BRYMWCS QF LEEBEUM 1R GERMAMUM Mil} COBQPENSATEB SILICON Thesis ‘30:? “no Degree 0% pk. D. WCMGW SEME BNE‘JERSETY Eric AEexanéer Bounce i972 This is to certify that the thesis entitled DYNAMICS 0‘5‘ LITHIUM I'\l GER‘JEAVIUl-l A‘ID COTx’EPE‘ISATED LITHIUM presented by ERIC ALEXANDER DC‘IWCE has been accepted towards fulfillment of the requirements for Ph. D. Jame in E. E. fl; w/f/X/I Major professor Date 2/10/72 0-7639 . s . "‘51-.'- H~"~m. 'fl-O—o '.l-‘—. 0- . no. r \ f‘. --. u ‘ I A-‘.,.A'_ : J---Lv‘. -‘ u ‘, A- ...A “Y" m. but: . .4.,..,.~ ‘P‘: "“"z 5..- u ' . a ‘ 1-‘-...,, V. “~.u-u. .. . ' .--.~."~ O‘- p. u-‘M-‘—H f.‘ ‘ 1 . . A‘ “ ‘hi V ‘L.A-‘. t... . .i‘. ‘A‘ a“ ‘-..‘D-V.‘ c" ‘. .~ ‘ ”-3 “A A ‘— §u_‘u " A .- n’fl ‘Pflfi \m..r._ .v‘ A I e... “\‘A“ “VA . “‘5, l ' ~ . v +“n- w u v‘ Q . ~ "“\1““‘ A- ‘R s.‘ aha C" . ‘v ‘ ‘ ll ~A‘ ‘ .H Y‘ ‘ Wu? ,3 .. “‘ - ~ “ n. e vu‘ afl§~ ‘» V“). A -_ ‘V‘~“. .- “r-L a»- ‘ u. . A I “y“: “ « S (\_‘ ~ V ABSTRACT DYNAMICS OF LITHIUM IN GERMANIUM AND COMPENSATED SILICON By Eric Alexander Dounce A study of the dynamics of lithium in germanium and silicon is important in helping to establish practical limits on the drift rate of lithium ions in silicons or germanium during the fabrication of nuclear detectors. The systems of lithium in germanium and lithium in silicon have been studied with this goal in mind. IsotOpic diffusion studies of lithium in germanium have been made at relatively low diffusion temperatures, 3.3., 2OD-HOOOC. Measurements of the ratio of 6Li to 7 Li diffusion constants provide evi— dence for the significance of quantum effects in the dif- fusion process at these temperatures. We have proposed a tunneling mechanism as the probable explanation and we compare our measurements with the results predicted by a general diffusion theory which includes tunneling effects. Our results, when extrapolated to detector drift temperatures, show that the lighter 6Li isotOpe drifts only 12% faster than 7Li. ii A? f‘ b. V 41": d ' .‘* -5 b ‘F .d A h C1 , ,.nv I .V'.‘ nrs -v ~ WIN .3“ - . -po- ,4 -«.- 0-- ;",F.: .A't - - awe-:O \r-‘ u - “\II‘ «0- 9‘: \‘ V—l-‘s... -‘ «s !A*.. .l \ . g A a 'v \“H‘ _ ‘-‘ F! v -. , ‘ . ‘ . a I fi‘~v~- 3”. EV "‘ ...g .4- ‘ Eric Alexander Dounce The detector fabrication process requires drifting lithium ions into the silicon or germanium to form pairs or complexes with negatively charged trap sites. The dynamics of the lithium-acceptor complexes then becomes important in fully understanding the compensation process. We have found the present model describing pair dynamics in a crystal lattice to be inadequate. A modification of the existing model is proposed which takes into account the actual C3V symmetry of the complex. The possibility of interactions between different local modes at the same physical site is also considered. Our analysis is, however, only approximate and a more rigorous formalism is indicated by comparison with measurements of line width and centroid variations with temperature. These measurements have allow- ed us to determine the coefficients of anharmonic terms in the local-mode potential. If a local mode of free lithium exists in silicon or germanium it is important to locate its frequency so that it may be excited directly. We did not find a local mode of free lithium in silicon and expect that the free lithium exhibits only resonant modes in silicon. A review of the literature indicates a probable free lithium local mode in germanium near 36H cm"1 for 7Li. Finally, the effects of uniaxial stress on the lithium- boron local modes in silicon have been studied. We observed that uniaxial stress near the fracture limit did not produce a detectable splitting of the two-fold degenerate lithium iii Lric Alexander Bounce or boron local modes. Similarly, it did not produce a measurable shift in the local-mode centroids. It is un- likely then, that uniaxial stress will produce significant variations in the lithium drift rate in silicon. iv DYNAMICS OF LITHIUM IN GERMANIUM AND COMPENSATED SILICON Bv Eric Alexander Dounce A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Electrical Engineering 1972 Q“; ~‘ ‘.. ACKNOWLEDGMENTS I would like to thank: Dr. William Hartmann for his help with the theoretical interpretations, my committee co-chairman Dr. David Fisher for his critical reviews of the manuscript, Dr. Gary Cloud for his help with the bi- refringence work, and to Mrs. Marcella Williams for typing the final draft. A special thank you is given to: my thesis advisor and committee co—chairman Dr. Charles R. Cruhn for his direction, help, and enthusiasm with the experimental phase of this project, Dr. D. J. Montgomery for his de- tailed evaluation of the manuscript on such short notice, and to Dr. G. E. Leroi for the donation of the FEE—112 spectrOphotometer and for the use of the FEE—225 spectro— photometer. Finally, I would like to thank my wife for typing the first drafts and for helping to make this project more meaningful. - 0" -qn'o A4 3". I." AI“ H.434 .11 TABLE OF CONTENTS Chapter Page 1 INTRODUCTION . . . . . . . . . . . . . . . . l 2 THEORETICAL HISTORY. . . . . . . . . . . . . 5 2.1 Lithium in Germanium - Diffusion Studies . . 5 2.2 The Boron Local Mode in Silicon. . . . . . . 9 2.3 The Lithium Local Mode in Silicon. . . . . . 10 2.u Lithium-Boron Complexes. . . . . . . . . . . 11 2.5 Anharmonic Effects on a Td Symmetric Local Mode . . . . . . . . . . . . . . . . . . . . 15 2.6 Experimental Difficulties in the Observation of Local Modes . . . . . . . . . . . . . . . 19 2.7 Complexing Donors with Acceptor Ions . . . . 22 3 SAMPLE PREPARATION . . . . . . . . . . . . . 26 3.1 Compensation Methods . . . . . . . . . . . . 26 3.2 Material, Mechanics and Preparation. . . . . 28 3.3 The Lithium Deposition . . . . . . . . . . . 29 3.u Diffusion and Lithium Solubility . . . . . . 31 3.5 The Hot-Point Probe. . . . . . . . . . . . . 38 3.6 Precipitation Kinetics . . . . . . . . . . . 39 u INSTRUMENTATION. . . . . . . . . . . . . . . H8 H.l The Spectrophotometer. . . . . . . . . . . . U8 u.1.1 The Fore Prism . . . . . . . . . . . . . . . MB u.1.2 The Source . . . . . . . . . . . . . . . . . Hg vi 1 ‘9-.v .4» «.1. n.4w TABLE OF CONTENTS (continued) The Grating Monochromator. . . . . The Infrared Detector. . . . . . . . . Double Beam Modification . . The Measurement System . . . . . . The Lock-in Amplifier. . . . . . . . . . LOCAL-MODE EXPERIMENTS AND EXPERIMENTAL RESULTS The Germanium System . . . . . . . . . The Silicon Svstem . . . . . . . . . . The Temperature Dependence Measurements. The Pressure Dependence Measurements . . DISCUSSION OF LOCAL MODES. . . . . . . Local—Mode Assignments in Silicon. . . . Lithium—Gallium Complexes in Germanium . Temperature Dependences of Centroids and Line Widths in Silicon . . . . . . . . Temperature Dependence of Integrated Intensities. . . . . . . . . . . . . . The Uniaxial-Stress Results. . . . . ISOTOPIC DIFFUSION STUDIES . . . . . . . Reasons for Measuring the Mobility Ratio The Ratio Measurement. . . . . . . Comparison of Experiment with Tunneling Calculations . . . . . . . . . . . . . SUMMARY AND CONCLUSIONS. . . . . . . . . BIBLIOGRAPHY . . . . . . . . . . . . . . APPENDIX A . . . . . . . . . . . . . . APPENDIX B . . . . . . . . . . . . . APPENDIX C . . . . . . . . . . . . . . A.PT)EP‘IDIX D O O O O O O O O O O O O O O 0 APPENDIX E . . . . . . . . . . . . . . . 57 6O 60 64 67 67 7O 7O 87 CO CO 96 98 99 101 107 111 11A 117 120 122 128 IBM Table II III IV LIST OF TABLES Line—Width Coefficients Experimental Results of Isotopic Studies Labeling of Local Modes. Phonon Spectrum. viii for Two-Phonon Decay Page 92 103 132 13H -I- .1 . ‘1: at. LIST OF FIGURES F'gure Page 2.1 Interstitial Mass M vs. Relative Local—Mode Frequency. . . . . . . . . . . . . . . . . . . . 12 3.1 The Evaporation System . . . . . . . . . . . . . 30 u.1 The Spectrophotometer. . . . . . . . . . . . . . 50 u.2 The Grating Equation . . . . . . . . . . . . . . 5H H.3 Indene Calibration of Modified Spectrophotometer 58 u.u Double-Beam Modification . . . . . . . . . . . . 61 U.5 The Overall Measurement System . . . . . . . . . 63 u.6 The Phase Sensitive Amplifier. . . . . . . . . . 65 5.1 Results of I-R Absorption Measurements in Germanium. . . . . . . . . . . . . . . . . . . . 69 5.2 The Heating Apparatus. . . . . . . . . . . . . . 71 5.3 Absorption Spectra at Different Temperatures . . 73 5.u Centroid Shifts vs. Temperature. . . . . . . . . 7H 5.5 Line Width vs. Temperature . . . . . . . . . . . 75 5.6 Time vs. Sample Temperature. . . . . . . . . . . 78 5.7 Relative Integrated Absorption Intensity vs. Temperature. . . . . . . . . . . . . . . . . . . 80 5.8 The Pressure Applicator. . . . . . . . . . . . . 81 5.9 Calibrations of Pressure Gauge (18%) . . . . . . 83 5.10 Line Width vs. Pressure. . . . . . . . . . . . . 8H 7.1 Tunneling Calculations Compared with Experi- mental Migh and Low Temperature Data . . . . . . 109 7.2 Ratio vs. Local-Mode Frequency . . . . . . . . . 110 ix B.1 C.1 D.1 LIST OF FIGURES (continued) Fraction of Paired Ions (6) vs. Temperature for Equal Concentrations of Lithium and Boron of lOlg/cm3 vs. Temperature. . . . . . . . . . 121 Birefringence Measurement Apparatus. . . . . . 123 The 6-Atom Model Used As A Basis for Lithium— Boron Local-Mode Calculations. . . . . . . . . 129 a fi'In". “up- 0 II‘ f‘f‘rl .._A\'V‘ .a ‘F‘ ,v I~ — .4 1 It . r” I.“ A; nu ‘x «3+ a.» r“ ‘ AU AV E. .. C S .1 C .1 C V. S R C c i A... T. 1.. F . us; 9. .1; Wu ”1.. M. u.. s « 0 us .s C. ‘N‘ $4.5 $. «3 . w. %. Li; L u L .. O; sflu ens ab 4“ *L 3 D. C T .i 2 a m... V... e e e E a D. :2 t 2: ”a n.“ o. A C A. :. ”a .l S a w. e 1‘ T‘ P. .J 1. .. C a: .1 a. .J G. .7. S .L C .3 .nn 1: L.” 1% .7... 8 Lu 3“ ~ k .1. a. 5.. as ..... K“ . . 13.: D... :u .Hm . . 7; A... Y.» C... . ‘Q :s a . o . a . .. m u. . .2 .~ I ~ .. n~.. EC”. "m I.” "P. :F‘. N.“ .salum. «awe» a: - CHAPTER 1 INTRODUCTION The work described in this thesis began with the snpecific goal of establishing limits on the practicality (3f increasing the diffusion rate for lithium ions in a senniconductor matrix. Two questions were fundamental to 'this goal: can some means of selectively heating the ilithium.ion in the semiconductor lattice be found? Or, iJi a more direct approach, is an advantage gained if a jlighter isotope of lithium is used? Our interest in the motion of the lithium ion in a senniconductor matrix was initially generated by the role gilayed by lithium in the fabrication of nuclear detectors. 'This fabrication involves a low-temperature drift of lgithium ions through silicon or germanium to compensate ruegatively-charged trap sites. Low drift temperatures Eire used so as to reduce the precipitation of lithium on riucleation sites. This drifting process consequently irequires considerable time and is limited by the low- ‘tenmerature diffusion constant. Therefore, any increase iJl the diffusion rate of lithium ions may be directly 'translated into a more efficient detector fabrication. In an effort to answer the question on selectively lqeating the lithium ion in the semiconductor lattice, we sset out to study the local—vibrational-modes of lithium jjl germanium and silicon by means of infrared absorption. 1 2 The introduction of lithium into silicon or germanium results in approximately equal concentrations of free electrons and lithium local modes. The free electrons have a much greater absorption coefficient than the local- modes and, therefore, mask them. The systems actually studied were electrically compensated with equal concen- trations of acceptors and donor ions. The acceptors and donors generally form pairs which tend to complicate the local-mode spectrum. Of these systems, lithium and boron in silicon has been the most extensively studied.l"10 However, relatively little has been done to evaluate the effects of lattice coupling on these local modes. Lattice coupling to the local modes is, however, fundamental to the question of selectively heating the lithium ion in a semiconductor lattice. The measurement of second harmonics by Waldner10 does give an indicator of lattice coupling effects; but, no indication of the lithium-boron pair dynamics is evident. We have, therefore, set out to learn about how the lithium ion reacts with its surroundings through measurements of the centroid shifts and line widths of the local modes as a function of temperature. The spectrum of lithium-boron local modes in silicon has one uncertain feature; that is, the designation of the local mode at 522 cm_1 as a singly or doubly-degenerate local mode of lithium in a paired configuration. It is generally thought to be a doubly-degenerate local mode 3a,3b,9a,9b of lithium paired with boron. However, Hayes8 ‘ ."l 1“- ..- ' m L1 had alrea: - ' l- . U5 ’ V ' Hui. ‘ an; Lassical ratj . V no ‘R ». kJe high tE .L. ratio of ditc. 5575,31 ‘ Lacie woul Her .92, use . V- In an 95's rump e .‘saL‘Vy‘ 4*e ”o l_\'. :31» I; 3 in a review article raises some questions as to its designation, and the application of uniaxial stress is 8 to determine the degeneracy of this mode. suggested7’ Since we are primarily interested in local modes due to the lithium motion, we set out to determine the degeneracy of the 522 cm"1 mode through the application of uniaxial stress. Prior to our infrared absorption studies, our work involved the more direct approach to the problem of an increased diffusion rate; we studied the difference in diffusion constant (or mobility) between 6Li and 7Li isotopes. The ratio of diffusion constants of 6Li to 7Li had already been measured in silicon at 800°C by 11 . . . His measurement showed no variation from the Pell. classical ratio. This result is not unexpected in view of the high temperature of measurement. The classical ratio of diffusion constants is only 1.080, and little advantage would be gained by drifting detectors with 6Li instead of 7Li. But for such light ions, the possibility of tunneling exists which may substantially increase this ratio at the low drift temperature involved. Evidence supporting this possibility was available in measurements of the low-temperature diffusion constant of 7Li in ger- manium by Sher}2 These measurements showed about a 15% decrease in activation energy in the range of the low- temperature measurement; such an effect might be accounted for by a tunneling mechanism. The parameter of greatest , CF DHV‘Q __,‘v. \v u ,1 rs so 0 1 n neafisui nfiODY‘. ‘ .:‘ a la- k-‘ 34““ 1'”2 on. O a.» *fi‘lp; .- u.... -‘s “3"‘1‘ 1“ .qu-LSO - s‘ ° n55“ ins rs“ New e " . .‘yu‘ ,:....‘.' C." ..u\.-\.3 are u uncertainty in subsequent diffusion-ratio calculation turned out to be the lithium local-mode frequency; this uncertainty further motivated our infrared absorption studies which followed. In order to make the account of these experiments more meaningful, the details of the necessary background material will first be presented. The sample preparation and instrumentation are then described, followed by a description of the local-mode experiments and their results. These results are then discussed in terms of theoretical calculations. Finally, the isotopic diffusion studies are presented and discussed. an' A: an. O‘QFI.‘ ‘ F .y..- bi“ ‘ RAF ‘- e“ I 1;“ ‘v '1‘" . -p- a .- .Ao-‘&.oo ¢~ I 1.1% a: \I A unn'a nvaVU vs V vs”, Ab-‘dgt a .. .u“"' ”on ... .. . R“ ..,.1 ~L_ 3.. , "n‘ VLA C. CHAPTER II THEORETICAL HISTORY 2.1 Lithium and Germanium Diffusion Studies The lithium atom enters a silicon or germanium lattice as a singly-ionized donor impurity. Most of the lithium 13 is ionized for temperatures above 500K. The diffusion constant for natural lithium as a function of temperature ll’lQ’ll‘lThese studies show that the has been well studied. lithium ion must move among interstitial lattice sites to account for the high mobility observed. Both silicon and germanium have a basic diamond structure, so that inter- stitial equilibrium sites can have only a tetrahedral (Td) or hexagonal symmetry. Various studies, including stress effects on electron excitations, show that the lithium must be in a Td-symmetry site.3’ls The potential well in which the lithium ion (Li+) is bound in the diamond-structure crystals is the key to under- standing the motion of lithium in silicon or germanium. We have considered two methods for studying this potential. The first method involves a study of tunneling effects through the well as indicated by the diffusion constant at low temperatures. The second method involves the infrared absorption in a transition from the Li+ vibrational ground state to the first excited state. Diffusion constants have been described classically by D(T) = Doexp(Q/kT), 2.1 mere 1 depen' 2P ilicon a: is the gas co; sole of diffu sition state. m. '. ‘-. ‘ ‘334‘45 at as: n' 5 H I’d 0 n .J: ‘ - -6. J): ‘1 “ ~59 lattwc. ;V‘p- ‘Afir‘no’flg c JV”; 7' O; 1 ‘:§: “'7?“ 'a' . -S I i” ‘1 - II‘ I" A h 15 Egg :1 ‘n‘: 6 where k is Boltzmann's constant and T is the temperature in °K. Q represents the classical activation energy, i.e. the height of a sinusoidal barrier between equilibrium sites. D0 is given by the expression16 D0 = Ad2v exp (AS/R), 2.2 where A depends on the unit—cell geometry, and equals 1/8 for silicon and germanium; d is the lattice constant; R is the gas constant, and AS is the change in entropy per mole of diffusing ions transferred to the activated tran— sition state. AS is related to the elastic modulus u and the activation energy by AS = —Q 5; du/dT; no is the elastic o modulus at absolute zero, i.e. "o = ll.2x10-ll dyne/cm2 and du/dT = —6.0x107 dyne/cm2 —deg. for germanium. These ideas were first put forth by Wert and Zener,l7 who assumed that the interstitial diffusant jumped from one hollow region in the lattice to an adjacent hollow region. The parameter of greatest uncertainty in this formulation is the vibrational frequency of the lithium ion.2 In a localized harmonic approxi- mation it is given by v = (Q/2Ma2)l/2 2. (A) where 'a' is the distance between adjacent interstitial sites, and M is the mass of the ion. This uncertainty in v further motivated our subsequent infrared absorption studies. The 0‘ I. .- .4. . “€55.33” “ '5 L v n I'V - 1'. ”3:: EV. #- ' -o:A 72F Atlvd g‘Ur-é 8.. ~ - I .gv ' . ‘ b ‘~"D‘ o v- I .11.: 3'. n -5.|s—. - P A . ' «01“: ev. 1 : he nave .LD “-1 ~‘-:f .n... avg- a»- fin I I met—1:1: s 0:” than the ~. 7 Li dlfoSlOn constant in Silicon follows this claSSical expression (2.1) for all temperatures for which it has been measured;" it is described by D0 = 23x10-1+ cm2/sec and Q = 0.655 ev}2 In germanium, however, a somewhat smaller activation energy is found at temperatures below 155°C than at higher temperatures. The classical expression at higher temperatures is given by D0 = 13x10—u cm2/sec and Q = 0.H6 ev.ll We have proposed that tunneling of the lithium through the activation energy barrier explains the low-temperature discrepancies of the Li+ diffusion constant in germanium. If this is indeed the case, the ratio of the diffusion constants of 6Li to 7Li at low temperatures should be larger than the 1.080 value predicted by the classical inverse- square—root mass ratio. We therefore set out to measure this ratio as a function of temperature. If this effect had turned out to be large, a time saving in detector fabrication would result if the drift were done with 6Li instead of 7Li. A general diffusion theory,18 beginning with a sinu- soidal potential and including tunneling through the barrier at low temperatures, is applicable to this problem. This sinusoidal potential is approximated by a harmonic potential well plus a parabolic barrier. The ion jump frequency 1/1 is taken as prOportional to a sum over all N oscillator levels 9 {This includes temperatures down to 2730K.52 -~ ‘I 's .23 resu-ti. :: “PSI term 9' 8 EN of a Boltzmann distribution times a tunneling probability T(EN), o-ap—a = g exp(—EN/kT)T(EN) 2.u The resulting diffusion constant is epreO(B—l/kT)]-exp[(2N-3)eO(B-l/kT)] l-exp[2eo(B-1/kT)I D(T) = B{eXp(-BQ) expE-(2N+3)eO/kT] I—exp(—2sO/kT) I } 2.5 where B = (2eO/kT)DO, e = (2n2(0.35a)/h)°(2M/Q)l/2, so = hv/2 and h is Planck's constant. Q is the full height of the sinusoidal potential and (a) is the distance between interstitial sites. The number of oscillator levels N is determined from the barrier height by N = (2Q/hv)-l/2. The first term in brackets in eq. 2.5 represents the contribution to D due to tunneling; the second term reduces to the clas- sical expression at high temperature. This expression for D(T) is compared with the full range of experimental diffusion constants in germanium; the ratio of diffusion constants for 6 Li to 7Li is then simultaneously compared with measured ratios. -‘- o 'N'rf'dy 'm‘“ the ion. 0 W ,3. 'L ‘s h "f“. L ""‘na. ‘0‘; - £.n\-_4 r... ~¢ J (D 3.3“. u": ' .. 1" “AI . et Q. “I: s89 ...€OI‘ -c.. ‘ A J :Y‘ r! eel-no-9 . - e ' +.° me aPOFOX‘.‘ ~ A I “I s‘ AA» ....stan‘¥:s were \. acnstants; t..e (‘5ng m..icient e (. ' A ‘ I ‘0‘»: W ’79» *‘L.-\ 5‘ .v‘ a Lu» HCF“ ‘ A 0 v ‘A r‘ &. R "A ‘VL 0 J ghato . ~~ H". W“ "A" p. a‘:,g C" ShC.J" 9 2,2 The Boron Local Mode in Silicon A local-vibrational-mode is a vibrational state of an :impurity ion and its immediate surroundings. The ground vibinational state corresponds to the zero-point motion of the ion. Crystals with point impurities may have other vibisational modes, called band modes, with frequencies belLJW'wM.6 The first studies of local modes in silicon were: theoretical, and they predicted reasonably well the locaLL-mode frequency of the substitutional boron ion.19 The sipproximation was made that the silicon-boron force constxants were the same as the silicon-silicon force consteants; the results were a local-mode absorption coefficient a(w) given by 2,2 a(w) = 2"n:e A |x23 kerg/cmz. Bellomonte and Pryce's calcu- lations, however, showed coupling to second nearest-neighbors to be larger than to nearest-neighbors, but the sum of the uration ,yii. ‘- :cnstants aI‘OU ‘- O- : 1' $33.32. w J? I , u + A 9-9 H 7‘ ‘7 :ZLEHJOs S, a ’ the boron and it: 1 between :crs. in; Inc fer: Mel to give shculd be keni 7'. once constan‘ rqnw . ......al coordii c. tne defect |.-' 19 interactions was in the range 20-30 kergs/cmz. Calculations by Elliott and Pfeuty7 for boron and lithium in the paired configuration required considerable modification of force constants around the pair. The model used considered force constants 8 between the lithium and its three nearest—silicon— neighbors, a+B between the lithium and boron, and e between the boron and silicon pair co-linear with the Li-B axis, and y between the boron and the three silicon nearest—neigh— bors. This model was chosen because it was the simplest model to give good agreement (0.3%) with eXperiments. It should be kept in mind, however, that second-nearest-neighbor force constants are important. In this model each of the normal coordinates of the defect space involves only one of the defect atoms. Only anharmonicity would result in terms in the hamiltonian which mix the atomic coordinates. The lithium line at 522 cm-1 was fitted with a B of 23 kerg/ cm2 for which the isotopic frequency shift agreed quite well with experiments. The boron local modes and isotopic frequency shifts were fitted by weakening the force constant ¢ of the Si-B bond. Variations of other force constants in the basic six atom model above did not give the prOper combination of intensity and isotopic shift of the paired boron local modes. Calculations of intensity were found to agree best with experiments when the lithium charge was replaced by an effective charge of 0.7e. . PV“ .- ’n"2'fi~‘l ‘*(‘ . J.;.I~/' : .u--‘ .0" . ’ ',‘, 6-- “.2 W‘J5. ‘ " 7 7:“ ' “,-.A~‘ v\-(““. - L‘s: ' .Aual..fl‘ fihafi- ‘ .vu»¢"u b4- . I atiarncnm for: LHHJV-wpr : .« ll‘lll‘. IAOVDl“ Tana been st‘..‘. N A!) ‘- r V ‘ .2-.. «10.9 2‘ l s \‘l ““ n- n «It man)», 8"“, .3 a ‘-.l‘* 1m ions ip . t -. .- .F‘ 5‘ w ‘.JL"‘aAp~ s “up... (D n5 ‘ AIL C *1. 15 2.5 Anharmonic Effects on a Td Symmetric Local Mode Line width and centroid shift with temperature are the result of anharmonic coupling of the free or paired ions to the lattice. Theoretical results for complexes in silicon have not as yet been presented, although residual line widths at zero °K have been discussed by Dawber and Elliottlga for the free boron local mode. They conclude that the residual line width is "probably" deter— mined by decay of the local mode into two band modes through anharmonic forces. Anharmonic effects on local modes in ionic crystals have been studied theoretically by Elliott et al.25 These calculations are for charged impurities in a Td symmetry, but many similarities exist with the paired lithium and boron ions in silicon; the pertinent consequences of their theory are presented here. Their calculations begin with an expansion of the potential out to fourth order in atomic displacement divided by interatomic distance. The symmetry of the Td site is used to determine which expansion coefficients are zero. Anharmonic thermal effects are the result of inter- actions between local modes (Operators b, b+) and lattice modes (Operators a(R)). Possible third-order interactions are represented by (b+b+)3 (b+b+)2(a(R) + a+(R)) + (b+b+) (a(R)+a+(R))(a(R’) + a (E'>> (a (K) + a+ > (a 2 (a(R) + a+(R))(a(R’) + a+(R’)) etc., 2.8 where R is the phonon wave vector and branch index. The resulting third- and fourth—order anharmonic Hamiltonians applicable to displacements quadratic in the local-mode coordinates are :3: I + 1/2 1 3 - Eja'ntém:‘( ’fi%+§’) { z (bx+b;)2i(a(k)+a+(k)) 2.9 xyz + +, 1/2 H = I C h(m(k);(k )> Z (bx+b;)2( + +, MM’QwM MN xyz (a(R)+a+(R))(a(R’)+a+(R’)). 2.10 The number of unit cells is given by N. The silicon atom has a mass M, and the impurity atom mass M’; w(R) is the lattice-mode frequency and Q is the local~mode frequency. The Spatial coordinates x,y and 2 refer to the local modes. Perturbation theory is applied to these Hamiltonians to determine thermal shifts of the local modes and line- broadening effects due to elastic scattering of band phonons. In the ionic crystal CaF2 the expansion coefficients c and .B have been measured to be c = 1021 erg/cml4 andfi = 2x1012 erg/CUP and similar orders of magnitude are expected for g I vyfl" -Eai-.~ ‘ .err siere 1.. S bets: 0. n o‘« Yu flit J3 an» Ca V A u l C \\ F. v . ‘J h l- ‘3'“,e TL l7 homopolar crystals. Elliott at aZ.found that the thermal shift in local-mode frequency is the result of fourth- order effects i.e. HQ. Hu was averaged over band modes, leaving local-mode operators unchanged to give .+ H = Z r fiw(k’ 2 (b b++b+b ) 2.11 H . 2 x x x x + MMM NQw xvz k M ’ where terms like a(k) a+(k’) have averaged to zero. Com- parison of eq. 2.11 with the ordinary oscillator Hamiltonian given by %fi02(bxb;+b;bx) gave a relative frequency shift of x M '12 where n(k) is the phonon number operator. This expression becomes linear in temperature at high temperature i.e. above 6d/2. Line width may also be explained in terms of anharmonic coupling to the band modes. Spontaneous decay of the ex- cited local mode into several band modes requires that the band—mode frequencies satisfy the energy—conservation re- quirement that {n(R) = Q, where the sum includes the energies of the several band modes created. The local-mode frequencies of lithium and boron in silicon are between wM and sz’ so that decay into two band modes through the third-order an- harmonic potential is likely. This is represented by the product (b+b+)(a(R)+a+(R)(a(R’)+a+(R’)). The resulting Hamiltonian given by Elliott at aZ.is l8 /2 , _ ’ h w(R)w(R’) + H3 7 2+ 2MN 2(fi 2M’Q Z (bx+bx) k,k’ “M xyz + ++ + ++ {a(k)+a (k)}{a(k’)+a (k’)} 2.13 Perturbation theory with this Hamiltonian gives rise to a two-phononwdecay line width of ,2 %.=45 "2; u I 3; w(R)m(R’){[n(R)+l][n(R’)+l] MM QM wM E,E, N - n(R)n(R’)}6(w(R)+w(R’)-Q) 2.1a where v is the local-mode oscillator level, taken as one for decay from the lowest excited state to the ground state. At low temperatures % becomes constant, whereas at high temperatures it varies linearly with temperature. In the ionic crystals the dominant line-width broadening was found to be due to elastic scattering of band phonons in transition Iv,n(R), n(R’)> + Iv,n(R): l, n(E’)¥ 1> with conservation of energy requiring w(R) = n(R’). The resulting width is 2 2 % = 2.(t - 23 2) (fiiliii%) I if w2(R)n(R) M 9 NM QMwM E,E,N [n(R’)+l] 6(w(R)-w(R’)) 2.15 C. c. r . ‘71 .9 '- n A V t u e a '1‘“ "A. C J. g A ne vocal m ‘ e O t ouzh mea kte ..5.| n ‘astlc n R "L ed ' I nfi " I - U L 0'5“ _8 t... 'x «~30 Lf‘ EH nng l .iLS 6‘ 3:. pa — I ,cv‘ ~M n bx-er-me. uIvJ or i ‘- v a ' s ‘sfv ion 0 L- 55 sa ; SETLCT e ¢. .fIA -~» 3 R :u v-AAA 1“.“ 'v. 19 This elastic scattering line width goes to zero as T7, and at high temperatures it varies as T2. One of the best experimental means of further vali- dating the above line width and centroid expression may be through measurements of line width and centroid of carbon local modes in silicon. Carbon enters silicon substitutionally; hence, it has a Td symmetry, and all basic conditions of the theory are satisfied. 2.6 Experimental Difficulties in the Observation of Local Modes Observation of the free—interstitial-lithium local mode in silicon or germanium presents several difficulties. The most serious of these is the presence of free electrons associated with the ionized lithium. These free electrons can satisfy energy and momentum conservation by the absorption or emission of a lattice phonon simultaneously with the absorp- tion of a photon.27 The resulting experimental absorption coefficient per unit lithium concentration in germanium varies as v-l’g from l.lxlO-16 cm-l/cm3 at 500 cm"1 to 7.0x10"l7 cm-l/cm3 at 800 cm-1. In silicon28 the absorption coefficient varies as v-2'0 from l.lxlO-16 cm_l/cm3 at 500 cm.1 to 7.0xlO— cm-l/cm3 at 800 cm-1. If the free interstitial local mode 17 is assumed to have an absorption intensity similar to that of lithium—boron-pair local modes then a value of approximately 19 cm—l/cm3 would be expected. The free-electron- 10x10- absorption coefficient would then be about 170 times larger. These absorption coefficients are additive in the attenuation factor e-ax, so that making the sample thin enough to transmit an Optical signal reduces the local-mode absorption strength .. - +'\ & 5A"'”O.€C» ‘” .v ".-‘\v fixer wave nu." - Q ' 1 ,2 fiA ‘ N V,’ “ U ..u'-.‘. ‘1 1" u‘ ~ 3-21; azcut l4 :,-.v-:-p°/4 1:+““I' .911-va ““‘l.‘ fie have < Oknfi‘un. ‘R A; --..vu'. ‘ ' c bnl \r. :cfiup“. h "u 'T "' .H- u ‘3, F‘OHR‘.‘ ‘ O ‘V‘la‘L" ;ev‘y'( "x "n ‘- "“' SLar. :‘~r\ ' ~ “\f~ ' n )v._‘.‘ n a S - : N «.“D‘ we Carr“ ep- r _&Hh. ..-er‘s 1'16" ”W emsat :‘a l h 7“ Des ‘Qh< "’K— ~‘~ ~A~y\‘\ N . ““rwt‘ A {U - ‘A At L. ‘\ T:‘ I ‘s ‘H “"C“ a“ I “ r‘ 'N 20 to undetectable levels. There is a temperature dependence to the free-electron absorption in germanium. At 700 cm“1 the absorption is constant with temperature, whereas at higher wave numbers it increases with temperature; at lower wave numbers, it decreases slightly with temperature. Cooling would therefore be of little benefit until temperatures below about u0°K where the lithium re-ionizes,l3 but the un— ionized lithium represents a different problem. We have considered the possibility of shining light through the charge—depleted region of a reverse—biased diode junction where lithium was used as one of the dopants. The depletion region is so small, however, that the resulting intensity levels would not be detectable. The standard method for reducing the free—carrier absorption is the introduction of an equal concentration of an acceptor impurity. Ideally this addition would per— fectly compensate the lattice electrically; in practice, the compensation is usually not complete, and a residual free-carrier-absorption background remains. These free carriers may be either electrons or holes, depending on the compensation processing. If the remaining free-carriers v-2.2 are holes in silicon, a wavelength dependence29 of is ~16 1 found, which varies from 9.2x10 cm.l/cm3 at 500 cm- , to ’16 cm-l/cm3 at 1000 cm-1. 2.0xlO A perfectly-compensated lattice will still have an absorption background, due primarily to the lattice itself. Silicon and germanium are covalent or homopolar crystals, .3: be ‘Vj.SZJéil: :cie induce S ‘ simultaneousl‘ an electric I? izgt the t1» 5:23 or dIfI‘E .ne mam asso 3.32m surma * “— :eutron spec: fiensit‘; of st :or both sili Jean spectre: ETA" A ' «ctEu out I ‘..L tfle refer“: Q . o ~Pmihlur § u “hf-l o J§\"ep to c 5:5; \‘ ‘ er 21 and have no net dipole moment per unit cell. In these crystals lattice absorption is a two-phonon process which can be visualized as follows: A first lattice—vibrational mode induces charge on the lattice atoms. A second mode simultaneously causes the charge to vibrate, thus producing an electric moment which couples to the radiation field.30 The structure of this absorption spectrum then varies accord— ing to the two phonon density of states, and represents sums or differences of two or more lattice phonon interactions. The main absorption peak in silicon, at 610 cm—1, is the two— phonon summation band of TA and TO phonons; similarly, the maximum absorption band in germanium at 360 cm.1 is a summation of TA and TO phonons. The dispersion relations obtained from neutron Spectrometry31 show these branches to have the highest density of states. Typical peak lattice absorption coefficients for both silicon and germanium are about 10 cm—l. In a double- beam spectrophotometer, the lattice absorption can be sub- tracted out by placing an equal thickness of undoped crystal in the reference beam. The system of lithium in silicon differs from that of lithium in germanium in two important respects. The Debye temperature in silicon is 658°K, in germanium 3620K. As a result, thermal broadening of the local-mode peaks occurs at much lower temperatures in germanium than in silicon. The germanium system must therefore be cooled below 2000K in order to observe the local modes. The two systems also differ in the diffusion constant of lithium. Hence, in ....¢ ‘ A u' I“ A ‘IHVCJ 9 0m. - ' w“ ' Litnnrn cc“... e. “V . P“ '5‘ o. . .-e.e. .. .. ‘. . n ‘ ‘r'D fi'flr‘p ‘:“:\.‘~ CM.A\.-I.».. .4 :I‘. LIEdCCGDTdLT. ":1 ' :4 . w‘Ls‘ I Of‘n.‘v‘£ h§ull-\lla \vb“ ‘ ”:7 ‘ A “h,“ ‘-"“‘eG \AC;u»/L ‘ I . “ “H ”‘38 have u, ..f ~QMN‘Q.,: l.‘ ‘ :‘~.}“- M‘ ' PK C. ‘ _ c. . * .\. *--.f‘r \ L'.‘ 1 .HQ 3*‘I . ‘M .‘A K sx‘ve ~‘ 1)‘ h“ I 13“.”) _ ‘ ~‘I .r‘n '_ ~V.| ‘ . ‘3 ”A .‘vfi v._ ‘ UDhACh n“~\-‘ ‘ r ‘l _ \ ~l I " ‘ '.4 \""‘ ' yqyj‘»‘_~“‘ k‘-‘ ‘ ‘Q 22 the rate of precipitation of lithium from super-saturated germanium is faster than that from super-saturated silicon, by approximately the inverse ratio of diffusion constants. The precipitation of lithium in germanium continues at temperatures down to room temperature; consequently, lithium compensated germanium must be maintained at a low temperature until all measurements on the sample are com- plete. The precipitation of lithium would otherwise leave large concentrations of uncompensated acceptors, and hence an unacceptable free-carrier background absorption. More will be said about the precipitation process later. The two systems are similar in several important ways. Lithium occurs in both silicon and germanium as a singly- ionized donor in an interstitial lattice site. Both lattices have the basic diamond structure with lattice constants of 5.H3A for Si and 5.663 for Ge. Furthermore, the outer electronic shells of silicon and germanium both contain two 5 electrons and two p electrons, hence similar repulsive forces on the lithium ion due to the exclusion principle may be expected. Many aspects of the local-mode studies on silicon can therefore be generalized to germanium through appropriate frequency or temperature shifts. 2.7 Complexing of Donors with Acceptor Ions The introduction of acceptor compensating ions into the lattice gives rise to the formation of positive and negative ion pairs or complexes. Calculations of the rela- tive percent of complexed ions, for any fixed temperature and substitutional acceptor concentration, have been done ' : trancn oi antor‘ ions, Luv-i“ " QRR v- v A .- 2(T), a T). H n M u( .2 I‘A ,f" ‘m CPS an “‘5! "J"‘ 23 . 22 , . . by Reiss et a1, They assumed that paired and unpaired ions were in dynamic equilibrium hence, a law of mass action could be written down. Thus, if P represents the concen— tration of paired ions, NA—P, the concentration of unpaired acceptor ions, and N -P the concentration of unpaired donor D ions, the mass action expression is P - (NA—P7 (ND—P7 9(T) , 2.17 where 0(T), an equilibrium constant independent of concen- tration, is given by 2 e n(T) = Mfr? exp (1??) dr . 2.18 E I‘ These calculations included only the effect of coulomb interaction between ions. The fraction of paired donors (P/ND) was then a function of the acceptor concentrations and n(T). In the case of nearly equal concentrations of donors and acceptors, (P/N) is given by ) . 2.19 Appendix B contains a somewhat modified form of this expression evaluated as a function of temperature for equal concentrations of lithium and boron in silicon. As would be expected, the fraction of paired ions increases rapidly with impurity con- centrations (N) for a given temperature. For a fixed N, p 5 A 110 che V _.o- .,. II' ‘u wfipnsat “ravat 0!] CC)“: D at t' . 3‘3 C uS‘ , X;- .4 “l. ‘ M' ‘Q} in .19 .i '7 Ike‘ § 0 | \ Cu l~ £ 2n the fraction paired decreases with temperature. The possi— bility of observing the unassociated Li+ or B- local mode in a compensated system must subsequently be considered. That is, is it possible to quench a compensated sample from a high temperature fast enough to prevent pair formation? Reiss et aL22 have made calculations of the average lifetime of an unpaired ion, and denoted it as the relaxation time T for ion pairing. They found T to be closely approximated P by T = —Ekg(§-M) 2.20 p une N D (T) 0 where M - JL Lip? ( e2 )dr 2 21 - u" exp EETF’ . . a The distance of closest approach (a) between Oppositely- 0 charged ions can be taken as 2.6A. The Debye length L is 1/2 given by L = (ekT/8ne2N) , and DO(T) is the lithium diffusion constant. For impurity concentrations of N = lOlglcm3 , M is small compared to N so that Tp varies as T/NDO(T). To freeze-in a sizable fraction of unpaired ions, i.e. 10% at ”00°K, it would be necessary to quench the sample in a time small compared with the relaxation time I at that temperature. At MOO°K, Tp has a value of 1.3x10-3 sec.; at room temperature Tp is 0.76 seconds. Hence, a quench fast enough to freeze—in unpaired ions is unlikely. At room temperature about 1% of the ions remain unpaired. As a result, the free-boron local mode a! [as—- I ~. an] 1‘. u- 25 has been observed in compensated silicon.u’5’ga’gb’10 However, a free-lithium local mode has not been observed, and it is thought that it must lie in the lattice continum.5’6’9a .F““ “1' 4-; ran aan QA bra L .LC‘A: :rcr to chat: 331:3; conce: “Virtnr 'vvsA ‘ :2; gh-ds‘ 1......"v‘r 5‘ :c-ze a:sorotio . n do; VOT‘Ser‘.Qa‘ Three 1“ a c I :‘.:Af\ f: ‘ —«c§\—Vn “0:8[3 A “A “Q9 “h .ucv_ A F :J‘O\.ES\D \:.~ 5 . .J A‘é- - P “3 3‘43» :C* - . . .5“~ ‘1; ~JI" CHAPTER III SAMPLE PREPARATION The infrared study of local modes in silicon and germanium required a sample uniformly doped with equal amounts of lithium and an appropriate acceptor impurity. Concentrations were required to be equal within 1% in order to eliminate the free—carrier absorption background. Impurity concentration and sample thickness were then of primary importance in determining the observed local- mode absorption intensity. 3.1 Compensation Methods Three basic methods for producing high-resistivity silicon doped heavily with lithium donors have been reported. These processes are all similar in that they introduce trap— ping sites for the donated lithium electron. The earliest method involved a radiation of the sample with nucleons.32 The silicon would first be diffused with lithium at high temperatures to produce the desired concentration.* The samples were then irradiated with nucleons to produce lattice defects. The defect site then captures one of the lithium donor electrons. In this manner the silicon was compensated to give resistivities near that of pure silicon. Spitzer5 later described a similar compensation technique using electron irradiation in the neighborhood of l Mev. This method is thought to leave a larger number V— ::For solubility of Li in Si see E. M. Pell, J. Phys. Chem. Solids, 3° 77 (1957) 26 ‘ + 5?. accelera LC: A k .- reputed by S. w- xztr. a sub st '4 l aCllEVEd by a El'v’e Uniform} 2339, of the S' t: be heaViep 27 of unpaired ions than techniques based on a final thermal diffusion of lithium to achieve compensation. This is because the nucleons, e.g. neutrons, will leave defects randomly distributed throughout the silicon and not neces- sarily near a lithium ion. The irradiation method, on the other hand, is not so convenient because of the need for an accelerator. A second method of compensation has been 5’33 It consists of growing the silicon reported by Spitzer. with a substitutional donor and a substitutional acceptor such that the donor concentration is about 10% less than the acceptor concentration. The final compensation is then achieved by a diffusion of lithium. This is reported to give uniformly—compensated samples; furthermore, no local mode of the substitutional donor will occur if it is chosen to be heavier than the host silicon atom.7 Another early methodau starts with silicon uniformly dOped with an acceptor such as boron. Lithium is then diffused in excess throughout the material. The final compensation is produced by annealing the sample, thereby causing the excess lithium to precipitate. More will be said about the precipitation process later. The diffusion-precipitation technique has many problems associated with it, but requires less apparatus than the other techniques. It does not cause nearly as much damage to the lattice as the irradiation methods and involves generally fewer uncertainties. The double-doping during crystal growth involved technologies not available to us; as a result, the diffusion-precipitation compensation tech— nique was chosen. These techniques apply equally as well for the compensation of germanium. q 0 W“ cms w u i. h5~45fi"¥ were not as r1 5‘. 1 11' : the ['41-] C‘Y‘ . d- A :rys .al , and -y¢:..__ru. .— Jf'fill ' £1 '1 ”Wunf- W ML“ h‘josaL. - #- Au.l Ln “31,0! ‘ ‘ “ v “:7 ts ‘Ar‘. ._. "h a. et'nha‘. ‘ “ “lOv‘ AA :Qufifi ""338, a7. -L 2‘ 28 3.2 Material, Mechanics and Preparation The initial material* was single—crystal silicon grown in the [111] direction and dOped with 1019 atoms of boron per cm3. A concentration of 1019 was chosen because it was reported?” to produce local—mode coefficients on the order of 10 cm"1 which could be readily observed. Higher concen- trations would produce larger absorption coefficients, but were not as readily available. The three symmetry axes of the [111] direction were evident along the sides of the crystal, and the orientation could be readily ascertained. X-ray diffraction photographs were taken in order to con- firm the orientation, and show that the material was single— crystal. The sample was then sliced in the desired orientation on a diamond saw. The resulting wafers were typically 0.7 to 1.5 mm thick. The infrared radiation is attenuated while (1X traveling through the silicon by e- , where a is the absorption coefficient and x is the sample thickness. The maximum atten- uation of the signalnu that we could easily work with and also the maximum sample thickness that didn't require excessive preparation time, techniques, and cost was 1.5 mm. The thinner samples, although easier to make, gave smaller signal variations a,- 0'0 0’. lb 0‘ l! between the local-mode peaks and the background. “Purchased from Materials Research Corporation. :': 2': . Additional signal losses were caused by reflection from the surfaces and the small cross-sectional area of the sample compared with the spectrometer beam. I'D Q'. J. 0 a. "This is primarily due to thermal surface effects dis- cussed later. . . r0 v" IF I .6“ .2- .‘ -~ a mu. ‘ .th'! 171a ed -A""‘ .id'44t5 .1 +1— .1 PD eye 3 a u V ‘A ustra' 1 1. .1 \JV IV 33 it w if A UP Cc L d haVQ t Edo ‘I s h ‘ h“? ““39 C111 «911 t“ ’V 29 3.3 The Lithium Deposition 7Lithium* was deposited on the boron-doped silicon by an evaporation process. The silicon was first lapped with #600-grit carborundum, and then with aluminum oxide. It was finally cleaned in an ultrasonic cleaner containing methal alcohol just prior to the lithium deposition. The lithium metal was placed in a tantalum boat, which was then mounted in a bell jar. A tantalum boat was chosen because it wouldn't react rapidly with the lithium at the vaporization temperature (1609°K). A grounding path to the sample was provided to prevent a possible surface charge build up due to deposition of thermally-ionized lithium atoms. Figure 3.1 illustrates the evaporation system. The bell jar was pumped down to two microns or below before the lithium was heated. Vaporization of the lithium was achieved by adjust- ing a current through the tantalum boat until the boat began to glow a faint red. If the boat was allowed to get too hot, or if it was placed too close to the sample, the deposited lithium layer was likely to boil off, and the sample would have to be reprocessed. Samples were typically kept 2.25" or more from the lithium boat. We generally tried to get a thick lithium deposition, i.e. greater than 10 mils, during the evaporation. The heating current was then turned off, and argon was let into the bell jar. When atmospheric pressure was reached, the sample was quickly removed from the vacuum chamber and placed in xylene. The amount of time that the evaporated sample spent in the hThe lithium isotopes were 99.99% pure obtained from Ortec. A\ Vacuum Belljar thhlum In Tantalum Bani rung! rE‘LIFI [ LE" \ ‘5‘ Sample (3mg) Meter Current Source.» FIGURE 3.| Evaporation system. farm-I " .L :o‘qaqfi‘ll‘ere ha.) wu- -; ' A ‘v T" sartace la.e. ' 1‘ u an: esceCLa-l. :5 found that surface when 1 II 1 C " :1..erai 01 l 'w c t) e remained (r “'1‘: ' ‘ -‘* Autism. The over. and about 3 .5 ie removed f0 ”9"" ' . rule prot‘vlCe: .. '1— 4 fl _;;vf\ r“ "a k, *A;e I 6 ‘-r\.,‘g ‘ 'nv' it Q..- a 9“ "' r 9 ( ’5 ME ‘ ‘4. qh'vfi .‘ . 24.1 F‘I P Q‘L_\’,’ ‘~ “(3! "‘3 ti‘o ‘3» 0V7 \— Vt ~h “n \1 ‘ I) ‘ - «ages A \ VI .‘s A a" g 31 atmosphere was kept to a minimum, because the lithium O CO 2’ 2’ 2 and especially water vapor in the air. Mineral oil was surface layer would react quickly with the N also tried as protection for the lithium surface, but it was found that cracks and peeling occured in the lithium surface when the sample was placed in the oven. If the mineral oil was burned off slowly at about 260°C, the sur- face remained quite good even though this temperature is above the melting point of lithium metal. 3.H Diffusion and Lithium Solubility The oven consisted of a quartz tube about 6 feet long and about 3.5" inner diameter. A joint allowed one end to be removed for insertion of samples. Gas-line connections were provided at both ends. At the high diffusion temperatures involved, lithium reacts readily with most substances except inert gases, consequently an argon atmosphere was provided during the diffusion. Three independent heating coils in the oven provide adjacent temperature zones of one foot in length and thermally regulated to :0.1°C. Insulation and electrical controls make up most of the main unit. This oven tube and the quartz sample mount were preheated to 375°C. The mounting pad was then withdrawn into the re- movable end of the furnace tube and the tube was Opened. An argon flow of 25 kcm3/min was maintained over the pad. Then, the evaporated sample was removed from the xylene and placed on the hot pad. The furnace tube was quickly closed up again, and after allowing two minutes to vent 4‘. . “2" 3.:Osf" 9": U \— \ bx ity u:: 4... lo 5 e 4. '\A the oven .2 the d :ate t3 301ub' ,Au 0". “w 32 the atmOSpheric gases from the furnace, the argon flow was reduced to 10 kcm3/min. If the sample was not initially heated from the quartz pad, it would have to be heated very slowly: otherwise, the lithium surface has a tendency to heat faster than the bulk silicon, and the lithium film is likely to boil, thus leaving only small beads of lithium on the silicon. The sample was left at 375°C for ten min- utes to form the initial lithium-silicon pre-alloy.3H The lithium was then firmly attached to the silicon, and the oven temperature was then raised to its diffusion temperature of 550°C. The argon flow rate was maintained at 10 kcm3/min. throughout the diffusion. Several factors had to be considered in the selection of the diffusion temperature. Perhaps the most important factor was the concentration of lithium desired in the sample. Clearly, the final lithium concentration must closely approxi- mate the boron concentration. A compensation to within 1% of equal concentration would reduce the number of free carriers to a manageable level of 1017/cm3. It is conceivable that the diffusion temperature could be chosen such that the solubility of lithium in the doped silicon was equal to 22’35 that an the boron concentration. But, it is known indeterminate amount of lithium will precipitate upon cooling, leaving the sample strongly p-type. Thus, it was decided to use the technique of diffusing in an excess of lithium and annealing out the excess at a lower temperature following the diffusion. o g ‘ ‘A A fill .\" ‘3 ' al- A‘ha—l A nob-v \— '5. t v-r A o O ‘ .- J l? a S h 0 11. «J; Dy v \: c . . C r ‘9‘ 01 d 4H ‘3 AV 7 . cl 0.. n.. 1 . e j . h. . .5“ .fi.. a | o. u a a c v. (r. .5 . . a» f. . A... ¢ \ «\u ‘6‘ P. o F‘. A.-. I~§ AP. I 071 -‘ :entrat . 33 It has been shown22 that an acceptor impurity in silicon or germanium has a pronounced effect on the solubility of lithium. This effect may be understood by considering the analogy with impurity ions in an aqueous solution repre- sented by the silicon. It can be seen in the following chemical-equilibrium expression that the presence of the + . . free hole "h " from the boron impurity greatly enhances the solubility of lithium; B 2::8 + h + + - Li ext.‘—-=7—Li‘-—:_-; Li + e J h+e— At the dOping levels involved, the dilute-solution condition is valid, and mass-action relationships can be written for each reversible reaction. The mass—action law states that in a dilute solution, at equilibrium, the product of the concentrations of the unassociated ions divided by the con- centration of initial associated reactant is equal to a constant. Such relationships can be written for the boron acceptors, the lithium donors and the associated electron- holes e-h+. Calculations based on these equations, including also the charge—neutrality condition, predict quite well the solubility of lithium in acceptor-doped silicon.22 These results show it to be a simple matter to match lithium and boron concentrations, but in order to reach uniform lithium 19 concentrations significantly above the 10 /cm3 boron level, O a"! a V ;:::“C C-..J-7' allc' v. -7 arivoT‘ A: P. :u .1 “Q p l -A no: ltfll‘lm CO. A re tcta all-OD Cd I . m ' n.v‘\¢ .-_.. O 1' ‘ is “\(m I. A '7‘“. "+ ““33 tant . t ‘ pa \n S fcun . 1 a. ”A. 31‘31‘] Ga 3 n ‘ A L. I u.“‘ ‘A 4 Q4. an a temperature near the eutectic point of the lithium-silicon system is indicated. At this temperature, lithium would diffuse away from a high concentration silicon-lithium sur- face-layer alloy. The temperature is high enough to allow sufficient lithium to remain in the silicon following pre- cipitation caused by cooling to room temperature. However, it would not leave the very large precipitation sites and resulting local strain of a much higher diffusion temperature. The eutectic point is reported2H to occur at 590 : 10°C and a lithium concentration of 58 i 5%. The total diffusion time also had to be considered in selecting the diffusion temperature. The diffusion time td may be approximated by td = d2/HD(T), where "d" is the dif- fusion depth and D(T) is the temperature—dependent diffusion constant. D(T) has been measured at high temperatures,“l and is found to closely approximate the temperature dependence given by the classical model D(T) : Doe-Q/kT , where D0 is 23x10-” cmz-sec, and the classical activation energy Q is 0.65 ev.2 Unfortunately, the diffusion does not take place 3a’3b This is partly due to non-uniform wetting uniformly. of the lithium to the silicon during the pre-alloying phase, and partly due to oxygen impurities33 in the original silicon material. Our own experiments show that 5td gives a reasonably uniform lithium distribution in the material that we used. During this time (Std), either a continuous and rather expen- sive argon flow must be maintained, or else the sample must be placed in a sealed argon atmosphere. The sealed vessel _.ach "an.“ 8" , ‘o-H +“i ‘~.‘ b_,... w +.» a; up; uated. O'rzc v ' ab ‘7" A out of its 512053.78 I5 , Hp da'x— ! t s q «3 DO #31 C. .1 «C CO 3. a v. A the 3‘. the 't.- San“ .pV 35 approach is reasonable, but somewhat difficult. It would be necessary to prepare a quartz vessel and a mount made of pure silicon.* If the sample were allowed to make con— tact with the quartz vessel, the quartz would tend to draw lithium from the sample at the point of contact. A sample coated with lithium and protected from the atmosphere with xylene could be mounted in a quartz vessel which is then evacuated. After allowing sufficient time for degasing, 1/2 atmosphere of argon would enter the tube. If the tube were left evacuated, the diffusion would be very difficult because of the evaporation of lithium.2u The tube would then be sealed, and care taken not to overheat the sample or knock it out of its mount. The whole quartz vessel could then be placed in the diffusion oven. If required, quenching of the sample, though difficult, could be done. Of course, any errors in the sealing process, particularly gas leaks, would not show up until after the diffusion. This process was found to be troublesome, and as expensive as maintaining a continuous argon flow. Hence, a higher diffusion temperature and a subsequently shorter td is obviously advantageous in terms of cost. High diffusion temperatures, above 600°C, introduce some new problems, however. Evaporation of the lithium21+ from the sample becomes serious, as does the increased reaction rate of lithium with impurities in the argon. The precipitation of lithium during cooling of the sample, is x“ u "SiC is also a good mounting material.3 “fix-hr 5 f-"1 ‘ . 1.\.lf‘ A '5'..r- fit a dl‘fv‘: fl 4 A : Ir V Tate O- 13 Act: ’/ 1231's was reculf ;.-1 ' ‘ 5:531:85. Dv eve.” #7") an; d:::. ‘-:V., 4-H Ln.- . I- ‘g .e..per~at.1re, th: 36 u . . also much more severe,2 ’3u’35 so that it is not at all cer- tain that sufficient lithium remains dissolved in the sample. Furthermore, the presence of a large number of precipitation sites in the sample could cloud the interpretation of the data. At a diffusion temperature of 550°C and an argon flow Pate of 10 kcm3/min., we found that of the order of six hours was required for the surface layer of lithium to be depleted by evaporation, reaction with impurities in the argon, and diffusion into the sample. At this time and temperature, the lithium could potentially diffuse through a one-ndllimeter-thick sample. The sample was quenched by dPOppiJig it into room—temperature mineral oil. It was then Preparwad for another lithium deposition, by a coarse lapping with #tSOO—grit carborundum to remove about 2-to-3 mils from the ditffusion surface. This step was taken because the sili- con inunediately below the evaporated lithium layer came out Of the diffusion as p-type material for a depth of 1- to—2 mils. 'Ihis is thought to be the result of diffusion of Vacancies into the silicon, which combine with the lithium. Best rwasults in the final lithium distribution were obtained when tlie first, second, and all other evaporations were made on the same surface. A£3 mentioned earlier, most materials in contact with the Séunple in the diffusion oven tend to draw the lithium from tflie sample; furthermore, many materials will diffuse into tflie silicon. We also found, and it was reported by ..Ja'»l"-T’.' '7" ' _. '44--- ~p‘ 5.1 w , . fl, .tn‘u- \4 ....A- AO‘. :33; Oil. C L . a ' .enratmn . ‘u :1- were fou 33171331"! C0,; ‘q" ~=L‘on 4‘ § 0 1 . g. 3 ."L ‘ ’M “it fag A.H DPO“\ § s. ::¢. “ 55“ ‘L . ‘\' ’I we A. I. ‘ . “‘Ou‘t *L 'Q l. "v- e. 3. of “"n »,.‘, ":3 3:. 37 others,5 that any free surface of the sample acted as a sink, or zero boundary condition, to the diffusion of lithium. Out-diffusion could occur as fast as in-diffusion, leaving unacceptably large lithium concentration gradients throughout the sample. Use of a piece of pure or lithium- doped silicon turned out to be the simplest solution. If good contact was made between the sample and pure silicon, the sample came out sufficiently uniform in lithium con- centration. Contact surfaces flat to within about one mil were found to be sufficient to remove the apparent zero boundary condition at the sample surface. The somewhat lower solubility of lithium in the pure silicon tended to prevent considerable lithium losses to the pure silicon mount. Another more expensive approach would have been to simply make the sample four to six times thicker than de- sired; then, when the diffusion process was complete, a piece of the desired thickness would be sawed off. The samples were quenched upon removal from the oven by first being moved to a cool--about 350°C-—section of the furnace tube for 15-30 seconds. They were then drOpped immediately into room-temperature mineral oil. Precipi- tation of lithium in silicon cooled from 550°C is not severe, but the probability of ending up with strongly n—type material after three to five diffusions would be considerably reduced without the quenching. At the diffusion temperature of 550°C, most of the lithium and boronis unassociated.22 Some thought was given to the possibility of freezing in the unassociated I u v!“ 0' ' ut I! ‘ oh I . t ' l-H' ‘ I .Appnfj relax ;:‘.VAA~A . T”the? cor: .-... --| »::c: ‘ ‘-.“Lv‘[~l + 4.” ‘U '1 rut: v 1‘.‘QS‘O 3‘ r \‘k (n r ‘- :‘A '--\JY,:Q’1‘ tL‘ “u ..at ‘3 r- u.:\ ”'1‘ ‘ ‘nx. ' no. .V _ .. Cher“ 2a.} 1 qui a 1‘ \ 38 lithium ions by rapid quenching. However, our quenching Cprocess turned out to be too slow compared with the milli- second relaxation times22 of the pair-formation process. Further complications would have been introduced by the necessity of annealing out the excess lithium. As a result of the short pair-formation relaxation times, the quenching can be safely assumed to have had little effect on the ratio of paired to unpaired ions. This was later confirmed by observation of relative absorption intensities in the local-mode spectrum. Much of the germanium sample preparation was exactly the same as the silicon sample preparation just described. Because of lithium precipitation and lack of solubility, we were not able to compensate germanium dOped to 2x1018 Ga/cm3. l7 Concentrations of 2.5-3x10 Ga/cm3 germanium could be com- pensated with a lithium diffusion, but the subsequent local- mode absorption coefficient would be less than 3 cm.1 and difficult to detect. Samples were compensated by lithium diffusion at u20-u60°C with total diffusion times of 1.25- l.5 hours. The relatively short diffusion time was possible because a signal-maximization technique (see Appendix A) showed that sample thickness should be typically 100—600 u. 3.5 The Hot-Point Probe A check of carrier type was necessary after the diffusion and also during the subsequent precipitation process. This check was performed with a simple hot-point probe. The probe .3 l 1 E s .. .9 a... - y . a. .._ s. T; ‘ _ a a. v . n. A u . c P, .C C e . . .. . S .3 .5 a D. 24 +.. D. .3 w o n“ vb” I . a: a. ¢ 2‘ w . Hu 0.. O. L H I . . . . .1 5.. 3.. t . ..w no t. c to 6. Av F. 3. s a J. L. r. W; 3. %. AU n. ah. .23 3. o . ... L.” ”a 2. Av a.“ ”a CV .,. A.” . . v. A... o 4 g: .3 :. o . a.” q .. 0.. .ya a: we. run. a . .\. q .. o.. A.. 5.19.5. 30o.” .4. 3 Fur. ‘ e ’0'! C o ‘ y i VQ.' 39 was a small pencil-type soldering iron powered by a variac set to 6U% of full output. The sample was placed on an etched metal plate, and the hot probe was placed on the surface of interest. The voltage between the probe and the metal plate was measured with a Keithley model 155 null-detector microvoltmeter with a minimum DC input im- pedance of 106 ohms. A positive voltage reading, relative to the etched metal plate, indicated an n-type surface under the probe, and a negative voltage indicated a p-type surface. This measurement was found to be most sensitive for lightly-doped or nearly compensated semiconductors; it was, therefore, especially useful for our purposes. The magnitude of the thermal voltage was found to depend strongly on the contact pressure. As a result, a mechanical— probe arm was constructed which allowed equal probe pressure for each measurement. 3.6 Precipitation Kinetics The kinetics of the precipitation process have been reviewed for lithium in silicon, and for lithium and copper in germanium.35 Only some of the highlights will be mentioned here. The precipitation kinetics for lithium in silicon is essentially the same as for lithium in germanium; the primary differences are in the rate at which precipitation occurs at any given temperature. Our main purpose was simply to remove all excess free carriers associated with the lithium ions in the silicon. It turned out, however, that one of our principal 3.. +. a“. a C l. C .1 S a: C .1 v. Av +L A: . . r. A—- Run a .- .\_. .va 5 a 118:9 O . o 1| 1Y9 tn g 0 9a“‘ ’4 n V-“ "‘urated a: ‘— 4‘ - 2 m . ‘ —A ‘S OCVLO‘ O + .5 O J. C. 1 . oanuc o a a vat - CA .. V H“ LJ§IQ n .N‘ ‘nn s-‘Vto Q --¢. r' no spectroscopic background problems could be attributed to a precipitation effect at the surface of the sample. The precipitation of lithium in silicon or germanium is a heterogeneous process in which the nucleation sites are lattice defects and lattice impurities. When lithium is diffused into silicon or germanium, and then quenched rapidly, the system is very much analogous to a super- saturated aqueous solution. The lithium tends to diffuse toward the nucleation sites and congregate about them. It is obvious that such a process is diffusion-limited, i.e. the precipitation time is primarily determined by the length of time that it takes for the bulk lithium to diffuse onto a nucleation site. Equally important in the precipi- tation process is the density of nucleation sites; this density may vary several orders of magnitude from one crystal to another. For crystals refined by the same process, the density of nucleation sites is approximately of the same order of magnitude. In the case of an equal number of nucleation sites in samples of silicon and germanium, the precipitation rates can be scaled as the ratio of diffusion constants. Furthermore, the rapid increase in lithium pre— cipitation with diffusion temperature can be attributed to the increase in the diffusion constant with temperature. The precipitation of lithium continues until the solution is no longer supersaturated. That is, precipitation stOps when the density of lithium in the solution becomes equal to the solubility of lithium in silicon at room temperature, ;. (D :3 F? (f ..L\. 1"! RA ““ .5 a §u~3 aft! ‘5qu “,3. a...;.. .u a- can. (‘i :0 . _.'v35 ‘1“. . Q‘A‘“ a“: 9L“ a 5 A..- azure a n. "- é-‘h .13; “.9 I' J!" - A‘l nu— -.. . "a Ay‘npq “w 3.1.413. « .2va t . . «my-"d :1! ~“~ '. Y’:f\-“ ""Vaioc‘., ul or when the lithium concentration becomes equal to or slight- ly less than the acceptor (boron) concentration. The principal nucleation site in silicon is thought to be a substitutional oxygen site.35 Oxygen is one of the most abundant impurities in silicon, and also one of the most difficult to remove. It typically exists in the sili- con lattice at interstitial sites. These sites will readily capture a lithium ion to form Li+O complexes, but these are not the nucleation sites for the precipitation of lithium. The substitutional oxygen (08) is formed by the reversible reaction of an interstitial oxygen (Oi) with a lattice vacancy (V) as follows: [Oi] + [V] <=> [OS] . 3.1 The corresponding equilibrium mass—action relation can be written as [O ] N ._§____ : exp (ASb/k) exp (—AHb/kT) = K = equilibrium 3.2 [OiJEV] constant where Ash is the change in entrOpy of the oxygen, AHb is the binding enthalpy of the vacancy, and N is the number of lattice sites/cm3. The concentration of vacancies has been given by Logan37 for the condition of thermal equilibrium to be [V] = 1.3x102” exp(-2.0/kT) cm'3. 3.3 rt. u... by ass ‘15.”: 9 J , to L: ' e“ual 'S \4 . II (A) (U N be htend on t. 3”9C‘7:>itation :0”. ‘h‘. o . ”a E Foximateu ‘J: ‘ “‘0' er “3k Snap . “ 7.7 “leg follows . W . I :5383 akl Dre 152a,, d‘il‘in ‘ #2 Then, by assuming that the number of nucleation sites N5 is equal to [08], it is seen that NS = 39 [O] exp(ASb/k) exp(-(Hb+2.0)/kT) cm.3 3.u Interstitial lithium then tends to congregate in a spheroid about the substitutional oxygen site with concentrations near ten atomic percent of lithium. Other shapes are possible for the precipitation center and the rate at which precipitation occurs can be shown to depend on the shape of these sites. For the spheroidal precipitation particles, the kinetics of precipitation is approximated by C-Cf n C-Cf E—T: exp (“t/T) n = 3/2 for m- > 0 5 , o f o f o-cf n - l for fit— < 0.5 , 3.5 o f where CO is the high-temperature (diffusion-temperature) concentration of lithium, Cf is the final annealing-tempera— ture concentration of lithium, and O is the average concen- tration of lithium at time t. In the case of precipitation particles having the shape of long rods growing only radially, or disk shapes only growing thicker, the precipitation kin— etics follows a simple exponential decay with n = 1. In any case, all precipitation kinetics follow a simple exponential decay during the latter phase of the process. Measured .regioitat ion re :13". the none I“ "‘Dflrg“ -....-..is:. occurs 5. ‘F‘An \ «w...» :ncto ; “:A :c , ‘ . .er toe 54. .__ : : «“3 . Y‘N‘, shl‘y n‘t‘!’ u3 precipitation rates35 and optical observations35 of these sites in silicon and germanium confirm that most of the lithium precipitation occurs on spherical sites. Further- more, the rate constant 1 governing the latter phase of precipitation has been measured to be inversely proportional to the 2/3 power of oxygen concentration. This is consistent with the number of precipitation sites being directly pro- portional to the oxygen content, and independent of the initial high-temperature lithium concentration. An additional but relatively minor precipitation mechanism occurs in silicon but not in germanium. The introduction of lithium into silicon causes a high density of stacking faults which are not found in the original silicon. Photographic evidence exist83§.indicating the precipitation of lithium onto these faults. These stack- ing faults, although not important in precipitation kinetics, may reduce considerably the stress which can be applied to such a sample without causing fracturing. After the lithium diffusion, our samples were usually uniformly n-type with low resistivity; they then required annealing to precipitate the excess lithium. The resistivity as a function of time was used initially to monitor the precipitation process. This resistivity measurement did not require an accurate measurement of resistivity, since the detection of a maximum of resistivity was sufficient. Hence, an ohmmeter was used to indicate relative changes in resistivity. The n-type sample was initially prepared for Q N. .Y. r . V a ‘1 . :l - H as Q s. :l S "I- a V V. S .C nu nu uh no nu nu ”a Au C . .l n. .h“ + . k1. 9 Fl kJ Po Q. h“ no t. .1; +9 who Wu 9 .‘l ab .C .1. Pu Ad 03 .1 A nun :1. 0.. n C... d O. .1; + . U. fiv AZ Do in“ A: a: F. ., x n) o .. n » Dc .1; A: .3 Wm .1 r“ . . r . .3 n . ..:v a: A.» .. . «I 2‘ .n. O. h. n. .o . p N. o . no: flu .,...m I . n v .he g to 38L; “lit“ M I In me 5‘5 + TL +. -o n an. .ant , cool .he P-ffh yu'... D on 'u f‘- -... ‘ ! qq .,..N ‘0th at f “nr~¢ vv“ 7:3“ FEB «a an the precipitation by lapping and cleaning in methanol. It was then placed in point-contact pressure clamps and placed in 162°C mineral oil. At this temperature the diffusion constant for lithium in silicon is about equal to the diffusion constant of lithium in germanium at BHOC. Since lithium was known to take several hours to precipi- tate in germanium at 3u°C, it was expected that resistance variations over five-minute intervals would be small. We observed, however, that within the first minute after im- mersion into the hot mineral oil, the resistance increased from 289 ohms at room temperature to 890 ohms; it then de- creased steadily to 7M6 ohms at four minutes into the ex- periment. The resistance then increased quite rapidly to 1,100 ohms at 6.5 minutes, where it remained relatively constant. The resistance did not change significantly upon cooling to room temperature. Charge-carrier type was then measured with the hot-point probe, and it was found to be p-type. This effect was not expected to be a bulk effect so the surfaces of the sample were lapped to remove approximately one mil and the carrier type was checked again. The bulk of the sample was still n-type without significant change in the resistivity. It is likely that the lithium was congregating on the high surface con- centration of nucleation sites. The effective density of lithium near the surface would then be considerably reduced. This results in a lithium concentration gradient, hence the out-diffusion of lithium. The depth of this surface +- rob ~i.,:oar§t a» My 'Jnub “ =s:ir.g to com: a. in” .xah . 1 . ‘ 1.“; tom, 8510.1-C ”V” 'r,‘ L..- FESIDL I I O u , 4- ‘- ::;1.a.. n pnas L “-3339 sample w I O O :n :zneral 011 a 'V‘ ‘k a. '515 the hot-p0 111:1]. the re s is 1:9 began to C 4' A Amp-PC; 3042‘ CE 0n the V" we 3;». “title ' us effect would then be determined by the lithium diffusion constant at the annealing temperature. It would be inter— esting to compare this effect in a similar sample which had been etched previous to annealing. This should reduce the number of defect nucleation sites on the surface which, in turn, should reduce the rate of out-diffusion. The resistivity of the bulk material during the pre— cipitation phase was measured in a multistep process. The n-type sample would first be heated for five to ten minutes in mineral oil at 200°C. It was then cooled, cleaned and lapped with #BOO-grit carborundum to remove surface effects. The resistivity was then measured with a four-point probe and recorded. Finally, the charge—carrier type was measured with the hot-point probe. This entire process was repeated until the resistivity became very large, or until the carrier type began to change. A four-point probe was now needed, in order to elimin- ate differences in contact potential and surface effects from one measurement to the next. A switch was provided on the four-point probe to reverse the polarity of the current source on the two outer probes. If the voltage reading had the same magnitude after switching the current source polarity, then contact potentials were not a problem. Further checks on surface effects were provided by varying the source current. Since the resistivity is independent of current, the measured value should not change with current unless surface problems occur. For the actual determination of resistivity, the ;. 0 . . :-~rcxmatlon -. JH' ' fl ,5 cw1 -:ascna3.v accl. this samples, a Shari“ w..-..Ctlng Surf ( (7‘ (D H- *1 c;- MUS ~ ‘ L .00., l_‘/4 L. h we‘r‘he .v‘ Y‘ ‘ear‘n: A*:~ ‘L—x ‘~ ~ ‘V‘:*\l+ * .4 A. kO ‘En V. .‘j:k ““1351“ M \Jr“. “ ‘ "n, \ . fin‘ , 5:1qu MB approximation for a thin infinite plane was used. This is reasonably accurate for measurement at more than 5d from the sample edges, where d is the spacing between probes. The resistivity p, in the case of all edges being far from the point of measurement, is given by p = 2ndV/I. For our thin samples, a correction factor based on the ratio of sample thickness w to probe spacing d is needed. The four-point probe used was a Dumas model G-25-C with a point spacing of .025" and loaded for 25 grams of contact force. For samples of one millimeter thickness, w/d is equal to 38 of 0.85, assuming non- 1.57, giving a correction factor conducting surfaces. Sample resistivities were found to increase monoton- ically, eventually becoming too high to be measured on our simple four—point probe system. Since we could make reason- able measurements of 0 up to 100 ohm-cm, the compensation for the best samples was comparable within one part in 105 or possibly better. Ohmmeter checks showed maximum resis- tance to be in the neighborhood of one megohm. The total precipitation time at 200°C generally was 15 to 30 minutes. One of the best compensated samples was used in our experi- mental determination of temperature dependences of line- width and centroid. Unfortunately, not all samples were this good. Sometimes a sample would begin to turn p-type before reaching a very high value of p. This change is thought to be the result of non-uniformities in the lithium diffusion. If the compensation was better than one part in 100, these samples were still useful. .1 Na 3 v -n 5.- ' aDJCC+1 D wan ”a .1 . S |\.er..nc ‘ ‘ ' n .. u 1». > D .AvaV .Aw 14 «6(4. My 0 AL by W“ H» .p C n. +L e A: ”a A: ”a 7.: O #L ”A H“ {L h. CL uJ Q m5 0 Van e v. a w. M . “x or. .r; nu n... A 5 . a 7 . «i. a: . i T C . a ... a. . . flh‘ . v, A m n4| x v ‘U iorx, O-E h T 91"de .sat vs . n ,, p-\ v :18 ”VD u. . '¢ 55"”. -atelv 1+7 The primary spectral background in the thermal studies experiment was an effect similar to the surface precipitation phenomenon described earlier. As the temperature was raised, lithium diffused toward the surfaces, leaving behind excess "holes" associated with the uncompensated boron. Since this effect was limited to a thin surface layer, the spectral signal was not totally blocked out, but instead was modulated with a free carrier type of absorption background. This background could, however, be readily subtracted out. The germanium samples were precipitated at room temp- erature for convenience, hence several hours were required. To prevent the precipitation process from continuing past compensation, it was necessary to cool the sample immed- iately to -60°C or lower. Consequently, precipitation in the germanium was monitored by measuring the intensity of the spectrophotometer signal. The sample was mounted on a dewar cold finger in the spectrometer; when the intensity stOpped increasing with time, the dewar was filled with liquid nitrogen. Samples, when not in use, were stored at -90°C to prevent further precipitation. a.” .. .¢ . 2 C_ r: s . c n.“ D» “U .V‘ A I“ QI.‘ ”1.“ S f; + » At; O n.. a S P O r e h. 0“ .3 to n. a I. S _ n.“ a . . .3 .nu I . . . 2. 3: D» r. n“ 2‘ o. a .pH 2‘ H . q 5. .5.“ . . .v a .. In. ".51.!“ .A' c .OI‘E Tectru: Ms .h A.[ L ¢ A O ..AD CHAPTER IV IN STRUMEI‘ITATION H.l The SpectrOphotometer Much of the initial exploratory work was done on a modified Perkin Elmer model 1126 grating spectrOphotometer. The original 1120 spectrometer consisted of three distinct units; the source section, the fore—prism, and a double- pass grating monochromator. It was designed primarily for high-resolution work over a narrow wavelength region. For the narrow spectral—scanning ranges, the optical background variations were small, hence the single beam, without back— ground cancellation, was sufficient. However, the local- mode spectrum of interest covers a much wider spectral range and, furthermore, does not require the very high resolution provided. It was apparent, then, that some major modifications were in order. u.l.l The Fore-Prism The fore-prism was intended to act as a sharply-tuned band-pass filter with a pass band on the order of two microns. Its function was to eliminate higher orders passed by the grating. This was done by passing the signal back and forth through a prism of KBr. The angle at which the signal struck the prism could be varied by means of an external drum; in this manner the central band-pass frequency could H8 A n 1 u .. an. 1 .V -' C.~..lals, .1 w, . l 'I' :3: Sulr‘e LO ‘ V ‘ ‘ .33- l -:‘.CC€ S‘ :‘7‘. on. ‘I _ n -.-. ‘--‘ ‘ . ‘Abe Isa; . .- L. mirror «911 A - We is we 3‘ 233 Yr hat-sf: ET .‘I‘ a Q N -. ..,_ {::~ *- ‘L 17'3‘ ‘sIQ. I? “\“')“:‘I ‘Q (a, P . ‘v u9 be varied. To avoid a large background effect from the fore—prism section, it would be necessary to change the fore—prism drum setting for nearly each micron scanned. Instead, the fore-prism section was removed and replaced with a set of three filtersfi covering the ranges 10.2 - 18u, lu.3 - 2H.3p, and 2u.3 - Now. The filters were con- structed by deposition of thin films of selected filter materials;39 they were consequently degradable by prolonged exposure to atmospheric water vapor. Most of the silicon local-mode spectrum could be covered with the lu- to-2up filter. The remainder of the spectrum was covered by the 10.2- to- 18u filter. The 2u-to—u0u filter allowed us to look for similar local modes in germanium. An additional benefit from the removal of the fore-prism section was a two-to three-fold gain in signal level. The increase was due to removal of the KBr prism, a decrease in the optical path length, and to lower atmospheric absorption and scatter- ing. u.1.2 The Source The source section, shown in Figure n.1, includes simply a black-body radiation source, viz., a Globar, a flat mirror m and a spherical mirror m2. The Globar 1’ source is heated to about 1600°K by an electrical power of 280 watts. The resulting radiated power is given by Planck's black-body radiation formula,”0 "The filters were models number 221-1790, 221-2009, and 221-2010 obtained from Perkin Elmer, Inc. mmaaoxo in: N III. N L m zoEomm momoom EOF 533m 39:.ch v.0 manor. 00m 00m 09 a. 02 £9: 22. D 25:23.3 III 28 EEoEtXxo l “NEVTEO ¢mw 3...: 18 3.8 it _.Eo «mm q _ aoonmwm u _~o_x~0mu~..u_ u . In rso\o§_~o_x¢_ «.2... u / / .2: D l .3; _U 1000 i000 I~_m l¢;n [0.0 1| 0.0 town INNm I¢Nm Immm com 3:: 03. 8369353 .m> £33 6:3 0.0 mmDmu—L 00m 00m 00. . _ a A \ . \1» H _ _ d \\.—\\\\ \ER tug \ \ \\ \\ \ \ M \ \ LU \ \\ n «U in a 9.6.5 flea N— .soxos..o_ .53 ... .em on 3:323 89¢ 23 D meo:o_:o_3 Ill 23 .quEtoqxo TEo ¢mw .Eo mfwm the NNm x0< 76 earlier, the free-carrier absorption forms a relatively smooth background and can be readily subtracted out. Another background effect has been observed to occur at the surface of the sample. As the measurement temperature approaches 500°K, the surfaces tend to become p-type due to a loss of lithium compensation. This surface effect is the same effect observed in the precipitation process. Since the effect occurs in such a thin surface layer, less than one mil in depth, the sample remains transparent to the infrared but has a free-carrier type background due to the p—type surfaces. A third possible background effect is the thermal breaking of Li—B complexes, causing an increase in the free-boron local mode and subsequent decreases in the local modes of the complexes. Calculations show a maximum effect of about 22% at 500°K and doping levels of lOlg/cm3. Calculations are given in Appendix B for the relative number of pairs as a function of temper- ature. This unpairing will not affect the local—mode centroid, but the decrease in amplitude will show up as a loss of accuracy in the line width. Finally, the background can be shifted by infrared emittance from the heated silicon; this effect is not expected to alter the local-mode spectrum. Immediately after the temperature data was recorded, the sample was cooled to 351°K and the spectrum was recorded. The resulting background had a similar appearance to the background at 082°K. This background was the result of the non-reversible loss of lithium compensation at the surfaces. 77 The surfaces were later checked with the hot—point probe and found to be p-type. Removal of about one mil of surface by lapping restored the sample to nearly its previous level of compensation. The line—width and centroids of this final spectrum fit the plot of the rest of the data well within the limits of error, indicating that background effects had been effectively subtracted out. The total time at “82° K, the highest temperature, was only 16 minutes, so that the total surface effect was not large enough to prevent a good measurement. A plot of time vs. sample temperature is shown in Fig. 5.6. As a further control on background effects, a spectrum of pure silicon was run in the same temperature range. Calculations of a background normalization factor based on the ratio of the doped-silicon-lattice peak centroid divided by the centroid shift of the pure—silicon-lattice peak showed no appreciable lepe with temperature, and therefore were not incorporated into the data. Immediately prior to recording the temperature data, but with the same spectrometer settings and resolution, a spectrum of indene was run. This gave us a calibration of the spectrometer and showed the need of adding 1 cm-1 to the wavenumber readings. The area under the local-mode peaks was observed to be very sensitive to background effects. For this reason we adopted a normalization factor which is the product of two terms. The first term is the ratio of the pure-lattice peak area to that of the doped-lattice peak area. The . oLSPMAoQEow magnum . m> oEHB m .m mmDmv—u 12;: as on. ow. o: 00. cm om ox. om om oc on ow o. o 78 _ _ _ _ _ _ _ _ _ a _ A _ fl com .I 00m .1 00¢ I... 000 AXOV 1. cow 79 second term is the effect of thermal breaking of the com- plexes calculated in Appendix B. The normalized local—mode peak areas and normalization factors are shown in Fig. 5.7. 5.2.2 The Pressure-Dependence Measurements The pressure measurements were made on the same spec- trometer (P 8 E 225) with a resolution of 0.8 cm-l. The pressure applicator is sketched in Fig. 5.8. The associated pressure meter was calibrated to an accuracy of 18%. A slightly compressible backing, such as bakelite, was needed to allow a more uniform distribution of the load across the sample.* The pressure gauge on our stress-applicator device was calibrated with two strain gauges attached to the sides of a small aluminum cylinder of about 1/2" in diameter and one inch long. The two strain gauges were connected in OppOSite arms of a four-arm bridge. The cylinder was first loaded with 5-pound weights on the end of a lS-to—l mechanical- advantage lever system. A plot of pounds loaded on the aluminum cylinder vs. bridge reading was then made for loading and unloading of the cylinder. The aluminum cyl- inder was placed in the pressure-applicator device, and several plots of the pressure gauge reading in psi vs. the bridge readings were made. A sizable hysteresis was ob- served between loading and unloading of the cylinder. This hysteresis accounted for the 18% error in the pressure deter- “Pieces of bakelite l/2" thick were necessary to prevent their breaking under pressure. 80 X 522. cm 654 cm“ Normahmtion tact r for ll 10 Tem I’Kl N SITE) \.0i 35\ \.0‘5 393 \.\2 444 \.52 2 2.60 b m R‘G‘ 300 ___><___- ____><_ _ _><— V FIGURE 81 Sampm HI Pressure Meter(psi) Hydraulic Jack 5.8 The pressure applicator. 82 mination. Both the loading and unloading plots were found to be linear. From the above measurements a plot of actual pounds of applied force vs. the pressure-gauge reading was made. This is shown in Fig. 5.9. A scale of kg/mm2 is also shown for a sample cross—sectional area of 10 mm2, which was in the range of most of our experiments. Our spectroscopic pressure experiments did not give us any conclusive results on line width or centroid vari- ations. Centroids were found to be constant with stress perpendicular to the [111] direction or parallel to the [112] direction. The line width variations were inconclu- sive. A plot of line-width vs. pressure is shown in Fig. 5.10 to illustrate this. The maximum pressure that we ob- tained in the [112] direction was 1.22x109 dynes/cm2 with unpolarized radiation propagating in the [111] direction. The sample was 10.5 mm in the [112] direction with a cross section of 9.65 by 0.79 mm. Another experiment was done with the applied pressure perpendicular to the [111] dir- ection and a maximum value of 0.9x109 dynes/cm2 was ob— tained. The sample dimensions were 15.0 mm in the dir- ection of pressure with a cross section of 12.0 by 1.09 mm. In each case the experiment concluded with the fracturing of the sample. 83 0_ 0. 0m mN samHV mmsmm mesmmmpa mo cowpmeofiamo GA.“ mmDmv—L OOm 00h 000 000 00¢ 00m OON OO. _ _ _ a _ _ _ _ :2: 333m 6960 meammoc no.8 62.36 *0 mucaom I 00. 00m 00m 00¢ 000 000 81+ .ocgwmmca .m> zpowz ocaq O_.m MEDOE _ _ A _ _ _ a mo. x $583580 3 av .v , .v .. H 7:3 ¢mw AV .5 33 CE “mg—Dot Jim _ 0 5:13 IVE]: 7.3.8... x 0. CHAPTER VI DISCUSSION OF LOCAL MODES 6.1 Local-Mode Assignments in Silicon Before discussing the results of our local-mode meas— urements, we review the assignments of the local modes in the silicon system. A good discussion of these assignments 9b This paper includes a study of is in a paper by Spitzer. isotOpic shifts in the local—mode frequency and a measure- ment of local—mode intensities as a function of the equal concentrations of lithium and boron. The free-substitutional- boron local mode had already been observed in silicon by Angress et al.;20 no other constituents had been added to the silicon. Spitzer96 observed this free-boron peak in reduced intensity at 620 cm-1 for 11B. This is between the two boron modes of the lithium-boron complexes, and is closer to the doubly-degenerate 65H cm“l peak. When the boron isotope was changed from 11 to 10, the 56H cm—l peak shifted to 58A cm”1 and the 65“ cm-1 peak shifted to 681 cm—1. When the lithium isotope was changed, these peaks shifted less than 3 cm-l. This identified these peaks as boron local modes. Their intensity was found to be directly proportional to the concentration of lithium-boron pairs. The degeneracy of the upper band was assigned as two-fold 85 86 degenerate, and the lower band as one-fold degenerate, on the basis of relative integrated intensities of 5 to 3. Axial symmetry of a paired configuration predicts an in- tensity ratio of 2 to 1, but local polarization due to atomic distortion may account for the difference.* The 522 cm“1 line was found by Spitzer to shift to 534 cm_1 when 6Li was substituted for 7Li. It is insensitive to changes in the boron isotope, and appears at 515 cm"1 for gallium- doped 7Li compensated silicon. The intensity of this mode is also directly prOportional to the number of lithium-boron pairs. The isotopic shifts and intensity ratio identified the 522 cm"1 peak as a lithium mode. By analogy with the two boron local modes, the 522 cm_1 peak is expected to be the two—fold degenerate peak of a paired lithium ion in a C3v configuration. Further studies on the breaking of the complexes with temperature have been done by Spitzer and Waldner.9a They found that as the temperature was raised to 150°C, the peaks assigned to lithium—boron complexes de- creased in intensity whereas the free-boron peak increased in intensity. This is in agreement with predictions on the thermal breaking of complexes given by Reiss et al.22 This comparison between experimental and theoretical thermal effects further reduced the possibility of the 522 cm"1 peak belonging to the free-interstitial lithium since such a peak would also have increased in intensity as the sample temper- ‘0 "For example, carbon impurities in silicon are substitutional and un-ionized but they have an effective local charge of about e.“7 87 ature was raised. The experiment also confirmed the identity of the 65A and 56” cm_1 as boron paired to lithium. The temperature at which the lithium is diffused into the silicon does not significantly alter the intensity of the observed local modes. Therefore, it is unlikely that the 522 cm"1 peak could be a lithium-vacancy complex local mode. The small but finite shift in frequency of the 522 cm-1 peak for different acceptor impurities confirms the designation of the 522 cm-1 peak as a paired lithium local mode. The free interstitial lithium must then be a resonance mode within the band of allowed lattice frequencies. We did not observe any such resonance mode in our samples. 6.2 The Lithium-Gallium Complexes in Germanium Our attempted measurements of lithium-gallium complexes in germanium were unsuccessful because of problems in sample preparation. However, some observations on work reported by A. E. Cosand33 will be made. The two local modes of lithium in a paired configuration have apparently been observed by Cosand33 in germanium at liquid-nitrogen temperatures. The lithium is paired to gallium, and 7Li local modes occur at 356 and 380 cm-l.“ When 6Li was used, these peaks occurred at 379 and HOS cm”l reSpectively. Cosand suggests, on the basis of integrated intensities, that the lower—frequency peak is the two-fold degenerate mode. His suggestion indicates that a free- .7 "The maximum single phonon lattice frequency in germanium is 300 cm-1. 88 lithium local mode does exist in germanium and should be between the two lithium complex local modes close to 36” cm-l. It was not observed in Cosand's samples because the solubility of lithium in germanium limited the amount of lithium; in fact, sample preparation was the major problem in this work. The complex local modes had only a 2 to H cm"l absorption coefficient, and at the liquid-nitrogen tempera- ture over 99.9% of the lithium was paired to acceptors. The location of the free-lithium local mode might be confirmed by slowly heating such a sample to break the complexes. How- ever, it appears that thermal broadening may mask such an effect. 6.3 Temperature Dependences of Centroids and Line Widths in Silicon Figures 5.5 and 5.6 show that thermal effects are most pronounced above room temperature.* Centroids are observed to shift to lower wave numbers with temperature, as expected. The H82°K data point in the centroid shift of the 522 cm"1 line seems to represent a rather sudden departure from the linear approximation. It is at about this temperature that the centroid has been shifted to the edge of the lattice- band modes. Therefore, the effect of the lattice phonons on the centroid appears to account for the deviation from linearity of the M82°K point of this line. The data on centroid shifts can be fitted with a linear approximation 7. "Low-temperature measurements of line width and centroids have be n done with :lcm‘l resolution by Balkanski and Naz- arewicz b 89 by lines varying in slope by 110% from the lines shown and still fall within the limits of the error bars. Variations in the fit to line-width data of :l5% in slope are within the corresponding error bar limits. The paired substitutional-boron and interstitial-lithium ions are each in essentially a perturbed Td symmetry. The resulting symmetry is C , but the dynamics of the system 3v are quite complex; the theory necessary to explain the centroid shifts and line-width variations with temperature has not as yet been fully developed. It is expected, however, that such a theory will take the form of the effects of an- harmonic terms in atomic displacement described by Elliott et al.25 for an isolated impurity, applied to the Green's function treatment of an impurity pair by Elliott and Pfeuty.7 A preliminary review of these ideas showed that such a cor- rectly done formalism was likely to result in third—and fourth- order Hamiltonians having the same form as those presented by Elliott et al. The interpretation, however, would be quite different, i.e. the local-mode operators must be re- labeled to apply to each local mode observed. The relabeling should also include a distinction between the possible one and two-fold degeneracies. Details of how such a relabeling of the local-mode Operators occurs, and what the relabeling means, are given in Appendix D. In this new representation, the two-fold-degenerate boron local—mode operator is subscript- 2 3 is labeled with F ed with F , and the singly—degenerate boron local-mode Operator 2 l . The 522 cm.1 lithium local-mode operator 90 is labeled with F31. The superscripts l and 2 refer to lithium and boron respectively; F1 and F3 refer to the singly and doubly degenerate infrared-active representations of the CBV symmetry. With this interpretation and relabeling of local-mode operators, equation 2.12 approximates the centroid shifts. Since our own measurements did not determine the zero-temp- erature local-mode frequencies 90, this parameter was varied to obtain a best fit to our data. Equation 2.12 is repro- duced here for convenience with the indicated labeling of the local modes. AQF i rr i _._l.. = 1 , {.1— Z’hm(k)(n(k)+1/2) 6.1 :2 l Mlm light) 2 N -> on 'Fj M k Evaluation of this expression requires a knowledge of the silicon phonon spectrum g(w). This has been evaluated numerically by E. A. Johnson (unpublished) for a mesh of 61 points on the interval Oiwib A modification giving M' closer agreement with experiments has been presented by Dawber and Elliott.19b These calculations represent one direction in the Brillouin, zone and all other directions are assumed to give the same density of states. The bracketed term of eq. 6.1 then becomes ”M y g(m)’fiw[n(w)+l/2]ng_ 6 2 (1) o M f g(w)dw O 91 where the number operator n(w) is given by l/(exp(hw/kT)-l). The factor of 6 occurs because the original sum over k in— cluded the sum over the 3 optical and 3 accoustical lattice dispersion branches. This expression can be easily pro- grammed to give the full range of temperature variations. The fortran program and further details are given in Appen- dix E along with the values of the silicon phonon spectrum taken from Dawber and Elliott's paper. The expansion coefficients eri and 90 were evaluated by requiring that the calculations agree in both magnitude and slope with the data. The results of these calculations are shown as dashed lines in Fig. 5.9. Calculations in the Debye approximation, for which g(w) = 3w2/wD3 *, have also been done. The results in all cases were very close to the results based on the silicon phonon spectrum. Agreement with our experimental results is very good for the 522 cm"1 lithium mode, and reasonably good for the two boron modes at 6H5 and 565 cm_l. The resulting expansion coefficients eri and zero-temperature local-mode frequencies 90 are summarized in Table 1. These calculations have not included such possible effects as the interaction between the various local modes. It is expected that the interaction between the two boron local modes would be the largest of such interactions be- .7 “The Debye frequency for silicon is ”57 cm.1 for a Debye temperature of 6580K. Line-Width Coefficients for Two—Phonon Decay TABLE 1 21 * l Local c (erg/emu) x 10 Q (cm_ ) F . 0 Mode 3 522 cm‘1 a l = 2.1a 525 r 3 55u.5 cm'l c 2 = 3.52 557.9 I‘ 1 65a cm'l gr 2 = 5.00 557.7 3 93 cause both modes are at the same physical location. Inter- action between the lithium mode and a boron mode would be a smaller effect because the ions are separated by 2.5 A. These effects would then be small for the 522 cm"1 line, possibly accounting for the good agreement of the data with the simplified theory. Similarly, such effects could account for some Of the disagreement between the data and the calculations for the boron local modes. The line-width processes also arise from the relabeled local-mode Operators described in Appendix D. Again, although the interpretation is different, the expressions for line width take the form of eq. 2.19 for the two-phonon decay process, and eq. 2.15 for the elastic-scattering process. To evaluate eq. 2.1M as a function of temperature, we first converted the sum over k’ to an integral over w’ making use of the phonon density Of states ng). The delta function allows this integral to be easily evaluated as the value of the integrand at w’ = eri-w. For example, ng) becomes g(9p.i -w). The sum over k is then converted to an integral 3 over w by making use of g(w). Equation 2.1M then becomes i 2 i 1 _ (231:3. > rm “M gummy]. -w) i .. .. if 36 — 03(521" . -w) T uM’Q l M2w 2 N 2 3 rj M o P X[(n(w)+l)(n(flpjl-w)+l)-n(w)n(9rjl-w)]dm . 6.3 The phonon density functions g(w) and g(erl-w) have been 9H 6 normalized so thatjf.h(g(w)/Np)dw = 1. Since g(iji-w) is zero for frequencigs below eri-wl’ the lower limit on the integral in eq. 6.3 can be replaced by (eri-mm). The program for the temperature dependence of eq. 6.3 is given in Appendix E. The elastic-scattering process has the form of eq. 2.15. The sum over k and k’ can be replaced by the following integral: “’I '2 ( ) 2 2 36/ (%—Ui—) w'n(w)[n(w)+l]dm 6.1+ {JP 0 The fortran program for elastic—scattering temperature- dependent line width is also given in Appendix E. When the parameter iji - ZCfifji)2/erji is evaluated to give agreement with the magnitude of the experimental results, the calculated line width is found to cross the experimental data sharply. The elastic-scattering mechanism is therefore considered an unlikely explanation of the observed line- width variations. The results of the two-phonon-decay calculations are shown as dashed lines on PigureSJi The expansion parameter .S’rji was fitted to give reasonable agreement with the magnitude of the experimental results. The calculations give approximate agreement with the measurements at tem- peratures above 3000K. However, agreement is poor below 300°K. As mentioned earlier, there are likely to be inter- actions between the two boron local modes. This would give a line—width contribution to the 65M and 56h cm_l lines proportional to AQ[n(AQ)+l][g(AQ)/NP][n(Q’)+l] , where Q’is the frequency Of the boron local mode which is not decaying and A9 is the difference between the two boron local—mode frequencies. This expression has a T2 dependence at high temperatures. However, the effects of the thermal population term [n(Q’)+l] are small below about 9500K or most of our temperature range of measurements. As a result the variations appear nearly linear in temperature between 300 and “500K. This temperature dependence is similar to that of the two-phonon-decay calculations and therefore cannot account for the large low-temperature line widths. Accurate prediction of the observed line widths will have to wait for a formal treatment of the problem based on the anharmonic factors of the actual symmetry of the lithium-boron pairs. 6.” Temperature Dependence Of the Integrated Intensity The relative integrated absorption intensity I/Io is shown on FfiynwaSJ as a plot of (1.0 - I/IO) vs. temperature. In this form, the dependence is linear on a log-log plot. The integrated intensity has incorporated into it the cor- rection factor mentioned in section (5.2.1). The relative integrated absorption intensity was evaluated by first plotting I vs. T on a linear scale. The zero-temperature 96 value IO was then extrapolated for each set of local-mode data. The dependence Of (1.0 - I/IO) is approximately Tn where n = u.2 through 5.1. This plot indicates that no local modes exist above a temperature Of about 5300K. This temperature dependence is, however, the least accurate of our data because of the correction factors involved, and because Of some uncertainty in IQ. Nevertheless, we do believe that a non-linear decrease in integrated intensity occurs 0 6.5 The Uniaxial-Stress Results The results of the uniaxial-stress measurements were inconclusive. The centroid shifts were constant for stress applied in the [112] direction and perpendicular to the [111] direction. A similar experiment by Hayes and MacDonald21 on H2 and D- ions in CaE2 had shown a linear variation of singly-degenerate local-mode centroids with uniaxial stress, and a splitting of multiply-degenerate modes. Line-width variations were expected to be of a similar order of magni- tude. It now appears that not enough stress can be applied to the lattice to shift the lithium—boron local-mode frequency by a measurable amount, i.e. more than 0.8 cm—l. The limit- ing factor is, of course, the fracturing of the samples. Evidence supporting this conclusion can be found in a paper by Hayes.“8 Hayes has applied uniaxial stress to oxygen doped silicon with a resonance mode due to oxygen at 29.3 cm-l. At stresses near the breaking point (2% - 30 kg/mm2) he Observed a frequency shift of only one cm-l. This small 97 a variation in the lithium—boron local modes would remain essentially undetected by our system. On the other hand, if the frequency shift scales with the impurity vibrational frequency, the results Of uniaxial stress should have been easily detected and the fact that we Observed no shift would be paradoxial. It appears that uniaxial-stress effects on Li-B local modes will require a high-resolution spectrometer (.2 cm"1 or better). The degeneracy of the local modes is unlikely to be determined by infrared-absorption tech- niques since the amount of splitting, near the fracture load, is much less than the local-mode line width. CHAPTER 7 ISOTOPIC DIFFUSION STUDIES Our isotopic studies on the mobility of lithium in germanium were initiated by a desire to obtain a faster drift rate of lithium in the fabrication Of nuclear "charged-particle" detectors. 6Li will drift faster than 7Li because mobility is inversely proportional to the square root of the ion mass. Measurements on the ratio of diffusion constantsfi were made primarily by means of a diffusion type measurement, although some effort was also put into a drift technique. Both techniques required correction factors in the evaluation of the magnitude of the diffusion constant. The diffusion constant of lithium in germanium, at temperatures in the neighborhood of 50°C, showed about 15% smaller lepe than at higher temperatures. In the classical picture this means that the activation energy has decreased by 15% with temperature. This suggests the possibility of a tunneling effect's becoming important at the low-drift temperatures. Such a tunneling effect would further increase the mobility Of 6Li compared to that of 7Li. The ciiffusion constant of 7Li has already been measured in the low-temperature range11 so that only a ratio of diffusion constants needed to be measured. “Extensive use has been made of the Einstein relationship between diffusion constant and mobility, applicable to the "dilute semiconductor solutions'" it is given by u = eD/kT. 98 99 7.1 Reasons for Measuring the Mobility Ratio Since an absolute magnitude of the diffusion constant was not required, the measurement was simplified for several reasons: First, a measurement of the diffusion constant would have required a determination of the internal enhance- ment field across the junction at the diffusion front. This effect tends to increase the diffusion rate by a factor of [1+ [2ni(T)/N)2+l]-l/2], where ni(T) is the density of in- trinsic carriers.ug Thus an accurate knowledge of the donor concentration N would be required. This, in turn, would re- quire an accurate determination of the diffusion temperature and some means of assuring its stability. Secondly, silicon and germanium are difficult to Obtain in ultrapure form; there are typically 1013 to 1012 traps/cm3 (usually oxygen) remaining in the best crystals.* Consequently, during the drift or diffusion process the lithium tends to form pairs, thereby effectively reducing the diffusion rate. This effect can be neglected if the lithium concentration is large compared with the concentration of trapping sites, as is the case for diffusion temperatures above 150°C. Finally, a measurement based on a drift technique would require an additional correction factor to account for that component Of mobility due to diffusion. In the de- termination Of the ratio Of mobilities of 6Li to 7Li, all 5. I‘ It has recently become possible to produce germanium crystals with less than 1010 traps/cm3; such material does not require a lithium drift to compensate trap sites in the production of nuclear detectors.S :i‘J'TE-n .. L. 100 of the effects mentioned above are equal for both isotopes and therefore cancel. It is only necessary to insure that both samples are drifted or diffused at the same temperature. In an effort to assure equal drift temperatures, a drift experiment was done with both 6Li and 7Li deposited on the surface of the same crystal. All germanium was 13 gallium/cm3. detector-grade, doped with typically 5x10 Half Of one surface was coated with 6Li and the other half with 7Li, followed by a lO-minute diffusion at 900°C. The sample was then drifted at about 25°C. The results showed nearly equal drift depths for 6Li and 7Li regions. A closer study showed two mechanisms which tended to equalize the drift depths: First, the field across the junction depended upon the drift depth. If 6Li had drifted further than 7Li, the resulting decreased E field would reduce the 6Li drift rate. Also, the 12R heating would be larger in the region of shortest drift depth, tending to thermally increase the drift rate. Differences in drift depth were, therefore, quickly reduced. Further drift measurements were tried based on the technique of Fuller and Severins;lH however, the CuSOu stain test used to define the junction region did not give sufficient accuracy. All drift meas- urements were further complicated by variations in contact resistance. This effect was minimized by coating the sample contacts with a liquid mixture of indium and gallium. 101 7.2 The Ratio Measurement Most of our ratio measurements were done with a dif- fusion process. Two samples Of 12 Ohm-cm gallium-doped detector—grade germanium were lapped and cleaned. One sample was evaporated with 6Li, then placed in mineral oil while the next sample was evaporated with 7Li.* The two samples were then placed side-by-side in the preheated diffusion oven. An argon flow of 10,000-15,000 cm3/min. was maintained throughout the diffusion. After completing the diffusion, the samples were quenched to room temperature. Temperature gradients in the oven were the most likely source of error, since the two samples could consequently be at slightly different temperatures. Efforts were there- fore made to eliminate or determine any such temperature gradients. For example, one diffusion was done on a single crystal, one surface of which was half 6Li and half 7Li. Another diffusion was done with four pairs of samples, each pair consisting Of one 6Li and one 7Li specimen. Each sample was typically Ux12 mm of surface area and about 2mm thick. Thermal variations were found to be small, but a possible temperature gradient may have existed along the direction of argon flow. Several measurements on single pairs of samples were done at 200°C and at 900°C. Diffusion times at 200°C were typically on the order of two hours or more. "The lithium isotopes were 99.99% pure and Obtained from ORTEC. 102 This gave diffusion depths on the order of 8 mils or more. Two measurements at 200°C for six hours were tried in order to improve the accuracy by diffusing to a greater depth, i.e. about 1H mils. At uoooc the diffusion time was generally 20 minutes, and diffusion depths of typically 25 mils were measured. The results of these measurements are given in Table II. The ratio of diffusion constants was taken as the square of the ratio of diffusion depths, i.e. the diffusion depth x was approximated as x = 2/Df so that D is prOportional to x2. The correction factors belonging to this expression then cancelled in the ratio. An accurate measurement of x required that the surface of the sample, upon which lithium was to be evaporated, be parallel to its opposite surface. After some practice, we found that two surfaces could be made parallel to within 1 mil by successive grinding and checking for variations with calipers. Better accuracy was then obtained by using a sensitive pressure gauge.2 This gauge had a full-scale variation of 12.0 mils; each mil was divided into 10 parts, so 0.1 mil accuracy was possible. Variations in the sample thickness were measured with this gauge; if the two surfaces Were not parallel, one of them would be lapped with more Pressure on the thickest area. The sample was then cleaned and the uniformity of thickness checked again with the Pressure gauge. In this manner we were able to Obtain surfaces parallel to within 0.1 mil. More typical variations, M The pressure gauge was a style 1000 gauge made by the Sheffield Corporation with serial number 2201RS. 103 Table II - Experimental Results of IsotOpic Studies Diffusion Temperature 673°K Notes Diffusion Depth (mils) Ratio DB/D7 Time (min) 7 6Li I 2Li 31.51 i .3 ,30.31 i .15 1.050 1 .012 23 35.00 i .15 I30.55 1 .15 1.091 1 .005 29 25.00 i .3 :20.70 i .20 1.125 1 .015 20 Average : Ratio 1.091 1 .02 Diffusion Temperature 073°K Diff. exp 7.50 i . : 7.09 i .2 1.095 1 .09 120 7.50 i . ' 7.01 i .2 1.120 1 .09" 120 9.71 + .1 | 5.53 + .05 1.210 + .02 120 _ , _ _ 5.05 i . ; 5.09 i 0.1 1.102 1 .00. 120 5.05 i . , 7.59 1 .1 1.155 1 .05" 120 10.30 i .2 113.50 1 .1 1.105 1 .05 350 2 pairs 5.51 + .2 I 5.35 + .2 1.07 + .10 diffused — — _ together 9.15 i .2 , 5.91 i .2 1.05 i .10 120 0 pairs 5.01 i .15 ' 7.50 i .15 1.11 i .05 diffused 7.59 i .15 : 7.35 i .15 1.05 i .05 simultaneously 7.37 i .15 ' 6.68 i .15 1.22 i .08 120 5.55 i .15 t 7.00 i .15 .95 i .05 average 7.16 ‘ 7.08 1.09 i .08 Average I Ratio 1.096+ .06 , _ Drift Temperature 318°K Drift exp. 50.20 i .3 ' 1.110 1 .02 23 hrs. I50.00 i .3 "This is a repeat of the data on the above line with a correction factor to account for room temperature drift and diffusion. 100 however, were 0.2 or 0.3 mils (0.1 mil = 2.50 0). At this point the thickness of the sample was measured. This meas- urement was made with the pressure gauge and a set Of reference blocks which varied in thickness by one—mil steps. The pressure gauge was first adjusted to read zero on some com— bination of reference blocks whose thickness was near that Of the sample. The reference blocks were then removed and the sample was inserted under the pressure gauge. The gauge then measured the difference in thickness between the refer- ence blocks and the sample. For samples which were not flat within 0.2 mil, the thickness was recorded at the center and four corners of the sample. The sample was then beveled at one edge to assure its orientation. This procedure was followed for both samples in the experiment. The samples were then ready for the lithium deposition without further surface treatment. Following the diffusion, we generally observed a sur— face layer Of black material one mil thick or less. Measure- ments indicated 0.5 mil increase over the original thickness before diffusion. This thin black layer was probably a Ge-Li alloy with some residue from the mineral Oil burned Off during diffusion; it could be easily rubbed off with CH30H and a Q-tip. Measurements were taken from a zero diffusion depth at the cleaned surface. The cleaned surface often had small pits which were avoided in the measurement. Mineral Oil was used to coat the lithium surface follow— ing the evaporation. Also, the sample evaporated first was 105 stored in mineral Oil for the nearly one hour needed to per— form the second evaporation. The amount Of mineral Oil on the sample surfaces was reduced to a minimum immediately prior to placing the samples in the oven, in order to mini- mize the time needed for it to be burned Off. One possible source of error may be due to the mineral oil's burning Off one sample sooner than the other, hence allowing that sample to become slightly hotter than the other. In retro- spect, it appears that xylene should have been used instead of mineral oil for this reason. The measurement technique was similar to that used by Pratt and Friedmann51 and consisted of measuring the sample thickness and the conductivity of the lithium surface be- tween successive lappings of the lithium surface. Each lapping removed 0.3 to 0.5 mil of material; the sample was cleaned before each measurement with trichloroethylene. The conductivity was measured with the hot—point probe, which was found to give readings that were strongly depen- dent on the applied pressure. Consequently, a mechanical support was arranged to apply equal pressure on each measurement. The hot-point probe, as described earlier, was connected to a voltmeter to indicate the thermally- induced voltage. The location of the junction at the diffus— ion front required only monitoring variations in the thermal voltage, hence the actual conductivity was not determined. The thermal voltage was found to be constant away from the junction; it began to decrease at about 1 to 2 mils from 106 the junction. At that point, the amount of material ground Off after each measurement was reduced to 0.1 to 0.3 mil. Grinding and measuring were continued until the thermal vol- tage changed polarity. In this manner we were generally able to measure the diffusion depth to 0.1 mil. The measure- ment process generally took one-tO—two hours. If work was done on one sample until the measurement was complete, the sample with the other isotope would be left standing. During this time, lithium continued to drift through the crystal due to the built-in field of the junction. Measure- ment Of the diffusion depth in one crystal obviously does not stOp this drift effect in the other crystal. An esti— mate of the largest effect to be expected can be made by assuming a step junction. The internal junction voltage is then about V = 0.0 volt and the resulting drift depth is given by w = quit)1/2. At room temperature the mobility u is 3x10"10 cm2/volt-secll, and a drift depth Of 0.0 mil results in one hour. The actual junction is graded so that drift depths of 0.2 mil or less may be expected in one hour. Similarly, the diffusion would continue at room temperature and produce an additional .13 mil in the diffusion depth in one hour. Most of our measurements, with the exception of a few of the earlier ones, were done on the 6Li and 7Li samples simultaneously. That is, a thin layer was ground off the 6L1 sample, and the thickness and thermal voltage were recorded. The same was then done for the 7Li sample, and the measurements alternated back and forth until the 107 junction depth of both samples was recorded. The difference in time between the location of the two junction depths was consequently reduced to five minutes or less. Measurements were also performed on crystals diffused at 100°C; however, the solubility of lithium in germanium at 100°C is only about 2x1012/cm3. This is not significantly different from the original acceptor concentration. The hot- point-probe measurements therefore showed large variations in thermal voltage, and the junction could not be located within 1 mil. 7.3 Comparison of Experiment with Tunneling Calculations The 200°C and 000°C measurements did show variations from the classical prediction in the ratio of diffusion constants. This led us to try and fit the tunneling theory (eq. 2.5) to the existing data on 7L1 diffusion constant simultaneously with our own data on the ratio. To do this we first needed the distance (a) between adjacent lithium interstitial sites in germanium. The value finally arrived at was scaled from existing data on lithium localemodes in silicon.3b This left us with the barrier height 0 and the local-mode frequency as parameters to be scanned. The best fit of the diffusion constant was obtained for Q in the range 0.515 to 0.530 electron volt, and any value of local- mode frequency in the neighborhood of 1013 Hz. The classical activation energy (Q) given by Fuller and Sevrienslu was 0.065 ev; Q determined by the tunneling expression, (eq. 2.5) is expected to be larger because it includes the energy between 108 the bottom of the potential well and the ground state of the ion. The fit of the tunneling expression to 7Li diffusion data is shown in Figure 7.1. Error bars indicate the range of variation in the calculated values as Q varies from 0.515 to 0.530 ev. The local—mode frequency (v) was scanned for a constant ratio of 06 to v7 equal to 1.08. The range of frequencies scanned was determined from the coefficient D0 in the classical expression for the diffusion constant (see eq. 2.2). We found that any frequency in this range would give agreement with the experimental data on diffusion constants but only a discrete set Of frequencies would fit both the magnitude and ratio of diffusion constants. These results are shown as a plot of ratio vs. local-mode frequency (v) in Figure 7.2. The shaded area represents the error in our measurements of the ratio. The best fit of the calcu- lations to all the data indicates a ratio (D6/D7) = 1.12 at room temperature. This effect is considerably smaller than hoped for, and represents a savings Of only 10 hours in a ten-day detector drift. 109 (°K) IOOO 700 500 400 350 300 250 I T I l I I I I I (9312) Solid curve I: tunnellng sec calculation Q= 5231.007“ '0-6 V7 = .494 XIO 233°C-| — .7 '— IO r- :0"8 — " I0'9 - I ' Fuller and Severiens '0 Physical Review 96, 21 _‘ m" r- 1950 X A. H. Sher Journal Of Applied Physics E, 2500 (1955) X‘ -ll " IO r- l I l l l L I I l l I 1 l LO 2.0 3.0 “91:92 oK-l) FIGURE 7,I Tunneling calculations compared with experimental high and low temperature data. 110 n.9x1::~> mm mm. _~m. mm. mm. .m. 00. _ d 05.??? , ‘ _ I4\\1 l .zocmsqmsm mp0s IHMOOH m> wpcmfimcoo cowmjmmfio fig Ob, a4 mo Oflpmm NF memv—h. b . w .vm. no. No. A //7%/// / CK 7 4 _ a 1.43.9 I 00.. V /V/V/MV VV /VV/VAF/ //// antmno £60.13. XOON¢HF // /V/.V/V/. 2.. 1 ON._ [s 10 I CHAPTER 8 SUMMARY AND CONCLUSIONS The purpose of our isotOpic-diffusion studies has been to determine whether a significant time saving could result in nuclear—detector fabrication if 6Li is used instead Of 7Li. A review of existing data showed that the apparent activation energy Q for 7Li in germanium increased with increasing temperatures. Such an effect might be caused by barrier tunneling at the lower temperatures; such a tunneling mechanism would tend to increase the ratio of diffusion constants D6/D7 at the lower temperatures. We have measured this ratio and found it to be 1.091 at 000°C and 1.096 at 200°C. These ratios are greater than the 1.080 value obtained from classical calculations and therefore indicate the possibility of tunneling. Calculations based on a diffusion theory which incorporated tunneling effects were performed. Good agreement was found to exist between the calculations and existing 7Li diffusion data with our measured ratios. These calculations and experimental data indicate a ratio of about 1.12 at room temperature. This effect is smaller than had been hoped for and translates into a saving of only 10 hours in a ten-day detector drift. These calculations did not evaluate the 7Li local-mode frequency or any of the effects of the lattice on the local mode. Consequently, the infrared-absorption studies were undertaken. 111 112 The study of local modes in germanium proved to be too difficult because of sample—preparation problems; we there- fore studied local modes in silicon. These local modes are similar in many important respects to the local modes in germanium. For this work it was necessary to compensate the lithium with boron in order to reduce the free-carrier— absorption background. We then set out to evaluate the effects of anharmonic terms in atomic displacements on the local modes through application Of a uniaxial stress. The uniaxial-stress experiments, however, did not show any splitting Of the local modes. In fact, the stress necessary to split the two-fold degeneracies appears to be considerably larger than the stress necessary to fracture the sample. The thermal experiments determined the line width and centroid shifts of the local modes as a function of temper— ature. A linear temperature dependence in line width and centroid shift was observed at temperatures above 300°K. A theory describing these anharmonic effects on local modes in a C3v symmetry has not yet been developed. We modified the theory of Elliott et al. in an effort to account for the actual C3V symmetry of the lithium-boron system. We then compared calculations based on these ideas with the Observed temperature dependences. Data and calculations for centroid shift were found to agree reasonably well. On the basis of the modified theory of Elliott et al., this agreement indicated that the centroid shifts were the result of anharmonic effects due tO fourth-order terms in atomic displacement. pv 113 Two processes were considered in calculating the local- mode line widths; two—phonon decay and elastic-scattering of band phonons. The elastic-scattering mechanism was re— jected on the basis of its T2 dependence at high temperature. The two—phonon-decay process was then fit to the experimental data at 300°K and above. The agreement between calculations and experiments at temperatures above 300°K was fair; however, below 300°K there was no agreement. It will therefore be necessary for a rigorous theory, based on the actual C3V sym— metry of the local mode, to be deveIOped in order to fully explain both the line width and centroid shift processes. BIBLIOGRAPHY 10. ll. 12. 13. BIBLIOGRAPHY R. Chrenko, R. McDonald and E. Pell, Phys. Rev. 38, A1770 (1965) P. M. Pfeuty, Proc. Int. Conf. on Localized Excitations in Solids, Irvine, p. 193 (1967) M. Balkansik and W. Nazarewicz a. J. Phys. Chem. Solids 25, 070 (1960) b. J. Phys. Chem. Solids 31, 571 (1955) c. and Pfeuty, Lattice Defects in Semiconductors, R. R. Hasiguti, ed., The Pennsylvania State University Press, Univ. Park, Md., p.3 (1968) V. Tsevetou, W. Allrod, and W. G. Spitzer, Proc. Int. Conf. on Localized Excitations in Solids, Irvine p. 185 (1967) W. G- Spitzer, NASA Report No. Cr-97309 (1968) L. Bellomonte and M. H. L. Pryce, Pro. Phys. Soc. 8:, pp. 967 and 973 (1966) R. J. Elliott and P. Pfeuty, J. Phys. Chem. Solids 28, 1789 (1967) W. Hayes, Proc. Int. Conf. on Localized Excitations in Solids, Irvine, p. 159 (1967) W. A. Spitzer and M. Waldner a. Phys. Rev. Letters 1i, 223 (1965) b. J. Appl. Phys. 36, 2050 (1965) M. Waldner, M. A. Hiller and W. G. Spitzer, Phys. Rev. 100A, 172 (1955) A. H. Sher, J. Appl. Phys 19, 2500 (1955) s. M. Pell, Phys. Rev. 119, 1010 (1950) L. J. Giacoletto, Proc. IRE 39, 921 (1961) 110 10. 15. 16. 17. 18. 19. 20. 21. 22. 23. 20. 25. 26. 27. 28. 29. 30. 31. 32. 33. 115 C. S. Fuller and J. C. Sevriens, Phys. Rev 96, 21 (1950) R. Aggarwal, P. Fisher, V. Mourzine and A. Ramdas, Phys. Rev. 138, A882 (1955) B. I. Boltaks, Diffusion in Semiconductors, Academic Press, Inc. ch. 3 71963) C. Wert and C. Zener, Phys. Rev. 76, 1169 (1909) . S. Yeremeyev, Phys. of Metals and Metallography 5, 139 (1968 m< . Dawber and R. Elliott . Proc. Roy. Soc. A273, 222 (1963) . Proc. Phys. Soc. 81, 053 (1963) U'DJ'U J. Angress, A. R. Goodwin and S. D. Smith, Proc. Roy. Soc. A287, 60 (1965) E. M. Pell, Appl. Phys. 31, 1675 (1960) H. Reiss, C. S. Fuller and F. J. Morin, Bell System Tech. J. 35, no. 3 535 (1956) F. A. Kroger, Chemistry 9: Imperfect Crystals, North Holland Publishing Co., Amsterdam (I969I’ch. 9, sec. 2 E. M. Pell, J. Phys. Chem Solids 3, 77 (1957) R. J. Elliott, W. Hayes, G. D. Jones, H. F. MacDonald and C. T. Sennett, Proc. Roy. Soc. Lon. 289A, 1 (1965) W. Hayes and H. F. MacDonald, Proc. Prog. Sec. (London) A297, 503 (1957) H. Y. Fan, W. Spitzer and R. J. Collins, Phys. Rev. 101, 566 (1955) H. F. Wolf, Silicon Semiconductor Data, Pergamon Press, (1969) pp. 1 2-110. “— Ibid., p. 115 M. Lax and E. Burstein, Phys. Rev. 97, 39 (1955) B. N. Brockhouse and P. K. Iyenger, Phys. Rev. 111, 707 (1958) R. J. Colling and H. Y. Fan, Phys. Rev. 93, 670 (1953) A. E. Cosand and W. G. Spitzer, Appl. Phys. Lett. 11, 279 (1967) 30. 35. 36. 37. 38. 39. 0O. 01. 02. 03. 00. 05. 06. 07. 08. 09. 50. 51. 52. 116 E. M. Pell, Symposium on Solid State Physics in Electronics and Telecommunications, Academic Press, New York, 6, 261 (1960) M. Kahlweit, R. A. Swalin, and R. D. Weltzin, Progress ip_Solid State Chemistry, (Reiss, ed.), 2, ch. 5 and'6 A. M. Sher and J. A. Coleman IEEE Trans. Nucl. Sci, NS-l7, 125 (1970) R. A. LOgan, Phys. Rev. 101, 1055 (1956) L. P. Hunter, Handbook g£_Semiconductor Electronics, second ed., McGraw—Hill —— Donald E. McCarthy, App. Optics 2, 591 (1963) Smith, Jones and Chasmar, The Detection and Measurement of Infrared Radiation, , Oxford Press, ch. 2—(1963) —— R. P. Bauman, Absorption Spectroscopy, John Wiley and Sons Inc., (1963) ch. 2 Fritz Kneubuhl, Appl. Optics 3, 505 (1969) Smith, Jones, and Chasmar gp. cit., ch. 3 Kneubuhl, 923 cit. Wolf, gp. cit., p. 200 G. L. Cloud and J. T. Pindera, Experimental Mechanics I, (May 1968) R. C. Newman and J. B. Willis, J. Phys. Chem. Solids 26, 373 (1965) W. Hayes and D. R. Bosomworth, Phys. Rev. Letters 23, 551 (1959) _‘ F. M. Smits, Proc. IRE 06, 1009 (1958) R. N. Hall and T. J. Solyts, IEEE Trans. Nucl. Sci. 05-15, 150 (Feb. 1971) B. Pratt and F. Friedmann, J. Appl. Phys. 37, 1893 (1966) J. P. Maita, J. Phys. Chem Solids 0, 68 (1958) APPENDICES Because of the small local-mode absorption coefficients in germanium, an Optimization technique was used to deter- Inine the best sample thickness. This technique was carried OLrt for the split-beam system described in section 0.2. lie ihave considered the system in which a difference measure- . O > 7 0 O nmant'ls formed with a compensated L1 sample in the upper b . . . 6 . «seam, and either a pure Silicon sample or a compensated L1 searnple in the lower beam. The intensity of the upper beam (5311) and (SL) of the lower beam, upon entering the monochro— nhaftor, can be expressed as and SQL are the Vvliexre x is the sample thickness and SO and aL refer :irliqtial upper and lower beam intensities; a Back— ‘tCD .absorption coefficients due to sample attenuation. XZrWDLnid effects are tuned to a zero on a spectral region of FND lIDcal-modes, i.e. we adjust SO by means of a variable we d 88 so that 117 The absorption coefficient aFC and ab, refer to the free— carrier-background absorption and lattice absorption respec— tively. We then have where : —. .. ‘( AA (“rcxu aFCXL+aL’Xu “5"L) ' 'Thue detected signal S is then the difference between Su and ESL.’ For a compensated sample in the upper beam and pure ggerrwanium, of the same thickness, in the lower beam 8 = SOu e-GBXIe-aLMX-l] , vvlfieere aLM is the local-mode absorption coefficient and GB $153 the background absorption, i.e. aFC +aL,. For aLMX< mEO\mHoH mo COLOQ cam Ejacuwa mo mCOapmspcwocoo Hmsom pom OLSPOLOQEOO. .m> on mcofl omswmd mo cowpomsm _.m MEDGE Axov com ooh 003 com A. _ _ ___ ___ _a__3___1__ APPENDIX C Use of bakelite contacts between the sample and the pressure clamp, raised the question of friction at the sample contacts avid the resulting possibility of large components of stress Exarpendicular to the direction of application. When uniaxial- stloess is applied to silicon, it expands in directions per- pueridicular to the applied force. But, friction at the pressure agbp>1icator contacts opposes this expansion. The result may 1362 local stress components considerably larger than the applied satiress. The amount of lateral expansion is indicated by Pois- sscxn's ratio, defined as the change in sample width perpendicular 1:c> the applied force, divided by the change in length along For stress applied to the ‘tlie directions Of applied force. 05 [:]_10] direction of silicon, Poisson's ratio is 0.33. It also possible that stress variations may exist in the ixs sseunple around regions of lithium precipitation discussed decided to evaluate the stress eéirlier. It was therefore Var‘iations across the loaded sample by birefringence tech- rliques.26 The test set-up for the birefringence work is shown in Irigyare C.1. In this system circularly polarized and nearly monoChromatic light is incident upon the sample such that t22e 'two components of the light are in the plane of the 122 123 .mspmsmmam useEOLSmmoE mocmmcHsmwsflm 296 352:; $2.28 cash StgcoooEPE _ . o mane E 3963 9.3 2 8.... 323.8 a II 858 a a. sum 888982. / _ fl W e fl / 502E m..— oSE 29:. 963 3:25 263 8:25 120 sample. The silicon crystal under applied stress is Opti— cally anisotropic, 1.3., it has different values of refrac- tive index if different directions, or in other words, the two components of light will prOpagate through the sample at slightly different velocities.2 When these two com- ponents are remixed to produce a plane polarized wave, an interference pattern results. The pattern of a loaded sample (one which has uniaxial stress applied) indicates the amount of applied stress when compared to the pattern of an unloaded sample. The components of the birefringence system are standard, and their function is easily explained. The incandescent source has approximately a black—body radiation characteris- tic. The filter is used to block most of the unwanted visible radiation which would not pass through the silicon sample. The first polarizer,22 set at zero degrees, is used to establish a polarization reference; it produces plane- polarized light which can be represented by PX. The x axis has been arbitrarily chosen as vertical, and the z axis is in the direction Of propagation. The first quarter—wave platefiflfi then produces a sum of left (P_) and right (P+) Circularly polarized light. Each circularly polarized wave can be represented by components PX and Py, of plane polari— zation as P+ = (l/2)l/2 (PX: i Py). The sample then retards If the crystal is viewed as a continum of refractive index .na the velocity of propagation is given by c/n. "WAhren's-type polarizers made of calcite crystals are used. LThey have typically 00% transmittance out to 2.3 microns. "“*The necessary retardation material is available from p<31aroid Corporation. 125 the propagation of Px with respect to PV as determined by the corresponding anisotrOpy in refractive index. The second quarter-wave plate combines the resultant field into «a plane-polarized wave. The second polarizer can be rotated tc> determine the angle of maximum intensity or to determine (wirdations in intensity vs. polarization angle. Since si_licon doesn't transmit in the visible, infrared radiation hand been used throughout. The photo-converter tube is neacessary to convert the infrared light to visible light fc>r1the measurements. The intensity maxima or minima are referred to as They arise as the x and y components of P+ and P_ If fxninges. ear‘e recombined following passage through the sample. ‘lc3ading produces a uniform stress across the sample, the tvliole sample surface would appear to change uniformly from 1Light to dark as the second polarizer angle is varied. A IWCDnuniform variation in applied stress gives rise to light aIId dark regions or fringes as the different polarization CKJmponents are delayed by different amounts. To evaluate tile total stress variation across a loaded sample, a fringe 1~irle is located at one edge of the sample by rotation of 'tI1e second polarizer. The polarizer angle 61 is then re- c3C>rded. The second polarizer is then rotated slowly so 'tlieat the fringe line appears to move across the sample. ‘Vifieen the fringe line appears to reach the Opposite edge (355‘ the sample, the polarizer angle 02 is again recorded. “7}1ifile the polarizer is being rotated, the number of times n 126 that the original fringe line position changes from light to dark to light again is also noted. This is usually ob- served to be less than one cycle; hence it doesn't enter the calculations. The total stress variation across the loaded sample is then simply %% = (n + 1%0) 100% . This technique was applied to pure silicon and to boron— doped lithium-compensated silicon samples. NO significant differences were observed. In particular, we did not Ob- serve any local stresses which could be attributed to lithium precipitation sites. We found that the stress variations were determined primarily by how uniformly the ends of the sample made contact with the bakelite. The sample ends were ground with aluminum oxide on a glass plate until they appeared to be parallel to the bakelite contacts in the pressure clamp. In order to ascertain an upper limit on the possible stress variations, a compensated sample was made up for the birefringence testing which had a more non-uniform contact with the bakelite than did any of the actual experimental samples. These results showed that the stress variations decrease with applied pressure. At 7.6 kg/mm2 the stress variations were :100%, while at 18 kg/mm2 the variations were 150% and at 23 kg/mm2 the variations were :01%. When some care was taken to fit the sample into the pressure clamp, the typical stress variation 127 at 23 kg/mm2 was :10%. The decrease in stress variations with pressure is likely to be due to compression of the bakelite causing more uniform contacts to the sample. Friction between the sample and the bakelite was found to be a relatively small effect, possibly as a result of expansion of the bakelite. APPENDIX D In their analysis of lithium—boron pairs in silicon, Elliott and Pfeuty7 have considered a basic six—atom model described in the text (see Figure D.l). The pair symmetry in C which has a mechanical representation for displace- 3v ment vectors given by Srl + F? + 6F3 for the six-atom model chosen. The ri's are irreducible representations Of the C symmetry with F and F being one dimensional and F 3v 1 2 3 two dimensional. Both F1 and F3 are infrared active. The symmetry coordinates given by Elliott and Pfeuty for F1 are C1 = -%f-(X1+Yl+Zl) , /3 c2 = J— (X5+Y5+Z§) , /‘§' \ c3 = ——l—— (X2+Y3+Zu) , V3 C0 = ._.__1 (Y2+27+23+X3+X0+Y0) /€ , . and is z; (X+Y+Z) . 128 129 Si 0 B (:flfl I V LI .5 : 1”[ FIGURE D.I The 6-atom model used as basis for Lithium-boron local—mode calcu- lations. 130 1 . .2 : —— \ + + ' n1 /§ (21 jYl 3 Z1 ’ _ 1 . .2 _ 1 . .2 n3 - Tr;— (X2+]Y3+j Z”) , _ 1 . .2 nM - —;— (Y2+jz3+j X“) , _ 1 . .2 T15 - 7:31- (Zz+jx3+j Y”) , arud _ l . .2 Mtith a second coordinate for F3 of Ci = nix, where i = 1.,2,3,0,5,6. The coordinates X,Y,Z are cartesian coordinates C>f the atoms, where each atom is represented by a subscript, 2;.3. the subscript 1 refers to the lithium ion and 5 refers ‘tO the boron ion: j represents the complex cube root of llnity, and 2,3,0 and 6 refer to the surrounding silicon éitons. The two-fold-degenerate boron local mode is then I‘epresented by one of the two dimensional r3's while the CHie-fold degenerate boron local mode is represented by one fo the F The two—fold-degenerate lithium local mode l's' ifs similarly represented by one Of the remaining five F3,S. ITt is observed that the above symmetry coordinates do not irlvolve a mixing of the silicon, lithium, and boron coord— lhates. Therefore, in a first-order approximation to the cactual dynamics of the lithium—boron system, each local> 131 mode may be considered independently of the other local modes. An analysis of anharmonic effects similar to that of Elliott et al.25 can be done for each local mode, taking care to note the appropriate representations as F1 or F3. Whereas, Elliott et al. find a Green's function which trans- forms like the irreducible representation F5 Of Td symmetry, we would have a Green's function which would transform like the irreducible representation F1 or F3 of the C3v symmetry. .After performing symmetric products on these representations and.by use of symmetry arguments to determine which terms «are zero, a third—and fourth—order Hamiltonian of the same .form found by Elliott et al. is expected. However, the local-mode operators and expansion coefficients should be labeled according to which representation and which ion they represent. Equations 2.9 and 2.10 would then appear as ..) H=Z(f-1§é}r<~)—)£B 2( ’I‘ )[S ,+ 2 3 4 Lhm 13 2” Q 2 X- (bxr 2 + b r 2) J k ‘ 2“r3 mu ’y 3 X 3 I1 +8 2 (~ )(b 1 2 F1 2M 0 2e. 21‘3 + sz 2) 2 r n 1 3 +8 1( 2‘ )[Z(b 1 + >21} r3 2M 9 2 x XF3 + bxr l 1. F3 (0M ’y 3 (a(k)+a+(k)) 132 , _ h(b(i€)b(fi'))1/2 CH2 .A .+ 2 L - 2 —~ __0_—— I (b 2 +b 2) 2 + + 00 “MW M 0 2 x y X123 xr3 k,k’ M " “2”r3 ’ Cr12 + 2 + (b F 2 +b F 2) 1(2st 2 1 z 1 1 (,1. l + “3 I“) 1" 1 + 5* 1)2}( (15+ (Em (k’)+ +010) M Q 1 x,y X 3 x? a ‘ a a a ’ l P 3 3 where M1 is the lithium mass and M2 is the boron mass. The notation of Elliott and Pfeuty has been changed slightly so that the boron ion is designated by a 2 instead of a 5. The labeling Of each of the observed local modes is summar- ized in Table 3. Table III — Labeling of Local Modes local _ mode (cm ) label 1 522 F3 2 560.5 F1 2 1 65+ F3 The temperature dependence of the centroid-shift and line width then takes on the form Of Eq. 6.1 and 6.5 for each individual local mode in this approximation. The presence of two boron local modes at the same coordinate location makes it seem likely that their inter— action is comparable with the above effects. By analogy with the development of Elliott et aZ., a line-width 133 temperature dependence is expected to be given by .) b(k)(n(E)+l)(n(n')+1)5(b(fi)_0 2 +0? 2) , I I“ k 3 l where 0’ is the boron local-mode frequency which is not decaying. APPENDIX E The programs for centroid shifts and line widths in- corporated a silicon phonon density of states taken from a paper by Dawber and Elliott.19a below for Aw = 8 1/3 cm-l. The values are given Table IV - Phonon Spectrum i g(iAw) i gCiAw) i g 1 0.0 21 0.20 91 1.u0 2 0.0 22 0.05 H2 1.80 3 0.10 23 0.25 H3 0.80 u 0.10 2H 0.25 an 0.90 5 0.0 25 0.u5 MS 0.35 6 0.10 26 0.30 H6 0.55 7 0.90 27 0.60 H7 1.80 8 0.00 28 0.u5 98 1.55 9 1.00 29 0.60 H9 2.30 10 1.35 30 0.70 50 2.80 11 2.90 31 0.85 51 2.85 12 3.50 32 0.95 52 3.00 13 7.80 33 1.10 53 3.05 1a 9.50 3H 1.u5 SH 3.20 15 10.60 35 2.00 55 2.30 16 12.15 36 2.55 56 1.95 17 9.55 37 2.85 57 1.u0 18 0.25 38 2.90 58 1.20 19 0.0 39 2.15 59 18.60 20 0.25 MO 2.u0 60 31.00 61 9.80 13H 135 For normalization purposes, it can be noted that Z g(iAw) = 173.55 . i=1 The equation to be programmed for centroid shift is ”M 1. g(w)fiw[n(w)+ l/2]dw w I M ng) dw O t) K) = constant. on] To program this equation it was first necessary to convert the integrals to sums; then dm becomes Aw, w becomes iAw for integer values of i, and w“ becomes 61Aw. Equation e1 then becomes 61 Z 8(iAwyhiAw[n(iAw+1/2]Am t t . i=1 CODS an 61 X g(iAw)Am i=1 which simplifies to i g(iAw)[n(iAw+1/2)] . The term c/M1002 HwM2 determines the leading constant and has been denoted as GAMMA in this program. It is evaluated in terms of the local—mode frequency OMEGA on line 5 of the program for centroid shift which follows. As a first try c 9 was taken as 10“1 ergs/cmu. GAMMA is multiplied by 6 on 136 line 8, to account for the six branches in the phonon- dispersion relation. The loOp over temperature, in 200K steps, begins on line 11; the sum over the lattice spectrum begins on line 13. The term AA represents the temperature- dependent factor. On line 17 the coefficients GAMMA and ‘hAw/Z g(iAw) are multiplied in to form AQ/QO. The program 1 was then run with lines 7 and 18 deleted. The results of this calculation were then expressed as AQ/QO = f(T) and compared with the experiments. The constants C and 90 were adjusted to fit our high—temperature data in magnitude and slope. The constant C then represented the variation of c from the value of 1021 ergs/cmu. For the purpose of calcu- lating Q, the zero temperature shift in AQ/QO was subtracted out by adding QOC f(T=0) to 00. This sum is denoted OMEGO in the following program where line 18 calculates Q a +9 Cf(T=0) - Q Cf(T) - Q C2f(T=0) f(T) O O O 0 ll 0 +0 Cf(T=0) - 0 CfCT) o o o The second-order term -QOC2f(T=0) f(T) represents an error of typically less than 1% in calculations of the high- temperature slope. The scale factor C can be conveniently multiplied into GAMMA on line 8. The program for centroid shift of the 659 cm-1 local mode is 10 11 12 13 1” 15 16 17 18 19 2O 21 22 23 data card. 30 MO 100 200 137 DIMENSION G(62),ANS(M0),GG(62) READ 100, (G(I),I=l,61) CENTROID SHIFT - 65a - SILICON FNONON SPECTRUM OMEGA = osu GAMMA (1.0/(10.81*1.6*28.09*1.6*((6*3.1ul6*5.18 C*6*3.1H16*OMEGA)**2)))*(10**25) OMEGO = 663 .0 GAMMA = 6*GAMMA GG(1) = 0.0 DP = (6.625*2.5/l.38) 00 no N=l,36 T=20*0*N DO 30 I=l,60 AA: (l.0/(EXP(I*DP/T)-l.0))+0.5 GG(I+1)=GG(I)+I*G(I)*AA CONTINUE ANS=CO<61)*CANNA*50*3.1u16*1.05u/(