A STUDY OF VACANCIES EN TUNGSTEN QUENCHED 2N SUPERFLUM HELIUM Thesis for the Degree of Ph. D. MECHiGAN STATE UfléVERSéTY RONALD 503EPH GREPSHQ‘VER 1969 This is to certify that the thesis entitled A Study of Vacancies in Tungsten Quenched in Superfluid Helium. presented by Ronald Joseph Gripshover has been accepted towards fulfillment of the requirements for M degree in 5 {C Betta“. a Major professor Date afgil 10;! 7441 0-169 k LIBRARY Michigan Sta“ University Barons av 7", 1 HMS & SIINS' A BOOK BINDERY IND. ‘ . qunnw omens - 3 . , I ~ I, l \ ABSTRACT A STUDY OF VACANCIES IN TUNGSTEN QUE NCHED IN SUPERFLUID HELIUM By Ronald Joseph Gripshover A system has been constructed which permits the refractory body- centered-cubic metals to be annealed and quenched in the ultrapure conditions which exist in and directly above superfluid helium. Studies were made on vacancies quenched into fine tungsten wires and the subse- quent annealing away of this quenched-in resistance. To determine the effective vacancy formation energy, BF, tungsten specimens are quenched from temperatures between 1500K and 3200K. The quenched-in resis— tance varies somewhat from specimen to specimen, but the effective formation energy obtained for the fast quench data is nearly the same for all specimens and equal to 3. 1i 0.2ev. This data, together with Schultz's data, establishes both the magnitude of the quenched-in resistance and the effective formation energy for fast quenched tungsten. Ronald Joseph Gripshover By studying how the quenched-in resistance varies with quench speed, itis shown that the equilibrium vacancy concentration is probably not being retained even with the fast quench speeds. By extrapolating the data to infinite quench speed we get an estimate of the equilibrium vacancy resis- tance. The extrapolation leads to an estimate of Ef = 3. 5i 0.2 ev. The extrapolated data is also used to test Flynn's theory of vacancy annealing during the quench. The theory fits the infinite quench speed data, but there is slightly too much scatter to state definitively that Flynn's theory is applicable. This theory yields avalue of 1. 4i 0.2 ev for the vacancy motion energy. To study the annealing away of the quenched-in resistance, both iso- chronal and isothermal anneals were performed. There is a single major isochronal recovery stage in the range 800-1000K in quenched tungsten, consistent with Stage IV recovery in radiation damage and cold work studies. This single major recovery stage suggests that there is a single defect present, presumably the vacancy. The isothermal recovery curve has an "S" shape suggesting complex annealingprocesses. Isothermal anneals with change-of—slope measurements yield an effective motion energy of 1. 5i 0.3 ev. This estimate of the effective motion energy is in good agreement with the estimate obtained from the Flynn theory analysis. If we take our "best values" for the effective formation and motion energies, we obtain a value of about 5 ev (3. 5ev + 1. 58V) for Q. This is in reasonable agreement with the 5.2 ev estimate of Danneberg. The maximum Q which would still be consistent with our data is Q N 6 ev. However, our data do not appear to be consistent with either Q = 6.‘ ev, or with Elin = 3. 3 ev obtained from recent radiation damage studies. A STUDY OF VACANCIES IN TUNGSTEN QUENC HED IN SUPERFLUID HELIUM By Ronald Joseph Gripshover A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOC TOR OF PHILOSOPHY Department of Physics 1969 7a «In» 7 c ,Q fr. ACKNOWLEDGMENTS I would like to thank my advisor, Dr. Jack Bass, for his invaluable guidance and assistance during the entire course of this work. I wish to thank Mr. John Zetts for assisting in taking much of the data and for many helpful discussions. I am indebted to Dr. J. M. Galligan for kindly supply- ing us with a very pure 2. 0 mil tungsten specimen. I would also like to acknowledge the financial support of the Atomic Energy Commision. Finally, Iwish to express my deepest appreciation to my wife, Sharon, whose support and encouragement throughout this work have been invaluable. I am also deeply indebted to her for performing much of the analysis and preparation of this manuscript. ii TABLE OF CONTENTS Page LIST OF TABLES ............................ v LIST OF FIGURES ........................... vi I. INTRODUCTION ......................... 1 A. Imperfections in Metals ................ 1 B. Vacancy Formation Energy, E? ............ 1 C. Techniques for Measuring E2] ............. 2 D. Vacancy Motion Energy, Ean ............ . 4 E. Techniques for Measuring Em ............. 5 F. Previous Work on Tungsten ............... 6 G. Previous Work on Molybdenum ........... 10 H. Present Experiment ................. 11 II. QUENCHING TECHNIQUE ................... 14 A. Specimen Preparation ................ 14 B. Quenching Procedure ................. 16 1. Room ’Emperature Measurements ...... 16 2. Lighting Procedure .............. 17 3. Temperature Determination .......... 19 4. Quenching under the Superfluid ........ 20 5. Determination of Quenched—in Resistance. . .21 6. Advantages and Disadvantages ......... 21 III. APPARATUS ........................... 23 A. Specimen Holder .................... 23 B. Cryostat ........................ 25 C. Circuits ........................ 27 1. DC Circuits .................. 29 3. Measuring circuit ............ 29 b. Heating circuit .............. 31 0. Potential circuits .............. 34 2. AC Circuits .................. 38 a. Constant current source ......... 38 b. Amplifier ................. 42 0. AC to DC converter ........... 44 d. Readout devices ............. 47 e. Calibration and evaluation ........ 49 iii IV. TUNGSTEN QUENCHING RESULTS .............. 51 A. Quench Speed Data .................. 51 B. Fast Quench Tungsten Data .............. 53 C. Dependence of Quenched—in Resistance on Quench Speed ................... 58 D. Additional Tungsten Data ............... 64 E. Experimental Accuracy of the Quench Data ..... 67 F. Extrapolation ..................... 70 G. Flynn Theory Analysis ................ 73 V. ANNEALING OF DEFECTS QUENCHED INTO TUNGSTEN. . .77 A. Isochronal Anneals .................. .78 B. Isothermal Anneals .................. 85 1. Isothermal Anneal ............... 85 2. Isothermal Anneal with Change-of— Slope Measurements ........... 87 VI. MOLYBDENU M DATA ...................... 92 VII. SUMMARY ............................ 94 LIST OF REFERENCES ......................... 97 APPENDD( A Significance of the Superfluid ............. 100 APPENDIX B Operational Amplifiers ................ 102 iv LIST OF TABLES Table Page 1 Summary of Past Work on Tungsten .......... . ............... 9 2 Summary of Past Work on Molybdenum ...................... 12 LIST OF FIGURES Figure Page 1. Isochronal Recovery of Irradiated Tungsten .......... 8 2. Simplified Schematic of the capacitance Discharge Circuit . . . 18 3. Specimen Holder ......................... 24 4. Cryostat .............................. 26 5. Block Diagram of Circuits ..................... 28 6. Measuring Circuit ........................ .30 7. Heating Circuit .......................... 32. 8. Quench Speed Control Capacitors ................ .33 9. Capacitor Charging Circuit .................... .35 10. Potential Circuits ......................... 36 11. Block Diagram of the AC Circuits ................ 39 12. AC Constant Current Source ................... 40 13. AC Amplifier ........................... 43 14. Twin T Network .......................... 45 15. AC to DC Converter ....................... 46 16. Diode Output to Resistive and Capacitive Loads ......... 48 17. Quench Speed Graphs ...................... 52 18. Quench Speed versus Quench Temperature for Specimen C (1.2 mil) ............................. 54 19. Quench Speed versus Quench Temperature for Specimen A (1. 0 mil) ....................... 55 vi Figure Page 20. Fast Quench Data for Specimen A (1. 0 mil) .......... 56 21. Fast Quench Data for Specimens A, B, C, and D ....... 59 22. Dependence of Quenched—in Resistance on Quench Speed for Specimen C ................ .60 23. Dependence of Quenched-in Resistance on Quench Speed for Specimen B .................. 62 24. Dependence of Quenched—in Resistance on Quench Speed for Specimen A . ................. 63 25. Dependence of Quenched-in Resistance on Quench Speed for Specimen A with Constant Subtracted ........ 65 26. Radiation Shield ......................... 69 27. Data Extrapolation for Specimen A ............... 71 28. Extrapolated Infinite Quench Speed Data ............ 72 29. 1/Tq q versus 1/TCl Plots for the Flynn Analysis ...... 75 30. Isochronal Anneals for Specimens Quenched from 2800K to 3000K ...................... 79 31. Isochronal Anneals for Specimens Quenched from Temperatures above 3200K ................... 81 32. Second Order Analysis of the Data from Figure 31 ....... 84 33. Isothermal Anneal ........................ 86 34. Isothermal Anneal with Change-of—Slope Measurements .......................... 89 35. Isothermal Anneal with Change-of—Slope Measurements .......................... 90 36. Preliminary Molybdenum Data ................. 93 vii I. INTRODUCTION A. Imperfections in Metals Many important properties of metals are affected by imperfections in the ideal lattice. A number of these properties are determined more by the nature of the imperfections than by the nature of the host crystal. Some well-known examples are: the electrical conductivity of semicon- ductors, the low temperature resistivity of metals, atomic diffusion, the mechanical and plastic properties of solids. To understand these properties, one must understand the nature and properties of the imperfections involved. The simplest imperfection is a missing atom, or lattice vacancy. It is the imperfection which we wish to study. The study of the properties of the single vacancy standardly focuses upon two quantities, the vacancy formation energr E}: and the vacancy motion energy El’n' B. Vacancy Formation Energy, E: The fractional concentration of single vacancies nV(T) in a pure mater- ial in thermal equilibrium at a temperature T and zero pressure can be 1_/ written in terms of the vacancy formation energy as nvn‘) - exp<-Gl’(r)/k'r) - exp(s¥/k)- exp(-El’/kT> (1) where T is the absolute temperature, k is Boltzman's constant, G}, is the Gibbs free energy per vacancy, and S}, is the entropy increase per vacancy. The pressure correction for metals at atmospheric pressure is negligible. If the equilibrium vacancy concentration could be measured 1 as a function of T, the slope of a plot of ln(nv) versus 1/T would be -E¥/k Typically, for pure metals, E¥~1 electron volt (ev) and the fractional vacancy concentration at the melting point is less than 10-3. Hence, to get an appreciable thermal equilibrium vacancy concentration, the metal must be at a high temperature. C. Techniques for Measuring}; In theory the vacancy concentration can be studied by measuring any property of the material which is affected by the presence of vacancies, such as electrical resistance, specific heat, hardness, length, etc. The difficulty is that one must know to high accuracy how this property behaves in the absence of vacancies. This problem was solved by Simmons and Balluffig/ They showed that for a cubic crystal the vacancy concentration could be written as * R.T. a‘R.T. ( ) where AL(T)/LR. T. is the fractional increase in the macroscopic length of the specimen from room temperature to T and A a(T)/aR.T. is the fractional increase in the microscopic lattice parameter. By taking this difference, they showed that all effects not associated with the presence of vacancies are removed. While this is a direct method for determining the vacancy con- centration, it has limited accuracy and it is not easily extended to high temperatures with presently available techniques. Also, since the vacancies are studied at thermal equilibrium, this method gives no direct information *Strictly speaking, the left hand side of this equation is nV -ni, where ni is the fractional interstitial concentration. At high temperatures in the metals of interest, ni is almost surely much less than nv, giving the above result. about vacancy mobility. To date, this method has been most useful in establishing the vacancy concentration at the melting point and the validity of other techniques, such as quenching. If a specimen is cooled sufficiently rapidly from a high temperature, most of the vacancies present before the quench are trapped in the material, allowing them to be studied at lower temperatures. Several methods can be used to study vacancies at these temperatures: 1) vacancy clusters canbe viewed by electron microscopyéJZ) individual vacancies, as well as clusters, can be seen and counted with a field-ion microscopeé/3) length change between the quenched and unquenched specimen can be measured; 4) changes in the heat output can be observed at temperatures where the quenched-in resistance begins to anneal away; 5) resistivity increases can be measured, usually, though not always, at liquid helium temperatures. The last method is most commonly used because resistances canbe accurately and precisely measured. Since it is used in this experiment, it will be described in some detail. To find the vacancy concentration as a function of temperature with this method, the specimen is first carefully annealed to get a stable, vacancy— free, base resistance. This resistance is measured at liquid helium tem- peratures to keep the normal lattice resistance from masking the vacancy contribution to the resistance. The specimen is then quenched from a lmown temperature, and its resistance at helium temperature again measured. The vacancy concentration is assumed to be proportional to the increase in resistance. If equation (1) holds, a plot of the legarithm of the resistivity increase against the inverse quench temperature will yield a straight line whose slope is the formation energy divided by the Boltzmann constant. In general, when quenching with a finite quench speed, one does not expect to retain all the vacancies originally present in thermal equlibrium at the quench temperature T The data obtained will therefore not be expected q' to reproduce equation (1) exactly. When dealing with quench data analyzed in terms of equation (1) we will speak of an "effective formation energy", to distinguish the experimentally determined numbers from the "correct" formation energy. The quenching technique has inherent in it several potential difficulties. It assumes that the increase in the base resistance is due solely to quenched- in vacancies, when, in fact, many things, such as changes in the internal structure of the specimen (grain boundaries, etc. ), different impurity contributions, strains in the specimen, etc. , could conceivably contribute to such aresistance increase. Each vacancy is also assumed to contribute the same amount to the resistivity increase. If there is vacancy-vacancy or vacancy-impurity clustering, this may not be true. Also it is assumed that all of the vacancies present at the quench temperature are quenched- in. This again may not be correct, especially when the specimen is quenched from very high temperatures. However if these difficulties are kept in mind, quenching experiments can be of great value. From them effective vacancy formation energies can be determined, and by measuring how the quenched-in resistance varies as a function of quench speed, one can obtain an estimate for the "correct" vacancy formation energy, as is described in Section IV of this thesis. D. Vacancy Motion Energy, EV,n It is usually assumed that at constant temperature an excess vacancy concentration will anneal down to the thermal equilibrium concentration é/ according to the equation -EVm/kT 33,, : -F(nv) e (3) dt where Elln is the vacancy motion energy, and F(nv) is a function of the vacancy concentration whose form is determined by the details of the annealing process. Ifsome function of nV is measured, such as the vacancy resistivityf), it is assumed that -E;',,/kT 11%: ’F(P) e (33) d This assumption introduces some problems; first, pmay not be proportional to nv, and secondly, this equation assumes that the only mobile entity is the single vacancy. In a real experiment, small vacancy clusters, vacancy- impurity complexes, and other entities such as interstitials may also be mobile, and by their motion to sinks or to large clusters contribute to a change in resistivity. Therefore, when analyzing annealing data using equation (3a) we will speak of an "effective motion energy" to distinguish the experimentally determined quantity from the real motion energy of single, free vacancies. IfF(nv) is proportionaltonv, the annealing process is said to be a "first order process" or to involve "first order kinetics". If F(nv) is proportional to (nv)2, the annealing process is said to involve "second order kinetics". We shall return to this equation in the analysis of the annealing data in Sections V and VI. E. Techniques for Measuring E; Efn can be estimated most directly by studying the annealing of vacancies from a specimen. There are two annealing procedures generally used to determine motion energies: isochronal annealing and isothermal annealing. An isochronal anneal consists in heating the specimen for a fixed time at each of a series of increasing temperatures until most of the quenched-in resistance anneals out. This gives the recovery as a function of the annealing tem- perature. An isothermal anneal consists in heating the specimen to a fixed temperature to obtain the recovery as a function of the annealing time. The detailed methods used for the calculation of El’n from annealing data will be discussed in conjunction with the analysis of the annealing data in Sections V and VI. The quenching technique is the best method of producing vacancies in a metal at low temperatures without producing large concentrations of other defects. However, point defects (vacancies, interstitials, etc.) can also be produced by radiation damage and cold working The difficulty with producing vacancies in these ways is that they represent only a fraction of the defects present, making it difficult to interpret the annealing data satisfactorily. The vacancy motion energy can also be estimated if the vacancy forma- tion energy E; and the activation energy for self diffusion, Q, are known. If diffusion takes place by means of single vacancies, which appears to be §/ the case for many pure metals, it can be shown that Elln = Q - E¥ F. Previous Work on Tungsten There is only one short paper in the literature which reports quantita- 7J tive quenching results on tungsten. In 1964 Schultz published the results of quenching a single 1. 2 mil tungsten wire using the superfluid technique. He obtained an effective vacancy formation energy of 3. 3ev. By extrapolating his data to the melting point and assuming that nV : 0. 02%;”, where ARq is the resistance that would be quenched—in during a quench from the melting point, he obtained a fractional vacancy concentration at the melting point of 1.1 x 10’.4 Vacancies and other defects in tungsten have been investigated by several experimenters using techniques other than quenching. From measurements of the high temperature specific heat, Kraftmakher and Strelkov§/<)btained aformation energy of 3.14ev and a vacancy concentration of 2. 7 percent at 3600 degrees Kelvin (3600K). In this technique, the low temperature specific heat is linearly extrapolated to high temperatures and any deviation of the data from the extrapolation is considered to be due to vacancies. It is difficult to make this extrapolation accurately, and it is not likely that all deviations from linearity are due solely to vacancies. The formation energy they obtained is in reasonable agreement with that obtained by Schultz, but their vacancy concentrations are two orders of magnitude greater than Schultz's. There have also been a number of radiation damage and cold work experiments with tungsten. On the next page is shown-94. typical isochronal recovery curve for radiation damaged tungsten. The fractional defect concentration is plotted as a function of the annealing temperature. The annealing stages have been labeled according to current practice. During each stage a particular defect becomes mobile and anneals away. a—I-ple I LO~ l | I if oloofififisbfismw Too—b D 0 Figure 1: Isochronal Recovery of Irradiated Tungsten Table 1 is a summary, in chronological order, of the previous work on tungsten. The remaining comments in this section will be in reference to this table. First note that there are no estimates for Ern’ or for the temper- ature at which vacancies become mobile, obtained from annealing studies of quenched tungsten. Since vacancies are the predominant defect in quenched tungsten, one would expect that annealing studies of quenched tungsten would yield the most reliable values for both the temperature at which vacancies become mobile and Evm. Until the mid-sixties it was thought that vacancies became mobile in Stage III but recent work, especially the field ion micro- scopy work of Galligan, et. al., suggests that vacancies are not mobile until Stage IV. Consequently, the estimates for Ei’n have recently been increased from 1. 7 ev to about 3 ev, and the estimate of the temperature at which vacancies become mobile has increased from the 600—700K range to the 700-1000K range. The experimental values for Q listed in Table 1 are not mutually con- sistant. This rules out for the present any estimate of the vacancy motion am am om ma ma 5H 0H ma «a ma NH dd 0H m.m N.m Nv.m mm.m Nm.m ¢H.m mu.H b.H h.H Ililcowumaseamo Hecaucnomsullil onmauomm .>H mmmum owaa coaumflpmnnw conusmc onmnomh .>H mmmum coaumflumuufl conusmc chad .>H wmmum cowumaomunw couusm: maflnocczw xuoz waoo coaumaommnuxm new: camauomm chhloom xuoz caoo omniomw.HHHmmmum coaunwonnuw conusmc onenesm.HHHommum one omm xno3 waoo xuo3 waoo cowumwwmuuw couusoa Amen eeaam.m.o coaunwvnnuw couusmcAwmvmmmuoz.b.£uHEm.m.b.m>mm.x.q Asmvemmaaamo.2.n a manoecmmb.o Ammv emmaaamo.2.n a commun¢.z Ammo eemm.z .uemnem.o.n.enamee<.q.m AmmveuaEm.m.n a mmmuoz.n Avon unannom.m Away Nuaseom.m Awwwaonum.o.m.umexmeummnm.a.mu Aamv muonmccmn .3 Aomvcflamzm.¢.m a Mnmeflcz.d.q Acme comm20a9.3.z Ammv Hmuuflceom.m.m Ammv Nuaszom.m Amm. oox.o.m Ammvnommfione.3.z a casucwx.m.0 Ammvommwmwuomom a >m.aamm>.m.> 1I.iiI!iI .mmm Axvmaanoz 0800mm mmwocmom> now£3 um cusumnmmswa monocmonv mqwosvoum mo vogue Amvuoummflumm>cH ZmBmUZDB ZO MMOS Bmdm m0 NMdZEDm H mqmdfi 10 energr from the relation, EX, = Q - E}, until the value of Q is more firmly established for tungsten. There have been only two theoretical papers published concerning the vacancy formation and motion energies for the body-centered cubic (bcc) transition metals. In 1963, Gregory published a short paper in which he used "reasonable" values for the quantities in equations (1) and (3) to conclude that "it may just be possible to quench in a measurable concen- tration of lattice vacancies in bcc transition metalszg/ In 1968, from an elastic theory of metals, Flynn calculated a vacancy motion energy of 2. 8 ev for tungsten. G. Previous Work on Molybdenum No systematic study of vacancies quenched into molybdenum has been published. Meakin et.al.g_4_/quenched a molybdenum single crystal from vacuum into molten indium. They then used electron transmission micro- scopy to examine what they interpreted as vacancy loops. From the number of loops they saw, they estimate a fractional vacancy concentration at the melting point of 5 x 10—5, from which they deduced a formation energy of 2 .4 ev. Nihoulzjjperformed " preliminary quenching experiments" by breaking the glass envelope of the specimen (immersed in water) when the specimen reached the desired temperature. He used the results of these experiments to strengthen his contention that vacancies do not become mobile until Stage IV in molybdenum. Kraftmakher, again using specific heat determinationsgfijobtained a formation energy of 2.24 ev and a vacancy concentration at the melting point of 4. 3 percent. The formation energy obtained by Kraftmakher is consistant with Meakin's, but the vacancy concentration is higher by three orders of magnitude. 11 As with the tungsten, there have also been a number of radiation damage and cold work studies of molybdenum. These results, a recent calculation of El’n by Flynn, and measurements of Q for molybdenum, are summarized in Table 2. H. Present Experiment Vacancies in the common face—centered cubic (fcc) metals, especially 2223/ gold, have been extensively studied. Most of the above techniques have been applied to these metals. All of the quench techniques, as well as the Simmons—Balluffi technique,yield values for the vacancy formation energy which are mutually consist nt. These studies serve to validate the quenching technique. The fee metals were studied first because they are much easier to purify and to work with. They can be resistance heated in air and quenched by plunging them into water or some other suitable liquid, or by merely turning off the current and allowing them to cool in air. This is not the case for bcc transition metals. Only recently have these metals been purified sufficiently for meaiingful quenching experiments tobeperformed on them. They also contaminate rapidly when heated in air, making the usual quenching technique of heating in air and plunging into a suitable liquid inapplicable. In 1963, atechnique was developed byRinderer and SchultzZ—g/which permits these metals to be annealed and quenched in the ultrapure conditions which exist in and directly above superfluid helium. They showed that fine wires could be resistance heated and quenched in the superfluid. This technique is used in the present experiment, which is the beginning of a planned systematic study of vacancies quenched into the refractory bcc metals (groups IVB and VB in the periodic table). Thus, a mm m.~ IIIIaOAumancho Hmoflucnoegullnl “may acmam .m.u «m m.m Edfiocfl Avmv. cmuaoa QH zonesw 00M.m.mea3mq.m.qflxmmz.n.b cowumaomnuuxo mm w~.~ new: UHMwommm Avov Hmnxmeumnux.¢ .mm m.v mm Io.v Ammv Hawxm¢.h em mm.H omv .HHH emmum cowumwomnufl couusmc Ammvcomcn0h.m.m a xooomom.m.o mm cos gunman Ammo anoaaz .n mewoanom> uoe HHH mmmum coaumaemuua coupes: “Nov Hsoeaz .n Aamv mm hm.¢ upsmum.m a mumnmccmo.3 co> Ammv>ouowmm.m.0 .>oafla>mm.>.q Hm o.m .eausno.q.m.>0mauom.>.m HH m.H owe coanaaemuun conusmcfimmvaOmmeoee.z.z a caeoaam.m.o mm omv xHOB pace Ammv nommaom.m.m om mm.a omv xuoz waoo Anny afluumz.m %” g A>mv A>mv A>mv Anomaanoz E eEoomm moflocmom> mmwocmom> .mmm 0 Wm >m noflg3 um musumuomEcB mcwoswonm mo vogue: Amvnounmflume>cH ZDZMDmMAOZ N mqmdfi ZO MmOB Bmdm m0 Mm>0n_ 0Q ~ W50... _ o 9.50.: 0 crew: meccomnoé 29 1. DC circuits a. Measuring circuit Figure 6 is a schematic of the measuring circuit. For the earlier experiments the power supply was a set of 6 volt storage batteries. Now a Princeton Applied Research (PAR) model TC 602R power supply is used. The output voltage of this power supply is more constant than the output of the batteries. Rv is a series of resistors and switches used to control the measuring current. A detailed schematic of RV is shown at the bottom of Figure 6. By appropriately varying the variable resistors and the switches, Rv is nearly continuously variable between 1 ohm and 52,221 ohms. The 1 kilohm and 50kilohm variable resistors are 10-turn helipots The fixed resistors are used to protect the variable resistors when they are turned to low re- sistance. With this system we can control a 1 ampere current to five places. The 1 ohm standard resistor is a Leeds and Northrup NationalBureau of Standards type resistance standard #4020B. It is used to measure the current when the specimen's liquid helium temperature resistance is being determined. The 100 ohm standard resistor is a Leeds and Northrup National Bureau of Standards type resistance standard #4030B, It is used to measure the low current usedin measuring the room temperature resistance. When- ever currents above a few milliamperes were used, the 100 ohm standard was shorted by S3 to avoid damaging it and to decrease the resistance of the measuring circuits. The standard resistors are both four—terminal resistors; the potential terminals are connected to the potential circuits as indicated. 30 ' ,h to h &g of‘ potential circuit Wétocgtdofg I n-s+d. 53 Co: potential circuit 9 $d ‘bo o i 82 I\ C — — -—DETAIL OF RV ABOVE- — ———n i 30 m mg mg 1. 1.1.4 : Figure 6: Measuring Circuit 31 b. Heating circuit Figure 7 is a schematic of the "heating" circuit. The power supply is a Kepco model KS- 60M power supply which can operate either as a constant voltage ora constant current source. For large currents (0.1- 10 amperes) the 0.001 ohm shunt is used to determine the heating current. For the lower currents, used in the tungsten annealing experiments, typically a few hundredhs cf manpere, the 0. 02 ohm shunt is used. Both of these shunts are four-terminal resistors with their potential terminals connected to the potential circuits as indicated. They were both accurately calibrated against the 1 ohm standard in the measuring circuit. Switch S4 is used to disconnect the 0.02 ohm shunt and other components in this branch from the circuits when the higher currents are used for other experiments; for all the experiments reported here this switch was positioned as shown. The 2 ampere fuse is used to protect the low current branch if S4 is in the wrong position and high currents are used. The 4 millihenry inductor and the 0.25 microfarad capacitor form a parallel tuned circuit to keep the AC circuits from being shorted by the quench speed capacitors and the power supply. The diode is used to keep the current impulse used to light the specimen from discharging through the power supply as discussed above. 85 can be used to short the diode; for the present experiments it was always left open. Switch 86 is used to disconnect the power supply and let the quench speed capacitors discharge through the specimen as described earlier. Figure 8 is a schematic drawing of the quench speed capacitors. With this circuit, the capacitance can be varied from zero to 4100 microfarads in twelve steps. 32 - IPOWER 1+ 1 LSUPPLY ] /__-_-___--___-__-_-- ’e 0.00m To POTENTIAL SWNT CIRCUITS 34,—} if ,e $092.0 To POTENTAIL SHUNT CIRCUITS fl-b Figure 7: Heating Circuit 33 muoufionmmu Honuaou eccmm Susana "m mndmflm F. U) ----4-- - --——-- 34 Figure 9 is a schematic drawing of the capacitor charging circuit which is used to light the specimen. T1 is an isolation transformer; T2 is an autotransformer. The meter is calibrated to read the charging voltage (0-100 volts), The 1 kilohm resistor limits the charging current. The 10 kilohm resistor is connected directly across the capacitor terminals to insure that it discharges when the system is turned off. It also allows the capacitor to discharge faster if the Variac output voltage is decreased (i.e., it decreases the time constant RC of the 120 microfarad capacitor). S is a single pole, single throw, spring loaded open switch used to dis— 7 charge the 120 microfarad capacitor through the specimen. c. Potential circuits Figure 10 is a diagram of the potential circuits. Switches S8 and S9 are double pole, twelve position switches made by Grayhill (number 44D30- 02-1-ADJ—N). These switches are connected in parallel so that either of the potentiometers can be used to read the potential of interest. PotentiometerA is a Honeywell model 2780 with a Leeds and Northrup DC Null Detector (No. 9834). Potentiometer B is a Honeywell model 2779 microvolt potentiometer. A photocell galvanometer amplifier (type 5214/ 9460) and a secondary galvanometer (type SR21/9461),made by Guildline Instruments Limited, are used as a null detector for this potentiometer. The reversing switch precedi ng the microvolt potentiometeris a Honeywell model 3565 thermal-free reversing switch. It is used to reverse the poten— tial when the current through the specimen is reversed during room tem- perature and helium temperature resistance measurements. The letters "a" through "h" correspond to the same letters on the standard and shunt resistors in the measuring and heating circuit diagrams; 35 uwsouflu mcfimumgu Houflonmmo "m cusmflm v.0. x0 20.90% V. ON Se 89 - 9 A I oozn STD in“ I- g a 4PT”* = mg 2. :3 0.00m STD? 3: = (0’ V 0—-——-—© #' i In 5T0 ’h‘ IN MEASURING CIRCUIL‘L v V v = - fi’y‘ VSPQC :5 {ll 4‘} e 6 VSPQJ V, Issv 50K DIVIDED HE“ VSpec IOK DMDED VSpeci‘e‘i 8 500K l2 I V(ca|) :9 r [’ .K) ICI . I 0 II II. Figure 10: Potential Circuits “'93 < POTE pv TIOMETERB; ,2, 0.0l 37 "i" and " j" connect to the potential leads of the specimen. These account for the first five positions of the switches. Position 6 reverses the poten- tial leads to the potentiometers. When the specimen is heated, the voltage across the gauge length is on the order of ten volts. Since this is too large for the potentiometers to measure, the voltage divider shown in the lower part of Figure 10 was designed and built. Voltages as high as several hundred volts can be mea- sured with this circuit and a potentiometer capable of reading 1. 5 volts. For the present experiments switch 812 is always closed, shorting the 500 kilohm resistor and permitting a voltage ratio of ten to be used. Before using the voltage divider circuit it must be calibrated. This is done by first opening 810 to disconnect the specimen and closing $11 which connects the mercury cell into the circuit. Potentiometer A is then switch- ed to position 9 and the voltage measured. This voltage is then divided by the desired ratio, potentiometer A is turned to position 7 and the 50 kilohm helipot adjusted until the potentiometer measures the divided voltage. The 50 kilohm helipot is then looked into place, and the readings checked again. 511 is then opened and S10 closed. The specimen voltage divided by the ratio is then measured by measuring the potential at position 7. Potentiometer A is used to measure the voltage across: the 100 ohm standard for room temperature resistance measurements; the 1 ohm standard for helium temperature resistance measurements; and the gauge length of the specimen when it is heated. Potentiometer B is used to measure the voltage across the gauge length of the specimen for room and helium temperature resistance measurements, and across the 0.02 ohm or 0.001 ohm shunt resistor for high temperature resistance measurements. 38 2. AC circuits Since the specimen's temperature is changed by varying the direct current through it, the DC voltage across the gauge length of the specimen is not directly proportional to its resistance. For fast quenches the current is turned off, hence there is no DC signal across the gauge length during the quench, making it impossible to determine the quench speed directly. To avoid these difficulties a small, constant AC current of 5000 hertz (Hz) is passed through the specimen continuously during the experiment. The AC signal across the gauge length of the specimen is then directly propor- tional to its resistance, regardless of the direct current flowing through the specimen. This technique could also be applied to conventional quench— ing experiments, and would be especially useful for "air" quenches when the specimen is quenched by turning off the heating current. Figure 11 is a block diagram of the AC circuit. It consists of four major sections: 1) a source of constant AC current which is passed through the specimen in parallel with the direct currents, 2) an AC amplifier, 3) an AC to DC converter, and 4) read out devices. To assist the reader in understanding these circuits, a brief description of operational amplifiers is given in Appendix B. a. Constant current source The AC constant current source drives a constant alternating current of several milliamperes through the specimen independent of the specimen's resistance. A schematic diagram for this device is shown in Figure 12. The generator "G" is a Hewlitt Packard model 202C Audio generator. The operational amplifier is a Philbrick P65AU. It has an open 100p gain of better than 20, 000 and a maximum output of 2. 2 milliamperes at 22 volts 39 mUHsoHHU Um use no Enummwn xoon “Ha madman _ 800m; — «088% m womaom - «.3358 3:224 w pzmmmao .AIIJ III, C 4 E hzflrszo _ $52 8— 00-04 0 w 04 _ $5259 .250. o - 40 OOH—90m HGOHHUU #GMPmCOU 04 W it». a L 3400 Q "NH wusmwm uh 41 peak to peak. The diodes are Sylvania 485 silicon diodes; R2 is a ten-turn helipot; T3 is a filament transformer (117v. to 24v. ). R1 and R2 form a voltage divider which permits fine control of the voltage em. Since point '8' is a virtual ground, the current input iin is ei n/ 1000. As described in AppendixB this current flows through the primary of the coupling transformer T regardless of the impedence seen 3 by the secondary as long as the operational amplifier is not overloaded (i.e., it is a constant current source). T3 is a step—down transformer usedto match the impedence of the specimen to the operational amplifier. Since there is a DC potential between the points A and B, whenever a DC current is flowing through the specimen, a capacitor C1 is placed in series with the secondary of the transformer to keep it from shorting the specimen. However, during a quench this capacitor must discharge through the spec- imen and T3. Hence, to get the fastest quenches and to minimize overload- ing of the operational amplifier, Cl should be small. On the other hand, C1 shouldbe large enough to pass the AC with little impedence. This problem is solved by using the series tuned circuit (tuned to the operating frequency) consisting of C1 and L1 (assuming that T is a "perfect" transformer and 3 that the operational amplifier is "resistive"). The diodes are used to protect the operational amplifier. If the spec- imen breaks or is disconnected, the operational amplifier will stillitry to maintaina constant current. This would severely overload the operational amplifierif it were not for the diodes which conduct when the voltage rises above a few tenths of a volt. In normal operation the voltage across the diodes is at most a few millivolts; with a forward bias this small, the diodes have a very high impedence (essentially infinite). 42 B. Amplifier Figure 13 shows a schematic diagram of the amplifying section of the AC circuit. T 4 is a United Transformer Company No. H-20 transformer. It must be capable of operating at 5000Hz, and the resistance in the windings must be low since it is part of the input resistance of the operational amplifier. Also, it must be well shielded since it passes small signals and rather large stray fields are present. The capacitor C3 is used only as an added safety to keep direct current out of the operational amplifier circuit. R 4 is the input master for the operational amplifier. The operational amplifier is a Philbrick P45AU. It has an open loop gain of 300, 000 and a maximum output of 20 milliamperes at 20 volts peak to peak. The transformer T4 enables one side of the signal to be grounded. As above, capacitor C2 is needed to block the direct current from the amp— lifier circuit. The AC signal is on the order of millivolts, whereas the DC potential is around ten volts before a quench. At the time of a quench the capacitor C2 must discharge through the diodes and transformer. Since the diodes don't "turn on"until there is a few tenths of a volt across them, the amplifieris overloaded with a pulse more than a hundred times larger than the signal. For this reason C2 must be kept as small as possible so that the pulse is over quickly, allowing the AC circuits to follow the resistance of the specimen as it cools. Many attempts were made to keep most of this pulse from the amplifier. The most successful was the series tuned circuit consisting of C2 and L as shown. L2 blocks most of the pulse 2 from the transformer thereby forcing most of it to flow through the diode. As in the constant current source the voltage across the diodes in normal operation is only a few millivolts, leaving them with essentially infinite impedence. 43 umwmaaflfid 04 and onsmflm ---———------ p o..- dfird Q) 0 SPECIMEN 44 To decrease noise and to help remove the spurious pulse discussed above,afrequency-selective network was used for the feedback impedence. As discussed in Appendix B' the gain of the operational amplifier is _I_tf_ . If. Rf is small for all frequencies except a narrow band of frequenciesfmdnly this narrow band will be amplified by the amplifier. A twin T network is used for this purpose. It is called a twin T network because of the shape of its schematic (see Figure 14). The resonant frequency is given by w: 1/RC. The twin T network was made by carefully mountingprecision components on a terminal strip. A 380 picofarad capacitor was used instead of a 360 picofarad (2 x 180 picofarad) because a 360 picofarad capacitor was not immediately available at the time. At the resonant frequency the impedence of the twin T was about 10 megohms. This network determines the operating frequency of the AC circuits. c. AC to DC converter The output of the amplifying circuit e 0 is a 5000Hz signal whose am- plitude is proportional to the resistance of the specimen. The circuit shown in Figure 15 is used to convert this signal to a DC signal which is directly proportional to the amplitude of the AC signal. T 5 is a 6:1 step-up trans- former, which steps up the output voltage to the diodes. Since the diodes are not fully turned on below approximately 1/ 2 volt, non-linear response results if the AC signal is not much larger than this. The typical output of the operational amplifier is 10 volts peak to peak. With the 6-fold voltage increase, the diodes typically rectify 60 volts peak to peak, and the non- linearity in the first half volt is negligible. The 40 kilohm resistor is placed before any capacitors so that the diodes see a resistive load. If the filter capacitor were placed before the major resistance of the load, 45 xnozumz a case “ea musmflm nmw xom ._maomn «\m whom. a. Lao x0m. 46 R, 1330 136 i Fl!) l Q) Li“.- AC to DC Converter Figure 15: 47 the diodes would be turned on only a short time with a small voltage drop across them. See Figure 16 and accompanying discussion. The capacitors C 4 and C 5 are used to smooth the rectified DC to pro- vide nearly pure DC for the read out devices. Some care must be taken to use low values for these capacitors for two reasons: 1) to measure the quench speeds, the time constant should be less than a few milliseconds and, 2) the overloading pulse through the amplifier at the beginning of the quench shouldbe quickly discharged from the capacitors so that the circuits can again indicate the specimen's resistance. d. Readout devices At this stage the signal is a DC voltage which is proportional to the resistance of the specimen. To "read" this signal four devices (connected as shown in Figure 11) are used, a DC meter, a digital voltmeter, a strip chart recorder, and an oscilloscope. The DC meter is a Weston model 271 which reads full scale at 50 microamperes. The meter shunt resistor RS is chosen to give full scale deflection when e is nearly at the maximum output of the Operational am— 0 plifier. The digital voltmeter (Data Technology Corporation model 323) was acquired after most of the present experiments were finished. Its read— ings are accurate to within :10. 01 per cent, much better than the DC meter. After calibration the AC circuits give the resistance ratio in the region of interest to within a few tenths of a per cent with the digital voltmeter; they were accurate to only a few per cent with the DC meter. The DC meter is now used only as a check on the output of the AC circuits. The recorder is a Honeywell model 153X11V-X-28 strip chart recorder. 48 AC INPUT R7? LOAD /\ /\A A ~ CO The diodes conduct for the full cycle for resistive loads AC 'NF’UT DC OUTPUT CAPACITANCE LOAD A A AV 7 Ev; ! ‘35 {\‘x . V ' j ’ l— I | l , I | When the output has a capacitance 6+6 load the diodes conduct only a C t d COhdUC'l' short time and then at a much condUCl'A smaller potential difference AV C "’0!” dISEharges than for resistive leads. through lOad Figure 16: Diode Output to Resistive and Capacitive Loads 49 A voltage divider must be used at the recorder input because the recorder gives full scale deflection for 10 millivolts input. The recorder is used to record resistance of the specimen as a function of time during a run. The recorder chart then allows one to have for future reference a permanent record of: the temperatures at which the specimen was annealed and for how long it stayed there, the time the specimen was at the quench temper- ature before a quench was made, and any unexpected changes in the resis- tance of the specimen, e.g., when a specimen breaks or potential leads come loose. The scope is a Tektronix model 531. It is used with a scope camera to provide pictures of the DC output as a function of time during a quench. The scope is triggered by the DC voltage change at point B (Figure 5) when the heating current is turned off. These pictures are then used to deter- mine the quench speed. The AC circuits were operated at a frequency of 5000Hz so that the electronics will respond sufficiently rapidly to accurately depict the cooling curve of the specimen during a quench. e. Calibration and evaluation The AC circuits must be calibrated each time a different specimen is used. The calibration is done in the following manner: 1) the specimen is heated to some intermediate temperature determined by eye; 2) the spec- imen's resistance is measured with the DC circuits, and divided by its room temperature resistance to give the resistance ratio; 3) the AC current through the specimen is adjusted to make the digital voltmeter (or DC meter) read this ratio. After this calibration, the digital voltmeter (or DC meter) indicates the resistance ratio of the specimen to within a few tenths of a (or a few) per cent. The AC circuits make it relatively easy to get 50 the specimen to any desired temperature by minimizing the number of hunt and test operations necessary to obtain this temperature. We intend ed to use this circuit in conjunction with other circuits to allow us to program the power supply to give slower quenches of any de- sired speed and shape. However, I found that slower quench speeds could be obtained quite satisfactorily by merely connecting a capacitor in parallel with the specimen and allowing the capacitor to discharge through the specimen, as discussed above. If capacitance discharge technique is not suitable for some future application, the programmed system could still be tried. IV. TUNGSTEN QUENCHING RESULTS A. Qgench Speed Data As stated in Section III C, the AC circuits are used in conjunction with an oscilloscope and Polaroid camera to determine quench speeds. In Figure 17 two of these pictures are drawn to scale. The vertical scale is proportional to the resistance of the specimen; since the resistance of tungsten is very nearly linear with temperature above room temperature, this scale is essentially proportional to the temperature of the specimen. The horizontal scale is proportional to time. A simplified fast quench speed picture, taken during a quench from 2905K, is shown at the top of the figure. The time scale is 20 milliseconds per division. The upper horizontal line is obtained by triggering the scope while the specimen is at the quench temperature; it marks this temperature on the vertical scale. The lower line is obtained by triggering the scope after the quench; it gives a base or zero temperature line. The curve between these lines gives the temperature of the specimen as a function of time during the quench. The quench speed is obtained by finding the slope of the first part of this curve as indicated by the dashed line in the figure. The first part of the curve is used because the vacancies are much more mobile at the higher temper- atures, hence most of the vacancies lost during the quench are lost during this time. The fast quench speed calculations are accurate to about 10 per cent. 51 (arbITr-arg Scale) SpeCImen Tamer-ature 52 l l 1 I l 1 I l l l I L i i l i l l 50 mscyflmaon F i g u r e 17: Quench Speed Graphs 53 A slower quench picture is shown at the bottom of the figure. This picture was taken during a quench from 2580K with an 1800 microfarad capacitor connected across the specimen to reduce the quench speed. The time scale was changed to 50 milliseconds per division. In this picture, the lines marked AC are equivalent to those in the upper picture. Here the slope is much easier to obtain because the curve is nearly linear; hence the slow quench speed determinations are accurate to a few per cent. The oscilloscope we use has a dual trace. The lines in the picture marked DC were obtained by using the second trace to monitor the DC voltage across the specimen. This gives a permanent check on the value of the capacitor used for a quench, since different capacitors discharge at different rates. For fast quench speeds there is no DC curve, of course, only an upper DC and lower DC line. These lines were left out of the upper picture. Figure 18 illustrates how the quench speeds vary with quench temper- ature for a 1.2 mil specimen. Notice that the fast quench points have con— siderable scatter at high quench temperatures. This will be discussed later. Figure 19 illustrates how the quench speeds vary with quench temper- ature for a 1. 0 mil specimen. The fast quench speeds rapidly increase with increasing quench temperature, while the slower quench speeds are nearly independent of the quench temperature. Again most of the scatter is in the faster quench speeds. The small numbers by some of the points will be discussed later. B. Fast Quench Tungsten Data Figure 20 shows the fast quench speed data obtained for a 1. 0 mil tungsten specimen. Here the quenched-in resistance is plotted on a 54 AHHEN.HV o seaflommm Mom musumnmmsma noccso wsmne> Ccmmm gonoso 7. . Nb 2. $342.02sz N0 On wN ON N N ON 0. "ma wusmflm _ _ q a] _ q d _ _ _ - _ q _ q . C m. .. - 0m «0:92.40 00m 4 zozuao ems. . . L L _ _ _ P h p L mommnm 202.50 Assam. 205023 .222 s L _ 55 AHHEo.HV d seaflommm you onsumnmmEmB nonmso momu0> 000mm nocmso "ma ensmwm Embimmazmmazwp fillednl A 0.” _|l_|4|ll:N .N 1 VN 102% .54“. o 08me $sz 79.12 2.50023 .22. 1L 0 N) l C) Ow .rel _ _ _ _ _ . . _ loss/9) €_OI x Suva 56 “TM 1 _ #4.. A. . :‘OA. ‘.A. 0‘. ‘ J .. . I __‘ _._ SENT DATA 9 ’- —ORRECTED SCHULTZ _ s C LINE 7 _ 3 ---UNCORRECTED SCHULTZ 6 _ LINE 5 r _ 4 .. _ 5 P 2 t 2 . _ %a 273.4 '94— 1 9 l- .1 8 b n 7 - 4 6 I- fl 5 I- _. 4 I— _ 3 ' 1 2 — _ 'WL " 5 _ 41 +X|O4 (K)L—> Figure 20: Fast Quench Data for Specimen A (1. 0mil) 57 logarithmic scale as a function of the inverse quench temperature. As stated above, the quenched-in resistance is divided by R273to cancel the effect of the specimen geometry. From the equation for the equilibrium vacancy concentration (equation 1, page 1), this data should fall on a straight line with slope -E}'/ k if the quenched—in resistance is in fact proportional to the equilibrium vacancy concentration. The circles are our data and the solid line is a best fit line for Schultz's data (converted to AR/R27 3 and to our temperature scale). The dashed line is Schultz's data without the temperature correction. Our data is in agreement with Schultz's data, except for the low quench temperature points. It is possible that this discrepency arises from a small quenched—in resistance, apparently independent of the quench tem- perature, which should be subtracted from each of the data points to give the vacancy contribution to the quenched-in resistance. This will be shown and discussed in Figure 24. The correction does not affect the higher points because it is only a small fraction of the quenched-in resistance there. The scatter at the low quench temperatures is due mostly to insta- bility in the base resistance. At the higher quench temperatures, the scatter is probably due to variations in the quench speed. The small numbers by the high quench temperature points correspond to the numbers in Figure 19, the quench speeds for this data. There is good correlation between the quenched—in resistances and the quench speeds. When the quenched—in resistance is less than average, so is the quench speed, and vice versa. This correlation was found in all specimens. The originof the quench speed scatter will be discussed in the next section. 58 Fast quench speed data obtained for two 1. 2 mil and two 1. 0 mil tung- sten specimens are given in Figure 21. The solid circles are the data shown in Figure 19. The slopes are nearly the same for each set of data, so that they give nearly the same effective formation energy, EE’ 2 3120.2 ev. However, the 1. 2 mil data lie systematically above the 1. 0 mil data. The most likely explanation for this shift is a systematic change in the tem- perature scales due to different resistance versus temperature curves for different specimens. A short paper published in 1962 by Davis:,18_/ indicates that errors of several hundred degrees are possible for tungsten wires which differ only in trace impurities. The data shifts more at lower temperatures than at higher temperatures as one would expect from a temperature shift. (Note in Figure 20 how Schultz's data changed for a constant temperature shift). If the data for the open circles are system- atically shifted to the left by 110 degrees and the data for the closed circles shifted to the left by 70 degrees, they both coincide, to within the scatter of the data, with the 1. 0 mil data. Another possible, but less likely, explanation for the data shift is that different impurities, grain size, etc, cause different vacancy sink den- sities in different specimens. More or less vacancies would then be lost to sinks during the quench. This would result in different quenched-in resistances for different specimens. C. Dependence of Quenched-in Resistance on @ench Speed Figure 22 shows how the quenched—in resistance varies with quench speed for a 1. 2 mil specimen (specimen C). The open circles and triangle were taken during the first run with this specimen. For some reason they 59 o 10 o .3»! e O .5 . 43 g, o oSPEC A IO ~ A oSPEC 8 "0m" O A ASPEC C " o I ml _ 3- A ASPEC D) '2 ' 2 7r- ‘3 4 6r- A ‘ 1 5r- _. 4.. o o. A .. O A 3r- ‘ _ 43A; 1 2r ‘ A '- AJ 0‘ A ‘ R273 0 0A‘ -4 0 IL 0 9; O . A‘ A j 8P- 0 . A -I 7 ‘ ‘ 6: o . 1 5i- _ C. ‘ ‘ 4F 6 45 _ 3 P ° - Q 0 2 - ° 1 A (9 I6 I 1 WW I 1 I 4.. 3T 35 3.7 3.9 41 45 +on4 («r—— Figure 21: Fast Quench Data for Specimens A, B, C,and D 7 r- 6 ._ _ 5 i- _ 4 - A 3 - _ 1.. _ QB. R27:5 -4 I04 _, SI— ._ s— _ 7 ~ _ 6 I. .— 5 P i 4 L - \ 3 - \ 0,11 181' RUN ~ 0 FAST OUENCH 2L A 900m CAPACITOR T . IBOO’uf \\ O U 4|OO#‘F II 8 BIOOFLI‘ n IWI I ll. 39 4a 4.3 3. 60 or. 2:.0 0) .0 no 0 Specimen C (1. 2 mil) llll l I “T‘XIO4 (Ki—— Figure 22: Dependence of Quenched—in Resistance on Quench Speed for Specimen C. 61 are systematically higher than the data taken on subsequent runs, hence they weren't considered when the fast quench speed line was drawn. The circles are fast quench data with the quench speeds shown in Figure 18. The data for the triangles were obtained by connecting a 900 microfarad capacitor across the specimen to lower the quench speed. The other points were obtained by increasing the capacitance as shown, thereby lowering the quench speed. The dashed line in this figure and the next will be discussed in Section IV. F. Figure 23 shows how the quenched-in resistance varies with quench speed for a 1. 0 mil specimen (specimen B). These data show that the quenched-in resistance is dramatically affected by the quench speed. In later sections these data will be used to extrapolate to infinite quench speeds, and are compared to the predictions of Flynn's theory of vacancy annealing during a quench. Figure 24 shows similar data for another 1. 0 mil specimen (specimen A). The quench speeds for this data are given in Figure 19. Data for a larger temperature range is shown here; the pre- vious data all ended at an inverse quench temperature of 4.3 x 10.-4 This data shows that the quenched-in resistance does not continue to decrease with decreasing quench temperature as the vacancy concentration formula would indicate. Apparently for this specimen there is a small, temperature independent, contribution to the quenched-in resistance. We suspect that it is due to some impurity, probably carbon, being quenched into the lattice. The solubility of carbon in tungsten varies with temperature, M reaching a maximum at about 2700K. When the specimen is slowly cooled to get a vacancy-free base, most of the carbon precipitates out of 3333 2 42 25%00 ZZFZZ I IBOOH.f Plat llll T l i?“ N on 5 Maxim ll 9 8 7 6 5 4 \ oFAST QUEMZH @553,le CAPACITOR 3 F \ A900pf " 2 _ AIZSOLL'F " Lllllll illllll I [fill] L53. 23 \ 5: \ 4— 37 o 2_ SPEC B (LOmII) \ \ -s ”at is 4‘0 4 ,,7fis Sb +XIO (K)“" Figure 23: Dependence of Quenched—in Resistance on Quench Speed for Specimen B 63 C———T 2842 25% 22 2 ZQOO 3333 ‘3' I I l I l I I7 I I I I I I l I I I lQ__ d 9— -I 8" .FASTQlENCH 1 E: Z A 900/1? CAPACITOR _I 5" .1800” H '"l 4‘ o4IOOpf " ‘ 3;— -I 2— ._ |-4L_ O _ g e .. 6- _ as..- _ - , .. R2734— - 3- o _ 2P .1 0 Us ' o .1 3L 2 6' -I 5'— —I 4- _ 3- .. 2L _ SPEC A team) IO6 I I I I _llllJllllllllllllL L 3.0 35 55 45 g 5.0 Figure +X|O4 (KY . 24: Dependence of Quenched—in Resistance on Quench Speed for Specimen A 64 the tungsten lattice to various sinks. However, during a quench the carbon is quenched into the lattice; it doesn't have time to leave the lattice before the temperature is too low for the carbon to move. This additional carbon in the lattice would increase the helium temperature resistance of the specimen. This is the quenched—in resistance which we propose to sub—- tract from the data for specimen A shown in Figures 20 and 21. We do not have sufficient data to accurately determine the quenched-in resistance to be subtracted from the other specimens in Figure 21, butafew low temper- ature data points on each specimen indicate that the quenched—in resistance is considerably lower than for specimenA at the lower temperatures. The three high points in the lower right in Figure 24 are the first two points of one run and the first point of another run, taken before the base resistance had stabilized. This illustrates how critical the base resistance is for low temperature quenches. Figure 25 shows the data of Figure 24 with this constant quenched-in resistance subtracted from each data point. The data now decrease nicely into the next decade. The dashed line will be discussed in Section IV E. D. Additional Tungsten Data We have some fast quench data not shown in Figure 21. This data falls into four categories. First, we have data from high purity tungsten specimens which broke after a few quenches. This data is in general agreement with the data in Figure 21. Secondly, we have data from 2.0 mil tungsten specimens obtained from the Mate rials Research Corporation. This was their highest purity (99. 999 per cent) tungsten. However, even . . 3‘ I 2 AHK 222K 2000K _ \ j i 0 FAST QUENCH j \ moon? CAPACITOR : 4 v \ . 'Bmp'f II _I \ 040011.? " - 21 \ >< AV. .of 6 0‘5 . t F4 \ - it ' - O \ _ AB. 3: A 2 R273. 5- \ - 4__ A . o _ 3 i- I . I \ . _ \ I- .1 - 'I g: n . Z 6 _ . 5 - .1 4... - 3 - _ 2 ,_ SPEC A (LOMII) CI '. _, D D ,‘6 g l_ 1_ Cl 3. 3 4.0 _4.5 50 Figure 25: Dependence of Quenched-in Resistance on Quench Speed for Specimen A with constant subtracted 66 after extended annealing directly above superfluid helium we could not get the resistance ratio to rise above 400; with ratios of 250 more typical. This data has more scatter than that given above. The low temperature data was a factor of ten higher than that in Figure 21; while the high tem- perature data was only a factor of three higher. Thirdly, data on some Westinghouse HRE grade (lamp filament wire) tungsten were also obtained. After long anneals above the superfluid we obtained ratios of about 200. This data fell about a factor of two lower than the data in Figure 21, and again had substantial scatter. Fourthly, we have data on a very pure 2. 0 mil tungsten specimen, kindly given to us by J .M.Galligan. This specimen was fabricated by carefully etching a high purity single crystal rod down to 2.0 mil diameter; it was not drawn. After being mounted on its holder, this specimen had an initial resistance ratio of 2670. A short series of anneals above the superfluid increased this ratio to over 5000. The data obtained from the first run with this specimen were in excellent agreement with the data from specimenA (1.0 mil). The data obtained from the second run were shifted upward by about 50K, in better agreement with the 1. 2 mil data in Figure 21. The data from this very pure specimen confirms both the magnitude and the temperature dependence of our other data, and indicates that the shifts in the quenched-in resistance are present even in the purest specimens. These additional sets of data suggest that for purity less than that of our 600 ratio specimens, both the magnitude of the quenched- in resistance and effective formation energy are dependent on the purity of the specimen. 67 E. Experimental Accuracy of the Quench Data The accuracy of the present quench data is limited by four factors: 1) uncertainties in temperature determination, already discussed on page 58; 2) carbon retained in solution as discussed on page 61; 3) changes in the base resistance, and 4) variations in the quench speeds. All other sources of error are insignificant compared to these four. Wehave checked two other possible sources of scatter. We found no difference in the quenched-in resistance for specimens annealed in hydrogen or helium Also, there was no difference in the data when it was taken with ascending quench temperatures as opposed to descendirg quench temperatures. The accuracy with which the specimen's temperature can be determined depends upon how accurately its room temperature and high temperature resistance can be measured, how constant its average temperature remains when it is hot, how uniform the temperature is along the specimen gauge length, and finally, how well its resistance ratio at a given temperature T, R(T)/R273, corresponds to the value on the NBS temperature scale used to estimate its temperature. The specimen glows uniformly along the gauge length as far as the eye can tell. The room temperature and high temper— ature resistance can be measured with a precision of better than 0 .1 per cent. Continuous readout on a digital voltmeter showed that short term fluctuations were less than 0.1 per cent of the high temperature resistance, and that long term variations occurred slowly enough so that they could be followed on the potentiometers. Taken together these three factors lead to an uncertainty in the specimen temperature of several degrees Kelvin. Therefore, the systematic difference in quenched-in resistance observed between 1. 0 and 1.2 mil diameter specimens must be ascribed to differences 68 in the temperature scales, or to differing sink densities as discussed on page 58 . For low temperature quenches the quenched—in resistances are only a few per cent of the base resistance, hence, if the base resistance varies by as little as one per cent, large errors are introduced. To minimize this, the low temperature points were taken only after the base was sta- bilized as much as possible. If one is very careful, it is possible to keep the base fairly stable. For example, the base resistance changed only by 0.008 per cent during the measurements of the four lowest temperature. points in Figure 24. These data were obtained after about fifty higher temperature points. When the base did change during a set of quench points, the base to which each quench point was referred was obtained by linearly interpolating between the initial and final bases, assuming that the base changed by the same amount for each quench. After most of the data was taken, we formulated a possible explanation for much of the quench speed scatter. As stated above, the specimen cools primarily by radiation during the first part of the quench. For this reason, the temperature of the surroundings seen by the specimen is important. If the specimen were in a perfectly reflecting container, its surroundings would appear to be the same temperature as the specimen regardless of the temperature of the container walls; consequently, it would not cool at all by radiation. The dewar used for this experiment was nearly a "perfectly reflecting container"; it was silvered except for two 1-1/4 inch slits, and of course, the top of the dewar. The rate at which energy was radiated, therefore depended on the orientation of the Specimen with respect to the slits in the dewar, and the depth of the spec- 69 imen in the dewar. When the specimen was perpendicular to the line of sight as viewed through the slits, it would see more room temperature than if it were rotated 90 degrees, in which case it would see mostly its own temperature reflected from the silvered walls. To test this idea a few quenches were tried with 1.2 mil specimens in the two extreme orientations described above. The quench from the first orientation gave a higher quench speed than the second orientation (by about 10 per cent). To further test this idea, I built a "radiation shield" out of sheet metal covered with black friction tape. It wentover the specimen as shown below. side view specimen I/\radiation shield holder k/ C Figure 26: Radiation Shield specimen Several quenches were tried with the radiation shield in place. The results were inconclusive;a more detailed study will have to be made to determine if the radiation shield gives faster, more reproducible, quench speeds. With the radiation shield in place, the helium consumption was roughly twice as high for a given temperature, indicating that more heat was being absorbed in the helium rather than radiating out of the dewar. 70 F. Extrapolation To estimate whether our fast quench speed was fast enough to quench in most of the equilibrium vacancy concentration, we extrapolated our data to infinite quench speed. For specimen A, quenched-in resistance values taken from smooth curves drawn through the data of Figure 25 were plotted on a logarithmic scale as a function of the inverse quench speed for a series of quench temperatures. The results are shown in Fig- ure 27. Each curve corresponds to a different inverse quench temper- ature, as indicated. Each of the four quench speeds shown in Figure 25 contribute one of the four points for the constant quench temperature curves in Figure 27. The slowest quench speed data in Figure 25 give the points in the far right of Figure 27. These points are less critical than the others for the extrapolation, hence the large scatter of these points in Figure 25 has no significant effect on this extrapolation. French curves are used to carefully extrapolate these curves to zero inverse quench speed (infinite quench speed). If this extrapolationis assumed to be valid* the quenched-in resistance corresponding to infinite quench speed can be read directly from this plot. The data along the line marked specimen A in Figure 28 show these quenched—in resistances plotted on a logarithmic scale as a function of the inverse quench temperature (same plot as the previous data). This data falls nicely on a straight line with an effective formation energy of 3. 52 0. 2ev. The error bars in Figure 28 represent the maximum possible errors in the quench speed and the quenched-in resistance. These possible errors introduce the :I: 0. 2 ev uncertainty. *For a critical discussion of the validity of this technique see reference 43. ail» 'N OIPUIOICOEU'I N uhmmmla 71 SPEC A ITIITI" .1 .. TI =3.O> Figure 28: Extrapolated Infinite Quench Speed Data 1 lllLll l L 11111 73 The dashedline in this figure is a best fit line to the fast quench speed data for this specimen. A similar analysis was also performed on specimen B (1. 0 mil) and on specimen C (1. 2 mil). Both of these specimens give extrapolation curves similar to these shown in Figure 27. The lines marked specimen B and specimenC in Figure 28 are best fit lines to the infinite quench speeds for those specimens. As in Figure 21, the 1.2 mil data lies above the 1.0 mil data, but they both give an infinite quench speed effective formation energy of 3. 5 ev. Again the re is an uncertainty of about :1: O. 2 ev associated with possible errors in the quench speed and quenched—in resistance values. We did not get enough low quench speed data on specimenD to perform this analysis on it. G. Flynn Theory Analysis We have also attempted to compare our data with the predictions of Flynn's theorygj of vacancy annealing during a quench. This theory assumes that the concentration of single vacancies is changed during the quench only by annihilation of vacancies at fixed sinks (eg. grain boundaries and dislocations). For quenches in which the temperature decreases linearly with time, the fraction of vacancies lost during a quench is shown to depend, to good approximation, only on the product DqTq’Tq, where D is the vacancy q diffusion coefficient at the quench temperature Tq, and Tq is the quench time. For the slower quench speeds the temperature of the specimen decreases nearly linearly with time as shown in Section IV A. For the fast quench speeds the temperature does not decrease linearly with time; but in the high temperature region where one expects most of the 74 vacancies to be lost, a linear approximation is not too bad; this is in fact the way the quench speeds were calculated. For a given fractional vacancy loss during the quench, eg., 50 per cent, the theory leads to the equation, : Z DqTqTq ’ where Z is a constant. Since Dq is proportional to exp (-E¥n/kT), the relation V V Doexp(—Em/kTq) = Z/Tq'rq follows. To test this theory, 1n(1/Tq‘Tq)is plotted against l/T if the data q, is consistent with the theory, a straight line will result. The data in Figures 22, 23, and 25 were used to test the theory. The dashed lines in Figures 22, 23, and 25 are the infinite quench speed lines obtained in Figure 29. Two separate tests of the theory were made. First, the fast quench speed data were assumed to represent infinite quench speed for the Flynn analysis. The upper half of Figure 29 shows plots of ln(1/Tq‘1;l) versus 1/ Tq resulting from this assumption, for all three sets of data. The error bars represent possible errors in the quench speed determinations. These results have considerable scatter, but if a "best fit" slope is determined, one obtains a' vacancy motion energy of 1.1 :h 0. 4ev If the fast quench speed effective formation energy is added to this vacancy motion energy, one obtains a value of 4. 4 ev for Q. This value is considerably lower than the lowest Q listed in Table 1 (5. 2 ev). The second test of Flynn's theory was made by using the extrapolated data in Figure 28 as the infinite quench speed data. The lower half of Figure 29 shows plots of 1n(1/Tq’7;1) versus l/T for this data, again for q _ l I I I l I T l 7 ' 3: oSPEC. A 1 g: OSPECB : 4— ASPECC -+ 3_ — 2__ —l IE- : BE : 6... — 5__ ._ 4— _ 3_ . 2' (a) '— l l l J l l J L 1 l I65 I l l T l l l l l qE OSPEC. : 6_ —- 5_ c-l 4— _ 3_ .— TT 2.. EXfilAtQZev _ qq Ipfi_ _ BE 3 6.- — 5_ u— 4_ .. 3- ._ 2— — (13) I65 1 J_ l 1 L l 1 l 1 . 52 if 40 4.4 48 -|-—XI&——> 701 Figure 29: l/Tq'l'q versus l/Tq Plots for the Fjlynn Analysis 76 all three specimens. There are three additional points on this graph be- cause the fast quench speed data is now included. These results fall nearly on a straight line, allowing the slope to be determined more pre- cisely than above. This data yields a vacancy motion energy of 1.4:!: O. 2ev. If the effective formation and motion energies (3. 5i 0. 2 and 1. 4&0. 2ev) used for this analysis are added together, a value of 4. 91-0. 4 ev results for Q. This is consistant with the 5. 2 ev value in Table 1, but consider— ably lower than the 5. 89 ev and 6. 6 ev values. We conclude that our data appear tobe generally consistant with Flynn's theory, but that scatter in the data, and the lack of a clearly established value for Q make a positive conclusion unwarranted at this time. If our upperlimit of E}, : 3.7 ev is used in the Flynn analysis, a vacancy motion energy of 1.9:1: 0.6 ev results. Hence, if the Flynn theory is valid, it wouldbe difficult to reconcile our data withavalue of Q much over 6.0 ev. V. ANNEALING OF DEFECTS QUENCHED INTO TUNGSTEN In addition to measurements of the quenched-in resistance, some studies were also made of the annealing out of this resistance in order to obtain information about the mobility of the quenched-in defects. For these studies the specimen was first quenched in the superfluid as described above, and the quenched—in resistance measured at the superfluid temper- ature. The helium pump was then turned off and helium gas bled into the dewar until the pressure was up to atmospheric pressure. This was done because it was much easier to get the specimen in and out of the dewar at atmospheric pressure, and because much less liquid helium was used when the experiment was done at 4.2K. The helium temperature resistance was then re-measured, with the specimen about an inch below the liquid sur- face where the liquid should have been warmed to nearly 4.2K by the warm helium gas bled in. This was done to obtain a 4.2K resistance reading, because all of the annealing data was taken at this temperature. The change in resistance between 1. 3 and 4. 2K was small, typically less than one per cent of the quenched-in resistance. The specimen was then removed from the cryostat and placed in the NRC evaporator, which was evacuated to less than 2 x 10"5 Torr. In this vacuum the specimen was resistance-heated to the annealing temperatures. After the specimen was heated to the annealing temperature, its resistance (hence temperature) was held constant to with— in a few tenths of a per cent. 77 r-i 78 A. Isochronal Anneals We first performed isochronal anneals to find the temperature range in which the quenched-in resistance anneals away. An isochronal anneal consists in heating the specimen for a fixed time (eg. 5 minutes) at a series of increasing temperatures until most of the quenched-in resistance anneals away. Figure 30 shows data obtained from four isochronal anneals on two different specimens quenched from temperatures in the 2800 to 3000K range. Here the fraction of the quenched-in resistance remaining is plotted as a function of the annealing temperature. The solid circles and squares are data obtained from the first run on each specimen; the open circles and squares are from the second run on each specimen. These data were taken using five minute holding times. There are three regions in the recovery: 1) the low temperature region; 2) the major recovery stage in the vicinity of 950K, and 3) the high temperature region above the recovery stage. The resistance of the specimen always increased when it was cycled from 4.2K to room temperature and back. However, the resistance usually returned to the original quenched resistance after the specimens were taken out of the cryostat and allowed to warm in the room. In all cases the resistance went back to the original quenched resistance after the first resistance-heated anneal (650-850). In addition, tests showed that for a given specimen this increase in resistance upon cycling from 4.2K to room temperature and back was very nearly the same whether the specimen had been quenched or not. For these reasons, we do not believe that this resistance increase is likely to be associated with vacancies. Rather, we are inclined to attribute it to a surface impurity effect. 79 Mooom on comm Eonm wmsocmso mawfiwommm How mamoacfi HMGOHAUOmH "om wusmam ‘ITV NO_ X 1’: 0.9 0.0. 0.3 Q N. 0.0. law 0.0 0.»... 0.N L I . I m . L . m_mNHmw_u.CDLgND o u Guam 0. - 0 on n ¥N_Onu.r Sin.- 0 was m. 8m~uwaaamo - . 8am xmmmmufiaafii . a. n o i o I COD 0 1 o 00.. Nanchang 0.02“? 80 The major recovery stage at about 950K, in which over half of the quenched-in resistance anneals away, occurred in each run for each spec- imen. The data indicated by the solid squares decreased until only a small fraction of the original quenched—in resistance was left, as one would expect when simple defects are annealing away. The data for the solid circles behaved similarly, but terminated abruptly because the specimen broke. The open circles and squares, the second run for both specimens (the solid circle specimen was spot-welded together outside the new gauge length), begin to increase after reaching a minimum of about 45 per cent. After the resistance of these specimens increased, we tried some quenches from 2000K to 2500K. The quenched-in resistances were one hundred times higher than the "normal" quenched—in resistances. This strongly suggests an impurity problem. After a high temperature (above 2700K) quench or anneal, further quenching gave "normal" quenched-in resistances, showing that this anomaly can be removed. Note that the quench temperatures are in the 2800 to 3000K range. Since the solubility of carbon in tungsten reaches a maximum at about 2700170] quite near the quench temperature, we suspect these increases are caused by carbon redistributing itselfin the lattice. To test this hypothesis we performed two more isochronal anneals on a specimen which had been quenched from over 3200K. At this temperature the carbon solubility decreases by a factor of three from its value at 2800K. The results are shown in Figure 31. The holding time was increased to ten minutes so that the time taken to get to the annealing temperature was a smaller fraction of the total annealing time, allowing more precise data to be obtained. The solid circles represent the first run and the open 81 Moomm m>onm All 0.x: » m. w. v. N. o. m w e N d _ - A q _ 1 - 4 .0 o - . 3 d .099 €60: < 8% .. o a 1 Av . o 1 a w - motive?" 9Q r [a .. 30.5mm... mud o . xwemnu ._.o o - . G O xmam . a». — _ _ monsumnmmfima Eoum monocmso d QoEflommm How mammccd HMGOH£00mH "Hm musmflm @100! 211%? 0.0209ch 82 circles the second run. In both runs the quenched-in resistance decreases to a small fraction of the original quenched-in resistance. This time the data are very reproducible from one run to the next. The recovery stage is now centered about 850K, approximately 100 degrees lower than for the specimens quenched from 2800 to 3000K. We conclude that most of the resistance quenched into tungsten anneals away in a single, Sharp recovery stage in the range of 800-1000K. This is in agreement with Stage IV recovery in radiation damage studies which Galligan et. fié—O/ showed to be due to vacancy migration. We have analyzed the data of Figure 31“ using the chemical rate equation 18,20,45 / frequently used in radiation damage studies. We do not propos e to justify this analysis; it is used only to compare the motion energy obtained from our data to those c 31 c ul ated from radiation damage studies. The 42/ chemical rate equation can be written as: g? - -V exp(-E}’n/kT) (4) where Pis the quenched-in resistivity, t is time, V is the frequency factor, Xis the order of the reaction. This is the same as equation (3a) except that F(o) is assumed to beVP? For an isochronal anneal, T is constant during a given heat pulse allowing (4) to be integrated directly. Fora first order process X: 1 and (4) integrates to mpg/of) = VTexp(-E}’n/k'r). (5) For a second order process, X : 2 and (4) integrates to “'1 -".1 ’ pf - Pi : VTexp(-E¥n/kT). (6) 83 Here Pi is the vacancy resistivity at the beginning of the jth_ heating pulse at temperature Tj , Pf is the vacancy resistivity at the end of this pulse; ’7' is the pulse time. If Pi is assumed to be proportional to ARi, the vacancy resistance (1. e. , [Oi = KARi), then equations (5) and (6) can be written as v lnEéBJJ- VT exp[-Em] (7) AR ' ‘ kT v (ARf)-1 - (ARi)—1 = KVTBXPE%¥1] 18) AR- AR against .1— and the resulting plot examined for linearity. Similarly, to T test for a second order process, ln[(ARf)"1 - (ARi)-1] is plotted against __1_ . We have performed both of these tests on the major recovery stage T in Figure 31 using the end of the major stage as the base. The first order respectively. To test for a first order process, ln[ln is plotted test gives linear results only over one decade for the temperature range 785K to 855K. If an effective motion energy is calculated, a value in the range of 2. 2 ev results. The second order test shown in Figure 32 gives nearly linear results in over three decades, covering the temperature range 785K to 950K. This analysis gives an effective motion energy of about 2. 5 ev. These effective motion energies lie about midway between the estimates of 1. 7 ev and 3. 1 ev obtained from radiation damage and cold work studies for Stage III and Stage IV respectively. As we will discuss below, isothermal anneals after quenches from about 3200K yield data which cannot be adequately described by equation (4). We therefore tend to believe that this second order fit is at least partly accidental, and that the effective motion energy obtained must be treated with extreme caution. 84 8.. T I I I .— 6: OB=32|5K : 5‘ -l 4b 4 3+— _ 2r— _ | _ Q - ll 3” j 6- _ 5e _ 4— _ 3- .. 1- - TA - 9.151% _, 8: 3 a? gi : a 4- . ’1." 2, ._ Si -4 m... j a: q 6: . Z 5 _ SPEC. A (LO mu) _, 4.. _ 3— - 2~ a mast 0.8 0.9 LO Ll l2 1.5 (VT) “03 (K7—> Figure 32: Second Order Analysis of the Data from Figure 31 85 The data are not consistent with the "first order" kinetics reported by Galligan w for Stage IV recovery in radiation damaged tungsten. Either the process by which vacancies are annihilated is different in radiation damage (1 specimens than in quenched ones, or one of these analyses obtain agreement with simple kinetics by accident. We believe these data caution against oversimplifying the interpretation of isochronal annealing data. To get a more direct estimate of the effective motion energy, involving less initial assumptions, we performed isothermal anneal s . B. Isothermal Anneals 1) Isothermal anneal For an isothermal anneal, the specimen is heated to a given temper- ature and the recovery measured as a function of time. For these anneals, the resistance of the specimen was measured, with a precision of a few hundredths of a per cent, 5 to 8 times during a typical 10 minute anneal. Using these measurements, the temperature was adjusted so that the positive and negative deviations from the temperature of interest were balanced. In this manner, the average temperatures from one 10 minute anneal to the next were held constant to within about 1K. The circles in Figure 33 represent an isothermal recovery curve for specimen A annealed at 798K. The specimen was quenched from 3230K to avoid the impurity problem encountered with the isochronal anneals after lower quench temperatures. The X's are data from another 1. 0 mil spec- imen which broke after the fourth data point was obtained. The most impor- tant point to be made about these two curves is that they are not consistent 86 Hmocnm HmsnmnuOmH "mm wusmflm IIWmSz .2. use 8 0m. 8. o: om. oo. 8 9. 8 d W 4 4 d . . I _. .__.eo._. < .omam 1N. I”. -v. .6 O Lmvflummw. 1o o 1 n. C .I O 1 . x m C . 0 VA 10 O P p — - m L - _ p x *0.— xmm No.0? x 053».qu o Sow-«c. x xmmk "4p 0 87 with equation (4) for any value of x Z 0 . Equation (4) predicts that the magnitude of the slope of the annealing data in Figure 33, should always decrease with increasing time (because P decreases with annealing time), i.e., a curve through the data should always be concave upward. However, for short annealing times the data curve is definitely concave down, and only later does the curve go through an inflection point and become concave W upward. Similar curves have been seen before, and categorized under the name, "S" shaped curves. In general, sucha curve is taken as an indica- tion that the annealing of the vacancies is a complex process, possibly involving vacancy-vacancy clustering and time varying sink densities. This "S" shaped curve was also observed in the isothermal anneals with change-of—slope measurements discussed in the next section. Other than this "S" shaped curve,the data decrease smoothly again consistent with the existance of a single major recovery stage. We had some trouble with temperature uniformity along specimen A during this anneal, hence no quantitative results were obtained from this data. gflsothermal anneals with change-of-slope measurements From equation (3a), we see that if we could measure dP/dt at two different temperatures T1 and T under conditions where F(P) is the 2’ same, we could obtain the energy of motion from the relation: 1 d . EV = HEELg/fltgl (9) m _1_ "1.) T2 T1 Here 3% is the slope of the T1 isothermal curve at a time to; flz (it t1 o l k 0 is the slope of the T2 isothermal curve at the same time to; 88 T1 is the first annealing temperature, and T2 is the second annealing tem- perature. Equation (9) forms the basis for the so—called "change—of—slope" _5_/ technique introduced by Bauerle and Koehler. In this technique, one begins an isothermal anneal at the temperature T1 and obtains enough points to accurately determine the slope of the recovery curve ~39 1. At time to, to the annealing temperature is then changed to T2 which changes the rate of annealing. Enough points are taken at the new annealing temperature T2 to enable one to extrapolate the data back to the point to, and thereby, ob— tain the slope £61 2. All of the quantities on the right side of equation (9) L are now known allowing the effective motion energy to be obtained. After the recovery curve is determined sufficiently to obtain its slope at the second annealing temperature, the annealing temperature can again be changed and the above process repeated to obtain another estimate of the motion energy. This technique has the advantage that only equation (3a) is assumed to hold. The additional restrictions required to obtain equation (4) are not necessary for this analysis. Figures 34 and 35 show the results of such measurements made on two independent 1.0 mil specimens. These specimens had 0. 3 mil , potential leads spot—welded on to minimize temperature non-uniformities; with these leads, no temperature variation was visible along the gauge length when the specimen was heated in vacuum until it barely glowed in a darkened room. Both of these sets of data also have the "S" shape discussed in connection with the isothermal anneals on page 87. The annealing temper— atures and the effective motion energy calculated for each change-of—slope point are marked in the figures. For these experiments to give meaning- ful results, the data must have very little scatter, and the curve connecting E3 02- — 0.1- specs (1.0 mil) O_'—‘2'oJ 46' 5'51 8'6l 1501120 TIME (MlNUTES)—> Figure 34: Isothermal Anneal with Change-of—Slope Measurements O.l SPEC H (1.01m) O1 l l l l l l l 265—4—‘4'0 so ' 80 70750 120 TIME (M1NUTES)——- Figure 35: Isothermal Anneal with Change—of-Slope Measurements 91 the data points must be carefully drawn so that the slopes can be accurately determined. The curves shown in Figures 34 and 35 were drawn through the data points with the aid of French curves. A mirror placed at the point of interest was positioned to make the curve and its image appear continuous. A line was drawn along the edge of the mirror giving the normal to the curve at the change of slope point. The normal to this line then gave the slope of the curve. In this manner, the slopes could be obtained accurately to a few per cent, except possibly where there is extreme. curvature. The data give an effective motion energy of 1. 5:1: 0. 3 ev. It should be noted that the effective motion energies calculated here are from specimens which were quenched from very high temperatures, and only the upper two-thirds of the recovery curve was used. The effective motion energy determined here is in good agreement with the motion energy (1. 4i- 0. 2 ev) obtained above with the Flynn theory analysis. However, these effective motion energies are much lower than those obtained for Stage IV radiation damage studies (see Table 1). They are even somewhat lower than the 1.7 ev motion energies obtained for Stage III. While the agreement on the vacancy motion energy between the Flynn analysis and change-of—slope analysis is suggestive, we do not believe that 5411/ it shouldbe taken ho seriously yet. In several other metals effective motion energies measured after high temperature quenches are found to be substantially lower than the single vacancy migration energy, presumably reflecting the effects of mobile vacancy-vacancy complexes. The "S" shaped isothermal curve suggests complexities of this type are likely to be found in the annealing of tungsten. It is likely that careful annealing studies from a series of quench temperatures will be necessary to obtain a satisfactory understanding of vacancy annealing in tungsten. VI. MOLYBDE NUM We have annealed and quenched molybdenum using the superfluid technique. A 1. 5 mil specimen and a 1.1 mil specimen were annealed for six minutes at about 2350K and cooled in steps, resulting in resistance ratios of 524 and 750 respectively. Further annealing and quenching in- creased these ratios to 620 and 870. Surprisingly, the thinner wire had the larger resistance ratio. Preliminary quench data for these specimens are shown in Figure 20. The 1.5 mil data is three orders of magnitude higher than the 1.1 mil data. Also the 1. 5 mil data is not very temperature-dependent; suggesting that something besides vacancies are being quenched into this specimen. The data for the 1. 1mil specimen appears more likely to be due to vacancies. However, the effective formation energy is only about 1.1 ev, which is w 2_6/ half of the value estimated by Meakin and Kraftmakher. These data are very preliminary and are reported primarily to give an indication of the difficulties which are likely tobe encountered in a systematic study of vacancies quenched into molybdenum. 92 93 8:, T r I T r I I j 6: o 7- 5,. _. 4— O a 3- 0 0 921) - 2 r O '5 O 0 LO..- 2 e-:- E 4 _ _ 3.. l.5mIlMOLYHDEMJM .. 2 1- .. 1 (a) be. 1 I 1 I 1 L 1 m 8 I I 1 r r .. 6 I 5 d 4 .. 3 - _ 2 — _ TlQ-fh E¥N|Jev _ 8:- 3 AR 6 _ .. R273 5 _— _. 4 L 4 3 r 1.1 mil MOLYBCENUM _ 2 r- .1 - (b) 11L6 l l l I l I 35 4.0 4.4 4.8 5'2 5.6 so 6.4 +X|O4 Kfl -—- Figure 36: Preliminary Molybdenum Data VII. SUMMARY We have described an apparatus and technique for superflu1d quenching, and have used this system to study vacancies quenched into fine tungsten wires. We have demonstrated that the quench speeds obtained are fast enough to allow studies of the dependence of quenched-in resistance on quench speed. We have found that the quenched-in resistance varies some- what from specimen to specimen, but that the effective formation energy determined for the fast quench data is nearly the same for all specimens and equal to 3. 1:}: 0.2 ev. Both the magnitude of the data and the effective formation energies are in satisfactory agreement with a similar study by Schultz. By studying how the quenched-in resistance varies with quench speed, we have shown that the equilibrium vacancy concentration is proba— bly not being retained even with the fastest quench speeds. By extrapo— lating our data to infinite quench speed we can get an estimate for the vacancy formation energy, BF. This extrapolation leads to the value BF: 3. 5:1: 0.2 ev. We have also used this extrapolated data to test Flynn's theory of vacancy annealing during the quench. We find that the theory fits the infinite quench speed data, butttwe believe our data has slightly too much scatter to state definitively that Flynn's theory is applicable. As- suming El.” : 3. 5 ev, this theory yields avalue of 1. 4:1: 0.2 ev for the vacancy motion energy. If the maximum value of 3. 7 ev is used, a motion energy of 1. 9i 0.6 ev results. 94 95 Isochronal and isothermal studies were made of the annealing away of the quenched-in resistance. Isochronal anneals were performed on spec— imens quenched from temperatures in the ranges 2800-3000K and above 3200K. There is a Single major lSOChI‘Ol‘lal recovery stage in the temper- ature range 800—1000K, which is consistent with the Stage IV recovery in radiation damage studies. This single recovery stage, in which most of the quenched-in resistance anneals away, suggests that there is a Single type of defect present, presumably the vacancy. The isothermal anneals were made on specimens quenched from above 3200K. These recovery curves have an "S" shape suggesting that the annealing process is complex and does not obey a simple chemical rate equation. Change-of—slope mea- surements yield effeCtive motion energies of 1. 5:1: 0.3 ev, in agreement with the value obtained from the Flynn analysis. Although this agreement is suggestive, it should be treated with caution. Previous experience with effective motion energies measured after high temperature quenches suggests that they are often somewhat lower than the single vacancy motion energies. If we take our "best values" for the vacancy formation and motion energies, we obtain a value of about 5 ev for Q. This is consistent with the lower value of Q reported in recent literature. The maximum value of Q which would still be consistant with our data is Q 29 6 ev. However, our data do not appear tobe consistent with either Q = 6. 9 ev,or with Bin: 3. 3 ev obtained from recent radiation damage studies. During the progress of this experiment, a number of interesting pro- blems were uncovered which require further study. These problems will merely be liSted here: 1) possible deviations from specimen to specimen 96 of the resistance versus temperature characteristics for tungsten; 2) resis- tance increase when a tungsten specimen is cycled from 4. 2K to room temperature and back; 3) increase in resistance during isochronal anneals after quenches in the range of 2800 to 3000K; 4) possible complex annealing processes suggested by the "S" shaped curve; 5) superfluid bubbling phenomena; 6) problems involved in molybdenum quenching. LIST OF RE FERE NCES 10. 11. 12. 13. 14. 15. 16. 17. 18. LIST OF REFERENCES . Girifalco, L. A. Atomic Migration in Cgstal_s, Blaisdell Publishing Company, New York (1964) . Simmons, R. O. and R. W. Balluffi Phys. Rev. 125, 862 (1962) Cotterill, R. M. J. Lattice Defects in Quenched Metals ed. by R. M. J. Cotterill, Academic Press, New York, (1965), p97 . Attardo, M. and J. M. Galligan Phys. Stat. Sol. 16, 449 (1966) Bauerle, J.E. and J. S. Koehler Phys. Rev. 107, 1493 (1957) . Lazarus, D. Phys. Rev. _9_3J 973 (1954) . Schultz, H. Acta Met. 12, 761 (1964) . Kraftmakher, Ya. A. Soviet Physics—-Solid State (USA) 4, 1662 (1963) . Thompson, M. W. Phil. Mag. _5, 278 (1960) Vasil'ev, V. P. and S.G.Chernomorchenko Zav. Lab. _2_2_, 688 (1956) Kinchin, G.H. and M. W. Thompson J. Nucl. Energy_6, 275 (1958) Koo, R.C. Reactive Metals, vol. 2, ed. by W.R.Clough, Interscience Publishers, Inc., New York (1959), pp. 265-274 Schultz, H. Z. Naturforschg. 143, 361 (1959) Schnitzel, R. J. Appl. Phys. E, 2011 (1959) Neimarlg L.A. and RA. Swalin Transactions , Metallurgical Soc.,Am. Inst. Mining, Metallurgical, Petroleum Engrs., _2_1_8 (1960), pp 82—87 Danneberg, W. Metall. _1_5_, 977 (1961) Schultz, H. Acta Met. _1_2, 649 (1964) Moteff, J. and J. P. Smith Flow and Fracture of Metals and Alloys in Nuclear Environments, Special Technical Publication #380 American Society for Testing and Materials, 171 (1965) 97 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 98 Andelin, R.L., J.D.Knight, and M. Kahn Trans. AIME (USA) 233, 19 (1965) Jeannotte, D. and J. M.Galligan Phys. Rev. Letters_1_9_, 232 (1967) Keys, L.K., J. P. Smith, and J. Moteff Phys. Rev. 116, 851 (1968) Flynn, C.P. Phys. Rev. _111, 682 (1968) Gregory, D. P. Acta Met. 11, 623 (1963) Meakin, J. D., A.Law1ey, and R.C.Koo Appl. Phys. Letters (USA)§, 133 (1964) Nihoul, J. Phys. Status Solidi. _2_, 308 (1962) Kraftmakher, Ya. A. Soviet Physics—-Solid State (USA) _6, 396 (1964) Cotterill, R. M.J. , ed. Lattice Defects in Quenched Metals, Academic Press, New York (1965) International Conference on Vacancies and Interstitials in Metals, Sept. 23—28, 1968, Kernforschungsanlage Julich, Germany Rinderer, L. and H. Schultz Phys. Letters (Ger.) _8_, 14 (1963) Martin, D.G. Acta Met. 5, 371 (1957) Borisov, E.V., P.L.Gruzin, L.V.Paulinov, and G. B. Fedorov Collection: Metallurgy and Metallography of Pure Metals (in Russian) (1zd. MIFI, Moscow, 1959), # 1 Peiffer, H. R. J. Appl. Phys. _2_9, 1581 (1958) Danneberg, W. Z. Naturforsch. 16a, 854 (1961) Peacock, D.E. Phil. Mag. (GB) 8, 563 (1963) Askill, A. and D. H. Tomlin Diffusion in Body—Centered-Cubic Metals American Society for Metals, Metals Park, Ohio (1964), p 247 Temperature, Its Measurement and Control in Science and Industry, Reinhold Publ. Co. , New York, (1941), p. 1318 Jones, H.A. and I. Langmuir Gen. Elec. Rev. E, 310 (1927) Davis, G.L. Nature (GB) 196,- 565 (1962) Kuhlmann, H.H. and H. Schultz Acta Met. _1_4, 798 (1966) 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 99 Goldschmidt, H.J. and J.A. Brand J. Less—Common Metals 5, 181 (1963) Kauffman, J. W. and M. Meshii Lattice Defects in Quenched Metals, ed. by R. M. J. Cotterill, Academic Press, New York (1965), p 77 Jackson, J. J. Lattice Defects in @enched Metals, ed. by R. M. J. Cotterill, Academic Press, New York (1965), p 467 Balluffi, R. W., R. W. Siegel, K. H.Lie, D. N. Seidman International Conference on Vacancies and Interstitials in Metals, Sept. 23-28, 1968, Kernforschungsanlage Julich, Ger. Flynn, C.P., J. Bass, and D. Lazarus Phil. Mag. _1_1, 521 (1965) Neely, H.H., D. W. Keefer, and A.Sosin Phys. Stat. Sol. _2_8, 675 (1968) Ytterhus, J. A. , R. W. Siegel and R. W. Balluffi Lattice Defects in Quenched Metals, ed. by R. M. J. Cotterill, Academic Press, New York, (1965), p 683 Federighi, T. Lattice Defects in Quenched Metals, ed. by R. M. J. Cotterill, Academic Press, New York (1965), p 217 de Jong, M. and J. S. Koehler Phys. Rev. 129, 40 (1963) Malmstadt, H. V. and C. G. Enke Electronics for Scientists , W. A. Benjamin, Inc. , New York (1963), pp. 341-383 APPE NDICES 100 APPENDIX 'A Significance of the Superfluid The fact that superfluid is not essential for this experiment was dis- covered accidentally during a test of the "radiation shield" described in Section IV E. The specimen was at a high temperature (about 3000K) in the superfluid; after 2 to 3 minutes the liquid around the specimen began to bubble indicating that the liquid near the specimen was no longer super- fluid even though the pressure was still in the 1-5 Torr range. Further tests indicate that bubbles form around the specimen and float to the top until the specimen is about 3/ 4 inch below the surface. If the specimen is lowered further, bubbles seem to form along the specimen, but do not leave it to rise to the surface. If the specimen is lowered still further (below about 1-1/ 2 inch), these bubbles form only at the ends of the spec- imen. They also appear to be smaller, and again do not float away from the specimen. If the specimen is placed slightly below the depth where the bubbles rise to the surface, there seems to be "pulsation", the bubbles rise to the surface for a while, then stop and start again. The period of these "pulsations" is of the order of a second. During all of these tests, as nearly as one can tell by eye, neither the average temperature nor the temperature uniformity of the specimen changed significantly. This fact was verified by direct readout of the spec— 101 imen resistance on a digital voltmeter. This is direct proof that bubble formation around the specimen does not burn it out, or ruin the experiment. This implies that the superfluid is not necessary. We believe that the superfluid around the specimen turns normal be- cause the radiation shield restricts the flow of the superfluid. This causes the superfluid to reach its critical velocity and turn normal when it passes through the holes in the radiation shield. More study of this phe— nomena is necessary before it can be adequately explained. 102 APPENDIX B Operational Amplifiers To assist the reader in understanding the AC circuits, a brief description of operational amplifiers is given. The symbol for an operational amplifier em :D— e. The principal characteristics of a typical operational amplifier are high input impedence, high gain (100, 000), and good frequency response. The ope rational amplifier alone is no better thana conventional amplifier. The addition of the negative-feedback path (control loop) makes the amplifier circuit stable and suitable for accurate control. A simple operational feedback circuit is shown below. o r _L 8° The feedback impedence Rf ?connects the output of thEamplifier to the input. The amplifier' 3 very high gain and phase reversal makes point 'S' (the summing point)_a "virtual ground',’ i.e., it is ideally at ground potential but not connected to the ground. Hence the input currrent iin is just ein/Rin. Since the voltage at 'S' is zero we see that the output voltage is just 'Rfiin Substituting for iin from above, gives the important result, 60 : -Rf/Rin) em. That is, the output voltage is directly proportional to the input voltage with the gain given by Rf/Rin Since these impedences can be made quite stable, 103 the gain of the operational amplifier can be made very constant. This is important in the AC amplifier system described in Section III C 2). For a constant current source, the load is put in the place of R f’ The output voltage will then always adjust itself so that the current through the load is ein/Rin as described above. This will remain true so long as the operational amplifier is not overloaded. TY "111111111111111111111111111111111116111111“