large as 1-3% when RRR is as small as 8 for a 9.4% K-Rb alloy (ref.65). The ppure(295K)s for K, Li, Rb, and Na are listed in Table 3-8. As described in section 3.1, there are two ways to study the temperature dependence of resistivity at low temperature. One is to measure P directly, the better one is to measure dP/dT. We measured dP/dT in this study. To obtain dP/dT of a sample, (e.g., R2), the current comparator was used again to compare R2 (as Rm) to R1 (as R5). The reference R1 was kept at a constant temperature T1 by regulating the mixing chamber temperature with a tempera- ture controller monitored by GRT4. The temperature of the sample R2 was changed from T2 to (T2 +70T) either by using heaters U2 and L2 alternately with the same power input into each or only L2 with two different powers. (These two methods were tested several times in each run, to make sure both of them gave consistent resultsd When the sample was at temperature T2, the ratio of its resistance to the refer- ence resistance R(Tl) was measured as C(Tz) = R2(T2)/R(T1). When it was at T2 + AT, the ratio was measured as C(T + AT) 3 R2(T ‘1' AT)/R1(T1)o Then, 94 C(Tz) = = —— = — 3’8 R2(T2) R2 92 ( ) since the change of L/A in AT, typically 0.1K, was less than 10-7%, which is negligible. Hence, dP _ . AP _ AC PdT " £15.30 PAT " CAT (3 9) and, (39 PAC dT CAT The current through the sample was usually 50 mA, or occasionally, to test for current dependence, 20 mA. There was never any measurable current dependence in dP/dT. In order to get optimal precision, the noise caused by the reference resistor must not be worse than that of the sample. For this reason, the two samples were normally made as similar as possible, and each of them was the reference for the other. Even so, the Johnson noise (x/IFTREf) in the reference would be larger than that in-the sample if the temperature of the reference was higher than that of the sample. To minimize this noise, the reference was cooled down together with the sample to a given temperature, and then kept at this temperature while the sample was heated to higher temperatures for taking data. The Johnson noise in 95 the reference resistor was thus always smaller than that in the sample. At the ends of most runs, 903 of the samples were measured by stopping the 1K pot pump and the circulation of the dilution refrigerator when the temperature of the sample was below 1K. When the standard resistor Rst' which was attached to the 1K pot, became normal (TC = 3.8K), while the temperature of the samples was still around 1K, the ratio of resistance of each sample to the resistance of the standard was measured using the current comparator. The resistivity at about 1K was then calculated as PULzK) was done above. According to van Kempen et al. (ref. 14), W.P. Pratt, Jr. (ref. 59), and our new results, when the temperature is below 1K, the electron-phonon scattering contribution to the resistivity was negligible and the electron-electron scattering contribution was at least 3 orders of magnitude smaller than no in K, Li and Na. Thus, P(lK) was chosen as 80' As already mentioned, temperature measurements were critical in this study. Even though the thermometers were already well calibrated (ref. 22), the Wiedemann-Franz (W-F) Law was used for in-situ checking to be sure that no systematic errors crept into the temperature measurements. The procedure was as follows. In Fig. 3-8, RLl and RLZ, both Ag-0.1 at% Au alloys, were used for weak thermal links between the mixing chamber and the samples. For each link, the U-heater was placed at one end (called the U end), which 96 was connected to the mixing chamber. Another identical heater, the L-heater, was placed at the other end (called the L end). The reference resistor and the mixing chamber were kept at a constant temperature. The sample was first heated to a temperature T by running a current through the U heater. In thermal equilibrium there should not be any heat flowing through RL (see Fig. 3-8). There should thus not be any temperature difference between the U and L ends. Then the sample was heated to T + AT by running the same current through the L heater, this time there was a continuous heat flow from the L heater through RL to the mixing chamber, which caused a temperature difference between the U and L ends. In these two cases, the temperature of the U end should not change since the heating power in these two cases were the same, so that the heat current flowing from the 0 end to the mixing chamber did not change. Thus, we can say that the temperature difference between U and L ends in the second case was equal tolAT. Applying the W-F Law, K/OT = L0, to RL' one has: RL = LOT AT/Q = LOT AT/RhIz (3-10) where Rh was the resistance of the heater, I was the current through the heater, and Tave = T + AT/Z was taken as T. The resistance of the alloy RL was dominated by its residual resistance R0 at low temperatures, so Tave AT/RhI2 should not change for every datum we got. This was used for in- situ checking the product T-AT and thus AC/CTnAT. The 97 uncertainty in T is estimated much less than 1% except near 0.1K where the uncertainty is closer to 1% (ref. 59). 3.5.2 Thsrmoelsstris_Ratic_§_and.lhermsscwsr_s. The thermoelectric ratio G of each sample was also standardly measured in each run, since it was easily obtained. ‘The only additional requirement was a G heater at the hot end of each sample (Fig. 3-8). As mentioned in section 1.1, c = -%— , i.e., = —¥— (3-11) q 8&0 Q E=0 G was measured by sending a current IG to the G heater to heat the sample from one end, and then sending another current I through the sample to cancel out the resulting thermal voltage. This voltage was detected by the SQUID. The master current was normally'set to be 50 x 0.1 mA, the cancelling current was supplied by the slave side of the current comparator, (i.e., the cancelling current was C x 5 mA. The resistance of the G heater, a Dale resistor, was measured at liquid helium temperature, and stayed constant (change less that 0.1%) below 4.2K. The heating power was calculated from RGIé (3-12) 0. ll thus 0 = c x 5 mA/RGIé (3-13) 98 The thermopower S of only one K sample (K - 1/2 H) was measured in this study. The G heater mentioned above was still needed. In order to measure the temperature differ- ence across the sample, two Germanium Resistance Thermom— eters GRT4 and GRTS were placed on the copper supports of potential leads (section 3-3). As mentioned in section 1.1, E V S a — ' i.e., S = ‘— S was measured by sending a current IG to the G heater to heat the sample from one end, and then sending another current I through the reference to cancel out the resulting thermal voltage. This thermal voltage was detected by the SQUID. 'The cancelling current was the same as in the G measurements mentioned above, i.e., I = C x 5 mA. So the cancelling IR drop was C x 5 mA x Rr' The different temperatures of the two ends of sample were measured by GRT4 and GRTS. Then AT was obtained. Thus, C x 5 mA x Rr 3 = AT (3-14) To determine S, it is thus necessary to determine Rr ('1‘). Rr was measured only at 4.2K. Since [Rr (4.2K) - Rr (OKil/R[(4.2K) = [p£(4.2K) - Pr(OK)]/Pr(4.2K), where Pr(4.2K) - Pr(0K) = 0.28 nflcm for potassium (ref. 12,59). Rr(0K) was then obtained as Rr(0K) = Rr(4.2K) [l — 0.28 nSZcm/Pr(4.2K)]. Furthermore, assuming R(T) = R(OK) + 99 Ce'ei‘r/T with 9* = 20K, where C could be obtained from R(4.2K) = R(OK) + Ce'20/4'2, R(T) was calculated for temperatures between 0K and 4.2K. The uncertainty of this calculated R(T) is estimated as 3%. 3.6 W 3.6.1 W In G and S measurements, our attention was focused on the temperature dependence of G or S. As mentioned in section BJLZ, G was calculated by G = C x 5 mA/RGIé. The 0.1 x 50 mA current setting of the current comparator had 0.16% uncertainty. The uncertainty in the resistance of the heater RG was 0.1%. The error in measuring IG was due mainly to the digital round-off of the DVM used to measure it; this could cause 0.5% error at the worst. Finally, the uncertainty of the resistance used to measure this current was 0.05%. Taken together, these uncertainties sum to, at worst 1%. The major source of uncertainty in G was in the deter- mination of C, due to thermal EMF noise and Johnson noise. This uncertainty was less than 1% when the temperature was above 1K, but the percentage error got worse as the magni- tude of C and the heat input got smaller at lower tempera— tures. At the lowest temperatures, the uncertainty in C could be as large as 5%. The major uncertainty in T measurement was random error due to thermal fluctuations and thermal drifts in the 100 system. The worst case was when T was around 3K, since the dilution refrigerator had not been completely started, the cooling power was varying with time, thus temperature regu- lation could not be used. This uncertainty in T is esti- mated as only 0.5% in the worst case. As mentioned in section 3JL2, S was calculated by S = (C x 5 mA x Rr)/AT. The uncertainty in 0.1 x 50 mA current setting was 0.16% (see above). The uncertainty in C was less than 1% in whole temperature region for this S measure- ment. The uncertainty in AT is estimated as 1% (section 3.5.1, ref. 59) in the worst case. The major uncertainty in S was in determination of Rr' This uncertainty is estimated as 3% (see section 3.5.2). 3.6.2 Uncertainties in P(4.2K),PQ, and dP/TdT Measurements P(4.2K) was determined from P(4.2K) = P(295K) x R(4.2K)/R(295K), where R(295K) was measured by R(295K) = R(T) (l + 26.8 xoz)/(1 + a(T - 273.2))or R(295K) = R(T) x 295/T (see section 2.4.1). 90 was determined similarly. So, the uncertainties in 1K4.2K) and 90 were dependent on three main factors: a) The uncertainty of R(295K) was mainly due to the uncertainty of the room temperature resistance R(T) which was always measured with better than 1% precision. in EH4.2K) or R0 were measured against Rst within 2%, due to the temperature dependence of Rst near 4.2K (ref. 22). RUL2K) was also affected, in lesser degree, by the small temperature variation of the liquid He bath due to changes 101 in atmospheric pressure. c) The main error was the systematic error due to ignoring the change of L/A when. the sample underwent thermal contraction from room temperature to 4.2K or 1K. The room temperature linear thermal expansion coefficients of the samples (in units of 10"6 K'l) were 83 for K, 71 for Na, 66 for Rb, and 45 for Li. These coefficients would stay constant from room T to about their Debye temperatures @Ds (330K for Li, 160K for Na, 114K for K, and 65K for Rb), then start turning down, at about 0.161) they would decrease as fast as T3. Calculations showed this caused about 2% errors in FK4.2K) and Do. The 0.5% opposite systematic error from using R(295K) = R(T) x 295/T as an approximation is negligible. In summary, the error in 1N4.2K) and 90 measurements could be as high as 4%, including a correctable systematic error of 2% due to c). But, any error in 1M4.2K) and Po would not affect the form of plots of the final quantity dP/dT versus T either as p(4.2K) AC/C AT or PO AC/C AT. The only effect of this error would be to multiply dP/dT by a constant close to 1. In the calculation of the quantity' PAC/C AT, C was ave shown to have precision better than 0.1 ppm earlier, so the main uncertainties came from 9. AC, and AT. a) P(1K) was chosen to be P0 in this calculation. As already mentioned, our attention was focused on the temperature region where electron-electron scattering was dominant, in this tempera- ture region the temperature dependent part of P was at least 3 orders of magnitude smaller than no or P(lK) (see section 102 3.5.1). So this caused less than 0.1% random error. Some- times P(4.2K) was used instead of Po; and 9(4.2K) could be 20% bigger than no. This, however, represents only a dif- ferent renormalization of the quantity of interest. The systematic errors in P(4.2K) and PO measurements were al- ready discussed above. b) In many cases of this study, AC could be very small or even zero, so the percentage uncertainty of AC could be very large. To reduce the uncertainty in AC when AC was small, a micro-computer averaging. technique was used to measure C to within 0.01 ppm. Typically, except at the lowest temperature, the un- certainty in AC was less than 1%. c) The uncertainty in AT, which was estimated using w-F law in-situ checking was within 1-2% (ref. 59). CHAPTER 4 Experimental Results and Analysis W As mentioned in section 1.2.1, C. W. Lee et al. (ref. 8) used the same technique as described above to measure free hanging, bare, high-purity, thick (d = 0.9-3.0 mm) K samples under Ar gas. Their dP/dT data varied closely as T, (i.e., pm = A132, from 1.3K to 0.35K. "A" did not change much from sample to sample, having a mean value of 0.24 i: 0.02 chm/Kz. This value is consistent with that expected for electron-electron scattering, 0.17 pacm/K2 (ref. 7,21). Below 0.35K, their data showed deviations from T2. Their G data had the approximate form C = C0 + B*T2 + (C*/T) x exp(-6*/T), with co = -0.03 i 0.03v‘1, 13* = -o.3o i 0.01v‘1K‘2, and 9* a 23 3: 2K. Using a better glove box, an improved sample can (see section 2.3), and a thin copper can attached to the mixing chamber to shield the sample can and the thermometers on it from heat radiation (see section 2.3), we measured 14 free hanging bare K samples with d = 1.5-2.0 mm; 4 were still measured under Ar atnh; 6 under He atm. (one among these was sample K-H4 annealed at 50°C for 30 minutes); and 4 under partial vacuum. 103 104 4.1.1 Ih£_B£SiSLiXi£¥. The results of resistivity measurements on these thick samples are shown in Fig. 4-1. Again, p(T) varied closely as AT2 from 1.1K to 0.2K-+0.4K (for different samples), with A = 0.24 1; 0.02 pflcm/K2 for Ar gas (the same as C. W. Lee found), A = 0.25 i 0.02 chm/K2 for He gas without annealing at 50°C, and A = 0.22 ¢_0.01 pflcm/K2 for partial vacuum. The sample annealed at 50°C showed the smallest p(4.2x) = 1.22 x 10"9 Gem and the smallest A = 0.19 chm/Kz. Below (L2K-+0.4K, all the samples showed an anomalous turn-up similar to that found by Lee. The parameters of the samples are given in Tables 3-1 and 4-1. In the data of our 6 samples cooled in He gas, the higher coefficient A corresponded to higher residual resistivity no. But our samples cooled in Ar gas and partial vacuum did not show any simple relation between A and Po. We plot A versus 100 [no a p(4.2K) - 0.28 chm/KZ] for 5 of our samples cooled in He (Fig. 4-2). These samples are consistent with the constant value 0.25 :_0.02 chm/KZ. However, they also fit A = A0 + cab (the solid line in Fig. 4-2) with A0 z 0.17s poem/K2 and c = 5.7 x 10‘5r’2. This latter fit would imply a Po-dependent part in A of about A1100) = A - A0 = 5.7 i 0.5 x 10‘590. This is somewhat larger than that expected for electron-defect inelastic scattering (section 2.2AD. With either fit, A has a dominant component AO due to electron-electron scattering. The linear fit in no would suggest an additional component due 105 110—- : K... x -~a o - K-9 K-H4 0 K—10 K -H5a $100 ' K-11 K -H5b E 0 K- H1a K - H63 c K- H1b K- H6b {SQOL . 0 K- H2 K - Hsc r—5 E -- K- 4a 2 2° }01 C.W. Lee \ :o —--- K- 53 CL 130 £801- 1.2:. fig ‘2: II . 73 1'2, 0 \JW {a o of act: 0 a 00?. ‘Q' : O o °’§.;‘.v~ . 6 v 0 v! o v0 I O —- - ‘ “-‘.O 60 no a\ :3..-‘.Qr;r:‘:) e-..o. '=.--o-- 1° ;..5°.......:.£. l V A A ‘o;“~~—--Q---...0-..l:-_..‘_ _....L--_._L_._ V v ' A A A A O 50— . ' . ' ' . . 4G .11., l l l i, .1 I 0 0.2 0.4 0.5 0.8 1.0 1.2 T [K] Fig. “-1.. (P(“.2K)/PT)(AP/AT) versus T for free hanging bare high- purity thick K samples. 106 Coefficient A and other parameters of thick K samples. To determine A, we assume Po = P(“.2K) - 0.28 nSlom. t = time at room temperature. Diam t P(“.2K) A Circumstance Sample (mm) (days) (nflcm) (chm/KZ) at room T. K-8 1.5 0.5 1.50 0.25 Ar atm. K-9 1.5 0.5 1.72 0.2“ Ar atm. K-lO 1.5 0.5 1.“? 0.2“ Ar atm. K-11 2.0 0.5 1.26 0.22 Ar atm. K-H1a 1.5 0.5 1.70 0.26 He atm. K-H1b 1.5 26.0 1.65 0.25 He atm. K-HZ 1.5 0.5 1.27 0.23 He atm. K-H3 1.5 0.5 1.97 0.27 He atm. K-H“ 1.5 0.5 1.22 0.19 He atm. annealed at 50°C for 30 min. K-H‘Sa 1.5 0.5 1.27 0.22 10 flHg He K-HSb 1.5 11.0 1.39 0.21 10 qu He K-H6a 1.5 0.5 1.23 0.23 He atm. K-H6b 1.5 2.0 1.33 0.22 10 {1H3 He K-H6c 1.5 3.0 1.35 0.22 100 pHg He + molecular seive 107 to electron—defect or electron-impurity inelastic scattering. However, in view of the small number of data points in Fig. 4—2 and the lack of any correlation of A with pO in samples cooled in Ar or vacuum, we cannot rule out the possibility that the observed variation with PC shown in Fig. 4-2 is just apparent, in that a large number of data points might end up randomly fluctuating the dashed line. The anomalous turn-up below 0.35K might be associated with defects in the samples. Comparing K-Hla (the first run of sample K-Hl) and K-Hlb (the second run of sample K-Hl after annealing at room temperature for 26 days), we see that K-Hlb has a much smaller turn-up than K-Hla. Similar- ly, sample K-HSb has a smaller turn-up than K-HSa. This result is consistent with what M. L. Haerle (ref. 22) found; deformation in pure K samples not only brought the coeffi- cient A up, but also enhanced the anomalous turn-up at low temperatures. More discussions of the anomalous turn-up will be given in section 4.6, together with the similar anomalous turn-up in Na, Li, and Rb. 4.1.2 The Thermal-M ' G measurements were made for samples K-ll, K-HSa, K- HSb, K-H6a, K-H6b, and K—H6c from about 4.2K down to about 70 mK, and for samples K-lO, K-Hla, and K-HZ from about 1K down to about 70 mK. All the data are shown in Fig. 4-3. They have the same form as C. W. Lee's data, except the data of K-HSa and K-H6b, which have small bumps at temperatures 108 .o: c_ po_00u mo_dEmm x xo_;u m to“ JR m> < . T: a: Trudi at 2 3 a... . _ . _ 2 law. V m... m w 0... Va llllllll O llllllllllllllllllllllll 1mm. 0 109 .mm_QEmm x xo_cu >u_tsa1£m_z scan mc_mcm£ worm sod p m> o c: ._. fie QM 9~ all .3... m A U I, .... 2. «Q II 8:12 .o , «>50 mezzmi--. made 5.--: ".0le c 002'! - NIIX QDI'X . “FT-Iv. earn: a 21x nmrlx a 31x c a 0.1- ’0 r I . I . u l to O I C. I .- ld row- / lo 940...; lilo J . ,7 0% n mnhk.c QN [mi 9 110 around 1.5-2.0K. At T < lK, the data of all the samples almost overlap. .Above 1K, the data diverge from sample to sample, but can be divided into two groups. In the first group, data of samples K-HSa, K-HSb, K-H6b almost overlap and are close to sample K-l of Lee from 2.0-4.2K. These data stay less negative than the others. These three samples were under 10 qu He gas without molecular sieve before being cooled down. The second group of samples, K-ll (under Ar atm. at room temperature), K—H6a (under He atm. with molecular sieve at room temperature), and K-H6c (under 100 qu He with molecular sieve at room temperature) almost overlap each other over the whole temperature range, and are more negative at T > 1K than the other data. Both groups have the general form G = G0 + B*T2 + (C*/T) Exp (~68/T) (see section 2;”. Assuming 9* 23K, for the first group we got cg = -0.17 i 0.01v‘1, 3* -0.314 1 0.0004v"1z<"2 and c* = 4100 i 100v‘1K for the second group 00 -0.075 1; 0.015v‘1, 3* = -0.484 33 0.005v‘1K‘2, and c* = 6200 i 140V'1K. These two fits are given in Fig. 4-3 by the dashed and solid line, respectively. The reason for the difference between these two groups might be explained as follows. Without the molecular sieve, the little amount of He gas (104qu at room temperature) could be incompletely adsorbed, the heat current from the G heater only partly passed through the sample, and part of it through the left—over He gas. But in the calculation, we still assumed that all the heat current passed through the 111 sample, (iJL, we divided the electric current I by too big a heat current, so we got a smaller absolute value of G). The small bumps could be explained as a kind of condensation of left-over He gas occurred from 2K down to 1.5K, so the data fell back to the position where they should be, and overlapped with the second group. The heat current passed through the left-over He gas mentioned above could be roughly calculated as follows: first we assume all the surface exposed to the 10 uHe gas inside the sample can can only adsorb one layer of He atoms at 4K. The total surface area is roughly calculated as: S = 25 cm x 10 cm x 2 (the height of the sample can z 10 cm, the perimeter of the sample can 2:25 cm, the areas of two bottoms, sample holders, and samples are all included in the factor 2). The atom spacing of liquid He is about 7.53. So, the number of He atoms that could be adsorbed is: The total number of He atoms inside the sample can is 23 -3 n = 6.023 x 10 x 10 x 10 mm x 500 cm3 22.4 x 103 cm3 760 mm z 18 x 1015. (where v z 500 cm3) So, the left-over He gas would be half as much as at room temperature. To find the thermal conductivity (I(= l/3 CVIV) of the left-over He gas, we need to calculate , 112 CV, and V: ‘l = 8 cm (the diameter of the sample can) Cv = n x l/2k 6.023 x 1023 10 x 10'3 mm = .5 x x x .5 x 22.4 x 103 cm3 760 mm 1.38 x 10"16 erg/K = 1.2 x 10‘9 J/Kcm3 Using the velocity formula of ideal gas, v = 1.6 x/RT/fl = 1.60 :v%.3128 x 4 x 107 erg/4g = 15 x 103 cm/s. Thus, x: 1/3 cvlv z 4.7 x 10‘5 watt/cm-K. So the heat current passed through the assumed left-over He gas would be approximately Q = KSAT/AX ==4.7 x 10"5 watt/cm-K x 3 cm2 x 0.1 K/4 cm s 3.5 x 10‘6 watt. This is comparable to the heating power of the G heater: 96 = 123 = (40 pA)2 x 400052 = 6.4 x 10'6 watt. 4.2 we (a u p ['1 Z C Ii 0 l . 111G-‘RTY 4.2.1 W111i! As mentioned in section 1.2.1, Rowlands et al., unlike other investigators, found that below 1.3K, their p(T) data of pure K samples were better fitted by p(T) at T3/2 than by p(T) at T2. Subsequently, their data were interpreted as evidence for electron—phason scattering (section 2.2.3). We noticed that their samples were thinner than ours, with d = 0.79 mm. After reviewing the old data of our group, we found that the plotting of dp/dT versus T for two pure K samples (d = 0.9 mm) of 1.1. Haerle cooled in He gas also showed deviations from a straight line, i.e. deviations from T2 behavior. In hOpes of discovering the reason for the difference between the data of Rowlands et a1. and our thick sample data, we set out to measure a set of thin samples with d = 0.5, 0.25, 0.16, 0.1, and 0.09 mm in three different circumstances, Ar atm., He atm., and partial vacuum. We found that data for wires cooled in He gas displayed a clear pattern of unusual behavior which is consistent with that reported by Rowlands et al. in the region of overlap, but more complex in form. The characteristics of all these samples are shown in Table 3-1. Fig. 4-4 shows a normalized dP/dT (ref. 59) plotted versus T from 0.07 to 1.8 K for selected samples with 0.0938515 mm, prepared in He gas and cooled down with 114 molecular sieve. They were, in sequence from t0p to bottom in Figure 4-4: samples K-Hla with d==145 mm 00, K—l/2H with d 0.5 mm (I), K-1/4H3 with d = 0.25 mm (x), K-1/4H2a with d 0.25 mm (c), K-1/4HZb with calculated d = 0.16 mm (o), K-l/10H2a with d = 0.1 mm (9), K-l/lOHlb with calculated d = 0.09 m (it), and K-1/10H2b with calculated d = 0.09 mm (o). The characteristics of these samples are given in Table 3-1. Two samples of each diameter were measured concurrently. The data for such sample pairs always agreed quite well with each other; one of the largest disparities is indicated by the different symbols ({tand 0. Thin samples had smaller r (iJL, smaller ylw ratio) and thus less contribution from any (r/x)2 term. According to the Rowlands et al. model, our thin samples should have had a more positive dP/dT, exactly the opposite of what is seen. Finally, none of the models of deviations from Matthiessen's rule used to describe size effects in other metals can generate a negative dP/dT (see section 2.2.5 for details). The only model we know, which might explain what we see, involves the combination of contributions from (a) UEES, giving a positive term, 2AT, in dP/dT; and (b) NEES plus surface scattering, giving a negative term in dP/dT, [suggested by Gurzhi] (section 2JL6), which at low enough temperatures increases in magnitude with increasing temperature [as calculated by Black (ref. 38)]. The main difficulty in directly comparing our data with this model is that the best estimates of 1126, (Normal electron-electron scattering mean free path) for K yield lge ~lO-100 mm at 1K (ref. 39). The diameters of our samples which show size effects were d = 0.09-0.9 mm. This means that we are not in the Gurzhi limit l§e<>d. A Monte Carlo calculation by 122 Black (ref. 38) can be used to estimate behavior in this limits.Table 4-2 shows some of the results from Black. Since 1 (mean free path of electron—impurity scattering) ei is temperature-independent in first approximation, and d changes only 2% from room T to lowest T, lei/(d/Z) is roughly constant. Unlike lei' 129 is temperature-dependent, lge T'Z. Thus, lge/(d/Z) decreases with increasing temperature. This means that in Table 4-2, the direction of decreasing lee/(d/Z) (to the right in Table 4-2) is the direction of increasing temperature. Since Pal/leff, then dP/dT a: -(l/leff)2 dleff/dT, where 1eff is the effective mean-free- path of electrons. This means that an increase of leff with increasing temperature gives a decrease in p and thus, a negative dP/dT. In Table 4-2, when lei“d' which is the situation for our thinnest samples, negative values of dp /dT persist to at least lg‘e/(d/z) 210. When lei/(d/Z) a 5, Table 4-2 shows a reduction in the total resistivity P of about 1% (the limit of accuracy of Black's calculations) from 1ge/(d/2) sec to lge/(d/z) = 10. In the case of our thinnest samples, lei/(d/Z) z 4.4, (d = 0.09 mm and lei = 0.2 mm calculated above), life/mm)- 200-2000 at 1K, and the reduction of p from T a 0K to T = 1K (i.e., from 126/(d/2) = aoto lg‘e/(d/z) z 200-2000) is only 0.003%. So, the calculation of Black could be consistent with our data. The fundamental theoretical questions are whether for lei ~ d the Gurzhi effect dominates the Knudsen—flow effect at such large values of lge/(d/Z) and, if so, whether the 123 «hqo.o ewv.o «m.m 3H.mH mNo.o ch.o vmo.o mmH.o mv.o om.o amv.H «m.N «N.v em.m no.0 .mHH ou mucusoom mum mosam> sonuo Ham .wNH ou oumusoom mum mmzfim> w Neo.o sec.o meo.o meo.o meo.c Nso.o emo.o eao.o eac.o emo.o mmo.o amo.o NNN.o omN.o NNH.o om~.c aNN.o mNN.o Ne... 35 and and and and cN.o NN.o oo.s mo.o No.o cc.o om.H 4H.H em.o mm.o mm.o Nm.o «NN.N om.H sm.a a~.~ HH.H oo.H «N.N NN.N mm.H em.~ NN.H NH.H we.e em.N cm.H mm.H NN.N Hm.H H.o N.o m.o N N m c\ooN N meo.o emc.c omH.c mm.o mm.o mm.c vc.H mH.H om.H AAmm .umuv xomam soummv mcfiumuumom ofim0uuomfi mow mama ."""i'--“"'"-'ii"-‘I‘-’ Neo.c mo.c eao.o H.e mNH.o N.e mm.o m.e am.o N om.c N mo.H m mN.N OH NN.H mos mes c\aoH N .Nue caste 124 Gurzhi effect is large enough to be seen. A much more accurate Monte Carlo calculation will be needed to answer these questions. If the Gurzhi effect is not the source of the behavior we see, then there currently exists no satisfactory explanation for what we see. As mentioned above, since we are not in the Gurzhi limit lge< 0Q .m.: .m_m . Tee. 3 0 ON . 0.0? om 0.0 QV O N O _ . l 2 g d i A _ . _ A q 1 1.“ 40H. -\ w 1 11...-111111H111\l1\ \ 1 11 . 111111 \ 1 3 111111 1.. \ l 1\l11111-1 1 1. . \11. 1 1cm \ 1 1 1 1 1 M 1 .d 1 1 1 109.. . 1 M. 1. - ,_ 1 _ ,, 1 3 dc) c. a I 10.9 .< c. _ w: E o - 132 not yet clear, the 90's of our thinnest samples (d g 0.1 mm) showed a lot of scatter, with most of the pb's much larger than expected. Possible explanations include: (1) Uncer- tainty in determining the diameters of these thinnest samples. The diameters were determined by the hole size of the dies. The smallest die (d = 0.1 mm) could easily get partially blocked or slightly damaged, which would reduce its hole size and make the extruded samples thinner. The 733 of these samples would thus be higher. (2) Corrosion of sample surfaces. This would thin the samples. For the thinnest samples, the thinning due to corrosion would be significant. [(1) and (2) together could make the real d half as small as the hole size of 0.1 mm dieJ (3) Impurity input from the sample surface during extrusion. This would give the thinnest samples the largest fractional concentra- tion of impurities, which would increase their figfls. 4 . 2 . 2 The Thermeilww ' G measurements were made for each thin K sample from about 4.2K down to about 70 mK. Fig. 4-10 shows G data for the same samples shown in Fig. 4-4, which were cooled in He gas. For K-Hla (d = 1.5 mm), G measurements were made only for T 1 1K, so data of another thick sample, K-H6a (d = 1.5 mm), are added to Fig. 4-10. All these data have similar form, and can be well fitted by G = 60 + B*T2 + (C*/T) Exp(-9*/T) . 133 otozdmosum m: m c_ uc_ooo x no mot_2 :_:u tom 5 m> o .c_1: .m_m c: e _ lam 14 _ I C 5 . / \ x .‘I ’l/ o .1. .... \ . .. x. o \x\ 1.111! I. \ / O ..\ , . o . 1oN- Edd: n..." .6 es sodas A. Sea: 5.20 .q 88 moduu rlHu Eco... Yawn Q 68 afiouu Eco: 34.... .e 55 one"... Soc... 2..“ u .q 88 mud no 110 —.1 800.: am; u .q 8:. mad at W74. 01* / 80%: 3..“ .q 8:. mduu 11m1x . - [I’M/7:17.76 Ends 3..." .e as 31.. 8:1: .4 V sods N31 .6 SE 31.. 2:15. + 134 But, the thicker samples show larger magnitudes of G at high temperatures and smaller magnitudes at lower temperatures than do the thinner samples. Table 4-4 shows the coefficients of the fits Go' 8*, C* for each sample with 9* fixed as 23K, lines in Fig. 4-10 show the fits for three samples: K-1/10H2a, K-l/4H2a, and K-H6a. The fit for sample K-H6a was already shown in Fig. 4-3 (section 4.1.2). Table 4-4 shows that as the diameter of a sample decreases: 1) the magnitude of Go gets larger, and 2) the magnitudes of B* and C* get smaller. "1" can be explained by the Gorter-Nordheim rule (Eqn. 2-45) 8 = (pisi + Pdel/P (see section ZdLl), where S is only the diffusion term of the thermoelectric power, S- 1 = thermopower due to impurity scattering, pi = electric resistivity due to impurity scat- tering, Sd = thermOpower due to another temperature- independent scattering (here considering only surface scat- tering) = pd + pi = total residual resistivity; When d decreases, pd increases. If Sd does not change in magnitude and always has the same sign as Si' then S would increase in magnitude. But, since 8 = GOT, this means that Go would also increase in magnitude. "2" can be explained as fol— lows. When a sample is thin, electrons and phonons are scattered from the surface more often, phonon drag (both normal and Umklapp) gets weaker, and the magnitudes of B* and C* get smaller. Fig. 4-11 shows G data of four K samples with d = 0.1 mm cooled in Ar gas, (ladata for these samples were 135 Table “-“. The coefficients from fits to the 0 data of thin K samples cooled in He gas Sample d(mm) 00(V'1) 3*(V‘1K’2) c'(V"K) K-H6a 1.5 -0.075 1 0.015 -0.“8“ 3; 0.005 6200 _-1_-_1“0 K-1/2H 0.5 -0.166 _-1_-_ 0.010 -0.“65 : 0.00“ 6090 _-1_-_ 120 K-1/“H3 0.25 ~0.301 ;1-_ 0.013 -0.“18 :t. 0.006 5360 i 160 K-1/“H23 0.25 -0.302 x 0.008 -0.382 1'. 0.00“ “8“0 ; 110 K-1/“H2b 0.16 -0.322 -_1-_ 0.010 -0.329 1 0.00“ “020 1,100 K-1/1OH2a 0.10 -0.“88 1 0.01“ -0.257 1; 0.00“ 3280 g 100 K-1/10H2b 0.09 -0.“83 i 0.023 -0.251 3; 0.011 3190 1 220 K-1/10H1b 0.09 -O.“87 1 0.022 -O.255 ; 0.011 3260 1 210 136 of .mtchmoeum t< m c. po_ooo x to mct_: c_£u to» b m> o 4.13.3... ohm C: ._. O_.N E00,: and u.q Ede: 56 «.61 Ewe: mm.» H .d 50%: 3..." u .Q 0.? 1. 6831.6 5:. cows 55 town 55 ..out 3pm.; 3..."— 15. Seq. 1: Emmi. d C D D 137 shown in Fig. 4-8). These data show similar form to the data shown in Fig 4-10, and can also be well fitted by G = 60 + B*T2 + (C*/T) exp(-e*/T). The fits of samples K-l/lOAl and K-l/10A4 are indicated by two solid lines in Fig. 4-11. Table 4-5 shows the coefficients of the fits of each samples with 6* = 23K fixed again. Table 4-5 shows that for four thin samples with d = 0.1 mm cooled in Ar gas, (1) unlike dP/dT, Go is almost sample-independent, and is close to the Go of samples with the same size cooled in He gas; (2) B* and C* show sample- dependence, and tend to behave like data for samples in He of slightly larger diameter. The reasons for (l) and (2) are not yet clear. A trial explanation is as follows. As described in section 4.241, for dP/dT, more impurity scattering and less diffuse surface scattering of electrons would make the samples behave like one with thicker size. For 60' things are different: from the Gorter-Nordheim rule (Eqn. 2-45) 1 S = '73:)" (9181 'i' Pde). More impurity scattering will bring 01 and no up, and less diffuse surface scattering will bring pd and 0 down. Both 0 together could keep S unchanged, and thus Go unchanged. For phonon drag, more impurity scattering of both electrons and 138 Table “-5. The coefficients from fits to the C data of thin K samples cooled in Ar gas Sample d(mm) 00(771) B*(V'1K‘2) C'(V'1K) K-1/10A1 0.10 -0.““5 ,1; 0.027 -0.259 1:. 0.010 3500 3, 2“0 K-1/10A2 0.10 -0.“66 : 0.019 -O.261 .t 0.007 3280 i 150 K-1/10A3 0.10 -0.“19 _~1_-_ 0.03“ -0.3“1 3 0.011 ““90 1 220 K-1/10A“ 0.10 -0.“65 3 0.0111 -0.312 1., 0.011 11070 i 250 139 phonons will make phonon drag weaker, and less diffuse surface scattering of electrons will enhance phonon drag, these two factors coming together could explain why B* and C* were unstable, but less unstable than dp/dT. Fig. 4-12 shows G data of four K samples with d = 0.10, 0.25 mm cooled in partial vacuum (10 “Hg He), (Ivdata of these samples were shown in Fig. 4-8). Each set of these data shows a "bump" in the vicinity of 1K. A possible explanation for these bumps was already given in section 4JL2. For thinner samples the heat conductance is smaller, so a smaller part of the heat generated by the G heater passes through the samples; thus the error in the G calculation (discussed in section 4Ju2) is bigger and brings the data further down in magnitude. Except for the bumps, the shape of these data is similar to that of the data in Fig. 4-10 and Fig. 4-11. 4.2.3 W The thermoelectric power 8 of one K sample (K-l/2H) was measured in this study to find the form of L(T) and the low temperature limit of L(T) = Lo (Eqn. 1.8) in K. The reason for choosing a sample with d = 0.5 mm is as follows. Since S is calculated from S = V/ATI 1:0, the sample must be thin enough so that its heat conductance is small enough to give an accurately measurable AT across the sample with a small enough heat input to keep the sample cold. But if the .E::oc> _m_utmd m :_ pm_ooo x we mmt_3 :_£u to; h m> u .~_1: .m_m ed on a: b oA.N 3 _ 140 A 7 code 8;; 583616 N>1m1x . sources": 85363 tfiwux l- sods amend seenowc «1an o + sods a: u .e 5:. one" >21 ++ + U n .m. X o 9 .++ + 11AHfiWI n x x + . . A + x N 1» + x x + x e 10w. 1*! x ix ++~ O O O O O O ”0. o . do . o o o o o o #1. o 00 0 o 0 +x $ 0 o o i .1 c w .o + N+~ 141 sample is too thin, it would show a strong size-effect. d = 0.5 mm was thin enough, but not too thin. Fig. 4-13 shows a plot of S versus T. The dashed line in Fig. 4-13 indicates a fit of s = A'T + 3'T3 + c'e'9*/T (section 2.3.1) with 6* fixed as 23K, as used for G data analysis above; the coefficients obtained were: A! = -0J78 i 0.15 x 10‘8 v‘l, 3' = -0.681 5; 0.034 x 10'8 v‘lx‘z, and c' = 11530 i 770 x 10"8 v‘lx. A better fit was found when 6? was left as a variable parameter; the coefficients were then: 9* = 17.4 i 0.4K, A' = -0.433 1 0.081 x 10‘BV'1, 3' = -0.957 3: 0.045 x 10‘8 vx‘z, and c' = 4070 i 250 x 10"8 V’lK. This fit is indicated by the solid curve in Fig. 4-13. Fig. 4-14 compares measured G with G* calculated by G* = S/LOT. A divergence occurs when T > 1K, this implies the temperature-dependence of L (see section 1.1) for K at T>1K. In order to see the form of L(T) more clearly, we plot L(T)/Lo versus T in Fig. 4-15. The values of L(T)/Lo were obtained from G*/G = (S/LOT)/(S/L(T)T) = L(T)/Lo. In Fig. 4-15, one can see L(T) z LO only for T < 1K. From about.lK up to about 4K, L(T) decreases with increasing temperature. Each value of L(T)/Lo in Fig. 4-15 was checked by the Wiedemann-Franz Law as follows. Since, 0 = xAAT/l, R = pl/A, and xp/T = L(T) (section 1.1); 142 .:~\_ux m_aEmm LOm h m> m .m_n: .m_m . . v: . oe on on Q_ o _ _ _ u 1O.m— S 5.2 1 m. 0. A8 M. on [ r/l/J/IIJ 3-5 143 .:~\_nx m_aEmm tom k m> Apob\m uv¥o wen u .57: .2“. o6 o_.m c: e ow. o_._ o _ . m .e 13.. . e . . 9 o o o - l._o.N.l . N, . m. -3- 144 .:~\_-x ..eEm. tot e m) ol\AeV5 o.— d .93 .3“. 145 SC, QR/TATLO = Kp/TLO = L(T)/Lo, where Q can be calculated by Q = R6175; as mentioned above. Table 4-6 compares two sets of values of L(TL/Lo, calculated from G*/G and from the Wiedemann-Franz Law. One can see these two sets of L(T)/LO are very close to each other. 4.3 NQMlflAL.31flfiLE:QBX§IAL.K_§AflELE 4.3.1 J2h3__BgS.i..3.‘s’..i.’:L.'L.t’..lL.w As mentioned above, in an attempt to get a longer mean- free-path of electron-impurity scattering, so as to see whether the size-effect could be observed in a thicker sample, we tried to make and measure a "nominal" single- crystal K sample with d = 1.5 mm: sample K-S. The word "nominal" is used here because, since only one X-ray spot on the sample was tested, the verification of the whole sample as a single crystal was not established (section 3am. Sample K-S had one of the lowest Po's and one of the largest RRR's for K in this study (p0 = 0.92 chm, RRR = 6150) (the details of its characteristics are given in Table 3-5). However, the mean-free-path of K-S was not much longer than the previous ones; lei = 0.5 mm was obtained from Eqn. (4-1), still much smaller than d (d = 1.5 mm). As expected, no clear size-effects were seen in this sample. Fig. 4-16 shows a plot of dP/dT versus T. The data at around T = 0.7 K look a little low, but not low enough to 146 Table 4-6. L(T)/Lo values calculated from G*/G and from Wiedemann-Franz Law for sample K-l/ZH T(K) L(T)/Lo from G*/G L(T)/LO from W-F Law 0.482 1.00 1.00 0.714 0.98 0.98 0.948 0.96 0.96 1.182 0.95 0.94 1.368 0.91 0.91 1.483 0.91 0.89 1.741 0.81 0.81 1.941 0.76 0.78 2.201 0.76 0.75 2.490 0.67 0.66 2.808 0.62 0.61 2.977 0.58 0.56 3.874 0.42 0.40 147 6,0— / _ / r / eso- ,/ C} / c: / / ’l: 4.0" / Z < / 2 A :4 3.0— / , Q. ,A g’ / / 2.0- / a / / 1.0— / / . ///A ()t’ A I l l l .J i O 0.2 0.4 0.6 0.8 1.0 1.2 T [K] Fig. 4-l6. (IKh.2K)/P)(AP%AT) vs T for the nominal single crystal K sample. ' 148 show clear evidence of size-effects; while the data at T < 0.2 K show a clear turn down. In order to see the anomalous turn down, dP/TdT is plotted versus T in Fig. 4-17. It is clearly seen in Fig. 4-17 that the anomalous turn down occurs when T < 0.3 K. This anomalous turn down is completely different from the size-effects described above. In section.4.4iwe will see similar turn-downs for K samples encased in polyethylene tubes. Further discussion of this turn-down will be given in section 4.4. 4.3.2 W G measurements were made for sample K-S from about 4.2K down to about 70 mK. Fig. 4-18 shows a plot of G versus T for samples K-S (A) and, for comparison, K-H6a (x) which was plotted in Fig.14én. These two sets of data almost overlap each other for T > 1.5 K. For T < 1.5 K, the data of sample K-S stay slightly less negative than K-H6a, a result consistent with the Gorter-Nordheim rule (Eqn. 2-41) (see section 4.2.2). 4.4.1 W57. As mentioned in section 1.241, van Kempen et al. (ref. 14) and Greenfield et al. (ref. 18) measured F>of high- purity K samples inside polyethylene tubes down to about 1K. The details of the results from these two groups are given in section 1.2Ju Tb investigate further the variations in 149 1 l 1 Fig. 4-17. I 0.4 0.8 1.2 T [K] (P(L+.2K)/PT)(AP/AT) vs T for the nominal single crystal K sample, 150 .m_asmm x .mum>tu e.mc_m .mc_eoc ecu to» k U: ._. 2, o .23 .2“. (I: 151 magnitude of the T2 coefficient reported by both of them, we measured four pure K samples clad in polyethylene tubes down to 0.07K. To test for annealing effects, each sample was measured several times, with intervening anneals for varying periods of time at room temperature under Ar gas, He gas, or partial vacuum. Sample K-PHl was in a 1.6 mm dia polyethy- lene tube and was measured 5 times under He gas and partial vacuum. After the second measuring run it was taken out of the sample can. It was remounted just before the third run with a new bare K sample as the reference. Sample K-PHZ was in a 0.9 mm dia polyethylene tube and measured 4 times under He gas with a reference K sample inside a 1.5 mm dia teflon tube. After the third run, one of its connections broke and the sample can was opened and the sample repaired. It was remounted just before the fourth run using the same refer- ence sample. Samples K-PAl and K-PAZ were in 0.9 mm dia and 1.6 mm dia polyethylene tubes respectively, and were meas- ured concurrently 5 times under Ar gas. To test for the rolling effect reported by Greenfield et al. (ref. 18), both of these samples were taken out of the sample can after the third run, rolled, left at room temperature for about 20 hours, and then remounted just before the fourth run. After the fourth run, both samples were again taken out of the sample can, and later remounted for the fifth run. The details of the preparation of these samples are given in section 3.4. The details of the characteristics of each sample in each measurement are given in Table 3-5. 152 Fig. 4-l9 and Fig. 4-20, in different scales, show a normalized dP/TdT (ref. 59) plotted versus T for five runs of sample K-PHl. Similarly, Fig. 4-21 and Fig. 4-22 for four runs of sample K-PH2; Fig. 4-23 and Fig. 4-24 for five runs of sample K-PAl; and Fig. 4-25 and Fig 4-26 for five runs of sample K-PA2. For comparison, the data of sample K- S plotted in Fig. 4-17 were added in Fig. 4-19; and the data of samples 2a, 2b, 2c from van Kempen et al., which were also measured under He gas, are added in Fig. 4-21. At T > 1.1K our data in Fig. 4-21 are consistent with the data for samples 2a and 2b of van Kempen et al., which had Po's comparable to ours. Sample 2c of van Kempen et al. had a lower F5, and its data stay well below our data (see Fig. 4- 21). The dashed curve in each of these figures shows, for comparison, typical behavior of a free-hanging, bare, high- purity, thick K sample (section 4.1.1). At T < 1K, the data presented in these figures, including the data of sample K— S, which was melted in clean paraffin oil during its growth, exhibit similar deviations from the dashed curve. For samples encased in polyethylene tubes, sample annealing at room temperature significantly brings both the normalized value of dP/TdT and Po down until dP/TdT becomes negative and large. To fully developlthe negative deviation requires about 10 days of annealing. The temperature at which dP/dT crosses zero corresponds to a resistivity minimum, which in the fully deve10ped samples occurs between 0.3 and 0.5K. When first cooled and measured, the samples appeared to have l l A \ A \\ A A A ‘ 0 8°0— \\ o I H ‘ \ 0 A N ‘\~~_‘__-__—-____-_.--_--_----- ' at g * ‘k * . O . 3 A E 00 I I ‘ e . _- . 1.74.0— *° Q - ‘ Q. A _ . £5 * o Nocfldays 1% g at roomT (ngcm) E A ‘ A K-PH'Ic 0.5 2.70 63.2.0L' . K-PHlb 11.0 1.42 0 . ° K-PHic 24.5 0.90 *A - K—PHld 26.0 0.95 ' K-PHle 27.0 0,95 0 e *r K-3 0 2.0— . u l 1 0 0.4 0.8 12 T [K] Fig. h-l9. (P(h.2K)/T)(AlnPh&T) vs T for sample K-S and five runs of sample K-PHl. 154 \“ ‘A-——---O~ “’A —————————— O.— —-AQ25—5-‘—°LA—OA— _‘ _____________ AA 00‘ A .- ‘ ‘1'" 8 ' X g' ‘ E ' . g 40.0- ““ A TS: _ Q A CL S éaoo— 4 p- \. SI. ‘ A K-PHla A K-PHlb 120.0- 1 0 K-PH‘I: ‘ . K-PHld I K-PHle ‘ 1 l 1 160.00 Fig. h-ZO. 0.4 0.8 [K] 1.2 T A version of Fig. 4-19 to larger scale. 155 F No of days 10.0" at room T p" ("9““) ° 0 K—PHZa 0.5 3.02 " i I K-PH2b 2,5 LIO A . A K-PHZC 13.0 |.l7 8.0- ‘ 0 K-PH2d 73.5 I.04 w ‘7; x * \ 0 £5 \ c ‘.‘ I O u— ‘\ o 0 " EMO“ " A ““~ ______ __,Q_._——-(-)———-I-— l2] 0 7° . \\ A * QE 0 ' A t 0 a ' . - ' I 4.0— A. < o . O Q; I V o A O 253'“ O G . Annealing P( ) o time(daysl ° chm A Data 11' 2a 2.39 from van it 2b 2 |.|3 O Kempen! 0 2c 80 l 0.88 _ l 2'0 0.5 1.0 1.5 T [K] Fig. 4-21. (P(L+.2K)/T)(AlnP/AT) vs T for four runs of sample K-PHZ. For comparison, the data of samples 2a, 2b, and 2c of van Kempen et al. are indicated. 156 10.0 \ \\\\“~-___O _______ O---S——.—A——-- o o . g I' O I 0 00 72.5; ‘L o “z E .- c —— A t- .- A IA % -2o.o—— o. '- 5 A 4 e a: __ I o K—PH2a \\Q . I K-PH2b _40.0__ A K-PHZC ' o K-PH2d 1 l J i 1 J i 0 0.4 0.8 Fig. 0-22. T (K) A version of Fig. h-Zl to larger scale. 157 __ No. of days R, 10-0 at room T (chm) o O K-PAla 0.5 5.|9 - K-PAlb 4,5 1,20 A K‘PAlc l6 Lid 0 K-PAld I9 L32 A 8.0- o K-PAle 97 |,|2 ‘ O l 3‘ O 7 \ ' =1 6.0— X o -' E “~__---._- 0 o c O -—-——-—---——---—--—-—-— A t o I O C 4.0— ° 0° e ° ' . E C < I .. . 3 0- o 0. \~ 20' . s“ ’ 1 - g ‘2 l 0i i l I A 1 l l -2.0 ' ' l 0 0.5 1.0 1,5 T (K) Fig. L1-23. (P(L+.2K)/T)(AlnP/AT) vs T for five runs of sample K-PAl. 158 201 o 00 E oIA o 5 __ 8 q.“ o 5: _ .2 E5 , C} _ q. “3 A 1— - __ .° Q 50 Q. E ‘— as Si __ .2 ‘\:: _. SE? __ 8 K-PAla K-PA1b 4001- : 4...... ._ K-PAld 14 1 1 l 1 1 l 0 0.4 0.8 T (K) Fig. A-ZA. A version of Fig. A-23 to a larger scale. 159 10. No. 0! days R, at room T (chm) 0 K-PAZa 0.5 5.20 o I K-PAZb 4.5 0.97 8.0— o K-PA2c I6 Ll? II I K-PAZd l9 0.98 0‘ K-PA2e 97 0.73 o 1 o \‘ O . A \ O I N \ O A I _ \ x 6.0 \\‘ Q I A O E ““““““““““““ .‘ """"" o C} . ° 0 u— 0 V o I A0 o O p 4.0- . A <1 o. 0 ° - E. S) ° - 1: ' ° \2I0_ 0‘. O V V C I A 0 o o A _2. 1 1 1 G0 0.5 1.0 1.5 T (K) Fig. A-ZS. (P(A.2K)/T)(AlnP/AT) vs T for five runs of sample K-PAZ. (meK'2) ( ”/T) (AInP/AT) 160 20 d) t"---1L---‘ .VK“‘§“‘ c 8 ° .3 0 I O A _ O I~A I 1— IA I A '40; 'A I I o K-PA2a — I A I K-PA2b A A K-PAZC ‘80P- <>K-PA2d 1 1 L 1 1 l 0 0.4 0.8 T Fig. 4-26, A version of Fig. 4-25 to a larger scale. 161 a rather large concentration of impurities or defects, as indicated by the relatively large value of P .About 3 o' days of room temperature annealing seemed adequate to bring Podown to its normal value for free-hanging, bare, thick K samples. For reasons not yet clear, sample disturbance during sample remounting at room temperature seemed to bring dP/TdT up, particularly for data at low temperature end, but did not affect 13 significantly. The effect of sample rolling was not simple. For K-PAl, after rolling, no in- creased, but dP/TdT did not change significantly; while for K—PAZ, 90 did not change significantly, but dP/TdT increased. For a possible Kondo effect, the electrical resistivity at low enough temperature is expected to have the form p= p0 + AT2 - BlnT (section 2.2.7). Thus, dP/TdT = 2A - BT72. In this case, a plot of dp/TdT versus T"2 should be a straight line of negative slope -B and with an intercept of 2A at T'2 = 0. Figures 4-27, 4-28, and 4-29 show plots of dP/TdT versus T"2 for samples K-PHl and K-PHZ, K-PAl and K- PA2 respectively; All the data in these figures are well fitted by dP/TdT = 2A-8T‘2, as indicated by the broken lines. It is also clearly seen in these figures that sample annealing at room temperature enhances the effect (brings the magnitude of coefficient B up) while sample disturbance appears to weaken the effect (brings the magnitude of B down). This can also be seen in Table 4-7. '162 0—27. Fig. 163 h —————— a ————— -— O’RE Q &\ \ \\\ %£> \ -20.- °\\\s \ \ \\ \ \\ \ ,. Wx 1 N 9 \ l 1" \\\\\\ 1 —-4o__ \ K \ l E \\\ \5 l C: Q \\ l 1: \\ \ ‘ \ \ i ‘ \ \ . \ \ \ 1 1 \ \ \ ’60?- \ \ \ \ :- ° \ \ ‘0 \ \ \ l <\] \ \ \ i % \ \ \ — \ \ o $1 ‘1 \ \ 1 ,._£3 L_. IQ'PAJ \ \ a o O \A .\ \ .' I \ \ A C \ \ 0 e \\ \\\ \ \ .- ‘4C“D" \\ A\ i \ \ \ \ \ \ \ \ \ \ \ . 10:1 _ l l l - O 20 40 60 80 T"2 (K-zl Fig. h-28. (P(h.2K)/T)(AlnPA5T) vs T.2 for sample K-PAl. 164 —o-——_—-1> —————— 0* _____ —0————_.a._ 0 ‘1 \ _____ \‘ \\o. \ \ \R \\ \O\ \ 'K \\\ \ °\\ \ \\ \\ a— X \ a 0x | -20 \\ \K \\ X \ \\ \ E \\ ‘1\ C} A \ l u— \ ‘\ \e. \ 9 40 \\ \K <" i \ Q- \\o \ I s \ \ g \\ \\ ; K‘PA2 \\ \\ \“6 o O \\ q \ Q: I b \A\ A C \ o e \\ b. -80 fi 1 l 1 0 20 240 2 60 T (K ) Fig, 4-29. (P(L+.2K)/T)(AlnP/AT) vs T-2 for sample K-PAZ. 165 Table 4—7. The coefficient B from fits to the 9 data of K samples encased in polyethylene tubes No. of d days at After sample Po B Sample Gas (um) room T disturbance (nflcml (farm) K-PHla 10 1.1119 He 1.6 0.5 Yes 2.97 0.14 K-PHlb 10 “Hg He 1.6 11.0 No 1.68 0.90 K-Pch He atm. 1.6 24.5 Yes 1.18 0.32 K—PHld 10 pHg He 1.6 26.0 No 1.20 0.47 K-PHle 100 pHg He 1.6 27.0 No 1.20 0.47 K-PHZa He atm. 0.9 0.5 Yes 3.02 0.16 K—PHZb He atm. 0.9 2.5 No 1.10 0.72 K—PH2c He atm. 0.9 13.0 No 1.17 0.87 K—PH2d He atm. 0.9 73.5 Yes 1.04 0.55 K-PAla Ar atm. 0.9 0.5 Yes 5.19 0.06 K-PAlb Ar atm. 0.9 4.5 No 1.20 1.39 K-PAlc Ar atm. 0.9 16.0 No 1.14 1.50 K-PAld Ar atm. 0.9 19.0 Yes 1.32 - K-PAle Ar atm. 0.9 97.0 Yes 1.12 1.23 K-PA2a Ar atm. 1.6 0.5 Yes 5.20 0.05 K-PA2b Ar atm. 1.6 4.5 No 0.97 0.83 K-PA2c Ar atm. 1.6 16.0 No 1.12 1.08 K-PA2d Ar atm. 1.6 19.0 Yes 0.98 -— K-PAZe Ar atm. 1.6 97.0 Yes 0.73 0.27 166 To test whether the effect described above only occurs for samples in contact with polyethylene and samples melted in clean paraffin oil, we measured samples encased in a teflon tube (K-THl) under He gas, a sample in contact with Kel-F pieces (K-KA) under Ar gas, and a sample coated by cold clean paraffin oil (K-O) under Ar gas. No similar effect was found in any of these three cases. To test whether there is some kind of impurity continuously diffus- ing from the polyethylene tube into the K sample and causing the ”turn down" effect, we measured a bare sample (K-PPA) with d = 1.5 mm and with potential leads encased in poly- ethylene tubes which were left in room temperature for 1.5 days before the sample was cooled down under Ar gas. No effect was found here either. Figure 4-30 shows a plot of dP/TdT versus T for these last four samples. Sample K-THl was measured 4 times con- currently with sample K-PH2 described above. The details of the characteristics of each sample in each run are given in Table 3-5. The data of samples K-KA, K-PPA, and K-O are similar to those of bare, free-hanging, thick samples. The data of K-THla (the first run of sample K-THl with d = 1.5 mm and large p0) stay more positive than those of bare thick samples. ‘When the annealing time at room temperature increased, the residual resistivity Do dropped, and the values of dD/TdT shifted downward with maximum shift at temperature around 1K, so that the data no longer formed a horizontal straight line. Since the smallest Db of sample 167 10.0- on 0 A O c?“ __ o x89 . E1 0 I ‘ ‘ g '— ..O ‘ a o o c 8 0 0 . x . o 0 if: X 5*: ‘ X ‘ x X ' ,0 X if 6. {3* 9* o o f: 0 oft * A |:- b t t a fit a A O Q _ ‘ .0 . . CL _ o o .5 s K-KA <1 01_ : 4. at K-PPA '- x K-O \N' _ No. of days Po .7 at room T (ngcm) 0 K-Tch 0.5 8.!0 2.04- o K-THlb 2.5 4.87 ‘ K‘TH‘C I3 L34 - o K-THld 73.5 0.94 01 l l l 0 0.5 1.0 1.5 T (K) Fig. h-30. (P(h.2K)/T)(AlnPAdT) vs T for K samples in contact with other plastics. 168 K-THl was 0.94 nacm, which yields an electron mean free path 0.26 mm (Eqn. 4-1) much smaller than the diameter of 1.5 mm, it seems unlikely that the deviation from horizontal line behavior is due to a size effect similar to that seen in the bare thin K samples. The reasons for the deviation from the horizontal straight line are not yet clear. Further experiments recently done in our lab (ref. 68) showed that application of an average magnetic field of 0.1T longitudinally'to polyethylene encapsulated samples shifts the resistance minimum to much lower temperatures. Thus, these samples seem to be exhibiting a Kondo effect due to a magnetic impurity which has an effective magnetic moment equal to or greater than one Bohr magneton. Identification of this magnetic impurity is still in progress. It seems to be a product of the reaction between potassium and hydrocarbons. 4.4.2 W G measurements were made for each sample above, in each run, from about 4.2 K down to about 70 mk. Fig. 4-31 shows a plot of G versus T for five runs of sample K-PHl; Fig. 4-32 for four runs of sample K-PHZ; Fig. 4-33 for four runs of samples K-PAl and K-PAZ. For comparison, the data of a free hanging bare thick sample (K-H6a) shown in Fig. 4-3 are also shown as a broken line in each figure above. For the samples which have been annealed for only 0.5 days, the plots have the same form as the broken line; the deviations from the broken line, more negative at low 169 ._:dlx o_dEmm com A m> o ._m-: .m_u 00 4.0 u . IN A. x O o , ~ Imlv. 9 I ) I / A z - a (l I I. /o o z I An I o z . , 3 ’0 0 II 0 I /4 I I I I pp 0 I /¢ . a / o .o 13 I . I/. .4.» I WK / l I ./ v 80’ I 170 .delx o—dEmm to» h m> o .lea .m_m ON our 0 \ I x \ 4 Q I/ x .. I/ \. l I / '1‘ 8 x . ,, .. on \n , x I \I dI/ Dd \ Ix. / n- I x N0 0 / szlx 10.09 (1.111 171 o x 8 3 0 ‘ O O OI“) .0 \ A O ‘ o I o . ‘ O \ o“ . \ A O. I 0 ‘ ‘ A ‘ \ \ O I 00 O “ 0 ° \‘0 \. \ t o ‘\ x I. ~ I ‘\ \O . .‘ . . \ \ 00 o O \ I‘ o o 0 ‘ \ ’ O O \ \ I— J .A‘ . 4;) ° V g \ \ 0.. \ ‘\ \ \ <5 ‘ ° \ I A— ' 'ZQ 0‘ ‘—l(-PA2 00 -b AC ed 00 0 ‘L0 Fig. “-33. G vs 2!) T for K-PAl -—I 00 -b AC od 09 O O a I I I . I I I .A I 0 ’ I I O l o I I \ I’ \ a ‘-v 0 O I I I I I I I ‘0 . I, I I \.. ’ “---". 0 samples K-PAl and K-PAZ. CD 172 temperature end and less negative at high temperature end, [i.e., with smaller magnitudes of Go' B*, and C* (see sec- tion 4.2.2)], can be ascribed to a high level of impurities (see section 4JL2 for details), as suggested by the large value of Po. Above 1K, further annealing causes the data in Figs. 4- 31, 4-32, and 4-33 to follow more closely the broken line, but below 1K they exhibit significant negative deviations which increase when the values of B obtained from Figs. 4- 27, 4-28, and 4-29, increase. A Kondo effect can produce such negative anomalies in.CL Below 1K, our data can be well fitted by G = G0 + B*T2 + D*/T, where the D*/T term is what is generally expected for a Kondo effect. Fig. 4-34 shows the data below 1K for two selected runs of each sample, the solid curves in this figure indicate the fits. The values of parameters Go, 8*, and D* obtained from com- puter fittings are given in Table 4-8. The small bumps at T z 1.5K in the plots of samples K- PHla and K-PHld in Fig. 4-31 are similar to the ones for their reference samples: K-HSa and K—H6b, discussed in detail in section 4.1.2. The plot of G for samples K-KA, KAPPA, K-O, and four runs of sample K-THl are very similar to those of free- hanging, bare, thick samples. They are not provided in this thesis. 173 .x. uaonm zo_on ocaumcanou um mo>cau m:_uu_m eo_z ~ o .sm-s .m_c a: e or md o QF md o _ _ _ 1 1_._- A _ A A _ A A _ A _ 4 4 _ _ 04+. A mv- - l - lV.—.l .. . -3- - . - - \ - - 1 - - T as; \. - 0.7 - \ - no-.. Ill-K .. - of - .// -. l \\ - - //// \. lmo-. n/ 4182+: \\. - h.AJ 174 Table 4-8. The coefficients from fits to the G data of K samples encased in polyethylene tubes sample G(V‘1) B*(V'1K'2) D*(V’lK) K-PHlb -0.228 g; 0.055 -0.l66 i 0.058 -0.065 a; 0.014 K-PHle -0.116 1; 0.007 -0.354 r. 0.010 -0.055 1; 0.002 x—mzc -0.115 3,. 0.006 -0.379 1 0.006 —0.052 3: 0.001 K—PHZd -0.060 1 0.013 -o.393 3,. 0.011 -0.061 at 0.003 K-PAlb -0.177 i 0.009 -0.333 3; 0.010 -0.126 1 0.003 K-PAlc —0.157 i 0.019 -0.358 i 0.017 -0.149 1 0.006 K-PA2b -0.125 g 0.010 —0.380 1 0.011 -0.075 1 0.003 K-PAZC -0.192 1; 0.011 -0.379 1 0.013 —0.089 1. 0.003 175 4.5 K:BD_ALLQX_§AMEL§§ In 1980 C. W. Lee et al. from our lab (ref. 8,23) found that P = PO + AT2 + AiPoT2 for K-Rb below lK (see section 1.2.1 for details). Assuming G = Go + G3 + G3, they also found that phonon drag terms (Gg and G3) are quenched in K- Rb more and more as the impurity concentration increases and that Go is positive and roughly obeys the Gorter-Nordheim rule G0 = Gi + FP/Po (Gp-Gi) (see section 1.2.2 for details) In this study we have improved the absolute accuracy of our temperature scale, and also our measurement precision by nearly an order of magnitude, so that we can now detect changes in p(T) of ~2 parts in 108. These improved conditions allow more detailed examination of the form of iflT) and G(T) in K-Rb alloys. Moreover, using improved glovebox facilities, including an accurate, built-in weighing scale, the atomic percentage of Rb impurity can be determined more precisely. Four K-Rb samples were measured in this study. Samples K-Rbl and K-Rb2 were 0.077 at.% Rb alloys with d = 0.25 mm; samples K-Rb3 and K-Rb4 were 9.4 at.% Rb alloys with d = 3.2 mm. The details of the characteristics of these samples are given in Table 3-6. 4.5.1 Wis-tints: Fig. 4-35 shows a plot of residual resistivity (Po) versus the Rb impurity concentration (c) for samples K-Rbl, K-Rb4, and two samples of M. L. Haerle (with 0.38 and 1.3 at.%Rb respectively). The stars indicate the two data 176 i—‘I-T-i 0.2L C (at%) Fig. 4-35. PO vs c for K-Rb samples. 177 points from M. L. Haerle; the solid straight line indicates the best literature value of no per atomic percent Rb impurity (= 0.13 uQcm/at%) (ref. 67); the dashed line indicates the value of no per atomic percent Rb impurity obtained by C. W. Lee in 1980 (ref. 8,23). It is clear in Fig. 4-35 that: (a) the value of no per atomic percent Rb impurity for our new dilute K-Rb alloy samples (including the two samples of M. L. Haerle) is now in closer agreement with the best literature value than what C. W. Lee previously obtained. (b) For our 9.4 at% samples, the value of no per atomic percent Rb impurity departs from the solid straight line. This departure might partly be due to the finite concentration effects which yield no = a(l-c)c in the simplest model. Only for dilute alloy the impurity concentration c is very small, so (1-c) z 1, and thus no = ac, yielding a straight line with a slope of a in a plot of no vs c; for c = 9.4 at.% as in our case, (l-c) = 0.91, ifa = 0.13 Mom/at% as given above, then a' = a(l-c) = 0.12 1.1.9 cm/at% should be the effective value of no per percent Rb impurity. Actually, the value deduced from Fig. 4-35, would be 0.11 mum/at%. Fig. 4-36 shows a plot of normalized dn/dT (ref. 59) versus T for sample K-Rbl (with 0.77 at.% Rb) and two samples of M. L. Haerle (described above). The broken line indicates typical data for a free-hanging, bare, high- purity, thick K sample. The small turn up on the broken line occurs for T < 0.3K, which we tentatively associated 178 40 ‘ x x x 1.3%Rb x x x X x X x 30— -— A X Nx X A Q T." .- < \ 20— —' i3: % 0.38%Rb D D E U DD U D C] D U D _._. u D D E! 10— -- \ 0.077%Rb A ‘\ A ‘ A AA A A A K o l l O 0.5 T(K) 1.0 Fig. 0-36. (P(A.2K)/T)(AlnPAAT) vs T for dilute K-Rb alloy samples. 179 with residual dislocations (see section 44LJJ. In contrast, in the K-Rb alloys there is a low temperature turn-down which becomes progressively more apparent as the Rb concentration increases. This turn-down is very obvious for the 9.4 at.% samples, samples K-Rb3 and K-Rb4 shown in Fig. 4-37. ‘ The coefficient (A) of the T2 term was deduced from the data in the region close to a horizontal straight line in Figs. 4-36 and 4-37. We tried to fit A by A = A0 + Aipo (see section 2.2.8). A0 = 0.24 i: 0.04 chm/K2 is obtained, consistent with the one for thick pure K samples (see section 4.1.1), and A1 z 11 1; 2 x 10"6K‘2 is obtained for all of our four K-Rb samples. This value of Ai is slightly larger than Ai 2:8.5 x 10"6K"2 obtained by C. W. Lee, and clo‘ser to the theoretical values of 13.7 x 10‘6K“2 from P. L. Taylor (ref. 24) and 12.5 x 10'6 from Kus and D. 9L Taylor (ref.25). In order to intercompare the Po dependence of dP/dT for all of the dilute alloys, we have normalized the data for each sample shown in Fig. 4-36 to the data for sample K-Rb3 by multiplying by pK_Rb3(4.2K)/p(4.2K) and then displacing each curve vertically to bring it into near coincidence with sample K-Rb3. Although the scatter in the data for the more dilute alloys is large, it is clear that the curves now roughly match. This means that the turn down is approxi- mately proportional to 90. Such a dependence is quite 3 5/2 different from the no and no expected from either 250 180 A. I C ‘ 'b [3 C] I x u C — x D A. .A “ X X X .A is ' .A ' K-Rb t I 9.4 % p=l.O4p.Q-cm . 94 7° p =].01pQ-cm 1001— old {X 1.3 % - data 00.38% A0077% 0 so ‘ I l 0 0.5 T(K) 1.0 Fig. h-37, NormalizedrflPAdT vs T for K-Rb alloy samples. 181 localization or electron-electron interaction effects (see section 2.2.9). We tried to fit the data by p = Po + A'T2 + f(T) with various different functions fKT). The best fit, shown as solid curves in Fig. 4-38, contains f(T) = -CT. Such a term would be expected from localization (see section 2.2.9). This fit is better than one with f(T) = -DTl/2, which would be expected for electron interactions (see seciton 2.2dn, or f(T) = -BlnT, which would be expected for a Kondo effect (see section 2.2.7), but the scatter in the data is such that neither of these two alternatives can be completely ruled out. Fig. 4-38 shows the data of samples K-Rb3 and K- Rb4 in plots of normalized dP/TdT versus T'l, T‘3/2, and T‘z, with these three alternative fits in the forms as: dP/TdT = 2A - C/T, dP/TdT = 2A - D/2T3/2, and dP/TdT = 2A - B/Tz. The coefficients deduced from Fig. 4-38 are C a 2.2 i 0.4 pocmx‘l, D = 2.0 i 0.5 pnemx‘l/Z, a = 0.3 i 0.1 poem for sample K-Rb3, and C = 2.3 i 0.8 chmK’l, D = 2.3 1; 1.2 chmK'l/z, B a 0.5 1; 0.3 pflcm for sample K-Rb4. Comparing the magnitudes of our anomalies with predic- tions for localization or interaction effects, we found that the measured anomalies for 9.4 at% samples are about one order of magnitude larger than predicted for the interaction effects and two to three orders of magnitude larger than predicted for localization effects. The calculations are as follows: -2 .2 T (K ) Fig. 4-38. Three different trial fittings for (P(L+.2K)/T)(A|nP/AT) of two K-9.4at.%Rb alloy samples. 183 According to Kaveh and Mott (ref. 69), the conductivity contribution of electron-electron correlation effects in disordered metals is: e 1 "2h I"int(T) «mm = f(x) where Lintu‘) = (hD*/kT)1/2 with D* = (1/3)VF1e' and f(x) is typically of the order of 0.5 in high-resistivity metallic glasses. Using VF = 0.9 x 108 cm/s for K, and 1e z 2 x 10"5 cm calculated from Eqn. 4-1 for our samples with poz 1 Mom, we got D* s: 6 x 102 cmz/s, then G(T) = (0.2 {Idem-1 K'1/2)T1/2. Since the resistivity contribution Ap= —pga-(T) = -DT1/2, so D = pg (0.2 n‘lem'lx‘l/Z) a 0.2 pocmrc'l/2 for our samples with 90 = 1 pacm. This calculated value of D is one order of magnitude smaller than our measured value of about 2 pacm/Kl/z. According to Howson (ref. 56), the conductivity contribution of localization for high-resistivity metallic glasses has the form 1 h Li(T) m G(T) =37 with Lfim = (l/2)1eli 7" In our case 1e 2 x 10'5 cm as above and 1i = mVF/nezAilf’oT2 with A1 z 10'5 K'2 (see section 2.2.8), thus om z 0.005 (fzch)'l T. Since the resistivity contribution AP= — pg em = -CT, (2 = pg (0.005(szch)’1) z 0.005 pncmx'l for our samples with p0 = lyncm. This calculated value of C is 184 two to three orders smaller than our measured value of about 2 pach’l. Considering the higher than linear power dependence on 90 predicted for electron correlation or localization effects, the discrepancy will be even larger for the more dilute alloys. A Kondo effect could be proportional to no, but we did not see any marked anomalies (e.g., see section 4.4.2) in the G's of our samples, such as normally occur in Kondo systems. Also, experiments more recently done in our lab (ref. 68) showed that the turn-down of dp/dT at low temperature end for K-Rb dirty alloy was not affected by a magnetic field of 0.2T; thus it cannot be a Kondo effect. In 1983 Cochrane and Strom-Olsen (ref. 70) used recent scaling theories of the metal-insulator transition, (which take account of both localization and electron-electron interactions), to analyze the low temperature resistivity of their Y-Al metallic glass samples and several other representative alloy systems. They argued that at finite temperature the scaling theories lead to a conductivity: o'(T) = 0(0) (1 + Tl/z/A). They also claimed that they found the correlation gap A~p"2, leading to a contribution to the conductivity: Ao(T) 2 6T]'/2 (Qcm)'1. Since the resistivity contribution Ap= 43(2) AO'(T), thus AP z -69(23T1/2. In our case, po 2: l pflcm, this gives us a coefficient D z 6 chmK‘l/z, which is of the same order as 185 our measured value of D 2 2 pach'l/z. Again, however, the no dependence of our data is not as predicted. 4.5.2 W G measurements were made for samples K-Rbl, K-Rb2, K- Rb3, and K-Rb4 from about 4.2K down to about 0.1K. Fig. 4- 39 shows a plot of G versus T for these samples. For comparison, data of two samples from M. L. Haerle et al., (with 0.38 at.% Rb' and 1.3 at.% Rb), are plotted in the same figure. The data of each pair of our samples with the same concentration of Rb impurity almost overlap each other over the whole temperature range. The data of samples K-Rbl and K-Rb2 with 0.077 at.% Rb are well fitted by a solid curve of c = so + 3*1'2 + (C*/T) exp(-23/T) with Go>° in Fig. 4-39. The coefficients obtained from a computer fitting are: G0 = 0.353 1, 0.051 v‘l, A* = -o.205 i 0.014 v‘lx‘z, 3* = 2570 i 270 v‘lx. Some other results for Go and phonon drag terms can be seen in Fig. 4-39. For Go: (1) The value is positive for each sample. (2) The magnitude increases with increasing p0, and seems to have a limit of about 0.5 V‘l. For phonon drag terms (both normal and Umklapp): (1) They are quenched more and more as the impurity concentration increases, i.e., the magnitudes of these two terms decreases and thus the G data below 1 K are more positive and the data above 3K turn up more slowly with increasing temperature. (2) the G data 186 0.6— 7m. OV. (331‘?\\ “23 __ \\.-:. ‘6 PF 02— \. .2, a 9.4% :4 _ \c -' a 'v :7 5 v. 0'— ‘3 o ’_ I - . 1.3% —0'2_ ' I I -o.4-— . _ 00.38% 0 ‘Qér— —o.8.-— . _ \omm . ‘10 l l \l A l ' 1.0 2,0 3,0“ 40 T [K] Fig. h-39. G vs T for K-Rb samples. 187 of K-Rb3 and K-Rb4 with 9.4 at.% Rb can no longer be well fitted by G = Go + Al‘T2 + (B*/T) exp(-6*/T). 4.6 W W According to the standard theory of rflT) of high-purity metals (see section 2.1), when the temperature is low enough, electron-electron scattering should dominate pun, yielding p(T) at T2. Approximate T2 behavior of p(T) has been reported for K (ref. 14,18,19), Na (ref. 18), and Li (ref. 20) in the vicinity of 1K. This variation has been attributed to electron-electron scattering. If electron-electron scattering were the only source of iKT) at very low temperatures, then the T2 variation of pCT) should persist to the lowest temperatures studied. For our thick high-purity K samples, INT) varied closely as T2 only from 1.2K down to 0.3K; below 0.3K significant deviations from T2 behavior occurred (see section 4.1.1). In order to see whether there are, as in K, anomalous deviations in Na, Li, and Rb, from T2 behavior of p at low temperatures, we measured four freely hanging, bare, thick, high-purity samples of each of these metals from 4.2K down to about 0.1K. G measurements were made for each sample, concurrently with fix The details of the characteristics of these samples, are given in Tables 3-2, 3-3, and 3-4. 188 4.6.1 Ihfi_Bfiai££ixiL¥. Fig. 4-40 shows a plot of (Po/pT) (dP/dT) versus T from 3.6K to 0.1K for four thick (d - 1.00 mm) Na samples from two different sources (see Table 3-4). The two dotted lines indicate, for comparison, the T2 coefficients reported by Levy et al. (ref. 18). The lowest temperature they reached was 1.1K. Each set of their data followed a horizontal straight line from 1.1K to 2.1K, above which the data turned upward due to electron-phonon scattering. The open circles indicate data for our sample Na-H3 with RRR a R(295K)/R(OK) 4700. The data follow approximately a horizontal straight line from 1.21: to 1.9K with a nominal T2 coefficient (A z 0.175 p(zcm/Kz) slightly smaller than those of Levy et al. Above 1.9K the data turn up, as did those of Levy et al., as electron-phonon scattering becomes more important. Below 1.2K, the data also turn up, as in the case of K. A few solid circles indicate data of sample Na-H4, the twin of Na-HB. The open and filled triangles indicate data of samples Na-Hl and Na-HZ with RRR 2:400. These samples behave qualitatively like samples Na-H3 and Na-H4, but with larger apparent T2 coefficients (A.z 0.215 pncm/KZ), and smaller turn-ups at the lowest temperatures. The observed variations in the nominal T2 coefficients of the different Na samples shown in Fig. 4-40, as well as the lack of a large temperature range over which a simple T2 dependence is observed, mean that considerable uncertainty remains concerning the magnitude of the "intrinsic" 189 14&) 12.0—3‘ o N a 0.0— “‘3 (film/K”) _. .m o l l [e/IllAInp/Arl .0» ‘? P (D l o l o Fig. “-40. (PO/T)(AlnPAdT) vs T for free hanging, bare, thick, high-purity Na samples. 190 electron-electron scattering contribution to the resistivity in Na. Fig. 4-41 shows a plot of (94.2K/p) (dp/TdT) versus T from.2.8K down to 0.07K for several Li and Rb samples and, for comparison, for typical K samples of diameter 1.5 and 3 mm. For Li, two samples were initially measured (Li-Hla:A and Li-HZa:v) and then remeasured after nine days at room temperature (Li-Hlb:A and Li—H2b:v). For all four sets of measurements, the data are consistent with a horizontal straight line from 4.2K (the highest temperature measured) down to about.l.6K, below which the data rise anomalously above the extension of the horizontal line. Our data are the same to within experimental uncertainty for all four sets of measurements and are also in good agreement with the data of Sinvani et al. (ref. 27) in the region of overlap above 1.2K. The horizontal dashed line in Fig. 4-41 indi- cates the T2 coefficient (A 2:3.0 pncm/Kz) reported by Sinvani et al. The high Debye temperature of Li (9D(Li)z 330K) (ref. 69) ensures that electron-phonon scattering is negligible below'4.2K, the highest temperature measured in the present experiments. In contrast, for Rb, which has a very low Debye temperature (6D(Rb) z 65K) (ref. 70), electron-phonon scattering remains important down to about 0.5K, below which we see anomalous behavior. All three samples measured (Rb- H2:-, Rb-Vl:o,.and Rb-V2:o) show qualitatively similar 191 130—-._ , ° 110—,_ - ° : I 90— _ ° % '- ' A 70_ o . o Rb < '- . E o 8 g SO—v ‘3 0° " =5 5; ~170 r—z—I . E 8. , 5. ‘. L I —l30 51 as. _ '3 ‘~ * t3?) .v H “‘90 ' V. i " “ 21+“ 4 “‘Wrfi'?‘ 3‘4—1‘+“F+FF‘+14 9.0—é To l" 7.04:0. K . b'rofi 0. °.,°. 3; . ' 50 l 1 ° 1.0 2.0 T(K) Fig. h-hl. (P(h.2K)/T)(AlnPAdT) vs T for Rb, Li, and K samples. 192 behavior, with no temperature range over which a T2 variation is dominant. For K, included for comparison, the data are horizontal from about 1.2K down to about 0.3K, as described in section 4JLJ, below which they rise anomalously above the extension of the horizontal line. No measurements were made on sample K-4A (I) of C. W. Lee (with d = 3.0 mm) and our sample K-Hla (o) (with d a 1.5 mm) between 1.1K and 2.8K, but measure- ments on our sample K—H6a (x) and some other samples (see Fig. 4-1) have established that at about 1.2K the data begin to rise rapidly above the horizontal line with increasing temperature as electron-phonon scattering becomes important. The Debye temperature of K (6D 2: 114K) (ref. 71) is in between those for Na and Rb, which is consistent with electron-phonon scattering becoming important in K at about 1.2K. The standard theory of ART) provides no explanation for the very low temperature anomalies in Fig. 4-40 and Fig. 4~41. One possible explanation involves low energy excitations associated with residual defects, such as dislocations, which can scatter electrons inelastically (see section 2.2.2). As noted in section 1,2;1, in their studies of deformed , K, M. L. Haerle et a1. (ref. 22) reported that the low temperature resistivity of the deformed samples was found to be fit reasonably well by adding to a T2 variation a model of inelastic electron scattering due to dislocation 193 vibration (see section 2.2.2). In this model, a single- frequency local-phonon mode of frequency Pb is associated with each dislocation and yields a resistivity contribution of P8 (C/4T) sinh (hwo/ZkT) (section 2.2.2). Another mechanism of inelastic electron scattering can be due to additional bound states for electrons with energy slightly larger than the Fermi energy existing near the cores of dislocations. This model yields a resistivity contribution of P: o:(l +5exp(e/kT))'1 (section 2.2.2). These two models yield maxima in both (l/T) (dP/dT) and dP/dT, with the maximum occurring at higher temperatures in the latter case. A maximum in dP/dT was found in deformed K at about 0.2K (ref. 22). Fig. 4-42 shows a plot of (p4.2K/P) (dp/dT) vs T from 0.1K to 1.4K for the Na, Li, Rb, and for comparison for K samples shown in Figs. 4-40 and 4—41. The K data shown in Fig. 4-42 do not display a maximum even in dP/dT. It is not yet clear whether this anomaly in undeformed K is due to residual effects of incidental plastic deformation during cooling, in part because this anomaly does not extend over a sufficiently large tempera- ture range to allow discrimination between different models fits. Also, we do not yet have a plausible mechanism for producing enough dislocations to cause such an anomaly in the undeformed K. The undeformed Rb data shown in Fig. 4-42 shows a hint of local maxima in dP/dT near 0.2K, but both the uncertainty 194 120“ 80—— o Rb . I, if 40?— .;,°° ' ' 7:80 0— 'v ' ' //’I “—60 E .‘."‘ /// —40 E .‘ ' Li l- " ///I 7 —20 Q /’ a: /’ 5 K/ o d A, _ Q2. "" o , JL 6.0.. o/' I“ 0” ‘V’( 40— K V’ ”0/ 20- 39V VJ” == _ "56.0 M a ....... i a ‘ ‘ ........... o [’3'” “—4.0 .. . 0 I ‘A . o O .L. ...... 6. I”’ .*° .° --/-°',":”’ ——2.0 f .I..;.”’ w”l l l 1 1 1.1 o 0 0 2 04 0.6 0.8 l 0 1.2 T(K) Fig. h-AZ. P(4.2K)(AlnPAAT) vs T for Rb, Li, K, and Na samples. The data are the same as in Fig. h-QO and Fig.24-hl. dashed lines indicated T The resistivity variations inferred from all of the available data. 195 in the data and the large contribution of electron-phonon scattering (the very-low-temperature form of which is not known for Rb) preclude any meaningful analysis until meas- urements on higher purity samples are extended to still lower temperatures. The undeformed Na data shown in Fig. 4-42 show maxima which occur at very different temperatures for Na from our two different sources. A plausible generation mechanism for the residual defects needed to produce the observed anomalies in Na and Rb could be the martensitic transformation that Na undergoes at about 35K and Rb might undergo at about 4.2K (ref. 72). In undeformed Li, the anomaly sets in at a high enough temperature (21.6K) that a meaningful fit can be attempted, and the simple straightforward T2 form of INT) above about 1.6K makes the fitting procedure highly selective between different alternatives. In addition, as recently suggested (ref. 20), the martensitic transformation that Li undergoes at about 75K provides a plausible generation mechanism for the dislocations needed to produce such an anomaly. We tried to fit our Li data in both dP/dT and dP/TdT forms with several alternatives. Fig.«4—43 shows 4 differ- ent fits for the data in dP/dT form, while Fig. 4-44 shows these fits in dP/TdT form for our data. First we tried to fit our Li data with a T2 term plus a single-frequency local-phonon mode term, i.e., P= po+ AT2 + (C/4T) sinh'2(ht./2kT) (see section 2.2.2), thus, 196 2 ./ a 120- single local er/T _, . [phonon mode ”l/ ’ ,,/ 0” ( ) ’/" so" 30‘ A ./C / ,.// —.120 I y/ / ',./ v/V/ / /_w/ I n ‘f/ / ""‘/ I I, ‘ I / / /v./"/ 40". *g’ ' ' / __80 ' /’ . /’ th} / My:// _140 22 (:4 B)-——-—> \. ‘3 single bound ? E electron state . “- . / :5 T X j: E ._ ”A ‘ <1. 120' two local "3"? ,. Q”. //. / phonon modes r/{/,.,z ./. v / ’ 80.“ (c ) ,/'// / (.4/ #120 / / ,,(/ v" / /‘//"/ /'// / y/ 40‘“ "‘3‘" / / / "‘80 f. / / 2’ / ,/ vf‘”// /’ 04“ ’ ~40 i / ——-—-> 3‘ f/[tfilo bound electron states] / L5 1 l l ‘ l 1 1 1 1 I 1 1 1 0 o 1.0 2.0 T [K] Fig. 4-h3. samples. Trial fittings for P(L+.2K)Alnp/AT of Li 197 if: —l70 v [single local] . phonon mode /\ (1‘ :_ —'l30 §~ I é' . i7o—-.:. 5° ' single bound -:L A 3 [electron state] '2 130— a, (2) E "tk 9 ~ \ v 90—— ’.\ Q Q 50—-—' 5 1 q HP é. a ' twol al \N. [phonogcmodes] c; ‘ -130 1: is,29 (31 “'- .° \\“fi -90 ii " 170— ' - ——=—+::-E:Lw it- -'-+ — [two bound ] .50 130‘— electron states .. 90— a“ “fig: _ 1 i i .- 1 i 1 l v 50 1.0 T [Kl 2.0 Fig. h-hh. Trial fittings for (P(l+.2K)/T)(AlnP/AT) of Li samples. 198 dP/dT ZAT + (C/4T2)sinh'2(hw/2kT) x ((hw/kT) ctgh(haM2kT) - 1) (A) and, dP/TdT = 2A + (C/4T3)sinh-2(hw/2kT) x ((hw/kT) ctgh(hw/2k'1‘) - l) (1) This model described our data reasonably well from 2.8K down to about 0.2K, see the solid curves in part (A) of Fig. 4-43 and part (1) of Fig. 4-44, below 0.2K the theoretical curves dropped off more rapidly than the data. As an alternative, we also tried a fit with a T2 term plus a term associated with bound electron states at dislocations, i.e., p: p°+ AT2 + a(1 +5exp(E/kT))"l, thus, dP/dT = 2m: + (aflE/k)exp(E/k‘1‘) (1 +flexp(E/KT))"2 (B)* and, dP/TdT = 2A + (afiE/kT)exp(E/kT) (1 +fiexp(E/kT))'2 (21* with ,8(the spin degeneracy of the additional level) fixed as 1, we have dP/dT = 2AT + (aE/k)exp(E/kT) (1 + exp(E/kT))"2 (B) and dP/TdT = 2A + (aE/kT)exp(E/kT) (1 + exp(E/kT))"2 (2) 199 These fits shown as the solid curves in part (B) of Fig. 4- 43 and part (2) of Fig. 4-44, are even worse at the lowest temperatures. The problem with both models is that at temperatures well below their maxima in dRVdT, their single exponential decay for dP/dT is too rapid for our data. These two fits can be improved by using two local-phonon modes, or two bound electron levels, with adjustable energies and coefficients, but still fail at the lowest temperatures, as shown in parts (C) and (D) of Fig. 4-43 and parts (3) and (4) of Fig. 4-44. This improvement could mean either that we have a range of dislocation lengths in our samples, or just that we have introduced enough parameters to describe almost any smooth, peaked behavior in dP/dT. The coefficients deduced from each fit above are given in Tables 4-9 and 4-10. The differences between the residual resistivities (measured at about 1K in this study) and 4.2K resistivities are very different for different alkali metals. As we mentioned above, this difference is about 0.28 nncm for potassium. For Rb, this difference is as big as about 6.2 nncm, since Rb has a much lower Debye temperature than K has, so the resistivity contribution of electron-phonon scattering in Rb samples at temperatures lower than 4.2K is much larger than for K samples in the same temperature range; and 1K might not be a low enough temperature for measuring the residual resistivity of Rb. On the other hand, since the Debye temperature of Na is significantly 200 oH.o H mm.c m.m H o.m~ o~.o H m~.~ m.m H o.om ee.o m as.em lav nmo.o H omm.c m.~ H o.e~ mmo.c H can.“ o.H H a.om cH.o H mo.e~ lav uniiuiu niiuiau mo.c H mo.H m.m m e.mm HH.o H an.em Ame 1111111 1111111 oo.c H ca.“ o.~ + e.mm om.o + em.n~ Am. Age A euc mauofi x. Axe . soc mfiuofi xv A~x\2uc males as none x\mm a x\~m a a moamemm as ho noon on roam HosoH cououoao csson on» scum mocoaoauooou one .CHia oHcos ll .1. "'i.l‘ om.o m m~.o m.H m ~.m n~.o m mm.o ~.m m o.cH mH.o m om.m~ .mv No.0 + mm.o o.v + m.m mm.o + ~o.~ «.5 + v.o~ wm.o + m~.mm ADV iiiiiiiiiiiiii mcv.o l oo.vH Hm.m~ AHV 11111111111111 mo.o H mN.o Hm.o + hv.m ma.o H va.v~ Acv Axe Assoc mmnofl x. Axe Assoc males x. .~x\euc ma:o~ xv muse x~\nsc o x~\es~ o a moamecm «an NO mummy Q Ou mufiw 000:— COQOSQIHwUOH 05H EOHM mucwfinvfimwwoo 05H. omlw OHDmE 201 higher than that of K, the resistivity contribution of electron-phonon scattering at 4.2K for Na is expected to be smaller: our results (Table 3-4) showed this contribution is less than 1% of residual resistivity. For Li, we did not measure the 1K resistivity, but the Debye temperature of Li is even higher than that of Na, so FK4.2K) should be usable as the residual resistivity. 4.5.2 Ih£_Ih£LmQ£l£QLLis_BaLiQ_9 G measurements were made concurrently with p from 4.2K down to about 0.1K for all samples described in section 4.6.1. The G data for the four Na samples are shown in Fig. 4-45. The data for samples Na-Hl and Na-HZ are similar to each other in both form and magnitude, and can be well fitted by G a G0 + B*T2, (since the Debye temperature of Na is as high as 160K, it is plausible that the Umklapp phonon drag term might be negligible below 4.2K). The lower solid curve in Fig. 4-45 indicates such a fit to sample Na-HZ, and the deduced coefficients are co = -0.296 3; 0.003 (V4) and 3* a -0.0176 1 0.0004 (V'lK'z). The data of samples Na-H3 and Na-H4, from another source, are similar in form to each other, and for reasons not yet clear, show a small turn down at the lowest temperatures. The upper solid curve in Fig. 4-45 indicates a fit of G = G0 + B*T2 for sample Na-H3. The deduced coefficients are G0 = 0.063 i 0.021 (v-1) and B* = -0.065 i.0-0027 (V'lK’z). In contrast to K samples, the Na samples with smaller 1%3 show more positive Gos. 202 .mo_aEmm m2 xo_cu tom h m> u 06 .m:.: .m_u . O .0 3+qu a «2-02 4 «1-02 0 2.702 0 203 The G data for all four Li samples are quite similar to each other, not only in the form but also in magnitude, as shown in Fig. 4-46. Since Li has an even higher Debye temperature than Na, the G data of Li samples are also expected to be fitted by G = G0 + B*T2. The solid curve in Fig. 4-46 indicates that the data of sample Li-HZb can be very well fitted by such an equation, and the deduced + 0.0005 (v‘1) and 13* = coefficients are G0 = 3.7367 -0.00707 1 0.00006 (V’IK'Z). The slight turn-up in the c; data of samples Li-Hla, Li-H2a, and Li-Hlb at the very lowest temperatures is not yet understood. In the Rb samples, on the other hand, the G data show different forms from those for Na or Li. Since the Debye temperature of Rb is as low as 65K, the G contribution of Umklapp phonon drag will not be negligible even in the vicinity of 1K. The data are then expected to be fitted by G = G0 + B*T2 + C*/T exp(-6*/T). The G data for all four Rb samples are shown in Fig. 4-47. The solid curve in this figure indicates that the data of Rb-HZ are almost perfectly fitted by the equation given just above. The deduced coefficients from the fit are G0 = 1.0566 i 0.0025 (v‘1), 13* = -o.211 i 0.006 (v‘lx’Z), c* = 561 1 32 (V'lK), and 6* = 9.71 i 0.12 (K). 204 4.0- O O 3.9—- .A '2 3.8" CD A o Li-Hla 3.6— . li-H2a A li-Hlb A i. i-H2b l l 1 1 0 1.0 2.0 3.0 4.0 T (K) Fig. h-h6. G vs T for Li samples. 205 .mo.aEmm am Lou h m> u md .Aa-s .m_u o 9 (km 206 4.7 i.. .:t k =::. . \ . .-- - . _. - E In section 4.2 we described the size effects in f>and G for free hanging, bare, thin, high-purity K samples cooled in Ar or Be gas. In order to test whether similar size- effects can also be seen in pure Na, we made 9 and G measurements for six thin Na samples cooled in Ar and He gas. Samples Na-l/4Al and Na-1/4A2 (d = 0.25 mm) were cooled in Ar gas, samples Na-l/lOHl and Na-l/lOHZ (d a 0.1 mm) were cooled in He gas. The details of the characteristics of all these samples are shown in Table 3-4. 4.7.1 The_BeaistixiL¥ Fig. 4-48 shows a plot of normalized (dP/TdT) versus T for these six thin Na samples from 3.6K down to about 0.1K. For comparison, the data of one thick Na sample made of Na from the same source (Na-H3 with d = 1.0 mm) are also shown in Fig. 4-48 by (x) with a solid curve through them. The data of samples Na-l/4Al (a) and Na-l/4A2 (o) are quite close to the solid curve, no obvious size-effects are shown. 0n the other hand, data of samples Na-l/4Hl Cfil and Na-l/4HZ (*) are lower than the solid curve for T 2.1K, showing a hint of the same size-effects seen in K samples. But, for reasons which are not yet understood, the data of still thinner samples, Na-l/lOHl (A) and Na-l/lOHZ (A), are a little higher than those of Na-l/4Hl and Na-l/4HZ, contrary to what is expected for the size effects. To see the possible size-effects more clearly, in Fig. 4-49 we plot normalized dP/dT versus T from about 3.3K to 207 A o Na-H3 " O . Na-l/4AI 12.0-E O Na-I/4A2 L Tr Na-I/4Hl r—\ «x — 1 * Na-I/4H2 A \E g . Na-I/IOHI c100"" A A Na-l/IOH2 :: _ ..° 0 ,_..‘ 8.0— A '21 i — to a u. .E '. d 60". f 1: _ i 0 ° \ 11... 0 OS *2‘ 0 . h—J 4.0L_ :0 O O 4- o - it - M" " AA W4. . it * 2.0— l l L J l l l l l l L l l l 0 10 2.0 - 30 T (K) Fig. ’-+-’-18. (PO/T)(AlnP/AT) vs T for thin Na samples. 208 10.0 o 9.0” Na - t 8.0- B/ / 7.0- /. o/ A $2 \\. .' t E 600- 0’ .A ‘5: .4. / A 5.0 x0 . A IA' i“- o/ * <‘ 4 o ' * I P /A 2:? .93/ * 3 A00 1;: /* * A .A.A .5 / % 3'0-‘flifooom */ d(rnm) Gas Qingcm) * t o ** / o Na-H3 1.0 He 1.01 3 / 2.0— // o Na-1/4A1 0.25 Ar 1.18 / / a Na-IAHz 0.25 He 2.57 / 1.O_ // A Na-l/lOHZ 0.10 He 8.10 / / ()/ 1 l 1 l 1 l 1.0 2.0 3.0 T [K] Fig. 4-49. (PO/P)(AFVAT) vs T for selected thin Na samples. 209 about 0.1K for three selected samples from above. Again, for comparison, the data of Na-H3 are indicated by Open squares, and the coefficient of the nominal T2 term is indicated by the slope of the broken line. The data of all three thin samples are lower than the broken line for temperatures between about 1K and 3.3K. According to the theory of inelastic scattering of electrons by defects and impurities, the sample with larger 1% is supposed to have a larger coefficient of T2 term, but here we see contrary results. These results can be explained by electron-surface scattering effects, (i.e., size effects) (section 4.2.1). On the other hand, all six thin Na samples showed no nega- tive dP/dT. This means that any size effects in thin Na samples are not as strong as these in thin K samples. A partial reason could be that the electron mean free-path in bulk Na samples is smaller, about 0.14 mm calculated from lei = (1/10,‘,)(rs/a0)2 x 92 A (Eqn. (4-1); ref. 4). But this could not be the main reason, because this is only a factor of 0.65 smaller than 0.2 mm, the electron mean-free-path in bulk K samples. 4.7.2 Ih£_Ih2LmQ§l££LLiQ_BaLiQ_§ G measurements were made concurrently with P for all six thin Na samples described in section 4JLJ.from about 4.2K down to about 0.1K. Fig. 4-50 shows a plot of G versus T for these six samples. For comparison, the G data of sample Na-H3 (made 210 .mo_a5mm oz c_gu Lou h m> u .omiq .mmu c: h s 0.1V O.m Ohm O_.—. O a _ _ O 1 lifiYFH 2.62 o 1 11m”. I $331; 5 a 1 N, :33 1oz 6 o nab - rU «rebiaz s. 340.410 :3: 1oz a U2 .. «(9} lmz o 1 3:31.“: o 1 211 of Na from the same source), originally shown in Fig. 4-45 are shown again in Fig. 4-50 by means of the best-fit curve. As in K samples (section 4.2.2), in general, the thinner Na samples show less negative G at high temperatures and more negative G at low temperatures than do the thicker samples. This means that the thinner samples have more negative Go and less negative 8*. The possible explanations of this are given in section 4.2.2. 4.8 Na_EAM2LES_ENQASED_IN_£QLXEIEXLENE_IHBEE In section 4.4 we described likely Kondo effects in P and G for K samples encased in polyethylene tubes. In order to test whether Kondo effects can also be found in Na in contact with polyethylene, we measured tiand G for two Na samples encased in 1.6 mm dia polyethylene tubes under Ar gas (Na-PAl and Na-PAZ). To test for annealing effects, each sample was measured twice, allowing it to anneal for two different periods. The details of the characteristics of each sample in each measurement are given in Table 3-7. 4.8-1 W Fig. 4-51 shows a plot of normalized dP/TdT versus T for Na-PAla, Na-PAZa, Na-PAlb, and Na-PAZb from 3.6K down to about 0.1K. For comparison, data of sample Na-H3 are indi- cated by open circles in Fig. 4-51. All data of these four samples almost overlap each other and overlapsthe ones of Na-H3 for temperature lower than 2.2K, no Kondo-effect-like turn down (see section 4.4) have been seen in Fig. 4-51. In 212 14!) o — A No -PAlo 12.0- A No -PAia O Na _ v No -PA2b v Na-PAQb 10.0— C; 3‘ _ v 2 s C'. 8 ' :sol— V .1: __ ‘Q <1 a, ' " o % L :aiifli %, v :1 40 ° .— \ _ o E: 0' O X’ a o 40'. v m A X"§Xd%?é? g’ 2.0- o N0-H3 1 1 v L,l 1 1 1 1 l 1 1 1 1 I 1 1 0 1 3.0 T(1<)2'o Fig. li-Sl. (Po/T)(AlnP/AT) vs T for Na samples encased in polyethylene tubes. 212 14.0 0 F A No -PAlo 12.0“ A No -PAlc . Na __ V No -PA2b v No-RAZb 1CHD—' '1 3‘ _- v E 4 Ci 8 ' :1 8.0— 'I3 __ ‘Q E; an ' v o .Ei£i°_' %_ V 1: .a o \\ __ o S". o, o X5 is! o 4.0— v o9 A P kig Q? Ad 2.0— o No-H3 1 1 1 l 1 1 1 1 l 1 1 1 1 l 1 1 0 1.0 2.0 3.0 T (K) Fig. h-Sl. (Pb/T)(AlnP%dT) vs T for Na samples encased in polyethylene tubes. 213 order to see the data at the low temperature end more clear- ly, we plot dP/dT versus T in Fig. 4-52 for samples Na-PAla and Na-PAlb. For comparison, data of samples Na-Hl and Na- H3, originally shown in Fig. 4-42, are also shown n Fig. 4- 52. Data of sample Na-PAla in Fig. 4-52 show a clear turn- down at temperatures lower than 0.3K, giving a hint of a Kondo-like effect. But for reasons not yet clear, data of the same sample annealed for a longer period show, unex- pectedly, a low temperature turn-up. 4.8.2 Ih2_Ih£LmQ£1££LLiQ_BaLiQ_§ The plots of G versus T for samples Na-PAla, Na-PAZa, NaePAlb, and Na-PA2b are quite similar to the ones of free— hanging, bare, thick, pure Na samples and no negative anomalies at low temperatures were seen. These plots are not given in this dissertation. 214 .monsu dam—Euoroq 5 6032.0 3353 oz to“. ._. .m> h<\Qc:VoQ .Nmua .m: 1“ «4. as a; a; o; 4... e; as a... to 1.2.. 1 (is . q q _ 14 a _ 1 1 on... T u . an 9..." a; 6:22.38 £20 .0 oz 4 a a . .0 use . . ill. IHII-IIIIIIII Ionv 4 I I I . . .0 do I / . 13V .4 . U Q a I do low 1) I7 0 H I w 0 lo.@/ ION d I .. a Z I I d CHAPTER 5 Summary and Conclusions The general outline of this present thesis work is described in section 1.3. The conclusions of this study are as follows: 1) For free hanging, bare, thick, high—purity K samples cooled in Ar, He, or partial vacuum, with improved techniques we found that a) the temperature-dependent electrical resistivity p(T) varied closely as AT2 from 1.2K to 0.2-0.4K (for different samples), with A varying from 0.19 to 0.24 pacm/K2 for samples under different circumstances; b) below 0.2-0.4K, all the samples showed an anomalous turn-up in a plot of dP/TdT versus T; c) the thermoelectric ratio G of these samples could be well fitted by the theoretical formula G = Go + B"’T2 + (C*/T)Exp(-9*/T). These conclusions agree well with what C. W. Lee et al. found for thick pure K samples cooled in Ar gas. With an improved sample can and other techniques, we believe that our G data have higher accuracy than C. W. Lee's data. We have co = -0.075 1; 0.015 v’l, 3* = -0.484 3; 0.005 v'lx'z, and c* = 6200 i 140 v‘lx, with 9* fixed as 23K. 2) For free hanging, bare, thin, high-purity K samples cooled in He gas we found that the data of normalized dP/dT 215 216 displayed a clear pattern of unusual behavior which is consistent with that reported by Rowlands et al. in the temperature and diameter regions of overlap, but more com- plex in form. dP/dT even became negative in the vicinity of 1K for the thinnest samples whose diameters were comparable to or smaller than electron mean free path (~0.2 mm). The more complex behavior of our data rules out both the simple T3/2 form for p(T) that Rowlands et al. originally proposed, and all previously published explanations for the data of Rowlands et al. The only model we know, which might explain the negative dP/dT we observed, is an interaction between surface scattering and normal electron-electron scattering as first prOposed by Gurzhi and then calculated by a Monte Carlo method by Black. However, the data are not in a regime where a closed form expression for P has been de- rived. We therefore tried various trial fits to our data. The best fit was obtained as: (P(4.2K)/P) (dP/dT) =- 2A'T - 55T4/3/x/d7'I—Tn'n7, where A' ranged from 0.59 to 0.70 pSzcm/K2 for samples with different diameters. For reasons not yet clear, similar samples cooled in Ar gas or partial vacuum tended to behave like data for samples in He of larger diameter and to show greater variability for fixed sample diameter. We also found expected size effects in the resi— dual resistivity po. #5 = .QD+-c/d was obtained for samples with d20.l6 mm, where c a (3.4 i 0.5) x 1011 Qcmz was com- parable to the best literatue value (c = 2.9 x 10"ll Qcmz). For reasons not yet clear, the £%'s of our thinnest samples 217 (d$0.l mm) showed a lot of scatter, with most of the 90's much larger than expected. All G data for these samples could still be well fitted by c a Go + 3*T2 + (C*/T)Exp(-9*/T) with 9 fixed as 23K, but the thinner samples show larger magnitudes of negative Go' smaller magnitudes of negative 8*, and smaller magnitudes of C*. This behavior could be explained by the Gorter-Nordheim rule plus a surface scattering effect on phonon drag. Further investigations on size-effects, including making thin samples under vacuum and measuring thin film samples etc. are in progress in our lab. The only S data taken on a 0.5 mm dia K sample, indicate that L(T) :1: Lo, i.e., S = LOGT, is approximately valid only at T 5 1K for potassium. 3) For pure potassium samples prepared and cooled in contact with hydrocarbons, such as encased in polyethylene tubes or melted in clean paraffin oil, a Kondo-type effect was seen in both Pand G measurements. The data of P can be well fitted by p a po + AT2 - BlnT where the -BlnT term is the Kondo effect term. The data for G show a negative anomaly below 1K. The data below 1K can be well fitted by G = G0 + B*T2 + D*/T, where D*/T term is attributed to the Kondo effect. A stronger Kondo effect was seen: (a) in K samples encased in polyethylene tubes compared to one melted in paraffin oil, (b) in samples encased in thinner polyethye lene tubes compared to ones encased in thicker tubes, and (c) in samples encased in polyethylene tubes and held at 218 room temperature for a long period compared to those held for only a short period. K samples in contact with non- hydrocarbon plastics such as teflon or Kel-F showed no Kondo effect in either l’or G. An average magnetic field of 0.1T applied longitudinally'to polyethylene encapsulated samples shifted the resistance minimum to much lower temperatures. Thus, these samples seem to be exhibiting a Kondo effect due to magnetic impurities. The magnetic impurities seem to be a product of the reaction between potassium and hydro- carbons. Further investigation attempting to identify the magnetic impurities is in progress in our lab. In addition to increasing the size of the Kondo-type effect, sample annealing at room temperature also brought both p5 and A down, as also observed by van Kempen et a1. Po changed by a factor of about 3, consistent with the results of van Kempen et al. However, the change in A was only about 30%, much less than the factor of 3.6 reported by van Kempen et al. 4) For K-Rb alloy samples, a plot of Po vs c shows that the value of Po per atomic percent Rb impurity is in close agreement with the best literature value of 0.13 imam/at% for our dilute alloy samples. Plots of normalized dP/TdT vs T showed that (a) each set of data follow a horizontal straight line roughly from about 0.5K to about 1.2K, yielding p= p0 + A'TZ. Here A' can be well fitted by A' = A0 + Aipo, with A0 found to be 0.24 i 0.04 p$2cm/K2 for an electron-electron scattering contribution, and A1 found to be 11 x 10"6 K"2 close to the theoretical values of 219 13.7 x 10’6 K"2 from P. L. Taylor and 12.5 x 10"6 K"2 from Kus and D. W. Taylor for inelastic electron-impurity scat- tering contribution; (b) a low temperature turn-down became progressively more apparent as the Rb concentration in- creased. The reasons for this turn-down are not yet clear, but our data can be fairly well fitted by adding to P 8 Po + A'T2 an additional term having any one of the alternative forms - CT, -DT1/2, or -BlnT. The -CT term would be ex- pected from the localization effect; the -DT1/2 term from the electron-electron interaction effects; the -BlnT term from the Kondo effect. However, our turn-down is approxi- ‘mately prOportional to p0, not proportional to pg or PCS/2 as expected for localization or electron-electron interac- tion effects. And, the G measurements and further meas- urements under a magnetic field of 0.2T did not show any behavior such as normally occurs in Kondo systems. More- over, for the 9.4 at.% samples, the calculated value of C from localization effects is two to three orders of magni- tude smaller than the deduced value of C from our data, and the calculated value of D from electron-electron interaction effects is one order of magnitude smaller than the deduced value from our data. Considering the higher than linear power dependence on Po predicted for electron-electron interaction or localization effects, the discrepancy will be even larger for the more dilute alloys. We also found that the G data of dilute K-Rb alloy samples could still be well fitted by G = G0 + B*T2 + 220 (C*/T)Exp(-9*/T), where Go is positive and increases with increasing po to a limit of about 0.5 V"1 , and the magnitude of both 8* and C* decrease with increasing pb, (iJL, the two phonon drag terms are quenched more and more as the impurity concentration increases). These variations in Go' 8*, and C* can be explained by the Gorter-Nordheim rule plus an impurity scattering effect on phonon drag. Measurements on still more concentrated K—Rb alloys and a search for a similar anomaly in K-CS alloy samples are planned. 5) For free-hanging, bare, thick, high-purity Na, Li, and Rb samples we made l7and G measurements from 4.2K down to 0.07K. A T2 dependence was found for Li from 1.6K to at least 4.2K (the highest temperature in the measurements) and, perhaps, for Na from 1.2K to 1.9K. No T2 dependence of p was found for Rb. We found anomalous turn-ups at low temperatures in a plot of dP/TdT vs T, as seen in K samples, for all Li, Na, and Rb samples. The turn-up is largest in Li and smallest in Na. Above 0.2K, the data of p for Li can be fairly well fitted by adding to a T2 variation a model of inelastic electron scattering due to local phonon mode of dislocation vibration or due to bound electron states associated with dislocations. Below 0.2K the theoretical curves dropped off more rapidly than the data. The fits can be improved by using two local phonon modes, or two bound electron levels, with adjustable energies and coefficients, but still failed at the lowest temperatures. 221 We also found the G data of Na, Li, and Rb to be well fitted by the theoretical form G = G0 + 3*T2 + (c*/T) x Bxp(-6*/T). For Na and Li, which have relatively high Debye temperatures, the third term due to Umklapp phonon drag was found to be negligible. Because different metals have dif- ferent Debye temperatures, and thus electron-phonon scatter- ing would die off at different temperatures, the differences between 1K4.2K) and p5 (measured at about 1K) were found to be very different for Rb, K, Na, and Li. We found this difference to be about 6.2 chm for Rb; about 0.28 nacm for K, (consistent with what van Kempen et al. found): and less than 0.01 nflcm for Na. For Na and Li, 4.2K is low enough for measuring the residual resistivity p0. On the other hand, for Rb, 1K might not be quite low enough for measuring Po accurately. 6) For thin Na samples we found possible size effects in both P and G, which, if they really exist, are much less strong than the ones in K samples. No negative dPVdT was found in thin Na samples. For Na samples encased in polyethylene tubes, no obvious Kondo effect was found in p or in G. 10. 11. 12. 13. 14. 15. LI ST 0F REFERENCES A. W. Overhauser, Adv. in Phys., 21, 343 (1978). J. M. Ziman, Electrons and Phonons, Oxford University Press, London, 1960. M. Kaveh, C. R. Leavens, and N. Wiser, J. Phys. F: Metal Phys., 2,, 71 (1979). Ashcroft/Mermin, Solid State Physics, Holt, Rinehart and Winston Press, 1976. D. K. C. MacDonald, W. 8. Pearson, and I. M. Templeton, Proc. R. Soc. A, 2_4_8_, 107 (1958). D. K. C. MacDonald, W. 8. Pearson, and I. M. Templeton, Proc. R. Soc. A, 255,, 334 (1960). A. M. Guenault and D. K. C. MacDonald, Proc. R. Soc. C. W. Lee, Ph.D. Thesis, Michigan State University, 1980. D. K. C. MacDonald, G. K. White, and S. 8. Woods, Proc. R. Soc. A, 215,, 358 (1956). G. G. Natale and I. Rudnick, Phys. Rev. 1.61, 687 (1968) . J. C. Garland and R. Bowers, Phys. Kondens Materie, 2,, 36 (1969) . J. W. Skin and B. W. Maxfield, Phys. Rev. 8.4, 4215 (1971). D. Gugan, Proc. R. Soc. (London) A, 125, 223 (1971). H. van Kempen, J. S. Lass, J. H. J. M. Ribot, and P. Wyder, Phys. Rev. Letters, 31, 1574 (1976). H. van Kempen, J. H. J. M. Ribot, and P. Wyder, J. Phys. F LL. M. Kaveh and N. Wiser, J. Phys. F: Metal Physics, .‘LQ, L37 (1980). 222 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 223 J. A. Rowlands, C. Duvvury and S. 8. Woods, Phys. Rev. Letters, 40, 1201 (1978). M. F. Bishop and A. W. Overhauser, Phys. Rev. Letters, 12, 1776 (1979). B. Levy, M. Sinvani, and A. J. Greenfield, Phys. Rev. Letters, 13,, 1822 (1979). C. W. Lee, M. L. Haerle, V. Heinen, J. Bass, W. P. Pratt, Jr., J. A. Rowlands, and P. A. Schroeder, Phys. Rev. B, 25, 1411 (1982). A. H. MacDonald and D. J. W. Geldart, J. Phys. F l_0_, 677 (1980); A. H. MacDonald, R. Taylor and D. J. W. Geldart, Phys. Rev. 823., 2718 (1981). W. E. Lawrence and J. W. Wilkins, Phys. Rev. B, 1, 2317 (1973). M. L. Haerle, Ph.D. thesis, Michigan State University, 1983: M. L. Haerle, W. P. Pratt Jr., and P. A. Schroeder, J. Phys. F 1.3. (1983) L243. C. W. Lee, W. P. Pratt, Jr., J. A. Rowlands, and P. A. Schroeder, Phys. Rev., 5.5,, 1708 (1980). P. L. Taylor, Proc. Phys. Soc., .89., 755 (1962); Proc. R. Soc. A215,, 209 (1963); Phys. Rev. 5132, 1333 (1964). 3. w. Kus and 0. w. Taylor, J. Phys. 310, 1495 (19301. G. Krill, Solid State Communications, Vol. 9,, 1065, 1971. M. Sinvani, A. J. Greenfield, M. Danino, M. kaveh and N. Wiser, J. Phys. F11 (1983) L73. F. Bloch, z. Phys., :12, 208 (1930). R. A. Brown, Can. J. of Phys., _6_0_, 766 (1982). V. F. Grantmakher and G. I. Kulesko, Sov. Phys. JETP, 4_0_, 1158 (1975). P. Fulda and I. Peschel, Adv. in Phys., 21, 1 (1972). S. G. O'Hara and A. C. Anderson, Phys. Rev. 8, .L0_, 574 (1974); Luigi, 8, 2, 3730 (1974). M. L. Bonack and A. W. Overhauser; Phys. Rev. B, 12 6454 (1978). 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 224 M. F. BishOp and A. W. Overhauser, Phys. Rev. B, 2_3_, 3638 (1981); Ibid., 13., 2447 (1978). M. Knudsen, Ann. Phys. (Leipzig) 28, 75 (1909). Saul Dushman, Scientific Foundation of Vacuum Technique (Wiley, New York, 1962), 2nd ed. R. W. Whitworth, Proc. Roy. Soc. London, Ser., A245,, 390 (1958). J. E. Black, Phys. Rev., B21, 3279 (1980). Carl A. Kukkonen and Henrik Smith, Phys. Rev., 8.8,, 4601 (1973). N. H. March and S. B. Woods, Solid State Communica- tions, 42, 993 (1984). R. B. Dingle, Proc. R. Soc., 291, 545 (1950). J. R. Sambles, K. C. Elson and T. W. Preist, J. Phys. F12, 1169 (1981). J. R. Sambles, Private Communication. R. I. Boughton and J. E. Neighbor, J. Low Temp. Phys. 1, 241 (1972). R. N. Gurzhi, Zh. Eksp. Teor. Fiz., 44, 771 (1963) [Sov. Phys. JETP 11, S41 (1963)]; Sov. Phys. JETP, 29_, 953 (1965). J. Kondo, Prog. Theoret. Phys., 32, 37 (1964); Solid State Physics, vol. 23, F. Seitz and D. Tumbull, eds., Academic Press, New York, 1969, p. 183. C. M. Hurd, Electrons in Metals, John Wiley & Sons Press, New York, (1975). S. Koshino, Prog. Theor. Phys., 24, 484 (1960); 1pm., 24, 1049 (1960); m” 3.0., 415 (1963). K. Frob6se and J. Jackle, Proceedings of BPS Study Conference, C-14, 1977. A. G. Aronov and B. L. Altshuler, 1978 Zh. Eksp. Teor. Phys. Pis. Red. 21 700; B. L. Atshuler, 1978 Zh. Eksp. Teor. Phys. 15,, 1331. V. N. Fleurov, P. S. Kondratenko and A. N. Kozlov, J. Phys. F 10, 1953 (1980). D. J. Thouless, Phys. Rep. 135., 93 (1974). 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 225 F. J. Wegner, z. Phys. 25, 327 (1976). H. G. Schuster, Z. Phys. 31, 99 (1978). E. Abrahams, P. W. Anderson, D. C. Licciardel lo, and T. V. Ramakrishnan, Phys. Rev. Lett. 42J 673 (1979). M. A. Howson, J. Phys. F 14 (1984), L25. J. M. Ziman, Phil. Mag., 4, 371 (1959). M. Bailyn, Phil. Mag., 5, 1059 (1960). W. P. Pratt, Jr., Can. J. Phys., 60, 703 (1982). D. Edmunds, W. P. Pratt, Jr., and J. A. Rowlands, Rev. Sci. Inst., 51, 1516 (1980). J. L. Imes, Ph.D. Thesis, Michigan State University, 1974. G. L. Neiheisel, IULD. Thesis, Michigan State University, 1975. National Bureau of Standard, Washington, D.C. Professor P. A. Schroeder, Michigan State University. J. Bass, Adv. Phys. 21, 431 (1972). S. B. Woods, private communication. J. Bass, In Landolt-Bornstein Tables, New Series III/15a. W. P. Pratt, Jr., et al. (unpublished). M. Kaveh and N. F. Mott, J. Phys. C: Solid State Commun.31,527,1981. R. W. Cochrane and J. O. Storm-Olsen, Phys. Rev. B, 22 1088,1984. D. K. C. MacDonald, Handb. Phys., 14, 137 (1956). I. M. Templeton, J. Phys., £12, L121 (1982).