'1 ,1 K s ’ ‘4 ’ '3. Ill‘I‘d-JL‘FSIJ‘M, » NEE." 1k”; .1" WY," 3:": 1‘71“,“ {NW an... 3 .1“. . .Wt'w 6%.}...1' *. 'l. 91 ”v.1 :.‘ '..‘ "WI." 1‘ N .fllnhufuhu. f;:? “'-3 lnqlgfrh. $IW "h ‘.i .1 REM . .-'y§v t '“':i'L M“ ‘45 1‘ WW". €2.er ’19:" MN Ii. . .1. f.. -r ‘: ' “”J m ~ ' .23 a . ")1, I I V H . 'r‘w‘fl ' ”MINN a L L_" _::z E ‘%" . I ' I H 3:. 1911.!" L" ‘I ' 3.13%? I‘I‘I‘ . .33.: IF?“ II.‘ LKI‘JFA‘LIHW‘ “ -‘ n i '>' y‘: ‘ iatzfi I H b..'~ . U ".5. I... P - v 3...... L‘ ’1 NW... .C- 5.. ,1. ‘33: 3s1.§3§|33.i{3.3 , :3“ 1., 5“ x..:~""h<1u.‘ .341. . '3’. i wr ML ‘ZHJ "3 m1 3‘“ ”mum; r‘ "‘5, “3"? ' 3".1 I :I _0 fi“ thII’I{}1,Ii' _.MEL-1',’:"L13:lj.1§l_."£§;s‘ .5“:" “’11 ' kw....13..“H'f"' {4. .H‘LJIII IRIIII’JI. :H‘I . IQI>¥3 .z u; ‘1’; fi.., ‘il," U§3191311 ... :_,.k3 W. 'm ...3; v“\ A LIL.JUET’ igE‘uIHai'l. ii?“ . .. _, . AP ~40 , .-‘ .2 . :z'u’ . '— -"-“’.T; m {4... N ..o|~— .’A:' ".. v V n .I .‘. ‘.°.\-,l;_-, I“). I T . n v 09 a 7 f ' 0" \-.A o I 4.‘ 5 N». o..— C to! N.“ I. ‘ r*' 3‘". . >0 Ba H'III “MW“ : I' .1. .13!” .‘ \ 3 Li). . "lib . 1.11:3”. 3‘“. .x'. . 3...... a“. 1 . .. ". .. '.33:, "4H « I“ -L”'.‘ ‘Llo ' l‘wm'lLMIL‘l 33LVLI333331‘.&/‘r-‘i a. I-' “‘ .¢.~:...=..;y . .._‘;. ..L.. . - 3 3H. [JIM- . . c I ‘ |. I". _ . I: \ . - l " - LEI". H M‘INI‘ ‘Ih'HwHH-‘L ‘I‘III'II'MIIII‘ I ’.\.v—- ”-35.64“ n F “.~ .. . w .... I"'. '.~.. -._., .4 —-. - y“ N. ._ ‘ - - ‘v _-.a- IIZBIZY“) IIIII III III IIII III III III III |III IIII III III III III III This is to certify that the thesis entitled THE MAGNETOTHERMOELECTRIC EFFECT IN SINGLE CRYSTAL ALUMINJM AND INDIUM AT LIQUID HELIUM TEMPERATURES presented by Barry Jay Thaler has been accepted towards fulfillment of the requirements for Ph.D . degree in _ED¥SJ$L5__' / Major professor Date 7 ,- i ‘1 "-7- 7 0-7639 LIBRARY Michigan State University THE MAGNETOTHERMOELECTRIC EFFECT IN SINGLE CRYSTAL ALUMINUM AND INDIUM AT LIQUID HELIUM TEMPERATURES BY Barry Jay Thaler A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics 1977 ABSTRACT THE MAGNETOTHERMOELECTRIC EFFECT _ IN SINGLE CRYSTAL ALUMINUM AND INDIUM AT LIQUID HELIUM TEMPERATURES BY Barry Jay Thaler We have studied the effect of a transverse mag- netic field H upon the low temperature absolute thermo- power S of very dilute aluminum and indium based alloys. Measurements were carried out on single crystal samples at temperatures T between 1.5 and SK and with magnetic field strengths up to 20kG. The aluminum measurements were performed on a sample grown along the [001] axis and containing 50 ppm gallium The indium measurements were performed on several samples containing from O to 500 ppm added impurities, which were grown along several crystallographic axes. S of both metals consisted of both a component which was oscillatory with increasing magnetic field and a non-oscillatory component. The non-oscillatory component of the aluminum sample was obtained for H along both the {010} and {110} families of axes. The data were consistent with the equation smean that these oscillations are caused by magnetic breakdown, the only other mechanism expected to cause this oscillatory behavior. For aluminum, the above interpretation is sup- ported by the fact that independent magnetoresistance measurements strongly suggest that magnetic breakdown is present for H directed along the [010] axis. In addition, s°SC was largest for H along this axis, being an order of magnitude larger than the non-oscillatory component at 20kG. As H was rotated away from the [010] axis the oscillations decreased in magnitude, but per- sisted throughout the (001) plane. The oscillations had periods corresponding to those of the B-orbit. The field and temperature variations of the [010] oscilla- tions were dominated by a term exponential in the ratio m*(T + TD)/H. Analysis using the de Haas-van Alphen pro- cedure for determining effective masses and Dingle temperatures yields m* = (0.093 : 0.01)me and T 2 l-4K. D * This value of m agrees with the value obtained from Barry Jay Thaler de Haas-van Alphen measurements for the same orbit, and corresponds to an electron-phonon mass enhancement of approximately 40%. Large amplitude, low frequency oscillations in S were observed in indium below 2.5K for H between 10 and 20kG and directed along either the [010], [I01], or [lIO] axes. The periods of these oscillations correspond with those obtained from de Haas-van Alphen measurements on the neck orbit of the B-arms of indium. Arguing by analogy, comparison of the [010] indium data with similar data for the [010] breakdown orbit in aluminum leads to the conclusion that these oscillations also arise from magnetic breakdown. This is the first evidence that there may be magnetic breakdown in indium. TO MY PARENTS ii ACKNOWLEDGEMENTS It is a great pleasure to acknowledge my thesis advisor Professor Jack Bass, whose suggestions and criticisms throughout the course of this research were invaluable. I would also like to thank Dr. Jon L. Opsal and Professor Robin Fletcher for numerous helpful dis- cussions. The time donated by Dr. E.L. "Terry" Stone III and C.W. Lee in helping to make measurements is also gratefully acknowledged. I am indebted to z. Katsiapis for growing the Al(Ga) single crystal. Specific thanks go to Mr. B. Schumaker for preparing the alloys, and to D.A. Cady for helping with the many aspects of sample preparation. I would like to thank all the guys of the machine shop for all their design tips and help in constructing the various pieces of apparatus. I would also like to thank all the secretaries who have been associated with the Physics Department during my stay, especially the two SSMS secretaries Jean Strachan and Delores Sullivan for all their help. And last, but certainly not least, I'd like to thank Diane Spero for helping to type this thesis. I would also like to acknowledge the financial support of the National Science Foundation. iii List List II. III. IV. TABLE OF CONTENTS of Tables of Figures INTRODUCTION A. The Experimental Problem B. Previous Work on the Magnetothermopower of Aluminum C. The Present Thesis EXPERIMENTAL TECHNIQUE A. "1310001 Voltage Measurement 1. Nulling Technique 2. Field Sweep Technique Thermometry Sample Preparation Alloys Crystal Growth Experimental Procedure and Data Analysis 1. Nulling Technique Measurements of S, K, and R 2. Field Sweep Technique Measurements of S, K, and R NON-OSCILLATORY THERMOELECTRIC PHENOMENA - THEORY A. B. Electron Diffusion 1. General Relations 2. Electron-Phonon Mass Enhancement in S. Phonon Drag NON-OSCILLATORY THERMOELECTRIC PHENOMENA - EXPERIMENT A. Aluminum 1. Introduction 2. Sample 3. Data and Analysis iv Page vi vii 33 33 33 40 48 56 56 56 59 59 Determination of the Non- Oscillatory Results AA for H Parallel to [010] and [001] AA for H Parallel to [011] Effects of Magnetic Breakdown Phonon Drag 4. Summary and Conclusions for Aluminum Indium 1. Introduction 2. Data and Analysis 3. Umkehreffect 4. Summary and Conclusions for Indium u C N H uww U) U100.) V. OSCILLATORY THERMOELECTRIC PHENOMENA - THEORY A. B. C. Variation of the Density of States at SF with H Estimate of S Oscillations due to Density of States Oscillations for Aluminum and Indium Magnetic Breakdown VI. OSCILLATORY THERMOELECTRIC PHENOMENA - EXPERIMENT A. B. The Fermi Surfaces of Aluminum and Indium Aluminum . 1. Introduction 2. Experimental Procedure 3. Data and Analysis 3.1 H Parallel to [010] 3.2 H Along other Directions in the (100) plane Indium - 1. Introduction 2. Experimental Procedure 3. Data and Analysis List of References Appendix I : Appendix II: Calculation of Anisotropy of r(k) for Al(Ga) 30(8) Determination of as Page 59 62 64 65 66 69 70 70 74 9O 92 92 95 102 112 112 114 114 116 116 116 125 132 132 133 133 143 148 150 Table II-1 IV-1 VI-l AI-l LIST OF TABLES Some characteristics of the samples used in this thesis Values of B(H) for Al(Ga) * Values of m /me and TD for two orthogonal directions of the magnetic field H and for different assumed values of n for Al(Ga) Average reciprocal relaxation times for Al(Ga) obtained from Sorbello's approximate phase shift mode1(7) vi Page 27 67 124 149 Figure 11.1 11.2 II.3 II.4 III.1 III.2 III.3 LIST OF FIGURES The Superconducting Chopper-Amplifier System: a) The superconducting chopper- amplifier circuit; b) Schematic (actual size) of the device (after Averback(3)). The variation with magnetic field of the gain of an unshielded chopper-amplifier system The variation with magnetic field of the gain of a shielded chopper-amplifier system The Vacuum Chamber: A) lucite block, B) heat sink for resistor leads, C) sample, D) Glastic backing, E) car- bon resistor, F) heater, G) cold sink (O.F.H.C. copper), H) binding post, I) flange for Pb O-ring, J) vacuum line, K) epoxy seal, L) support rod, M) vise clamp, N) sample support (brass) which is attached to the cold sink e vs. k relations for free electrons (dashed curve) and for quasi-particles (solid curve) The change of electron wavevector accompanying the annihilation of a phonon by a (a) Normal process, (b) Umklapp process (after Barnard(23) ) Schematic diagrams of sections through the 2nd (upper) and 3rd (lower) zones of Al showing electron-phonon scattering events (after Gripshover, et. al. (2 )) vii Page 13 15 18 46 51 54 Figure IV.1 IV.2 IV.3 IV.4 IV.5 IV.6 IV.7 IV.8 IV.9 The variation of S with magnetic field at 3.55K for the Al(Ga) sample. The magnetic field is directed along the [010] axis. The maxima and minima of the quantum oscillations are repre- sented by circles, and the non- oscillatory component of S is repre- sented by the crosses The temperature dependence of the non- oscillatory component of S of the Al(Ga) sample for: H = lSkG and parallel to [010] (squares); H = lSkG and parallel to [100] (triangles); H = lSkG and parallel to [110] (crosses), H = OkG (circles). The full symbols indicate data obtained with H reversed. The temperature dependence of the non- oscillatory component of S for sample In(Ga) with H directed along [010] The temperature dependence of the non- oscillatory component of S for sample In(Ga) with H directed along [001] The temperature dependence of the non- oscillatory component of S for sample In(Sn)-II with H directed along [001] The temperature dependence of the non- oscillatory component of S for sample In(Sn)-II with H directed along [110] The temperature dependence of the non- oscillatory component of S for sample In(Sn)-I with H rotated 12° from the [101] in the (101) plane The temperature dependence of the non- oscillatory component of S for sample In(Sn)-III. The magnetic field dependence of S for sample In(Sn)-II with H directed along [110] viii Page 61 63 75 76 77 78 79 80 81 Figure Page IV.10 The temperature dependence of the Lorenz ratio L divided by the ideal Lorenz ratio L0 for samples: A) In(Sn)-III, with H directed along [001]; B) In(Sn)-II, with H directed along [001]; C) In(Ga), with H directed along [010] 84 IV.ll Variation of S with rotation angle 6 at H = lOkG and T = 4.7K for sample In-I. H was rotated about the [101] axis. 6 = 213° corresponds to H directed along [010] 87 IV.12 Variation of R (upper curve) and 8 (lower curve) with rotation angle 6 at H = lOkG for sample In(Sn)-I. H was rotated about the [101] axis. 6 = 202° corresponds to H directed along [010]. The R curve was measured at 4.2K and the 5 curve at 4.3K 88 V.l Variation of the Fermi surface with field, showing the increasingly spherical shape as the field becomes weaker. The field decreases going (a) to (d) (after Adams and Holstein(43)) 93 v.2 (a) Free electron orbit in a magnetic field. (b) In the periodic potential of the lattice, the orbits are recon- nected at the zone boundary; but in a strong magnetic field may jump back to the free electron path (after Ziman(51)) 104 v.3 Ashcroft's 4-OPW pseudopotential model for the electron-like 3rd zone of the Fermi surface of aluminum (after Abele and BlatthZI) 106 v.4 Central [100] section through the (63) aluminum Fermi surface (after Ashcroft )107 V1.1 (a) The Brillouin zone for face-centered tetragonal indium; (b) the indium free- electron model hole surface; (c) the indium third zone electron B-arm sur- face (after Hughes and Shepherd(52)) 113 ix Figure VI.2 V1.3 V1.4 V1.5 VI.6 The variation of S with magnetic field at T = 3.55K for the Al(Ga) sample. H is along the [010] axis. The solid line labeled S = 0 represents the x-y recorder output when there is no ' applied temperature gradient. The broken line indicates the non- oscillatory component of S. The oscillating curve represents raw data, uncorrected for a field dependent amplification of the measuring system. The magnitude 9f this correction is in- dicated by the manner in which the S = 0 curve and the broken line approach each other as H increases. When the data are corrected, these two curves are equi- distant at all fields above about llkG to within experimental uncertainty The variation of the quantity 3/2 osc . . H S /T (plotted on a logarithmic scale) with H”1 along [010] The variation of the quantity SOSC/T (plotted on a logarithmic scale) with H- for a series of temperatures for sample Al(Ga). H was directed along [010] The variation of the quantity H-B/ZSOSC/T (plotted on a logarithmic scale) with H"1 along [010] The variation with temperature of the slopes of the lines shown in Figure V1.4 for H parallel to [010] (circles) and of similar slopes of similar lines for H for a series of temper- atures for sample Al(Ga). H was directed for a series of tempera- tures for sample Al(Ga). H was directed Page 117 120 121 122 parallel to [100] (triangles). The upright triangles represent initial measurements with H parallel to [100]. To evaluate the effects of changes in sample treatment and alignment, after these initial measure- ments were completed the sample was re- moved from the holder, its potential leads X Figure V1.7 V1.8 V1.9 V1.10 V1.11 Page were removed, it was reannealed, the potential leads were reattached and the sample was returned to the holder. The resulting data are indicated by inverted triangles 123 The variation of S with magnetic field at T = 2.4K for the Al(Ga) sample. H is rotated 20° from the [010] axis in the (001) plane. The solid line labeled S = 0 represents the x-y recorder output when there is no applied temperature gradient. The oscillatory curve repre- sents raw data, uncorrected for the field dependent amplification of the measuring system 125 The variation of S with magnetic field at T = 2.8K for the Al(Ga) sample. H is along the [110] axis. The solid line labeled S = 0 represents the x-y recorder output when there is no applied tempera- ture gradient. The oscillating curve represents raw data, uncorrected for the field dependent amplification of the measuring system 127 The variation of S with magnetic field at T = 2.3K for H along a series of directions in the (001) plane of the Al(Ga) sample. 0° corresponds to H directed along [100] 129 Variation of the oscillation period of the third zone B-orbit with rotation angle in the (001) plane. This figure displays our present magnetothermopower data, the de Haas-van Alphen data of Larson and Gordon,(32) and the results of a pseudo- potential calculation by Ashcroft(53) 130 The magnetic field dependence of S for H aligned along the indicated crystallo- graphic axes. The lines labeled S = 0 represent the recorder output when there is no applied temperature gradient. The oscillating curves represent raw data, uncorrected for the field dependent xi Figure V1.12 V1.13 amplification of the measuring system. The amplification decreases by about 30% over the field range shown. The curves for H aligned along [101] and [010] are from sample In(Sn)-I and the curve for H along [110] is from sample In(Sn)-II The variation of the quantity HSOSC/T (plotted on a logarithmic scale) with l/H for sample In(Sn)-I at T = 1.67K. H was directed along [010] Variation of S with rotation angle measured from the [010] axis, at H = lSkG for aluminum and indium single crystals. The aluminum data are from sample Al(Ga). The indium data for H rotated about the [001] axis are from sample In-II and the data for H rotated about the [lOl] axis are from sample In(Sn)-I xii Page 134 136 140 I. INTRODUCTION A. The Experimental Problem Thermoelectric effects are usually classified in one of three categories: the Seebeck, Peltier; or Thomson effects. However, all three effects are related by the (1) Kelvin-Onsager Relations, and therefore provide equi- valent information. The experiments forming the basis of this thesis are measurements of the Seebeck effect. This effect is characterized by the generation of an electric field E by means of a temperature gradient VT which is applied along an electrically isolated sample. Using (2) dyadic notation, the thermopower tensor 4S is defined as ‘E = E (’v’T)'1 . (1.1) . + + . Since E = -VV, we may write ‘3 = -VV(VT)-1 . (1.2) This is equivalent to H _ -4 g! as S - 1 dT 3 (1.3) A where i and j are unit vectors in the directions of the voltage and temperature differences respectively. 1 Since measurements are performed over paths of finite length, the experimentally measured thermopower, ‘Séxp, is defined as where AV and AT are measured voltage and temperature differences. If AT 5 dT then S 5 Sexp; this may be attained when AT << T. Since the present measurements were performed at temperatures between 1.5 and SK, temper— ature differences of approximately 0.1 to 0.3K were used. A typical value of S for aluminum or indium is lO-BV/K, 8 9 which implies that 10- - 10- V must be measured. To make measurements with 1% accuracy required a voltage 10 - lO-llv. This is not a trivial sensitivity of 10- task in the presence of a large magnetic field (20kG.). However this problem can be solved by using a supercon- ducting chopper-amplifier. B. Previous Work on the Magnetothermopower of Aluminum The thermopower is the most difficult transport property of a metal to properly calculate. A major reason for this is that thermopower is the transport property which is most sensitive to the details of the scattering processes. However, Averback found a property of the electron-diffusion component Sd of the thermopower of aluminum which was insensitive to the scattering details and depended only on the host meta1.(3’4) Using a number of polycrystalline samples of very dilute aluminum alloys, he showed that although both the zero field and high field values of S (note that S saturates in the high field limit for aluminum or any other uncompensated metal with no open orbits) varied from sample to sample, the difference between these two values, ASd = Sd(H + w).- Sd(H = 0), was approximately constant -- for all samples ASd/T varied between 2.1 - 2.6 x 10-8V/K2. Averback and Wagner(s) calculated ASd using no adjustable parameters and obtained the value ASd/T = 1.6 x 10-8V/K2, about 30% smaller than experiment. Some possible sources of this 30% discrepancy are: (l) The measurements were performed on poly- crystalline samples, while the calculation is appropriate for H directed along a four-fold symmetric axis of a single crystal. (2) The impurities in the samples might have been anisotropic scatters in aluminum, while the calculation assumed isotropic scattering. (3) The 1-OPW Fermi surface used by Averback and Wagner in their calculation might be in- apprOpriate for comparison with experiment. (4) The effects of mass enhancement were not in- cluded in Averback and Wagner's calculation. (Averback and Wagner neglected effects of electron-phonon mass enhancement because Prange and Kadahoff‘G) had claimed that such effects were not present in the electronic -properties of metals.) C. The Present Thesis This thesis began as an attempt to measure the transverse magnetothermopower of an aluminum sample which satisfied the requirements of the calculation of ASd made by Averback and Wagner. We hoped that such measurements would isolate the source of the discrepancy between theory and experiment. A single crystal foil, oriented so that the mag- netic field could be directed along either a four-fold symmetric [010] crystallographic axis or a two-fold symmetric [110] axis, was used. This crystal contained SOppm Ga, an impurity which should be a nearly isotropic scatterer in aluminum.(7) The data obtained for ASd with this single crystal are compared below with the data from polycrystalline samples reported by Averback, (4) (5) et. al.; with the calculation of Averback and Wagner; with an improved, 4-OPW version of this same calculation (8) by Opsal and Wagner which was stimulated by the experi- mental results to be described; and, finally, with the Opsal and Wagner calculation including the effects of mass (8) enhancement. The results obtained with this aluminum single crystal stimulated us to extend magnetothermopower measure- ments to indium single crystals to see whether ASd be- haved the same as in aluminum. It was anticipated that such measurements would be interesting because although aluminum and indium are both Group III metals, they crystallize with different structures; aluminum in a face centered cubic (FCC) structure, and indium in a face centered tetragonal (FCT) structure. The slight tetragonal distortion of indium causes some axes which are four-fold symmetric in the FCC structure to be only two-fold symmetric in the FCT structure. Thus indium would allow a study of the effects of this small distortion on ASd. Unfortunately, this study yielded results which did not allow the electron- diffusion component of the magnetothermopower to be un- ambiguously separated from the phonon drag component. The indium data will be presented and will be analyzed to de- termine the source of this lack of separation. Some indium samples exhibited a significant magnetothermopower umkehreffect; this will also be discussed. In the process of making the aforementioned studies, giant quantum oscillations were discovered in both metals for H along a number of crystallographic directions. For H along the [010] axis in aluminum these oscillations could easily be distinguished from the noise for T between 1.7 and 4.8 K; therefore the temperature dependence of the oscillation amplitudes was systematically studied. In indium the oscillations could only be distinguished from the noise for T between 1.5 and 2.5 K; therefore an oscillation amplitude study similar to that performed for the [010] direction in aluminum was not attempted. The experimental results will be compared with theories which predict oscillatory behavior. Also, the periods of the magnetothermopower oscillations will be compared with the same periods obtained from independent magnetoresistance or de Haas-van Alphen measurements. The remainder of the thesis is organized as follows: Section 11 provides a description of the experi- mental apparatus and measurement techniques. Section 111 furnishes the theoretical background 'for the calculation of magnetothermopower. Particular emphasis is placed on the role of electron-phonon mass en— hancement in thermoelectricity. In section IV the experi- mental results for the non-oscillatory component of the magnetothermopower for both aluminum and indium, are pre- sented. Section V presents the theory of the oscillatory component of the magnetothermopower. These oscillations can be caused by either oscillations in the density of states at the Fermi level, or by magnetic breakdown. Both of these pictures are developed. A simplistic calculation is used to estimate the magnetothermopower oscillation amplitude expected from oscillations in the density of states. In section VI the experimental data for the oscillatory component of the magnetothermopower of both aluminum and indium are presented and compared with the simple calculation. II. EXPERIMENTAL TECHNIQUE A. VOltage Measurement The voltage measuring system is shown in Figure 11.1. This chopper-amplifier system was designed and de- SCIiDEdeldetail by R.S. Averback.(3) In view of this, only the techniques of voltage measurement will be de- scribed hithis thesis. The system was operated in two modes: (1) At steady magnetic fields the signal gen- erated by the sample was nulled out, using the chopper- amplifier as null detector, (2) In continuously varying magnetic fields the signal generated by the sample was directly measured using the chopper-amplifier as a linear amplifier. Operation in the first mode will be described in the next section entitled "Nulling Technique". Opera- tion in the second mode will be described in the section entitled "Field-Sweep Technique". 1. Nulling Technique As can be seen from Figure 11.1, the chopper- amplifier system contains a reference resistor which is in the liquid helium bath. This reference resistor was a small segment of copper wire having a resistance of about BuQ, which remained constant to within our ability to measure it (< 2%) at all temperatures and fields employed in these measurements. 8 \D SEFIHHS‘IIIEVE Low "0'88 I parecrop , AMPLIFIER I FREQUENCY A CURRENT» DOUBILER SUPPLY AUDIO ‘ Pl-OTENT OSCILLATOR OMETER IF:- fir R 9 o z « - E‘ RESISTOR i—dwwww—A 1 W | CHOPPING ELEMENT I (a) l,4 Current Lead: V 5 6 2.3 Potential Leads I: 5 Oscillator Lead 0 6 7 Ground R 8.9 Secondary Leode I0,ll Sample Leads 3 4—"'1 R Reference Reeietor 7__ 1 fl T.T Transformers - c Chopping Element T 1- ‘l V Vector Board 9 8 l0-—’ "¢—-ll (b) Figure 11.1: The Superconducting Chopper-Amplifier System: a) The superconducting chopper-amplifier circuit; (b; Sche- matic (actual size) of the device (after Averback( ). 10 When a known current (produced by a series of 12V dry cell batteries) is passed through the reference resistor, a known voltage is generated in the chopper- amplifier circuit. This voltage is used to null out the voltage which is being generated by the sample. At null condition there will be no net current flowing in the circuit, so the panel meter (output) of the phase sensitive detector will show no deflection. The accuracy of the technique is limited by the circuit noise, which appears as a small oscillation of the needle of the panel meter. Using a time constant of 3 seconds on the phase sensitive detector (PAR model JB-4), this noise was usually 2 - 4 x io'llv in zero field and increased to 10 1 - 3 x 10' v at 20kG. 2. Field Sweep Technigge At all magnetic fields used in these experiments the output of the chopper-amplifier system was quite linear (to within 2%). So after independently calibrating the panel meter, the voltage generated by the sample was measured by reading the deflection of the needle of the panel meter. The panel meter was calibrated by passing a known current through the reference resistor under con- ditions where the sample was not generating a voltage, and then measuring the resulting deflection of the needle. In addition to monitoring the output of the phase sensitive detector on the panel meter, the analog output 11 was fed directly into the y-axis input of an x-y recorder (HP model 7004B). Sweeping the magnetic field at a con- stant rate and feeding the analog ouput of the Hall probe gaussmeter (F.W. Bell model 660) into the x-axis, resulted in a continuous recording of the voltage generated by the sample as a function of H. Since a changing magnetic field induces a voltage within our measuring system, the magnetic field had to be swept up twice - once with the sample gen- erating a voltage, and once with it generating zero voltage (to define a zero voltage baseline). The difference be- tween these two lines gives the voltage generated by the sample. Data taken using this field sweep method agreed with data taken independently using the nulling technique. The field sweep technique is advantageous for recording data when the voltage generated by the sample varies rapidly with field (e.g. giant quantum oscillations). There are two reasons why it is important not to sweep the field too rapidly: (1) If the field is swept at a rate such that the thermally induced voltages are changing faster than the time constants allow the Circuit to respond, then the voltage variations will be smoothed out (i.e. reduction of amplitude of the quantum oscillations). (2) If the field is swept too fast, the voltage induced by the changing magnetic field may 12 cause a component of the voltage measuring system to saturate, thus giving "false" data. The field sweep technique has some disadvantages compared to the nulling technique: 1) 2) The field sweep technique offers less accuracy because the quantity of interest must be ob- tained by taking the difference between two lines on a recording, each of which contains noise. Additionally there may be slow varia- tions in both the signal and the zero baseline which lead to systematic errors. Comparisons between equivalent data suggest that the field sweep technique is about a factor of three less accurate than the nulling technique. A second problem with the field sweep technique is that the gain of the chopper—amplifier system is field dependent (at 20kG the gain can be as much as 50% less than at lOkG). Figure 11.2 shows a typical field dependence of the amplification of the system. The raw data must be corrected for this field depen- dent gain. Toward the end of the present thesis research, the effect of shielding the chopper-amplifier with 5 mil thick lead foil was investigated. When the shielding ex- tended from above the chopper-amplifier to about 8cm below 13 (N I- t: Z I“ :3 a3 .. (r 3 <1 I Z I-- A: .oou uuommsm AA .Hmmm hxomm AM .mcwa Essom> Ab .mcwuio no How omcmam AH .umom mcfiocfin Am .Aummmoo .O.m.m.ov xcwm oaoo A0 .Houmos Am .Houmflmmn conumo Am .mcwxoon oeummaw An .oHnEmm kw .momoH Houmwmou How xCAm now: am .xooHn muwosa Am "Honfimno Essom> one ”v.HH musmflh .I 19 of the sample to a cold sink, which is in direct contact with the helium bath, the sample is thermally isolated. The end of the sample attached to the cold sink will be called the "cold" end and the other end will be called the "hot" end. When the heater at the "hot" end of the sample (the "hot" heater) is turned on, a temperature gradient is produced along the sample; this also causes the tem- perature of the sample at the position of the "cold" resistor to be higher than the bath temperature. After measuring RC and AR the "hot" heater is turned off and the "cold" heater is turned on. The "cold" heater heats the sample to a uniform temperature. The heating power of the "cold" heater is adjusted so that Rc has the same value as when the sample was heated only by the "hot" heater. Rc and AR are measured. The inverse of eq (11.1) is used to calculate the temperature at either end of the sample, and therefore T is determined. If the resistance of the two resistors drifts with time this drift should show up as a false temperature difference when only the "cold" heater is on. By substracting this false tempera- ture difference from the temperature difference obtained using only the "hot” heater, the correct AT is determined. The false temperature difference was usually less than 5% for temperature differences greater than 0.03K. If the sample is not completely isolated, the "cold" heater will produce a real temperature difference, 20 which should be indicated by a thermally induced voltage. In our measurements, such voltages where typically about 1% of those produced by the "hot" heater. It was there- fore concluded that the real temperature difference pro- duced by the "cold" heater was about 1% of that produced by the "hot" heater. C. Sample Preparation Figure 11.4 shows how the measuring probes and other apparatus were mounted on a sample. In order to have enough surface area to accommodate all these elements and also to maintain a sufficiently large temperature difference, the samples were about 7cm long, and 2 to 5mm wide. The following procedure was utilized to quickly mount the elements on the sample while introducing a minimum of cold work. First the potential and current leads were attached. These leads were 3 mil copper-clad NbTi wire, chosen because NbTi remains superconducting at all tempera- tures and magnetic fields employed in these experiments. For aluminum samples, the leads were spotwelded on. In preparation for spotwelding, the sample was glued to a template using Duco Cement. The insulation and copper- cladding were removed from both ends of each NbTi wire. The leads were then varnished, using G.E.-703l varnish, to the sample, so that one end of each lead was in posi- tion for attachment. The leads ran from these points 21 to the center of the sample from where each pair of leads was twisted tightly together as it led away from the sample.. The potential and current leads were then spot- welded to the aluminum sample. The sample was removed front the template by soaking in an acetone bath, and then glued to a Glastic backing, again using Duco Cement. Glastic was chosen as a backing material because it is an in- sulator whose thermal expansion is similar to that of a metal. Since indium is a much softer material than aluminum, the potential and current leads were either "acid-welded" or soldered to the indium samples; this avoided the damage to the sample inherent in spotwelding. Indium has the property that when the surface is well cleaned with hydrochloric acid, it will bond to another indium surface which has been similarly prepared. This permitted NbTi leads which had been previously "wetted" with indium to be "acid-welded" to the sample. Alter- natively the leads could easily be soldered to the in- dium samples using Rose's metal and Superior #30 liquid soldering flux. Tests showed that the data were the same for both methods of lead attachment. For both of these methods the copper-cladding was not removed from the lead ends which were to be attached to the sample. This was done because it was much easier to "wet" the copper- cladding with either indium or solder than it was to "wet" 22 NbTi. Since tests showed that the thermOpower of indium with either method of lead attachment was zero (to within experimental uncertainty), when the indium was in the superconducting state, it is concluded that the copper- cladding did not introduce any measurable spurious vol- tages. Since the potential and current leads were not spotwelded to the indium samples, these samples could be glued to the Glastic backing before the leads were attached to the sample. Once the leads were attached and the sample mounted on the Glastic, the sample was placed in a brass sample holder which held the sample parallel (to within 1°) to the magnet pole pieces. The sample was isolated from the brass holder by 2 small lucite blocks, one at either end of the sample. These blocks were varnished between the Glastic and the brass holder using G.E.-7031 varnish. In the case of aluminum, one end of the sample was then mechanically clamped to the cold sink. 1n the case of indium, a small piece of very pure indium with about the same cross-sectional area as the sample was clamped to the cold sink, and the end extending from the clamp was melted and attached to one end of the sample. This was done to avoid the strains which would be intro- duced if the indium single crystal were directly clamped to the cold sink. 23 The other ends of the potential and current leads were then spotwelded to permanent leads which ran out of the sample chamber through an epoxy seal (which was con— structed in accordance with the procedure outlined by Anderson‘ll) ). Since the insulation and copper cladding had already been removed from the last centimeter or so of these ends, only the superconductor, NbTi, remained. Superconductors are poor thermal conductors, and therefore these leads shouldn't provide a significant heat leak. This was verified by the fact that the thermal emf gen- erated when the sample was heated only with the "cold" heater, was typically 1% or less of that generated by the "hot” heater. Next the carbon resistors and heaters were mounted. The insulation on the carbon resistors was ground off on one side. Cigarette paper saturated with G.E.-7031 varnish provided electrical insulation between the resistor and sample, and the varnish held the resistor in place. The leads to the resistors were 3 mil. manganin wire. To pre- vent large heat leaks through these leads, the manganin wires were attached toarsmall heat sink at a temperature within 0.01 to 0.02K of each carbon resistor before being brought out to connections external to the sample. These small heat sinks were made by varnishing several turns of the manganin wire to copper posts which were either var- nished to the aluminum samples or soldered (using Cerrolow 24 #117 solder) to the indium samples, about 0.5cm from the carbon resistors. The heaters were constructed by wrapping about 5000 of Evanohm wire (~130/cm) around a copper post and varnishing the wire to the post. The copper post was then attached to the sample in the same manner as the heat sinks. The heaters were operated using a d.c. current of l to lOmA. The several leads from the heaters and resistors were soldered to permanent leads on a vector board which was mounted on the brass sample holder. The permanent resistor leads were 3 mil. managanin wire and the per- manent heater leads were 38 awg. copper wire. All these leads entered the sample chamber via the vacuum line and were heat sunk by varnishing them to a copper post which was tightly screwed into the main frame of the cryostat (which was in contact with the helium bath). Finally, the brass vacuum can was attached. The seal between the can and the sample chamber was made by squashing a Pb + 0.05%As O-ring between the can and the sample chamber with 12, 2-56 x 3/8" stainless steel Allen head screws. The O-ring was made by wrapping Pb + 0.05%As wire, which had already been greased with Apiezon M grease, around the flange (see Figure 11.4) and twisting it tight. Not a single O-ring prepared in this manner leaked. A few times tempered steel (black) screws were 25 substituted for the stainless steel screws. In almost every instance these seals leaked at 4.2K. This entire sample preparation procedure could be completed in 1-2 days. This procedure was conceived to introduce a minimal amount of cold work to the rather delicate samples. D. Alloys Both the aluminum and indium alloys were mixed by Mr. B. Shumaker, using the procedure described by R.S. Averback.(3) The only difference in procedure was that the indium master alloys were prepared with a con- centration of about 1% solute material. The "pure" aluminum was 69 grade aluminum supplied by Cominco Inc. The "pure" indium was either 69 grade indium supplied by Cominco Inc. or 59 grade indium supplied by the Indium Corporation of America. The solute materials, gallium and tin, were 69 grade. The aluminum alloys were rolled into 20 mil. thick foils, while the indium alloys were rolled into rods with a 2mm square crosssection. E. Crystal Growth The Al(Ga) single crystal was grown by 2. (12'13) Before Katsiapis using a strain anneal technique. mounting, the single crystal was annealed in air for 1 hour at 400C. 26 The indium crystals were grown in an atmosphere consisting of 90% argon and 10% hydrogen using a simple zone refiner and seeds of known crystallographic orienta- tion. First the indium alloy and seed were placed in a crucible made from Spectroscopically pure carbon. A heat lamp housed in an elliptical reflector was used to pro- duce a molten zone about 1.5 to 2.5cm. long. The molten zone was first produced at the gnseeded end of the indium alloy. This zone was then slowly moved along the entire length of the alloy until a small portion of the seed melted. The zone was then moved slowly away from the seed back along the entire length of the alloy. This two-pass procedure was adOpted with the intention of achieving a homogeneous impurity distribution, since an odd number of passes is expected to produce an impurity concentration gradient along the sample. Once grown, the indium single crystals were not further annealed. During initial trials the indium sometimes separated into two pieces within the molten zone. This problem was eliminated by completing the etching (using concentrated HCl) and rinsing (using double distilled H20 followed by methyl alcohol) procedure between 5 and 10 minutes before the indium was placed in the crucible. The crucible was then immediately placed in a tube which was evacuated so that the argon-hydrogen atmosphere could be introduced. 27 All samples were characterized by x-ray (Laue) pattern analysis. Table 11—1 shows the intended impurity concentration, the measured resistance ratio, RRR (R (300K)/R(4.2K)), and the crystallographic axis along which the crystal was grown, for all crystals used in this thesis. TABLE 11-1: Some characteristics of the samples used in this thesis Nominal Impurity Orienta- Misori- SAMPLE Conc. RRR t1on entae t1on Al(Ga) 50 ppm 1,400 [001] < 1° In-I 0 15,400 [101] 2° In-II 0 4,350 [001] 3° In(Ga) 50 ppm 7,400 [100] - In(Sn)-1 50 ppm 2,300 [101] 3° In(Sn)-II 100 ppm 1,650 [110] 2° In(Sn)-III 500 ppm 390 [110] 2° The rather high RRR of the In(Ga) sample indicates that considerably less than 50 ppm gallium was in solution. Single crystals were considered satisfactorily oriented when the crystallographic axis along which the sample was grown lay within a few degrees of the sample axis. The misorientation was determined from X-ray pattern analysis. The misorientation for each sample is also given in Table 11-1. Since the sample holder aligned 28 the sample axis (not the crystallographic axis) trans- verse to the magnetic field, the magnetic field could not be aligned exactly along the crystallographic axes. How- ever, in the interest of simplicity, the magnetic field will be referred to as being aligned along a given crystallographic axis when it was aligned as closely as possible to that axis. F. Experimental Procedure and Data Analysis 1. Nulling Technique Measurements of S, K, and R Thermopower measurements using the chopper- amplifier as a null detector were performed in the follow— ing manner. The measurements were always begun with the helium bath at 4.2K. The "hot" heater was turned on, producing a temperature gradient along the sample, and the power was recorded. Within a few seconds the sample reached a steady state condition, and the thermally gen- erated emf was then nulled by passing current through the standard resistor. The current necessary for nulling was recorded. RC and AR were also recorded. The "hot" heater was then turned off and the "cold" heater turned on. The power input to this heater was adjusted so that Rc had approximately the same resistance as it did when only the "hot" heater was on; this power was then re- corded. RC and AR were again recorded. Sometimes a relatively small thermal emf was generated by the "cold" heater, this is probably indicative of a small heat leak. 29 When this occurred, this voltage was nulled and the current necessary for the nulling was recorded. When analyzing the data, this latter voltage was subtracted from the thermal emf generated by the "hot" heater. This series of measurements was repeated for a few different power inputs to the "hot" heater and at several magnetic fields. The electrical resistance of the sample could be measured by sending an electrical current through the sample and nulling the voltage which was generated (both heaters were off when making this measurement). The electrical current and the nulling current were both re- corded. To correct for "thermals" the electrical current was then reversed, the voltage was again nulled and both currents were again recorded. The bath was pumped down in 0.2K intervals and the carbon resistors calibrated. Complete sets of the types of measurements just described were normally made at bath temperatures of 4.2, 3.0, 2.1, and 1.4K. From the above measurements the thermally gen- erated voltage V the temperature difference across the sample AT (including the corrections described in the thermometry section), and the average temperature of the sample could be determined. With this information we could compute ILR V V H,T 30 where 2 is the distance between the 2 carbon resistors R and 2V is the distance between the two voltage probes. 2R The correction factor I— was needed because the tempera- V ture difference was measured over a larger distance (10% - 30%) than the voltage difference. This correction factor assumes that the temperature gradient is uniform between the carbon resistors. As a by-product of the 8 measure- ments, the thermal conductance K is also obtained: (11.3) ”I so where Q is the power input to the "hot" heater. Having measured the resistance of the sample R, the thermal conductance K, and the average temperature T, at which the thermal conductance was measured, the Lorenz. number L can be computed from the relation 2 L = 1%; (11.4) V 1R The correction factor E—-is included for the same reason V as when computing S. At both H = 0, and in the high field limit L should be equal to the ideal Lorentz number, 2.443x 10-3 watt-Q/K2 in the limit where elastic impurity scattering is dominant. 2. Field Sweep Technique Measurements of S, K, and R Measurements utilizing the field sweep method were performed using the following procedure. The analog 31 output of the monitor of the lock-in amplifier in the chopper-amplifier circuit was connected to the y-axis of an x-y recorder, and the analog output of the gauss- meter was connected to the x-axis. The field was then swept up at a rate of 7G/sec. After the field had in- creased approximatley lkG, the "hot" heater was turned on for the remainder of the sweep. Upon completion of the sweep, the "hot" heater was turned off and the field re- duced to its initial value. The field was swept up again, this time with the heater off. If initially the output on the x-y recorder had shifted from the output of the previous sweep when the heater was off, the y—axis off- set was adjusted to align the two recordings. Following completion of the second sweep, two sets of calibration measurements were made with station- ary magnetic fields. First, the output signal was measured at several magnetic fields for a constant input current to the reference resistor. A smooth curve was drawn through these data points to determine the amplifications factor at all fields of interest (see Figures 11.2 or 11.3); this determined the correction of the raw data for the field dependent gain of the measuring system. Second, Rc and AR were measured in the usual manner, with the "hot" heater adjusted to the same power as during the first sweep. Since the temperature dif- ference increased only about 10% over the field range 32 between 10 and 20 k6, these resistances were measured only every 1 or 2kG. These two sets of calibration mea- surements were made after each thermopower recording. Eq. (11.2) was again used to calculate S. At each magnetic field, V was determined by measuring the difference between the two lines on the x-y recording and then multiplying this difference by the appropriate amplification factor. AT was found by first converting the Rc and R values to temperatures and temperature dif- ferences appropriate to the calibration fields and then linearly interpolating to obtain AT at intermediate fields. III. NON-OSCILLATORY THERMOELECTRIC PHENOMENA - THEORY A. Electron Diffusion Electron diffusion Sd refers to the thermopower that would be measured if the phonon spectrum remained in local thermal equilibrium. In section III.A.l a gen- eral expression for ASd of any system of non-interacting fermions (e.g. free electrons or quasi-particles) will be derived. In section III.A.2 both Averback and Wagner's(5) (8) and Opsal and Wagner's results for the evaluation of this general expression will be presented. Both of these calculations neglected the effect of the electron-phonon mass enhancement. Also in section III.A.2 we shall follow 1(14) to show how this enhance- the work of Bass and Opsa ment should be included in the evalution of the general expression for ASd. 1. General Relations The following derivation of Sd is primarily based (5) on the work of Averback and Wagner. In this section it is advantageous to define the transport coefficients (tensors) according to the following scheme: ‘3 ‘E - (-'v"r) (111.1) + + J= ~E+ 33 34 -+ 4+ -+ H -) where 3 and 6 are the electrical and heat currents/unit 1 area respectively. IS is defined as E(VT)- 3:0, so that ‘3 and‘z are related by the equation 4+ ++-]_ H S = 0 ° 8 . (III.3) We assume a system of non-interacting fermions (electrons or quasi-particles) whose motion in the presence of electric and magnetic fields is governed by the rule: _) n§E=EE+§3xfi (111.4) + I where v is g1ven by n i; = i7}; a]; (111.5) with ”h Plank's constant divided by 2n e = electronic charge c = speed of light ER = energy of an electron in state k _ 3 3 3 I - ( 3k ' 3k ’ 8k ) x y 2 It can then be readily shown within the framework of the linearized Boltzmann equation for E and 5 along the x axis, and H along the z axis that: e2 3+ 0 = ——§ f d k v f' x (III.6a) xx 4n + _ B 3+ r xx - ——3-4 I d k v fE,T (III-6b) n ek T _ B 3+ _ + "XX '— IF I d k (5E 8F) Vx fl-E'E (1:11.76) - kgT f d3? ( - ) I f' (111 7b) Kxx - 3 5k EF vx K,T ' ° Here kB is Boltzmann's constant, e is the Fermi energy, F and [éE f'+ ([-kB$T f'E T]) is the non-equilibrium por- I I k E] tion of the distribution function for an electron in state k when only an electric field (temperature gradient) is present. The above integrals are over k space, but we can just as easily integrate over surfaces of constant energy and then integrate over all energies. Performing this change of variable gives:15 fd3k = f d f dzs (111 8) E 457' - where S is a surface of constant energy. For convenience deS will be written as de, with the understanding that de is a 2-dimensional surface integral. It is now con- venient to define the function WE by (-3f°§) f]; = TWK (111.9) 1 where f3 = [l.+ exp(ei - eF)/kBT] .’ k In the ensuing analysis the subscript R will be dropped, except where essential for clarity. Combining equations (111.6a, 8, 9) we see that af° OXX = I de (" 5?) Oxx(€) (III.10) where e2 dS Oxx(€) = 1;? I‘D; waE (111.11) Combining equations (111.7a) and (111.11) we see that k T o - B _ 3f _ "xx - e I de ( as ) (e 8F)oxx(e) (111.12) k T _ 2 2 d B 3 - e LO T [3; OXX(€)]€=€F + 0(E;—) nzk: where L = -——7 (the ideal Lorenz number). From the 0 3e (16) «+ MI-fi) Onsager relations we know that 5(H) = E-T———q so we may write 4+ _ (_i_4++ _ 8(fi) - eLOT [de 0 ( fin€=€ (111.13) F where?+ is the transpose of the tensor‘E. Using the Onsager relation oij(H) = oji(-H) it follows from equation (III.3) that Sd is given by: + _ 4++ _c_'i_++ Sd(H,T) — eLoT p(H) do o(fi) (111.14) where43(H) is the resistivity tensor,‘3(H) =‘3-1(H). It is worth noting that eq. (111.14) holds for arbitrary scattering mechanisms. In the experimental arrangement described in section 11, Se is measured under adiabatic boundary XP conditions (i.e., 6 = (Q,0,0)), with H along the z-axis. 37 Sexp can then be related to the components of‘S by the relation 8 = s + s —1— . (111.15) provided that H is along a 2-fold or higher symmetry axis. Neglecting thermoelectric contributions we see from eq. (111.2) that the quantity VyT/VXT is related to the components of the thermal conductivity tensor,4z, by the relation V T K V11 = - —X§ . (111.16) K x yy Combining equations (111.15) and (111.16) we obtain K s = s - s —¥§ . (111.17) exp xx xy Kyy Evaluating the first term on the right hand side of eq. (111.17) in the high field limit gives: _ d Sxle + a0) - eLoT [pxx(H + ) de Oxx(H + )I€=€F + p (n + m) 9— ln 0 (H + m)| XY _ d6 YX €=eF 1 1 d = eL T [0( ) - m '- O (H + 00)I _ o H: oxy(H + ) d8 yx E-EF = eL T 9— 1n 0 (H + 00)I . (III 18) 0 de xy €=€F ' For an uncompensated metal in the high field limit 38 [ne(e) - nh(e)] ec _ 1.. oxy(e) - H + 0(H3) (111.19) where ne(e) and nh(e) are the number of electrons and holes contained within the surface at e. Noting that the electronic specific heat is just 2 yT — e LOT ID.- [ne(8) - nhIEIJIE=EF (III-2°) D.) e we see that eq. (111.18) reduces to m = YT Sxx(H + ) (“e _ “n’e (111.21) If we now assume that the scattering is predominantly elastic, then we can use the relation (111.22) to evaluate the second term on the right hand side of eq. (111.17) in the high field limit and obtain xy yy = eLoT [p x(H + m)::i:: : :I %; oxyiH + 0°)I€=€F + pxy(H + w) I:::: : :I g; oyy(H + m)|€=EFI = e101 [g— in oxy(H + m) - gg.oyy> 9 expected to be much shorter than Tpe' sequently So) and Tpe are independent of temperature, and Tpx is inversely proportional to T (for T >> GD, the number of phonons is proportional to T). We thus see that the phonon drag thermopower, as given by eq. (111.39), should vary inversely with temperature. At low temperatures, T << GD, 1 is much longer than 1 and therefore eq. px (111.39) reduces to eq. (111.38). pe We will now briefly examine the effects upon S9 of some departures from the ideal free electron gas be- havior which may be expected in a real metal. A real metal has Bragg reflection planes. Consequently, Umklapp scattering processes are allowed, and these can have a pro- found effect on 89. To illustrate this point we consider both Normal and Umklapp scattering processes for a spherical Fermi surface. Umklapp scattering is distinguished from Normal scattering by the appearance of a reciprocal lattice + vector G in the momentum conservation relation, i.e., E - 1' = a + 6 (111.40) where q is the phonon wavevector. For a Normal scattering .+ event G = 0. Figure 111.2(a) shows a phonon-electron Normal scattering event. From this figure we see that the change in the electron's momentum Ak is in the same 51 \ I \ \ ~ I .‘ 5 ‘- ‘s ‘ H ‘ \—_ o’ ‘- (a) (b) Figure 111.2: The change of electron wavevector accom- panying the annihilation of a phonon by a I?) Normal pro- cess, (b) Umklapp process (after Barnard‘2 ). 52 direction as 3. Since the electron's velocity is directed normal to the Fermi surface, i.e., in the direction of k for the cases considered in Figure 111.2 16 is in the same direction as a, so the electrons will move in the same direction as the phonon flux, (i.e. down the temperature gradient), thus contributing a negative 89. Figure 111.2(b) shows a phonon-electron Umklapp scattering event. From this figure we see that after transforming the momentum of the electron after the interaction k' by a reciprocal lattice vector, Ak is now approximately anti- parallel to 3; thus the electron will move up the temper- ature gradient producing a positive 89. So for a spherical Fermi surface,Norma1 scattering events produce a negative 89’ while Umklapp scattering events usually produce a positive 89. However, most metals, including aluminum and indium, have complex Fermi surfaces. Bailyn(24) and Ziman(25) both employed variational techniques to cal- culate S9 for a complex Fermi surface taking both Normal and Umklapp scattering into account. They each obtained quite general expressions for 59. A discussion of their expressions would take us far afield from the major topics of this thesis, so only the results concerning the sign of S9 will be considered. Ziman's result leads to the approximate rule that "the contribution to S will be 9 negative or positive according to whether the chord q 53 between the points k and k' passes through occupied or un- occupied regions of the Fermi surface."(25) Bailyn's analysis leads to a similar rule. To illustrate the significance of Ziman's rule for a complex Fermi surface, we consider aluminum, a nearly free electron-like metal. The Fermi surface areas of aluminum lie only within the second and third Brillouin zones. Figure 111.3 shows a schematic representation of two dimensional slices through the second and third zones showing the occupied regions. A number of phonon-electron scattering events (both Normal and Umklapp) are indicated, and their contributions to S9 can be discussed within the framework of Ziman's criterion. In Figure III.3 all A type scattering events should contribute negatively while the B type should give positive contributions. All the results stated so far concerning Sg have been confined to the H = 0 situation. Almost all the work in the presence of a magnetic field has concentrated on the single component (a ) . Blewer, et. al.(27) have XY 9 shown for an isotropic phonon gas that %c , (111.40) an expression which is quite similar to eq. (111.37). Contrary to the electron diffusion terms, Opsal‘zg) has shown that none of the phonon drag terms of the thermo- electric tensor are enhanced by the electron-phonon 54 Filled region .. ._ ._ ._ _, N process *— U process Figure III.3: Schematic diagrams of sections throu h th e 2nd (upper) and 3rd (lower) zones of Al showing elegE n- phonon scatter1ng events (after Gripshover, et. a1.CZ g?). 55 interaction. Currently, calculations of the phonon drag terms of the thermoelectric tensor in the presence of a magnetic field have not been performed for a single metal. IV. NON-OSCILLATORY THERMOELECTRIC PHENOMENA - EXPERIMENT A. Aluminum 1. Introduction Measurements of the magnetothermopower of poly- crystalline samples of "pure" aluminum and very dilute aluminum based alloys (25 ppm to 350 ppm impurity concentra- tion) are reported by Averback‘3) and Averback, et. al.(4) Their results span a temperature range between 2 and 6K, and a magnetic field range up to 20kG. The results may be summarized as follows: (1) For all samples the data were consistent with the equation S(H,T) = A(H)T + B(H)T3, allowing a separation of the electron diffusion thermopower component Sd(H,T) A(H)T from the phonon drag component Sg(H,T) = B(H)T3. This separation assumes that in the presence of a magnetic field Sg retains its zero field form i.e. Sg a T3. (2) Upon application of a transverse magnetic field A(H) (i.e. Sd) first became more positive and then appeared to saturate in high fields. (3) Although A(H = 0) and A(H + co) varied from impurity to impurity, the quantity 56 57 AA = A(H + 00) - A(H = 0) was nearly impurity in- dependent; AA varied only from 2.1 - 2.6 x lO-BV/KZ. This insensitivity to the nature of the impurity suggests that ASd is determined primarily by the properties of the host metal aluminum. Using no adjustable parameters, Averback and (5) Wagner calculated AA for alumiunum. They obtained AA = 1.6 x 10-8V/K2, which is approximately 30% smaller than the experimental value. However, Averback and Wagner's calculation assumed that the magnetothermopower was got influenced by the electron-phonon mass enhancement, that a l—OPW Fermi surface would adequately describe the electronic properties of aluminum, and that the applied magnetic field was directed along an axis of four-fold symmetry in a single crystal containing an impurity which scatters electrons isotropically. The experimental re- sults, on the other hand, had been obtained with poly- crystalline samples containing impurities for which no independent information concerning scattering anisotropy is yet available. It thus seemed possible that the 30% discrepancy might arise from any of three sources: (1) Incompatibility between theory and experi- ment, in that the measurements were not per- formed on appropriate samples under proper con- ditions for comparison with the calculation. 58 (2) Inadequacy of a l-OPW Fermi surface for calculation of the thermopower in aluminum. (3) Neglect of the electron-phonon mass en- hancement. In an effort to experimentally test the via- bility of source (1), measurements were performed on an oriented single crystal foil of aluminum containing 50 (7) ppm gallium. Recent calculations by Sorbello indicate that gallium in aluminum should scatter electrons nearly isotropically.* The single crystal was oriented so that the magnetic field could be directed along both the four- fold symmetric [010] axis and the two-fold [110] axis. The data obtained from this crystal will be compared with the polycrystalline data of Averback, et. al.,(4) and also with the theory presented in the previous section. As already indicated in section III.A.2, Opsal (8) and Wagner recently ruled out source (2), by showing that a 4-OPW Fermi surface still leaves a significant discrepancy between theory and experiment. It will be shown that source (3) is the major cause of the dis- crepancy between theory and experiment. The single * Sorbello showed explicitly that the scattering of the third zone electrons by gallium was isotropic to within a few percent, and indicated that the scattering of second zone electrons was also nearly isotropic. Using his phase shifts and amplitude factors, it was found that the scattering of second and third zone electrons is the same to within a few percent (see Appendix I). 59 crystal data can be explained by the theory of section III.A.2 provided an electron-phonon mass enhancement of 45% is used. 2. Sample The aluminum foil was 0.6mm thick, 1.6mm wide, and the superconducting NbTi potential leads were 30mm apart. The crystal was oriented so that the [001] axis lay within 1° of the longitudinal axis of the foil and the [010] axis was rotated 20° from the normal to the foil. Once the sample was mounted in the sample holder, the heat current flowed along the longitudinal axis, and the mag- netic field H was directed transverse to the heat current. The axis of rotation of the magnet coincided with the longitudinal axis of the sample. Therefore the magnetic field could be aligned along either the four- fold symmetric {010} family of axes or the two-fold symmetric {110} family. Thus, it was possible to deter- mine whether AS was sensitive to the direction of H' d relative to the crystallographic axes of aluminum. 3. Data and Analysis 3.1 Determination of the non-oscillatory results In the absence of any applied magnetic field, the thermally generated voltages were measured using the nulling technique described in sec. II.A.2. However, for H > 10 k6, the presence of giant quantum oscillations when H was directed along a four-fold symmetric axis‘29-3l) 60 (see section VI.A.3) made the field sweep measurement technique (see sec. 1.A.3) advantageous. Figure 1V.l shows the maxima, minima, and the derived values of the non-oscillatory component of S for a typical field sweep after making the corrections outlined in sec. II.G. The non-oscillatory component of the thermopower was obtained by averaging each minimum with the average of its two adjacent maxima, and each maximum with the average of its two adjacent minima. From figures similar to Figure IV.1 the non-oscillatory of S could be extracted as a function of H at a constant temperature. As indicated in Figure IV.1, the non-oscillatory component saturated in value, to within experimental uncertainty, by 15kG*. 1n the ensuing analysis the lSkG value will always be used as the high field limiting value of S (above lSkG the measur- ing uncertainties increased rapidly with increasing field). Since the thermopower oscillations are so large, small uncertainties in their magnitudes are magnified into large uncertainties in the magnitude of the non-oscillatory component. With care and patience the non-oscillatory component was extracted with an uncertainty of about 10% (at a 70% confidence level). The bulk of this uncertainty arises from slow variations of the zero baseline and from * In some cases S had not quite saturated by lSkG, but the resulting small increases do not significantly change any of the conclusions reached in this thesis. 61 .mmmmouo can an omucmmonmou me u no ucmcomfioo muoumHHaomOIco: one one .moaouwo an owucmmoumou one mcofiumaaflomo Esucmsv may no masses one oEchE one .mflxm Hoaog on» macaw omuoonflo ma macaw ommmcmmfi owe .oeoson Andean one now smm.m so oeoem oeuodoos gees n no doeuneen> one “H.>H mesons .mux.\« ON mp O— m P _._.e___________1____ e IV: a .1 S o / o :1 IL I X 0e. 1 III. eeeeeeeeeoeeoee e e e icw one... Ii \ll ee A e I: III cc. VA e 1% Z O ( e .10 62 electrical noise created by vibration of the sample holder in the magnetic field. On the basis of the polycrystalline aluminum (4) measurements by Averback, et. a1. it is expected that the single crystal data should be of the form S(H,T) = A(H)T + B(H)T3 (1v.1) Therefore if we plot S/T vs T2, we should obtain a straight line with slope B(H) and intercept (with the T2 = 0 axis) A(H). We are interested in the quantity ASd ‘1'— : AA = A(H + co) — A(H = 0) (1v.2) 3.2 AA for ngarallel to [100] and [010] axes In Figure IV.2, the ratio of the non-oscillatory component of S to the absolute temperature T is plotted as a function of T2 for H = 0 and H = lSkG. The error bars reflect the 10% uncertainty which was previously assigned to each point. In each case the data are consistent with straight lines. If there is any misalignment of the potential leads on the sample, the value of S should change when the magnetic field is reversed. To test for such an effect, measurements were made with the magnetic field both forward (open symbols) and reversed (filled symbols). To within experimental uncertainty, no systematic differences were observed. Straight lines were therefore drawn through 63 ’-~ 10 . A. ‘3 N A . D x ‘A‘ ‘\~ > Ib— ~_a 00 X O _ o >< h- —— ‘\~ (I) T2IKZI Figure IV.2: The temperature dependence of the non- oscillatory component of S of the Al(Ga) sample for: H = lSkG and parallel to [010] (squares); H = lSkG and parallel to [100] (triangles); H = lSkG and parallel to [110] (crosses); H = 0kG (circles). The full symbols in- dicate data obtained with H reversed. 64 all data for a given field direction. The quantity AA can be obtained directly from Figure IV.2; it is simply the difference between the intercepts for H + w (here taken to be 15kG) and H = 0. To test whether the results were dependent upon the orientation of H relative to the foil axes, ASd was measured with the magnetic field along the two crystallo- graphically equivalent [010] and [100] axes. Figure IV.2 shows that the data for these two directions are systematically different, but the differences are compar- able to the uncertainties in the data. The average of the [010] and [100] values gives a "best value" of AA = (2.23 i 0.25) X 10-8V/K2. This average value agrees to within experimental uncertainty with the AA values obtained for each direction individually. The average value of AA is in good agreement with (4) the polycrystalline data of Averback, et. al. It is also in agreement with the calculation of Opsal and Wagner,(8) provided that a 45% electron-phonon mass enhancement is included. A 45% electron-phonon mass enhancement is con- sistent with values inferred from both de Haas-van Alphen‘32) (33) and low temperature specific heat data on aluminum. 3.3 AA for H parallel to {110} axes We have noted in section III.A.2 that Opsal and (8) Wagner predicted that ASd should be somewhat smaller for H directed along a two-fold axis than a four-fold axis 65 in aluminum. In this section results will be presented for H directed along each of the four, two-fold symmetric axes (the {110} family) lying in the (001) plane. The data for H along any of the {110} axes are shown in Figure IV.2. Since the data for all four equi- valent axes were the same to within the specified un- certainties, we will not distinguish between these axes in Figure IV.2. Also, for these axes the magnetothermo- power oscillations were an order of magnitude smaller than for the {010} axes. This accounts for the smaller error bars in Figure IV.2. From this figure we see that AA = (2.0 i 0.2) X 10-8V/K2 for H along these axes. This value is about 10% smaller than that for H along the {010} axes, in agreement with the calculation by Opsal and (8) Wagner. 3.4 Effects of Magnetic Breakdown The previously mentioned giant quantum oscilla- tions in S for H along the {010} axes (Figure IV.l) are (29,30) How_ believed to arise from magnetic breakdown. ever, Opsal and Wagner's calculation of the non-oscillatory component of S completely neglects breakdown. The follow- ing argument is made against the presence of significant effects of breakdown on the non-oscillatory component. (34) effects According to Balcombe and Parker, of magnetic breakdown should appear primarily for H in the (001) plane and should be most pronounced for H along the 66 [010] axis. In such a case the largest effect of break~ down on 8 would be expected for H along [010], a smaller effect for H along [110], and the smallest effect for polycrystalline samples. Indeed this is just what is ob- served for the oscillatory component: there are giant quantum oscillations for H along [010]; oscillations an order of magnitude smaller for H along [110]; and no evidence of quantum oscillations in comparable purity polycrystalline samples. On the other hand, as regards the non-oscillatory component little difference is found between the values of AA for H along [010] and along [110] and both values are consistent with those obtained with polycrystalline samples by Averback, et. al.(4) There- fore, it is concluded that magnetic breakdown does not significantly affect this component, and thus the values for AA should be amenable to the calculation of Opsal and (8) Wagner, provided that the electron-phonon mass enhance- ment is included. 3.5 Phonon Drag Besides measuring the electron diffusion contribu- tion to S, our experiments also provide information con- cerning the phonon drag contribution. Recalling from eq. (1V.l) that S9 is simply B(H)T3, it follows that the slopes of the lines in Figure IV.2 give B(H). The values of B(H) for the aluminum single crystal are shown in Table IV-l. 67 Table IV—l: Values of B(H) for Al(Ga) Sample B(x 10'1OV/K4) H -3.2 i 1.5 lSkG, H ||{110} -2.1 :_o.9 lSkG, H ||[0101 -o.5 : 1.5 lSkG, H III1001 -1.0 i 0.6 0.0kG. The variance of B with magnetic field and also the measured values of B are similar to the results reported by (4) Averback, et. al. The fact that B(H) is negative may indicate that the A type scattering events of Figure III.3 predominate over the B type events (see section III.B). Averback, et. a1. noted that they were unable to find any quantitative feature of the variation of B with H which was impurity independent. They therefore concluded that the magnitude of the observed variation of B with H is not an intrinsic property of the host metal aluminum, but is determined primarily by the details of the scattering of electrons by impurities. No detailed explanation of the observed variation of B with H has been offered to date. One must be cautious when interpreting B as solely a measure of phonon drag. If the scattering is partially inelastic, then the electron diffusion magneto- thermopower, in the presence of a magnetic field may also (35) contain terms which are non-linear in T. If these 68 contributions are small compared to the linear term, then an S/T vs. T2 plot will incorporate these contributions into B (the slope). The magnitude of inelastic scattering may be estimated by examination of the variation of the Lorentz ratio with temperature. Lorentz ratio data from (36’shows an aluminum sample of comparable purity by Willot that this ratio changes about 10% between 0 and 4K. This could cause Sd/T to vary by at least several percent over this temperature range; moreover, the variation is expected to be field dependent. Since S/T varies by only 25% over this temperature range, we conclude that inelastic scatter- ing could make a non-negligible contribution to B. Nielson and Taylor(37) have shown that second order multiphonon scattering processes may also give a T3 contribution to Sd; the so-called "phony phonon drag." Bourassa has estimated the size of the "phony phonon drag" for aluminum in the absence of a magnetic field and finds it to be orders of magnitude smaller than the ob- (4) served values of B. Currently the effect of a mag- netic field on "phony phonon drag" is unknown. In con- clusion, although the coefficient A(H) gives a quantita- tive measure of 8d in the zero temperature limit, the experimental value of B(H) does not necessarily measure only 89. We therefore believe that a detailed theoretical analysis of B(H) solely ins terms of phonon drag, may be a bit premature at this time. 69 4. Summapy_and Conclusions for Aluminum Transverse magnetdthermopower measurements of a single crystal aluminum sample (containing 50 ppm Ga) have been made with H along both the {010} and {110} families of axes. In each case S was consistent with the form S(H) = A(H)T + B(H)T3, allowing a separation of Sd from $9, and therefore permitting a determination of the quantity AA. For H along the four-fold symmetric {010} an average value of AA = (2.23 i 0.25) x 10-8V/K2, was obtained” and for H along the two-fold symmetric {110} axes a value of AA = (2.0 i 0.2) X lO-BV/K2 was obtained. These values differ only slightly, and they agree with the values pre- viously reported by Averback, et. a1.(4) for polycrystalline samples. We have argued above that these values are appropriate for comparison with the calculations of Opsal (8) and Wagner. This comparison leads to good agreement, provided that a 45% electron-phonon mass enhancement is included. A 45% enhancement factor is consistent with that independently obtained from other types of measure- ments. The fact that AA is somewhat smaller for H along the two-fold symmetric axis than for H along the four-fold symmetric axis is also in agreement with Opsal and Wagner's calculations. Data was also obtained for S9 = B(H)T3. The values of B(H) are shown In Table IV-l and agree with the values previously reported by Averback, et. al.(4) for 70 polycrystalline samples. We have noted that these values of B(H) may not be solely a measure of S . B. Indium 1. Introduction Indium and aluminum are both Group III metals. Aluminum crystallizes in a face centered cubic structure for which the {100} family of axes are all four-fold symmetric. In contrast, indium crystallizes in a face centered tetragonal structure (c/a = 1.08) for which the [001] axis is four-fold symmetric but the [100] and [010] axes are only two-fold symmetric. Thus, it seemed quite natural to extend the magnetothermopower measurements made on aluminum, to indium single crystals for H aligned along both the [010] and [001] directions, in order to try to observe the effect of the slight tetragonal distortion on ASd. Although the Fermi surfaces of aluminum and indium are quite similar, previous measurements of their magneto- thermopowers by Caplin, et. a1.(38) gave apparently dif- ferent results. They measured the magnetothermopowers of aluminum and indium polycrystals over the temperature range 4.2 to 60K in magnetic fields up to 50kG. The significant difference between the results of the two metals showed up in the quantity AS(H) (AS(H) E S(H) - S(H = 0)). Only in aluminum did AS(H) change sign from positive to negative 71 as the temperature was increased. In indium, AS(H) re- mained negative at all temperatures investigated. Using magnetic fields less than 50kG., Caplin, et. al. also noticed that S tended to saturate at high fields for T.< 21K for aluminum and for T < 9K for indium. Saturation indicates that the high field limit (wcr >> 1) has been reached. They also measured the thermopower of indium single crystals over the temperature range 4.2 to 300K in zero magnetic field. They observed that S was always more positive for VT along [100] than for VT along [001]. Averback and Bass(39) performed measurements on polycrystalline indium samples in the temperature range 2 to 5K in magnetic fields up to 20kG. Comparison of the data of Averback and Bass with that of Caplin, et. a1. shows a discrepancy in the sign of S. Measurements per- formed during the course of this thesis indicate that the sign given by Averback and Bass is incorrect.(4o) Averback and Bass also observed that S saturated in high magnetic fields. In addition, they noted that S did pg; fit the simple form: S(H,T) = A(H)T + B(H)T3. Two possible reasons are: (l) The low temperature approximation of S9 (= BT3) may not be adequate. However, if elastic impurity scattering is still the major scattering mechanism then 72 S(H,T) = A(H)T + Sg(H,T) . (IV.3) For this case, if a term linear in T can be extracted from the data then Sd can be separated from S9 (assuming Sg does not contain a term linear in T). (2) If there is a significant amount of in- elastic scattering, then S will no longer be d linear in T. In this case Sd cannot be un- ambiguously separated from S . 9 (41) measured the electrical Blatt, et. a1. resistivity of polycrystalline indium wires. The RRRs of Blatt, et. al.'s samples were comparable to the RRRs of Averback and Bass' samples. Blatt, et. al.'s data show that the electrical resistivity is strongly temperature dependent between 2.0 and 4.2K. This indicates that there is a significant amount of electron—phonon (inelastic) scattering at 4K. Thus, when the present study began it seemed possible that reason (2) could be the cause of de- viation of S from the simple form of eq. (IV.1). In this thesis, measurements similar to those of Averback and Bass were performed on oriented single crystals of dilute indium alloys. It was anticipated that added impurities would increase the electron-impurity scattering to where it would finally dominate the in- elastic scattering, causing Sd to become linear in T for T < 5K, and thereby allowing a separation of S from 89. d 73 On the other hand, the alloys must be very dilute in order to reach the high field limit with H < 20kG. In addition to measuring S, both the electrical and thermal con- ductivities were measured as functions of H and T. This allowed calculation of the Lorenz ratio L as a function of H and T. The divergence of L from L6 (= 2.443 x 10'-8 watt-ohms/Kz) gives a measure of the amount of inelastic scattering present at a given T. The variation of L with T gives a more sensitive measure of the amount of inelastic scattering than does the electrical resistivity alone. To summarize, the aims of the magnetothermopower measurements on indium were: (1) To see the effect of the slight tetragonal distortion on 48d by measuring S with H aligned along both the four-fold symmetric [001] axis and the two-fold symmetric [010] axis. ' (2) In order to achieve goal (1) it is necessary to separate Sd from 89. We hOped to do this by adding impurities to increase the electron- impurity scattering to a point where it finally dominates the inelastic scattering, causing Sd to become linear in T for T < 5K. (3) By extending the magnetothermopower measure- ments on indium down to 1.5K we wanted to see whether AS(H) would change sign from negative to positive as the temperature was decreased, as it does in aluminum. 74 2. Data and Analysis Figures IV.3 - 1V.8 show magnetothermopower data for a variety of indium crystals. The data are plotted in the form S/T vs. T2 at constant H. Although in some cases for a particular H the data lie on straight lines, not a single sample produced data which lay on straight lines for 211 H. This demonstrates that the magneto- thermopower of indium, in contrast to that of aluminum, does not obey the simple form: S(H,T) = A(H)T + B(H)T3 throughout the temperature range 1.5 to 5K. Figure IV.9 shows a plot of S vs. H at both 4.7lK and 3.50K for a sample containing 100 ppm tin in indium (In(Sn)-II). The data appear to be approaching saturation, but have not reached saturation at H = 20kG. Presumably this non-saturation by 20kG stems from the fact that the sample contains 100 ppm tin; this places an upper limit of about 100 ppm on the impurity content of the indium alloys if the high field limit is to be attained with the present measuring system. The fact that the data appear to be approaching saturation, is in agreement with the behavior found by both Caplin et. a1.(39) (40) and Averback and Bass. A goal of this research was to test the hypothesis that S would fit the simple form of eq. (1V.l) throughout the temperature range 1.5 to 5K for sufficiently high impurity concentrations. Therefore measurements were 75 F_ v - 0.02 KG. 4" ,v o 0.50 KG. ‘7 C] loC>C>I<<3¢ Xx iv XZOeOOKGe 2...... .r‘ _ 8 o o E 3%; cy__..__.___0 __. fl...E}-El..£2_.._£352_ __. C. \\ U) " X v -2— Vi v P- X ‘7 ‘7 _4____ x I I I I J 0 I0 20 T2(K2) Figpre IV.3: The temperature dependence of the non- oscillatory component of S for sample In(Ga) with H directed along [010]. 76 I <1x S/TxIOBCV/KZ) I xc><0 X ' 0002 KG. 0 0.50 KB. El I.OO KG. - v 5.00 KG. AI0.00 KG. x2 0.00 KG. ._£3 J I .I I l l 0 IO 20 30 TZCKZ) I J) I .894“ I :KB> <3 Figure IV.4: The temperature dependence of the non- oscilIatory component of S for sample In(Ga) with H directed along [001]. 77 F. 4L— x 0 CD. 0 bill 0 d3 063 D 0 xx CD 953 l x 2_ 9A. 0 (CD) 0 G "x a." v x E- x V E; I! ‘7 ‘7 \J X 0 0...... _I_ 9. 4x " - 0.02 KG. I" x a t. o 0.50 KG. 9 x A -2... O 3.00KG. . A V 5.00 KG. Xx I I. AI0.00 KG. x .. II5.00 KG. X --.4_. x20.OO-KG. X I I I I 0 IO 20 TZCKZ) Figure IV.5: The temperature dependence of the non- oscillatory component of s for sample In(Sn)-II with H directed along [001]. 78 5r“ . p 6):.) <2. <2] 8 (5% I— O I o o __ g V O O A O V— _— N A v ¥ ”" v 9 L_ . V V A “’9 ._ - 0.02 KG. A f: __o 0.50 KG.’ I \\ (0-5....0 |.00 KG. x A A __o 3.00 KG. x . . K . r.AIOOO G _'O___...x20000 KGo I I I I 0 IO 20 T2(K2) Figure IV.6: The temperature dependence of the non- oscillatory compopent of S for sample In(Sn)-II with H directed along [110]. 79 55'—‘ I . 0.02 KG. 7 II? . o 0.50 KG. _’ §<1C> <3 0 _- Xx x20.00 KG. 0 _ ______ (\lx __ S __ mu 2 __ l: _. ‘\ (1)-5— 7“ v _— X Y .40— l l I l l 0 IO 20 T2(K2) Figure IV.7: The temperature dependence of the non- oscillatory component gf S for sample In(Sn)-I with H rotated 12° from the [101] in the (101) plane. 80 mamEmm you w mo muoumHHfiomOICOG 0:» mo mocoocommo munuoxomfimu one .HHHiAdmvcH "m.>H ousmwm €31 on o. o _ d i _ _ T - Ix). IIIIIIIIIIII 0 (did AV d I. xx aw x x x x xx x x x xx x x x44 I I I 93:... .oxoooNd a I In: a a. m . .3506? an n a . ‘ .3506.- ono In D0831 . . o o 3.00 o 0 COO Iv .oxmoo . . Doc 0 O In . .o .o lo (ZN/A > 8le 1/8 81 N C).— r\ n X: \\ :> PS? “I h— 5 x T=5.5K a) a -2__ ‘3‘— T:4e7K _4___ I I I J 0 IO 20 H(KG) Figure IV.9: The magnetic field dependence of S for sample In(Sn)-II with H directed along [110]. 82 performed on a sample containing 500 ppm tin in indium (1n(Sn)-III). Figure IV.8 shows that the data for H 15kG and 20kG with H directed along the [001] axis lie on straight lines. This tempts one to make the hypothesis that all the data for H > lSkG will also lie on straight lines. However, one must also note that for this sample H = 20kG is a relatively low field (mcr < l); and as illustrated by Figure IV.6 the form of an inter- mediate field (H = 3kG or SkG) curve over a limited tem- 2 < 25K2) can be quite different perature range (10 < T from the high field curves (H > lSkG in Figure IV.6). In fact these intermediate field curves over this limited temperature range are quite similar to the curves for H = 15 and 20kG with H directed along the [001] axis of Figure IV.8. Therefore, since intermediate field curves are not reliable indicators of the high field curves, and also because the H = 0.50kG data do not lie on a straight line, we believe that S will not fit the simple form of eq. (IV.l) at higher fields over our available temperature range. Figure IV.8 shows that the data for H directed along the [110] axis deviate strongly from a straight line already at 20kG. So it is concluded that for an arbitrary H, S will not fit the simple form of eq. (IV.l) throughout the temperature range 1.5 to 5k for indium samples with impurity concentrations less than 500 ppm. 83 The nonlinearity of Figures IV.3 - IV.8 is most 2 pronounced in the temperature region above 3K (T = 9K2). The experimental data of Caplin, et. a1. show a strong negative peak in the vicinity of 10K. They ascribe this peak to phonon drag. It seems likely that the available temperature range is too close to the phonon drag peak and therefore Sg cannot be adequately described by the low temperature approximation (S = BT3). As a rule of thumb,‘ the phonon drag peak occurs between BD/S and 00/7 (0D is the Debye temperature). Averback and Bass tentatively attributed the inadequacy of the low temperature descrip- tion in the temperature region above 3K, to the low Debye temperature of indium (SD = 110K);aluminum which has a much higher Debye temperature (0D = 425K), begins to deviate from the simple form: S(T) = AT + BT3 above about 6K. ‘42) The other apparent cause of the non-linear S/T vs. T2 plots for indium concerns the effects of the in- elastic scattering on S For this case, it would be d' expected that a sample with a significant amount of in- elastic scattering would produce a different S/T vs. T2 plot when compared with that of a sample with a smaller amount of inelastic scattering. Figure IV.10 shows the Lorenz ratio as a function of T for a few samples. The deviation of L/Lo from 1 gives an indication of the amount of inelastic scattering present. This deviation is 84 x‘,,xx 100’— 88 "xx C”) ”x I O X (3 no 00) ‘ A " Q00 00 "99,I 00:0 09 I x“ (la .8)—- O xlxo D l X0 8 O ‘0 _x___I _ O D O 3'6I— V U o x O v D 0 )-— x D 00 D V O X on 00 V DD .4?- xv DOD XV o 0.50 KG. ”V C D I.OO KG. V ‘, Vvv V 5.00 “6' X‘ .21 x20.00 KG. "" l l l I l O 5 T(K) Figure IV.10: The temperature dependence of the Lorenz ratio L divided by the ideal Lorenz ratio L for samples: A) 1n(Sn)-III, with H directed along [001]; B) In(Sn)-II, with H directed along [001]; C) In(Ga), with H directed along [010]. 85 largest in our purest sample, In(Ga) (RRR = 7400). The fact that L deviates most from L0 is expected since this sample should have the least electron-impurity (elastic) scattering. Figure IV.10 also shows that sample In(Ga) exhibits significantly more inelastic scattering than sample In(Sn)-II. However, Figures IV.3 - IV.6 show that these two samples produce similar S/T vs. T2 plots. Thus it is concluded that the major cause of the nonlinearity of Figures IV.3 - IV.8 in the region T > 3K is the break- down of the low temperature description of 89, rather than the influence of inelastic scattering on Sd' Probably the most interesting region in Figures IV.3 - IV.8 is below 2.5K (T2 < 6K).‘ For all samples ex- cept In(Sn)-III, the high field set of data (H > lOkG) either becomes more positive than the 0.50kG set (essen- tially H = O) or it appears that this will happen at a temperature somewhat below 1.5K (the lowest temperature attainable with the present system). This behavior was not suggested by any previously published data. However, such behavior would be expected when Sd is the dominant component of S. Moreover, the only abnormal case, In(Sn)-III, can be understood by noting that this sample has not reached the high field limit by 20kG, and we pre- viously argued that the intermediate field curves over a limited T range do not reliably predict the high field S/T vs. T2 curves. The general low temperature behavior 86 of these samples is qualitatively similar to that of aluminum, which is only slightly affected by phonon drag. Perhaps this low temperature region is far enough removed d is comparable to $9, in which case S might be described by from the large negative phonon drag peak, so that S the simple form of eq. (IV.1). However, this is only a conjecture concerning the scattering mechanisms responsible for S. The true test will come when magnetothermopower measurements are performed at still lower temperatures (0.3 < T < 1.5K). At such low temperatures elastic, electron-impurity scattering should be dominant, so Sd should be linear in T, and moreover, Sg should be small compared with Sd' 3. Umkehreffect Samples In-I and In(Sn)-I demonstrated an inter- esting umkehreffect. The umkehreffect is characterized by the magnetothermopower changing drastically upon re- versal of H.‘ Figures IVJLland IV.12 show this effect. For H near [100] the magnetothermopower changes by a factor of two upon field reversal, Figure IV.12 also shows the angular variation of the electrical resistance. It is clear that the resistance, in contrast to the magneto- thermopower, is well behaved, i.e. not affected by field reversal. Such a result can be understood (at least qualitatively) by considering the properties of the‘E and 87 9“ A+I0.00 KG. _ “\. ’“.-IOOOO KGO .yf' \‘ ‘LA \ h'/ \ A ‘f \ / n 6" \ k: \ _ / s _ "‘ ‘ / c \ 1 N9 ’— I \ / .\ €13 \ / | A \ /O 3— - - ‘(A \_.__'_—/ A " [0:0] U01] 0 l l l l l I L I L Ill 1 J I I] l L I O ICC 200 300 9(DEG) Figure IV.11: Variation of S with rotation angle 0 at H = lOkG and T = 4.7K for sample In-I. H was rotated about the [101] axis. 6 = 213° corresponds to H directed along [010]. ' 88 8.5r-— A+l0¢00 KG. 5.. .-'0000 KC. g t. ”9 8.0—— ; _. )—-— (— 7.5%— 6— -sno7(vn<) [0:0] [10:] lllJJllllJLlJlllllel O ICC '200 300 B(DEG) Figure IV.12: Variation of R (upper curve) and S (lower curve) with rotation angle 0 at H = lOkG for sample In(Sn)-I. H was rotated about the [101] axis. 0 = 202° corresponds to H directed along [010]. The R curve was measured at 4.2K and the S curve at 4.3K. 89 the‘S tensors. For either tensor, symmetry requires that there be no umkehreffect for H directed along a three- fold or higher symmetry axis. As a consequence of the :13) ‘16) Onsager relation pik(H) there can be no = pki( magnetoresistance umkehreffect for H directed along a two- fold symmetry axis either. However, since there is no such Onsager relation connecting components of the‘S tensor, H must be directed along a three-fold or higher symmetry axis for the magnetothermopower umkehreffect to vanish. Since the [100] axis is only a two-fold symmetric axis, it follows that the magnetothermopower may exhibit an umkehreffect. Although such symmetry arguments allow the existence of an umkehreffect, they alone do not pre- dict its magnitude. Therefore, one cannot on theoretical grounds alone unambiguously attribute the observed magnetothermopower umkehreffect to an intrinsic property of indium as opposed to an experimental artifact (e.g. lead misalignment or sample preparation). To see if the observed umkehreffect was due to a small misalignment of the voltage leads attached to the sample, the leads were removed and reattached, aligned as accurately as possible. This had no significant effect on the measured value of S. Therefore, it is concluded that lead misalignment was not the principal cause of this irregular behavior. Samples In-I and In(Sn)-I were both grown from the same seed. Figures IV.11 and IV.12 show that both 90 samples exhibit the largest umkehreffect when H is in the vicinity of [100]. H could also be aligned along [100] with sample In(Sn)-III. The umkehreffect exhibited by In(Sn)-III, as well as all other samples, was less than 20%; therefore, it is concluded that the very large umkehreffect exhibited by In-I and In(Sn)-I is not intrinsic to H along [100]. Although it seems unlikely, one cannot rule out the hypothesis that this large effect is a consequence of 5 being directed along [011]. There- fore, further measurements on independently grown single crystals are necessary to determine whether the afore- mentioned hypothesis is correct or whether the large umkehreffect was an experimental artifact due to sample preparation, which did not show up in the X-ray (Laue) patterns.* It should be noted that Figure IV.7 shows data for a direction in which the umkehreffect was less than 10%. 4. Summary and Conclusions for Indium We briefly summarize the qualitative results for indium: (1) Magnetothermopower measurements performed on indium single crystals containing as much as 500 ppm impurity showed that S was not consistent with the simple * It is unclear whether or not higher quality X-ray photo- graphs would have shown a difference between these two samples and the other indium samples. 91 form: S(H,T) = A(H)T + B(H)T3, over the temperature range 1.5 to 5K. It is concluded that the deviation from this simple form is not primarily due to the effects of in- elastic scattering on S but rather to the influence of d’ the large negative phonon drag peak (near T = 10K), which causes S9 to substantially deviate from the form S9 = B(H)T3. (2) The magnetothermopower of samples with impurity con- centrations of 100 ppm or less appear to be approaching saturation for H = 20kG and T < 5K. (3) At high temperatures, 3 < T < 5K, S becomes more negative with increasing H. This is presumably due to S9 dominating S. (4) A low temperatures, 1.5 < T < 2.5K, the high field limit of S is more positive than the low field values. This can be understood by assuming that-Sd is increasing relative to S9 with decreasing T. The temperature at which S starts becoming more positive instead of more negative with increasing field varies somewhat from sample to sample. V. OSCILLATORY THERMOELECTRIC PHENOMENA - THEORY A. Variation of the Density of States at SF with H The state of a free non-interacting electron gas can be described by a sphere in single particle momentum space (k-space). At T = 0 the interior of the sphere has an electron occupying every single particle state, while the exterior states are all empty. The Fermi surface separates the occupied and unoccupied regions. In the absence of a magnetic field the single particle states of the electrons are evenly distributed throughout all of k-space. However, the application of a magnetic field redistributes these states on to the surfaces of cylinders aligned parallel to the magnetic field. Figure V.l shows such a redistribution of states. This figure also shows the H dependence of this redis- tribution. An electron, in any particular state, orbits around the circumference of the cylinder ("Landau cylinder") with an angular frequency wc (the so-called cyclotron frequency), given by O.) = ___ 0 (Vol) where me is the mass of the electron and c is the speed 92 93 (a) (b) flu. . .--..o‘ - .u..‘..V-.-..-.. o .3'.J'. ' _ ’- '.. ' . o ' .O..\‘......O... . ' '. a O . a" (C) (d) Figure V.I: Variation of the Fermi surface with field, showing the increasingly spherical shape as the field be- comes weaker. The field decreases going (a) to (d) (after Adams and Holstein(43)). 94 of light (in a vacuum). Thus, the effect of the magnetic field is to quantize the electron orbits in k-space. The radii of these cylinders increase with H. As the outer cylinder passes through the Fermi surface, the electrons on this cylinder "condense" back onto avail- able states on lower energy (smaller radius) cylinders. This process repeats as each Landau cylinder passes through the Fermi surface, giving rise to an oscillatory magnetic field dependence of the density of states at the Fermi level. This dependence has been observed in measurements of a number of high field (wCT >> 1) electronic properties such as magnetization (de Haas-van Alphen effect) or .electrical resistivity (de Haas-Shubnikov effect). The exact solution of the free electron gas in a uniform magnetic field is given in many solid state physics texts.(44) One finds that the oscillations in the density of states at the Fermi energy are periodic in l/H, with a period A(l/H) {2‘23 A (v.2) where A is the cross-sectional area enclosed by the electron's orbit. What happens when we are dealing with electrons in real metals? In some metals, such as indium and aluminum, the valence electrons can be treated as nearly free electrons (NFE). The NFE approach treats the crystal 95 lattice potential as a perturbation. The important changes introduced by NFE treatment are: (l) The electron orbits (and consequently the Fermi surface) are no longer spherical. (2) The Fermi surface can be made up of pieces from more than one band. (3) The dynamics of the electrons in their orbits are described in terms of an "effective mass" * m where * m = lg I (v.3) .1. 2“ e e=€F Q) Using the Correspondence Principle, it can be shown that electrons within a single band will still ex- hibit an oscillatory density of states at the Fermi level (44) The period is still given which is periodic in l/H. by eq. (v.2), but now A is the extremal cross-sectional area of the Fermi surface in the plane perpendicular to the magnetic field. B. Estimate of S Oscillations Due to Density of States Oscillations for Aluminum and Indium There do not exist any rigorous calculations of the oscillatory component of the thermopower of a metal which are suitable for quantitative comparison with ex- periment. Only simplified calculations of limited appli- (45) (46) al.,(47) cability by Zil'berman, Horton, Grenier, et. 96 (48) are available. All of these cal- and Long, et. a1. culations predict for the particular orbit of present interest within our experimental range of temperature and magnetic fields, that the thermopower oscillation ampli- tudes should consist of one or more terms of the form: 1 n 2 A1,nT H exp[-2n kB(T + TD)/BH] (v.4) where A1 n is independent of H and T. Here 1 and n are I eh * O m c In this section we will estimate the size of the constants 2‘: 1‘: -1 and 3/2 Z.“ 1 -3/2; and B = quantum oscillations of Sd induced by the oscillatory com- ponent of the density of states, under a series of assumptions which are necessary to allow the estimation to be performed with data that currently exist in the literature. The assumptions are indicated at the points they are made. Only one of the assumptions, if incorrect, might be expected to have a dramatic effect on the esti- mate; the limits of this assumption are indicated in a footnote following the assumption. From equations (111.20) and (111.24) we see that 2eL T e o d Sd(H + 00) e h F YY where 97 If we assume that the electrical conductivity can be described by a two band Sondheimer-Wilson model(49) then en H en H o =—_2————7ee +T‘7hh (v.6) yy H+H H+H e h where * m.c Hg= 1 . 1 e T. 1 3Hi 3Ti If we further assume that 3E— (i.e. 52-) gives a negligible contribution to the oscillatory component of * os Eyy' eyyc, then H 3n H an 83:6 3 -e2LoT[%2 e2 (aee)osc + 2h 2 (36h)osc] . (V 7) +He H +Hh e=eF If we still further assume that He = Hh = Ho then H an an osc 2 o [ e osc h osc e = -e L T ——- (-—) - (-—-0 I] (v.8) 2 YY o H 38 36 8:6 and t (50) . Recent measurements by Thaler, et. al. on a high purity (RRR = 7000) aluminum polycrystal revealed that under conditions where S displayed substantial quantum oscillations, the Nernst-Ettingshausen coefficient (i.e. exy) revealed none. Since exy is not dependent on 31/38, and eyy i§_dependent on 31/38, any significant oscillatory component in 3T/3e would cause oscillations in S but not in the Nernst-Ettinsghausen coefficient. Therefore, Thaler et. al.'s result suggests that the assumption that 31/38 contributes a negligible oscillatoryr component may not be completely valid. 98 eHo Oyy = ;7_ (ne + nh) . (v.9) Dividing eq. (v.8) by eq. (v.9) gives a eLoT 3ne osc 3n £0 = " ___—n + n (38 I ' (3e ' . (v.10) e h - Measurements of the electrical resistivity of the single crystal aluminum and indium samples used in the experiments comprising this thesis show that for osc H : 20kG., g < .01. From these measurements it may be inferred that ngsc and nfisc are negligible compared with ne and n at all fields of interest. De Haas-van h Alphen results show that both a1uminumI32'51) and . . (52,53) . . indium have very small third zone electron-like orbits and much larger second zone hole-like orbits. This means that the third zone orbits will produce 5 _ 107 relatively low frequency (~ 10 G.) oscillations compared to the higher frequency (~ 108G.) oscillations due to the second zone. From an experimental viewpoint, the highest frequency of oscillation which may be ob- served is limited by the homogeneity of the magnetic field over the entire length of the sample. In an in- homogeneous magnetic field, different portions of the sample will oscillate out of phase with other portions. This will wash out the oscillations in the thermal emf generated over the length of the sample. The highest 99 frequencies observed with present apparatus were about 5 x 106G. Therefore, only oscillations due to density of state variations of the third zone orbits can be ob- served. In view of this, the estimate of the quantum oscillations in 8d will be confined to these electron- like third zone orbits. We therefore write g eL T 3n (_yy)osc = _ o ( eI _ )osc . (V 11) Oyy ne + nh 36 5 -EF In both aluminum and indium the volume of the second zone hole-like portion is much larger than the volume of the third zone electron-like portion. This means that n + n 5 n 5 n - n = l/atom , (v.12) and consequently (Eyy)osc _ I Oyy ne - nh de e=eF ) . (v.13) Combining equations (v.5) and (v.13) gives osc eL T dne _ o osc sd (H,T) — n _ n ( ) , w T >> 1. (v.14) e h d8 l€=€F c dn We now must evaluate the quantity (522' )osc s=eF Within the framework of the Lifshitz-KosevichI54) theory for a NFE gas, GoldISS) obtains an expression for nose. For the fundamental frequency his expression (in MKS 100 units) reduces to 1 _. nosc = ——-l—-§-(e“hH)3/2 2" 4 IlKl sin(§% - 2wy+ %) (v.15) n where y is an undetermined phase factor, Q) 3’ is the curvature of the extremal cross-sectional muo area of the Fermi surface in the direction of H, X I1 = m' kBT << fiwc' * 2n2m k X’ = TB T H I * 2n2m kB K1 = exPI' ehH TD” . . (56) TD 18 the Dingle Temperature. The 1 factor gives the temperature dependence of the l amplitude of the oscillations in the carrier density. Since I1 decreases as T increases, the effect of a non— zero temperature is to decrease the oscillation amplitude. The oscillation amplitude will be further reduced if the electrons suffer collisions. This means that the electrons will possess a finite (not infinite) relaxation time. The K1 factor gives this correction because T is inversely D proportional to the relaxation time. Grenier, Zebouni, and ReynoldsI47) claim that “ ‘ anosc the most important term in 3? ls-s arises from the ‘ F derivative of the phase (recall that A = A(e)). This 101 assumption gives osc * l/2 _ e E—EF w'h '3 Al 2 (v.16) BkH osc d I the effect of a finite relaxation time will be neglected, In an effort to estimate an upper bound for S i.e. set K1 = l. The calculation will be performed for H aligned along the [010] axis in both aluminum and in- dium. For the B-orbit in aluminum the values: m* = 0.095 m (32) 8 32A _ 7(57) '7?" ‘ a H -1 _ _ -29 3 n = 15.0kG. T = 1.72K give an oscillation amplitude. SESC(A1) = 0.28 x lO-llV/K. (v.17) For the neck orbit of the B-arms in indium the values: * 58 m = 0.2 m I ) e 32A ,_ 5(57) '7' - 0b 3kH 102 15.5kG. H T = 1.67K give an oscillation amplitude 535°(1n) = 1.0 x 10-11V/K . (v.18) It is hoped that eqs. (v.17) and (v.18) give order of magnitude estimates of the amplitude of mag- netothermopower oscillations arising from oscillations in the density of states at 8F of aluminum and indium. These estimates suggest that such S oscillations should d be roughly the size of the signal noise in our measure- ments and much smaller than the non-oscillatory component of S. C. Magnetic Breakdown Up until now we have assumed that the electrons can make transitions only between states within the same band, i.e. we have ignored interband transitions. How- ever, in large magnetic fields electrons can tunnel be- tween bands, giving rise to the phenomenon of magnetic breakdown. When this happens, the semi-classical quantization scheme described in section V.A breaks down. The review article by Stark and Falicov‘Go) gives a more comprehensive study of this phenomenon than will be presented here. To understand magnetic breakdown we will first consider a free electron gas in a magnetic field. 103 Neglecting interactions among the electrons, the electron wavefunctions are those of a particle in a circular orbit in the (kx, ky) plane, which is normal to H. We then introduce a lattice perturbation of the form V(x) = Z VGelcx. When an orbit passes through a zone boundary G (i.e. its wavevector in the x direction, kx, is equal to 1 1/2 G), there is the possibility of Bragg reflection. Using Figure V.2(a) as an example, this means that in- stead of continuing along AB, the orbit may switch to the direction AC, and so on. If the strength of the perturbation is increased, then the trajectories at A will be split apart in energy, and the route AC will be favored. The electron is now moving on an ordinary (open) orbit in the repeated zone scheme. This is shown in Figure V.2(b). The part B of the circle in Figure V.2(a) has now been joined up into a separate branch of the Fermi surface, and is traversed separately. As the magnetic field is increased, the effect of the lattice perturbation will decrease, so the circular orbit will become more and more favored. In- stead of the electron going along AC, it may break through the energy gap, ("tunnel through" the region separating the two orbits in k-space) and wind up going around orbit B. 104 .11 scams.s Hmummv spam couuomam mmum may ou xomn mesh awe uflnuo mnu macaw owumcmme coupm m CH pan .muwocsoa mcoN on“ um ompomccoomu mum muwnuo may .moauuma map mo Hmwucmuom oflUOflHmm on» CH ADV «vamwm oaumsmmfi m cw uflnuo couuomam mmum Amy "N.> musmwm 3v 3 105 In practice such behavior occurs at portions of the Fermi surface where the band splittings are very small (10'2 - 10‘3 eV.). This usually happens at symmetry points of the Brillouin zone. The best known cases correspond to splittings due to spin-orbit coupling. The description of magnetic breakdown as illus- trated by Figure V.2(b) can be extended to real metals like aluminum and indium. Figure v.3 shows the third zone pieces of the Fermi surface of aluminum and Figure v.4 shows a central (001) section of this Fermi surface. With a magnetic field along the [001] axis, electrons travel clockwise on orbits like FABH... (of Figure v.4), thus they are hole-like. In the absence of magnetic breakdown, a hole arriving at B from A will continue to H: otherwise it could break through the energy gap and "jump" on to the small electron—like B-orbit and proceed to C. Here it may "jump" to the hole-like orbit and proceed to D, or remain on the B-orbit and continue to B, then to H or C, etc. The net result of this, ignoring complications due to the pos- sibility of scattering while on the B-orbit, is that either the hole remains on its original (second zone) trajectory, or it moves across the zone boundary on to another second zone trajectory and moves in the opposite direction. When the breakdown probability is large, the hole is "reflected" near the corner W' and thus travels 106 v x oasrr ‘3‘ ”‘8" , N I ’I \ I, ’ I) I / \ I I I / \ I / § ,’ \ , 4 l \\ / \ a ’ f \ I \ a \ \\ )’ ‘\ I \ \ ’ \l \ ’ I \ \ ’ / \ / \ / ‘\ \” ‘/ \.\ I! \ \ , .— \\ .—- .o- ’ Figure v.3: Ashcroft's 4-OPW pseudopotential model for the electron-like 3rd zone of the Fermi surface of aluminum (after Abele and Blatt(52)). 107 +v—“’. ‘ - . Figure V.4: Central [100] section through the aluminum Fermi’surface (after Ashcroft(53)). 108 over the lens shaped, electronélike orbit. Similarly a hole arriving at C from G will move to either D or B: if it reaches B, it then goes toward H or returns to C. When the breakdown probability is large the resultant orbit GCBH... is circular and electron-like. Pippard(64) has derived an expression for the breakdown probability for a junction between orbits of the same topology as in Figure V.4. If R is the prob- ability of a particle proceeding along CDE after arriving at B from A, or proceeding along BH after arriving at C from G, then 2 R = P 2 (v.19) l - 2cose + Q * where P = exp(-HO/H), Ho =_egm c/effie, 69 being the energy gap between the bands at the point in k-space where the break through occurs, Q = l-P, and e is defined as 6 = 6 + -—— (v.20) where 60 is a constant and A is the area (in k-space) through which breakdown occurs (the B-orbit in our case). Thus, the breakdown probability has an oscillatory com- ponent with a period (in l/H) of Zne/dfiA. This is the same period as would be found in the ordinary (i.e. in absence of magnetic breakdown) de Haas-van Alphen oscillations associated with the small connecting orbit. 109 As the breakdown probability becomes alternately stronger and weaker with increasing field, the orbits of the central (001) section of the Fermi surface become alternately more and less electron-like and both the equil- ibrium and transport properties of the metal fluctuate accordingly. So far we have only described a simple micro- scopic picture of magnetic breakdown, and have not con- sidered the effect of this phenomenon on the macroscopic properties of the material. The equilibrium properties (e.g. de Haas-van Alphen effect) are the easiest to treat, sovuawill consider them first. Once the electron systemis free energy is known as a function of H, all the equili- brium properties may be calculated from it via the appropriate thermodynamic relations. To calculate the free energy requires a knowledge of the magnetic field dependent density of states p(e,H). Falicov and (65) Stachowiak have proposed a Green's function method for calculating the oscillatory part of p(e,H). The Green's function has been exactly calculated for the free electron case, and it is believed to provide an accurate description of the NFE case (after making the appropriate changes(60)). The important result of this procedure is that oscillations, periodic in l/H, will occur in the equilibrium properties. These oscillations will have a frequency corresponding to the area of the breakdown 110 orbits: this frequency will be the same as that observed in the ordinary de Haas-van Alphen type oscillations. However, when the breakdown nears completion new fre- quencies will appear: these frequencies reflect the new and larger orbits formed by breakdown (similar to the free electron circle shown in Figure V.3(a)). Such giant orbits have been seen in a number of metals, e.g. magnesium.(66) Currently the description of transport phenomena in the presence of magnetic breakdown is far from com? plete. The effects on electrical conductivity have been considered by a number of authors, some being Pippard,(67) (68) (60) Young, and Stark and Falicov. They all agree that magnetic breakdown will cause oscillations, periodic in l/H, which will have the same frequencies as those in the equilibrium properties. Stark and Falicov have per- formed a calculation for an idealized case, which in- dicates that the oscillation amplitude can be comparable to the non-oscillatory component of the transverse magnetoresistance.(60) At the present time very little work has been done on the effects of magnetic breakdown on magneto- thermopower. YoungIGg) has considered this problem for a different junction topology than that found in aluminum and indium. He obtains an oscillatory component of the magnetothermopower, which again exhibits the same 111 frequencies as those of the equilibrium properties. We might note’in passing, that if parameters appropriate to the particular orbits of present interest are substituted into Young's expression, then within the experimental range of temperatures and magnetic fields, this expression may be approximated by one or more terms of the form of eq. (v.4). Young also notes that for an "opening" net- work (similar to the junction topology of aluminum and indium) no actual calculations have been made, because a simple expression does not arise. However, there should be an oscillatory component. The present state of theory of the effect of magnetic breakdown on magnetothermopower oscillations is not far enough advanced to allow calculation of the amplitude of these oscillations. However, independent experimental evidence obtained prior to this thesis in- dicates that magnetic breakdown is present in all metals that have exhibited giant magnetothermopower oscillations* (this will be discussed in more detail in section VI.C.1). These giant oscillations have frequencies that correspond .to the small connecting orbits which facilitate the mag- netic breakdown. * The fact that the oscillatory component is comparable to, or much larger than the non-oscillatory component en- genders the designation "giant oscillations". VI. OSCILLATORY THERMOELECTRIC PHENOMENA - EXPERIMENT A. The Fermi Surfaces of Aluminum and Indium 'The Fermi surfaces of aluminum and indium have been extensively studied; and consequently the Fermi sur- face dimensions are known in detail. The Fermi surfaces of both these metals consist of a large second zone hole surface and a small third zone surface which is made of "arms" of electrons. Figure V1.1 shows the Fermi surface of indium. The second zone hole surface (Figure V1.1(b)) is very adequately represented by the free-electron model, and is quite similar for both aluminum and indium. The third zone electron surface of indium shown in Figure V1.1(c), consists only of B-arms, and these 8 arms are joined together in four-sided rings in the (001) (but not the (100) plane) by small junctions near point T.(59) The third zone surface of aluminum is similar to that of indium, but with a few significant differences: 1. The third zone of aluminum consists of both a and 8 arms (see Figure v.3). 2. The third zone of aluminum consists of four sided rings of "arms" in both the (001) and (100) planes (a consequence of cubic symmetry). 112 113 EM! 23 m couuooam macs chap» snaps“ may on fine “Eafiosfl accommnumu commucmonmomw .Afimmvonmsmmsm can mwnmzm Hmummv mUMMHSm “mUMMHSm «do: H0008 couuomamlmmum sawmcm How macs swooHHHHm one Ame "H.H> musmfim 114 3. In aluminum there is a swelling at the junc— tion where the B-arms connect: in indium this swelling is either smallISZ) (59) or completely absent. This is a consequence of the much stronger spin-orbit interaction present in indium. The free electron model predicts that the second and third zones would be degenerate at certain high symmetry points of the Brillouin zone. However, a non-zero spin- orbit interaction breaks this degeneracy, thereby closing both surfaces (i.e. there are no open orbits). The oscillations observed in this thesis have frequencies which are associated with the third zone B and y-orbits in aluminum, and the neck and belly orbits of theE3"arm" in indium. B. Aluminum 1. Introduction Since the spin-orbit interaction is quite small in aluminum, the second and third zones are almost but not quite degenerate at certain points in the Brillouin zone. Therefore, aluminum has a Fermi surface which is favorable to magnetic breakdown. Balcombe and Parker‘34) have observed oscillations in the magnetoresistance of aluminum for H along the [010] axis. These oscillations had the same frequency as the electron-like 8 orbits. Using a simple magnetic breakdown model, which allowed 115 holes (second zone) to "jump" from one orbit to another in an adjacent Brillouin zone using a third zone electron- like B-orbit as a bridge (see section V.C), their calculated. values for pxx and pxy agreed "tolerably well with the experimental results". Recently, Bozhko and VOl'skiiI7o) have observed oscillations in the magnetoresistance of aluminum for H along the [110] axis, when the temperature was below 1K. They also attributed these oscillations to magnetic break- down. Concurrent with the research constituting this thesis, two other research groups: Kesternich and Papastaikoudis,(29) (34) and Sirota, et. al., independently discovered the existence of giant quantum oscillations in the magnetothermopower of aluminum for H along the [010] direction. The oscillation frequency was the same as that observed by Balcombe and Parker.(34) Both of these groups attributed these giant oscillations to magnetic breakdown. Additionally, Kesternich and Papastaikoudis noted that the amplitude of the.thermopower oscillations exhibited a simple exponential decay in l/H over the entire field range (H < 48kG). Both of the aforementioned groups constrained their thermopower measurements to having the cold end of the sample remaining at about 4.2K, while producing temperature differences along the sample of about 1K. 116 The great sensitivity of the chopper-amplifier system allowed the measurements of this thesis to be performed using relatively small temperature differences (typically 0.1 - 0.2K). Therefore, the temperature variation of the oscillations could be studied. The thermopower measure- ments were extended down to 1.7K. 2. Experimental Procedure All measurements were performed on the Al(Ga) sample (see Table I-l). The measurements were performed using the chopper-amplifier in the field sweep mode (de- scribed hlsections II.A.2 and II.F). This procedure pro- vided a continuous record of the thermopower as a func- tion of the magnitude of the magnetic field, while the sample temperature remained essentially fixed. 3. Data and Analysis 3.1. H parallel to {010} Figure V1.2 displays an example of the raw data; it shows the relative magnitudes of the oscillatory and non-oscillatory (broken line) components, and indicates the precision of the data. In this figure the solid line labeled 8 = 0 represents the recorder output when there is no temperature gradient along the sample; the oscillating curve is uncorrected for the field dependent amplification of the voltage measuring system and also for the slight increase in AT along the sample as H is 117 A m .2: c v I 3 __———-— 1-..---‘Dd-(--.---.~-s o D ., .- ,_~ ‘5 S... U 0) I l 1 I 1 I l I '2 '4 '6 I8 20 fiIkG) Fi ure V1.2: The variation of S with magnetic field at T = 3.55K for the Al(Ga) sample. H is along the [010] axis. The solid line labeled S = 0 represents the x-y re- corder output when there is no applied temperature gradient. The broken line indicates the non-oscillatory component of S. The oscillating curve represents raw data, uncorrected for a field dependent amplification of the measuring system. The magnitude of this correction is indicated by the manner in which the S = 0 curve and the broken line approach each other as H increases. When the data are corrected, these two curves are equidistant at all fields above about llkG to within experimental uncertainty. 118 increased. One can get a feeling for the size of the corrections by noting that after correction, the S = 0 curve and the broken line remained equidistant at all fields above llkG to within experimental accuracy. The oscillations in Figure IV.2 are described by a single frequency, characteristic of the third zone B-orbit of aluminum. This is consistent with all previous measure- ments of oscillatory transport properties.(29'30’34) The oscillation pattern was insensitive to small mis- alignments (~ 1°) of H relative to [010]. In section V.B the amplitude of oscillation in S due to variation of the density of states at the Fermi energy was estimated, and it was concluded that these oscillations should be smaller than the signal noise in- herent in the voltage measuring system. The variation in the S = 0 curve in Figure V1.2 gives a measure of this noise. It is quite clear from this figure that the oscillation amplitude estimate of section V.B is much too small to explain the data. In fact at 15.0kG, the experimental oscillation amplitude was more the four orders of magnitude larger than the estimate given by eq. (v.17).* However, previous oscillatory magneto- (34) resistance data suggest that magnetic breakdown is * For comparison with experiment the value of S (Al) given in eq. (v.17) must be multiplied by four, because there are four independent 8 orbits which contribute to 8. 08¢ 119 present in this direction in aluminum. If we assume that breakdown is also the cause of the magnetothermopower oscillations then we should expect a discrepancy between the calculation of section V.B and experiment; and indeed the discrepancy is large. In section V.A it was noted that all calculations predict for the orbit of interest within the experimental range of temperatures and magnetic fields, that the thermopower oscillation amplitudes should consist of one or more terms of the form of eq. (v.4). If only one term is present or dominant, then the data can be directly analyzed in terms of eq. (v.4). According to this ex- pression, if ln(SH-n) is plotted against l/H for a series of fixed temperatures, straight lines with slopes ankB(T + TD)/B should result. Figures V1.3 - V1.5 show such plots for n = -3/2, 0, +3/2 respectively. The best straight lines are for n = 0 and n = -3/2. Continuing in the spirit of eq. (v.4), plotting these slopes against T should also yield a straight line with slope 2n2kBm*c/efi and intercept at T = OK, ankBTDm*c/eh. Thus, from knowledge of the slope and intercept both m* and TD can be experimentally derived. Figure VI.6 shows such a plot for n = 0; similar plots are obtained for n = t 3/2. Table VI-l contains the values obtained for m* and TD for two perpendicular (but crystallographically equivalent) directions of H. 120 6 ITTIW I H3’ZSOSC/ T x10 (03’? V/Kz) II-— __ T=5.55K T=4.57 K I I I I I I I 0.! 4 6 8 IO H" x105 (0") Figure V1.3: The variation of the quantity H3/ZSOSC/T (plottedcnia logarithmic scale) with H‘1 for a series of temperatures for sample Al(Ga). H was directed along [010]. 121 IO --- - it \ :> ”9 I Tzlo72 K C: c>‘ ‘y can 7:2.43 K T=5.55K T34057K 0,“. 1 1 I I I II 4 6 8 IO H"' x 105 CG") Figure V1.4: The variation of the quantity Sosc/T (plotted on a logarithmic scale) with H'1 for a series of tempera- tures for sample Al(Ga). H was directed along [010]. 122 H-3/2 soSC/T X {0'2 (6-3/2 V/KZ) T=3.55 K 0"4 6 8 IO H" x10“3 (0") Figure V1.5: The variation of the quantity H-B/Zsosc/T (pIOtted on a logarithmic scale) with H'1 for a series of temperatures for sample Al(Ga). H was directed along [010]. 123 1) x 10'“ (G) 11‘H2 IOQISOSCIHII/SOSCIH2IJ/(H 7"(K) Figure VI.6: The variation with temperature of the slopes of the lines shown in Figure V1.4 for H parallel to [010] (circles) and of similar lepes of similar lines for H parallel to [100] (triangles). The upright triangles rep- resent initial measurements with H parallel to [100]. To. evaluate the effects of changes in sample treatment and alignment, after these initial measurements were completed the sample was removed from the holder, its potential leads were removed, it was reannealed, the potential leads were reattached and the sample was returned to the holder. .The resulting data are indicated by inverted triangles. 124 * Table V121: Values of m /me and TD directions of the magnetic field H and for different assumed values of n for Al(Ga) for two orthogonal * n m /me TD(K) H parallel to [010] -3/2 0.096 1 0.01 2.5 i 0.6 0 0.093 1 0.01 1.1 :_0.6 +3/2 0.086 _t 0.02 ~ 0 i 0.6 H parallel to [lOO] -3/2 0.093 : 0.01 4.2 i 0.6 0 0.092 :_0.01 2.4 :_0.6 +3/2 0.095 1 0.01 0.7 i 0.6 We see that the values of m* are the same to within the specified uncertainties for both directions of H and for all values of n in the predicted range, whereas the values of TD are not in agreement. This demonstrates that TD can only be determined for an assumed value of n, but reliable values of m* may be determined without such an assumption. Additional studies showed that m* did not change significantly when the magnetic field was reversed or when the sample was removed from the holder, re-annealed and returned. From Table IV—l it is concluded that m* = (0.93 : .01)me. The only other measurements of m: for the B—orbit of aluminum are de Haas-van Alphen measure- (32) * ments. These give m = (0.102 : .006)me, in agree- ment to within the specified uncertainties with the value 125 derived from thermopower measurements. We might note in * passing that theory predicts the m derived from de Haas- van Alphen data to be phonon-enhanced by about 50% over * (71'72) Therefore, the m derived the bare band mass. from thermopower measurements is similarly enhanced. Finally, the thermopower data shows 1 i 1 5 2, again in accordance with available predictions. 3.2. H along other directions in the (001) plane Magnetic breakdown in aluminum is expected to be greatest for H directed along [010]. In this section data will be presented for H in directions other than [010]. As previously noted, the S oscillation pattern was insensitive to small rotations (~1°) of H from [010]. However, the oscillation amplitude did decrease as H was rotated away from [010]. Figure V1.7 shows that the oscillation amplitude has been reduced to about the size of the non-oscillatory component (at T = 2.4K, H ~ l7kG) when H is rotated 20° from [010] axis in the (001) plane. Figure V1.8 shows that the oscillation amplitude has be- come much smaller than the non-oscillatory component of S when H is directed along [110]. Thus, it may be con- cluded that although magnetic breakdown via the B-orbit exists for H aligned along any direction in (001), the probability of magnetic breakdown decreases as H is rotated away from [010] (i.e. the energy gap between the 126 .Ewummm mCflHSmme on» no :oHumowmemfim usmncmmwo pamflm may now omuomuuooc: .muwm 30H musmmwummu m>uso huoumaawomo one .uswwvmum mus» ImHmmEmv cwfiammm on ma wumnu c033 unmuso Hmnuoomu mix on» mucmmwummu o u m cwamnma mafia endow was .msmam Assoc ms» :0 mflxm Hose. was 2000 com omumuou ms m .0Hm53m Amovad may now xv.m u e um oaoam oapwsmme spas m mo :ofiumwum> one "5.H> musmHm 2331 0m 0. o_. _ A _ A _ fl _ A _ _ a 3...: 8 (Sian 'BHV) 127 .Emummm msfiusmmme on» no coflumoflmemEm unmosmmmo Gamma on» now omuomuuoocs .mumo 3mm muswmmumwu o>uso mew lumaawomo one .usmficmum wusumuwmswu owflammm on ma mums» c033 usmpso Hwnuoomn aux on» musmmmummu o u m owawan mafia wflaom one .mflxm HOHHH on» mGOHm ma m .wamEmm Amwva< mnu new Mm.~ n B as oamflm owumsmme sufi3 m mo sofiumfinm> one "m.H> musmfim 23:: ON 0. o. _ _ i _ _ _ _ _ _ _ S (8.1.an 'BHV) Ouw 128 second and third Brillouin zones increases as H is rotated away from [010]). For a particular orbit, a magnetic field directed in an arbitrary direction in the (010) plane will in gen- eral intersect different extremal cross-sectional areas (normal to H) on the individual third zone "arms" which form the ring of arms. Since the cross-sectional area determines the oscillation frequency, the general oscillation pattern due to a particular orbit will be composed of oscillations of several frequencies. Also the oscillation amplitude is expected to be frequency dependent. Thus, the oscillation pattern for the magnetic field in an arbitrary direction may be quite complicated, as shown in Figure V1.9. Only for H along either [010] or [110] will the frequencies of all B-orbits be equal, so that the oscillation pattern may be described by a single frequency. Figure V1.10 shows frequencies of the major low frequency oscillations obtained from S data. Also shown in Figure V1.10 are the frequencies obtained from de Haas- (32) van Alphen measurements and Ashcroft's predicted fre- quencies for the B-orbits based on a pseudopotential cal- (63) which was fit to de Haas-van Alphen data. culation This figure shows that the frequencies for a particular orbit obtained from either de Haas-van Alphen or magneto- thermopower oscillation measurements agree to within their 129 (ARB. UNITS) S .b (N o 34° Iv, Awfuwe/V”n/\/ J J I I I I L I I I I ICD l5 2() H(KG) Figure 21.9: The variation of S with magnetic field at T = 2.3K for H along a series of directions in the (001) plane of the Al(Ga) sample. 0° corresponds to H directed along [100]. 130 .Ammvumononmm an coflumasoamo Hmwusmuomoosmmm M NO muasmmu map was mv.coouow cam somumq mo sumo cmnmad sm>lmmmm mp map .mumc szom IoEstuoumcmmE ucmmmum uno mamHmmflU musmfim mane .mcmam Aaoov on» ca mamas cowumuou nuwz uflnuonm msoN cues» map mo ooflumm COMDMHHHomo 0:» mo COHOMHHm> "OH.H> mnsmwm 8mg; om on os on o N o_ o 90 a r _ q _ s 8:“. moo; d w I1 Q _ Mu AU .33 [05:3 I m 2088 a 20mm} 9 I .0. on «:3 Fzmmmma a m. U D u .u . . I IN I flu Ilulmn u - 131 mutual uncertainties. It is interesting to note that the magnetothermopower oscillations for 0 between 0° and 20° appear to be due to 81 orbits, in contrast to the de Haas- van Alphen measurements which are sensitive to the 82 orbits in this range of angles.* Thus, magnetothermopower oscillations apparently provide information regarding the shape of the Fermi surface which is complementary to that obtained from de Haas-van Alphen measurements. If one carefully looks at the 46° and 49° oscillation patterns of Figure V1.9, one discerns a small, high frequency component above l7kG. The period of this high frequency component is (3.4 i .3) X 10-76-1. This _ _ ** is in agreement with the value 3.5 x 10 7G 1 obtained (32) from de Haas-van Alphen measurements of the larger y-orbits of the third zone Fermi surface of aluminum (see Figure v.3). We might note in passing that Bozhko and Vol'skiiI7o) have observed magnetoresistance oscillations with a period characteristic of the y orbits of aluminum for H along the [110] axis. * . The geometrical arrangement of the 81 orbit with respect to the 82 orbit is shown in Figure 5 of C.O. Larson and W. L. Gordon, Phys. Rev. 156, 703 (1967). **This frequency is actually due to the y orbit. The geometrical arrangement of the y with reépect to the other orbits is shown in Figure 5 of C.O. Larson and W.L. Gordon, Phys. Rev. 156, 703 (1967). 132 C. Indium 1. Introduction It is apparent from pseudopotential models of the Fermi surface of indium,(52'53'59'73'74) that magnetic breakdown could occur for a sufficiently large magnetic field along the [010] axis. However, until now, no ex- perimental evidence of such breakdown has been obtained. In this section experimental data showing giant, low fre- quency magnetothermopower oscillations are presented. The oscillations are attributed to magnetic breakdown. This attribution is supported by a detailed comparison between the indium data and similar data for the [010] breakdown orbit in aluminum (presented in section VI.B). Aluminum is one of only four metals (aluminum,(29-3l) (75) (69’76) and zinc(47)) for which giant magnesium, tin, magnetothermopower oscillations have been observed. In all four metals, these oscillations have been attributed to magnetic breakdown. In each case, this attribution is supported by independent magnetoresistance(34'77'78) de Haas-van Alphen(66,79) measurements and also by analysis of the known Fermi surface of the metal. Aluminum was chosen for the comparison because its electronic structure and Fermi surface are most nearly similar to those of indium. In addition to the giant oscillations for H along [010], large, low frequency magnetothermopower oscillations 133 for H along [101] and [110], and small, medium frequency oscillations for H along [110] were also observed. These low frequency oscillations are also tentatively attributed to magnetic breakdown, but the medium frequency ones are not. 2. Experimental Procedure Oscillatory behavior was found in samples In-II, In(Sn)-I, In(Sn)-II, and In(Sn)—III. As the temperature was raised the oscillation amplitude fell off rapidly, becoming comparable to the measuring uncertainty at about 2.5K. Searches for oscillatory magnetothermopower were not performed on samples In-I and In(Ga); presumably these samples would have exhibited this behavior when H was aligned along the proper crystallOgraphic axis. The measurements were performed using the chopper- amplifier in the field sweep mode (see sections II.A.2 and II.F). 3. Data and Analysis Figure V1.11 shows the output of the x-y recorder (essentially a plot of S vs H) for samples In(Sn)-I and In(Sn)-II for H along [101], [110], and [010]. No oscillations were observed for H along [001]. In Figure V1.11 four sets of oscillations are present, each periodic in l/H. The oscillation frequencies are 1.4 i 0.1 x 105G for H parallel to [101]; 1.55 i 0.1 x 1056 and 6 4.6 i 0.1 x 10 G for H parallel to [110]; and 134 HI) [1011 T=1.68K { Hutfl'OJ >: v T=1.98K Q 2 03 HIIEOlOJ T=1.67 K I I I' 10 15 20 H (kGJ Figure V1.11: The magnetic field dependence of S for H aligned along the indicated crystallographic axes. The lines labeled S = 0 represent the recorder output when there is no applied temperature gradient. The oscillating curves represent raw data, uncorrected for the field de- pendent amplification of the measuring system. The ampli- fication decreases by about 30% gver the field range shown. The curves for H aligned along [101] and [019] are from sample In(Sn)-I and the curve for H along [110] is from sample In(Sn)-II. 135 1.04 i .04 x 105G for H parallel to [010]. All four frequencies agree to within mutual uncertainties with values obtained for indium from de Haas-van AlphenISZ'Bo) and ultrasonic absorptionIBl) measurements; measurements which can be understood without invoking magnetic break- down. In an effort to see whether the oscillation amplitudes were consistent with eq. (v.4), graphs of ln(Hn SOSC/T) vs l/H with n = 0,1 were plotted for H along the [010] axis. Figure V1.12 shows a typical graph. The data clearly diverge from a straight line. Thus, in contrast to aluminum, the oscillation amplitudes in indium are not consistent with eq. (V.4). The relatively small signal to noise ratio for H along either [101] or [110] precluded any similar oscillation amplitude analysis for these directions. Samples In-II and 1n(Sn)-III also exhibited oscillations for H directed along the [010] and [110] axes respectively. For H along the [010] direction, the data for the purer sample In-II were similar to those for 1n(Sn)-1, but the oscillation amplitude was about 25% smaller for a field of about lSkG. at 1.7K. One also obtains a similar comparison with the data of the purer In(Sn)-II with those of In(Sn)-III for a magnetic field of about lSkG. directed along the [110] axis at 1.9K. Therefore, it is concluded that the oscillation 136 50M (gt 9 \x o > 9 I0— ¢ )— 9 + x r- F— I- 6‘ ‘8 5— w I _ ' I I I I I 4 6 8 IO l/H x no5 (0") Figure V1.12: The variation of the quantity HSosc/T (plotted on a logarithmic scale) with l/H for sample In(Sn)-I at T = l.67K. H was directed along [010]. '— fl,“ .m. 1.4 137 amplitudes were limited by extended crystal lattice imperfections (mosaic structure, lattice dislocations, etc.) rather than by impurity scattering. Such a result is reasonable since indium is a very soft material which can be easily damaged. Like aluminum, the oscillation pattern for these indium crystals were not significantly affected by 1° rotations of H relative to the major symmetry directions; and similarly the oscillation amplitudes gradually de- creased as H was rotated away from the major symmetry directions. The remainder of this section will concentrate on the giant oscillations for H along [010], both because these are the largest oscillations and because they can be compared directly with breakdown oscillations in aluminum for the same field orientation. As noted earlier, electronic transport oscilla- tions are expected to arise from either magnetic break- down or oscillations in the density of states at SF. Clearly, the most convincing proof that the giant thermo- power oscillations which are observed are a result of magnetic breakdown would be a demonstration that their amplitude could be predicted from theory. Unfortunately, the theory of the effect of magnetic breakdown on the thermopower of metals is not sufficiently far advanced to allow direct quantitative comparison with experiment. 138 Therefore, the breakdown hypothesis was indirectly tested in two ways involving comparison of the indium data with similar data for the [010] breakdown orbit in aluminum. First, the estimation of the oscillation amplitude due to the other mechanism which might cause oscillations: the variation with H of the density of states at the Fermi energy in the absence of magnetic breakdown given by eqs. (V.l7) and (v.18), was compared with the experi- mentally measured amplitudes for both aluminum and indium. L”fl:.-’rl€.";m 5' . . ' .4 -.. For aluminum the calculated oscillation amplitude for H ——- along [010] was four orders of magnitude smaller than the experimental amplitude, and for indium it was three orders of magnitude smaller. Since the oscillations in aluminum result from breakdown, a discrepancy between this calculation and experiment is expected and indeed it is very large. The fact that indium also exhibits a very large discrepancy is taken as evidence in favor of magnetic breakdown in indium. The large oscillations for H parallel to [101] and [110] are also two or more orders of magnitude larger than the calculated value of eq. (v.18),* 2 * * In eq. v.18, the values of |§—%| and m appropriate to H 3k H along [010] was used. Proper calculations would use the * - _. 3—%| and m appropriate to H along either [101] or [110]. 3k H However, it seems unlikely that these changes will signif- icantly reduce the discrepancy between calculation and ex- periment. 139 making them candidates for breakdown too. On the other hand, the small medium frequency oscillations for H parallel to [110] are within one order of magnitude of the value predicted by this simple model, and it is there- fore unnecessary to invoke breakdown to explain them. Second, the angular variation of the thermo- powers of the two indium samples was compared with that of the aluminum single crystal as H was rotated away from the [010] direction. Agreement between the forms of these two angular variations might provide additional evidence for magnetic breakdown in indium. The angular variations are shown in Figure V1.13. The field was rotated around the [001] axis of the aluminum sample and of sample In-II, and around the [101] axis of sample 1n(Sn)-I. Indeed, the forms of the rotation patterns are nearly the same for both aluminum and indium. Similar patterns were also obtained for rotations with H near [101] and [110]. Having shown that the low frequency data can be plausibly attributed to magnetic breakdown, this section will be concluded by considering the ramifications of the presence of such breakdown for the most recent pseudopotential model of the Fermi surface of indium by (59) Holtham. For H along [010], this model yields a breakdown field H0 in the range of 200-400kG,(S7) more than an order of magnitude larger than the largest field 8 —— 6 - AI. T: 2.26K 4 — ROT. AXIS-[001] 2 __ O _ 4... _ In A 2 T21.76 K x O __ ROT.AXIS-[001] \ 3 ‘6 -2r- 2' 0) 6h In 4 1.. 131.71 K ROT. AXIS-[101] 2 __ 0 _— _2 _— I I I I I I I -60-40-20 0 20 4O 60 £010] 9 (DEG) Figure V1.13: Variation of S with rotation angle measured from the [010] axis, at H = lSkG for aluminum and indium single crystals. The aluminum data are from sample Al(Ga). The indium data for H rotated about the [001] axis are fronI sample In-II andtfluadata for H rotated about the [101] axis are from sample In(Sn)-I. 141 used here. In view of the exponential dependence of break- down probability on the ratio H/Ho,(64) it seems unlikely that the experiments performed in this thesis would show significant effects of breakdown for such a large HO. The experiments suggest that H0 is roughly an order of magnitude smaller. As Holtham points out,(59) this would imply that the second and third zones are closer together than in his model. In conclusion, large, low frequency oscillations have been found for H along the [010], [101], and [110] axes, while small, medium frequency oscillations are found for H along the [110] axis. All frequencies agree to within mutual uncertainties with the values obtained by de Haas-van Alphen and ultrasonic absorption measure- ments on indium. Arguing by analogy with aluminum, a metal for which there is independent evidence that magnetic breakdown exists in directions where there are giant magnetothermopower oscillations, the large, low fre- quency oscillations are attributed to magnetic breakdown. This attribution implies that the energy gap between the second and third zones is smaller than predicted by the latest pseudopotential model of indium.(59) To confirm the existence of magnetic breakdown in indium, it is suggested that a search for its effects on other properties be begun. Because of the apparent sensitivity of magnetothermopower oscillations to 142 (29) breakdown, other properties are likely to require both magnetic fields well above 20kG. and temperatures below 2.5K. This dual requirement is the probable rea- son for the lack of prior evidence of magnetic breakdown in indium, since measurements of less sensitive properties such as magnetoresistence have only been made at either high fields or comparably low temperatures, but not both at once. REFERENCES 10. 11. 12. 13. 14. LIST OF REFERENCES J.M. Ziman, Electrons and Phonons, (Oxford Univer- sity Press, London, 1960). see for example K.R. Symon, Mechanics, (Addison- Wesley Publishing Co., Inc., Reading, Massachusetts, 1960), 2nd ed., p. 408 ff. R.S. Averback, Ph.D. Thesis, Michigan State Univer- sity, 1971 (unpublished). R.S. Averback, C.H. Stephan, and J. Bass, J. Low Temp. Phys. i2, 319 (1973). R.S. Averback and D.K. Wagner, Sol. State Comm. ii, 1109 (1972). R.E. Prange and L.P. Kadanoff, Phys. Rev. 134, A566 (1964). R.S. Sorbello, J. Phys. F: Metal Phys. 4, 1665 (1974). J.L. Opsal and D.K. Wagner, J. Phys. F: Metal Phys. 6, 2323 (1976). H.A. Fairbank and J. Wilks, Proc. Roy. Soc. (London) A231, 545 (1955). J.A. Rayne, Aust. J. Phys. 9, 189 (1956). A.C. Anderson, Rev. Sci. Instr. 39, 605 (1968). T.J. Tiedema, Acta Crystallogr. 2, 261 (1949). G.K. Williamson and R.E. Smallman, Acta Met. i, 487 (1953). J. Bass and J.L. Opsal, Proceedings of European Studpronference on the Transport Properties of Normal Metals and AlIOys We11 Below 6D (to be published). 143 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 144 C. Kittel, Introduction to Solid State Physicg (Wiley, Inc., New York, 1971), 4th ed., p. 210-1. A.A. Abrikosov, Introduction to the Theory of Normal Metals (Academic Press, New York, 1972). J.L. Opsal, B.J. Thaler, and J. Bass, Phys. Rev. Lett. 33, 1211 (1976). S.K. Lyo, Phys. Rev. Lett. 39, 363 (1977). G. Grimvall, Physica Scripta (Sweden) 32, 63 (1976). T. Holstein, Ann. Phys. 39, 410 (1964). D.K.C. MacDonald, Thermoelectricity: an introduc- tion to the Principles (Wiley, Inc., New YorkT' 1962). 1.1. Hanna and E.H. Sondheimer, Proc. Roy. Soc. (London) A239, 247 (1957). R.D. Barnard, Thermoelectricity in Metals and Alloys (Taylor and Francis Ltd., London, 1972), p. 132. M. Bailyn, Phys. Rev. 157, 480 (1967). J.M. Ziman, Adv. Phys. 39, l (1961). R.J. Gripshover, J.B. Van Zytveld, and J. Bass, Phys. Rev. 163, 598 (1967). R.S. Blewer, N.H. Zebouni, and C.G. Genier, Phys. Rev. 174, 700 (1968). J.L. Opsal, to be published. W. Kesternich and C. Papastaikoudis, Phys. Stat. Solidi 33, K41 (1974). N.N. Sirota, v.1. Gostishchev, and A.A. Drozd Dokl. Akad. Nauk, SSSR 220, 818 (1975). (English trans- lation: Sov. Phys.-Dokl. 39, 116 (1975).) B.J. Thaler and J. Bass, J. Phys. F: Metal Phys. ‘3, 1554 (1975). c.0. Larson and W.L. Gordon, Phys. Rev. 156, 703 (1967). D.A. Dicke and B.A. Green, Phys. Rev. 153 , 800 (1967). 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 145 R.J. Blacombe and R.A. Parker, Phil. Mag. 33, 533 (1970). R. Fletcher, J.L. Opsal, B.J. Thaler, J. Phys. F: Metal Phys. (in press). W.B. Willot, Phil. Mag. _1__6_, 691 (1967). P.E. Nielson and P.L. Taylor, Phys. Rev. Lett. 33, 893 (1968). A.D. Caplin, C.K. Chiang, J. Tracy, and P.A. Schroeder, Phys. Stat. Solidi 26A, 497 (1974). R.S. Averback and J. Bass, Phys. Rev. Lett. 33, 882 (1971). R.S. Averback and J. Bass, Phys. Rev. Lett. 33, 631 (1975). F.J. Blatt, A. Burmester, and B. LaRoy, Phys. Rev. 155, 611 (1967). A.R. DeVroomen, C. Van Baarle, A.J. Cuelenaere, Physica (Utrecht) 33, 19 (1960). N.E. Adams and T.D. Holstein, J. Phys. Chem. Solids 32, 254 (1959). see for example J.M. Ziman, Principles of the Theory of Solidsd(Cambridge University Press, New York, 1972), 2“ ed. G.E. Zil'bermann, Zh. Exp. Teor. Fiz, 39, 762 (1955). (English translation: Sov. Phys.-JETP 3, 650 (1956).) P.B. Horton, Ph.D. Thesis, Louisiana State Univer- sity, 1964 (unpublished). C.G. Grenier, J.M. Reynolds, N.H. Zebouni, Phys. Rev. 129, 1088 (1963). J.R. Long, C.G. Grenier, and J.M. Reynolds, Phys. Rev. 140, A187 (1965). A.H. Wilson, Theory of Metals (Cambridge University Press, New York,’I953), 2nd ed. B.J. Thaler, R. Fletcher, and J. Bass, J. Phys. F: Metal Phys. (in press). wuss-2...: -- mun '1‘! I 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 146 M.G. Priestley, Phil. Mag. _7_, 1205 (1962). A.J. Hughes and J.P.G. Shepherd, J. Phys. C: Solid St. Phys. 3, 661 (1969). J.H.P. vanWeeren and J.R. Anderson, J. Phys. F: Metal Phys. 3, 2109 (1973). I.M. Lifshitz and A.M. Kosevich, Zh. Exp. Teor. Fiz. 39, 730 (1955). (English translation: Sov. Phys.-JETP 3, 636 (1956).) A.V. Gold, Solid State Physics: Electrons in Metals, Ed. J.F. Cochran and R.R. Haering (Gordon and Breach, New York, 1968), v.1, p. 39. R.B. Dingle, Proc. Roy. Soc. (London) A211, 517 (1952). P.M. Holtham, private communication. This is an estimate (expected to be slightly high) based on Holtham's calculated value for m*. 59) Holtham's calculation neglects the electron-phonon mass enhancement. P.M. Holtham, J. Phys. F: Metal Phys. 3, 1457 (1976). R.W. Stark and L.M. Falicov, Progress in Low Tem- perature Physics, Ed. C.J. Gorter (North—Holland Publishing Co., Amsterdam, 1967), v. 5, p. 235. J.M. Ziman, Principles of the Theory of Solids (Cambridge University Press, New York, 1972), 2nd ed., p. 327. J.C. Abele and F.J. Blatt, Phys. Rev. 33, 1298, (1970). N.W. Ashcroft, Phi1. Mag. 9, 2055 (1963). A.B. Pippard, Proc. Roy. Soc. (London) A287, 165 (1965). L.M. Falicov and H. Stachowiak, Phys. Rev. 147, 505 (1966). M.G. Priestly, L.M. Falicov, and G. Weisz, Phys. Rev. 131, 617 (1963). A.B. Pippard, Phil. Trans. Roy. Soc. (London) A256, 317 (1964). 147 68. R.C. Young, J. Phys. C: Solid St. Phys. 3, 458 (1971). 69. R.C. Young, J. Phys. F: Metal Phys. 3, 721 (1973). 70. v.1. Bozhko and E.P. Vol'skii, Zh. Exp. Teor. Fiz. 39, 703 (1974). (English translation: JETP Lett. 39, 325 (1974).) 71. N.W. Ashcroft and J.W. Wilkins, Phys. Lett. 33, 285 (1965). 72. A.B. Meador and W.E. Lawrence, Phys. Rev. 815, 1850 (1977). 73. N.W. Ashcroft and W.E. Lawrence, Phys. Rev. 175, 938 (1968). 74. W.J. O'Sullivan, J.B. Schirber, and J.R. Anderson, Phys. Lett. 27A, 144 (1968). 75. C. Papastaikoudis and W. Kesternich, J. Low Temp. Phys. 33, 517 (1975). 76. J.A. Woolam and P.A. Schroeder, Phys. Rev. Lett. 33, 81 (1968). 77. R.W. Stark, Proc. 1Xth Int. Conf. Low Temoeratugg Physics, Ohio 1964 (Plenum Press, New York, 1965), p. 712. 78. R.C. Young, J. Phys. C: Solid St. Phys. 3, 474 (1974). 79. F.A. Buot, P.L. Li, and J.O. Strom-Olsen, J. Low Temp. Phys. 33, 535 (1976). 80. 6.8. Brandt and J.A. Rayne, Phys. Rev. 132, 1512 (1963). 81. J.E. Cowey, R. Gerber, and L. Mackinnon, J. Phys. F: Metal Phys. 3, 39 (1974). 82. R.S. Sorbello, private communication. 83. E.M. Purcell, Electricity and Magnetism (McGraw- Hill, New York, 1964), p. 55. APPENDICES APPENDIX I CALCULATION or ANISOTROPY OF 1(1) FOR Al(Ga) We will use Sorbello's approximate phase shift model(7) to calculate the anisotropy of 1(k) of Al(Ga). Within the Born approximation (52 << n) the average value of l/To(k) around an orbit on the Fermi surface, ,may be written (D = gficg z (22 + l)(6:A)2 . (AI.1) ‘F 0 i=0 SEA are the phase shifts obtained from the Born approxima- tion, and depend only on the impurity scattering potential. The quantity F2(k), called the "amplitude factor"’measures the square of the magnitude of the 1 component of ¢K(;)(¢k(;) is the pseudowavefunction associated with the Bloch state k) when ¢k(;) is expanded in spherical harmonics about a lattice site. In the free electron (P.E.) l for all 3. III approximation F£(k) Using the 63A obtained from the Heine-Abarenkov- Animalu form factors in the Born approximation, which are given in Sorbello's table 1; and the average amplitude factors for orbits on the Fermi surface of 148 149 aluminum, which are given in Sorbello's table 3, we can calculate the ratio <1/Io(k)>/(1/To) for Al(Ga). P.E. Table AI-l shows the results for the third zone y, a, B and orbits and second zone [110] orbit on the Fermi surface of aluminum. Table AI-l: Average reciprocal relaxation times for Al(Ga) obtained from Sorbello's approx- imate phase shift model 7 ORBIT /(l/TO)F.E. 1.10 1.10 1.19 1.08 Table AI—l shows that the average reciprocal relaxation times of the third zone of Al(Ga) are isotrOpic to within 10%. Sorbello also claims that the average reciprocal relaxation times of the second zone should be more isotropic than the third zone times.(82) There- fore from table A1-1 it is concluded that the average relaxation time is nearly isotropic (to within an accuracy of about 10%) over the entire Fermi surface (which only consists of second and third zone pieces) of dilute Al(Ga) alloys. APPENDIX II 30(8) DETERMINATION OF 38 From eq. (III.3la) we see that 2 v. e f dS —$ T v 0"“) S(e) v j' 13 4flzfi which may be rewritten as 2 _ e E Oij‘e’ ' ggzg [S(e) as v - a. T v., (AII.1) A . . . .th where ei is a unit vector in the l (83) direction. Apply- ing the divergence theorem to eq. (AII.1) gives _ e 3+ . A e 8 ds A = d ' ' -—— . . . 0 AI 02 4fl2h f0 8 15,8 ) hv 9k (e1 1 v3) ( I ) So it follows that 30..(€) 2 1 _ e dS . A 38 _ 4fl3fi fS(e) hv 6k (ei T vj)’ (AII‘3) We recognize that V - (ei T vj) = [$(t v)]ij so 30..(e) 2 13 = e dS + + 150 151 For cubic symmetry eq. (AII.4) becomes 2 30(8) e dS + 38 12655