va FINE§z ,c-u. .0) murmur its: ,1. ”a“. ‘ . .u .. RETUMIIG LIERARY MTERIAL§z ‘.d":“' A,” '. ' :4" *‘g’i'w ~; ‘ . ‘ Mac. in book return to move chm. fro. circuhtion records 4 r,-- .«uA EXPERIMENTAL AND THEORETICAL MODELING STUDIES OF THE SIEGFRIED MASS IDENTIFICATION SYSTEM By Marcello Michael DiStasio A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY 1980 ABSTRACT EXPERIMENTAL AND THEORETICAL MODELING STUDIES OF THE SIEGFRIED MASS IDENTIFICATION SYSTEM By Marcello Michael DiStasio The electrostatic mass identification system, SIEGFRIED was used to identify short-lived nuclear reaction products. A combination of Helium Jet Recoil Transport and Time-of- Flight methods was employed for Al, KCl, Ti, Sm and NaF targets. Both B-mass and y-mass coincidence experiments were performed. All measurements were made using the 70—MeV 3He beam from the Michigan State University Sector-Focused Cyclo- tron. Theoretical modeling of the Time-of-Flight system was accomplished by numerical methods. This modeling was used to explain observed mass peak broadening in terms of the recoil nucleus initial kinetic energy which results from B decay. The possibility of measuring B-decay Q values from observed peak broadening is demonstrated. ACKNOWLEDGMENTS Thank you to Professor William C. McHarris for serving as my thesis adviser and for all his support through my graduate career. I would also like to thank Dr. William Kelly for all the time and trouble he spent in fulfilling his role as my second reader. To Ray warner and Richard Firestone I want to express my appreciation for much help and guidance. Thanks to Dr. D. O. Riska for invaluable discussions and suggestions concerning theoretical matters. I would also like to thank Richard Au and Barb Woodward. Special thanks to Wayne Bentley for countless hours of help on the experimental work for this thesis. Thank you to Peri-anne Warstler for an excellent Job of typing this work. Thanks to Roy Gall and Sandy Koeplin for the special friendship they extended to me. My sincerest thanks to my mother and father for con- tinuing faith and love, all that I accomplish is but a small acknowledgment of their devotion. To my brother and sisters, thank you for the love and support you have always given. Finally, a continuing and loving thanks to my wife, Ernie, for her love and understanding over all these years. ii Chapter LIST OF TABLE OF CONTENTS TABLES. . . . . LIST OF FIGURES I. INTRODUCTION . II. DETAILS OF TOF SYSTEM. III. EXPERIMENTAL RESULTS . IV. RECOIL ENERGY FITTING. V. MODELING THE TOF SYSTEM. . . VI. THEORETICAL MODELING RESULTS VII. CONCLUSIONS. APPENDICES I. SIMPLE TOF MODEL . II. RELATION BETWEEN B-DECAY AND RECOIL ENERGY. . . . . . . . III. DOUBLE 3-POINT INTERPOLATION . IV. FINITE DIFFERENCE APPROXI- MATIONS. . . . . . . . . . BIBLIOGRAPHY. iii Page iv vi 16 75 101 1142 175 177 18L! 187 . 191 195 LIST OF TABLES Table Page III-l LSQ Fit to Al TOF Spectrum. . . . . . . . 25 III-2 Predicted Cross-Sections and Integrated Areas for Al Peaks . . . . . . 28 III-3 Results of Kinetics Study of Al Spectrum. . . . . . . . . . . . . . 37 III-A(a) Properties of Reaction Products from 23Na + 70-MeV 3He. . . . . . . . . . Al III-4(b) Properties of Reaction Products from 19F + 70-MeV 3He . . . . . . . . . . Al III-5 LSQ Fit for KCl Spectrum. . . . . . . . . A9 III-6 Properties of Products of “5T1 + 70—MeV 3H8. . . . . . . . . . . . . . . 55 IV-l Results of SAMPO Fitting of KCl Spectra . . . . . . . . . . . . . . . 8l IV-2 Calculated TOF'S for KCl Spectrum (in nsec). . . . . . . . . . . . 8A IV-3 Fit to KCl Spectra Recoil Energies (no F-GT Correction) . . . . . . 9O Ivsu Fit for Al Spectrum ~- No F-GT Correction . . . . . . . . . . . . . 93 iv Table VI-l VI-2 Fermi-Gamow-Teller Parameters for Peaks in KCl Spectrum Recoil Fit for KCl Spectrum (With F-GT Correction). Fit to Al Spectrum Recoil Energies (With F-GT Correction) Comparison of Relaxation and SOR Results for Infinite Coaxial Conductors. . . . . Comparison of Speeds for Runge- Kutta vs. Predictor-Corrector Routines. Comparison of Bound Values. . . Comparison of Peak Widths for Various Distributions Page 99 100 100 116 luO 15A 168 LIST OF FIGURES Figure Page 2-1 Schematic Diagram of SIEGFRIED. . . . . . 7 2-2 Electronics for TOF mass measure- ment. . . . . . . . . . . . . . . . . . . 10 2-3 SIEGFRIED'S vacuum system . . . . . . . . l2 2-A Schematic for CEMA failsafe . . . . . . . 15 3-1 TOF spectrum for products of 27Al + 70-MeV 3He . . . . . . . . . . . . 2A 3-2 ALICE predictions for 27Al + 3He. . . . . 26 3-3 ALICE predictions for 23Na + 3He. . . . . 39 3-u ALICE predictions for 17F + 3He . . . . . no 3-5 TOF spectrum for 70-MeV 3He on NaF target. . . . . . . . . . . . . . . . A3 3-6 ALICE predictions for 39K + 3He . . . . . as 3-7 ALICE predictions for 35C1 + 3He. . . . . A6 3-8 TOF spectrum for 70-MeV 3He on KCl target. . . . . . . . . . . . . . . . A8 3-9 ALICE predictions for I"STi + 3He. . . . . 5A 3-10 ALICE predictions for ”7T1 + 3He. . . . . 57 3-11 TOF spectrum for products of 70-MeV 3He on Ti target. . . . . . . . . . . . . 58 3-12 ALICE predictions for 58Ni + 3He. . . . . 61 vi Figure Page 3-13 TOF spectrum for products of AS-MeV 3He on Ni target. . . . . . . . . . . . . 62 3-1A TOF spectrum for products of 70-MeV 3He on Ni target. . . . . . . . . . . . . 65 3-15 TOF spectrum for 27Al + 70—MeV 3He reaction products (using NaI as start detector). . . . . . . . . . . . 67 3-16 NaI y-spectrum in coincidence with spectrum of Figure 3-15 . . . . . . . . . 70 3-17 ALICE predictions for 3He + 1""Sm . . . . 71 3-18 TOF spectrum for products of 70-MeV 3He on 1""Sm target . . . . . . . . . . . 73 A-l Comparison of peak width for KCl spectrum. . . . . . . . . . . . . . . . . 76 A-2 Decay of 28Al . . . . . . . . . . . . . . 91 A-3 Fermi-Gamow—Teller recoil energy distributions . . . . . . . . . . . . . . 96 5-1 Two-dimensional x-y grid with dij's at free points after 1 iterations. The grid spacing is the same in the x and y directions. x = 1A, y = 5A, where 1,3 are integers. . . . . . . . .‘. . 111 5-2 Cross section of infinitely long coaxial cylindrical conductors. The inner conductor has radius a = 1.0 and vii Figure 5-3 544 5-5 5-6 547 5-8 Page is held at 83 = 100; the outer conductor has radius b = 10.0 and is held at ob = 0.0. The analytic solution is given in the text. A¢= ¢a - ob . . . . . . . . . . . . . . . . . 115 Finite cylindrical conductor held at ¢ = 100, enclosed in larger grounded cylinder. . . . . . . . . . . . . . . . . 118 Equipotentials resulting from SOR solution to problem geometry of Figure 5-3. . . . . . . . . . . . . . . . 119 Finite cylindrical conductor (with o = 100) enclosed in grounded outer conductor that abruptly expands radius by a factor of two. . . . . . . . . . . . 120 Equipotentials resulting from SOR solution to problem geometry of Figure 5-5. . . . . . . . . . . . . . . . 121 Equipotentials in the acceleration zone calculated by overrelaxation method. . . . . . . . . . . . . . . . . . 125 Example of Z and p components of the calculated electric field in the acceleration zone. For Ez curve x = z, for ED curve x = p. . . . . . . . . . . . 126 viii Figure Page 5-9 Equipotentials in the drift zone calculated by overrelaxation method . . . 131 6-1 Approximate model of the TOF system. . . . . . . . . . . . . . . . . . 1A8 6-2 To - T¢ bound curves for several initial positions . . . . . . . . . . . . 155 6-3 To — T¢ hit efficiency curves . . . . . . 157 6—A Graphs of final 0 position as function of T¢ (for fixed initial 9 velocities) . . . . . . . . . . . . . . 159 6—5 Graphs of final p position as func- tion of T (for fixed initial p velocities) . . . . . . . . . . . . . . . 160 6-6 Recoil energy distributions for pure Fermi and pure Gamow-Teller decays. . . . . . . . . . . . . . . . . . 162 6-7 Recoil energy distribution for mixed decay . . . . . . . . . . . . . . . . . . 163 6-8 Theoretical TOF peak predicted from flat momentum distribution . . . . . 167 6-9 Theoretical TOF peak predicted from Fermi distribution. . . . . . . . . . . . 170 6-10 Theoretical TOF peak predicted from Gamow-Teller distribution . . . . . . . . 171 ix Figure Page 6-11 Observed TOF spectrum of products of 27A1 + 70-MeV 3He reaction . . . . . . 172 CHAPTER I INTRODUCTION The development and refinement of many new and ingenious instruments for investigations of nuclei from the region of B-stability has made such endeavors challenging and re- warding. Experiments that were very difficult ten years ago are fast becoming relatively easy and almost routine. Among the new techniques are the Helium Jet Recoil Trans- port system (HeJRT), the Rabbit system, and a number of new time-of-flight (TOF) mass identification systems. The HeJRT and Rabbit systems were developed to provide a means of fast transport of short-lived products of nuclear re- actions to a low background counting area. This is a very important function when one is dealing with reaction products that have half-lives in the tenths of second range. Besides their short half-lives, another difficulty one must deal with is interfering products. Generally the result of the interaction of a projectile beam with a chosen target is never exclusively the product of interest. One is always faced with the problem of differentiating the species of interest from a crowded field of interfering reaction products. As a means of separating out and studying a particular isotope of interest, on-line mass separation systems have been developed. The time—of-flight spectrom- eter is a conceptually simple and increasingly popular instrument for isotope identification. Our system, SIEG- FRIED, was constructed toward this end. The basic principle of SIEGFRIED, as of all TOF systems, is that nuclei of varying masses may be differentiated ac- cording to the time each mass requires to traverse a given flight path length. In our system each mass is initially accelerated to a known velocity and then allowed to drift a known length. Measurement of the time-of-flight yields the associated mass. In contrast to many other TOF systems, SIEGFRIED was developed as a means of labeling a B-decay event with the associated mass resulting from the decay; it was not originally intended to provide highly precise mass measurements. In this respect SIEGFRIED is analogous to the fast chemical separation techniques used for the study of short-lived nuclei. These techniques were in- valuable in dealing with complex mixtures of radioactive reaction products. Serving a similar function SIEGFRIED is a fast, mass identification system. Our mass identification system is actually a result of coupling a HeJRT system, for transport, to the TOF system, for mass identification. Some of the original work along these lines was accomplished by a group at Texas A & M (Ju7l). They used their MAGGIE system to analyze recoils from a-decay. To demonstrate that recoils from B-decays could be mass analyzed by TOF methods the MSU SIEGFRIED project was initiated (Ed76). As with so many supposedly simple instruments unexpected results often manifest them- selves. Closer inspection of the details of our TOF spectra leads one to suspect that a great deal of interest- ing information about the nuclear decay associated with an observed recoil mass, is being preserved in the observed mass peaks. In order to confirm or refute such suspicions, it is necessary to gain a clear understanding of precisely how the SIEGFRIED TOF system works. In a large measure, this will be a main thrust of our discussion. In this thesis we will attempt to dovetail experimental results in various mass regions with a rigorous theoretical modeling study of the system. Of course the model will not be an exact simu- lation of the system yet we will strive for the most real- istic representation possible while still practical. Before proceeding to the bulk of our presentation I would like to point out that our system has resulted directly and indirectly from the work of many researchers in the area of TOF measurements. As is the case with many experimental systems, SIEGFRIED has provided us with a number of suc- cesses, a fair share of frustrations, and a plethora of intriguing potentialities. Among the successes we should first count the fact that SIEGFRIED has proven that B-recoil ions can be routinely mass analyzed. This is something that only a few years ago was an uncertain hypothesis. Second, as will be shown in later chapters, we have employed SIEGFRIED for mass identification over a range of dif- ferent mass regions with fair success. These experiments have been performed, in part, to investigate the utility. of, and difficulties associated with our system when applied to a variety of cases that one would expect to obtain reason- able results. Third, we will demonstrate in this work that a prime bit of information that can be obtained from our mass peaks is the initial recoil energy of the observed B-decay product. This is a quantity that is directly related to the B-decay energy. CHAPTER II DETAILS OF TOF SYSTEM Since a detailed analysis of our Time—of-Flight (TOF) system will comprise a significant portion of subsequent discussions it is worthwhile at this point to provide a brief exposition concerning the design and use of the SIEGFRIED spectrometer. Many aspects of our instrument and the TOF measurement process are not unique to SIEGFRIED but are common to many TOF systems. In the following presentation the only details relating to SIEGFRIED that will be stressed are those that represent changes from the original construction. They are modifications effected by the author that should be noted by future users of SIEGFRIED. For a full discussion of the design and construc- tion of SIEGFRIED, see reference (Ed76). In a "typical" experiment for TOF measurements, nuclear reaction products are generated with the MSU Sector-Focused Cyclotron and transported to SIEGFRIED by means of a Helium Jet Recoil Transport System (HeJRT). In this HeJRT system a gaseous mixture of He, impurities, and reaction products flow through a capillary tube and into an evacuated chamber of SIEGFRIED that is called the skimmer chamber. The gaseous mixture is directed at a conical skimmer which acts to remove most of the helium gas (See Figure 2-1). The reaction products are attached to large molecular "clusters" which have a smaller divergence as they flow from the end of the capillary and so have a higher prob- ability of passing through the skimmer assembly and striking the stainless steel collecting plate in the collection chamber. These molecular clusters stick to the collecting plate and provide a very thin source. The metallic collect- ing plate is held at a static, positive high voltage, usually +6 kV with respect to SIEGFRIED ground potential. When one of the radioactive species undergoes a B-decay, a simple sequence of events occurs that is the basis for our time-of-flight measurements. As the nucleus of interest emits the 8 particle, it will recoil to conserve momentum and leave the surface of the collector. Also, when a nucleus emits a 8 particle, there is a resulting sudden change in the nuclear charge from Z to Z i 1. This very fast change causes a pertubation on the electrostatic po- tential in which the atomic electrons move and often results in the ejection of an atomic electron. Studies (Ca63) have shown that in the majority of ionizing events a +1 charge state results. The positive ions so created are then accelerated across a region of nearly uniform electric field, trans- forming the 6 kV of potential into kinetic energy. The .QmHmmcmHm co anemone oHooEocom .Hnm cnowfig _ utmgl _ no Ommm) _ .Hvov ..oEE.xlev|/ / 525:...” 000.com 3.82.8 acceleration zone is in effect a circular plate capacitor with the collector at 6 kV as one plate and a disc formed of fine wire mesh (90% transmission mesh) acting as the ground plate. After passing through the wire mesh disc, which ef- fectively completes the ion acceleration, the recoiling ions enter a weak field region we will refer to as the "flight tube". In this portion of the instrument there is a very small diameter (0.005 cm) wire held at a small nega- tive voltage (-110 V) concentric with a relatively large radius (5.25 cm) pipe that is held at ground. The result of this configuration is a region of logarithmic potential: the wire, commonly called an Electrostatic Particle Guide (ESPG), acts to focus the recoil ions in the radial direc- tion and onto a set of Channel ELectron Multiplier Arrays (CEMA) that serves as the ion detector. A set of two CEMA's is often called a Chevron detector. The CEMA's are positioned close to the end of the flight tube, roughly 1 m away from the collector plate and the source. The electronics necessary for a mass measurement is made up of standard NIM modules that are widely employed by nuclear experimentalists. The basis of the measurement is as follows: (1) a radioactive species emits a 3 particle that strikes a plastic scintillator, which provides a start pulse for a Time to Pulse Height Converter (TAC); (2) the recoil ion, after acceleration and drift down the length of the flight tube, strikes the CEMA detector and generates a pulse that stops the TAC; (3) the TAC then gives a voltage output that is proportional to the recoil ion time of flight. In Figure 2—2 is shown a minimum electronics block diagram for a mass TOF measurement. The function of this set-up is to perform a simple delayed coincidence measure- ment. The anode output from the plastic scintillator photomultiplier tube is fed to timing filter amplifier (TFA) for shaping and amplification. The TFA output is then pro- cessed by a Constant Fraction Timing Discriminator to pro- vide a very sharp, fast start pulse for the TAC. The output from the CEMA preamp is sent through an identical set of NIM modules and so generates the stop signal for the TAC. The output of the TAC is sent to an analog-to-digital con- verter (ADC); the digitized output of the ADC is then stored in a computer as the TOF spectrum of interest. The basic operation of SIEGFRIED has remained virtually unchanged since the original construction was completed in 1976 by M. Edmiston. Nevertheless, we have made a few modifications related to the system that are necessary to document for future use of SIEGFRIED. The two most im- portant changes are: (1) design and inclusion of a safety control unit for the power to the CEMA's and (2) a new set of CEMA detectors. The CEMA's are a rather delicate set of detectors, and lO SCIn‘IiIlO‘IOl’ I CEMA PHOTOMULTIPLIE ANOOE OUTPUT PreO mp I I TIMING TIMING FILTER FILTER AMPLIFIER AMPLIFIER cms'mm CONSTANT FRACTION FRACTION TIMING TIMING DISCRIMINATOR - DISCRIMINATOR TIMING TO PULSE HEIGHT 310”. Stop CONVERTER ADC Figure 2-2. Electronics for TOF mass measurement. 11 the manufacturer (Bendix Corporation) warns that they should not be powered in any vacuum worse than 10-5 torr at the risk of their destruction. In the original operation .mode of SIEGFRIED the only safeguard against the destruc- tion of the CEMA detectors during an experiment was through interlocks between two vacuum valves, VA and V6 (see Figure 2-3), and the power supply for the detectors. Actually, these interlocks were not meant to serve as a safeguard system against problems arising during the length of an experiment. Originally the interlock system was intended to ensure only that the CEMA detectors were not powered before the vacuum in SIEGFRIED was below a level safe for operation of the detectors. Well, true to form this interlock system did not function as the failsafe it was never intended to be, as the old CEMA detectors met with a rather quick and complete demise during the course of an early experiment. This sudden, distasteful development is what led to the second of the aforementioned modifications related to SIEGFRIED -- acquisition of a new set of CEMA detectors. The new channel plates were purchased from the Bendix Corporation, product number 302-B-005-MA. In many respects the new CEMA's are much the same as the old ones, both sets functioning according to the same ingenious theory of opera- tion which is well described in the Technical Application Note provided by the manufacturer. The new CEMA's are thinner and draw less current than the old set, but a most l2 SKIMMER CHAMBER X4 Air Inlet TOF CHAMBER TOF LINE [3) IOCm <51 DP V4—¢-)( X ' A V5 25C In )L———VI 0p 5 ll: |884 VANE PUMP U) BOOSTER 25 I/‘s D VANE PUMP 95 ”S X = Vacuum Valve BLOWER [_'_'> = Vacuum Gauge - - LN2 Trap f 7 ll: VANE PUMP I Figure 2-3. SIEGFRIED's vacuum system. l3 important difference between the old set and the new that cannot be overemphasized is that the new CEMA detectors require -2000 V bias. The old set required -2700 V for operatiOn. This particular difference is especially im- portant to note carefully since using the wrong high voltage setting can easily lead to serious damage of the new CEMA's -- if not their complete destruction as functional detectors. Another difference between the old and new sets of CEMA detectors is of much less importance than the power require- ments but interesting nonetheless. The bias angle of a CEMA is the angle that the channels make with the normal to the detector face. Greater bias angles alleviate a problem that is known as ion feedback. The new CEMA's have bias angles of 8° as compared to 0° and 5° of the old CEMA chevron set, this means that the new set will have a better signal—to- noise ratio than the old set. As a result of our experience of destroying the CEMA's, it was decided that a fast and reliable means of safeguard- ing the new set of detectors was a rather obvious necessity. A safety-control unit was designed, constructed, and put into operation as a reliable means of quickly shutting down power to the CEMA detectors in the event of a sudden bad vacuum and/or a large current flow through the CEMA's -- both of which can damage or destroy the multiplier array detectors that are vital for the operation of SIEGFRIED and are very expensive. 1A In Figure 2-A is shown the circuit schematic for the Safety control unit for the CEMA detectors. Power is from an ORTEC A59 power supply that has a remote shutdown feature and so can be effectively turned off by the control unit. The 10 uA meter shown monitors the current drawn by the CEMA's; if this current exceeds the setpoint (usually =3 uA), the meter will cut off the power to the CEMA's. In many tests of this system, intentional or not, the failsafe has been highly successful. The vacuum interlock system was not so useful as the current interlock portion. There were ‘ two reasons for this. First, the only region of vacuum that is of value to the safeguard system is around 10-5 torr. The only reliable means to monitor the vacuum in that range is utilizing vacuum gauges that generate ions in such quanti- ties as to create severe noise problems in the CEMA's. Second, the vacuum interlock system only provides a redun- dancy of the overcurrent safeguard that, in view of the noise problem, is of reduced value. l5 ORTEC ‘59 0'5 KV HV OUT TO CEMA 2 Ii OUT TO CEMA To nzLav. ORTEC I 459 I INTIRLOCK I I ‘V I I I 24 vac I rJ I i i I L Figure 2-A. Schematic for CEMA failsafe. CHAPTER III EXPERIMENTAL RESULTS An important aspect of any experimental instrument is its range of utility. In order to make the greatest use of any apparatus it is necessary to investigate and define the types of systems that are most suited to the capabilities and limitations of the method of measurement. This is es— pecially true of the SIEGFRIED, and the following discussion will attempt to deal with these aspects in a manner that will aid in further use of this time-of-flight system. The results of a number of TOF measurements for several different targets will be presented and discussed in a manner that will display the reasoning and logic required to design, run, and interpret a mass-identification experiment. In addition, these measurements will present a survey on the use of SIEGFRIED in a number of different mass regions. The coupling of the HeJRT system to SIEGFRIED presents the experimentalist with a number of advantages that makes such a combination very well suited to the study of short- lived nuclei that fulfill certain criteria. Naturally, species that do not have the appropriate characteristics yield poor or null results. The aforementioned criteria 16 17 are unfortunately not quantitative conditions, but the following discussion should serve as a guideline for deter- mining whether a particular TOF measurement is feasible. The obvious first step in designing an experiment with SIEGFRIED is to choose an appropriate target. Because of the total efficiency of the HeJRT-SIEGFRIED system it is im- portant to select cases with high reaction cross-sections and targets that can withstand maximum beam on target. Foil targets are well suited to such conditions. In a number of runs with chloride and oxide targets it was found that a large quantity of powdery substance had been generated in the target area and had clogged the input end of the HeJRT capillary, which effectively reduced the transport efficiency to zero. The nature of this powder and the mechanism for generating it are both unknown but do consti- tute a real problem with oxide and chloride targets. Having decided on the target most suited to the reaction of interest, it is worthwhile to estimate the products of the chosen reaction and their relative cross-sections. This is a task that is best accomplished by using the computer routine called ALICE (B170). ALICE is Fortran code written by M. Blann and F. Plasil and is available on the MSU 27 computer as a user's code. The results of ALICE computa— tions give the reaction cross-sections as a function of energy for a number of different reaction products. A number of figures in this section are the result of such 18 computations which provide a convenient tool for estimat— ing the most probable products from the reaction of interest. After determining the probable reaction products, one must consider the half-lives of these products. The time re- quired for transport of a product from the target area up to the collection plate of SIEGFRIED has been estimated to be on the order of hundreds of milliseconds, so species with half-lives less than a few hundred milliseconds will decay away before reaching SIEGFRIED. Using the HeJRT system at MSU, Edmiston (Ed76) was able to measure the half—lives of ”7Cr, “6V and 55Ni. He found these to be A60 i 1.5 msec, A20 i 3 msec and 219 i 6 msec, respectively. These measurements can be used to gain a rough estimate on the upper limit of the transport time. Species with half- lives greater than a few minutes are also difficult to measure with SIEGFRIED. In such cases one must optimize counting and collection intervals as is discussed in this chapter. The chemical properties of the reaction products and the associated molecular clusters are an important aspect- of the transport and measurement process. Though less under- stood than many details Of the system, these properties should be considered in designing an experiment. For the sake of this discussion I will divide the relevant chemical properties into two categories: (1) transport efficiency and (2) sticking efficiency. (This division may well be 19 partially artificial but will aid in understanding two chemically related aspects of our system.) Kosanke (K073) studied the efficiency of transporting a number of different reaction products with the HeJRT system. It was found that the transport efficiencies varied from 2A% to 60% and was also a function of the type and concentration of impurity added to the pure He. The situa- tion now is that the actual nature of the recoil-cluster association has not been elucidated to the point where it is possible to predict the transport efficiency for a given species, but it remains an important point. Once a species has been transported from the target area to the counting area, it is sprayed onto a surface and if the sample adheres to the surface, a thin film source results. However, it is by no means certain in all cases that the nuclei of interest will stick well enough to remain on the surface until they decay. This "sticking efficiency" was also looked into by Kosanke (K073) but only with regard to varying capillary-collecting surface angles and distances. No studies thus far have dealt with this efficiency with respect to differing chemical species, since it has been assumed that the "clusters" stick to the surface and the nuclei of interest somehow adhere to the clusters. Whatever the actual chemical nature of the transport and sticking process, a few practical observations have been made. In no instances have any noble gases (i.e., He, Ne, 20 Ar, Xe) been observed even when the appropriate reaction cross-section is fairly high. This seems to be a rather obvious chemical effect, since the noble gases are known to be relatively chemically inert. In a number of experiments per- formed it has been noted that chlorine and fluorine products had very reasonable cross-sections but were not observed. This may suggest a poor transport and/or sticking efficiency for halide ions. Once a set of reasonable reaction products have been predicted, one should calculate the time of flight for the heaviest mass expected. This is necessary in order to choose the appropriate TAC scale to be used in the measurement. The ‘time of flight for a nucleus of mass number A for our system can be calculated from the formula TOF(usec) = 2.536 /A7HV (3-1) vvhere HV, the voltage applied to the collector plate is in lciloVolts. For a more complete discussion of this equation, see Appendix I. Having chosen the appropriate time scale for the TAC, :lrt is very important to calibrate the TAC. A convenient method of calibration is to use the Tennelec Model 1030 UDILme Calibrator, a NIM module that generates start and stop $3Zlgnals suited for most TAC's. A second means of calibration 1;53 to use some standard target with a well-known TOF spectrum. 21 In all of our experiments an Al foil target was used for calibration, since it gives high statistics, well separated peaks, and is simple to construct. This latter method is actually preferable to the former, since it gives a calibra- tion for the system -- not just the TAC. It has been found that a combination of the two calibration procedures is preferred in most cases. (The Time Calibrator allows deter- mination of a precise number Of nsec per channel.) After completion of the data acquisition, one uses the time calibration to extract the measured time of flight; that is, transform channel number to time units. This trans- formation is a simple linear one. Once the TOF's for each peak are known, the mass number can be extracted by the simple relationship A 0.155 (HV)(T0F)2 (3-2) After determining the mass number corresponding to each Lbeak, one must determine the particular isotope that is Eissociated with the known mass number. If a Y-mass coinci- mzlo> + H (e‘Mt + e'3AAt) t = (2j-l)At N((2j-1)At) = g(At) :E: e-k(2xAt) t: = 2jAt N(2jAt) = N((2j-l) t) e‘Mt We count the number of disintegrations in each count int erval , D(2jAt) = N<<2I-1)At)(1-e‘*At> <3—6) The total number of disintegrations counted after M counting int ervals is D: M M AAt DM = z D(2jAt) = z (1-e‘ )N(2 much simpler form: (Gr65) 321 e-kX g -(k-1)x 1-e-JX k=0 k=l 1-e"x 1M = z (1‘e x ) 1_x { z 1 - z e'Jx} 3:1 1-ep l-e j=l 3:1 A“ M M ’Mx z 1 = M 2 e-JX = e-X. 2 e-(J-l)x e-x {l-e } J=1 3:1 j=l 1-e“x -M I = l {M e-x(l'e x)} M -x -x l-e l-e So we have succeeded in collapsing a finite double sum 3A to a simple algegraic form. Taking x = 2At, we have _ § _ ->.At 2 _ . DM - A (l e ) IM - s FM(A,At) - S (l-e'AAt)? e-2AAt(l_e-2AAt) D - X M (l-e'2AAt) (i-e"ZAAtI A much more concise form than the double sum. Turning to the mechanistic side, we see that the total number of disintegrations is linearly proportional to the source term S, which is intuitively reassuring. If we wish to compare calculations with observed results, a model of S my st be hypothesized. This is very difficult, since the source term S, representing the mechanism for adding fresh sample to the collection plate, is an unknown but certainly Cornplicated quantity! Realistically, it should be function 015' the relevant reaction cross-section, transport ef- f1C31ency, transport time, and sticking efficiency. We can use the ALICE calculations to provide reaction cross- Sections. In order to take the transport time into account, we- suppose that the radioactive species with a decay constant A will decay during the total transport time 1:; thus, a factor exp(-AT) reduces the source term from the start of tr‘ansport. Different species might have different transport 35 times, but if we accept a molecular cluster transport mechan- ism, we would expect I to be species independent, since the massive clusters should have roughly the same T. To compare with the observed peak areas, Ai we will assume: - i _ where e is the efficiency of SIEGFRIED itself, which should be independent of species. For source term we use the form Yi is the transport efficiency for species 1, Ci is the reaCtion cross-section, I is the beam current on target, and N is the number of target atoms. We are actually in- ter‘ested in relative quantities, so we will normalize to a spec ific case. For the case of the Al spectrum I used the laI‘gest peak “Mg for the normalization. If we take a1 Ai/A23 and ”i = Y1/123 (A23-A1)T Fmoi) —- e -———3- 36 From the last form we can estimate the relative trans- port efficiency ”i from ai, an experimental quantity, and ”’1’ a calculated parameter. For the spectrum of Figure 3-1 we collected for two hours with A-sec counting intervals, so At = A.0 and M = 900. The results of this calculation are given in Table III-3. All parameters have been normalized to the 23Mg peak, the largest in the spectrum. The last column gives the relative transport efficiency according to our model. The values for the 25A1, 26A1, and 27Si cases app ear reasonable in the absence of any corraborative fig- ures. The value for the 28A1 is excessively high, perhaps reflecting an unusually low value for the reaction cross- section (2.6 mb). This dependence on the calculated cross- SCCtion may also explain the difference in the transport co- efficients for the 25Al and 26Al cases. Since both are aluminum isotopes, and so chemically equivalent, their transport coefficients should be the same. N__3~F + 3He In an attempt to observe some light mass TOF spectra and to obtain a high multiplicity of peaks it was deemed WC>3:“thwhile to use targets made up of binary compounds. We wanted to use compounds whose components both seemed likely '50 result in products that would be suited to measurement by SIEGFRIED. 37 Hm.H emo.o moo.o owm.o moo.o cmo.fi NMH Ha.“ m:.o mom.o cmc.o Hem.o no.0 m:m.o m.e em.~ mm.o mme.o mm.m omm.o mm.m mem.o e.c Hemw :m.o mmo.o msm.o mmm.o mm.o Hem.o m.e Hem“ o.H o.H o.H o.H o.H o.H H.NH mzm~ A: He He gmmce Emmefiev :Hfmmco who AHAVG seesaw 2 .3 seem ones .3 mass .3: .28 38 Our first choice was a NaF pressed powder target. The excitation functions resulting from ALICE calculations for 23Na + 3He and 19F + 3He are shown in Figures 3-3 and 3-A, respectively. In Table III-Aa and III-Ab are given the expected products, their associated half-lives, calculated cross-sections, B decay energies, and associated recoil energies for 23Na + 70-MeV 3He and 19F + 70-MeV 3He, respec- tively. All the cases tabulated possess a number of attrac- tive characteristics that make them well suited for TOF measurements by SIEGFRIED. The majority of half-lives are on the order of tens of seconds. The decay energies are large enough to result in large recoil energies. In the cases of 16N and 2°F, the very large decay energies give rise to huge recoil energies -- A keV and 1.5 keV, respec- tively. Such recoil energies are quite appreciable fractions of the usual 6-kV accelerating voltage employed in our system. As will be shown in following discussions, the broaden- ing effects from such large initial kinetic energies would be very pronounced in the resulting TOF spectrum. This would provide us with a highly graphic demonstration of the dependence of TOF peak widths on the initial recoil energy. Furthermore, the NaF target should provide us with a real multiplicity of TOF peaks with very little inter- ference from recoils with the same A. As a result, we should obtain a light mass spectrum that is easily 39 KDC>- so — "'- "’F °F 25 Al 23Mq 0(mb) ' 2 INC ./ 3‘" 2°... ..-N K3-- 5 _ 22Mg I90 I l l J l I I0 20 30 4O 50 60 70 E3 (MGV) He Figure 3-3. ALICE predictions for 23Na + 3He. A0 IOO - 20F- I50 so — 2'N0 0’ ‘ IGN (mb) V» \ I9 N IO — I6 5 __ F L I I l L l 20 30 40 so 60 70 E3 (MeV) He Figure 3-A. ALICE predictions for 19F + 3He. Table III-Aa. A1 Properties of Reaction Products from 23Na + 70-MeV 3He. t E Recoil 1/2 0 B Energy Isotope (3) (mb) (MeV) (eV) 16N 7.1 7 10.u2 A000. 17F 6A.5 A7 1.7A 152. 190 27 u A.82 795.5 20F 11 A2 7.03 1519.3 21Na 3.9 16 2.52 228.2 22Mg 12.1 A 3.77 AAl.0 23Mg 7.2 37 3.03 280.1 Table III-Ab. Properties of Reaction Products from 19F + 70-MeV 3He. t E Recoil 1/2 a B Energy Isotope (5) (mb) (MeV) (eV) 150 122 68 1.7A 172.0 1°N 7.1 23 10.u2 u000. l7F 6A 22 1.7A 152.0 19Ne l7 19 2.22 203.0 2°F 11 7 7.03 1519.3 A2 calibrated and contains some interesting peaks. For this eXperiment we calibrated the system by first obtaining an 27Al + 70-MeV 3He spectrum from two hours of counting. After cleaning the collection plate, we col- lected data on the NaF spectrum for l-l/2 hours with 10A of beam current on target. Shown in Figure 3-5 is the result of the l-l/2 h counting from the NaF target. Ob- viously we didn't need to worry about overlapping or inter- fering TOF peaks! The single mass peak occurring in the spectrum occurs exactly where the 23Mg peak occurs in the Al calibration spectrum. The results of this run are slightly disconcerting, as we mentioned that the pertinent characteristics of the reaction products seem to be almost tailored to SIEGFRIED. It goes almost without saying that we must have overlooked something important! In a few cases it might be that the real cross-section is much smaller than the ALICE predic- tions but it is very improbable that all but one cross- section were overestimated. The characteristic that we did not take into consideration was the transport prOperties 0f the reaction products. Except for the Mg and Na isotopes, all the reaction products are potentially volatile species that would have poor transport efficiencies. In studies using a system very similar to SIEGFRIED, H. Wolnik (W076) found that elements like the noble gases, Br, and I are hardly attached to the molecular clusters but still pass A3 8402-— 23Mg .J UJ Z 2 <1: I L) $\ 09 I... Z I) c: C) 0-J —~Aee - 5500 000 CHANNEL NUMBER Figure 3-5. TOF spectrum for 70-MeV 3He 0n NaF target. AA through the capillary with reasonable efficiency. Never- theless, at the collector foil it was found that the total efficiency is very low. Evidently the same phenomenon is occurring for the F, N, O, and Ne isotopes that we expected to observe. KCl Target The next binary compound used as target material for a possible TOF measurement was potassium chloride (KCl). For these experiments pressed powder targets on a thin Al backing were used. Figure 3-6 gives the calculated excita- tion functions for the 39K + 3He reaction for the energy range 20 - 75 MeV. The excitation functions for 35C1 + 3He are given in Figure 3-7. As can be seen from these plots we can expect a fair number of suitable reaction products especially at the higher energies for the incident 3He. Another consideration favoring higher energies is the transport efficiencies of the products. The excitation functions for the 39K + 3He reaction show that at the lower energies (E3He < 50 MeV) two chloride isotopes account for the majority of the reaction cross-section. If our ex- perience with the NaF target taught us anything, poor transport efficiency should be expected with potentially volatile species such as chlorides. In view of these facts, we ran the KCl experiments with a beam energy of 70 MeV. Since the heaviest possible product would have a A5 H30 34%“ 50L 6 (mb) \ 29p IC)- 3| / s 28AI 5— ' 33CI 39to I I I I I I 20 3O 4O 50 60 7O E3He (MGV) Figure 3-6. ALICE predictions for 39K + 3He. A6 IOO - 50 - “A! 275‘ 0' (mb) V I0 - ”AI 9,: 5 I... l l l l l l E3 . (MGV) H Figure 3-7. ALICE predictions for 35C1 + 3He. time of flight of about 6.8usec we used an 8 usec TAC range. This is also a convenient range for the Al Cali- bration Spectrum. We collected data on the KCl target for approximately ten hours. The count rate with luA of 70 MeV 3He on target was decidedly lower than the A1 calibration. Nonetheless after an hour of counting, a multiplicity of TOF peaks was observed. Using the Al calibration we determined the lepe, K and the intercept, t for the relation: 0 TOF(CN) + K . CN + tO (3.3) This allows one to calculate the time-of-flight as a function of channel number (CN). From the TOF(CN) we are able to determine the corresponding mass number from Equation 3.2, with RV = 6 kV. The accumulated spectrum is shown in Figure 3-8. Using the K and t0 from the Al spectrum and Equation 3.3 we determine the correct mass numbers. The fact that masses 26 and 27 occur at precisely the same position on both the calibration and KCl spectrum streng- thens these assignments. Finally as a double check, a linear least squares fit to the assigned TOF's, as a func- tion of Channel number, was run. The results of this fit are given in Table III-5. Thexz for the fit was 2.3 x 10‘5. Assignment of the correct Z corresponding to each peak in the KCl spectrum is more difficult than in the previous .pomgmp Hos co omm >ozlo> pom sapwoodm mos .wIm ogswfim mumZDZ JMZZozlo> + He mmeDZ JMZZozlos mo mposoogogom ESLCOOQm mos .wHIm mgsmfim 833 do» I, erI. o: 00. on 00 os 00 On 0e onIIo IIJ . 1 _ _ q _ A L A .3} iottxsitigggéé it .5133... I w... w .. ... ... ... ... v e «a w a a v m cw v w — ’_—-—‘ _ WENNVHD/SINOOO 7A from two hours of counting of the products from the 1HSm + 3He. At the long TOF end of the spectrum are two large, very broad peaks. The first of these two peaks occurs at the position of A = 1A1 but, as expected it is broad enough to encompass masses 138-1AA. Obviously it is impossible to resolve meaningfully the individual mass com- ponents of this peak. The only thing one can surmise from this feature of the spectrum is that we have measured the TOF's of a number of species with masses between A = 138 and 1AA. The second large peak shown in the spectrum seems to allow some resolving of the individual components. The masses most prominent in the peak correspond to 155, 158 and 159. Yet how can such masses result from a compound nucleus of maximum mass of 1A7? A number of experimenters (Ne78), (N879), (W077) have observed very similar phenomena in their TOF spectra and attribute the heavy mass peaks to molecular species that are somehow volatilized from the col- lector surface. The fact that the second large mass peak corresponds to the earlier masses + 16 amu, which is the mass of 160, suggests that the components of this peak are the same as the lighter masses except there is an attached oxygen (A = 16). In the remainder of this spectrum can be found a large number of small TOF peaks that are not explained as simply as the peak centered around A = 158. They may correspond to radicals (CnHmOk’ etc.) that have been ionized by collision with the primary recoils. CHAPTER IV RECOIL ENERGY FITTING A close inspection of the mass time-of-flight peaks acquired with SIEGFRIED reveals a number of phenomena that are not exactly what one would expect for simple measurement of an ion flight time. Referring to Figures A-l, 3-1, and 3-5, it is easy to notice that the peaks are all rather wide -- much wider than any electronic timing contributions might account for, as will be shown. Per- haps more striking to the eye is the very peculiar shapes exhibited in the spectra; all have long nearly exponential tails on the low time side, implying some mechanism that suppresses higher energy events and favors longer TOF events. It seems reasonable to View the source of the recoils as a thin film deposited on the collecting plate, since massive clusters provide the means of transport of the product nuclei from the target area. A simple picture of recoils passing through a surface layer of cluster deposit seems to favor a higher probability of escape for high energy recoils than for low energy recoils. However, higher 75 76 00¢: _M 1 comm 3885 EOE. Ohm.» conenMII _ .Ezppoogm Hos 90% cups: some mo comHLmQEoo .HI: Gasman nnO. _ 0G0: m I' ‘II +1 '13 NNVHO/ SlNflOO OON 77 energy recoils should transform into shorter TOF events, and this is not what one observes in the accumulated TOF spectra; in fact, the converse seems to be true. That is, the low energy events seem to be the more probable events to be detected. The entire puzzle is especially intriguing, since in the early days of development of SIEGFRIED and its forerunners it was not known for sure whether recoils from B decays would be energetic enough to escape the surface on which they were deposited (Ed76), (Ma7A). It appears, however, that not only are the recoils energetic enough to free themselves, but also the recoil energy is a significant contribution to the peak broadening. Just how important are broadening effects such as electronic timing contributions in our measurements is a reasonable concern that we will deal with briefly here. A very rough argument for nanosecond range precision can be made by consideration of the performance specifications for the various NIM modules used for the TOF measurements. This is not quite so convincing as an internal measure or upper bound on intrinsic broadening. Happily, nature sees fit to provide us with just such a means to estimate an order of magnitude figure of merit. Referring to Figure A-l, it is found that a narrow peak occurs at exactly the position predicted for an H+ ion. Obviously, a proton is the result of a neutron undergoing a B decay and it seems a bit far-fetched to hypothesize neutrons being transported 78 . and deposited on the collection plate; however, the cluster molecules are somehow formed from saturated hydrocarbons (e.g., CGHG), so there is certainly an abundant source of hydrogen atoms at the collection site. The exact mechanism for how the H+ ions result from decay-related events is certainly not clear. However, the recoil energies involved in a "typical" 8 decay of medium-light nuclei are on the order of a few hundred electron volts. When one compares such recoil energies with typical carbon-hydrogen bond energies, which are roughtly 5 electron volts, the dis- sociation and ionization of a hydrogen atom from a recoil- bearing cluster is a plausible occurrence. Furthermore, in a number of spectra taken by a group at Orsay on a TOF system much like SIEGFRIED (Be78) a large, prominent spike is also assigned to the H+ ion. Accepting this assignment we can use the full width at half maximum (FWHM) to obtain an estimate of the intrinsic broadening effects due to the electronics. As shown, the H+ ion has a FWHM of 8 nsec which represents a real limit on the peak broadening. In comparison, the mass groups at longer times of flight have FWHM's of about A-5 times that of the H+ peak. As we intend to show, the peak widths result from the spread in recoil energy that the daughter nucleus has after the B decay. Another aspect of the TOF Spectra that is especially intriguing is the overall shape of the peaks. This is 79 most easily seen in the large, isolated peaks correspond- ing to 39Ca in Figure A-1 and 23Mg in Figure 3-1 and 3-5. All exhibit a long, exponential-type tailing on the short TOF side with a very sharp drop-off on the higher channel side. This distinctive feature seems to just elicit further investigation. As a first step in the analysis, we decided to attempt to fit the TOF spectra with the program SAMPO, since it is a fairly flexible and easily run data analysis routine. SAMPO is a Fortran routine written and developed by J. T. Routti and S. G. Prussin (R069) to perform automatic data- analysis of y-ray spectra. The routine is familiar to spectroscopists involved in Ge(Li) work but is probably rather foreign to mass spectroscopists in general. In order to elucidate our fitting results, a brief description of the analysis routine and the associated terminology will be given here. In SAMPO the peak shape is approximated by a function that is basically a Gaussian with possible high— and/or low-channel tailing that "goes" as exponentials.- The amount of tailing is determined by the distance in channel number from the centroid to the points at which the Gaus- sian is joined to the appropriate exponential. At these junctions the function and its first derivatives are con- tinuous; as a result, the complete peak shape, aside from the normalization, is specified by three parameters: 80 (1) CW, the width of the Gaussian, (2) CL, the slope param- eter of the low-channel exponential tail and (3) CH, the parameter of the high-channel tail. The signifiCance of the magnitudes of these shape parameters is very straight- forward: (1) a large CW means a wide Gaussign contribu- tion, (2) a large CL yields a fast rise on the low-channel side, and (3) a large CH gives a rapid fall-off of the peak on the high-channel side. Shown in Table IV-l are the results of a SAMPO fit for the KCl spectrum of Figure A-l. As indicated by the X2 for each, the fitting routine has proved to be well suited to deal with the rather asymmetric shapes. The CW's for all the identified species except the 28Al case are very large. (CW as was discussed earlier, is the width parameter for the Gaussian contribution to the peak shape.) The magnitudes of the CW's shows a significant broadening effect in the TOF peaks. The low-channel tail parameter (CL) for the peaks are generally small and reflect a long, slow drop-off on the shorter TOF side. Especially striking is the size of CH for masses 26 through 39, excluding the 28A1. All the CH's for this set are greater than 5; in fact, all except the 27Si case are greater than 10! This indicates that the exponential fall-off on the high- channel side of the peak is very, very rapid. In fact, the large size of the CH's implies some type of cutoff -- a limit on the maximum time-of-flight allowed. If we 81 30 x mmm.m n Ezzm om. IIIIIIIIII IIII mz.m N.m m.m mo.H mmmo m: Hz. IIIII IIIII IIII NH.m mH.H ms.m mH.H mmmw H: m.mm mam. 0.30: m.m m.w mm.mw Hw.m NN.H meow momN as.: mam. m.moz Nm.: m.m :w.wa mm.m mam. mmmm man :.s o.H m.:m: :z.: 3.5 w.:H o.m Ho.H mmmm mom :m.m mam. 0.9mm mm.m m.s H.3m w.: wo.H momm mow m.m 0.0 :.mam mm.m mm.m m:. :H. mm. zosm H.m :s.: sm.m :O.H mo.H mawm Hmew w.w oo.H w.owm Hm.m 0H.m m.mH Hw.H mo.H mHmm amen AO\:U HELom A>ov A>mzv 30 30 do 0 mpfiopu < s m mm Icoo .mpuooam Hox mo wcfiuufim OmzH canoe 82 equate the time of flight with the total energy of the recoil ion, as shown in Appendix I, the TOF is inversely proportional to the kinetic energy of the ion. In light of this, it is easily seen that the maximum TOF corresponds to a minimum allowed kinetic energy, that acquired by ac- celeration across the gap region. To sum up the results of the fitting briefly, the peak shapes predicted by SAMPO will display a long tailing for the shorter TOF's, a wide Gaussian portion and a very fast fall-off on the high time side. A glance at the KCl spectrum readily exhibits this predicted behavior. Ob— viously there must be a reason for these asymmetric peak shapes, although the causes may not be immediately pin- pointed. Nonetheless, the simplest hypothesis concerning the TOF peaks seems to be that the initial recoil energy of the daughter nucleus is somehow related to the observed spectral shapes. If this is so, then we can develop a quantitative relationship between the observed peak shape parameters and the recoil energy of the daughter nuclei. Obviously, the broadening of the mass peaks is not an electronic effect. Is it reasonable to attribute the effect to the initial recoil energy of the daughter nuclei? As is shown in Appendix I, using a gross model of the SIEG- FRIED system, we can estimate the time-of-flight of a singly charged ion of mass number A and initial recoil energy R by the expression: 83 TOF = ml/2 {[2(E:;?]l/2 + 11%18 [/EIE _ Rl/21} (A-l) Rigorously R = l/2MV3 where V2 is the initial velocity in the z direction, that is, in the direction of the flight tube axis. In Table IV-2areethe results of times of flight calculated for the mass set observed in the KCl spectrum. R is the maximum initial recoil energy, T(R) the corresponding TOF, T(O) the TOF for zero initial recoil, and AT the difference between T(O) and T(R). The TOF dif- ferences are all in the neighborhood of a few hundred nano- .seconds. Considering our timing resolution to be of the order of nanoseconds, it is imminently plausible that the recoil energy effect is a significant contribution to the broadening. We have argued that the peak shape, especially the broadening is because of the initial recoil energy of the measured mass species; if that is so, then it seems reason- able to attempt to develop a quantitative relationship between some measured peak shape parameters and the recoil energy of the ion. First, we need to have some rough form -of the parameterization. If we consider the ions drifting along the flight tube, with negligible time of acceleration (correct to ~3%, see Appendix I) then we can write: 8A Table IV-2. Calculated TOF's for KCl Spectrum (in nsec). A R TOF(R) TOF(O) AT AT/T mean 26 381 5055 52AA 189 .036 27 362 5160 53AA 18A .03A 28 213 5323 5AA2 119 .022 29 359 53A9 5539 190 .031I 3O A3A 5AO6 5633 227 .OAO 31 A09 5507 5726 219 .038 39 A9A 613A 6A23 289 .OA5 85 E = l/2mv2 = l/2m —— (A-2) E is the ion kinetic energy, v its velocity, m the mass, A, the flight length and t the time of flight. Now suppose an ion starts with zero recoil energy, then B = qV 5 E0 = 1/2mvg, where V is the electrostatic potential on the collection plate and q is the charge on the recoil ion. If the ion starts out with the maximum recoil energy, with zero transverse momentum (pg/2m = R, R the maximum recoil energy), then E = EO + R = l/2mv2 2 l/2m22/ti. So we have _ 2 2 EO - 1/2m2 /t0 (“-3) = 2 2 El l/2m2 /tl (A-A) 81-80 = R = 1/2m22(;1§ - 52-) (II-5) l o If we make the assumption that the measured TOF is linear in channel number and the spread is proportional to the FWHM (or equivalently to CW from SAMPO fitting), we can write: 86 t0 = K(CP + CW) (A-6) t1 = K(CP - CW) (A—7) t: = K2(CP2 + 2CP cw + CW2) (u-8) = K2CP2(1 + 2CW/CP + CW2/CP2) (u-g) Similarly, ti = K2CP2(1 - 2CW/CP + CW2/CP2) (A-lO) Since typical values of CW/CP are on the order of 10-3, we can drop the (CW/GP)2 terms. Using the binomial theorem and retaining terms to first order, we have 1 = l (l + 2CW/CP) (A-ll) :5 K2CP2 l .33 z _§l—§ (1 — 2CW/CP) (A-l2) t K CP 0 2 2 m 1 1 ml 1 ACW R = -— E - ——l = ( ) (A-13) 2 g?" t2 2 K201,2 CP 0 R = 0A EHL (A-lA) 87 This last expression is the form that will be used to re- late the recoil energy to the TOF peaks parameters CW and CP. We now have a parametric form to test. Next we need some experimental cases to apply it to. Since we are really working in a new area, it is obvious that there are certain criteria for the spectra we will attempt to fit. We require fairly high statistics in each peak and that the decays be fairly simple; also we prefer as high a multiplicity of peaks in the spectrum as is obtainable. The reasons for these conditions are fairly obvious; we are in the position of hypothesizing a relationship that we believe can be expressed by a simple functional relation- ship. Now, in order to prove or, at least further, this view, we must start with the most reasonable set of tests available. The KCl and Al spectra come closest to ful- filling our rather loose conditions. The Al target has been the standard for our runs and is well understood, while the KCl spectrum has a fair multiplicity of single species, peaks, good statistics and the "characteristic" shapes of TOF peaks occurring in a number of instances. The actual form of the equation used for the fitting was y1 = ax + b (A-15) '88 xi = CW/(CP)3 - A (CP in kilochannels) (A-16) y1 = Recoil energy of species i A linear least squares fit was performed on three known cases, and the resulting parameters were used to predict the recoil energy of the members of the set not included in the least squares fitting. The predicted values were then compared with the known values. The maximum recoil energy of a daughter nucleus with mass number A was calculated using the following equation: where E A R R = 5&1 (EO + 1.022)) (A-17) is the maximum kinetic energy of the 8 particle in MeV is the mass of the recoil in amu is the maximum energy of the recoil in eV (For the derivation see Appendix II.) The first example of the fitting is for the KCl spectrum. Our procedure was to choose two sets of three peaks as the "known" energies and predict energies for the remaining peaks in the spectra. The first set chosen for the 89 calibration corresponds to 26Si, 2781 and 3OS. In Table lV-3 are the results of a linear least squares fit to the recoil and decay energies for this set. The predicted decay energies are calculated from the fitted recoil energies according to the formula .. ‘/ £5- - EO - 0.261 + 537 0.511 (A 18) (with R in eV and E0 in MeV) The fit for this set is fairly good, X2 for the standards is 1.1, and the largest deviation is A%. The results of applying the equation R1 = axi + b to the remaining peaks in the spectrum are given in Table IV-3: the errors in the recoil energy are all in the neighborhood of 15%, the fit to the decay energies closer to 10%. The worst fit is for the 29P case, which is not well separated from the tail of the 3°S. The mass-28 peak actually gives a somewhat better fit than was expected, considering the decay scheme of 28A1, shown in Figure A-2. The B decay of 28Al is to the 2+ excited state of 28Si which decays via a y ray to the 0+ state. Also the decay of the 28Al is by negatron (B-) decay whereas the rest of the decays we observe are by positron decays. Since the collector plate is lO-mil Table IV-3. 9O Fit to KCl Spectra Recoil Energies (no F-GT Correction). Calibrated R yfit EO Efit A (eV) (eV) % Error (MeV) (MeV) % Error 26 380. A97. A.5 3.81 3.90 2.A 27 362. 355. 1.8 3.79 3.7A 1.3 30 A3A. A2A. 2.A A.AA A.38 1.3 = 1. Predicted R yflt E0 Efit A (eV) (eV) % Error (MeV) (MeV) % Error 28 273. 302. 11. 2.86 3.08 7.7 29 359. A28. 19. 3.92 A.32 10. 31 A10. A73. 15. A.37 A.7A 8.5 39 A9A. A23. 1A. 5.5 5.05 8.2 91 2.3! min ZBAI Op = 4.635 7 W ZBSI Figure A-2. Decay of 28Al. 92 stainless steel, it will stop 8's of an energy of less than about 1 MeV, which means that the lower energy events are discriminated against; however, the plastic (NE-102) has roughly a 30% efficiency for detection of y rays of energy 1.8 MeV, so that if the y ray is emitted in the "right" direction, it has some probability for serving as a valid start signal and can result in a detectable TOF event. The estimation of the y detection efficiency was accomplished by using a nomogram developed by Roulstan and Naqui (R057). In the case of the positron-emitters, even if the low-energy B's are stopped by the collector plate, the accompanying annihilation radiation is suitable for serving as a start event. Using the same procedure mentioned earlier, the gamma efficiency for 511-keV annihila- tion radiation is about 50%. The second case we attempted to fit with our recoil energy parameterization was the TOF spectrum resulting from 70-MeV 3He on a pure aluminum target. Unfortunately there are only four peaks with reasonable statistics and shape. The 28A1 peak is very broad and shaped distinctly dif- ferently from the common TOF peaks, and the 29P peak has very poor statistics. The peaks chosen for the "known" energies are the three corresponding to masses 23, 25, and 27, respectively; the results of these fits are given in Table IV-A. The x2 for this fit was 7.A, and,-as shown, the fit errors are in the neighborhood of a few percent. 93 Table IV-A. Fit for Al Spectrum -- No F-GT Correction. R fit EO Efit A (eV) (eV) % Error (MeV) (MeV) % Error 23 287. 286. 0.3 3.0 3.03 1.0 25 299. 331. 11.0 3.26 3.A5 6.0 27 362. 325. 10.0 3.79 3.56 6.0 9A Fermi and Gamow-Teller Distributions One assumption that is implicit in the parametric treat- ment of the recoil energy is that all allowable energies are equally probable and all the various decays have the same initial energy distributions. This is not actually true, for, as is known from B-decay theory, the distribu- tion function for the recoil energy is dependent upon the "type" of decay the parent nucleus undergoes. I will not endeavor to present a full treatment of B-decay theory here, there being a large number of well-known and more appropriate texts for such a task. However, I will give a somewhat pedestrian and brief discussion of general results and terminology from what is called allowed B-decay theory. In the B-decay of a nucleus, an electron and an anti- neutrino are emitted, and each of these particles has a spin l/2. If the total spin of the decaying nucleus is unchanged, then, in order to conserve total angular momentum the electron and antineutrino must have antiparallel spins. This type of decay is often called a "Fermi" decay. If the total spin of the parent changes by l (in units of h), then the electron and antineutrino must carry off one unit of angular momentum, ;;g;, they have parallel spins. This type of decay is known as "Gamow-Teller" decay. Common terminology in B-decay theory attributes Fermi decay solely 95 to a "vector" interaction and Gamow Teller to an "axial- vector interaction." Because these two interactions are "effectively" different, the probability that a recoiling nucleus has a given energy is dependent upon the type of I interaction that gives rise to the decay. In Figure A-3 are shown the distribution functions for the recoil energies corresponding to the two types of decay considered here. The distribution functions for both cases are skewed towards higher energies, but the one correspond- ing to the so-called Fermi decay is obviously more sharply peaked, favoring higher-energy recoils. When one takes into consideration these two distributions, it becomes reason- able to ask whether the effects of the initial energy distribution are in some way contained in the observed peak shapes. If that is the case, then, even if there is only a gross relationship, it is very useful and highly intriguing. Assuming the initial energy distribution effects are some- how preserved, we need to quantify the influence of the distribution upon the TOF peak shapes. The majority of decays observed are actually mixed, that is, they result from a combination of the Fermi and Gamow-Teller decay modes. In an attempt to take this account, we have included as a multiplicative fixed parameter the quantity, D 5 (fF + «fGT) where fF and fGT are the fractions of transi- tion probability due to the Fermi and Gamow-Teller modes, respectively. The factor a is a strength parameter that 96 PROBABILITY Figure A-3. Fermi-Gamow-Teller recoil energy distribu- tions. 97 is a measure of the relative contribution of the Gamow- Teller mode to the FWHM of a composite energy distribution. (It is estimated by taking the ratio of FWHM's for the two distributions shown in Figure A-3. A rough estimate in most cases gives a = 0.5.) The quantities fF and fGT can be found if the ft value of the decay is known. If we define IMFI and IMGTI as the matrix elements for the vector and axial vector interaction, and Cv and CA as dimensionless coupling constants for the vector and axial vector interactions, then _ 61A3 _ ft 5 IM I2 + (CVI2 IN I2 (u 19) F '6; GT (with CA CV = 1.25) Now, fF and fGT are defined as MF 2 (u-20) c 2 2 2 IMF! + (51) IM A cr' So in order to obtain fF and fGT’ we need the ft value for the decay and a value for either IMFI2 or IMGTI2. In 98 most cases of interest to us, the decays will be simple mirror transitions or at most decays that are classified as allowed, in which instances there are simple rules to Calculate IMF . The ft values can be calculated, (see (Wu66)) or obtained from the literature. The combination ft and IMF] allow one to calculate fF and fGT’ which repre- sent the fraction of decay strength due to the Fermi and Gamow-Teller decay modes, respectively. In Table IV-5 are given the ft values and calculated values of IMF], fF, fGT, and D for all the known decays observed as described earlier in this chapter is followed, CW 3 CP3 where DA is totally determined by the particular decay. except now the independent variable is DA - A - The results of this new parameterization applied to the KCl spectrum are given in Table IV-6. X2 for the linear least squares fit to the "known" cases was 1.8, not sig- nificantly different from the simpler initial parameteriza- tion, perhaps a little worse. The same seems to hold true for the fit to the "unknowns." However, referring to Table IV-7 where the results of the new parametric fit for the A1 spectrum are given, a very obvious improvement in the fit to the "known" peaks is evidenced. In fact, x2 for this fit is 0.0A, a very large and significant reduc- tion from the x2 of 7.A for the simpler parametric form. But the fit to the 27Si "unknown" is very poor in this instance. 99 Table IV-5. Fermi-Gamow-Teller Parameters for Peaks in KCl Spectrum. A ft IMFI2 fF fGT D x' 26 3162 2 1.0 0.0 1.0 0.97 27 3981 l 0.65 0.35 0.85 0.61 28 79A3 0 0.0 1.0 0.58 0.26 29 5012 1 0.82 0.18 0.92 1.05 30 3162 2 1.0 0.0 1.0 1.12 31 5012 l 0.82 0.18 0.92 1.28 39 3981 l 0.65 0.35 0.85 0.9A 100 Table IV—6. Recoil Fit for KCl Spectrum (With F-GT Cor- rection). Calibrated R Rfit EO E:f‘it A (eV) (eV) % Error) (MeV) (MeV) % Error 26 381. A01. 5. 3.81 3.92 3.0 27 362. 356. 1.6 3.79 3.75 1.0 30 A3A. A20. 3.A A.AA A.36 1.8 Predicted 28 273. 308. 13. 2.87 3.53 23. 29 360. A11. 1A.O 3.92 A.23 8.0 31 A09. AAO. 8.0 A.37 A.55 A.0 39 A9A. 397. 20.0 5.5 A.88 11.0 Table IV-7. Fit to Al Spectrum Recoil Energies (With F-GT Correction). R Rfit EO Efit A (eV) (eV) % Error (MeV) (MeV) % Error 23 287. 28A. ' .7 3.0 3.01 .A8 25 299. 297. .8 3.26 3.2A .61 27 362. 362. .1 3.79 3.788 .05 CHAPTER V MODELING THE TOF SYSTEM I. Introduction The results of the empirical fitting to the observed recoil energies in the previous section were encouraging enough to elicit further investigation. As a means of theoretically justifying our assumption about the recoil energy contribution to the peak broadening, we decided to mathematically model our TOF system and the resulting particle trajectories. The first step in modeling our TOF system was to obtain a reasonable set of electric fields. The system is made up of elements of finite lengths and edges whose effects should not be totally ignored. As with most practical problems, analytic solutions exist only for idealized cases. For real geometries we often have to settle for approximate solutions. To this end numerical methods are especially well—suited. To obtain the electro- static potentials for our system we have to solve Laplace's equation subject to the appropriate boundary conditions. The iterative technique of relaxation is a standard method 101 102 for obtaining approximate solutions to Laplace's equation. Once we have a reasonable set of fields, we can calculate particle trajectories. This means that we have to inte- grate equations of motion for charged particles. To ac- complish this we used a combination of Runge-Kutta and Predictor-Corrector algorithms. Since these numerical methods are an integral part of our modeling study, we will present a discussion of them in some depth. This is especially true for the relaxation methods we employed because of what we feel is their growing importance in this age of digital computers. II. Relaxation Techniques A. Preliminary The solution of Laplace's equation for a system of conductors is one of the most common problems in electro- statics. Students are often taught the usual analytic methods of solution: images, Green's functions, and ex- pansions in orthogonal functions. The powerful yet in- elegant numerical techniques are rarely presented as alter- nate approaches. For the more complicated numerical tech- niques, this exclusion is understandable. However, the relaxation technique, which we shall discuss here, is so simple and well—suited to boundary-value problems, that once learned, it provides a valuable tool for constructing 103 solutions for many common and specific problems. When used in conjunction with digital computers available today, relaxation techniques are able to handle problems that are beyond the scope of practical analytic solution. In this section we present a straightforward discussion and derivation of relaxation and the related over-relaxation techniques. We then apply these techniques, both to simple problems such as would arise in a classroom and also to slightly more complicated problems such as have arisen from our own experimental work. Occasionally, in discussions of Laplace's equation, students are told that it implies that the potential at a given point is the average of the potential on a surface enclosing the point. Then, to demonstrate this, the poten- tial at the center of a charged, hollow, spherical conductor is shown to be the same as the average on the surface. Such an example serves the purpose well enough for the specific case of spherical symmetry, but the implications of Laplace's equation need not be so restricted. We feel that it is reasonable to demonstrate to students that the general equation var = 0, (5-1) can be interpreted as a differential statement of the fact that the solution at a specific point is just the average 10A of the solutions over a surface of any shape surrounding thegmdxu;of interest. If such a demonstration can be easily given, it will greatly enhance and supplement specific examples. (As will be shown, application of relaxation techniques to boundary-value problems in electrostatics is intimately and obviously dependent upon the differential validity of the averaged nature of solutions to Laplace's equation.) The presentation of relaxation methods in electricity and magnetism texts is nearly nonexistent. A notable ex- ception is the introductory text, Volume II of the Berkeley Physics Course. Yet even here the student must delve into the recesses of an appendix on advanced problems. There a recipe, applicable only to a two dimensional Cartesian problem, is given. The basis for the prescription and its extension to other orthogonal systems is not discussed. With a few elementary preliminaries we will attempt to correct such exclusions and present relaxation techniques in a manner that students of electrostatics can appreciate and employ, having access to only modest sized computer. B. Mathematical Preliminaries To those of us who worked our way through differential calculus with mild chagrin, the definition of a derivative is perhaps familiar yet seldom used. However, for numerical solutions the finite divided-differences that define a 105 derivative as a limit are essential to replacing a dif- ferential equation by an analog algebraic expression. Recalling the definition of the derivative at a point x0, we have f(X +Ax) - f(X ) df(x) = 11m 0 0 (5-2) dx Ax = Ax+o x xO This expression gives an approximation to the deriva- tive when Ax is small; that is, df x)l f(xO+Ax) - f(xo) dX Ax ° (5-3) Of course, this is an approximation and the associated error is on the order of Ax. As shown in the Appendix IV, a better approximation is given by df;x) f(xO+Ax) - f(xO—Ax). (5 u) 2Ax ’ - Q. H II and for the second derivative, 106 d2f(x) f(xO+Ax) - 2f(xo) + f(xO-Ax) ( ) = 5'5 x=x (AX)2 These last two expressions have associated errors on the order of (Ax)2, which shows the advantages of choosing Ax small. The extension of these expressions to partial deriva- tives is straightforward and given in Appendix IV. Now let us consider Laplace's equation for a two- dimensional Cartesian system, at the point (xo,yo): 32¢(x ) 82¢(X.y) _ ____E£Z_ + -——_7T-_' - 0. (5-6) 3x 3y - x=xo X‘XO y=yo y=yo The analogous expression in terms of finite differences is ¢(xO+Ax,yO) - 29(xo,yo) + 9(xO-Axo.yo) (Ay)2 + ¢(xo.yo+Ay) - 29(x03y0) + ¢(xo.yO-Ay) = 0. (5_7) (A102 ' Rearranging this last expression, we have 107 9(xo,yo) = J; 2 {©(xO+Ax,yO) + 9(x0-Ax,yo) 2 = % {(xo+h,yo)+¢(XO-h,yo)+¢(xo,yo+h)+¢(xo,yO-h)} <5—9) this makes it clear that the potential at x ) is equal o’yo to the average of the potentials at points around (xo,yo). Furthermore, it is simple to obtain equivalent expres- sions when we need to deal with different coordinate systems. Consider the Laplace equation for an axisymmetric system in cylindrical coordinates, O = ¢(o,z): 2 v20 = 3 No.2) + 30 I\.) (0.2) + a 3 (0,2) = 0. (5-10) 0 dz 0' 3 DIH ”I At the point (90,20) the equivalent finite-difference expression is 108 (b - _ .(oO+A0,zo)+¢(oO Ao.zo) 2¢(OO,ZO) (Ap)2 9(00+Ao,zo)=¢(oo=Ao,zo) 2A0 +._1_ pO ¢(p ,z +Az)+¢(p ,z -Az)-2¢(o ,z ) + O 0 0 3 ° ° = 0 . (5-11) (AZ) Rearrangement yields 9(00.20) = J; 2 {9(00+Ao,zo)(1+Ao/200)+¢(CO-Ao.zo)(l-Ao/200) 2(1+(E%) ) + <%%)2 (¢(oO,zO+Az)+9(oo,zO-AZ))}. (5—12) In this last equation it appears that we have a singular- ity for 00 = 0, but for axial symmetry, 0) '6- II 0 (5-13) 0) 0 Therefore, by L'H8pita1's rule, 109 “"122 ..ail o +0 o 80 302 o o po 00 At 0 = 0, LaPlace's equation is 32¢ 32¢ _ 27+ 2-0, 30 Bz ¢(O+Ap,zo) = ¢(0-Ap,zo) (5 15) (5-16) (5-17) Rearrangement of the finite-difference equation equivalent to Equation 5-15 gives, for p0 = 0. 1 ¢(O,ZO) = {A¢(Ap,zo)+(%%)2(¢(O,zO+Az) 2(2+(%%)2) + 9(0,zO-Az))} , (5-18) 110 which yields the starting point for applying relaxation tech- niques. C. Techniques for Calculation If one employs the appropriate expressions for the potential given in the preceding section, boundary-value problems become very amenable to iterative techniques of solution such as relaxation and its extension, overrelaxa- tion. To employ either of these, our first step is to replace the continuous region of interest with a grid net- work as shown in Figure 5-1. For convenience we set up the grid so that conductor and boundary surfaces lie along grid lines. Next we make guesses for the value of the potential at every point that is pg§_fixed by the problem. Such points will be called "free" points. A convenient initial guess for the free points would be all zeroes; although this will work, it is very inefficient. Intuitively, it seems that an initialization of the free points based on the analytic solution to a similar but less difficult problem is a reasonable first guess. Relaxation consists of replacing the value of the potential at a particular free point by the average of its neighboring points according to Equation 5-8, or 5-12 and 5-18, whichever is appropriate to the coordinate-system. This procedure is applied to each free point in the network, 111 3 6.3 7.3 8.2 I j 2 5.2 6.2 7.2 I 4.0 4.0 4.0 0 o I 2. 3 | -—D Figure 5-1. Two-dimensional x-y grid with 9 points after one iterations. ég's at free grid spacing is the same in the x and y directions. x = 1A, y = jA, where i, j are integers. 112 constituting one iteration. Iterations are continued until a reasonable convergence criterion is met. We have monitored the maximum percent change effected through one iteration. For the Laplace problem it has been found that a modi- fication of relaxation, called successive over-relaxation (SOR), may significantly accelerate convergence (Am69), (F060). Here the change in the distribution of values ob- tained from each iteration of relaxation is similar to a diffusion process; i.e., SOR acts to enhance the diffusing corrections by "amplifying" the averaging process. To utilize SOR one must choose a relaxation factor, 1 < m < 2. For the case of equal 0 and z mesh steps, an estimate of the optimum m can be obtained from the expres- sion ~ 2 wept " W , (5'19) where A is the mesh step size. Denoting ¢ij as the value at (i,j) obtained from the 8th iteration of ordinary relaxation, we get an "over- SOR relaxed" value of 013 by applying SOR = 1-1 A ij ' c (l-u)lJ + 0013 (5-20) 113 Then we replace ¢ij by ¢§3R° Choosing w = 1, reduces over- relaxation to relaxation. Perhaps an example will more fully demonstrate. Ref- ererring to Figure 5-1, suppose the values shown at each point were obtained after five iterations. Consider the point, (i = 2, j = 2); 922 = 6.2, and using Equation 5-8, we obtain Relaxation Step: 032 = 1/A(A.07+7.3+5.2+7.2) = 5.925 . (5-21) If we are using only relaxation, we would move to the next point and compute the average of its neighbors. However, if we are using SOR with w = 1.5, our next step is Over-Relaxation Step: SOR 6 °22 5 (l-l.5)o22 + 1.5 922 -.5-6.2 + 1.5-5.925 5.7875 (5-22) Next we take 932 to be 5.7875 and move on to the next point to repeat the procedure. 11A D. Applying Relaxation and Successive Over-Relaxation A typical problem in electrostatics that is easily solved analytically is the case of two coaxial infinitely long cylinders held at different potentials. Referring to Figure 5-2, the closed form of the solution is found to be 9(0) = (AV/£n(a/b))°£n(o/b) . (5-23) This case provides an instructive application of the tech- niques and a simple means of comparing the convergence rates of relaxation and SOR. We mock-up the infinite length by fixing the potential, at both ends of a finite coaxial length, to be the analytic solution. An equal 0 and z mesh size of 0.25 was employed. One hundred iterations were performed with a Xerox 2-7 computer, first using ordinary relaxation and then SOR with w = 1.75. In both cases an initialization of 9(0) = 50/(l+p) was used. The results for each, with their respective errors, at representative p points are given in Table V-l, after 10, 50, and 100 iterations. In the right- hand column the exact values from an analytic solution are given for comparison. Inspection of the table makes it apparent that the SOR method converges much more rapidly than the relaxation technique (w = 1.0). The time saved in computing can be significant. 115 .08 I me n eq .pxou 0:» CH co>fiw ma coHuSHom ofipsamcm one .o.o H me up 0H0: ma pom 0.0H n o mzfiomp no: gonosocoo gonzo one mooa u we no cap: mH 0:0 o.H u m msfiomg mm: gonosocoo Lassa one .msooosocoo Hmofigocfiaso Hofixmoo wcoa saouficfimcfi mo coauoom mmoLo .mlm opswfim 116 Table V-l. Comparison of Relaxation and SOR Results for Infinite Coaxial Conductors. Numerical % Numerical % Value Error Value Error Exact Solution 0 = 1.0 w = 1.75 After 10 Iterations 1.5 53.67 3A.9 72.92 11.5 82.33A 5.5 7.7A 70.2 13.30 A8.8 25.96A 9.5 2.36 6.1 l.Ao 36.9 2.228 After 50 Iterations 1.5 69.90 15.1 81.31 1.3 82.33A 5.5 8.92 65.6 23.92 7.9 25.96A 9.5 1.2 A6.2 2.05 7.9 2.228 After 100 Iterations 1.5 75.09 8.9 82.33 0.069 82.33A 5.5 l2.A0 52.3 25.80 0.62 25.96A 9.5 1.08 51.6 2.21 0.6A 2.228 . 117 As an example of the usefulness of the SOR technique the example of a finite cylindrical conductor coaxial with a larger radius and longer closed cylinder held at a dif- ferent potential has been solved. This geometry is shown in Figure 5—3. Such a deceptively simple looking problem is beyond closed form solution, but by employing SOR methods (Equation 5-12 and 5-18) the problem can be solved in short order. The matrix map of solutions results in the equi- potentials plotted in Figure S-U. In Figure 5-5 we show a second cylindrically symmetric system that is intractable by analytic methods, yet presents no real difficulties for an SOR treatment. (This particular case is actually a prototype of an electrostatic particle focusing system that we employ in a recoil—mass time-of- flight spectrometer.) The potentials obtained from an SOR treatment of the appropriately scaled system will be used in calculating trajectories in such a region. The equipotentials that result from an SOR solution of this electrostatic problem are shown in Figure 5—6. After gaining some experience in field calculations and the estimation of the best relaxation factor to use, we began the calculation of the potentials that correspond to our TOF system. The calculations were accomplished in two parts: (1) the acceleration zone and (2) the drift zone. This was possible because the grounded mesh separating the acceleration zone provides a natural boundary surface for 118 Lowpmfi Ca oomoaoco .ooa .mecfiazo oo©:309w n e um cam: pouospcoo Havapccfiamo opaCam .mum opswfim 9 II a 119 .mlm opsmfim ho mauoEoow Emanopq on COHpSHOm mom Soak wCHpHSmmh mHmecmuoqfisum .nlm opswfim O ”9\ ON. 00 120 .03» mo nonomm m an mzfiomp mommaxm zHuQSAQM awn» you036200 Lopzo cooczogw Ca comoaoco Aooa I e nuazv pouosocoo HQOfiaocfiazo mpficfim .mum magmas 1 ON 1‘ w n: w .. . clllrgwwm- ................ VMM. ........................ ..colgw v n to. oo_.e\ .. own .. owns a M .. 0.9 ~ 121 .mum mpswfim mo zapoEomw Emanopa on COfipSHom mom Scum mafipazmmh mamfipcmpogfizdm .wlm ohsmfim 122 both regions. This allows us to use the mesh plane as a boundary surface for both portions of the calculation. Acceleration Zone The acceleration region consists of a circular stain- less steel plate held at 6 kV, a fine wire mesh disc at ground and a boron nitride (BN3) insulating support connect— ing the two plates. Basically it is a cylindrical capacitor with a dielectric "ring" connecting the two plates. The radii of the charges plate and the mesh disc were both 1.83 cm. The length of the dielectric and therefore, the separa- tion of theplates was 1.67 cm. The thickness of the insulat- ing support was o.u7 cm. The inclusion of the eielectric in our model complicates the calculation somewhat since we no longer have pure Dirichlet boundary conditions on all surfaces. At the surface of the dielectric we have Neumann boundary conditions. Dirichlet boundary condi- tions mean that the value of the function is specified on a particular surface while Neumann boundary conditions mean that a condition on the derivatives of the function are specified. In electrostatics the relevant Neumann boundary condition is that the normal component of the electric displacement vector, Dn must be continuous. As a result the averaging equation for the potential presented in the previous section is not valid at the surface of the 123 dielectric. However the continuity of the normal component of the electric displacement can be translated into an equivalent expression using the finite divided differences of the potentials. Using this method one can then obtain a new expression for the potential at the surface of the dielectric. The relevant equation is ¢ = (Q + 1,3 i-l,J €D ' ¢i+1,3)/(1 + 8D) (5-2u) where so and ED are the dielectric constants for vacuum (60 = 1) and BN3 (5D = “.08), respectively. This expres- sion was used at the inner and outer surfaces of the di- electric. A further complication in this treatment was that the Neumann condition on the outer surface leaves the potential unfixed on any surface in the rho direction which is an unstable situation. There are a number of ways to deal with this problem (see (Ac70)), the easiest and most straightforward is to "enclose" the entire acceleration zone in a large cylindrical box at a fixed potential. Physically this is very reasonable since in actuality the acceleration zone is encased in a vacuum chamber that is held at ground. This was the approach we chose to utilize. Our mesh step was .0235 cm enabling us to place 20 nodes in the insulator along the c direction. Two hundred iterations were performed until the maximum fractional change in the mesh was less than a part in 10“. Shown in 129 Figure 5-7 is the equipotentials resulting from the over- relaxation calculations for the acceleration zone. The first nine lines represent drops in potential of 0.6 kV as one moves from left to right —- the direction of motion for a positively charged species. The spacing of these equipotentials is fairly regular which indicates the electric field strength is uniform. This is especially true along the axis of symmetry (O = 0) as can be seen from Figure 5-8. The top curve is a plot of the Z component of the electric field (at o = 0) as a function of distance from the charged plate. This component is relatively constant throughout the acceleration zone, the total drop in field strength is about 5% of the maximum. From these considerations one would expect a positively charged ion to undergo a uniform acceleration in the Z direction. Now if we follow one of the equipotentials in the first half of this zone it begins to bow slightly as we move away from o = 0. This indicates the presence of a p component of the electric field. The lower curve in Figure 5-8 shows the 0 component of E as a function of the distance from the symmetry axis. This particular curve was calculated at Z = .5 cm. As required by the Neumann boundary conditions the electric displacement normal to the surface is con- tinuous at the interface of facuum and the dielectric. However, the electric field component normal to the surface has a discontinuity at this interface. This implies the 125 . .Uonpoe cofip Imxmaogzo>o an nonmazoamo mcom cofipwpoaooom on» CH wamfipcouoafisvm .nlm ogsmfim 5088 92:382. mzm / 1::S l. Z.:.mwhm 4‘ m0 __“.. .Euan fl: o --- --=== 126 .Q n x m>pso am now .N u x m>pso um pom .ocou coaumpoamoom one cfi oaofim oapuomam oopmHSQHmo on» no mquCOQEoo a cam N no oHQmem .mlm mpswam 653 X ”.1 Orv.- no . Oh. mm. 0.0 1 _118 .911 1.50%.; 3: mm 5.1.01 10 n . loom m 0:66.06 9.: ICON. B 823.... . coca. loom. AEU\>V [OOVN [000m 1 0 a I D o 0 b D 0 0 I l 0 ”N D o 0 u 0 l .0 0 [00mm 8.01:0: m LOON¢ 127 presence of a surface charge. There is indeed a surface charge, it-is the induced polarization charge resulting from the external applied E field. The 0 component of the field is much weaker than the Z component, its maximum value is about .08 times the Z part. Nonetheless it is a defocusing mechanism through much of the acceleration zone. This effect is most pro— nounced in the first three quarters of the region. Also the field in the 9 direction increases as the distance from the center increases. Therefore particles starting out far from the center of the collection plate should be most strongly affected. This is one reason for focusing the spray on center as best as possible. Drift Zone Once the recoil ions have passed through the wire mesh disc it enters the drift zone region. The region is made up of a small radius wire (0.0025 cm) concentric with a much larger radius pipe. Ideally this was meant to provide a logarithmic potential to focus the recoils onto the CEMA detectors at the end of the flight path. It was hoped that only an electric field in the p direction would result from this configuration. Unfortunately the wire can't be in— finitely long so it has to start and end somewhere. This means that near the end of the wire the field are not purely functions of 0, they have a dependence on Z that is very 128 difficult, if not impossible, to determine analytically. Another nonideal aspect to the real geometry is a sudden expansion in the radius of the outer conductor. This ex- pansion, which occurs about 1/3 m from the acceleration region, results in a change in radius from 1.83 cm to 5.2 cm. This discontinuity preserves the symmetry in the 0 direction but definitely destroys the Z symmetry in the area of the expansion. Hereafter I will refer to the area of the radius change as the expansion region. Obviously we must again employ numerical methods to solve Laplace's equation for the potentials in the flight tube. In setting up the problem there are immediate differences from the calculation on the acceleration zone. The first difference is an advantage. In the flight tube region there are only conducting boundaries to deal with, so only Dirichlet condi— tions are relevant. This means the relaxation codes are a bit simpler. The second difference is a practical dis- advantage. As we pointed out earlier the step A should be scaled to characteristic linear dimension of the problem geometry. However in the flight tube region the most obvious scale is the radius of the wire -— 0.0025 em. But if we used a A = 0.0025 cm for our .5 cm x 100 cm system we would require an array of nearly 80 million elements -- totally unrealistic for our computer. Increasing the step size alleviates the problem of array size only at the ex- pense of potentially large errors creeping into the 129 calculation since the error goes as the square of the step size. In our studies on the relaxation technique applied to the geometry of Figure 5-3 we found that the potentials approached the idealized infinite wire case as we moved away from the wire end. At a distance on the order of the separation of the wire from the ground plane the potentials calculated are well approximated by a logarithmic form. In the flight tube the separation of the wire and the ground disc is about 7 cm so after roughly 10 em down the wire we thought that the use of a logarithm in 0 would sufficiently describe the potential field distribution. This effectively reduced the number of array elements required for the Z direction. We still had to contend with the problem of the small radius of the wire for the remaining regions. Following a more intuitive than rigorous argument we decided to use the actual wire voltage as the boundary value for the appropriate p = 0 points (i.e., points in the wire) and to use a boundary at one small step size from the p = 0 points. The fixed value of the potential on this surface was obtained from the logarithmic form discussed previously. Basically we replaced the "real" wire with a "pseudo"-wire of a more manageable radius. This final approximation allows us to reduce our array size to a realistic number for computation. The final mesh step size was 0.0235 cm. In the expansion region we again used the method of employ- ing the logarithmic form for the potential at a reasonable 130 distance from the radius expansion. The results of the calculations are shown in Figure 5-9 as a number of equi- potentials. The effect of the wire ends are clearly dis- played at both ends. At the end nearest the acceleration zone the equipotentials resemble cylindrical waves, the cor- responding electric fields will have both 0 and 2 components. So in this region the ESPG is not focusing only in the radial direction. In fact, near the axis of the wire the effect of the particle guide is to accelerate predominately in the z direction. At the other end the field acts to focus the particles on the CEMA detectors and also to deaccelerate them in the z direction. At the expansion region we can easily see the field "response" to the sudden radius change, the equipotentials seem to "expand" to fill the suddenly enlarged region of space available. The effect of this part of the fields will be a perturbation on particle tra- Jectories that had been established as stable by the time they reach this part of the system. In a sense we can view the expansion as resulting in an easing of the attractive force that a charge particle "feels" as it passes through this region. The net effect is a defocusing of particles that had made it down a third of the flight path length. In the remaining regions the equipotentials are parallel to the wire with no real 2 dependence as we should expect. These regions, away from the discontinuities, make up about 80% of the system length and act to focus in the radial direction only. 131 .Uonme :ofipmxmaohpw>o an cwpwHSono ocou p.338 on» :H mamfiucmpoafiscm .mlm madman .n-\o~.\.~.\o..\=.\n.\..\ 132 Calculation of Particle Trajectories The relaxation calculations Just described allow us to obtain the electrostatic forces acting on a charged particle passing through our system. We are now in a position to calculate particle trajectories. This means we will inte- grate an appropriate set of equations of motion. For our case this set will be the Lagrange equations for a charged particle in an external electric field. In cylindrical co- ordinates we have m -—§ = D¢ ‘ q -— (5-25) m -—§-

(5446) 1+1 121 1+1 p1+1 Final Value y1+1 = c1+1 ‘ p1+1 (5'u7) The term ei+l’ the truncation error, is checked at each step to determine whether the step size should be changed. If we compare the two algorithms just presented it can be seen that the Runge-Kutta routine requires four separate 139 evaluations of the derivative term while the Hamming pre- dictor-corrector only needs two evaluations of f(x,y). Generally this difference will mean that the predictor- corrector algorithm should be faster than the Runge-Kutta. In an attempt to see how significant the difference in speed of the two algorithms, we ran our computer routines for each algorithm on a number of different analytical functions. In Table V-2 are the results of the comparison, in every case the Runge-Kutta algorithm required at least “0% more time than the predictor-corrector routine. This is a significant difference. When we begin calculating tra- Jectories in our calculated fields the derivative evalua- tions involve a number of interpolations in the field map. These interpolations can be very time—consuming so we want to keep these derivative evaluations to a minimum. This is another advantage of the predictor-corrector method. For the actual trajectory calculations we used the Hamming method with a variable step size. Whenever the truncation 5 error term was greater than 1 x 10' the integration step size was cut in half andtflueprocedure continued. If the error term was less than 1/50 of the upper limit on the error then the integration step size was doubled. This was very useful for our problem since the particle traverses long regions of slowly varying potentials on much of its flight path so the step size can be very large. In the more rapidly changing regions the step size reduces in 1H0 Table V-2. Comparison of Speeds for Runge-Kutta vs. Predictor-Corrector Routines. Time of Calculation for 5000 Steps (in seconds) Function Predictor-Corrector Runge-Kutta RK/PC tanh(x) - X 15 23 1.5 eXp(x) - x - 1. 15 21 l.A sech(x) - 1. 16 22 1.“ cosech(x) + l 19 31 1.6 cosh(x) - 1. ' 15 21 l.“ sinh(x) - x . 15 21 1.” log(x) - x + l. 16 22 l.“ x + l/X 16 22 l.“ cos(x) 15 21 1.“ sin(x) 15 20 1.3 1Ul response. In addition to the error monitor we checked the constancy of the particle's total energy. Since we have an electrostatic problem the sum of the kinetic and potential energy should be a constant. This was true for our calcula- tions to one part in 105. Special care had to be taken in crossing from the acceleration zone into the drift zone. If one does not insure that one of the integration steps "lands" on the grounded disc position then it appears that the electric field in the acceleration region continues to act on the particle a short distance into the weak field zone. The result of this is strong nonconservation of energy and very incorrect results. After integrating down the length of the flight tube the particle's 0 position was checked to determine whether it was within the radius of the CEMA detectors (“CEMA = 1.25 cm). If it was, then the time, p position and velocities are recorded and the event is considered a successful hit. CHAPTER VI THEORETICAL MODELING RESULTS In the previous chapter we presented the tools that will be required to model a specific time-of-flight peak that has been experimentally observed. In the following section we shall present the formal aspects of the problem. This will include clearer definitions of 1) our approach, 2) the particular isotope chosen for study, and 3) the most effective use of the mathematical tools discussed previously. 1. Approach Basically our problem is that we have, as a physical observable, a distribution of events as a function of their associated times-of-flight through the SIEGFRIED system. Our goal is to relate the observed distribution to a reason- able set of initial conditions that will be related to the particular isotope of interest. Since we are interested in correlating the particles' initial recoil energy to the observed peak broadening, the kinetic energy is an important initial parameter that must be specified. Also, the source of the recoils is a finite size spot, so we should include 192 1H3 this fact in the initial distribution. Fortunately the source spot is approximately circular and well-centered on the collection plate. After passage through the SIEGFRIED TOF system the initial particle distribution in position and energy is transformed into the observed distribution in time-of-flight. If we denote the initial distribution as NO(E,DO) and the final distribution Pf(T), then the process of measurement can be represented as shown below: NO(E,oO) ———- SIEGFRIED —-—~ Pf.(T) Mathematically we will consider the action of SIEGFRIED on the initial distribution as expressed by an integral transform of the form: Pf(T) = fffS(Egpo:Tspf)NO(EapO)dedEdof (6‘1) Here S(E,pO,T,pf) represents the transformation of a particular kinetic energy E and position p0 distribution in the energy and position interval dEde. (The effect due to the finite size of the detector is taken into account by integrating over final positions with of less than or equal to Rdet’ the radius of the CEMA detectors.) Ideally S(E,pO,T,pf) and NO(E,pO) would be expressible in a closed form. Then we could merely integrate and solve the problem easily. Even more appropriate would be to 14M invert our approach: suppose we express our integral trans— form, Equation 5-1, as Pf(T) a§ENO(E,oO) (6-2) Here §Z7represents the integral transform operator. Now if an inverse oqug, exists, then we can invert the problem. Denoting the inverse bycgy'l we get from Equation 6-2 _ -l NO(E,OO) -Q me (6-3) This last form says that, by using 3‘1 on the observed final distribution Pf(T), we can reconstruct the initial distribution function. That is precisely what we would like to do -- if we could. The problem is that the kernel of the transform, 8 is not expressible in a simple form. We do know that solving the equations of motion for a set of initial condi- tions that result from a particular distribution will allow us to follow the time evolution of the distribution. This is the functional effect of S on the initial distribution, but we cannot "find" an inverse, so we are basically solv- ing the problem in an unavoidably roundabout manner. We will assume a given initial distribution, predict a test distribution F(t) and compare this with Pf(T), the ob- served distribution. 145 2. Choice for the Study The particular isotope I have chosen for this modeling study is 23Mg. It is a positron emitter with a half-life of 12.1 sec. The decay energy is typical of isotopes seen by SIEGFRIED (QEC = b.056 MeV). 91% of the decay proceeds from the 3/2+ ground state of 23Mg to the 3/2+ ground state of 23Na. The decay has 70% Fermi fraction and 30% Gamow-Teller fraction (as defined in Chapter IV). Experimentally 23Mg is seen very strongly in the TOF spectrum resulting from 27A1 + 70-MeV 3He (See Figure 3-1). Its TOF peak is the largest feature in the spectrum and is very well separated from other peaks in the spectrum. The shape is typical of TOF peaks we have observed, and the high number of counts in this peak allows us to fit it very well with SAMPO. This gives us good estimates of peak parameters such as the width and the tailing. Finite Source To take account of the finite source size we considered the source to be made up of 8 concentric rings. Each has a mean radius of n°(.075 cm) with n = 1,2,...8. 1M6 Initial Conditions An important part of our modeling will be the integra- tion of a large number of particle trajectories. While the integration codes are fast (roughly six complete integrations per minute) we want to be as efficient as pos- sible. A very significant increase in efficiency can be obtained by eliminating initial conditions that result in trajectories that strike the boundaries of the system or are otherwise unreasonable. This is an important considera- tion and will be the focus of this section. The maximum initial recoil energy, R is fixed by the decay energy as shown in Appendix II. If we denote the maximum kinetic energy of the electron by T0 (in MeV) and the mass of the recoil by A (in amu) then the maximum kinetic energy of the recoil R (in eV) can be obtained by the following equation from Appendix II R — iilT _ A T + 1.022) (6-4) 0(0 This is a bound on the square of the velocities. For 23Mg the maximum kinetic energy of the electron is 3.03 -MeV. (For positron emitters T0 = QEC - 2mec2). The result- ing maximum recoil kinetic energy is 290 eV. By using a gross model of the TOF system and some 1“? simple energy conservation arguments we were able to ob- tain rough limits on the transverse velocity and the angular momentum. To demonstrate this we use the geometry of Figure 6-1. In region I the electric field has only a Z component, at Z = d is a ground plane and region II has a potential that is purely logarithmic in p. This means that a particle of mass m and charge q starting from Z = 0 experiences only accelerations in the Z direction in region I and only 0 accelerations in region 11. The system is cylindrically symmetric and so the angular momentum about the Z axis (L) is constant. Consider a particle of mass m and charge q starting from Z = 0, O = O with an initial kinetic energy R. In 0 cylindrical coordinates .2 . R = 1/2m(p + L2/(m p )2 + 22) (6-5) 0 o 0 Since it starts from the plate held at 6 kV its potential energy is 6 keV. Denoting the potential energy by U, we have U = 6 keV The total energy, E of the particle is conserved so E = R + UO = constant (6-7) 1118 .Eoumzm hoe 639 no Hmpoe mumEonpdd< .le opswfim >x® URN OuN .__". O >C:Iu >/ v .1 E J. 149 In crossing the region I only the velocity in the Z direc- tion is changed. When the particle reaches the ground plane it has a total Z velocity, Z(d) 2(a) = z: + /"2U'O/—m (6—8) The time necessary to cross region I is td, which is ob- tained in Appendix I as . . 2U = £11751- [-z +122 + ——9 ] (6-9) 0 The p velocity is not constant if the angular momentum is nonzero, at Z = d and t = td we have O m 2 2 L 1 1 - _- _ - +—<—§-—§> <6—1o> 0(td) - ad -‘/0 2 po 0 There is drift across the region in the direction given by Equation AI-19, which for t = td gives [(6§ + (L/moo)2)t + 600012 + (L/m)2 pd = 0(td) = p0 ' 2 2 (popo) + (L/m) (6-11) 150 Now we consider the particle to be infinitesimally far from the ground disc at Z = d but now in region II where it fields a force due to the logarithmic potential in 0. We assume the p position is given by Equation 6-11 and the O velocity by Equation 6-10. Again the total energy is constant but in addition the velocity in the Z direction- is also constant so defining T as follows: T = 1/2m (52 + L2/m292) (6-12) Then the quantity T + W = constant = Ed where W is the potential energy. In region II, W is V _ o W - EHTE737 £n(p/b) (6-13) V0 is the potential on the center wire, a is the radius of the outer conductor. Thus, at the beginning of region II ‘2 E mod mL2 V d - 2 + 3—2- 4' mm a0 b £n(Od/b) (6-1“) O d 151 Defining new quantities T¢ and To as 2 L T (p) = L“— (6—15) ¢ 202 m52 To = T (6-16) we have Vo If p = b, the particle has struck the outer boundary and its p velocity 6 and potential energy W are both zero, so: Ed = T¢(b) (At 0 = b) (6-18) From Equation 6-17 we have V Ed = Tp(d) + T¢(d) + EaTé%BT-£n(pd/b) = T¢(b) (6-19) From Equations 6-10 and 6-16, Tp = Tp + T¢2> <6—2o> 152 T¢(b> = T¢(oo)'(oO/b)2 (6-21) Rearranging Equation 6-19 and substituting, we have -v Tp(oo) + T¢(oo)(1-(oO/b)2) = aob £n(pd/b) (6-22) n which gives bounds on the initial 0 position and velocity and angular momentum. Suppose initially the angular momentum is zero, then Tmax -V 0 (p0) = E57526? 2n(pd/b) (for T¢(po) = 0 (6-23) with This fixes the bound on the maximum initial 0 velocity. If L = 0, any initial velocity less than this maximum will not hit the outer wall. If we want the corresponding maximum angular momentum for 50 = 0, then Tp(po) = 0 For Tp(po) = 0 max

480 omx_s_ Ammv n. 16M NO(E,DO) is ours to vary. In representing the distributions, we divide the energy range bounded by Equation 5-M into a set of ten subintervals of a set length. So, in practice, our approximation to Equation 6-1 can be expressed as 10 Pf(T) = 3:1 aiS(Ej,Tj) NO(EJ)(AEJ) (6-27) and 10 ( P (T) = X p T ) (6-28) f j=l f J We have suppressed the p arguments for convenience. The index i refers to the ith bin in the energy interval and AEi is the width of the ith bin. The term “i is a weight factor. For the flat distribution NO(E1)AEi is constant, so we have aiSJ (6-29) 165 with SJ 5 S(EJ,TJ) (6-30) Thus, by evaluating the predicted Pglat for a flat distribu- tion, we are able to obtain the elements SJ which are the most important and elusive quantities. The combination of Equation 6-27 and 6-28 allow us to recast our problem in the form of matrix equations. P = S N (6-31) We construct the S matrix from Equation 6-29 and then proceed to apply Equation 6-31 to the various initial distribution functions of interest. Results of Modeling The determination of S was accomplished as just dis- cussed and led to a number of reasonable conclusions. First, the TOF of a particle is completely a function of the Z velocity within a few nanoseconds of the 5 usec TOF. While not surprising, it is something that could not be stated definitively before the calculations because of the field nonuniformities. Second, the effect of the kernel function S will not transform the flat energy distribution into the observed form. The result is basically still a 166 square wave as we input. In a way, the resolution of the difficulty seems to be hinted at by the first conclusion. This tells us that the TOF is dependent on the component of the momentum in the Z direction. Instead of using a flat energy distribution, it seemed reasonable to attempt a flat momentum distribution. Actually we used a flat distribution in the Z component of the momentum since the results are independent of the other components. In addi- tion to the independence of the observed distribution on the transverse components, the initial distribution in these components are severely restricted by the limit curves of Figure 6-2 and reduce to a constant, multiplying the distribution in the Z momentum. The results of the calculation using the flat momentum distribution gives a set of functional values of P(Ti) at unequally spaced values of T1, the time-of-flight for the ith bin. Treating these as functional values at un- equally spaced base points (Ti's), I used a Newton Divided Difference Interpolation (See (Ca69)) to generate inter- polate values at equally spaced time intervals. These values were then used to generate the plot shown in Figure 6-8. A similar procedure was followed for the other distribution functions. For the Fermi and Gamow-Teller. initial distributions, the distribution functions that are the momentum equivalents of Figures 6-2 and 6-3, respec- + tively, were used as the NO(P,pO). The distribution 167 2000‘— 21mg .J UJ Z 2: <1 I L) \\ f2 - ~46 Z :3 O C) 75-: ' I35 ——-4 O _- - - I ., - -11-.. -.I nsec 1 _-..--.-..l mi] 4.5 4.7 4.9 5.| TOF (psec) Figure 6-8. Theoretical TOF peak predicted from flat momentum distribution. 168 Table VI-2. Comparison of Peak Widths for Various Distribu- tions. Obs. flat Fermi GT mix Peak No No No Nn Base Width (nsec) 1&5 135 13 135 135 Width at 1 Max 90 75 120 130 120 (nsec) Width at 5 Max 20 16 70 105 100 (nsec) 169 corresponding to the correct admixture of F and GT decay ror 23Mg was constructed by a linear combination of the F and GT distribution. The form used was Pmix = a PF + (l - oc)PGT (6-32) here a = .7 for the case of 23Mg. The predicted TOF distributions corresponding to the Fermi and Gamow-Teller initial distributions are shown in Figures 6-9 and 6-10, respectively. In Figure 6-11 is shown the observed TOF spectrum resulting from the reaction 27A1 + 70-MeV 3He which was discussed in Chapter III. The focus of interest is the large 23Mg peak that we are attempting to simulate. On each of the 23Mg peaks in Figures 6-8, 6-9, 6-10, and 6-11 are given the full widths at the base, at .1 peak maximum and half maximum, all in nanoseconds. A visual comparison of the three predicted TOF dis- tributions with the experimental 23Mg peak heavily favors the peak shown in Figure 6-8 as the best reconstruction. The peak resulting from the Gamow-Teller distribution (Figure 6-10) has a tail on the short time side as the experimental peak, but is clearly the wrong shape. In addition, this peak is far too broad except at the base. The Fermi distribution gives rise to a peak (Figure 6-9) that is actually a rough mirror image of the experimental. 170 zooo FERMI DECAY 23 _J Lu 2: 2: wZIo> + H¢h~ mo mpospopo mo Sappooam hoe ©m>nmmno .Hatm mhswfim 335 “.0... 09m 0000 Ono—mm 00.0 oom¢ 005? 000? . p h _ mam :WN oznu no.3 173 By this I mean that it has a sharp cutoff on the short TOF side and tailing on the longer TOF side which is the opposite of the experimental 23Mg peak. The peak in Figure 6-8 results from a flat momentum distribution and is an excellent reproduction of the experimental peak. On the short TOF side there is the long bowed-tailing that is so obvious in Figure 6-11. Then the distribution comes to a sharp spike with the sudden dropoff on the long TOF side in the same manner as the experimental peak. The peak resulting from the initial distribution from Equation 6-32 has a shape similar to that of Figure 6-9, except that it is broader at .1 maximum and half maximum (See Table VI-2). Quantitatively the Nglat (flat initial momentum distribu- tion) gives the best predictions also. Referring to Table VI—2, the base width predicted is within 6.9% of the experi- mental width, the width at 0.1 maximum is within 16% and the full width at half-maximum (FWHM) is 25% off. This com- pares very well with the various widths predicted from the Fermi, GT and mixed distributions. These last three distributions result in widths at .1 max and half—max that are from 50% to 500% in error! The base widths for all the distributions are all the same since they correspond to the same definite limiting type events. That is, the base widths are set by the time differences between a particle starting out with 0.0 initial recoil energy and a particle starting out with the maximum velocity in the Z l7“ direction. Both cases can only occur in one way and so are independent of the initial distribution assumed. In many experimental instances, one uses the position of the peak maximum as the best estimate of the position of the centroid and then uses the associated time-of-flight to calculate the mass from Equation 3-2 (in Chapter III). The position of the peak maximum for the observed 23Mg is M910 nsec. The "flat momentum" peak has a predicted posi— tion of the maximum at 4915 nsec and the Gamow-Teller peak at “908 nsec; both are in good agreement with the real peak. The Fermi distribution peak of Figure 6-9 has its maximum at #800 nsec, which is well off the correct position. In fact, if we use Equation 3-2 with TOF = “.8 msec and HV = 6 kV, we get A = 22.0 —- off by one complete mass unit. This is not the case for the observed peak but does demon- strate the care that should be taken in choosing positions (on the peak of interest) to calculate the mass. Although a little surprising, the simple flat distribu- tion is the most successful in explaining the observed 23Mg peak. The success is fairly impressive; (1) the peak position is correct to 5 nsec, which is a .1% error, (2) the base width is correct to 6%. Thus, theoretical broadening accounts for nearly 94% of the observed peak broadening. The simple model demonstrates the importance of the recoil energy on the observed peak broadening very clearly. CHAPTER VII CONCLUSIONS In this thesis I have tried to present experimental results and theoretical analyses concerning the SEIGFRIED Recoil Mass Identification system. Before closing, I will briefly summarize the work presented and the conclusions drawn throughout this work. First, it was demonstrated that the system works well in a number of different mass regions, although the chemi- cal nature of the HeJRT system can be a real concern. Second, the observed TOF peaks were shown to contain in- formation about the recoil ion's initial kinetic energy, which is a direct measure of the associated decay energy. To demonstrate the relation between the recoil energy and the peak shapes, I adopted a double-edged approach. First, I showed that a parameterization of the observed peak parameters yields reasonable predictions of the recoil energy and the decay energy. Second, I developed a reason- able theoretical model using numerical methods and used this model to simulate a theoretical peak corresponding to 23Mg which I observed experimentally. The results of the modeling study were very successful. Using a simple 175 176 distribution, the experimental peak shape is reproduced very closely. The broadening of the peak was shown to be due to the recoil energy, and the theoretical broadening from this was 95% of the observed broadening. The overall success of our study in the experimental and theoretical studies proves the viability of using TOF system such as SIEGFRIED to identify mass products (measure A) and measure decay energies through the recoil energy. Combining these two measurements in a simple system provides the nuclear experimentalist with very useful and different tool for the study of nuclei far from stability. APPENDICES APPENDIX I As an approximate model of the time-of4flight system, we will use the geometry shown in Figure AI-l. In region I, the acceleration zone, there is a uniform electric field E with only a Z component. In region II there is a field- free drift zone of length 2. Thus, a particle of mass m and charge q is uniformly accelerated across region I and then drifts through region II. Using cylindrical coordinates (p,¢,Z), the classical Lagrange equations of motion are: fi- (m6) = moi»2 - q 3—‘3- (AI-1) §%-(m92$) = Oq %% (AI-2) £5- = -q 3% (AI-3) In region I E = g 2 (AI-u) ¢ = V(l - Z/d) (AI-5) ad}- (mz) = q g (AI-6) 177 178 .H-H< mpswfim I- *-- --- N ’ 179 Integrating twice with respect to time, 2 - - 9.12. z - Z0 + Zot +md 2 (AI-7) where Z0 is the velocity in the Z direction at t =-O. We consider all particles starting from Z - 0 so ZO = 0. Z = Z t + QY~§E (AI-8) o md 22 Solving for t, we have _I£1_° 42 2gVZ t - qV IZO + Z0 + md (AI‘9) When Z = d, the end of region I, the particle undergoes no further acceleration. Defining T = t(z=d), the acceleration time, we have from Equation AI+9 md ° 2 2 V =— - + + - r qV { ZO J20 -%— (AI 10) Since 0 is independent of o and p in both regions I and II, the remaining equations of motion are simple. 180 First, 3¢/B¢ = 0, so Equation AI-2 tells us that mo2$ is a constant of motion. Actually, this is the angular momentum about the Z axis. Thus, we have L E mp2¢ = constant (AI-ll) Using L to eliminate ¢ from Equation AI—l, we have for the 9 equation 2 2 g_% = £5 3? (AI-l2) dt m 0 Equation AI—l2, in spite of its simple form, is an example of a second-order nonlinear differential equation. While nonlinear differential equations of any order are notoriously difficult in general, this particular equation can be solved. Since appropriate references were not found we will solve this in a step by step manner! Defining a new independent variable x, Lt X = IT (AI-l3) Equation AI-12 becomes __5 = __ (AI-1A1 181 Defining u = g%, we have du l 0 Integration gives 2 _ 2 _ 3; _ j; u uo - 2 2 (AI-l6) oo p or 9.9.- 2 .1. ._1._ dx - uO + 2 2 (AI-l7) O 0 Integrating from x = 0 to x, we have x = 2{ .J 013 +013) 2 -1 - uopo} (AI-18) 1+(uO OOp ) Inverting to solve for p, we get [(u2 + 1/p2) x+u p ]2 + l p(x) = ___o 0 0 0 (AI-19) u2 + l/ 2 o 0o L, and Do or in terms of t, 182 JE(5303+(L/M)2)t/p:+éoool2 + (L/m)2 p(t) = 00 (AI-2o) (popo)2 + (L/m)2 In region II there are no external forces acting on the particle, so the equations of motion are trivial. The only one we need to look at is Equation AI-3. For a par- ticle "starting" from Z = d we have 2 g_% = o (AI-21) dt Z = VZ a constant (AI-22) and Z = vzt + d (AI-23) where t is measured from time of passing poind d. From the discussion concerning region I, we know that when Z = d its velocity Zd is 2 = J 22 + 3%?- (AI-2A) Thus 183 We are interested in the time a particle takes to drift through region II. Calling this time T from Equation AI—23, we have T = (l - d)/VZ (AI-26) T = (z — d)/ z: + §%y_ (AI-27) This will be called the drift time. The total time of flight is the sum of the drift time and the acceleration time: TOF = T + T (AI-28) (2 d) md 2 2 v TOF - + av {-20 + ‘Jz + —9—- } (AI-29) APPENDIX II A 8 decay is actually a three-body event, involving the decaying nucleus, the 8 particle, and a neutrino. If we define the following momenta and energies: F momentum vector of daughter nucleus of mass M E electron momentum vector 3 neutrino momentum vector Eo total decay energy E total electron energy Ev total neutrino energy R recoil energy then the appropriate conservation laws take the forms: E0 = E + Ev + R (AII-l) q = Ev/C (All-2) E2 = p2C2 + m2cu ' (AII-3) R = r2/2M (AII-A) 18A 185 ? = E + E (All-5) 2 2 + 2 q + 2pq cose (All-5) since R << E + Ev so = E + EV (AII-7) 2 E3 (EO-E)2 C C According to relativistic energy-momentum relation for the electron, p2 = (E2 - m2o“)/c2 (All-9) M II (E -E) 2 2 u (E2-m2cu)/02 + (EO‘E)2/C2 + Co JE 'c'm C 0039 (AII-lO) 2 . R = gfi" 1'2 (E2-m 0“) + (EC-E)2 + (EC—E) \1E2-m2ci cos 2M0 (AII—ll) We want R in terms of the kinetic energy of the electron, since for a“ decay the QB in the literature represents 186 the maximum kinetic energy of the electron. Defining l as the electron kinetic energy, we have R = 1'2 T(T+2mc2) + (TO-T)2 + (TO—T) VH