ham? in: . .a Dr: .. r.. .1. an, 4..." an: .n . Caruh‘ . 3g 3. . J. I 3.4.1.. . .fznw. .4 .m . )1... .. .311 Ed. .. .f z. .3... as a .. use .. ..: d. 2. iii ‘ ... 4...... L... ..:. .2. .. an». .32...“ )4 .2 JAG .. 4.14%... i . 131.1 {4. KR: mvmmflm... .w t.- A 5Y2. l .. . 13:! a...) . .4 £12.13 w n 4 5!: Links: 1. 4.. xi... in“... “I 1 1. :fl-qr . l .1)“. 3255...; I... .. {1:15 fi ..: r 4.51.13 55’ I: . #3? i . . .1 .5. 61.2.3: 34.. ‘55:;0‘lvntal.“ Diflhwwvlfl: . _ 3.2:. ”$13.2. fitnfi‘ :. 3.58.. 40...: 4: . I... .. 3.0.2.341... z. 35.... ....«u..fiHJ.3t32:,px IR...» 3.. .7 2 4 . . . . 9.4m? a»? 213.34....“ .. p . . £249. :7 . . . . ysaamnininoumauvs )5, t... .4..........._.. 14 ..1 L .wl,. ”.2.- ; . 1.. 3':..3.v.\....4 1.1:. 2.4:. ..4hn..,:_4.a.d1: 3.5%. a!» $1.44.. ..Y .55.: : . . ; : . r :L. .. . .m...»...§.:.13 . I 1‘ it.“ lizmmwfismwhfisé . . ., . . . ”Hymn! .1 .. . L 1 an. . .. f .M...voy.\.......:fnvu.w&u 3.591. was»: .Mfiwfifi. . , IT; in}. LIBRARIES MICHIGAN STATE UNIVERSITY EAST LANSING, MICH 48824-1048 This is to certify that the dissertation entitled THE STUDY OF (P,D) REACTIONS ON 13C, 118, AND 1088 IN INVERSE KINEMATICS presented by Xiaodong Liu has been accepted towards fulfillment of the requirements for the Ph.D. degree in Physics and Astronomy ’LhimAkiflbflg Ikyuapy Mala Professor 5 SignatureU 5/6/05 Date MSU is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2mm THE STUDY OF (p, d) REACTIONS ON 13C, ”B, AND “’Be IN INVERSE KINEMATICS By Xiaodong Liu A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the Degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 2005 ABSTRACT THE STUDY OF (p, d) REACTIONS ON 13C, “B, AND “’Be IN INVERSE KINEMATICS By Xiaodon g Liu . . . ll . . This work studied the one neutron transfer reactions on 13C, B, and 10Be In Inverse kinematics using detectors with high angular and high energy resolution. The (p,d) reactions were performed using secondary beams of 13C, 11B, and 10Be on polyethylene targets (CH2)n. The experiment was performed at the National Cyclotron Superconducting Laboratory (NSCL) at Michigan State University. Neutron spectroscopic factors have been extracted for the transfers from the ground states of 13C, II 10 12 10 9 13 B, and Be to the ground states of C, B, Be and from the ground state of C to the first excited state of 12C. The theories of DWBA and ADBA were reviewed and the approximations of zero-range, finite range, and non-locality were examined. Sensitivities of the optical-model potentials in the extraction of the spectroscopic factor were analyzed. The results of this work indicated that a transfer reaction in inverse kinematics provides a unique tool for the study of the structures of the radioactive nuclei and that reliable spectroscopic factors must be extracted with a systematic and consistent approach using global optical-model potentials. To my wife - Yu Liang and my daughter - Liying Liu iii ACKNOWLEDGMENTS First of all, I would express my deepest gratitude to my advisor, Dr. Betty M. Tsang. I could not have completed this thesis without her tremendous support and guidance. She taught me how to solve somehow impossible problems by detailed inspections and unremitting effort. The experience of making progress through gloomy zenith has been wonderful and invaluable. I also learned much from her enthusiasm for life as well as for the nuclei. I own my sincere thanks to C. K. Gelbke, P. Hosmer, T. Liu, W. G. Lynch, R. Shomin, W. Tan, G. Verde, A. Wagner, M. Wallace, H. Xi and H. Xu. I am also indebted to L. Beaulieu, B. Davin, Y. Larochelle, T. Lefort, R. T. de Souza, R. Yanez and V. E. Viola from Indiana University, as well as R. J. Charity and L. G. Sobotka from Washington University. I would also express my gratitude to the REU students, K. Chalut, S. Farges, M. D. Lennek, M. Ramos, and S. Seun for their massive work. I would like to thank Dr. F. Nunes, Dr. M. Famiano, and Dr. F. Delaunay for their great help on theoretical calculations and discussions. I am deeply indebted to Dr. J. A. Tostevin who offered the TWOFNR program from Surrey University and gave us invaluable help on theoretical analyses. Great thanks to Dr. Daniel Bazin who taught me a lot about the S800 spectrometer and beam transmission. I would also thank Dr. Marc-Jan van Goethem and Michal Mocho for their help during my analyses. Special thanks go to Dr. William G. Lynch who always gave me excellent ideas and philosophical understanding of physics throughout my thesis experiment and data analyses. iv I would like to acknowledge the excellent support at the National Superconducting Cyclotron Laboratory at Michigan State University. I appreciate the help from J. Yurkon of the detector laboratory, L. Morris of the design group, J. Vincent of electronics group, R. Fox and B. Pollack of the computer group. I would also like to thank Dr. S. D. Mahanti, Dr. B. Sherrill, and Dr. C. P. Yuan for reading my dissertation and also serving on my guidance committee. Last but not least, I would like to thank my wife and my daughter for their love and belief in me. TABLE OF CONTENTS LIST OF TABLES ................................................................................ viii LIST OF FIGURES ............................................................................... x CHAPTER 1 INTRODUCTION ................................................................................. l 1.1 Motivation ................................................................................. 1 1.2 Inverse Kinematics ....................................................................... 3 CHAPTER 2 THEORETICAL DESCRIPTION ............................................................... 9 2.1 Overview ................................................................................. 9 2.2 Theoretical Spectroscopic Factor ..................................................... 10 2.3 Reaction Theory ......................................................................... 14 2.3.1 Distorted-Wave Born Approximation (DWBA) ............................. 14 2.3.1.1 Optical-Model Potential ............................................. 18 2.3.1.1.1 Overview .................................................. 18 23.1.1.2 Global Optical-Model Potentials ....................... 19 2.3.1.1.3 Proton Global Optical-Model Potentials . . . . 2O 2.3.1.1.4 Deuteron Global Optical-Model Potentials .......... 28 2.3. 1.1.5 Nucleon-Nucleus Optical-Model Potential .......... 35 23.1.1.6 JLM Optical-Model Potential .......................... 43 2.3.1.2 Zero-Range Approximation .......................................... 53 2.3.1.3 Finite-Range Approximation ....................................... 54 2.3.1.4 Non-Locality Correction ............................................ 58 2.3.1.5 Neutron Form Factor ................................................ 64 2.3.2 Adiabatic Deuteron Breakup Approximation (ADBA) ................... 70 2.4 Momentum Matching .................................................................. 75 2.5 Summary ................................................................................. 76 CHAPTER 3 EXPERIMENTAL SETUP AND CALIBRATION ......................................... 79 3.1 Overview ................................................................................ 79 3.2 MWDC Detectors ....................................................................... 84 3.2.1 Principle of MWDC Detector ................................................ 84 3.2.2 Position Calibration ............................................................ 86 3.3 LASSA Detector Array ................................................................ 94 3.3.1 Overview ........................................................................ 94 3.3.2 Geometric Setup ................................................................. 95 3.3.3 Silicon Strip Detector Array .................................................. 95 3.3.3.1 Overview ............................................................ 95 3.3.3.2 Energy Calibration .................................................. 97 vi 3.3.3.3 Particle Identification ................................................ 98 3.3.3.4 Position Calibration ................................................ 99 3.3.4 CSI(T1) Crystals ................................................................ 101 3.3.4.1 Detectors ............................................................ 101 3.3.4.2 Energy Calibration ................................................ 102 3.3.4.3 Particle Identification ............................................. 104 3.4 8800 Spectrometer ................................................................... 114 3.4.1 Overview ...................................................................... 114 3.4.2 Cathode Readout Drift Counters ........................................... 115 3.4.3 Ion Chamber .................................................................. 116 3.4.4 Plastic Scintillators .......................................................... 116 3.4.5 Summary ...................................................................... 117 3.5 Data Acquisition Electronics ....................................................... 124 CHAPTER 4 EXTRACTION OF AN GULAR DIFFERENTIAL CROSS SECTIONS ............... 126 4.1 Overview ............................................................................... 126 4.2 Deuteron Spectra ...................................................................... 126 4.3 Extraction of Angular Differential Cross Sections ............................... 128 CHAPTER 5 EXTRACTION OF SPECTROSCOPIC FACTORS ....................................... 141 5.1 Overview ................................................................................ 141 5.2 12C(d,p)l3C (g.s.) Reaction ........................................................... 142 5.3 l3C(p,ti)'2C (g.s.) and p(‘3C,d)‘-2C (g.s.) Reactions ................................ 154 5.4 13C(p,ti)‘2C (2") and p(l3C,d)12C (2*) Reactions .................................. 158 5.5 l"I3(ti,p)l ‘B (g.s.), “B(p,d)‘°B (g.s.), and p(‘ ‘B,d)‘°E (g.s.) Reactions 162 5.6 9Be(d,p)10Be (g.s.) and p(loBe,d)9Be (g.s.) Reactions ............................ 168 CHAPTER 6 SUMMARY ...................................................................................... 173 BIBLIOGRAPHY ............................................................................... 178 vii 2.1 2.2 2.3 2.4 2.5 2.6 2.7 3.1 3.2. 3.3 3.4 4.1 4.2 5.1 5.2 5.3 LIST OF TABLES An overview of the input parameters and options for TWOFNR ................ 13 The parameters of global nucleon potentials ......................................... 23 Deuteron global parameters. N=neutron number, Ezdeuteron laboratory energy in MeV. For Daehnick potential, ,6 = 4&6) 2 , 11,-: (54421132 , where M ,- = magic numbers (8,20,28,50,82,126) ................................... 30 Parameters for the global nucleon-nucleus optical potential of CH89 [Var91]... 38 Parameters for J LM potentials ......................................................... 48 The neutron potentials and the binding energies of the neutron form factors in the reactions in this experiment ..................................................... 65 Summary of the input parameters used in TWOFNR ............................... 78 The properties of the secondary beams and the targets ............................. 81 Geometric setup of the telescopes and the configuration of the silicon strip detectors ............................................................................ 105 The calibrated parameters in Equation 3.3.1 .......................................... 105 Characteristics of the S800 spectrometer ............................................ 120 Contributions to the energy resolution of the single neutron pickup reactions in inverse kinematics ......................................................... 135 Experimental angular differential cross sections and statistical errors .......... 136 List of references and Spectroscopic factors for the 12C(d,p)13C (g.s.) reactions ................................................................................... 147 List of references and spectroscopic factors for the ‘3C(p,d)12C (g.s.) and p(13C,d)12C (g.s.) reactions ........................................................ 157 Extracted spectroscopic factors of the lp3/2 neutron from the l3C(p,ti)‘2C (2”) and p('3C,d)'2C (2*) reactions .................................... 160 viii 5.4 5.5 Extracted spectroscopic factors of the 1pm neutron from the reactions of ”B(p,d) l"I3 (g.s.), ‘°B(d,p)“B (g.s.) and p(“B,d)‘°B (gs) Extracted spectroscopic factors of the 1pm neutron from the reactions of p(loBe,d)9Be (g.s.), and 9Be(d,p)]0Be (g.s.) ix 166 .171 1.1 1.2 1.3 2.1 2.2 2.3 2.4 2.5 2.6 LIST OF FIGURES Velocity diagrams for normal kinematics (a) and inverse kinematics (b) as in (p,d) reactions. V00" is the velocity of the center of mass in the laboratory frame; leab and Vd cm are the deuteron velocities in the laboratory frame and the center of mass, respectively; Blob and 66", are the emitted angles in the laboratory frame and the center of mass, respectively ............................................. 6 The deuteron emitted angles in center of mass vs. emitted angles in the laboratory frame. The solid line presents the inverse kinematic reaction of p(l3C,d)lzC (g.s.) at bombing energy of 47.9 MeV per nucleon; the dashed line stands for the reaction of l3C(p,d)12C (g.s.) at proton energy of 48.3 MeV ........................ 7 Kinematic broadening vs. angles in the laboratory frame. The solid line presents the inverse kinematic reaction of p(l3C,d)lzC at bombing energy of 47.9 MeV per nucleon; the dashed line stands for the reaction of l3C(p,d)12C at proton energy of 48.3 MeV . ................................................................................. 8 The target nucleus A is composed of the core nucleus B and one neutron n. The proton picks up a neutron to form the deuteron. O is the center of mass of nucleus B and 7,, points to the neutron; O’ is the center of mass of nucleus A; F and 7,, are the proton coordinates relative to the neutron and the center of mass of nucleus A, respectively. R is the coordinate of the deuteron center relative to nucleus B .............................................................. 17 Proton global optical-model potentials of 13C at incident energy of 12.5 MeV... 24 Calculations by different global optical-model potentials for the reaction of 13C(p,p)13C (g.s.) at incident energy of 12.5 MeV compared with the data [Wel78] ..................................................................................... 25 Proton global optical-model potentials of 13C at incident energy of 30.95 MeV ........................................................................................ 26 Calculations by different global optical-model potentials for the reaction of l3C(p,p)l3C (g.s.) at incident energy of 30.95 MeV compared with the data [Bar88] .................................................................................... 27 Deuteron global optical-model potentials of 12C at incident energy of 11.8 MeV ......................................................................................... 31 2.7 2.8 2.9 2.10 2.11 2.12 2.13 2.14 2.15 2.16 2.17 2.18 Calculations by different global optical-model potentials for the reaction of 12C(d, d)12C (g.s.) at incident energy of 11.8 MeV compared with the data [Fit67] ....................................................................................... 32 Deuteron global optical-model potentials of 12C at incident energy of 34.4 MeV ........................................................................................ 33 Calculations by different global optical-model potentials for the reaction of ”C(d, d)‘2C (g.s.) at incident energy of 34.4 MeV compared with the data [New67] ................................................................................... 34 Comparison of the CH89 proton potentials of 13C with the proton potentials of Menet and Perey & Perey at incident energy of 12.5 MeV ...................... 39 Calculations for proton elastic scattering on 13C at incident energy of 12.5 MeV using the potentials of Menet, Perey & Percy, and CH89 compared with the data [Wel78] ...................................................... 40 Comparison of the CH89 proton potentials on 13C with the proton potentials of Menet and Perey & Perey at incident energy of 30.95 MeV .................... 41 Calculations for proton elastic scattering on 13C at incident energy of 30.95 MeV using the potentials of Menet, Perey & Perey, and CH89 compared with the data [Bar88] ....................................................... 42 Comparison of the JLM proton potentials of 13C with the proton potentials of Menet and CH89 at incident energy of 12.5 MeV ............................... 49 Calculations for the l3C(p,p)‘3C (g.s.) reaction by JLM, Menet and CH89 potentials at incident energy of 12.5 MeV compared with the data [Wel78] ....... 50 Comparison of the JLM proton potentials of 13C with the proton potentials of Menet and CH89 at incident energy of 30.95 MeV ............................... 51 Calculations for the l3C(p,p)l3C (g.s.) reaction by JLM, Menet and CH89 potentials at incident energy of 30.95 MeV compared with the data [Bar88] 52 Calculations for reaction of l3C(p,d)12C (g.s.) using finite-range approximation (solid line) and zero-range approximation (dashed line) at incident energies of 15 MeV and 48.3 MeV. The cross sections at proton energy of 15 MeV have been multiplied by 10 so that the calculations at the two energies can be seen more clearly ............................................. 56 xi 2.19 2.20 2.21 2.22 2.23 2.24 2.25 2.26 2.27 2.28 2.29 3.1 Finite-range DWBA calculations for reaction of l3C(p,d)12C (g.s.) by TWOFNR (solid line) and DWUCKS (dashed line) at incident energies of 15 MeV and 48.3 MeV. The cross sections at proton energy of 15 MeV have been multiplied by 10 ............................................................. 57 Comparison of the local (dashed line) and non-local (solid line) proton potentials (CH89) of 13C at proton energy of 48.3 MeV ........................... 61 Calculated differential cross section for l3C(p,d)12C at proton energy of 48.3 MeV by non-local proton potential (solid line) increased 12% at the forward angles compared to that by local proton potential (dashed line) 62 Calculated differential cross sections for l3C(p,d)]2C at incident energy of 15 MeV using non-local proton potential (solid line) and local proton potential (dashed line) ............................................................................... 63 Neutron form factors for the reactions of (a) p(13C,d)12C (g.s.), (b) p('3C,d)‘2C (2"), (c) p(”B,d)'°B (g.s.), and (d) p(‘°Be,d)9Be (g.s.) 66 Calculations on the variation of neutron radius parameter r0, where the neutron diffuseness a0 is fixed to 0.65 fm and the spin-orbit strength Vso is 0.0 MeV. The cross sections at proton energy of 15 MeV are multiplied by 10 ............ 67 Calculations on the variation of neutron diffuseness do, where the neutron radius parameter r0 is fixed to 1.25 fin and the spin-orbit strength Vso is 0.0 MeV. The cross sections at proton energy of 15 MeV are multiplied by 10 ..... 68 Calculations on the variation of neutron spin-orbit strength V50, where the neutron radius To and diffuseness a0 are fixed to 1.25 fin and 0.65 fm separately. The cross sections at proton energy of 15 MeV are multiplied by 10 .......................... 69 Comparison of Daehnick global deuteron potential (dashed line) with the adiabatic deuteron potential (solid line) constructed by CH89 potentials, for 12C at Ed: 49.2 MeV ................................................................ 72 Comparison of ADBA (solid line) and DWBA (dashed line) calculations for reaction of l3C(p,d)12C at proton energy of 48.3 MeV. The ADBA increases the cross section at the forward angles and faster fall off than DWBA .......... 73 Comparison of ADBA (solid line) and DWBA (dashed line) calculations for reaction of l3C(p,d)12C at proton energy of 15 MeV ........................... 74 Schematic diagram of A1200 ........................................................... 82 xii 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 3.10 3.11 3.14 3.15 Schematic diagram of the facilities in 3800 vault .................................. 83 Schematic drawing of wire plane of MWDC detector. All the anode wires are connected to a micro-strip delay line, which has two timing outputs T1 and T2. Every other field wire is connected together to form two groups and gives the left—right ambiguity signals E1 and E2 respectively 88 An enlarged drawing of the ions drifting between the wires. The positive ions drift toward the field wires and the negative ions drift toward the anode wire. The term dt is the drift time of the negative ions from the hit point to the anode wire ................................................................ 89 Spectrum of the wire sequential number k. Each individual sharp peak indicates the wire with its sequential number ......................................... 90 Spectrum of drift time (it, where the time is scaled in units of microseconds ..... 91 Spectrum of E2 versus E1. There are two groups of particles. The left-right ambiguity can be clarified by which side yields larger signal than the other 92 (a) Patterns on the mask used to calibrate the MWDC detector. The distance between adjacent small holes is 2.54 mm; the distance between adjacent big holes is 10.2 mm; (b) The two-dimensional position Spectrum of the mask. The corresponding areas are surrounded by the dotted rectangle in (a) and (b) ...................................................................................... 93 Structure of LASSA telescope . ...................................................... 106 Schematic of the geometric setup .................................................... 107 One double-sided silicon strip detector with the flat printed circuit board cables ..................................................................................... 108 Calibration curve for silicon strip detector, by which the channel readout of the silicon strip detector is converted to particle energy in units of MeV. This curve stands for the No. 6 strip of EF detector in telescope 3 108 Particle identification using the energy spectrum of BBB vs. E51: for telescope 7 ............................................................................... 109 Geometry calibration system composed of one optical telescope and a mirror mounted on a turntable with two orthogonal axes that rotate in horizontal and vertical planes. The center of the mirror is the position of the target in the experiment and the optical telescope is mounted in the beam line ............... 110 The shape of CsI(Tl) crystal ........................................................... 110 xiii 3.16 3.17 3.18 3.19 3.20 3.21 3.22 3.23 4.1 4.2 4.3 4.4 Schematic of the CsI calibration. The deuteron emitted angle is determined by the pixel on DE and EF(EB) silicon strip detectors. The deuteron emitted energy is obtained by kinematic calculation. The deuteron deposited its energy into the CsI crystal after going through target, window foil, DE and EF Silicon strip detectors .. ......................................................... 111 Calibration for CsI detector, by which the channel readout of the CsI Detector is converted to particle energy in units of MeV. This figure shows the calibration for the No. 3 crystal in telescope 3 .......................... 112 Particle identification in the energy Spectrum of 4E DE + E EF vs. Emm, for telescope 3 ........................................................................... 113 Schematic of the focal plane detector of 8800 spectrometer. It consists of two CRDC detectors, one ion chamber, and four plastic scintillators ........ 119 Schematic of the CRDC detector. A particle ionizes the gas as it passes through the detector. The electrons drift to the anode wire where they are collected. The induced image charges on the cathode pads provide horizontal position information. The drift time of the electrons to the anode wire provide vertical position information ................................... 121 (a) Patterns on the mask. (b) Position spectrum of the mask placed in front of the first CRDC detector ...................................................... 122 Spectrum of the energy deposited in E1 scintillator versus the time-of-flight for the p(1 lB,d)lOB reaction ........................................... 123 Schematic of the electronics .......................................................... 125 Deuteron energy spectrum of the p(13C,d)12C reaction at the laboratory angle of 19° measured by telescope 7 ................................................ 132 Deuteron energy spectrum of p(1 iB,d)10B reaction at the laboratory angle of 13° measured by telescope 7 ................................................. 133 Deuteron energy spectrum of p(loBe,d)9Be reaction at the laboratory angle of 15° measured by telescope 7 ................................................ 134 The measured angular differential cross section of p(13C,d)12C (g.s.) (open circle) reaction compared to the published data of Ref. [Cam87] (closed circle) and Ref. [Sco70] (diamonds) .......................................... 137 xiv 4.5 4.6 4.7 5.1 5.2 5.3 5.4 5.5 5.6 5.7 The measured angular differential cross section of p(13C,d)12C (2+) (open circle) reaction compared to the published data of Ref. [Cam87] (closed circle) and Ref. [Sco70] (diamonds) ........................................ 138 The measured angular differential cross section of p(1 lB,d)loB (g.s.) reaction ................................................................................... 139 The measured angular differential cross section of p(loBe,d)9Be (g.s.) reaction .................................................................................. 140 Spectroscopic factors for 12C(d,p)l3C (g.s.) and l3C(p,d)]2C (g.s.) reactions extracted from the literatures (see Table 5.1 and Table 5.2) ..................... 148 Angular distributions for 12C(d,p)l3C (g.s.) reactions for beam energy from 7 to 56 MeV: solid lines present ADBA (JLM); dotted lines present ADBA (CH89); dashed lines present DWBA. Each distribution is displaced by factor of 10 from adjacent distributions. The overall normalization factor is 1 for the 19.6 MeV data. References are listed in Table 5.1 .............................................. 149 Extracted spectroscopic factors in the present work for 12C(d,p)l3C (g.s.), l3C(p,d)12C (g.s.), and p(l3C,d)'2C (g.s.) reactions. The dashed lines represent the shell model prediction of Cohen and Kurath [Coh67] of 0.62. See text for detail explanation ........................................................................ 150 Comparison of the existing measurements of 12C(d,p)13C (g.s.) reaction for deuteron energy at 4.5 MeV, a [Gur69], b [Gal66], and c [Bon56] .............. 151 Comparison of the existing measurements of 12C(d,p)l3C (g.s.) reaction for deuteron energy at 11.8 MeV [Sch64], 12.0 MeV a [Lan88], 12.0 MeV b [Sch67], and 12.4 MeV [Ham6l] ..................................................... 152 Comparison of the existing measurements of 12C(d,p)]3C (g.s.) reaction for deuteron energy at 14.7MeV [Ham61], 14.8MeV [MchS], and 15 MeV [Dar73] ................................................................................... 153 Angular distributions for l3C(p,d)12C (g.s.) and p(l3C,d)12C (g.s.) reactions for beam energies from 35 to 65 MeV: solid lines present ADBA (JLM); dotted lines present ADBA (CH89); dashed lines present DWBA. The calculations have been normalized by the spectroscopic factors. Each distribution is displaced by factor of 10 from adjacent distributions. The overall normalization factor is 1 for the 65.0 MeV data ........................................................................... 156 XV 5.8 5.9 5.10 5.11 5.12 5.13 5.14 6.1 Calculations for reactions of l3C(p,d)12C* (4.439MeV) and p(13C,d)]2C* (4.439MeV) [Toy95, Cam87, Tak68]; solid lines present ADBA (JLM) potentials; dotted lines present ADBA (CH89). The calculations have been normalized by the spectroscopic factors. Each distribution is displaced by factor of 10 from adjacent distributions. The overall normalization factor is 1 for the 55.0 MeV data ...... 159 Extracted spectroscopic factors of the reactions 13C(p,d)12C* (4.439MeV) (circle) and p(13C,d)lzC* (4.439MeV) (square). The dashed lines represent the shell model prediction of 1.12 by Cohen and Kurath [Coh67] ............... 161 Calculations of 10B(d,p)”B (g.s.) reaction at 10.1MeV [Hin62], 12.0 MeV [Sch67], 13.5 MeV [Bar65], and 28 MeV [Slo62]: solid lines present ADBA (JLM); dotted lines present ADBA (CH89). The calculations have been normalized by the Spectroscopic factors. Each distribution is displaced by factor of 10 from adjacent distributions. The overall normalization factor is 1 for the 28.0 MeV data ......................................................... 164 Calculations of llB(p,d)10B (g.s.) (closed points) p(l lB,d)mB (g.s.) (open points): solid lines present ADBA (JLM); dotted lines present ADBA (CH89). The calculations have been normalized by the spectroscopic factors. Each distribution is displaced by factor of 10 from adjacent distributions. The overall normalization factor is 1 for the present data ............................... 165 Extracted spectroscopic factors for the reaction of p(l lB,d)'0B (g.s.) (open square), llB(p,d)mB (g.s.) (open circle), and l0B(d,p)l lB (g.s.) (closed circle). The dashed lines represent the shell model prediction of 1.09 by Cohen and Kurath [Coh67] .......................................................................... 167 ADBA calculations based on CH89 potentials for the reactions of 9Be(d,p)lOBe (g.s.) (closed circle) [Ze101, Gen00, Va387, Sch64, Dar76, And74, 81062] and p(lOBe,d)9Be (g.s.) (open circle). The calculations have been normalized by the spectroscopic factors. Each distribution in closed circle is displaced by factor of 10 from adjacent distributions. The overall normalization factor is 1 for the data at 11.0 MeV. The present data is reduced by a factor of 100 ............................................................. 170 Extracted spectroscopic factors for the reactions of p(10Be,d)9Be (g.s.) (open square), and 9Be(d,p)'oBe (g.s.) (closed circle). The dashed lines represent the shell model prediction of 2.35 by Cohen and Kurath [Coh67] .............. 172 Comparison of the extracted spectroscopic factors with the predictions of the modern shell model [BroO4] for 79 nuclei ranging from Li to Cr [Tsa05]. Good agreement with most isotopes except Ne, F, and Ti isotopes .............. 177 xvi CHAPTER 1 INTRODUCTION 1 . 1 Motivation The Study of nuclei far away from stability has been the focus of nuclear study in recent years. We want to know the extent to which the nuclear shell model theory is valid for nuclei beyond the stability limits. Such understanding is especially important since unstable nuclei are essential components in the nuclear synthesis process. Since the discovery of the shell model, which explains many structural properties of the nuclei, transfer reactions have been used to study the configuration of the valence nucleons. Spectroscopic factors (SF) are important quantities that tell us the structure of the single nucleon orbit. In this work, we define the Spectroscopic factor as the ratio of the experimental cross section from the transfer reaction to the theoretical calculation based on a reaction model that assumes the orbital fully occupied by the transferred nucleon. Since unstable nuclei cannot be made into targets, the transfer reactions must be performed in inverse kinematics using rare isotope beams. Currently there are unanswered questions in the extraction of spectroscopic factors. In the reaction theory, which uses the Distorted-Wave Born Approximation, DWBA, a fast one-step direct process of less than 10'22 sec is assumed. Elastic scatterings are used to describe both the entrance and the exit channels. It is usually believed that an accurate optical-model potential, which is derived from the best fitting of the elastic scattering data, would give the correct incoming and outgoing wave functions and hence the correct extraction of the spectroscopic factor. Unfortunately, such practice has failed to provide a consistent extraction of the spectrosc0pic factors in part due to the ambiguity in the parameters needed to describe the optical-model potential. In contrast, there are also arguments that superior results would be obtained if global optical-model potentials that describe a range of nuclei and incident energies are used instead. There are statements in literature that such an average optical-model potential tends to give more reasonable Spectroscopic factors than individual potential [Sch67]. However, such statements have not been well quantified. One purpose of this thesis is to compare the different strategies and find a reliable method to extract consistent spectroscopic factors. Another goal of this work is to study the structure of deformed unstable nuclei such as 10Be via the (p,d) reaction. It was envisioned that this would become the starting point of a series of studies of the N=6 isotones in inverse kinematics. The valence neutron of 10 . . . 9 10 . Be had been prevrously studied vra the Be(d,p) Be reaction. The extracted spectroscopic factors, however varied from 0.97 to 2.07, in some cases differing from the theoretical value of 2.35 based on the shell model. We want to know if there is new physics in 10Be that makes it different from the shell model expectation. Furthermore, understanding the structure of 10Be may help us to understand the structure of more neutron-rich isotopes of Beryllium such as 11Be. This is the first time that the secondary radioactive beam of l0Be was used to perform the (p,d) reaction in inverse kinematics. For these experiments, we used a high-angular and high—energy resolution detector, Large Area Silicon Strip Array (LASSA), to detect the deuterons emitted in the reactions. In addition, we also measured (p,d) reactions on another N=6 isotone llB as well as '3C. This letter reaction was used to obtain energy calibrations. High quality data were obtained in this reaction and used in our systematic studies to find a strategy to extract the spectroscopic factor. Various properties of the inverse kinematic reaction are analyzed in the next section. The theoretical background is presented in chapter 2. This latter chapter includes the description of the theoretical spectroscopic factor (Section 2.2) and of the reaction theories (Section 2.3). Specifically, the theories of distorted-wave Born approximation (DWBA) and adiabatic deuteron breakup approximation (ADBA) are introduced in Section 2.3.1 and Section 2.3.2 respectively. Detailed descriptions of the experimental setup are provided in chapter 3, which includes the descriptions of the various detectors and electronics. Chapter 4 describes the analyses of the deuteron spectra and the extraction of the angular differential cross sections. Theoretical calculations and the extraction of the spectroscopic factors are explained in chapter 5. Chapter 6 gives the summary of this thesis. 1.2 Inverse Kinematics Nuclear reactions involving nucleon transfer between stable beams and target nuclei have been a very useful source of nuclear structure information, and many theoretical tools have been developed to extract spectroscopic information. However, for the Study of radioactive nuclei far from the stable region, which has become the new focus of studies in nuclear astronomy and nuclear structure beyond the shell model in recent years, inverse kinematics becomes necessary since the radioactive nuclear targets, especially those with a very short half-life, are usually not available. Thus, transfer reactions induced by radioactive beams on proton and deuteron targets have great potential for probing single-particle structures in new regions [For99, Win01, Reh98, Oga99]. One advantage of the inverse kinematic reaction is that it is relatively easy to cover the forward scattering angle in the center of mass. In the normal kinematic reaction, where the light projectile bombards the heavy target, the small scattering angle in the center of mass can only be covered at the most forward angle in the laboratory frame. Figure 1.1(a) shows the velocity diagram for the (p,d) reaction in normal kinematics, cm lab where V0 is the velocity of the center of mass in the laboratory frame; Vd and Vdcm are the deuteron velocity in the laboratory frame and in the center of mass; 61a}, and 66m are deuteron emitted angles in the laboratory and in the center of mass. In the inverse kinematic reaction, as shown in Figure 1.1(b), the deuteron scatters backward in the center of mass. Smaller (96m can be obtained at relatively large 610,). Figure 1.2 shows the relations between the deuteron emitted angles in the laboratory frame and the emitted angles in the center of mass for the reactions of p(13C,d)12C g.s. (solid line) and 13C(p,d)12C g.s. (dashed line) at the equivalent bombing energy respectively. The detector in the inverse kinematic reaction covers smaller angles in the center of mass than that in a normal kinematic reaction at the same laboratory angle. One disadvantage of inverse kinematics is the kinematic broadening. Figure 1.3 shows the kinematic broadening vs. the emitting angle in the laboratory frame. The solid line presents the inverse kinematic reaction of p(]3C,d)12C at bombing energy of 47.9 MeV per nucleon; the dashed line stands for the 13C(p,d)]2C reaction at a proton energy of 48.3 MeV. Except for the very forward angles, the kinematic broadening is much more severe for the inverse kinematic reaction than the normal kinematic reaction. For example, the kinematic broadening increases dramatically from 650 keV at 30 degree to 1.27 MeV at 35 degree Therefore in this reaction, deuterons should not be detected beyond 35 degree in the laboratory frame. For the forward angles, detectors with high angular resolution as well as high energy resolution are required. Silicon strip detectors are widely employed to achieve high angular resolution and high energy resolution. The techniques of using these detectors will be discussed in Section 3.3. I W \ I i ‘ \ ’ s I x ’ \ r ’ lab ‘3 .’ Vd 6 \ : lab V cm \ I r I g Beam t cm I \‘ V0 8 cm ’r \ I \ I \ l x l/ \\ I, ‘ r s ‘ I \ ‘ ~ 4.. ’ a ’ ( a ) 2” “\ I ‘ \ ’ \ I \ I \ ’ \ ’ \ lab ’ cm \ Vd , Vd r‘ I I L ' i | r I I I Beam 9 t‘ V cm 9 i lab r 0 cm I .. I \ I \ l x l/ x \ I x I ’ \ ‘ ’ \ I (b) Figure 1.1: Velocity diagrams for normal kinematics (a) and inverse kinematics (b) as in (p,d) reactions. VOC'" is the velocity of the center of mass in the laboratory frame; leab and Vd cm are the deuteron velocities in the laboratory frame and the center of mass, respectively; 61a}, and 6m are the emitted angles in the laboratory frame and the center of mass, respectively. 60 9am (deg) \ \ 20— / 0 10 20 30 40 elab (deg) Figure 1.2: The deuteron emitted angles in the center of mass vs. emitted angles in the laboratory frame. The solid line presents the inverse kinematic reaction of p(l3C,d)12C (g.s.) at bombing energy of 47.9 MeV per nucleon; the dashed line stands for the reaction of l3C(p,d)12C (g.s.) at proton energy of 48.3 MeV. Kinematic broadening (MeV/deg ) Figure 1.3: Kinematic broadening vs. angles in the laboratory frame. The solid line presents the inverse kinematic reaction of p(l3C,d)12C at bombing energy of 47.9 MeV per nucleon; the dashed line stands for the reaction of l3C(p,d)12C at proton energy of 48.3 MeV. CHAPTER 2 THEORETICAL DESCRIPTION 2.1 Overview The main goal of this thesis is the extraction of the neutron spectroscopic factors from measurements. The experimental spectroscopic factor is defined as the ratio of the experimental differential cross section to the calculated differential cross section based on a reaction model that assumes the relevant orbit is fully occupied. The extraction of the experimental differential cross sections measured in this thesis will be described in Chapter 4. This chapter mainly describes how the theoretical differential cross sections are calculated. The theoretical nucleon spectroscopic factor will be introduced in Section 2.2. The most widely used models, the distorted—wave Born approximation (DWBA) and the adiabatic deuteron breakup approximation (ADBA), will be discussed in Section 2.3. The effects of different input parameters including the choices of optical-model potentials in DWBA will be covered in the subsections of Section 2.3. At the end of this chapter, a list of standard input parameters for the DWBA and ADBA calculations will be proposed. 2.2 Theoretical Spectroscopic Factor In the theory of the shell model, the valence nucleon in the nucleus is described as a single-particle state of a particular orbit. Each orbit is assigned the number n, l, and j corresponding to the node number, the orbital momentum, and the total spin momentum of the nucleon. The assumption that the nucleon occupies a pure single-particle state is an idealization, which is true only in few cases in real nuclei. Due to the interactions among other nucleons, each nucleon may occupy several single-particle states. The occupation of a nucleon in a pure single-particle state is called the spectroscopic factor, which contains the information of the nuclear structure and how well the shell model theory describes the real nuclei. Thus the nucleon Spectroscopic factor is among the most fundamental tests of shell model theory [Ban85]. In a nucleus composed of A nucleons, the spectroscopic factor can be deduced from the expansion of the wave function I// (A) in terms of a summation over the complete set of single-particle states ¢,,,j(f,,) and the states III (B) of the residual core nucleus B is composed of A-l nucleons [Gle04]: wfij‘tam = Bzflflnfl (A. B')A[¢nlj(7n)lfl13'(3)] 31" (2.2.1) where A is an antisymmetrization operator, ,BnIJ-(A,B') are coefficients of fractional parentage and their values depend on the detailed structure of the nuclear wave function. 10 The square bracket denotes vector coupling: [Q10 (mil/13' (3) MA = Z CJB'jJA ¢m - M' M'mMA nlj(r") WJB'(B) (2'22) Mm The spectroscopic factor for a specific single particle state (n1 j) is: 5",,- = fl,,,,-2(A,B') (2.2.3) For the pickup (p,d) reaction, the spectroscopic factor is related to the experimental angular differential cross section and the theoretical calculation by: “(6') = "U (“a”) (2.2.4) d9 d9 theory do(6) . . . . where rs calculated assuming the neutron m the exact state (121]). The theory theoretical calculations are performed either in the distorted-wave Born approximation (DWBA) theory or in the adiabatic deuteron breakup approximation (ADBA) theory. The introduction of the theories will be in Section 2.3. In this work, all the theoretical calculations for (p,d) and (d,p) reactions are performed using the code TWOFNR, which was initially developed by M. Igarashi in 1977 [Iga77]. This code is relatively easy to use since it supplies multiple options with default values for every step and component in the calculations. All the inputs, including the parameters and option choices, are converted automatically into a standard input file for TWOFNR by a partner program FRONT. The input parameters and options are listed in Table 2.1. For example, the integration ranges and the number of partial waves can be specified by the user or the default values can be adopted. The user can choose the global optical-model potentials for proton and deuteron or specify the parameters for individual potential. When the ADBA theory is employed, the Johnson-Soper adiabatic potential for deuteron can be constructed using three 11 different nucleon-nucleus potentials. For the application of JLM potential, users can input their own parameters for the target density and potential scaling factors following the prompts of the program. There is a switch either to zero-range approximation or to finite- range approximation. If the finite-range approximation is chosen, the finite-range factor can be the default value or be specified. The same strategy is also applied to the options of neutron binding potential, the vertex constant, and non—locality correction. We choose to use TWOFNR because there are many default options available and it is easier to perform many calculations in a systematic study. Another popular finite-range DWBA code is DWUCKS [Kunz], which performs finite-range calculations with deuteron wave function instead of the finite-range approximation in TWOFNR. We compared them in Section 2.3.1.3 and found that the results from these two programs are very close to each other for the same input parameters (see Figure 2.19). We believe essentially the same results will be calculated if different codes than TWOFNR are used. 12 Table 2.1: An overview of the input parameters and options for TWOFNR Integration ranges Specified or default value (30 fm in 300 steps) Number of partial waves Specified or default value (70) Proton potential Choose built-in options of global optical-model potentials: Bechetti-Greenlees; Chapel-Hill 89 (CH89); Perey & Perey; Menet; JLM; Or Specified parameters for V,, r,, av, WV, W5, rw, aw, V50, r50, ago, and RC ’ Deuteron potential Choose built-in options of global optical-model potentials: Lohr-Haeberli; Perey & Perey; Daehnick; Johnson-Soper adiabatic (ADBA); Or specified parameters for V,, r,, av, WV, W5, rw, aw, V50, r30, ago, and RC Johnson-Soper adiabatic potential Choose built-in options of global optical-model potentials: Bechetti-Greenlees; Chapel-Hill 89 (CH89); JLM; Choose built-in options: Target density Negele form; for JLM potential Specify rrns radius; Modified Harmonic oscillator form; JLM potential scaling A Specified or default values (Av=1.0, Aw=0.8) Neutron binding potential Specified or default values (r0=1.25 fm, ao=0.65 fm, VSO=6 MeV) Zero-range approximation Use or not Finite-range approximation Use or not If use, finite range factor can be specified or choose default value of 0.7457 fm Vertex constant D02 Specified or default value of 15006.25 MeV2 - fm 3 Non-locality correction Use or not. If use, non-locality range can be specified or choose default value (0.85 fm for proton; 0.54 frn for deuteron). l3 2.3 Reaction Theory 2.3.1 Distorted-Wave Born Approximation (DWBA) Transfer reactions have been an important tool in the study of nuclear structure. The results obtained from the studies of the pickup A(p,d)B and stripping B(d,p)A reactions, involving single neutron transfer, help to validate the nuclear shell model by identifying the single-particle states. To a large extent, the (p,d) reaction can be understood as one in which the neutron is removed from a single particle state of the target nucleus A. In the (d,p) reaction the neutron in the deuteron is deposited to a sin gle-particle state of the final nucleus A. Thus, the theoretical description of the (d,p) reaction is similar to that of the (p,d) reaction. In the pickup reactions, A(p,d)B, where A=B+n, a neutron in a single-particle state in A, is picked up by the incident proton to form the deuteron. The process is illustrated in Figure 2.1. The transition amplitude for this reaction under the distorted-wave Born approximation (DWBA) theory is written as [Gle04]: T(#pJAMA_)#dJBMB):ZCJBjJ/I C(1/2)(1/2)(1)Cl(1/2)j 1' MijMA Hunt/Jr! mlpmnj 1 (2.3.1) l n]; 14 where Bml(kp,kd)=i-l(21+l) —((-l/2)J‘Z) *(kd RWMIU'; ) .(v +Va—U mam-1‘ ’(k .r )dFd? W! p p p p p n (232) 51/2 A _ _ Snlj :JIW: *(A)A[WJB(B)¢an(;}190 531).“ dBdrn where A and B refer to the nucleon coordinates and Spins of nucleus A and B; 7,, and ’17 are the coordinates of neutron and proton; 7,, and R are the relative and center-of- mass coordinates of the deuteron; 1P(+)(i€p,Fp) is the distorted-wave function describing the elastic scattering of the incoming proton by the proton optical-model potential U I); the distorted-wave function zd(-)*(l;d,R) describes the elastic scattering of the emitted deuteron by the deuteron optical-model potential U d? i/I(A) is the wave function of the target nucleus A; (11(3) is the wave function of the core nucleus B; and ¢d (f) is the internal wave function of the deuteron. The term VP" + VpB — U p is called the residual interaction, where V is the interaction between the proton and neutron, pit and VpB is the interaction between the proton and the remaining B nucleus. The term ¢nlj(fn ,6") is the neutron wave function in Specific single-particle state (n1 3' ), which is also called the neutron form factor: 61"" (7n 5n)= [(0,210 )Xt/2(0,.)l"' (2.3.3) nlj raj/(r,)=u,,,(r,,>Y;”’(r,,) (2.3.4) 15 Assuming VpB and U p approximately cancel each other in Equation 2.3.2 [Aus70, Sat71], the 8;", (12p , 12d ) becomes 3;"‘(i'r‘p kd)=i_l(21+l)_(1/2)JZ()*(kd ,R)¢""(r) ‘ (+) __ (2.3.5) Vpn'¢d(r)'Zp (kp’rp)d?nd;p The cross section is do mmd p de —(fl JAMA *fldJBMB)= (2.3.6) where rm; and mp are the reduced masses. In order to obtain the distorted-wave functions of 2’de and 1p”) , we need the optical-model potentials for the deuteron and proton. In the next section, the choices and detailed descriptions about the optical-model potentials will be discussed. 16 Figure 2.1. The target nucleus A is composed of the core nucleus B and one neutron n. The proton picks up a neutron to form the deuteron. O is the center of mass of nucleus B and 7,, points to the neutron; O’ is the center of mass of nucleus A; F and 7,, are the proton coordinates relative to the neutron and the center of mass of nucleus A, respectively. R is the coordinate of the deuteron center relative to nucleus B. 17 2.3. 1. 1 Optical-Model Potentials 2.3.1.1.1 Overview The Schrodinger equation of the collision system of a + b can be written as: (H - E)‘I’ = 0 (2.3.7) where H includes the intrinsic energy HO, the kinetic energy T, and the potential U between a and b: H =H0+T+U (2.3.8) The Schrodinger equation is separable into the nuclear intrinsic coordinates and relative coordinates so that the solution w can be written as a product of the nuclear intrinsic wave function WaWb and a relative wave function ¢(f), which satisfies the optical- model Schrodinger equation: (T+U —E)¢(F) =0 (2.3.9) Since U depends only on the relative coordinates of the two nuclei, it produces no change in the nuclei and describes only the elastic scattering. As the nucleon force is Short- ranged, and since the density p of nucleons in the nucleus is fairly constant in the interior and falls smoothly to zero at the nuclear surface, it is reasonable to assume that U has a radial shape that is similar to the density. Usually the optical potential is expressed in the Woods-Saxon form [WooS4]: l8 U(r) = —va(r,Rv,av)—iWVf(r,RW,aW) +4iWsaW—al-f(r,Rw,aw) d’ 1 d (2.3.10) + 2.0(v50 +iwm)(—r——d—r f(r,R,0,a,0)Z-a) + Vc where the Woods-Saxon shape function f (r, Rk ,ak) is : 1 l r,R, = , R = -A‘ 2.3.11 fl kak) l+exp[(r—Rk)/ak] k r" 3 ( ) Here rk is the radius parameter and ak is the diffuseness parameter; V, and W, are the depths of the real and imaginary potentials, respectively; Ws is the depth of surface term of the imaginary potential. Vso and Wso are the depths of the real and imaginary parts of the spin-orbit potentials; Z is the orbital angular momentum of the relative motion of the scattered particle; and 6' is the spin operator. VC is the Coulomb interaction, which is taken for a uniformly charged sphere of radius RC with different expressions inside and outside the radius RC . 2 Z Z e a b , r>Rc r V = 2.3.12 C i ZaZbez r2 ( ) 3— ), rSRc 2R6 R02 2.3. 1. 1.2 Global Optical-Model Potentials In principle, all the parameters of the optical-model potential can be obtained by fitting them to the experimental data of the elastic scattering. For the best fit to individual 19 nucleus at specific energy, all parameters could be optimized. However, the parameters of the optical potential usually vary smoothly with energy and are similar for neighboring nuclei. Thus global optical potentials could be obtained by fitting a group of nuclei with a total of N points in a certain energy range by minimizing 12 : (2.3.13) 2 12 : N[ath(6i)-Uexp(6i)] l N E AUCXP(6i) where 0' and 0' are the calculated and ex erimental values of the cross sections at an th exp P angle of 6,- and Aoexpis taken to be the experimental error; N is the number of data points. Over the years, many global optical potentials have been develOped for both protons and deuterons. In the following sections, we discuss mainly those potentials which have been provided as options for TWOFNR [Iga77] that we have used to calculate the theoretical differential cross sections. 2.3.1.1.3 Proton Global Optical-Model Potentials In this section, we introduce three sets of proton global optical-model potentials developed by Bechetti-Greenlees [Bec69], Menet [Men71], and Percy & Perey [Per76]. The Becchetti-Greenlees [Bec69] global proton potential has been developed for A>40 nuclei and proton energies up to 50 MeV. Menet [Men7l] developed a global proton potential for 1240 12 _4() — /.,~’/ —: Menet : xc/ —————— : Perey8cPerey — ' l ------------ : Bechetti—Greenlees —60 . I 1 L g : I _ Cl) , -. E —10 f; ........ ,x i: —20 E g : _3OP..II.......IILL 0 2 4 6 8 Figure 2.2: Proton global optical-model potentials of 13C at incident energy of 12.5 MeV. 24 105 ' ' ' 4 i ' v r r I r r . 4 13 m C(p,p) C (g.s.) Ep=12.5MeV 104— . ————:Menet A 3 ------ : Perey8cPerey L; 10 ’ """""" :BechettI—Greenlees 4 to .. \ ..Q g 102— C2 33 10 b "O 1.0- 2 10_1‘***#‘AL.1...L 0 50 100 150 9am.(deg) Figure 2.3: Calculations by different global optical-model potentials for the reaction of 13C(p,p)l3C (g.s.) at incident energy of 12.5 MeV compared with the data [Wel78]. 25 O_ /..- t // —10} ’9 i CD —20_— 5 . ”a -803- .. > : / (fr —: Menet —40 E: T ///' — — —: Perey8cPerey / ------- : Bechetti—Greenlees ' 1 . . a .L 9 <1) 5 3 —20,— 3 i _3OF....1E...I....I.... O 2 4 6 8 r (fm) Figure 2.4: Proton global optical-model potentials of 13C at incident energy of 30.95 MeV. 26 5 10....T..rtl 13C(p,p)13C (g.s.) Ep=30.95MeV 104— r —— : Menet A 3 —————— : Perey&Perey $53. 10 ~ ---------- :BechettI-Greenlees \ "Q 2 g 10 — C2 3 10 b "C 1.0 r 10‘1 r . r r I r . L L l . r I r \‘r 0 50 100 150 9am.(deg) Figure 2.5: Calculations by different global optical-model potentials for the reaction of l3C(p,p)l3C (g.s.) at incident energy of 30.95 MeV compared with the data [Bar88]. 27 2.3. 1.1.4 Deuteron Global Optical—Model Potentials There are three widely-used deuteron global optical-model potentials: Lohr-Haeberli [Loh74], Perey & Perey [Per76], and Daehnick [Dae80]. These three potentials are available as options in the code of TWOFNR [Iga77]. The Lohr-Haeberli deuteron global potential is for nuclei with A>40 and for deuteron energies from 8 MeV to 13 MeV; the Perey & Percy deuteron global potential is for nuclei with 2212 and deuteron energies from 12 MeV to 25 MeV; the Daehnick deuteron potential spans the energy range from 11.8 MeV to 90 MeV and includes nuclei ranging in mass from 27Al to 238Th. The parameters of the above three global optical-model potentials are listed in Table 2.3. As a comparison, Figure 2.6 Shows the three global deuteron potentials of 12C at incident deuteron energy of 11.8 MeV. Figure 2.7 shows the elastic scattering calculations based on the above global potentials, together with the experimental data at incident energy of 11.8 MeV [Fit67]. Unlike Figure 2.3 and Figure 2.5, Figure 2.6 plots the ratios of scattering differential cross section divided by the Rutherford differential cross section. The calculations agree with each other at the forward angles (less than 25°), but there exist slight deviations from the data. Figure 2.8 shows the three global deuteron potentials on 12C at incident deuteron energy of 34.4 MeV. Figure 2.9 shows the elastic scattering calculations with the experimental data at 34.4 MeV [New67]. The calculations agree with each other within the standard error of 5.3% at the forward angles (less than 20°), and they fit the elastic scattering well up to 36° in the center of mass. 28 Based on the above comparisons, we see that the present deuteron global optical- model potentials describe the deuteron elastic scattering better at higher energy than at lower energy and at smaller scattering angles better than at larger scattering angles. At higher energy, The Daehnick potential gives better fitting than others, so we choose Daehnick deuteron potential in our DWBA analyses. For the ADBA, we use the adiabatic deuteron potential that will be introduced in Section 2.3.2. 29 Table 2.3: Deuteron global parameters. energy in MeV. For Daehnick potential, ,6 = —(I§0) 2, fl, =(M—i2fl) 2, where M i: magic numbers (8,20,28,50,82,126). =neutron number, E=deuteron laboratory Potentials Lohr-Haeberli Perey & Perey Daehnick (deuteron) (deuteron) (deuteron) Parameters [Loh74] [Per76] [Dae80] A A>40 2212 27 I // —— : Lohr—Haeberli —80 777/ ------ : Perey8cPerey :/ ............. I Daehnick —100 _ % 20 : 5 : 3 —-40 i 3 i —60 h A I E . l . l A O 2 4 6 8 r (fm) Figure 2.6: Deuteron global optical-model potentials of 12C at incident deuteron energy of 11.8 MeV. 31 (dU/dOI/UknkflDnnh 10 'I f f r T I T I 1 I I T Y I I I Y ‘1' I I I I T 4 12C(d,d)12C (g.s.) Ed=11.8MeV : Lohr—Haeberli 3 """ : Perey8cPerey """""" : Daehnick 10—2 .1.n..1....1.1111. O 20 4O 6O 80 9am.(degl Figure 2.7: Calculations by different global optical-model potentials for the reaction of ”C(d, d)12C (g.s.) at incident energy of 11.8 MeV compared with the data [Fit67]. 32 O . E‘ Q) 5 E :> -—-— : Lohr—Haeberli - - — : Perey8cPerey . -------- : Daehnick — 100 _ IT: 20 Z 5 : 3 —40 I 3 i _60' A 1 . 1 . In L 0 2 4 6 8 T (im) Figure 2.8: Deuteron global optical-model potentials of '2C at incident deuteron energy of 34.4 MeV. 33 (dU/dQ)/(dU/dQ)Ruth |._s O H O I...r[....rrTrrf. 12C(d,d)12C (g.s.) Ed=84.4MeV 1 ti. l! : Lohr-Haeberli lj' " — —: Perey8cPerey ' """" : Daehnick l L L .l l l 11 1 L L l A g A l l 20 40 60 80 9am. (deg) Figure 2.9: Calculations by different global optical-model potentials for the reaction of 12C(d, d)12C (g.s.) at incident energy of 34.4 MeV compared with the data [New67]. 34 2.3. 1.1.5 Nucleon-Nucleus Optical-Model Potential The global optical-model potentials discussed above are derived from the fitting to the elastic scattering data in particular mass and energy regions. One consequence is that the derived global optical-model potentials cannot cover all the nuclei over a wide energy region. Thus, derivation of an optical-model potential using a much more extensive database of elastic scattering than previously used is desirable. A parameterization of the nucleon-nucleus optical-model potential based on data for A from 40 to 209, proton energies from 16 to 65 MeV and neutron energies from 10 to 26 MeV, was developed by R. L. Vamer et. a1. [Var91]. This parameterization, which is called Chapel-Hill 89 (CH89), is based on the current understanding of the basis of the optical potential, such as the folding model and nuclear matter approaches instead of the determination of optical- model potentials phenomenologically. The extensive database includes nearly 300 angular distributions (9000 data points) of proton and neutron differential cross sections and analyzing powers, which is significantly more accurate and complete than previous analyses [Per76, Men71, Bec69]. This parameterization adapts the basic Woods-Saxon form of Equation 2.3.10 but some parameters have Slight modifications. One special feature of the parameterization of CH89 is that, based on the parameterization of nuclear charge radii [Mye73], offset values are added to the conventional radius parameters: RV = rvAl/3 + rvw), Rw = rwA1/3 + rwm) R50 = rsoAl/3 + rig), RC = rcAl/3 + r20) (2.3.14) 35 0 0 0 . . . . . where rvw) , rso( ) , rw( ) , and rc( ) are offset radrus of the real, Imaginary, sprn-orbrt, and Coulomb potentials. The other special feature of CH89 is that the depths of the potential have more complex dependence on the energy and proton-neutron number. v, =vO :v, +(E—EC)Ve (2.3.15) 6Z6 2 f , or proton EC =4 5Rc (2.3.16) l O, for neutron W, W, — W ‘0 E E (2.3.17) l+exp[ veO_( — C)] erW N - Z W30 i Wst T w, = E E w (2.3.18) l+exp[( _ C)_ 5‘30] 1 W56”) where ‘+’ is used for protons and ‘—’ for neutrons. The parameters used in potential CH89 are listed in Table 2.4. Figure 2.10 shows the shapes of CH89 proton potentials on 13C at incident energy of 12.5 MeV. The global potentials of Menet and Perey & Perey are plotted together for the convenience of comparison. Figure 2.11 shows the calculations for proton elastic scattering on 13C at incident energy of 12.5 MeV using the potentials of Menet, Perey & Perey, and CH89. The potential CH89 gives better fitting to the data. Figure 2.12 shows the shapes of proton potentials on 13C for CH89, Menet, and Perey & Perey at incident energy of 30.95 MeV. The surface regions of the real parts are close to each other. Figure 2.13 shows the calculations for proton elastic scattering on 13'C at 36 incident energy of 30.95 MeV employing the potentials of Menet, Percy & Perey, and CH89. It is obvious that the potential CH89 gives better fitting to the data. Based on the above comparisons, we adopt the potential of CH89 in our calculation in a wide energy region. 37 Table 2.4: Parameters for the global nucleon-nucleus optical-model potential of CH89 [Var91] ‘ Parameters Value Parameters Value V0 52.9 MeV aso 0.63 fm vt 13.1 MeV wv0 7.8 MeV v, -0299 wv,O 35 MeV rv 1.250 fm erw 16 MeV r,‘°) -0225 fm wso 10.0 MeV av 0.690 fm w,, 18 MeV rc 1.24 fm wseO 36 MeV r30) 0.12 fm wsew 37 MeV vso 5.9 MeV-rm2 rw 1.33 fm no 1.34 fm rW‘O’ -0.42 fm rsom) .1.2 fm aw 0.69 fm 38 0 —10_— ”>7 i (D —207 2 . v _ ”t? —307 v _ /// :> // —: Menet —40 h /-’,’ ------ : Pere &Pere /.. Y Y /2/’ _._. ..... . CH89 Ara/1:04.; l . A K“\\ T) 10' 2 ; """"""" v _ A ’ t... —20;‘ v _ 3 . _30’....1....1....1...L O 2 4 6 8 Figure 2.10: Comparison of the CH89 proton potentials of 13C with the proton potentials of Menet and Perey & Perey at incident energy of 12.5 MeV. 39 104 A 3 5.} 10 ' (I) \ PO 2 g 10 c: E 10 o "O 1.0 10‘1 I I Y T I l 13C(p,p)13C (g.s.) Ep=12.5MeV —— : Menet ______ : Perey8cPerey _.-.-._.- :CH89 l I I I A l 50 100 98m.(deg) Figure 2.11: Calculations for proton elastic scattering on 13C at incident energy of 12.5 MeV using the potentials of Menet, Perey & Perey, and CH89 compared with the data [Wel78]. 40 0_ /,/»~ ' /./ —10} “ g i CD ~20? E . 2: —3oi- ;> I /// ——:Menet —40 f" T / — — —: Perey8cPerey ?/’// — —-n CH89 _.50’..._\ . 1 . 1 I 1 A LLL‘ \\ / > : 7" \ _,/ / /' Cl) _ — \ / 2 10: \w/ v i E —207 3 I O 2 4 6 8 r(fin) Figure 2.12: Comparison of the CH89 proton potentials of 13C with the proton potentials of Menet and Perey & Perey at incident energy of 30.95 MeV. 41 5 10.4vrrtvvat E 13C(p,p)13C (g.s.) Ep=30.95MeV 104? _ - —— :Menet q ’ j ------ : Perey&Perey 7 E 103?— 7°} :CH89 —. \ . . PO 2 E 10 E‘ C C E 10 r b E “O 1.0 E— 10_1* . . L . 1 . . . . 1 . . . . N: O 50 100 150 0am. (deg) Figure 2.13: Calculations for proton elastic scattering on 13C at incident energy of 30.95 MeV using the potentials of Menet, Percy & Percy, and CH89 compared with the data [Bar88]. 42 2.3. 1. 1.6 J LM Optical-Model Potential Instead of fitting elastic scattering data phenomenologically, the optical-model potential could be determined from nuclear matter theory, which may supply more microscopic understanding of the nuclear interior and overcome the uncertainties of the geometry parameters in the global optical—model potentials described previously. One such optical-model potential developed via realistic nucleon-nucleon interaction and nuclear matter density is the JLM (the initials of the threeauthors: Jeukenne, Lejeune, and Mahaux) potential [Jeu77]. The JLM potential started from the Brueckner-Hartree-Fock approximation and Reid’s hard core nucleon-nucleon interaction, which was folded with the nuclear matter density. The complex optical-model potential in infinite nuclear matter is parameterized for nuclei with mass numbers 12 S A S 208 and for energies E up to 160 MeV. For the nucleus whose nuclear matter densities are available experimentally, the JLM potentials may model the shape of the optical potentials more accurately than the phenomenological ones. The real and imaginary JLM potentials are expressed as: 4 4.2 — _ —3 V50) . ’"l 3. VE(r)_/t,-(bJ?r_) —p—(—r)—Ip(r)exp -——bZ—— d r _ _3WE(r) IF-f'l2 3 WE(r)=/1w-(b\/;t_) ——jp(r')ex — d r' (2.3.19) p(r) b2 where b = 1.2 fm, corresponding to the range of effective interaction [Gil71]; the scaling factors 2., and xlw is 1.0 and 0.8 for the real and imaginary potentials, respectively 43 [Pet85]; the VE(r) and WE(r) are the real and imaginary nucleon potentials derived in the local density approximation (LDA). LDA implies that the value of the potential at each point of the nucleus is the same as in a uniform medium with the same local density. In the case of a neutron with energy E, the LDA potential in uniform nuclear matter with density p and neutron excess 8 is given by: V,,(p.E) =Vo(p.E)+5 -Vt(p.E) W,,(p,E) =W0(p,E)+6-W1(p,E) (2.3.20) where the neutron excess is measured by the asymmetry parameter 5 6_pn—pp _ (2.3.21) pn + pp There are different models to parameterize the density distributions of protons and neutrons. One that is used in this work is the modified harmonic-oscillator model [Dej74] as defined by: ptr) = 700[1+0’('2)2] exp[—[-;-]2] (2.3.22) The parameters a and or can be read from Ref. [Dej74]. The density p0 is [Neg70]: p0“): 3 3'; 2 2 , k=NorZ (2.3.23) 472Cp (1+7: ap /Cp ) where ap= 0.54 fin, and Cp = (0.978+0.0206A“3)A“3 (fm) (2.3.24) The quantity V0 (p, E) is parameterized to 44 M Di “he 3 . . V0(p,E)= z aijp'El“ (2.3.25) i,j=l The coefficients aij are listed in Table 2.5. The parametric form of the imaginary potential W0 (p.15) is 4 . . 2 are/2'5!“ i,j=l W0(,0, E) = (2.3.26) D 1+-———2 (E—é'F) Where D = 600 MeV 2, 3F (,0) = p(—510.8 +3222p—6250p2), the coefficients d),- are listed in Table 2.5. The function V,(,0,E) and W1 (p, E) have the forms: 171 ,E Vl(p9E) :—(”%—2R6N(p,E) W1(p.E) = _ ’" ImN(p,E) (2.3.27) m(,0,E) where iii adapts the form: WEE) 3 i ‘—1 —=1— 2 cl-jp E] (2.3.28) m i,j=1 The coe fficients Cij are listed in Table 2.5. The IT: is calculated by , * m=mf a3») m where th e effective mass m * is defined as a: M). = 1——d—V0(p, E) (2.3.30) m (IE 45 N(p,E) is the auxiliary function in Brueckner-Hartree-Fock approximation [Jeu77]. The real part of N is parameterized by: 3 . . ReN: z bijp'EJ-l (MeV) (2.3.31) i,j=1 with coefficients by- listed in Table 2.5. The imaginary part of N is parameterized by 4 . . 2 ftjplEj l i,j=l Im N(,0, E) = (2.3.32) F E—8F 1+ where F=l.0 MeV. The coefficients fij are listed in Table 2.5. In the case of a proton with energy E in the additional presence of a Coulomb field Vc, the corresponding real and imaginary potentials are given by Vp(.0,E) = Vo(p.E) +Ac(p,E)-§ V1(;0.E-Vc) Wp(p,E) = W0(p,E) +WC(p,E) —5 Wl(p,E —VC) (2.3.33) where AC(P,E) = Vo(,0,E—Vc)—Vo(,0,E) WC(,0,E)=W0(p,E—VC)—W0(p,E) (2.3.34) Figure 2.14 shows the JLM proton potentials on 13C at incident energy of 12.5 MeV. For comparison, the global potentials of Menet and CH89 are plotted in the same figure. The surface regions of the real potentials are similar to each other but the interior part of the real J LM potential is deeper. 46 Figure 2.15 shows the elastic scattering calculations based on the above potentials. All of these calculations are quite similar at the forward angles (less than 10°). The calculated angular distribution by the JLM potential gives good fitting up to 125°. Figure 2.16 shows the JLM proton potentials on 13C at the incident energy of 30.95 MeV together with the global potentials of Menet and CH89. Again, the surface regions of the real potentials are similar to each other but the interior part of the real JLM potential is deeper. In addition, contrary to the other potentials, the imaginary JLM potential in the nuclear interior is positive. Figure 2.17 shows the elastic scattering calculations based on these potentials. The calculated angular distribution by the JLM potential looks similar to that by the CH89 potential. All of these potentials give quite similar results at the forward angles. Based on the above comparison, we can see that both the JLM and CH89 potentials are better than other global optical-model potentials; the JLM is even better than CH89. The disadvantage of the JLM potential is that it requires the information of nuclear density. When the nuclear density is available, we perform calculations with both the JLM and CH89 potentials; if the nuclear density is not available, only the CH89 is used. 47 Table 2.5: Parameters for JLM potentials b0 (a) aij for Vo(P,E) (b) dij for Wo(p,E) i . j=1 i=2 j=3 , j=1 j=2 j=3 j=4 J J i = 1 -974 11.26 00425 i = l -l483 37.18 -0.3549 0.001119 i = 2 7097 -125.7 0.5853 1 = 2 29880 -93l.8 9.591 -0.0316 = 3 -19530 418 -2.054 1 = 3 -212800 7209 -77.52 0.2611 1 = 4 512500 -l7960 198 06753 (C) Cij fOI' 171(p, E) (d) fij fOI' MN i j=1 j=2 j=3 j=1 j=2 j=3 j=4 L—nv 4.557 -0.00529l 0.6108E-5 546.1 -ll.2 0.1065 -3.54113-4 _ do -2.051 -0.4906 0.001812 h ll ~847l 230.0 -2.439 0.008544 WN— -65.09 3.095 -0.01190 51720 -1520 17.17 -0.06211 II #0319— -1 14000 3543 -41.69 0.1537 — (e) bij for ReN b0 j=1 j=2 j=3 360.1 -5.224 0.02051 -2691 51.3 -0.247 MINI- 7733 ~17l.7 0.8846 48 0. _10_. i A i > —207 Q) t E : —30r A _ La . v > —40— -—50 7:} l' _GOl...i....1.,_H._.12.LL A »::.‘__‘::~..\~‘ 1%— > : 1“‘\\\“ -__,<,"’" Q) t ‘~__ " 2 ~10: V P A t é; —20: _3OWIILL 0 2 4 6 8 Figure 2.14: Comparison of the JLM proton potentials of 13C with the proton potentials of Menet and CH89 at incident energy of 12.5 MeV. 49 5 10 . . , ' I Y , v r I 13C(p,p)13c (g.s.) Ep=12.5MeV 104— A 3 L; 10 ' (I) \\ PO 2 g 10 — C2 31 10 b '0 Lo—— _ 10—1 i . . . l a i . . I i . l . o 50 100 150 90m.(deg) Figure 2.15: Calculations for the l3C(p,p)13C (g.s.) reaction by JLM, Menet and CH89 potentials at incident energy of 12.5 MeV compared with the data [Wel78]. 50 OF 3 I / E _40 —/ A h / L. : / > 60L // —:Menet : —-——:JLM ; _._.-:CH89 m . . i l . 1 L i L i r J r l —80_\\ A - \ i ' ‘\ 2 o i \ ,_. 1? : E _20 r 11.1, L r 1 [1 I O 2 4 6 8 r (fm) Figure 2.16: Comparison of the JLM proton potentials of 13C with the proton potentials of Menet and CH89 at incident energy of 30.95 MeV. 51 t; 10 m i \ PD 2 g 10 c: E 10 b "C . 1.0E 10‘1 l?"rl 13C(p,p)13C (g.s.) Ep=30.95MeV — : Menet ------ : JLM : CH89 Figure 2.17: Calculations for the l3C(p,p)]3C (g.s.) reaction by JLM, Menet and CH89 potentials at incident energy of 30.95 MeV compared with the data [Bar88]. 52 2.3.1.2 Zero-Range Approximation The DWBA expression for the transition amplitude in Equation 2.3.2 involves a 6- fold integration over 7,, and Fp after the integration of the nuclear coordinates B. The 6- fold integration has been discussed by [Au564] and Sawaguri [Saw67]. To simplify the integration, it is usually assumed that the transition amplitude receives contributions only from the region where the coordinates of the proton and neutron coincide so that we have the zero-range approximation: 0(7) 2 v,,,, (rm (f) = Do 66) (2.3.35) The value of Do can be obtained by integrating this equation over f : _ 2 — _ 1)0 _ j r Vpn(r)¢d(r)dr (2.3.36) and the vertex constant 002 is [Lee64, Knu75] : 002 =15006.25 MeV2 . fm3 , (2.3.37) When the zero-range approximation is made, the coordinates are transformed to: B 7 (2.3.38) __.) 3+1" rn—>R, rp The term Brl(iP,Ed)becomes mz- - _.—1 —(1/2) (—)*- - ml — (+)- B — — Bl (kp,kd)—z (21+1) DOI/rd (kd,R)¢nl(R)zp (kp,B+lR)dR (2.3.39) 53 2.3.1.3 Finite-Range Approximation In general, deuteron has finite range of radius and the interaction between proton and neutron exists in a finite range. The zero-range approximation over-emphasizes contributions coming from the nuclear interior. A means has been found to approximate the finite-range effect so that it reduces to the form of the zero-range approximation multiplied by a radial dependent factor A(R) [But64]. ml- - _.-1 —(1/2) (—)*- - ml — (+)‘ B - - B, (kp,kd)—z (21+1) Dojzd (kd.R>A(R)¢n,(R)zp(k,,.B+1R)dR (2.3.40) The factor A(R) is A(R) = l—(a/fl)2(l/Ed)[Ud(R)-V,,(R)—Up(R)-Ed] (2.3.41) = h 2 1/2, : mdMB a (#Ed/ ) ‘u md + MB where B is the finite range parameter with the value of 0.7457 [Knu75], Ed is the deuteron binding energy, U d and U p are the deuteron and proton optical potentials, and V" is the neutron potential that binds the neutron to the core nucleus B. Figure 2.18 shows the comparison of the DWBA calculations in this finite-range approximation (solid line) and zero-range approximation (dashed line) for the reaction of l3C(p,d)12C at proton energies of 15 MeV and 48.3 MeV. The calculations use CH89 as the proton potential and Daehnick global potential for the deuteron. Thus finite—range approximation increases the cross section by 4.8% at the peak region for the proton energy of 15 MeV. For incident proton energy at 48.3 MeV, the enhancement is 8.9% at 54 forward angles. Thus the effects, although not negligible, are not very large for the reactions that we studied. It should be noted that results from this finite-range approximation are very close to the exact finite-range calculations with numerical solution of the deuteron wave function. The latter kind of calculation is available by another widely used finite-range DWBA code: DWUCKS [Kunz]. Figure 2.19 shows the comparison between the calculations from two codes for the l3C(p,d)12C reaction. They are very close to each other especially at the forward angles. At proton energy of 15 MeV, the curve from TWOFNR is just 1.5% lower than that from DWUCKS at the region of first peak. At proton energy of 48.3 MeV, the curve from TWOFNR is 2.0% higher than that from DWUCKS at the forward angles. However, TWOFNR is more user friendly with many options to choose from. 55 ' ..- — \ 13 12 f \ C(p,d) c (g.s.) « A e ‘ \ /,2\ m 10. : \ / —. r-\ \ _, / : E = \ 15 MeV \9; Cl ’ \\ ' "C5 \\\ 48 3 MeV \ - \ \° b 10“— U Finite range — - — Zero range 10‘2 1 1 1 1 l 1 1 1 L l L 4 m 1 1 1 1 O 20 4O 60 80 Gem (deg) Figure 2.18: Calculations for reaction of l3C(p,d)12C (g.s.) using finite-range approximation (solid line) and zero-range approximation (dashed line) at incident energies of 15 MeV and 48.3 MeV. The cross sections at proton energy of 15 MeV have been multiplied by 10 so that the calculations at the two energies can be seen more clearly. 56 dU/dQ (mb/sr.) TWOFNR (Finite—range) “““““ DWUCK5 10—2 . . . 1 1 . . . . 1 1 1 . . l 1 . . 1 1 . O 20 4O 6O 80 (deg) 6 CH] Figure 2.19: Finite-range DWBA calculations for reaction of l3C(p,d)12C (g.s.) by TWOFNR (solid line) and DWUCKS (dashed line) at incident energies of 15 MeV and 48.3 MeV. The cross sections at proton energy of 15 MeV have been multiplied by 10. 57 2.3. 1.4 Non-Locality Correction The optical-model potential is usually taken to have simple local form, which means that, at the point r, the particle feels the potential only at that point. The Schrodinger equation reads h 2 —-2——V2 +UL(f)—E 1/1(f)=0 (2.3.42) ,u The real situation is more complicated and the optical potential should be non-local, which means that the wave function at point 7 is affected within the range of non-local potential. U L (FM/(f ) in Equation 2.3.42 should be replaced by IU(?,?')1//(F')d?' (2.3.43) where U (?,F')is the non-local potential. Non-locality can be expected wherever the potential is energy dependent that comes from the exchange terms required by the asymmetry of the overall wavefunction. This effect has been studied by Percy and Buck [Per62]. They separated the non- local kernel U (7, f') into a potential fonn U times a Gaussian non-locality function. 7+? U(F,f')=U[ )H (F — 7') (2.3.44) where H (F — 'r") was chosen to be a Gaussian function: H(? — 7') = ex"(— ((7 - PM” )2) (2.3.45) 3 [#2131113] 58 where ,BNL is the range of non-locality. The value of flNL that Percy and Buck found to yield the best fit to the data of neutron scattering on Pb over an energy range from 4.1 MeV to 24.0 MeV is 0.85 fm [Per62]. For the deuteron, flNL is 0.54 fm [Per74]. Figure 2.20 shows the comparison between the local and non-local proton potential from CH89. The local potentials are generally weaker than the non-local potentials, |VL| <|V~L| and IWLI <|WNL| , especially within the interior of the nuclei. Thus, non- locality reduces contributions to transfer reaction from the interior of the nucleus. The main change produced in the cross section is the reduction of the large angle scattering while increasing the forward or peak cross section [Phi68]. Figure 2.21 shows the calculations for the reaction of l3C(p,d)12C at incident energy of 48.3 MeV, where proton potential chooses the global potential of CH89 and the deuteron potential adapts the global potential of Daehnick [Dae80]. Finite-range approximation is employed in these calculations. The dashed line shows the result from the local proton potential; the solid line shows the result when non-locality correction is applied to the proton potential. We can see that the cross section by the non-local proton potential increases 12% at forward angles, where the spectroscopic factors are extracted. Similar effect is obtained when the non-locality correction is applied to the deuteron potential. The total effects are cumulative when the non-locality correction is applied to the proton and the deuteron simultaneously. Thus it is important to include non—locality corrections in transfer reaction calculations. The non-locality correction at lower energy is also examined. Figure 2.22 shows the calculations for the same reaction at incident proton energy of 15 MeV. The calculations with and without non-locality correction are very close at the region of first peak. 59 Therefore, the extracted SF are not strongly affected by the non-locality correction at low energy. 60 v(1~) (MeV) T; (l) 5 E B _4OF111111111L111L11111 O 2 4 6 8 r(fm) Figure 2.20: Comparison of the local (dashed line) and non-local (solid line) proton potentials (CH89) of 13C at incident proton energy of 48.3 MeV. 61 2 10;'rrfi1""1""1""1'3 13C(p,d)12C (g.s.) Ep=4eeMeVi —— Nonlocal — — — Local 10-2 1 1 1 1 l 1 1 1 L 11 1 1 1 I 1 1 1 1 O 20 4O 60 80 Gem (deg) Figure 2.21: Calculated differential cross section for l3C(p,d)12C at incident proton energy of 48.3 MeV by non-local proton potential (solid line) increased 12% at the forward angles compared to that by local proton potential (dashed line). 62 ‘13 101 \ E C 1.0; "C . \ b "C —— Norrlocal — -— — Local 1L11l l l l l l l l 1 1 L I l. l l l J 20 4O 6O 80. Gem (deg) Figure 2.22: Calculated differential cross sections for 13C(p,d)12C at incident energy of 15 MeV using non-local proton potential (solid line) and local proton potential (dashed line). 63 2.3.1.5 Neutron Form Factor In most analyses of single neutron transfer reactions, it is assumed that the neutron is picked up or deposited into a shell model single—particle state. In the standard energy separation procedure, the corresponding single-particle wave function ¢nlj (7,1), called the neutron form factor, is usually taken to be an eigenfunction of a Woods-Saxon potential whose geometry is fixed (ro=1.25 fm, ao=0.65 fm, R = r0 -Al/ 3 fm) and depth is adjusted so that the eigenvalue is equal to the experimental neutron separation energy [Pin65, Ber65]. Table 2.6 lists the information of the neutron form factors for the four reactions we studied in this thesis. Figure 2.23(a) shows ¢an (r) in the 1 pl ,2 orbit for the reaction of p(l3C,d)lzC (g.s.). The depth Vn is adjusted to be —39.779 MeV corresponding to the neutron binding energy of —4.946 MeV. The neutron form factor for the reaction of p(l3C,d)12C (2+ , 4.439 MeV) is shown in Figure 2.23(b). In this case, the neutron orbit is 1173/2 and the neutron separation energy is —9.385 MeV. Similarly, Figure 2.6 (c) and ((1) show the neutron form factors for the reaction of p(”B,d)lOB (g.s.) and p(10Be,d)9Be (g.s.), separately. The well-depths and the binding energies are listed in Table 2.6 It is important to examine the sensitivity of the calculations to the parameters r0 and 210 of neutron potential. Figure 2.24 shows the dependence on the neutron radius parameter for the reaction of l3C(p,d)12C (g.s.) at proton energies of 15 MeV and 48.3 MeV. CH89 and Daehnick potentials are used for proton and deuteron respectively. (Finite-range approximation and non-locality correction are employed in all following 64 calculations). The neutron radius parameter r0 is changed from 1.2 fm to 1.3 fm while the spin-orbit strength is fixed to zero and the neutron diffuseness is fixed to 0.65 fm. The change of 0.1 fm (corresponding to 8% change) changes the cross sections at forward angles by 16% at higher incident energy and 11% at lower incident energy. Figure 2.25 shows the dependence on the neutron diffuseness parameter, where the neutron diffuseness changes from 0.6 fin to 0.7 fm with the r0 fixed to 1.25 fin and Vso fixed to zero. The increase of 0.1 fm (corresponding to 16% change) in the neutron diffuseness increases the cross sections at forward angles 20% at higher energy and 17% at lower incident energy. Figure 2.26 shows the dependence on the spin-orbit strength. When the spin-orbit strength of the neutron potential changes from 0.0 MeV to 6.0 MeV, the cross sections at forward angles decrease by 6-8%. Since this effect is small, the spin-orbit strength is set to zero in all of the following analyses. Table 2.6 The neutron potentials and the binding energies of the neutron form factors in the reactions of this experiment Reaction N222?“ Emdgllc’lgeflnfrgy (181g, ) (f1: ) (331) p(‘3c,d)‘2c g.s. 1191/2 -4.946 -39.779 1.25 0.65 p(‘3c,d)‘2c 2+ 1123/2 -9.385 48.257 1.25 0.65 p(”B,d)'°B g.s. 1103/2 41.455 -56.853 1.25 0.65 p(lOBe,d)9Be g.s. 1103/2 -6.811 -50905 1.25 0.65 65 110 E 1 -2 10 5 10"3 Form factor I jTTIVTI 10 5 10 R (fm) Figure 2.23: Neutron form factors for the reactions of (a) p(l3C,d)12C (g.s.), (b) p(‘3c,d)‘zc (2*), (c) p(”B,d)‘°B (g.s.), and (d) p(‘°Be,d)9Be (g.s.). 66 do/dQ (Inb/sr.) 6cm (deg) Figure 2.24: Calculations on the variation of neutron radius parameter to, where the neutron diffuseness a0 is fixed to 0.65 fm and the spin-orbit strength Vso is 0.0 MeV. The cross sections at proton energy of 15 MeV are multiplied by 10. 67 3 10 ""I'fi7rl 102 — ”'5 5—1 —\ g 10. Ni I \ E \ V 10 __ “a C: E "o 1 E ‘ 0‘1— - 'U 1 ao=0.6 frn i - — -— a0=0.7 fm l 10‘2 1 . . - ' . 0 20 Gem (deg) Figure 2.25: Calculations on the variation of neutron diffuseness a0, where the neutron radius parameter r0 is fixed to 1.25 fin and the spin-orbit strength Vso is 0.0 MeV. The cross sections at proton energy of 15 MeV are multiplied by 10. 68 ,5 L; U) \ E c: _ . "o \ 9 10"—- ‘~ U vso=oo MeV ‘ _ — — — VSO=B.O MeV 10‘2 1 1 1 1 I 1 1 1 m l 1 1 1 1 1 1 1 1 1 o 20 4o 60 80 Gem (deg) Figure 2.26: Calculations on the variation of neutron spin-orbit strength V50, where the neutron radius re and diffuseness a0 are fixed to 1.25 fm and 0.65 fm separately. The cross sections at proton energy of 15 MeV are multiplied by 10. 69 2.3.2 Adiabatic Deuteron Breakup Approximation (ADBA) Deuteron is composed of two loosely bound nucleons, a proton and a neutron. Since the separation energy between the proton and neutron is 2.224 MeV, deuteron breaks up easily in the field of core nucleus. Thus the extraction of the spectroscopic factors from (p,d) and (d,p) reactions using the DWBA calculations is usually not very reliable especially at high incident energy because of inadequate treatment of the breakup effect of deuteron [Pea66, But67]. Johnson and Soper [Joh70] extended the DWBA theory involving deuterons to adiabatic deuteron breakup approximation (ADBA), which requires only a specification of the nucleon-target interactions. In this approximation, the effective two-nucleon- nucleus interaction is assumed to be the sum of the nucleon optical-model potentials evaluated at one-half the incident deuteron kinetic energy. The deuteron adiabatic potential is defined as: 4 1 _ 1_ - 11 1 _ Ud(R)=——j Un(R+—r)+Up(R——r) Vpn(r)¢d(r)d? (2.3.46) D0 2 2 where U" and Up are the neutron and proton optical potentials at one half the deuteron bombarding energy, R is the coordinate of the deuteron center of mass and f is the relative coordinate between proton and neutron, VP" ('r‘) is the interaction between proton and neutron, ¢d (7) is the deuteron wave function, and D0 is defined in Equation 2.3.36. The exact (d,p) and (p,d) transfer reaction amplitudes require knowledge of the adiabatic three-body wave function only at small neutron-proton separations. There, the adiabatic distorting potential governing the center of mass motion of the deuteron is well 70 described by the sum of the neutron- and proton- target optical potentials. It is important to stress that this adiabatic distorting potential generates the three-body wave function in that limited region of configuration space needed to evaluate the transfer amplitude, and it does not describe deuteron elastic scattering at the beam energy. Figure 2.27 shows the comparison of two deuteron potentials of 12C at Ed = 49.2 MeV. The dashed line presents the Daehnick global deuteron potential and the solid line represents the adiabatic deuteron potential constructed by the CH89 nucleon potentials. The adiabatic potential based on CH89 is deeper in the interior and is shallower at the surface. The effect on the stripping or pickup cross section is to cause a faster fall off with angle and to create stronger oscillations at higher incident energy. Figure 2.28 shows the calculations of ADBA (solid line) and DWBA (dashed line) for reaction of l3C(p,d)12C. At the forward peak, the ADBA calculation is 29% larger than that of the DWBA calculation. The situation is different at lower incident energy. Figure 2.29 shows the calculations for the same reaction at proton energy of 15 MeV. The ADBA (solid line) has similar peak value as the DWBA (dashed line) and thus has little effect on the extraction of spectroscopic factors. However, to be consistent in the use of deuteron potentials throughout the range of energy, we choose ADBA calculations. 71 V(r) (MeV) ’>‘ Cl) 5 . E a )- _50111111111111111 0 2 4 6 8 r (fm) Figure 2.27: Comparison of Daehnick global deuteron potential (dashed line) with the adiabatic deuteron potential (solid line) constructed by CH89 potentials, for 12C at Ed 2 49.2 MeV. ‘ 72 A P 18C(p,d)12C gs. Ep=48.3MeV : ‘13 101 —: \ "E 'U - \\\ : \ Z \ g \ 1.0—1:— — I —— ADBA — — — DWBA 10‘2 1 1 1 1 l 1 1 1 n l_1 O 20 4O 6 (deg) CH1 Figure 2.28: Comparison of ADBA (solid line) and DWBA (dashed line) calculations for reaction of l3C(p,d)12C at proton energy of 48.3 MeV. The ADBA increases the cross section at the forward angles and faster fall off than DWBA. 73 do/dQ (mb/sr.) 0 20 4O 60 80 Gem (deg) Figure 2.29: Comparison of ADBA (solid line) and DWBA (dashed line) calculations for reaction of l3C(p,d)12C at proton energy of 15 MeV. 74 2.4 Momentum Matching Assuming PI) is the incident proton momentum and Pd is the momentum of the deuteron, the transferred momentum P, of the neutron is given by conservation of momentum [Fes92]: P, = Pd — Pp (2.4.1) From this equation one can immediately determine the magnitude of P,: Pz-P2+P2—2PP 6 , — p d p d cos (2.4.2) Where 6’ is the angle between the direction of the final deuteron and the direction of the incident proton. The angular momentum transferred, hl, , must be less than P, R, where R is the projectile-target separation at which the reaction occurs. Hence 1 h 2(1, +5) 2 3 [1,2197- (2.4.3) So that (19,119)2 +(kpR)2 —(1, 41%)2 cos 6 S (244} 2(kd R)(kpR) where hk as usual equals p. This equation expresses the approximate relation between the angular position of the first peak in the angular differential cross section and the transferred orbital momentum so that the angular position of the first peak in the measured angular differential cross section will tell us the value of the transferred orbital momentum. 75 The transferred momentum is bounded by the momentum of the transferred neutron by [Fes92]: 2m|€| h 2 A -——k —k < A+l p d (2.4.5) where 6‘ is the binding energy of the neutron. Good momentum matching that satisfies the Equation 2.4.5 gives slow radial oscillations and large overlaps in the nuclear surface region [Au387]. Based on the neutron binding energy listed in Table 2.6, the upward limits of the proton incident energies in l3C(p,d)12C reaction to the ground state and the first excited state are 32 MeV and 66 MeV respectively; for reactions of llB(p,d)mB (g.s.) and 10Be(p,d)9Be (g.s.), the upward limits of the proton incident energies are 82 MeV and 46 MeV respectively. 2.5 Summary As one of the fundamental tests of the shell-model theory, the spectroscopic factor measures the occupancy of a nucleon in a pure single-particle state. It can be derived from the ratio of the measured cross section to the calculated cross section assuming pure single-particle state. The theoretical cross sections are calculated via DWBA and ADBA models. The proton and deuteron global optical-model potentials have been discussed. The nucleon-nucleus potentials of CH89 and JLM usually give better descriptions to the proton elastic scatterings than other global potentials. An adiabatic deuteron potential can be constructed based on the nucleon-nucleus potentials such as CH89 and JLM. ADBA 76 calculations give better predictions than the regular DWBA calculations especially at higher incident energy where deuteron break-up effect is significant. The wave function of the transferred neutron (neutron form factor) is obtained by adjusting the depth of the neutron potential to match the neutron separation energy to the experimental value. The radius parameter and the diffuseness of the neutron potential are usually fixed to 1.25 fin and 0.65 fm respectively. The sensitivities of the calculations to the geometry parameters of neutron potential have been examined. We chose target densities to have the form of modified Harmonic oscillator [Dej74] for the JLM potentials. The scaling factor )1 of the JLM potentials were chosen to be 1.0 and 0.8 for the real and imaginary parts respectively [Pet85]. The momentum matching is discussed in Section 2.4. All the standard input parameters used in TWOFNR for our calculations are listed in Table 2.7. We adopted the value of 15006.25 MeV2 - fm3 [Knu75] for the vertex constant 002. Finite-range approximation is employed in the calculations with the Hulthen finite-range factor of 0.7457 fm [Knu75]. Non-locality correction is also employed with the non-locality range IBNL to be 0.85 fm [Per62] and 0.54 fm [Per74] for the proton and deuteron potential respectively. 77 Table 2.7: Summary of the input parameters used in TWOFNR DWBA Adiabatic CH Adiabatic JLM Proton potential Chapel-Hill [Var91] Chapel-Hill [Var91] JLM [Jeu77] Deuteron potential Daehnick [Dae80] Adi abatic [J oh70] Adiabatic [J oh70] from CH ‘ from JLM Modified Harmonic Target densities oscillator density [Dej74] Woods-Saxon, Woods-Saxon, Woods—Saxon, r0=l.25 fm, ro=l.25 fm, ro=1.25 fm, n-binding pctential ao=0.65 fm, a0=0.65 fm, ao=0.65 fm, depth adjusted, depth adjusted, depth adjusted, no spin-orbit no spin-orbit no spin-orbit Hulthen finite range factor (frn) 0.7457 0.7457 0.7457 [Knu75] Vertex constant Do2 ( Met/2 . fm3) 15006.25 15006.25 15006.25 [Knu75] JLM potential Av=1.0 and 71w=0.8 scalin g A N/A N/A [Pet85] p: 0.85 fm; p: 0.85 fm; p: 0.85 fm; Non-Locality range d: 0.54 frn; d: 0.54 fm; d: 0.54 frn; n: N/A; n: N/A; n: N/A; 78 CHAPTER 3 EXPERIMENTAL SETUP AND DETECTOR CALIBRATION 3.1 Overview This experiment was performed at the National Superconducting Cyclotron Laboratory (NSCL) at Michigan State University. One experimental objective is to study the nuclei with neutron number, N=6, such as 1'B and 10Be which can be produced by bombarding the production target of 98c with a primary beam of 13C (E=80.4 ~A MeV) produced from the K1200 cyclotron. In addition to HB and 10Be beams, a secondary beam 13C was also produced and the reaction p(l3C,d)12C was used for energy calibration of the CsI detectors as explained in Section 3.3. The schematic diagram of the beam fragment separator A1200 is shown in Figure 3.1. Fragmentations from the collision of the primary beam with the production target of 9Be are bent by the two dipoles. A momentum slit at dispersive image #1 selects the desired particles according to their mass and momentum. A wedge at dispersive image #2 could be used to further disperse the particles according to their energy loss. There is another momentum slit at the final achromatic image that select the desired secondary beam. The thickness of the production target (98e) and the beam intensity are listed in Table 3.1. The experiment was carried out in the S800 vault at NSCL. Figure 3.2 shows the schematic of the facilities in the 8800 vault. The secondary beam particles produced after 79 the A1200 were transported to the 8800 line. They were bent through the analysis line consisting of Sextupoles, an Intermediate Image, Dipoles, and a Quadrupole Triplet before the target chamber. The (p,d) reactions took place inside the target chamber which contains the reaction targets of polyethylene (CH2)n foils. The thickness of the (CH2)n targets are listed in Table 3.1. The total detection system includes the Multi Wires Drift Counter (MWDC) detectors, the Large Area Silicon Strip Array (LASSA) [Wag01, DavOl], and the S800 spectrometer. The following sections will describe each of these detectors separately. 80 2 sex ma «.2 4.23 was ad on 2 62wa 2:. 2% a: 5 on 2 msxnm ma... owe. me. SW 82 2 $.an 2.4 fix... an. Ex m: $2295 23 911% 2.165 mane 393 NED . Bombs 283 $55 .888 »w 90—): Hawks omo 832 Co $05.25. @656 58m 52335 ~28 .88. 2809 38:0 mo swear—E: Succuom . Ewen Evacuom .6 $2365 Bombs 2: use 253 b22508 2: Co 85885 2a. ”_.m 2an 81 8009 309.008 09:... 20082.3( .mcfi .oomjx mo EEwflu oszocom ”_.m 03mm an 090...: 020.0006 E 09:: 020.305 8.6.26 085. E0: snow 41% «090 1.. 0mm 82 23> oowm E 82:08 05 we EEwEv 0:080:0m ”Nd ocswfi Eunice N “Encom— 20950sz 858020 woman. ~0120901/BEEP Boga—00:0 /\\.1fim1:&w ... \ wo_om_olV@ goes 020958 » x... \ ND O~OQ_QI\@’\~C~ owns: cum—flogged. /m_a=m_lmum 111111 Eoooaoor\\0% 85:00 3003850 25 E _aoom 83 3.1 MWDC Detectors 3.2.1 Principle of MWDC Detector The Multi Wire Drift Counters (MWDC) are used to measure the beam positions and angles at the intermediate image chamber of the 8800 vault. Each MWDC detector has an active area of 11.2xl 1.2 cm2 covered by PPTA (p-Phenylene Terephthalamide) [Fuj89] foil of 50 pm thick at the front and back windows. It is filled with P30 gas (70% Ar and 30% CH4) at a pressure of 500 torr. Beam particles transversing through the production targets and the MWDC detectors lose energy due to interactions of the beam with the target and other materials used to construct the counters. Taking into account the energy loss which could be substantial with thick Be targets, the beam energies before the (CH2)n targets are calculated and listed in Table 3.1. Each MWDC detector has two orthogonal wire planes for the X direction and the Y direction, respectively. The front and back of these two wire planes are covered by 12 um thick mylar foil at front and back. Another mylar foil with the same thickness is placed between the two wire planes. A schematic drawing of one wire plane is shown in Figure 3.3. In each plane, there are 14 anode wires and 15 field wires. The working voltage of the anode wires is 580 Volts. The separation of the adjacent anode wires is 8.0 mm, and all the anode wires are connected to a micro-strip delay line. This delay line has two timing outputs T1 and T2. The field wires are separated into two groups. Every other field 84 wire is connected together and gives the left-right position signals E1 and E2, respectively. An enlarged drawing for the ions drifting between the wires is shown in Figure 3.4. The positive ions drift towards the field wires and the negative electrons drift towards the anode wire. The time signals T1 and T2 are expressed as: T1 = dtl + (k — 1) - Z0 + dz (3.2.1) T2 = dtz +(l4—k)-ZO +dt (3.2.2) where 20 is the delay time between adjacent anode wires, dtl and dtz are the delay times from the left and right wire ends to the preamplifier respectively, dt is the drift time of the negative electrons from the hit point to the anode wire, and k is the wire sequential number. Subtracting T2 from T1 yields : T1 —T2 = dtl —d12 +(2k —-15)-ZO (3.2.3) Then the wire sequential number k is (T1 —T2 —dtl +d12) +1—5— 220 2 k: (3.2.4) A spectrum of the wire sequential number k is shown in Figure 3.5. By adding T1 and T2, we get T, + T2 = dtl + dt2 +1320 + 2dr (3.2.5) Then the drift time dt is : T1+T2 -dtl -d12 _1320 2 dt (3.2.6) A spectrum of the drift time dt is shown in Figure 3.6, where the time is scaled in units of microseconds ( ,u sec). The sharp peak near the zero drift time and the tail at the drift 85 time around 80 us correspond to the nonlinear electron drifting. Analysis indicated that 5% of the total particles are affected by this nonlinear behavior. The left-right ambiguity is resolved by comparing the amplitudes of the two signals E1 and E2 from the field wires. When the negative ions move close to the anode wires, they produce a significant multiplicative effect and produce lots of positive ions around the anode wire. Some of the positive ions will drift to the field wires. Since more positive ions are produced on the side that the negative ions drift from, the particles hitting at one side of anode wire produce larger signals on field wire in this side than the other. Figure 3.7 depicts a typical spectrum of the signals from one side of field wires versus the signals from the other side. Since alternate field wires are connected together, the particle position is calculated as following in units of millimeters (mm). k dt 1, E11~32 (3 ) where k is the wire sequential number, dt is the drift time, D is the alignment center in value of 60.0 mm, and ('9 is the left-right ambiguity function of E. and E2. 3.2.2 Position Calibration To ensure that all the equations used in the position deterrrrination of MWDC detectors are correct, a mask made of a 3.1 mm-thick brass plate is used to calibrate the position. Figure 3.8(a) shows the pattern on the mask. The distance between adjacent big holes is 10.2 mm; the distance between adjacent small holes is 2.54 mm. There is one L- 86 shaped slit with the width of 2.0 mm. The area in the middle surrounded by a dotted rectangle is the area that detected the passing beam particles. Figure 3.8(b) is the reconstructed two-dimensional position spectrum of the mask using the 11B beam, where the beam was focused on the middle position between the first and second MWDC detectors. A position resolution of 0.4 mm was achieved. 87 \ \ i J T2 r-’ Tl Figure 3.3: Schematic drawing of wire plane of MWDC detector. All the anode wires are connected to a micro-strip delay line, which has two timing outputs T1 and T2. Every other field wire is connected together to form two groups and gives the left-right ambiguity signals E1 and E2, respectively. 88 El E2 (JD—r dt A 0 V Figure 3.4: An enlarged drawing of the ions drifting between the wires. The positive ions drift toward the field wires and the negative ions drift toward the anode wire. The term dt is the drift time of the negative ions from the hit point to the anode wire. 89 5000 *- ~ - Counts 2500 r - 1 1 1 AL 11‘ 7 8 9 10 91) OB Wire Number Figure 3.5: Spectrum of the wire sequential number k. Each individual sharp peak indicates the wire with its sequential number. 90 1500 1000 - - Counts 500 — O l l l l 1 J 0 20 40 60 80 100 Drift Time (,as) Figure 3.6: Spectrum of drift time dt, where the time is sealed in units of microseconds. 91 250— 200- 150 — E2 (Channel) 100- 50- r n r I r 0 50 100 150 200 250 E1 (Channel) Figure 3.7: Spectrum of E2 versus E1. There are two groups of particles. The left-right ambiguity can be clarified by which side yields a larger signal than the other. 92 O O O O O / ) 0 000000000 0 ooooooooo 000000000 0 OoooOoooO 0 000000000 000000000 000000000 V O O O O O A a: v .0; Q»... : $312 0 vi a 9 f '5‘“. .'.. II". 131940;. «is: o o -0 03“ "iii 19:! I . . A 5' v Figure 3.8: (a) Patterns on the mask used to calibrate the MWDC detector. The distance between adjacent small holes is 2.54 mm; the distance between adjacent big holes is 10.2 mm; (b) The two-dimensional position spectrum of the mask. The corresponding areas are surrounded by the dotted rectangle in (a) and (b). 93 3.3 LASSA Detector Array 3.3.1 Overview The Large Area Strip Silicon Array (LASSA) was designed and constructed to provide excellent energy, angular, and isotope resolution for the detection of charged particles. It has been used successfully to study the isospin degree of freedom in heavy ion multifragmentation experiments at NSCL [Tan02, LinS]. It consists of 9 individual telescopes, which can be arranged into various configurations. Each LASSA telescope consists of two silicon strip detectors and a cluster of four CsI(Tl) scintillators. The side view of the telescope assembly is shown in Figure 3.9. A 5 mg/cm2 SnPb foil covers the top window and provides a dark environment to the Si detectors. In addition, the foil also protects the Si detectors from electrons produced in nuclear collisions. The two silicon strip detectors are mounted inside the top frame below the SnPb foil. Right behind the two silicon detectors are four CsI crystals with light guides and photodiodes mounted on the back. The preamplifiers, and their motherboards, for the CsI detectors are placed directly behind the photodiodes. A cooling bar is mounted at the back of each telescope to keep the telescope at constant temperature. 94 3.3.2 Geometric Setup In this experiment, LASSA is used to detect the deuteron particles emitted from the (p,d) reactions. The 9 telescopes were separated into three groups and mounted on three independent frames. A schematic drawing of the detector setup is shown in Figure 3.10. The following coordination in the laboratory frame has been adopted: the beam direction is defined as the z axis and the upward direction is the x axis; the polar angle 0 defined the angle of the particle direction with respect to the beam axis; the angle Otis the angle between the particle projection on the z-y plane and the z axis; the angle B is the angle of the particle direction with respect to its projection on the z-y plane. For reference, each telescope is assigned a number as labeled in Figure 3.10. The geometry of the centoid of each telescope is specified in Table 3.2. The detectors cover the angular range of 3.6°<0<36.9° in the laboratory frame, which covers mainly the first peaks of the transfer reactions. Beyond this region, the kinematic broadenings increase dramatically in inverse kinematics. 3.3.3 Silicon Strip Detector Array 3.3.3.1 Overview Silicon strip detectors are widely used in nuclear experiments because of their excellent energy resolution and linear response for charged particles. Both layers of silicon strip detectors used in LASSA are ion-implanted, passivated devices, Si(IP), 95 obtained from Micron Semiconductor [Micr]. For all the strip detectors used in our experiment, the width of each strip is 3 mm, and there is a 0.1 mm wide gap between adjacent strips. The active area on each surface is about 50x50 mmz. The front silicon detector, which is labeled as DE, is about 65 um thick. It has 16 strips on the front surface. There are two different types of silicon strip detectors for the second Si detector. One is double-sided and about 500 um thick, and the other type is single-sided and about 1000 pm thick. In the double-sided detector, the strips on the front are perpendicular to the strips on the back. For convenience, we refer the front strips and the back strips of the double-sided detector as EF and EB respectively. For the single- sided second detector, only the label EF is used. The close-packed design of the telescopes right next to each other with a minimum dead area required the development of a highly flexible flat printed circuit board cable connecting the silicon strip detectors with the pre-amplifier housings. Figure 3.11 shows the picture of a double-sided silicon strip detector with the flat printed circuit board cables. The combinations of the silicon strip detectors are listed in Table 3.2. The reason for choosing 1.0 mm silicon detector for telescope 2, 4, 5, and 9 is that the deuterons emitted at smaller angles have lower energies and will be stopped inside the 1.0 mm silicon detectors. The deuterons emitted at larger angles with higher energies (E>10.8 MeV) would punch through the 500 um silicon detectors and be stopped in the CsI(Tl) crystals. One advantage of silicon-strip detectors is their position resolution. For the double- sided detectors, we can use the orthogonal strips on EF and EB to obtain each particle’s (x, y) pixelwise position. When the 1.0 mm detector is used, its strips are oriented orthogonally to the strips in the DE silicon detector to provide a two-dimensional 96 position. As the strips are 3.1 mm apart, the 50mmx50mm lateral dimensions of each telescope are divided into 256 (16x16) square pixels with a resolution of 3x3 mmz. At a distance of 205.3 mm, the angular resolution of the pixel is 0.43 deg. Telescopes 4, 5, and 7 were placed at a greater distance of 392.6 mm and their angular resolution is 0.22 deg. All the position and angular information is summarized in Table 3.2. 3.3.3.2 Energy Calibration One advantage of silicon detectors is their linear and largely particle independent energy response. In this experiment, the relevant deuteron energies range from 9.0 MeV to 20 MeV. The silicon energy response in this range is very linear. However, due to the nonlinearity of electronic system including preamplifiers, shapers, and ADCs, energy calibration must be performed. A BNC (Berkeley Nuclear Co.) precision pulser generator was used to calibrate the silicon detectors. The pulser has a group of attenuation switches to change the amplitude of the output signal. Three attenuation settings were chosen corresponding to three different dynamic ranges. An absolute calibration was obtained from the measurements of 228Th or source for these three settings. The linear relation between the pulser dial value and its equivalent energy was obtained: E=a~W+b (3.3.1) where E is the equivalent energy of the pulser signal in the units of MeV, and W is the dial value in the Volts. The values of a and b for the three different settings, as listed in 97 Table 3.3, show that the output of the pulser is not strictly proportional to the dial voltage and the offsets are not zero. Right after the experiment was finished with all the electronics setup intact, the calibrated pulser signals were sent as inputs into the preamps of each strip. Then one-by- one, the pulser-calibration was carried out for all 352 silicon channels. The signals were read by the DAQ program and analyzed with the analysis program SMAUG. A linear fitting was performed to the channel reading C and the energy E converted from the pulser dial value W by Equation 3.3.1. The relation between C and E was obtained for each strip: E=gi ‘C‘l‘hi (3.3.2) where i stands for each strip. Equation 3.3.2 was used to convert the channel readout from each strip into particle energy in units of MeV. There is a total of 352 calibration curves for the Si-strips. As an example, Figure 3.12 shows the calibration curve for the No.6 strip of the EF detector in telescope 3. 3.3.3.3 Particle Identification In a heavy ion collision, many kinds of particles like protons, deuterons, tritons, and other fragments from the projectiles are detected in LASSA detectors. Since we are only interested in the deuterons, a particle identification (PHD) must be performed to distinguish the other particles. The PID can be performed by a combination of AE and E detectors. 98 For particles passing through a detector, the energy loss is approximated by the Bethe formula [Bar96]: kAZ2 ~ E AE dx (3.3.3) where dx is the detector thickness, k is a proportional constant, A is the mass number, and Z is the atomic number of the particles. For a fixed dx, a plot of AE versus E yield a family of contours with AE cc l/E . Each line corresponds to an integer value of Z and A. Figure 3.13 shows the AE-E spectrum of telescope 7. The x axis is the particle energy deposited in the EF strip detector; the y axis is the energy loss deposited in the DE strip detector. By vetoing the particles stopped in CsI(Tl) crystals, we were able to separate the deuterons and protons as well as tritons. 3.3.3.4 Position Calibration The position determination of the emitted deuterons is critical in this experiment. To determine the position of each pixel of the telescope accurately, we need to perform the position calibration. The angular position (61,6) of the center of each telescope was determined optically by using a system composed of a optical telescope and a mirror as shown in Figure 3.14. The mirror was mounted on a turntable which can be rotated in both horizontal and vertical planes and the rotation angles can be read from the turntable. The center of the mirror was placed at the center of the target in the experiment. The optical telescope was mounted in the beam line. The mirror was rotated until the center of each detector is visible and aligned with the optical telescope, then the angular positions 99 (a, ,6) of the detector center are: a = 2 - a0 13 = 2 - flo (3.3.4) where do and ,60 are the angles of the mirror rotated in the horizontal and vertical planes. The rotation angles can be read to the accuracy of 0.01 degree so the accuracy of the measurement is 0.02 degree. The angular position of the center of each detector is listed in Table 3.2. The coordinates 7(x. y. Z) of each pixel of the LASSA telescopes are obtained by: f(x. y, z) = R1(a)-R,1(fl)- fo(xo.yo.zo) (3.3.5) where 20 is the distance between the target and the center of the detector; x0 and yo are the vertical and horizontal distances between the pixel and the center of the detector, respectively; R,C (a) and Rv(,6) are the rotation matrices along the x axis and y axis. They are defined as: 0 0 O Rx(a) = 0 cosa sina 0 — sin (1 cos a cosfl 0 - sin ,6 Ry (,5) = 0 O 0 (3.3.6) sin ,6 0 cosfl We performed the calibration before and after the experiment. The positions did not change during the experiment. 100 3.3.4 CsI(Tl) Crystals 3.3.4.1 Detectors The CsI(Tl) crystals are produced by Scionix [Scio]. A non-unifonnity in light output of CsI(Tl) crystals better than 1% was obtained via crystal selection and a quality control procedure [Mic99, Tan02]. Each crystal is trapezoidal in shape as shown in Figure 3.15, and the length of the crystal is 6.0 cm. The front and back areas are 2.64x2.64 cm2 and 3.38x3.38 cmz, respectively. To allow compact packing, the sides between adjacent crystals are at right angles to each other while the sides next to the frame are cut at an angle of 7.09 degrees. Each crystal is wrapped with two layers of cellulose nitrate on the outer surfaces (next to the frame) and one layer on the inner surfaces. One layer of aluminized mylar foil (0.15 mg/cm2 mylar + 0.02 mg/cm2 Al) is inserted between adjacent crystals to ensure optical isolation. Each crystal is optically coupled to a clear 1.0x3.5x3.5 cm3 acrylic light guide with optical epoxy BC600 [Bicr]. This light guide is in turn optically connected to a 2.0x2.0 cm2 Hamamatsu S3204 photodiode [Hama] with clear silicone rubber compound RTV615 [Gene]. To prevent light leak and cross-talks between adjacent crystals, the outer sides of the light guide and the photodiode are painted with a reflective white paint BC620 [Bicr]. To reduce the noise level, the charge-sensitive preamplifiers are mounted right behind the crystals to reduce the length of the input cables and minimize the capacity input to the 101 preamplifiers. There are two motherboards in one telescope. Each one hosts two preamplifiers. Fig 3.9 shows the internal mounting structure, the outside of the detector box is Open and the two silicon strip detectors are placed on the side. An aluminum mylar foil covers the top of the wrapped crystals to maximize light reflection and improve the energy resolution. One motherboard of the preamplifiers can be seen under the crystals. Three major precautions are taken to reduce the cross-talks between the preamplifiers of the CsI detectors. The first is to place a grounded copper shielding between the two motherboards to minimize broadcasting; the second is to put a 110 Q resistor on the test input line connecting the two preamplifiers on the same motherboard to terminate each amplifier; the third is to use shielded coaxial cables instead of twisted—pair cables to reduce the broadcasting between cables. With this set up, the cross-talks are reduced to the level of 0.1% [Mar98]. 3.3.4.2 Energy Calibration The fluorescence emitted by the CsI(Tl) scintillator has two major components of a fast (500 ns) and a slow (7 us) decay time constants. The relationship of light output and energy is mass and charge dependent. Therefore the CsI calibration cannot be performed by different kind of particle, neither by pulsers. In addition, the light output of a CsI crystal also depends on the T1 doping of CsI crystals. Since every CsI crystal may have different doping during manufactory, it is necessary to perform calibration for each CsI crystal individually. 102 For heavy ions at low energy, the light output of a CsI crystal shows non-linear response to the deposited energy [Lar94, BirSl]. However, for the isotopes of hydrogen, Tan [Tan02] and Handzy [Han95] found that linear functions result in good fitting. However, previous calibrations did not extend deuteron calibration below 20 MeV, so the deuteron response function for the CsI crystals was not known in our energy region of interest. Calibration of the CsI(Tl) crystals was achieved by the reaction of p(l3C,d)12C. 13C is the primary beam with high beam density. The emitted deuterons corresponding to the ground state and the first excited state of 12C can be identified clearly in the energy spectrum. As the deuteron scattering angle is known from position calibration of the pixels, the deuteron energy is obtained by kinematic calculation. As shown in Figure 3.16, the emitted deuteron goes through target, SnPb foil on the window of telescope, DE silicon strip detector, and EF(EB) silicon strip detector before being stopped in C51 detector. The deuteron energy deposited into the CsI detector is: ECsl = Ed — AEtar ’ AESnPb ‘ AEDE — AEEF (3-3-7) where Ed is the emitted energy of the deuteron from the target determined from kinematics, AEm, and AEsnpb are the deuteron energy losses in target and SnPb foil respectively. These energy losses are obtained using the program ENLOSS [Enlo] according to the material components and thickness. AE DE and AEEF are the energies deposited into DE and EF(EB) silicon strip detectors. Then the channel readout CCsl from €81 detector can be calibrated to ECsI by a linear fitting: ECsl =41 'CCsl +101 (333) 103 where 1 stands for each CsI crystal. This equation was used to convert the readout of the CsI detectors into particle energy in units of MeV. Figure 3.17 shows the calibration curve for the No. 3 crystal in telescope 3. Clearly, the linear fitting works very well for the deuteron calibration from 1 MeV to 14 MeV. This result is consistent with the observations of Tan [Tan02] and Handzy [Han95]. Our fitting results in a precision of the calibration better than 2%. 3.3.4.3 Particle Identification For the particles that stopped in CsI(Tl) crystals, the particle identification can be performed by the combination of silicon strip detectors and CsI(Tl) detector. Figure 3.18 shows the AE-E spectrum of telescope 3. The x axis is the particle total energy, including the energies deposited in DE, EF(EB) silicon strip detectors and CsI detector; the y axis is the sum of 4 times the energy loss in DE detector and the energy loss in EF(EB) detector. The deuterons and protons are well separated. 104 Table 3.2: Geometric setup of the telescopes and the configurations of the silicon strip detectors. Thickness Thickness of of EF/EB Telescope 9 or B Dist. Angular DE silicon No. (deg.) (deg.) (deg.) (mm) resolution silicon strip strip detector detector (um) (pm) 1 21.5 -14.2 -l6.3 205.3 i0.43° 67 480 2 14.2 -14.2 0 205.3 i0.43° 68 978 3 21.5 -l4.2 16.3 205.3 i0.43° 64 500 4 7.0 0 -7.0 392.6 i0.22° 64 913 5 7.0 0 7.0 392.6 i0.22° 65 982 6 27.9 23.0 -16.3 205.3 10.43° 67 481 7 16.7 0 16.7 392.6 i0.22° 66 476 8 27.9 23.0 16.3 205.3 i0.43° 70 482 9 23.0 23.0 0 205.3 i0.43° 66 993 Table 3.3: The calibrated parameters in Equation 3.3.1 Attenuating Maximum energy a b setting range x 2 30 MeV 5.0592 01081 x 5 16 MeV 2.0347 -0.0933 x 20 3 MeV 0.5115 -0.l223 105 Figure 3.9: Structure of LASSA telescope. 106 A X Entrance of S800 Figure 3.10: Schematic of the geometric setup. 107 Figure 3.11: One double-sided silicon strip detector with the flat printed circuit board cables. 10 m ' ' ' ' T E = 0.0021C-0.0265 S‘ 0 a DJ 5 - -l 0 . . . . 4L . 4% 0 2000 4000 Channel Figure 3.12: Calibration curve for silicon strip detector, by which the channel readout of the silicon strip detector is converted to particle energy in units of MeV. This curve stands for the No. 6 strip of EF detector in telescope 3. 108 EDE (arbitrary unit) 100 l | t 1 0 0 E5]: (arbitrary rmit) 150 200 Figure 3.13: Particle identification using the energy spectrum of BBB vs. EEF for telescope 7. 109 Detector Telescope Mirror Eve Telesmpe \ \ <1 ‘ [I] i /— ll 'I‘m‘ntable Figure 3.14: Geometry calibration system composed of one optical telescope and a mirror mounted on a turntable with two orthogonal axes that rotate in horizontal and vertical planes. The center of the mirror is the position of the target in the experiment and the optical telescope is mounted in the beam line. a= 2.64 cm Figure 3.15: The shape of CsI(Tl) crystal. 110 13C beam CH2 target Figure 3.16: Schematic of the CsI calibration. The deuteron emitted angle is determined by the pixel on DE and EF(EB) silicon strip detectors. The deuteron emitted energy is obtained by kinematic calculation. The deuteron deposited its energy into the CsI crystal after going through target, window foil, DE and EF(EB) silicon strip detectors. lll 15 E = 0.0188C - 0.1777 E (MeV) 0 l 1 fi 0 200 400 600 800 Channel Figure 3.17: Calibration for CsI detector, by which the channel readout of the CsI detector is converted to particle energy in units of MeV. This figure shows the calibration for the No. 3 crystal in telescope 3. 112 5‘5 200 .— E. E. 150 - ha m [:1 + 100 — a, '5‘ '9'. a I: ' “nigh _ _ .39.“. 50 “- -J _ ' H . .. p 3 ' 331.913: 9‘. - .- H 0 I r l " l" 0 50 100 150 200, Brant (arbitraryrmit) Figure 3.18: Particle identification in the energy spectrum of 4E DE + E EF vs. Emm, for telescope 3. 113 3.4 8800 Spectrometer 3.4.1 Overview The working principle of magnetic spectrometer is following: A particle with charge q and mass m, traveling at speed v, passing through a uniform magnetic field with strength B, will travel in a circular path with radius p given by m3: 8,0 (3.4.1) q Relativistically, the mass is m, where m is the rest mass and y is the Lorenz transformation factor. Thus, for a given magnetic field setting, particles with identical momentum to charge ratios are deflected the same amount by the magnet. A schematic of the 8800 spectrometer is shown in Figure 3.2. It stands behind the target chamber and consists of one quadruple doublet, two dipoles, and one focal plane detector. The advantages of S800 spectrometer are the high energy resolution and large solid angle acceptance [Zha97, Yur99, Cag99]. Some of notable characteristics of the S800 spectrometer are listed in Table 3.4. Figure 3.19 shows the schematic of the focal plane detector of 8800. It consists of two Cathode Readout Drift Chambers (CRDC), one ion chamber, and four plastic scintillators. The CRDC detectors measure the two transverse positions and angles of the particles; the ion chamber measures energy loss in the gas; the plastic scintillators measure the particles energies. The particle flight time is measured relative to the cyclotron radiofrequency (RF) pulses. Different species of particles emitted from the reactions have different velocity, 114 and hence different flight time to the focal plane. This flight time measurements can then be used in conjunction with the energy loss measurements or total energy measurement to identify the particle species that arrive at the focal plane. 3.4.2 Cathode Readout Drift Counters The CRDC detectors have an active area of 30 cm x 59 cm and an active depth of 1.5 cm. They are filled to a pressure of 140 Torr with 80% CF4 and 20% C4H10. Figure 3.20 shows a schematic illustrating the principles of their operation. Ions traveling through the gas create ionizations. A constant vertical electric field in the detector move the electrons toward an anode wire, where charge amplification takes place in the high electric field close to the wire. The anode wires are placed below a ground Frisch grid and held at a constant voltage, typically 1400 Volts. The electrons are collected on the anode wire. Cathode pads are Tocated in front and back of the anode wires. The charges collected on the anode wire induce positive charges on the cathode pads. There are 224 pads in each CRDC detector. The centroid of the Gaussian fit to the charge distribution is used as the horizontal position in the detector. The vertical position is determined by the drift time of the electrons to the anode wire. The typical drift time of the electrons to the anode wire is 0-20 us, depending on their vertical position. Measuring the time between the scintillator signal and the anode wire signal provides a direct vertical position measurement of the particle track. Masks with well-defined holes and slit patterns as shown in Figure 3.21(a) are placed in front of the CRDC detectors to calibrate the detector positions. Figure 3.21(b) is the 115 position spectra taken with 10Be beam with the mask placed in front of the first CRDC detector. The position resolution of 0.2 mm is achieved. 3.4.3 Ion Chamber Immediately following the CRDCs, the beam particles pass through an ionization chamber. The ion chamber (IC) used in the S800 is a standard Frisch grid ion chamber [Yur99]. It is designed to measure the energy loss as the beam particles ionize the gas in the detector by sampling the signal generated along sixteen anode strips. The gas used is P10, which is composed of 90% argon (Ar) and 10% CH4 (methane). The energy loss in the ion chamber combined with the time-of—flight or the energy deposited in scintillator detectors can provide particle identification. 3.4.4 Plastic Scintillators There are four plastic scintillators in 8800 spectrometer. They are made of BC-408 scintillant plastics manufactured by Bicron [Yur99]. In the order from first to last, with respect to the beam direction, the scintillators are labeled as E1, E2, E3, and E4 in Figure 3.19 with the thickness of 3 mm, 5 cm, 10 cm, and 20 cm respectively. Light guides are mounted on each end to enhance the collection of the light in the photomultiplier tubes (PMT). The light travels through the plastic as well as the light guide and is collected in the PMT’s on the top and bottom ends of the scintillator. The energy deposited in the scintillator is calculated by: 116 E, = J 15,-”? x EiDOW" (3.4.2) In the meanwhile, we get the number of particles that enter into S800. We performed normalization run at the beginning of each kind particle beam, when the targets are moved out of the position and the beam particles enter 8800 directly. We measure the beam transfer efficiency by the ratio of particle number in 8800 to the particle number going through BLT2 scintillator. The normalization procedure will be introduced in Section 4.3. Since both energy degraded beam particles and the residual particles from the (p,d) reaction enter 8800 spectrometer simultaneously, we need to separate them out by the combination of deposited energy in E1 vs. time-of—flight. As an example, Figure 3.22 shows the spectrum of the deposited energy in E1 versus the particle time-of-flight for the reaction of p(“B,d)loB. The residual particle of '0B is separated from the incident beam ll of B. 3.4.5 Summary Originally, we plan to use S800 spectrometer to detect the recoiled residual nuclei in coincidence with the deuterons detected by LASSA detector to perform complete kinematic measurement. Based on the above analyses, 8800 supplies excellent particle position determination via CRDC detectors and good particle identification via the combinations of energy loss in ion chamber, energy deposited in scintillators, and time- of—flight of particles. Unfortunately, there were errors in writing the data from 8800 spectrometer onto the tape in this experiment so that some 8800 data were lost. In the 117 present work, the S800 spectrometer was used only for the normalization, when the data from 8800 are complete. 118 El. E2. E3. E4 Scintilators \l 4,41. 1 . ' III 4. . , lon Chamber l/Q’RM‘ #2 Figure 3.19: Schematic of the focal plane detector of S800 spectrometer. It consists of two CRDC detectors, one ion chamber, and four plastic scintillators. 119 Table 3.4: Characteristics of the S800 spectrometer AE _ Energy resolution 7?— =10 4 AP - Momentum resolution 7; = 5 x 10 5 Energy range 11.6 % Momentum range 5.8 % Solid angle 20 msr Angular resolution 52 mrd Horizontal detector . 0.3 mm resolution Vertical detector . 0.3 mm resolution Maximum rigidity 4.0 T-m Maximum dipole field 1.42 Tesla 120 .1. ‘ lllul 1' - ls-~i~i-s-sl~\1 1§R§N§§§§N§S§§§N 1....l~.x‘..\‘x....\.1|\\u.\l..\l.\ .1.‘ .x, ..s .x .x. ..c. .\ 1.. ..4 .,. ....\. ~§§\ _ x... 1.. . \ . \x . .. ..1 x .\ 1... x. 1.. x \ x \ ..x \1 n! I . . 1 . . .1 x. 1 a .. u \ ~ \ 1 \ a —. \ )0 x xx . x . . . 4 as. . . \ K .2. . x. .. .- ~ . .\ v .. ~ 4 a! \ _. .. x . \ \. . x .. .. . .. . x . .\ ll/ 4. . .~ . .. .. x x. .\ .. H \ _ . H 1.. . . . w. x . . ‘ \x I . 1. \ R . .. . . . 4 .x \ \ .‘ ax 1.. ..r'sll.\ ..‘.\.l!. '. x i Anode wire Figure 3.20: Schematic of the CRDC detector. A particle ionizes the gas as it passes through the detector. The electrons drift to the anode wire where they are collected. The induced image charges on the cathode pads provide horizontal position information. The drift time of the electrons to the anode wire provide vertical position information. 121 0.10 A bi'l'l'llTr IUIIIIIIIIITI IIIIIIIUIIITUI UTTITITII'IT' WI .1 12 E ° 3 2 0.05 —- ° . . — .5 I o o c j 1: I 3 3 Z :3 0.00 -_— o o o o o o o o o o .1: 3. : ° 3 {>4 “0.05 [_- ° 0 o T. _0_ I 0 :r_Lu11_L_r..r_r_r_Lr. W .r.r..r_LL_r_r_L.Lr_r_r_r_rur_r_r_1_‘. -0.3 -0.2 -O. I 0.0 0.1 0.2 0.3 X position (meters) (8) 0.05 0 0 t . t ‘ . E. . ‘ ' . l l l -0.05 r 1 1 1 -0.15 —0.l -0.05 0.0 0105 0.1 0.15 0.2 (b) Figure 3.21: (a) Patterns on the mask. (b) Position spectrum of the mask placed in front of the first CRDC detector. 122 CE __ 5 E” .E 140 — .fl 3 E _ B H 100 —— 60 | l l l l I 120 140 160 200 Time of Flight (arbitrary unit) Figure 3.22: Spectrum of the energy deposited in E1 scintillator versus the time-of—flight ll 10 . for the p( B,d) B reaction. 123 3.5 Data Acquisition Electronics Figure 3.23 is a schematic of the electronics used in this experiment. The signals from the up and down PMT’s of the first scintillator El are sent to constant fraction discriminator (CFD) module. The outputs of the CFDs are AND-ed to give the S800 premaster signal. The S800 premaster signal and the CFD outputs of the anode wires of CRDC detector are used as the start and stop for the drift time in the CRDCs. The TAC output is input to module of analog-to-digital converter (ADC). The cathode pads are read by the fast encoding and reading ADCs (FERA). The gate for the FERAs is given by the AND of the 8800 premaster and the anode pulse. The signals from the silicon and CsI(Tl) detectors are digitized in Phillips Scientific peak-sensing ADCs (7164H). The signals from the EF silicon strip detectors are sent to Shaper-Discriminator-TFC dual modules. For this module, the shaper outputs are sent to ADCs; the TFC outputs are sent to Lecroy 4300B fast encoding and reading ADCs (FERA) to give the time signals; the trig outputs from all 9 telescopes are OR—ed to give LASSA Premaster signal. The LASSA Premaster will be AND-ed with 8800 premaster to give coincidence Premaster signal. The LASSA Premaster will also be delayed and downscaled to give LASSA trigger signal. The Master signal is logically AND-ed with the Busy signal from the computer, coincidence Premaster, S800 premaster, and LASSA trigger. The Master signal is the start signal to the computer and stop signal for the TFC. The gates for the modules of ADCs and FERAs are also supplied by the Master signal. 124 Down scale {Delay ELUP L Coincidence E1_Up ~ CPD ssoo LASSA Premaster Pr emaster 0 Pre mas ter (AND) :9: (AND) 3 ELDown CFD “‘ E LASSA Trigger ELDN Busy start 3801 TAC ADC / stop TAC CRDC Anode —{>— CFD H I ‘ AND gate FERA Pads Cathode Pad FD" Shaper in out Master : ‘ FIFO DGG Start To LASSA TF C l\\ CsI Detectors l/ Shaper ADC 1\ DE Detectors L/ Shaper ADC l\ {— EB Detectors V Shaper ADC Shaper j BF Detectors —{>_ Disc ADC TFC 5 3 Disc FERA Master Stop Trig "I :: LASSA _ Down scale __ LASSA ‘ —p OR Premaster {Delay Trigger —I~ —' T c ' T o om. o Premaster Master Figure 3.23: Schematic of the electronics. 125 CHAPTER 4 EXTRACTION OF AN GULAR DIFFERENTIAL CROSS SECTIONS 4.1 Overview This chapter will discuss the extraction of the deuteron spectra and analyze the contributions to the energy resolution in Section 4.2. In section 4.3, the procedure to extract the differential cross sections will be introduced and the measured data are presented. 4.2 Deuteron Spectra Applying the PID gates obtained in Section 3.3, we can pick out the deuterons and obtain their energies in laboratory frame via Equation 3.3.2 and Equation 3.3.8 for silicon strip detectors and CsI(Tl) crystal detectors respectively. The deuteron energy in the center of mass is obtained by converting the measured deuteron energy in the laboratory frame to the center of mass frame. EC," = %mV2 + -:—mV02 - mV vO cos a (4.2.1) where m is the deuteron mass, V is the deuteron velocity, V0 is the velocity of the center of mass, and 6 is the emitted angle of deuteron. Figure 4.1 shows the deuteron energy spectrum of the p(l3C,d)12C reaction at the laboratory angle of 19°. The peaks of ground 126 state (0+) and first excited state 4.439 MeV (2+) can be distinguished clearly. The peaks at 7.654 MeV (0+) and 12.71 MeV (1+) do not have enough statistics but still can be identified. The peaks at 15.11 MeV (1*) and 15.44 MeV (2") cannot be resolved completely. To estimate the full width at half maximum (FWHM) of the peaks in the laboratory frame AEL, we need to take into account that the emitted deuterons emitted from the reactions have to go through the remainder of the target and the SnPb foil before reaching the Si detectors. AEL = ‘55,," 2 + AESnPb 2 + A592 (4.2.2) where AEtar is the rms width of the deuteron energy loss distribution in the target, AESnpb is the deuteron energy straggling in the SnPb foil, and the AEB is the kinematic broadening due to the angular resolution of the strip detectors. AE.ar is larger than the width given by energy loss straggling because of the variation of energy loss in the target depend on how much of the target is traversed before the reaction occurs. The beam broadening and the beam straggling in the target are not included in Equation 4.2.2 because they contribute little to the deuteron resolution. The FWHM in the center of mass derived from Equation 4.2.1 is AEcm =AEL—mV0cost9-AVL +mVLV0 sin 6-A6 (4.2.3) where A0 is the angular accuracy of pixelation (i0.l°), AVL is the FWHM of deuteron velocity in laboratory frame. Table 4.1 lists the contributions to deuteron resolution. The energy straggling are calculated by the program SRIM [Srim]. The energy broadening 127 AEe due to the angular resolution and the energy loss in the target contribute most to the final energy resolution. The experimental FWHM for the peaks of 0.0 MeV and 4.439 MeV are 800 keV. The estimated energy resolutions are pretty close to the measured OIICS. Figure 4.2 shows the deuteron spectrum of the p(“B,d)loB reaction at the laboratory angle of 13°. The energy resolutions are about 640 keV for the ground state and the first excited state at 1.74 MeV. The other states cannot be evaluated because of low statistics. From Table 4.1, we see again that the kinematic broadening due to the angular resolution and the energy loss in target contribute most to the energy resolution. A smaller kinematic broadening and a thinner target will result in a better energy resolution. Figure 4.3 shows the deuteron spectrum of the p(lOBe,d)9Be reaction at the laboratory angle of 15°. The energy resolution is 800 keV for the ground state. The excited states cannot be distinguished because of low statistics. Contributions to the energy resolution are also listed in Table 4.1. 4.3 Extraction of Angular Differential Cross Sections Before the extraction of the angular differential cross sections, we need to know the beam transfer efficiency f and the total beam particles Npar that hit the target. The beam transfer efficiency is measured in a normalization run at the beginning of each experiment, where the target is moved out of the beam line. The beam particle delivered before the target is measured by the BLT2 scintillator and the beam particle through the target is measured by the El scintillator of S800 spectrometer. The beam transfer efficiency is: 128 .1 Na Nam 'RLT f (4.3.1) where RLT is the life time of the data acquisition system. The total beam particles that hit the target Npar are calculated by the summation over all the runs: Npar :ZNBLTZI 'RLTl 'f (4.3.2) I where N 3112'. , and R [1" are the number of particle detected by the BLT2 scintillator and the life time of the data acquisition system for each experimental run, i. The angular differential cross section in the laboratory frame is obtained for each (do) i Nd — = (4.3.3) d9 6L dQL'Ntar 'Npar telescope individually: where i denotes individual telescope, BL is the angle in the laboratory frame, Nd is the number of deuterons detected in the interval of il.0° relative to BL, dQL is the solid angle in the laboratory frame, and NW is the target thickness in number of hydrogen atoms per centimeter square. The statistical error for each telescope is calculated by A [do I _ do I 1 do 6L do 6L ,/ Nd (4'34) The average angular differential cross section and statistical error are obtained by (L0) 1’ 2 d9 6L . 2 A(LU) 1 (dd) _ dQ 6 L 6L _ 214311.104 L" .1 1 .. A(dfljflz Z[A[do) i]_2 (436) (4.3.5) 129 The angular differential cross section and statistical error in the center of mass are [12) _.(iz) d9. 6,", do 6L A (fl—9) = y - A (£9) (4.3.7) d9 60,, d9 19L 1+ flcos 6’6," m2 y : 2 2 m m 1+ ——1—2— + 2—l—cos66m "12 m2 where y is the ratio of dQL / dflcm , m1 and m2 are the mass of projectile and target nuclei. For the p(l3C,d)]2C reaction, the angular differential cross sections to ground state and first excited state at 4.439MeV have been extracted. For the reaction of p(“B,d)mB and p(loBe,d)9Be, only the angular differential cross sections to the ground state have been extracted because of the low statistics of the excited states. The data and the statistical errors are listed in Table 4.2. The Open red symbols in Figure 4.4 show our measured angular differential cross sections of p(13C,d)12C (g.s.) reaction. It is compared to the published data of Ref. [Cam87] at proton energy of 41.3 MeV (solid red circles). There are additional data in Ref. [Sco70] at proton energy of 50 MeV. However, the latter set of data was published in arbitrary unit. We match Scott’s data [Sco70] at 121° to the data of this measurement and get the normalization factor of 2.45. The three sets of data show fairly good agreement especially when the difference in beam energies is taken into consideration. 130 Figure 4.5 shows the comparison of the differential cross section of p(l3C,d)12C to the first excited state from our measurement (open circle), data of ref. [Cam87] (closed red circles), and data from ref [Sco70] (diamonds) with the same normalization factor of 2.45. Again, our data agree with the past measurements fairly well suggesting that the experimental procedures we used for measuring angular distributions for inverse kinematic reactions with high resolution strip detectors work rather well. Figure 4.6 shows the angular differential cross sections of p(”B,d)lOB (g.s.). Unfortunately, the 11B data were taken with relatively short time so that the total statistics we have collected are low. Only telescopes l, 2, 3, and 7 yield significant counts to the measurements. The measured data have large error bars. Figure 4.7 shows the angular differential cross section of p(loBe,d)9Be to the ground state. However, we have problem with the absolute normalization. In this particular reaction, the 8800 trigger some time did not fire. When that happens, the LASSA trigger fired alone but with a downscale factor of 5. We have to add the events by S800 trigger together with 5 times of the events by LASSA trigger. This problem only happened in the beam of 10Be. We still do not understand the reason of this problem thus there are unresolved questions about the absolute value of the cross section. 131 30 l * r I r'f)‘ 2 4 i 3 2 <13 >0; i g) B 20— 2L2 2 a S 1 $ 0. 0 mi 0 c.) ._. 3 2 5 l i g 2 1 10— oi 8 — *" Z¥ 6.1111111111111111”. 1 1100111071 iirni. .n‘ 24 3 16 2 40 E d(MeV) Figure 4.1: Deuteron energy spectrum of the p(l3C,d)12C reaction at the laboratory angle of 19° measured by telescope 7. 132 30 ' l r l r l " > 0 2 O. > o m _ _ a 20 1 9 l a 9 as — o is 2.4 ‘13 sl" L9. to) co "3 10- ‘ m _ Oat—L11 Ilfl 1 1 1iflfl r0 16 20 24 28 32 Ed(MeV) Figure 4.2: Deuteron energy spectrum of p(1 lB,d)lOB reaction at the laboratory angle of 13° measured by telescope 7. 133 — 3011111 B 20 — — C. . :3 o C.) > g >> go 22 10 7 1‘. LOCI) — _ (0 OD- F‘— l _ 11 :AM 00101111111111.411111111 11111111111 0111111 16 24 4O Ed(MeV)2 Figure 4.3: Deuteron energy spectrum of p(lOBe,d)9Be reaction at the laboratory angle of 15° measured by telescope 7. 134 m com os ‘va s: 2 SN of ad A18 od $2 0% gm SN 2: ow v2 3: 86 be e: 9% :v 8N m2 mm m: :2 od be od m: 8w 0% 2m m8 ow RN Ci :5 8 3% 8w NS Sn 5 mm mom o2 od re ed on $8: 98: Se: 99: $8: $8: 93 a; m4 .8 m4 1. m4 a m4 =3 m< 5; m4 c 932V 39: mo 888 3.2: we 5E8 088“. .93 choumos =8 pmcm Ewes E .3 a £23 .33 88m 52:? 52:3.» 535% gem awczwwgm aolwscm _23 ©2332 3385mm noumESmm 35cm .mosanEx oEo>E E 20:88 @335 5:50: BwEm 2.: go cons—88 $55 of 9 32:58:00 :6 2an 135 Table 4.2 Experimental angular differential cross sections and statistical errors (a) p(l3 C.d)12c. Ex: 0.0 MeV (10' da' _ A _ 6cm d9, d9 (deg) (mb/ sr) (mb/sr) 4.0 1 1.54 1 .07 5.2 11.48 1.19 6-4 1 1.88 1.59 7.7 1 1.09 1.80 8.9 10.07 0.58 10.2 9.60 0,49 1 1.6 8.70 0.61 13.0 7.85 0.45 14.5 6.68 0.41 16.1 5.39 0.37 (c) p(' ‘B.d>‘°B. Ex: 0.0 MeV do' do _ A _ 6cm d9 dg (deg) (Mb/ Sf) (mb/ sr) 6.8 6.12 1,49 3-4 6.73 1.28 10.0 7.59 1,27 1 1.7 8.44 1.16 13.5 7.45 1.2 15.4 6.84 1.25 136 (b) P 2‘ g? I E E O 'I— I 4— “ :-——-*———_1__—————————————:I_ ———————— 4 H 1.0—IT -1- T O '1‘ “ .2 ”53:9. 4 9 i E 5 ‘ g, 4.2 I T 3 T CH 5 L 8 .———2—‘2—;-£———-—-—_% ——————————————— . 8 10:733— — o e — 4 H —————— a —_,— — ~— 4:; ______ 2 _ .. .1:— CD 0 5— I 2'. I 2 9 - Cg, ' -- -— JLM OO . l L . 4 . l . . . i l . . i i 1 1 . . . l . . . . 10 20 30 4O 5O 60 Ed (MeV) Figure 5.3: Extracted spectroscopic factors in the present work for 12C(d,p)13C (g.s.), 13C(p,d)nC (g.s.), and p(]3C,d)12C (g.s.) reactions. The dashed lines represent the shell model prediction of Cohen and Kurath [Coh67] of 0.62. See text for detail explanation. 150 4 . . 1 . . . . I . . . . I . . . . I 12C(d,p)13C Ed=4.5 MeV ? 20 b O: a — m 1. X: b 3 X c E x . g :1 O 1: x b 10 _ X O X — Pd 1- D 1 D o x x x E] 01210 CDC) 0 O O . . . . a . . . 1 1 . . L . 1 . . . . 1 . O 20 4O 6O 80 Gem (deg) Figure 5.4: Comparison of the existing measurements of 12C(d,p)wC (g.s.) reaction for deuteron energy at 4.5 MeV, a [Gur69], b [Gal66], and c [Bon56]. 151 3O . . r . T T l 12C(d,p)13C Ed~12.0 MeV 4 1: 1‘2- _ O: 11.8 MeV E 20 _ l: + [3+ X: 12.0 MeVa z E j T" +: 12.0 MeVb v 141 ,_ + 1:1: 12.4 MeV c: ._ ET 1:: ’ + \ “ 1 b -- 1 Pa 10 — my _ >11 13.; T“ X15 3%: >111) X5 E] [:1 O . . . i l l O 20 4O 60 Gem (deg) Figure 5.5: Comparison of the existing measurements of 12C(d,p)BC (g.s.) reaction for deuteron energies at 11.8 MeV [Sch64], 12.0 MeV a [Lan88], 12.0 MeV b [Sch67], and 12.4 MeV [Ham6l]. 152 1 . . . . 1 . . - 12C(d,p)13C Ed~15.0 MeV __ T __ O: 14.7 MeV E >< __ X: 14.8 MeV E X O X E]: 15 MeV V 6 9 C: «4 "d” E3 13 3 5 1 b 10 L— T T U 0 El T .l 0 i1:] [:10 CED [:10 C O L 1 L L r J r O 20 4O 60 9m (deg) Figure 5.6: Comparison of the existing measurements of 12C(d,p)nC (g.s.) reaction for deuteron energies at 14.7 MeV [Ham61], 14.8 MeV [Mcg55], and 15 MeV [Dar73]. 153 5.3 l3C(p,d)lzC (g.s.) and p(13C,d)12C (g.s.) Reactions Systematic analyses are performed to the measurements of 13C(p,d)lzC reaction to the ground state [Toy95, Cam87, Tak68, H0880] listed in Table 5.2. The proton energies range from 35 MeV to 65 MeV. Same parameters listed in Table 2.7 and same procedure as described in section 5.2 are employed. The data and calculations multiplied by the corresponding spectroscopic factors listed in Table 5.2 are shown in Figure 5.7. The angular distributions at 35 MeV, 41.3 MeV, and 65.0 MeV do not have data at forward angles. These data without the first peak may not give reliable SF. The data at 55.0 MeV have data at forward angles but the shape is different from that of the calculations. The extracted SF at 65.0 MeV is almost twice the expected value. Thus the data at 65.0 MeV may not be correct. The extracted SF from the data at 35.0 MeV, 41.3 MeV, and 55.0 MeV are plotted in Figure 5.3 as open circle points. As the existing measurements do not give reliable SF, a new measurement to cover the first peak in inverse kinematics is desirable. The data and calculations for the inverse kinematic reaction of p(l3C,d)12C (g.s.) performed in the present work are plotted as the third set of data (open points) and lines in Figure 5.7 The extracted spectroscopic factors, as shown in open squares in Figure 5.3, are 0.74, 0.91, and 1.18 for ADBA (JLM), ADBA (CH89), and DWBA calculations, respectively. One possible reason that the (p,d) reactions give higher spectroscopic factors than the (d,p) reactions in Section 5.2 is that the (p,d) reactions are performed at higher energies. The overall averaged spectroscopic factors from all the (d,p) and (p,d) reactions are 0.62i0.09, 0.761011, and 0.89:0.20 for ADBA (JLM), ADBA (CH), and 154 DWBA calculations, respectively. The ADBA calculations based on JLM potentials give the best result compared to theory. Again, the SF values from the DWBA calculations are higher. As the importance of the deuteron break-up effects has been demonstrated, we will not discuss the DWBA calculations in the remaining part of this chapter. 155 dU/dQ (mb/sr) F 1 r V T I I I r I 0 13C(p,d)lzc (g.s.) o p(13C,d)12C (g.s.) — 35.0 MeV Figure 5.7: Angular distributions for l3C(p,d)lzC (g.s.) and p(13C,d)12C (g.s.) reactions for beam energies from 35 to 65 MeV: solid lines present ADBA (JLM); dotted lines present ADBA (CH89); dashed lines present DWBA. The calculations have been normalized by the spectroscopic factors. Each distribution is displaced by factor of 10 from adjacent distributions. The overall normalization factor is 1 for the 65.0 MeV data. 156 Table 5.2: List of references and spectroscopic factors for the l3C(p,d)lzC (g.s.) and P(13C,d)12C (g.s.) reactions Reaction SF SF SF SF Ref (MeV) (Liter) (JLM) (CH89) (DWBA) ' 0.7 ”C(p,d)”mge.) 35 0.8 0.66 0.85 1.16 [Toy95] 1.0 ”a d)12C( s) 413 0'91 078 098 131 [Cam87] l3C(p,d)”c (g.s.) 55 0.82 0.66 0.82 1.05 [Tak68] 0.26 13C(p,d)12C(g.s.) 65 0.31 1.22 1.57 1.33 [H0580] 0.43 ( 3 0.70 0.88 1.17 °V°rag° i007 i009 :013 p(13C,d)12C(g.s.) 483* 0.74 0.91 1.18 * Equivalent proton energy 157 5-4 l3C(p,d)12C(2+) and p(l3C,d)12C (2+) Reactions Theoretically the transferred neutron in the reaction of l3C(p,d)lzC to the first excited state 2+ at 4.439MeV of 12C is predicted to be in pure lp3/2 orbit [Cam87]. The theoretical spectroscopic factor from Cohen and Kurath [Coh67] is 1.12. This clearly identified state provides another opportunity to test our strategy to extract the SF. Systematic analyses are performed to the existing measurements of l3C(p,d)lzC reaction to the first excited state [Toy95, Cam87, Tak68]. The proton energies range from 35 MeV to 65 MeV. The same procedure as described in section 5.2 is employed. The data and calculations for the present measurement of p(l3C,d)12C* (4.439MeV) are plotted as the third set (open points) in Figure 5.8 together with the data (closed points) from the literatures [Toy95, Cam87, Tak68]. Among the published data, only the data at 55.0MeV have reasonable coverage at forward angles. The data at 35.0 MeV and 41.3 MeV are not reliable since they did not include the first peak. However, the extracted spectroscopic factors by fitting the slope of these two data sets may provide consistent checks. The extracted spectroscopic factors are listed in Table 5.3 and plotted in Figure 5.9. The averaged spectroscopic factors from ADBA calculations based on JLM potentials and CH89 potentials are 0.92:0.09 and 1.08:0.13, respectively. The extracted SF for present measurement are 1.03 and 1.2 from ADBA(JLM) and ADBA(CH89), respectively. These values are listed in Table 5.3. The SF values from CH89 potentials are usually higher than that from J LM potentials. 158 W,,.... - 13c(p,d)1°c* (4.439 MeV) o p(130,d)120* (4.439 MeV) 35.0 MeV dU/dQ (mb/sr) l 0 20 40 60 em (deg) 0.1 Figure 5.8: Calculations for reactions of 13C(p,d)12C* (4.439MeV) and p(l3C,d)12C* (4.439MeV) [Toy95, Cam87, Tak68]; solid lines present ADBA (JLM); dotted lines present ADBA (CH89). The calculations have been normalized by the spectroscopic factors. Each distribution is displaced by factor of 10 from adjacent distributions. The overall normalization factor is 1 for the 55.0 MeV data. 159 Table 5.3: Extracted spectroscopic factors of the lp3/2 neutron from the l3C(p,d)lzC (2+) and p(l3C,d)]2C (2+) reactions. Proton SF SF Reaction energy ADBA ADBA Ref. (MeV) (JLM) (CH89) l3C(p,d)12C* (2*) 35 0.92 1.08 [Toy95] I3C(p,d)'2C* (2*) 41.3 1.01 1.2 [Cam87] '3C(p,d)l2C* (2*) 55 0.84 0.95 [Tak68] (average) (0.92i0.09) ( 1 08:0. 13) p('3c.d)'2c* (2*) 433* 1.03 1.2 * Equivalent proton energy 160 2.0? - a r r - r 1.5} -— z: T - CH T o C] s ------- s ————— 1 —————— - ------ T - — O 1.0j 4— o ‘ .53 ; 4‘ O “ 4 "-t 1.5 1 3* : JLM l 0 l. — e — e e e e H — -.s —————— e. ————————— . 8 l.0~ - s - L. f ET] AL 0 +2 l e 4 O r 13 12 :I: l a 0‘5? O C(p,d) C (4.439MeV) _ m g o p(13c,d)lzc* (4.439MeV) 0.0b . l l l l l l l l l A . l . 30 4O 5O 60 Ed (MeV) Figure 5.9: Extracted spectroscopic factors of the reactions ”C(p,d)IZC“ (4.439MeV) (circle) and p(l3C,d)12C* (4.439MeV) (square). The dashed lines represent the shell model prediction of 1.12 by Cohen and Kurath [Coh67]. 161 5-4 103&1,pr (g.s.), 11B(p,d)1°B (g.s.), and P(llB,d)l0B (g.s.) Reactions Systematic analyses are performed for the reaction of 11B(p,d)loB (g.s.) [Leg63] [Ku168] [81062] and its inverse reaction of ‘°B(cl,p)“B (g.s.) [Hin62] [Sch67] [Bar65]. The transferred neutron is in lp3/2 orbit and its form factor is plotted in Figure 2.23. The theoretical spectroscopic factor from the shell model is 1.09 [Coh67]. Applying the consistent procedure as described above, the ADBA calculations are performed based on CH89 and JLM potentials respectively. The modified harmonic oscillator densities (a=0.837 fm, a=l.7l fm for "’B; a=0.811 fm, a=l.69 fm for ”13) compiled in Ref. [Dej74] are used in the JLM potentials. The data and calculations are shown in Figure 5.10 for the (d,p) reactions and in Figure 5.11 for the (p,d) reactions, respectively. The ADBA calculations based on CH89 and JLM potentials give similar results. The spectroscopic factors have been extracted by fitting the first peaks as described in section 5.2. The results are listed in Table 5.4 and plotted in Figure 5.12. The spectroscopic factors from the published data present a trend with larger value at 30 MeV and lower value at lower (10 MeV) and higher (50 MeV) energies. Particularly, the measurement of llB(p,d)mB at incident energy of 19 MeV (not plotted) gave a much higher SF compared to other experiments. In general, the absolute cross sections increase with incident energy. However, instead of lower cross section, the measured cross sections at 19 MeV are nearly twice as large as the cross sections measured at 33.6 MeV and 44.1 MeV. Thus, we believe this data set has normalization 162 problems and disregard it. The average value of the extracted SF except the data at 19 MeV is 1.37i0.34 and 1.34:0.31 by using potentials of JLM and CH89 respectively. The theoretical angular distribution calculations for the p(“B,d)lOB reaction are shown in Figure 5.11 as open symbols. The spectroscopic factors calculated from ADBA based on JLM and CH89 potentials are 1.05 and 0.97 respectively. Due to low statistics, the data of this measurement have larger error bar. The statistical uncertainty of the extracted SF is 17.2% for both the JLM potential and CH89 potential. However we cannot determine systematic errors due to our concern about the absolute normalization of those cross sections. Some data from the 8800 scintillator are missed. Although the data for present measurements seem reasonable, we are not sure the normalization is absolutely correct. 163 do/dQ (mb/sr) E 10.1MeV f 103:— 0 —: : 12Mev ° 3 o -' d 2 .' 10 :— . —. ” ' o. i .. 13.5MeV , o '_ . .l . q 10; 28Mev o . ‘5 : ' _.-- "0 .70 : _._. g . 10 11 . B(d.p) B (g.s.) 01 r . I l . r . l . 4 O 20 4O 60 (deg) Figure 5.10: Calculations of loB(d,p)“B (g.s.) reaction at 10.1 MeV [Hin62], 12.0 MeV [Sch67], 13.5 MeV [Bar65], and 28 MeV [81062]: solid lines present ADBA (JLM); dotted lines present ADBA (CH89). The calculations have been normalized by the spectroscopic factors. Each distribution is displaced by factor of 10 from adjacent distributions. The overall normalization factor is l for the 28.0 MeV data. 9 cm 164 19.0 MeV 3 — s‘, \ i — L. . . U) E ° E 102 g‘ . x 38.6 -5 V : - i g no \ ' O O r b ”d 10 ": 1.0 E ...... . o 11B(p. , .......... .0 ........................... 7“ ......... D ..... . 0 1.— o ‘ CD : Q-I L. _ U? e O ’ 1 1 l O 20 4O 60 Ed (MeV) Figure 5.12: Extracted spectroscopic factors for the reaction of p(”B,d)10B (g.s.) (open square), llB(p,d)lOB (g.s.) (open circle), and loB(d,p)“B (g.s.) (closed circle). The dashed lines represent the shell model prediction of 1.09 by Cohen and Kurath [Coh67]. 167 5.6 9Be(d,p)lOBe (g.s.) and p(10Be,d)9Be (g.s.) Reactions Systematic analyses are performed on the angular distributions measured from the 9Be(d,p)mBe reactions that were published in the literatures [ZelOl, GenOO, Va587, Sch64, Dar76, And74, $1062]. The transferred neutron is in the 1pm orbit and its form factor is shown in Figure 2.23. The SF value obtained from the theoretical prediction of shell model is 2.35 [Coh67]. The ADBA calculations are based on CH potentials as described in section 5.2. The JLM potential is not used since the nucleon radius information of 10Be is not available. The experimental data and calculations are shown in Figure 5.13. The extracted spectroscopic factors are listed in Table 5.5 and plotted in Figure 5.14. The systematic studies do not give a consistent value of spectroscopic factor. The data from 6.0 MeV to 11.0 MeV came from one reference of [GenOO], which gave the spectroscopic factor value around 1.0 for deuteron energies from 7.0 MeV to 11.0 MeV. The other data yield spectroscopic factor values from 0.97 to 2.59. There is big difference in the measured cross sections at 15 MeV and at 15.3 MeV, which give spectroscopic factor of 1.83 and 1.19 respectively at nearly the same energies. The average value (to give the same weight for different systems, only one set from Ref. [GenOO] at 11.0 MeV is included) is 1.401041 with rather large uncertainty. The data and the calculations for the present measurement in the inverse kinematics of p(loBe,d)9Be are shown as open points in Figure 5.13. Our ADBA (CH89) calculation gives spectroscopic factor of 2.99, which is 27% higher than the theoretical value and nearly a factor of two higher than the values obtained from seven (d,p) reactions measured by different groups. Since we have problems in the absolute normalization as 168 discussed in Section 4.3, it is not clear if the discrepancies arise from problems in our measurements. 169 10 IYIUTIITII’Wfirr 'VIUIT'YIIWYUU YI'YIUIY'IYYTY ' 9Be(d,p)mBe (g.s.) 106 — 6.0MeV - op(1°Be,d)9Be (gs) — 11.8MeV *‘ 9.0MeV dU/dQ (mb/sr‘) 10—2 W11....l....l..11 O 20 40 0 20 40 Gem (deg) Figure 5.13: ADBA calculations based on CH89 potentials for the reactions of 9Be(d,p)'°Be (g.s.) (closed circle) [ZelOl, GenOO, Vas87, Sch64, Dar76, And74, 31662] and p(10Be,d)9Be (g.s.) (open circle). The calculations have been normalized by the spectroscopic factors. Each distribution in closed circle is displaced by factor of 10 from adjacent distributions. The overall normalization factor is l for the data at 11.0 MeV. The present data is reduced by a factor of 100. 170 Table 5.5: Extracted spectroscopic factors of the 1pm neutron from the reactions of p(lOBe,d)9Be (g.s.), and 9Be(d,p)mBe (g.s.) Incident SF Reaction Energy ADBA Ref. (MeV) (CH89) 9Be(d,p)loBe (g.s.) 6 2.05 [GenOO] gBe(d,p)lOBe (g.s.) 6.5 1.43 [GenOO] 9Be(d,p)wBe (g.s.) 7 1.3 [GenOO] 9Be(d,p)lOBe (g.s.) 7.5 1.04 [GenOO] gBe(d,p)10Be (g.s.) 8 1.12 [GenOO] gBe(d.p)loBe (g.s.) 8.5 1.01 [GenOO] 9Be(d,p)IOBe (g.s.) 9 0.97 [GenOO] 9Be(d,p)mBe (g.s.) 9.5 1.01 [GenOO] gBe(d,p)loBe (g.s.) 10 1.07 [GenOO] 9Be(d,p)mBe (g.s.) 10.5 1.08 [GenOO] 9Be(d,p)‘°Be (g.s.) 11 1.03 [GenOO] 9Be(d.p)lOBe (g.s.) 11.8 1.44 [Sch64] 9Be(d,p)‘°Be (g.s.) 12.5 1.29 [Vas87] 9Be(d,p)lOBe (g.s.) 15 1.83 [Dar76] 9Be(d,p)loBe (g.s.) 15.3 1.19 [ZelOl] gBe(d.p)loBe (g.s.) 17.3 0.97 [And74] 9Be(d,p)loBe (g.s.) 28 2.07 [81062] (average) (1.40:0.4 l) p( mBe,d)9Be (g.s.) 49.8* 2.99 * Equivalent proton energy 171 . I - T I - 2 8 4— CH89 — o 3 T .2 " 0 ~ 0. . L 1 8 r ....................................................... . <03) 2 L 0 fi_ E E 5 + _2 8 1' 3.0: 3‘ m C L O i . . . . 1 . . . . l . . . . O 20 4O 60 Ed (MeV) Figure 5.14: Extracted spectroscopic factors for the reactions of p(mBe,d)9Be (g.s.) (open square), and 9Be(d,p)loBe (g.s.) (closed circle). The dashed lines represent the shell model prediction of 2.35 by Cohen and Kurath [Coh67]. 172 CHAPTER 6 SUMMARY This experiment is originally designed to study the structure of the valence neutron of 10Be and 1'B by extracting the spectroscopic factors. The angular differential cross . . 10 ll 13 . . sections of (p,d) reactions on Be, B, and C are measured, wherein the reaction on 13C is performed as a calibration system and later used as a systematic study to devise a strategy to extract spectroscopic factors using the (p,d) and (d,p) reactions. Since target of . . 10 . . . . . . the radioactive nucleus Be is not available, the reaction is performed in inverse . . . ll 13 . . . . kinematics. The reactions on B and C are also performed in inversed kinematics to keep all the three experiments similar to reduce systematic errors and to learn about the new technique of using reverse kinematics of radioactive beams. All three secondary beams are produced by bombing a thick 9B6 target with the 13C primary beam. This experiment provides a learning experience of how to study (p,d) transfer reactions using secondary beams. The characteristics of the reaction in inverse kinematics are analyzed in this work. The advantage of the reaction in inverse kinematics is that the emitted light particles can be easily detected at forward angles (Section 1.2). The disadvantage of the reaction in inverse kinematics is that the energy broadening requires high angular resolution of the detectors (Section 1.2). The contributions to the energy resolution are analyzed in Section 4.2, which states that thin target and high angular 173 resolution of the detectors are the essential keys to achieve high energy resolution for the reactions in inverse kinematics. The angular differential cross sections are measured for the reaction of p(l3C,d)12C to the ground state (0+) and the first excited state (2+). The distributions are in good agreement with the published data in the literatures at adjacent energies (Section 5.3 and 5.4). The extracted SFs are in good agreement with the theory expectation. This means that the experimental techniques are good and the strategy to extract the SF works fine. The angular differential cross sections are measured for the reaction of p(“B,d)loB to the ground state. The extracted SF is 1.05. Past measurements give higher SF values even though this experiment in inverse kinematics presents the best agreement with the shell model prediction. The angular differential cross sections are measured for the reaction of p(mBe,d)9Be to the ground state. The extracted SF is 2.99, which is 27% higher than the theoretical value of 2.35. This value is not confirmed as we had problems in the beam normalization. Since the published data give lower values of spectroscopic factor, it is desirable to re- measure the differential cross section. The measurements of the differential cross sections to the excited states of 10B and 9Be are not performed because of the low particle counts. Therefore higher intensity of '1B and loBe beams, which are available from the new Coupled Cyclotron Facility, are desirable in future measurements. The energy resolution in this work is around 600 keV to 800 keV, which may not be high enough for the separation of some other excited states. Higher angular resolution achieved by placing the detector further away or by using smaller spacing of the strips is 174 desirable. This is currently under development in the construction of the HiRA (High Resolution Array) at NSCL. In the course of this study, we have developed the strategy to extract the spectroscopic factor by using a standard set of input parameters listed in Table 2.7 for the adiabatic deuteron breakup approximation (ADBA) calculations. We find that the Optical-Model Potential (0MP) obtained from fitting individual data of elastic scatterings do not give consistent and reliable spectroscopic factors due to the ambiguity of the 0MP; however, global optical-model potentials for proton and deuteron give consistent good “relative” spectroscopic factors. The 0MP based on the folding model and the effective nucleon-nucleus interactions such as CH89 and JLM potentials seem to give better agreement with data. Based on this work, specifically the analysis procedure provided, recent extraction of ground state neutron spectroscopic factors of 79 nuclei for elements ranging from Li to Cr [Tsa05]. These values are in consistent agreements with shell model predictions [BroO4]. Figure 6.1 shows the comparison of the extracted spectroscopic factors with the predictions of the modern shell model. Good agreements are achieved except for Ne, F, and Ti isotopes. Such agreement raises the possibility that the extracted spectroscopic factors are not only relative but absolute values. Furthermore, the agreement between the extracted values and the shell model predictions suggest that long-range n-n and n-core interactions can be described by modern day shell model. The disagreement between the spectroscopic factors extracted from transfer reactions and knockout reactions using the electron probe could be explained by the short-range nucleon-nucleon interactions since the electron probes the interior of the wave function where n-n interaction is more 175 important than that at the surface of the wave function where the transfer reactions are more sensitive. The present work has stimulated a lot of interest in the use of transfer reactions to extract spectroscopic factors, not only for rare nuclei but for stable nuclei as well. 176 .moaouoa E. 98 mm .02 Eooxo 888$ “88 53, 30808? @000 .Foamb “U 8 5 Eat wfiwcfi Eon: on c8 ”383 Boos. :2? =8on 2: mo 328605 2: 5:5 E98“ 030888on 38858 05 mo 2859.80 :6 Eswfi l: I "_m .2905. :25 S F to a") OBD