3 1:24" ’ ... 44¢” , rpm-u...“ NV Thane}: 4%, x»: is .7 ‘ \. . 1‘ .9.) 7;! E .3 q,..‘. ‘ lgv. u i L. .. . :5... n‘fikygfidzéxi 3.2.3.3233591 a . G KI: Burs. “4:319 /. illllill‘llfllllllllllxllilillwll’IHI‘W‘lllllllilUllllfl 3 1293 01774 616 LIBRARY Michigan State Unlverslty This is to certify that the dissertation entitled LOW-LYING COLLECTIVE EXCITATIONS IN NEUTRON-RICH EVEN-EVEN SULFUR AND ARGON ISO’I‘OPES STUDIED VIA INTERMEDIATE—ENERGY COULOMB EXCITATION AND PROTON SCATTERING presented by Heiko Scheit has been accepted towards fulfillment of the requirements for PhD. degree in Physics lo. 4M Major profeur Date December 17, 1998 MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 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 ma mu LOW-LYING COLLECTIVE EXCITATIONS IN NEUTRON-RICH EVEN-EVEN SULFUR AND ARGON ISOTOPES STUDIED VIA INTERMEDIATE-ENERGY COULOMB EXCITATION AND PROTON SCATTERING By Heiko Scheit 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 1998 ABSTRACT LOW-LYING COLLECTIVE EXCITATIONS IN NEUTRON-RICH EVEN-EVEN SULFUR AND ARGON ISOTOPES STUDIED VIA INTERMEDIATE-ENERGY COULOMB EXCITATION AND PROTON SCATTERING By Heiko Scheit The energies and B(E2; 0;... —> 2f") values for the lowest J’r 2 2+ states in the neutron-rich radioactive nuclei 38’40'423 and 44’46A1‘ were measured via intermediate- energy Coulomb excitation. Beams of these nuclei were produced by projectile frag- mentation with energies per nucleon of E /A z 40 MeV and directed onto a secondary Au target, where Coulomb excitation of the projectile took place. The subsequently emitted de-excitation photons were detected in an array of NaI(Tl) detectors, which allowed the identification of the first excited 2+ states in these nuclei. The ener- gies of the first excited J’r = 2+ states in 40’428 and 44Ar were established, while a previously observed state in 46Ar was assigned a definite spin. For all isotopes the B(E2; 0:3. —> 2?") was established. The results for 40'428 provide the first evidence of moderate collectivity (52 z 0.29) near N = 28, while the effects of the N = 28 shell closure persist in the Z = 18 nucleus 46Ar (,82 = O.176(17)). The deformation parameter of 44Ar was measured to be ,82 = 0.241(14). A proton scattering experiment was also performed on the unstable nuclei 42’44AI‘. The measured cross sections have been compared to optical model calculations as- suming rotational and vibrational excitation. The resulting deformation parameters indicate an isoscalar excitation. Meiner Familie gewidmet iii ACKNOWLEDGMENTS First and foremost I would like to thank Thomas Glasmacher. Thanks Thomas! You were a great boss. You gave me plenty of freedom (especially in my working hours) and you always knew when I needed a “kick in the butt” to get me going again. I want to especially thank you for sending me to the summer school in Erice, Sicily. (I won’t elaborate...) I would like to thank Walt Benenson, Wolfgang Bauer, Wayne Repko, and Bernard Pope for serving on my Ph.D. committee. Special thanks go to Maggie HellstrOm, since she got me started on the Coulomb excitation experiments. It all sounded so easy in the beginning... Rich and Sharon, thanks for the nice times at your parties. I really enjoyed a lot; throwing darts, bottling beer, playing foosball, doing the BBQ, watching Wallace and Gromit . . . Your house was often the meeting point for the “international crew”, which I had the pleasure to know: Greg, Heather, Marielle, Roy, Razvan, Kurt, Justine, Peter, Marcus, Navin, Anu, Andreas, Thomas A., Yorick, Tiina, Theo, Bertram, Takashi, Valentina, Marilena, ...(I am sure I forgot many more). I wonder who will keep things together after you leave. I am thankful to the people that helped me during my graduate studies: Mike Thoennessen, Paul Cottle, Kirby Kemper, Rick Harkewicz, Rich Ibbotson, Dave Mor- rissey, Mathias Steiner, Peter Thirolf, Yorick Blumenfeld, Tiina Suomijarvi, Alex Brown, Gregers Hansen, Jim Brown, Keith Jewell. Of course there are also my fellow students: Don — thanks for the good deal on my car, it wasn’t so bad after all, Jon, Mike R, Pat, Declan, Chris — sorry for going overboard in taking messages (from Sara or Jen? I forgot. . . ), Joelle, Erik, Joann, Jing, Barry, Raman, Luke, Sally, Gerd ...Maybe we will see each other at iv some “boring” physics conference again. (J ac, was it 305 Schuss Village?) Thank you, Barry, for the times we spent (jointly) together and the many discussions. Daniel and Jac, thanks for introducing me to IGOR. Rich, Boris, Marcus, and Barry: thanks for proof reading my thesis. Unfortunately they didn’t check this section. I thank the cyclotron staff members for their support. I really enjoyed doing the experiments here at the NSCL. I am also grateful for the generous financial support during my graduate studies. The Sandhill Soaring Club also helped me to survive during the last 4 years and thanks, Jeff, for bringing me in contact with them. I would like to thank everybody that went to fly with me. It was a pleasure. “Moi Droog” Boris, I know “In russia, we ...” I hope I wasn’t too rude to you at times, but I am really glad to have you as a friend. Well, now I need to say something about Marcus. Without you it would have been a lot more boring in East Lansing. I will never forget the many evenings and nights at the RIV and many other establishments in the area. Also, our discussions about physics and in general (“Did you have meat in the East Germany?”) have been very enlightening and useful. The analysis of my data without the “Driftellipsenmethode” would have been much more difficult. Your soccer expertise is very much appreciated. I would like to thank my family, especially my mother. Their support made this work possible. Finally, thanks to Simona, her cell phone, e-mail and the TeleGroup calling card. CONTENTS LIST OF TABLES LIST OF FIGURES 1 Introduction 1.1 The Atomic Nucleus ........................... 1.2 Nuclear Models .............................. 1.2.1 Microscopic Models ........................ 1.2.2 Collective Models ......................... 1.3 Experimental Probes of Nuclear Structure ............... 1.3.1 Coulomb Excitation as an Electromagnetic Probe ....... 1.3.2 Proton Scattering as an Hadronic Probe ............ 1.4 Heavy sd-Shell Nuclei ........................... 2 Intermediate-Energy Coulomb Excitation 2.1 General Description ............................ 2.2 Excitation Cross Section ......................... 2.2.1 Approximations .......................... 2.2.2 The Excitation Cross Section .................. 2.2.3 Electric Versus Magnetic Excitations .............. 2.2.4 Equivalent Photon Method .................... 2.3 Basic Parameters ............................. 2.3.1 Impact Parameter and Distance of Closest Approach ..... 2.3.2 Sommerfeld Parameter ...................... 2.3.3 Adiabaticity Parameter ...................... 2.3.4 Excitation Strength ........................ 2.4 Experimental Considerations ....................... 2.4.1 Projectile Excitation ....................... 2.4.2 Radioactive Nuclear Beam Production ............. 2.4.3 Detection of De-Excitation Photons ............... 3 Coulomb Excitation of 3840423 and 44"‘6Ar 3.1 Introduction ................................ 3.2 Experimental Procedure ......................... 3.2.1 Secondary Beams ......................... vi >4 X nu I—| OOCOCDanmI—lI-i ...—l 13 13 15 16 17 18 18 19 19 20 2O 24 25 25 26 28 34 34 35 35 3.2.2 Experimental Setup ........................ 36 3.2.3 Particle Detection ......................... 37 3.3 The NSCL NaI(Tl) Array ........................ 40 3.3.1 Mechanical Setup and Principles of Operation ......... 40 3.4 Electronics ................................. 42 3.4.1 Electronics for the NaI(Tl) Array ................ 44 3.4.2 Calibration and Gain Matching of the Detectors ........ 45 3.5 Analysis .................................. 51 3.5.1 Stability of Calibration ...................... 51 3.5.2 Angular Distributions ...................... 51 3.5.3 Photon Yields ........................... 57 3.5.4 Error Analysis ........................... 57 3.6 Results and Discussion .......................... 60 3.6.1 Observations ........................... 60 3.6.2 Comparison to Theory ...................... 63 4 Direct Reactions 68 4.1 Elastic Scattering ............................. 70 4.1.1 Optical Model ........................... 70 4.1.2 Effective Optical Potentials ................... 71 4.1.3 Microscopic Optical Potentials .................. 72 4.2 Inelastic Scattering ............................ 75 4.2.1 Coupled Channels ......................... 75 4.2.2 Nuclear Deformation ....................... 76 4.2.3 Mn/Mp .............................. 77 5 Proton Scattering of 36“42"‘4Ar 79 5.1 Experimental Setup ............................ 80 5.1.1 Proton Detectors ......................... 81 5.1.2 Beam Particle Tracking ...................... 91 5.1.3 Electronics ............................. 94 5.1.4 Kinematic Reconstruction .................... 96 5.2 Simulation ................................. 100 5.2.1 Examples ............................. 100 5.2.2 Efficiency ............................. 101 5.3 Analysis .................................. 101 5.3.1 Excitation Energy ......................... 101 5.3.2 Problem with the Cross Section ................. 108 5.3.3 Deformation Parameters and M,,/Mp .............. 115 6 Summary 121 vii A Coulomb Excitation 123 A.1 Semi-Classical Approach and Perturbation Theory ........... 123 A.1.1 Cross Sections ........................... 126 A.1.2 Relation Between Impact Parameter and Deflection Angle . . 127 A.2 Some Formulas .............................. 128 A21 Relations between 62, Q0, B(E2) .............. 128 A22 Constants ............................. 130 B ’y-Ray Angular Distribution Following Relativistic Coulomb Excita- tion 131 B.1 General Structure ............................. 132 B2 Detailed Derivation ............................ 133 B.2.1 What is (IffofkalH,|Ifo)? .................. 133 B.2.2 The Angular Distribution .................... 136 B3 Angular Distribution Using the Winther and Alder Excitation Ampli- tudes .................................... 139 BA Summary of Formulas .......................... 143 B.4.1 Angular Momentum Representation of the Rotation Matrix . 143 B.4.2 Some Properties of the Clebsch - Gordan - Coefficients and 31'- Symbols .............................. 144 B.4.3 Some Properties of the 6 j-symbols ............... 144 B.4.4 In", ................................. 145 B.4.5 7 — '7 Correlation Function .................... 145 C Relativistic Kinematics 146 C.1 Notation and Preliminaries ........................ 146 C2 Collision of Two Particles ........................ 148 C.2.1 Relation Between 06m and 9,0,, .................. 151 C22 Solid Angle Relation ....................... 152 C3 Decay of A into B,{,-=1,2MN} ....................... 154 CA Summary of Formulas .......................... 155 C.4.1 General .............................. 155 C.4.2 Total Energy in the Center of Mass System Em, ........ 155 C.4.3 Individual Energies in the Center of Mass System ....... 155 C.4.4 Velocity of the Center of Mass in the Laboratory 6cm ..... 156 C45 Relation Between 06", and 0M, .................. 156 C.4.6 Solid Angle Relation ....................... 156 C47 Invariant Mass .......................... 157 D Simulation of the Proton Scattering Experiment 158 D.1 Structure Of Computer Code ....................... 158 D2 Description of Input File ......................... 160 E Differential Cross Section 162 viii F A New Method for Particle Tracking 165 El Traditional Tracking ........................... 165 R2 New Method ................................ 166 F.2.1 Some Beam Physics ........................ 166 F.2.2 Determination of Position and Slope .............. 169 E23 Practical Method ......................... 171 G Fitting of Spectra 173 LIST OF REFERENCES 175 ix LIST OF TABLES 3.1 Parameters a0, a2, a4 of the angular distribution. ........... 52 3.2 Summary of uncertainties. ........................ 59 3.3 Experimental parameters and results ................... 67 4.1 Becchetti-Greenlees optical potential parameters. ........... 72 5.1 Detector thicknesses in am. ....................... 85 5.2 Experimental Parameters. ........................ 108 5.3 Deformation parameters 62 Obtained with the various ECIS fits. . . . 118 5.4 Summary of properties of argon isotopes ................. 120 0.1 Momenta of particles before and after scattering. ........... 149 D.1 Explanation of input file for simulation .................. 161 LIST OF FIGURES 1.1 1.2 1.3 2.1 2.2 2.3 2.4 2.5 2.6 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 3.10 3.11 3.12 3.13 3.14 3.15 3.16 3.17 3.18 3.19 3.20 Low-lying energy levels of 2°9Bi. ..................... 4 Typical vibrational (114Cd) and rotational (238U) level schemes. . . . 7 Energies Of the first excited 2+ state E (2?) for even-even isotopes. . . 11 Classical picture of the projectile trajectory. .............. 15 Adiabatic cutoff. ............................. 21 Cross sections for Coulomb excitation of 40S as a function of beam energy for 40S impinging on 197Am ................... 23 Secondary Beam Production. ...................... 27 A typical experimental setup for intermediate energy Coulomb excitation. 29 Contributions to the photon energy resolution .............. 31 Setup around the secondary target. ................... 36 Arrangement of the position sensitive N aI(Tl) detectors. ....... 37 Illustration of pulse shape discrimination ................. 39 Energy-loss vs. total energy after the secondary Au target. ...... 39 Arrangement of the position sensitive NaI(T1) detectors. ....... 41 Position calculated from the two PMT signals over the actual position Of the collimated source. ......................... 42 Electronics Diagram. ........................... 43 Lead housing and photon collimator for the position calibration of the NaI(Tl) detectors .............................. 45 Sample energy calibration curve ...................... 47 Illustration of the position dependence. ................. 47 Typical spectrum illustrating the correction for the Doppler shift. . . 49 Efficiency calibration ............................ 50 Photon angular distributions for pure E2 transitions. ......... 53 Angular Distribution of de-excitation photons .............. 54 The angular distribution in the laboratory system and in the center of mass ..................................... 55 The observed photon spectrum for 40S. ................. 57 Observed energies of 'y-rays as a function of position without correction. 61 Background subtracted photon spectra .................. 62 Shell model space and interactions. ................... 65 Results. .................................. 66 xi 4.1 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 5.10 5.11 5.12 5.13 5.14 5.15 5.16 5.17 5.18 5.19 5.20 5.21 5.22 5.23 5.24 5.25 5.26 5.27 5.28 6.1 C.1 C.2 F.1 F.2 E3 E4 Fit by Becchetti and Greenlees for the proton scattering cross sections. 73 Particle identification. .......................... 80 Setup around the secondary target. ................... 81 Schematic 3D view of the setup around the secondary target ...... 83 Position sensitive silicon detector telescopes. .............. 84 228Th a-source spectrum .......................... 86 Energy calibration of strip detector backside ............... 87 Position calibration of the strip detector. ................ 88 Signals that are registered by the individual detectors in the telescope as a function of proton energy ....................... 89 Particle ID spectra in the telescopes. .................. 92 Calibrated PPAC position spectrum. .................. 93 Effect of target orientation and beam spot size on the angular resolution. 95 Electronics Diagram. ........................... 97 Kinematics for the reaction 36Ar(p, p’) at E /A = 33.6 MeV ....... 99 Simulated spectra under ideal conditions ................. 102 Simulation including target effects. ................... 103 Simulation including target and detector effects. ............ 104 Actual data ................................. 105 Efficiency. ................................. 106 Excitation energy of 36Ar. ........................ 107 Excitation energy spectra for 42Ar and 4“Ar. .............. 109 Comparison of the measured cross section to previously published data. 111 Relation between the proton laboratory energy and the center of mass scattering angle ............................... 112 Ratio of the present data to the ones by Kozub. ............ 114 Different fits are obtained depending on the chosen shape of the nucleus.115 Measured cross sections and ECIS fits. ................. 116 Deformation parameters obtained with the various fits. ........ 117 A comparison of the extracted Mn/Mp values. ............. 119 A comparison of the extracted Mn/Mp values. ............. 119 Summary of results ............................. 122 Schematic drawing of a collision of two particles ............. 149 Relation between z,y, and 0 ........................ 153 Setup for particle tracking ......................... 166 Phase-space—ellipse ............................. 167 Comparison of old and new method. .................. 170 Measured excitation energy. ....................... 172 xii Chapter 1 Introduction 1.1 The Atomic Nucleus The atomic nucleus is a many-body system and consists of neutrons and protons which interact mainly via the strong interaction. Protons and neutrons are spin-% particles and therefore obey Fermi-Dirac statistics. Nuclear sizes are on the order of 10 fm (= 10"14 m). The structure of the nucleus is extremely complicated, mainly because of the small size of the nucleus and the strong (but short range) interaction between nucleons, which renders perturbation theory not applicable in most cases. Typical velocities of the nucleons in a nucleus are on the order of 0.2 -— 0.3 c, and the corresponding de-Broglie wave length is approximately /\ 2 2—7E z 4.5 fm. (1.1) mv This wavelength is on the order of the size of the nucleus and hence a quantum mechanical description of the nucleus is essential. 1.2 Nuclear Models The previous section hints at the complications one encounters while attempting to describe the nucleus theoretically. In general one has to solve the SchrOdinger equation 77011 = E\II . (1.2) The subject of nuclear structure theory is the determination of ”H is, and the resulting wave-functions look. Even with an exact knowledge of 71, it is impossible to solve equation 1.2 exactly, except for the lightest nuclei. Therefore one introduces a model that focuses on certain aspects of the structure and neglects others. A broad range of nuclear models exists; depending on the physical problem some models are more convenient than others and offer more insight into the structure of a particular nucleus. Nuclear models can be divided into two major groups according to the choice of coordinates used to describe the nucleus. Seemingly the most natural coordinates are those of the individual nucleons (i.e. the position r, the spin 5, and isospin 1'), hence the nuclear states are described by \Il=\Il(r1,S1,7’1,r2,...) a (13) and the Hamiltonian ’H is A p2 1 1 7-1:Z—'+—Zv(z',j)+—:v(i,j,k)+..., (1.4) i=1 2m 2 i,j 61“,“: where the sums go over all nucleons. These models are called microscopic models. The other major group Of nuclear models uses shape coordinates. These models are based on the collective degrees of freedom such as the center of mass R and the quadrupole moment Q20 of the nucleus 1 A A R = Z 2 Pi Q20 = 2722360620 (15) i=1 i=1 These are collective models. 1.2.1 Microscopic Models The best known microscopic nuclear model is the shell model. In the shell model the nucleons in the nucleus are considered to move independently of each other in the central potential produced by all other nucleons. Thus the multitude of interactions between the nucleons is replaced by an external mean field. Originally this model was motivated by the observation of magic numbers of protons and neutrons. The numbers are 2, 8, 20, 28, 50, 82, and 126. It was observed that nuclei in which the neutron or proton number (or both) corresponds to a magic number were particularly stable. Discontinuities (peaks or dips) in other quantities e.g., energies of first excited states, neutron absorption cross sections, and binding energies were also observed. The properties of many nuclei in the vicinity of closed shells could easily be explained by the shell model. Figure 1.1 (taken from [1]) shows the excited states of 209Bi. In the simplest shell model this nucleus consists of an inert 208Pb core (a doubly magic nucleus), which creates a mean field, plus one proton in the h % shell. Hence the spin and parity of the ground state is J1r = g-. A number Of excited levels can simply be explained by putting the extra proton into higher and higher orbits. The spins, parities, and even the excitation energies of these states can easily be predicted. However, one can also see in figure 1.1 that there is a group of levels around 2.6 MeV that can not be explained in this simple way. These states correspond to an excitation in the 208Pb core coupled to the extra 5; proton. The first excited state in 208Pb is a J 7' = 3" state with an excitation energy of 2.6 MeV. Hence these states (in 2°9Bi) have an energy of about 2.6 MeV, positive parity, and spins ranging from g- to 12,—. In order to explain these states within the shell model, the whole lead nucleus would have to be included into the shell model calculation, and the resulting wavefunction would correspond to a superposition of many single particle states. However, this coherent motion of many nucleons can be described more easily zoo ”8|,“ Spy: 3.64 V;- 3": 3.12 y; at“ an SI;- Wmfim {—} ~25 W-w; ”W: 1.61 W; 2!,“ 0.90 7/; ”’1: o 9/; man ENERGY 5m. mm PARITY Figure 1.1: Low-lying energy levels of 209Bi. The group of states near 2.6 MeV are excitations of the 208Pb core. Nicol OfIlN I“ nuck‘ readi Inodc 1.2. ltiva OfIn. DUle‘ beha‘ coHe( can 1 IS 011 209 B1 hunt (Jenn be Cc Darti by h Heigh Shun. hlat 3 ant in collective models, in which this 3‘ state is considered to be an octupole vibration of the nuclear surface. The mean field is only the average smooth part of the interaction between the nucleons and does not include all interactions. The remaining part is called the residual interaction. This interaction gives rise to collective behavior within the shell model. 1.2.2 Collective Models It was pointed out in the previous section that many features of nuclei in the vicinity of magic numbers can most easily be described by the shell model. Properties of nuclei in the mid-shell regions, however, can not be explained in this simple way. The behavior of these nuclei is best described in a collective model. The best examples of collective motion are the giant resonances, which occur in all nuclei. These excitations can be considered to be an oscillation of all neutrons and all protons. The emphasis is on all, since this is the most extreme contrast to the shell model, where, as in the 209Bi example, only one nucleon is responsible for the excitation. Besides the giant resonances, nuclei in the mid-shell regions show low-lying states (much lower in energy than corresponding single particle states), with large matrix elements that couple these states. A nucleus that exhibits these features is said to be collective since the large matrix element can only arise if more than one nucleon participates in a coherent collective motion. The energy spacing between levels is by itself only an indication of collective motion: One can for instance guess that neighboring even-even nuclei with the same energy for the first excited state are similarly collective, but this might be misleading. The best measure of the collectivity in a transitions between two states is the reduced transition probability B(7r/\), where 7t and /\ denote the parity and multipolarity of the transition operator. The B(7rA) is proportional to the square of the absolute value of the transition matrix element between states Ii) and | f) Bax) o< l|2, <16) and can e.g. be deduced from the lifetimes of nuclear states or from measured exci- tation cross sections. Equation 1.6 shows why the reduced transition probability is such a good indicator of collective motion; it depends directly on the wave functions of the involved states. Two major mechanisms are responsible for the low-lying collective states: One is the rotational motion of a statically deformed nucleus and the other is the vibration of the nuclear surface (without change in matter density within the nucleus). Two examples of collective nuclei are shown in figure 1.2, where the excited states of the even-even nuclei 238U and 114Cd are depicted. The level energies for 238U follow a very distinctive J (J +1) pattern (spacing increases linearly with J) and the spins are spaced with AJ 2 2. This spectrum is typical of a rigidly-deformed rotating body. The level scheme of 114Cd shows a completely different structure. One observes a 2+ state at 0.56 MeV and at about twice this energy a set of states with quantum numbers 0+, 2+, and 4+ are seen. This is typical of a nuclear surface vibration. The first level corresponds to a one phonon state carrying angular momentum 2 and positive parity. Two phonons (corresponding to the first harmonic vibration) can then couple to 0+, 2+, and 4‘“; other values of J are not allowed because the phonons Obey Bose-Einstein statistics. In the rotational case mainly electric quadrupole transitions (E2) occur between the states because of angular momentum and parity selection rules. One can show that the E2 transition Operator is identical to the quadrupole moment Operator and therefore the B(E2) value is related to the intrinsic quadrupole moment Q0 of the spacings that 0.777 0.5" 0.045 O 23.” ROTATIONAL 2* W F.” 1.2. 2’--—--I.zl “‘ca VIBRATIONAL Figure 1.2: Typical vibrational (”4Cd) and rotational (238U) level schemes. Level spacings that increase linearly with angular momentum indicate a statically deformed shape, whereas equal spacings indicate a vibrational nature of the excitation. nucleus The factor and also or shape will shape where .33 i quadrupol deformati Where R9 The d COllGCllYii the Perio ground SI SH 18 ( Collective defgrmat: deformat: the dOUb] Time IOI Collem nucleus B(E2) O< Q3 . (1.7) The factor of proportionality depends on the angular momentum J of the initial state and also on the quantum number K that characterizes the orientation of the deformed shape with respect to the quantization axis [2] Assuming a symmetrically deformed shape 12(6) = Ro(1+ mow, «b» , (1.8) where ,82 is called the quadrupole deformation parameter. Using the definition of the quadrupole moment yields the following relation between the B(E2) value and the deformation parameter (to first order): 4 52 = §\/3(E2;0+ —> 2+) (1,9) 1 ZeRg ’ where R0 is given by 1.2 fmA1/3. The deformation parameter given in equation 1.9 is a convenient measure of the collectivity in a nucleus, which allows the comparison of different nuclei throughout the periodic table, even if the nucleus is not statically deformed. For instance the ground state and the first excited state in the proton magic (non-collective) nucleus 122Sn is connected by a B(E2;0‘L —> 2+) Of 1930 ezfm", which is much larger than collective values for nuclei in the A z 40 region, which are typically 400 ezfm“. The deformation parameter, on the other hand, for 122Sn is 0.1, which is comparable to deformation parameters of non-collective nuclei throughout the periodic table (e.g. the doubly magic 40Ca has a 62 of 0.122). Typical deformation parameters range from 0.1 for non-collective nuclei to 0.5—0.6 for collective nuclei. 1.3 Experimental Probes of Nuclear Structure The nucleus can be studied passively, e.g. by Observing the radiation from radioactive decay, or by studying the nucleus’s electronic environment through the observation of the hyperfine structure in atomic spectra. The most fruitful approach, however, is to apply an external field and to study the reaction of the nucleus to this field. This is done by directing a beam of particles onto a target and observing the reaction products (other particles, 7 rays, electrons, etc.). Depending on the target-projectile combination and the relative velocity, different fields can be applied and the nucleus can be probed in various ways. The best probes are the ones that interact the least with the nucleus to be studied, so that it remains almost undisturbed. In addition, if the probe is weak, perturbative methods can be applied to extract the structure information. A possible drawback of a weakly interacting probe is a low reaction cross section. 1.3.1 Coulomb Excitation as an Electromagnetic Probe The best understood probe available is the electromagnetic interaction. For instance, electron scattering is used to map the charge distribution deep inside the nucleus, where no hadronic probe can give reliable results. Another method, which has long been applied to stable nuclei, is Coulomb excitation, where a nucleus is excited in the Coulomb field of another nucleus. The energies of the subsequently emitted photons (conversion electrons or other particles) reveal the spacings between the levels in the nucleus and the cross sections can be related to nuclear matrix elements connecting these levels. 1.3.2 Proton Scattering as an Hadronic Probe Coulomb excitation only probes the charge distribution of the nucleus since only the protons in the nucleus interact electromagnetically (neglecting the magnetic moments Of the neutrons). Because of the Pauli principle, the like-nucleon interaction is about 3 times weaker than the unlike-nucleon interaction in a nucleus. Therefore, proton scattering (p, p’ ) in the energy range of E /A z 10 — 50 MeV, which corresponds to typical kinetic energies Of nucleons in a nucleus, is mostly sensitive to the neutrons [3]. Information on both the neutron and the proton motion can be obtained by a comparison of the hadronic and the electromagnetic probe. The drawback of the hadronic probe is that the interaction is strong, and the structure information ob- tained is somewhat model dependent in contrast to the Coulomb excitation results, which are largely model independent. 1.4 Heavy sd-Shell Nuclei One of the frontiers in nuclear structure research is the investigation of shell evolution far away from stability. The nuclei in the region of N 220—28 and Z =14—20 (heavy sd-shell nuclei) have attracted particular interest for several reasons. These nuclei are located in between two major neutron shells (N =20 and N =28), which are very pronounced for the stable isotopes. This is shown in figure 1.3, which includes the results of the present work and shows the energies of the first excited 2+ states of the even—even nuclei for N up to 38 and for Z up to 28 are shown. The possibility of obtaining high-intensity secondary beams of these nuclei makes them particularly attractive experimentally. In addition, the nuclei in this region are large enough to show collective properties but are also sufficiently light that the neutron drip line can be approached. From the theoretical perspective full (no truncation) or almost full 10 Flgilr that 1 ll] thf; Figure 1.3: Energies of the first excited 2+ state E (2?) for even-even isotopes. Results of this thesis work are included. The magic numbers are indicated and one observes that the energies of magic nuclei are much larger than neighboring nuclei. Changes in the nuclear structure are expected, however, as the driplines are approached. 11 shell model . The \ms An introdui ‘2. and the 1 gives an int: proton scan. compares the shell model calculations can be performed. The unstable nuclei 38’40’428 and 44"“SAr have been studied via Coulomb excitation. An introduction to Coulomb excitation at intermediate energies is given in chapter 2, and the experimental details and results are presented in chapter 3. Chapter 4 gives an introduction to proton scattering, while the results of elastic and inelastic proton scattering on 36’42'44Ar are presented in chapter 5. Chapter 6 summarizes and compares the results obtained. 12 ... Cha Inte Exc: Following Valldlly C Pertainin 2.1 COUIOmb 0f the Dr OUlSide i in terms elicited I; SGCtiOn It The Con are largf'] 1 In the b} eXch 8.119 Chapter 2 Intermediate-Energy Coulomb Excitation Following is a general description of intermediate-energy Coulomb excitation. The validity of the assumptions made is discussed, and an overview over the parameters pertaining to Coulomb excitation at intermediate energies is given. 2.1 General Description Coulomb excitation is the excitation of the target nucleus in the electromagnetic field of the projectile, or vice versa.1 For pure Coulomb excitation, where the nuclei stay outside the range of the strong force, the excitation cross section can be expressed in terms of the same multipole matrix elements that characterize the 'y decay Of excited nuclear states. Therefore, a determination of the Coulomb excitation cross section leads directly to the determination of basic nuclear structure information. The Coulomb excitation process, as outlined below, is well understood, and results are largely model independent. 1In the following target excitations will be considered. Projectile excitations can be considered by exchanging target and projectile quantities (i.e. Z, H Zp,. . . ). 13 Keepl” dear exill MeV/“ml E/A S 4 such bfall fragmelliii far from 5 MeV/hurl at the $81 the use of beaniinte In the way to (‘II don.ist0 onh'eveni imumdis1 reaethons thatthe i Their matrix eh (hseussed. El lousin Keeping the bombarding energy below the Coulomb barrier ensures that no nu- clear excitation can take place. This beam energy is typically of the order of a few MeV/nucleon. For example for 40S impinging on a 197Au target, a beam energy Of E/A g 4 MeV is sufficiently small to exclude nuclear excitations. Unfortunately, such beam energies are impracticable for radioactive beams produced by projectile fragmentation, the mechanism employed at the NSCL to produce and study nuclei far from stability. The beams have energies Of several tens or even hundreds of MeV/ nucleon, and if the beams were slowed down, they would lose quality and flux at the same time. Furthermore, the long range of the ions at high energies allows the use of thick secondary targets, which helps to compensate for the potentially low beam intensity. In the intermediate energy range, defined here as E /A = 10—100MeV, the easiest way to ensure the dominance of Coulomb excitation, as compared to nuclear excita- tion, is to limit the scattering angle of the projectile to small angles. This means that only events are considered in which the impact parameter is larger than a certain min- imum distance defined by the maximum scattering angle. Also, events where violent reactions occur (corresponding to more central collisions) are rejected by requiring that the incoming beam particle be observed after the target. The next sections outline how the measured cross sections are related to nuclear matrix elements. Basic parameters that are involved in Coulomb excitation are also discussed. The main references are [4, 5]. A description of the experimental setup follows in chapter 3. 14 Figure 2.1: Classical picture of the projectile trajectory. The (in this picture stati- cally deformed) target nucleus can be excited due to the tidal forces exerted by the electromagnetic field of the projectile. 2.2 Excitation Cross Section Assuming that the projectile follows a Rutherford trajectory, the Coulomb excitation (CE) cross section is given by (more details are given in appendix A) do) (do) (d9 CE d9 Ruth where PH; is the probability of excitation from the initial state Ii) to the final state If). Treating the electromagnetic interaction potential V(r(t)) as a time-dependent perturbation, PH; is obtained as 00 meta-4|? with ai—Jzleg / e‘“’*‘dt- (22) -00 The amplitudes a,_,, can be expressed as a product of two factors Cit—+1 = iZXiflffalgl I (2-3) A where the excitation strength X is a measure of the strength of the interaction and the function f (.5) measures the degree of adiabaticity of the process in terms of the adiabaticity parameter 5. Section 2.3 gives a detailed discussion Of these and other parameters involved. 15 2.2. 1 Approximations Winther and Alder [5] obtained a closed formula for the Coulomb excitation cross section by assuming straight line trajectories and a static target. As shown in appendix A.1.2, the momentum transfer to the target perpendicular to the direction of motion is given by 2Z Z e2 77mm 2 Apit) = ——l:—J£—, (2.4) p where Z120) is the proton number Of the projectile(target), b is the impact parameter, ’0’, is the (incident) projectile velocity, and at is the target recoil velocity after the collision. 7 is the relativistic factor and mt is the mass of the target nucleus. For collisions considered in this work (Zt z 80, Z10 z 20, mt as 200 u up as 0.3 c, and b = 15 fm) recoil velocities of ’Ut < 0.2%c are obtained. This by itself might seem quite high, but typical collision times are on the order of 50fm/c z 1.5 - 10‘228 and the corresponding flight path of the recoiling target nucleus of 0.1 fm is much less than the nuclear radius of about 7 fm.2 Therefore, the assumption that the target nucleus remains at rest during the collision process is justified, which allows the use of a coordinate system with the target nucleus located (fixed) at the origin. A correction for the small target recoil can be introduced [5]. The deflection angle of the projectile in the laboratory is given by 2 glab = —- b— . (2.5) Inserting the values given above, deflection angles of only a few degrees are obtained. Hence the assumption of a straight-line trajectory is justified. 2Nuclear radii can be estimated as R = 1.25 fm Ai with A being the mass number of the nucleus. 16 2.2.2 The Excitation Cross Section The excitation cross section can be obtained by integrating the excitation probability from a minimum impact parameter bmm, determined by the experimental conditions (e.g. a maximum scattering angle), to infinity. An approximate result is obtained by introducing the adiabatic cutoff and integrating the absolute square of the excitation strength (equation 2.3) [le from bmin to bmax, instead of integrating H, = [X f (§)|2 from bmin to infinity:3 00 bmax 0 = 27r / P,fbdbz 27r / |x|2bdb. (2.6) bmin bmin bum can be estimated as 712 _ yl'w ~ 7197 bmar : '23; — AB ~ AE MeV fm , (2.7) where AE is the energy of the transition and to], = 935-. This leads to an approximate expression for the excitation cross section of parity it and multipolarity /\ 22 /\—1‘1 for A>2 a“ z (Zte ) B(m,0 —> MMQUA,’ ( ) _ (2.8) 2 min hc e 2ln(%:l::) for A=1, where bmaz >> bmm was assumed. B(7r/\, 0 —> A) is the reduced transition probability, defined as BOWL—>0) = ZIUIMIIMWAMIJz-MeHz I‘M] I _ , 2 — 2Ji+1|| , where M(7T/\/.l.) is the multipole operator for electromagnetic transitions. The exact expression for the excitation cross section, summed over parities and multipolarities, derived in [5], is Z e2 2 B m, 1,- —> I OH, = ( £6 ) Zk2(’\'1) t( 62 f) 2 9u(€(bmin)) - (2-9) C Ge“ (3) «Au 3i.e. the function f is approximated by a step function. 17 t Here k = 7 etgt. and B. The \Tinth- ll l0iil.i\\' ' transition ["1 Thus. the B 2.2.3 E] Electric and Therefore f 0370-3) In (“hem-[59‘ Here k = 5%}, Z, is the proton number of the projectile, E7 2 AE is the excitation en- ergy, and Bt(7r)\, I,- —-) I f) is the reduced transition probability of the target nucleus. The Winther and Alder functions G and g are explained in appendix A. It follows that the excitation cross section is directly proportional to the reduced transition probability 0'1'_,f OC Bt(7l')\,Ii ——-) If) . (2.10) Thus, the B(rrA) value can be extracted from a cross section measurement. 2.2.3 Electric Versus Magnetic Excitations Electric and magnetic fields of a moving charge are related through 22 IE! = -C-|E| (211) Therefore for intermediate- or high-energy Coulomb excitation, of interest here, (i z 0.3—0.5) magnetic excitations are possible and must be considered if not forbidden otherwise, e.g. by selection rules. In contrast, in low-energy Coulomb excitation (E z 0.1) only electric excitations are of importance. 2.2.4 Equivalent Photon Method In principle, Coulomb excitation can be viewed as the absorption of virtual photons by the target nucleus. These virtual photons are produced by the moving projectile and the equivalent photon number (the number of real photons that would have an equivalent net effect for one particular transition) is related to the Fourier transform of the time-dependent electromagnetic field produced by the projectile. One can express the Coulomb excitation cross section as dd) 0,-_,f : Z/Nfl(w)0§“)(w)—, (2.12) «A w 18 where t' photon The phi: where p nuclear lnsertin number Since th SPCtion IS USC’d q COUIOTTT 2.3 2.3.1 AS Shim rleatpS t tern. Z the pffjj that Str E where the spectrum of photons Of multipolarity A is determined by the equivalent photon number N..,\(w) and the photoabsorption cross section is given by UgfiANw). The photoabsorption cross section for real photons is given by [6] 0.(7rA)(w) _ (271F)3(A + 1) ‘ A((2A —1)u)2 7 p(e)k2’\+lB(7r/\) , (2.13) where p(e) is the density of final states and is usually given as a 6-function for discrete nuclear states p(€) = 6(E, + c — E,) = on — 13,). (2.14) Inserting equation 2.13 into 2.12 and comparing the result to equation 2.9 gives the number of equivalent photons of multipolarity 7r). [6] a... (5) Since the number of equivalent photons can be determined the photoabsorption cross 2 _ 2il((21+1)!!)2 9.46). (2.15) NMW) _ ”he (27r)3(/\ +1) ,, section can be related to the Coulomb excitation cross section and vice versa. This is used e.g. to derive astrophysically important photodissociation cross sections from Coulomb excitation results [7]. 2.3 Basic Parameters 2.3.1 Impact Parameter and Distance of Closest Approach As shown in appendix A the equation 2Zth C2 b—1 2 7m,,,vp Grab = (2-16) relates the impact parameter b to the deflection angle 0,0,, in the laboratory sys- tem. Z”. are the proton numbers of the target and projectile, mp is the mass of the projectile and up is its incident velocity. It was shown in the previous section that straight-line trajectories are a good approximation, and therefore the distance 19 0i closest n larger lilal; of Coulont? i The nut‘lear of the nude the sentient». 2.3.2 5. The South mp [5 ll)( Tallies a Small Cr. along tt 2.3.3 Iflhe< Dmitri “Ti l (h of closest approach is nearly equal to the impact parameter. The latter has to be larger than the sum of the two nuclear radii plus 2—4 fm [8] to ensure the dominance Of Coulomb excitation: DzbgRt+Rp+A3 with A,~2—4fm. (2.17) The nuclear radii can be estimated as R = 1.25 fmAi, where A is the mass number of the nucleus [9]. A minimum distance can be ensured experimentally by limiting the scattering angle of the projectile to be below a certain maximum scattering angle 0 3 6m, => b 2 bmmwmx). (2.18) 2.3.2 Sommerfeld Parameter The Sommerfeld parameter 1) compares the physical dimensions of the classical orbit, in this case the impact parameter b, with the de Broglie wavelength X of the relative motion of the two particles b b n=_:____’ympvp with 7: _ 3’2 X h and ,6— c . (2.19) 1 W mp is the projectile mass, 2),, the projectile velocity in the laboratory system. Typical values are r) z 1000, meaning that a wave packet containing several waves is still small compared to the dimensions of the trajectory. Such a wave packet will move along the classical trajectory, justifying the use of the semi-classical approach in the calculation of the Coulomb excitation cross section. 2.3.3 Adiabaticity Parameter If the time-dependent perturbation potential changes slowly the nucleus follows the perturbation adiabatically and no excitation is possible. This is the adiabatic cutoff, which is schematically illustrated in figure 2.2. This cutoff is parameterized in terms 20 Figure : adiabarj aHows a “UC‘leus the nud‘ “0 9Xcit. Cases. —> Vb vb Vb Figure 2.2: Adiabatic limit: In panel a) the collision time is short enough for the adiabaticity parameter 5 to be small and excitations are possible. Indicated by the arrows are the force vectors acting on the nucleus. Because of the orientation of the nucleus a torque is generated which gives rise to excitations. However, in panel b) the nucleus is able to follow the motion of the projectile, no torque is generated and no excitations can occur, even though the field strengths (~ x) are similar in both cases. 21 of the adiabaticity parameter 5, which is defined as the ratio of the collision time rm“ to the time of internal motion in the nucleus Tnucl T coll 5: (mm 7-nucl If f is large, because the projectile velocity is low or the impact parameter is large, then no excitation is possible. In appendix A it is shown that the electric field component in the x-direction E,c (perpendicular to the direction of motion) produced by the projectile at the target position is given by E b Z a: 7° 3mmr=——mdm=iL (no (1+(t/T)2)§ 7v. b2 Therefore 7' defines the collision time b Tcoll : _ (2.22) 7UP and the time scale for the nuclear motion is given as Tnucl = wgl = 3%. Hence 5 AE b = — == . 2.23 C can 7’01: 717% ( ) For I) ~ 0.3 c and b ~ 15 fm it follows that AE N , 2.24 6 5 MeV ( ) 6 should be smaller than unity for excitations to occur, and thus low-lying collective states with energies of several MeV can be excited. Going to even higher beam energies, e.g. 1) ~ 0.5 c, states in the giant resonance region with excitation energies of 10 — 20 MeV can readily be excited. Figure 2.3 shows the Coulomb excitation cross section as a function of beam energy for low-lying collective states and giant resonance states for dipole and quadrupole excitations for 40S impinging on 197Au. In both cases the giant resonance states are assumed to exhaust the electromagnetic 22 4 10 I IIIIIIII I IjITIIIl I IIFrIIrr r I IIII- + + (lflzl = 0.234) 3 102 - GQR E f d? , I 1 mi ;' 1 hm: 15 fm 10-1 1.....‘llllll L llllllll L l lllllll l lllLlll 10‘ 102 103 10“ 105 Beam Energy E/A (MeV) Figure 2.3: Cross sections for Coulomb excitation of 4”S as a function of beam en- ergy for 40S impinging on 197Au, with constant minimum impact parameter cutoff bmm = 15 fm. The energy region accessible with radioactive beams at the NSCL is indicated. The rising giant resonance cross section can be traced to the reduction of the adiabaticity parameter. The drop of the cross section for the low-lying collec- tive state is caused by a diminishing excitation strength parameter, due to a shorter collision time. sum rules [5]. It follows that in the region between 30 and 50 MeV, which is the energy region where the highest secondary beam yield can currently be achieved at MSU, only low-lying collective states can be excited. At higher energies, however, giant resonance states will also be populated [10]. For Coulomb excitation below the Coulomb barrier, where E/A ~ few MeV, the adiabatic cutoff essentially limits the possible excitation energies to be below 1— 2 MeV. The excitation to higher lying collective states is nevertheless possible, since the strength parameter (see next section) can become large. In this case first-order perturbation theory is no longer valid and multiple excitations can occur. 23 2.3.4 Excitation Strength As mentioned in section 2.2 the excitation amplitude can be expressed as a product of two factors: ai—1f = 2: XiiszAlgl- (2-25) ,\ f (g) is a measure of the adiabaticity as a function of the adiabaticity parameter, as stated in the previous section, and x measures the strength of the excitation from (A) state Ii) to state If) Classically, Xi—d is a measure of the strength of the interaction of the monopole field of the projectile with the A-pole component of the field created by the target nucleus. The monopole-monopole interaction gives rise to the Rutherford trajectory (and is treated classically). The next higher order is monopole-dipole interaction, which is not important for low energy excitations, because nuclei can not have a static electric dipole moment.4 Dynamic dipole-moments are nevertheless possible and they give rise to the giant dipole resonance, which can be found in all nuclei. Most nuclei, though, have a static (intrinsic) quadrupole moment, making the monOpole-quadrupole interaction the most important for low-lying (collective) states. Classically the field of the projectile exerts a tidal force (for A = 2) on the two lobes of the target nucleus and the resulting torque can excite the nucleus (see figure 2.2). x depends on the internal structure of the target nucleus, e.g. its quadrupole moment for A = 2, the strength of the monopole field of the projectile nucleus,5 and the duration of the interaction, the collision time Tm“. Hence x can be estimated as (1.1) ~ VA(b)Tcoll N Zt€(f|M(7rAp)|z') x (b) N —— ~ A ’ h hyvb (2.26) where Vy(b) is the monopole—A-pole interaction potential, which has a b‘lHl) depen- dence. 4Parity and time-reversal invariance forbid the existence of static electric dipole moments. 5The quadrupole moment of the target nucleus interacts with the derivative of the external electric field. The interaction energy is given by -%Qij g—fj [11]. 24 The matrix element (f |M(7rAp)|i) can be estimated as the square root of the B(E2T). A collective transition in the A240 region, which is of interest in this work, might have a B(E2) = 300 e2fm4 [12]; assuming furthermore v = 0.3 c and b = 15 fm it follows that, even for collective states and heavy projectiles, values of only about X z 0.15 are obtained and perturbation theory is justified. X can also be thought Of as the number of (7rA) quanta that are exchanged in one collision. Using equation A.10 one can also calculate the excitation probability at the minimum impact parameter, which has the highest probability. Typical values are on the order of 1%. Hence, Coulomb excitation at intermediate or relativistic energies is mostly a one step process. 2.4 Experimental Considerations In the following sections some experimental matters that need to be considered for a successful experiment will be discussed. Attention will be focused on projectile excitation of radioactive nuclear beams and the detection of '7 rays as a method of measuring Coulomb excitation cross sections. 2.4.1 Projectile Excitation In traditional Coulomb excitation the target nucleus is studied by bombarding it with heavy ion beams with energies below the Coulomb barrier. Since targets can only be produced from stable isotopes —— with few exceptions where long-lived isotopes have been formed into targets — the powerful method of Coulomb excitation has been limited to stable isotopes and the vast majority of nuclei (only about 10% of all known nuclei are stable) have been inaccessible by it. This changed with the construction of radioactive nuclear ion beam facilities, as described in the next section. Using projectile excitation (i.e. the interest lies 25 in exciting the projectile nucleus instead of the target nucleus; see figure 2.5) one can study all nuclei that can be made into a beam with sufficient intensity. This has opened up the possibility to study many previously inaccessible nuclei far from stability. It is also possible to use a cocktail beam consisting of many different isotopes if an event-by—event particle identification is possible. 2.4.2 Radioactive Nuclear Beam Production There are two main methods for the production of radioactive nuclear beams. One method is known as Isotope Separation On-Line (ISOL). Facilities that are working or are coming on line in the near future are HRIBF at Oak Ridge, SPIRAL in France, ISOLDE at CERN, and the REX-ISOLDE experiment, which will accelerate radioac- tive nuclei produced by ISOLDE. In the ISOL approach a high-energy, light-ion beam is stopped in a production target, where fragmentation of the target occurs. The re- sulting fragments diffuse out of the target, are transported to an ion source, and are then accelerated to the desired energy. An alternative method employed at MSU, RIKEN in Japan, GANIL in France, and GSI in Germany is projectile fragmentation. More details on these and other methods of radioactive beam production are given in [13]. Projectile Fragmentation The principle of projectile fragmentation is illustrated in figure 2.4. The primary beam impinges on the target, and even though most of the beam particles pass through the target Without reacting, a considerable number Of primary beam particles collide with the target nuclei and fragmentation occurs. A variety of fragments are produced and a system Of magnetic dipoles and quadrupoles together with an energy degrader (a piece of material in which the projectile loses energy) is used to reduce the number 26 l—> O O =’ 0 => 0 .N O A’ .\ Momentum Slits Slits for Beam Wedge Shaped Selection Production 1‘ ,1 Degrader Target . "' '" / Secondary Prunary Beam B cam Fragmentation Products &-' QT- s~% . p0“ [ w. w. _2_25° Dimes $115- First Intermediate Image: Momentum Focal Plane: Production Target: usually selection 98c Slits to select Second Intermediate Image: beam of interest Achromatic degrader (wedge) for beam punfication Figure 2.4: Secondary Beam Production: Illustrated here is prOJectIle fragmentation on the example of the A1200 fragment separator at the NSCL. The primary beam strikes the target and several nucleons are removed from the projectile nucleus. Due to the different energy losses of the fragments and the primary beam in the target, the resulting fragments have a large (longitudinal) momentum spread. Slits are used to reduce this spread. The various ions have different energy losses in the degrader and therefore have different magnetic rigidities after the degrader. The second set of dipoles spatially separates different ion spec1es 27 of unwanted fragments and to produce a secondary beam. The desired ions are transported to the experimental area where a secondary target is located, which serves as the Coulomb excitation target. The production of secondary beams at the N SCL using the A1200 fragment separator is now very common and the initial identification and setup takes, depending on the nuclei wanted, from a few hours to at most one day. For more details on the A1200 fragment separator see [14]. Notable properties of beams produced by projectile fragmentation are: a) high beam velocities, typically between 30-50% the speed of light, b) potentially low sec- ondary beam intensities, especially for the most exotic6 nuclei, which are often the most interesting ones, and c) a poor beam quality, due to the production mechanism; the large transverse and longitudinal momentum spreads (up to ApH/p” = i1.5% at MSU) result in a large beam spot size (around one cm) and a considerable beam emittance. The time from the production of the secondary isotopes to the arrival at the experimental station is on the order of a few hundred us to one as. Therefore it is possible to study extremely short-lived isotopes. 2.4.3 Detection of De-Excitation Photons After Coulomb excitation to a bound state the excited nucleus decays back to the ground state by emitting a 7 ray. The measurement of the y-ray energy readily reveals the spacing between the involved energy levels and the number of detected 'y rays can be used to determine the Coulomb excitation cross section which is directly related to transition matrix elements of the nucleus (see section 2.2). 6i.e. very neutron or proton rich compared to stable nuclei with the same mass number. 28 a“ Photon Detector 2 H ‘29; 5v Coulomb _ $3 Excitation Y may 5 o C O Figure 2.5: A typical experimental setup for intermediate energy Coulomb excitation: The beam particle is detected in a detector covering only forward angles. When particle—7 coincidences are measured, this limited extend of the detector ensures the dominance of Coulomb excitation by limiting the impact parameter to be above a certain minimum bmin. For each Coulomb excitation at least one 7 ray is emitted, which can be detected in a suitable detector. Energy Resolution and Dappler Broadening The beam velocities of particles produced by projectile fragmentation are very high, about 30—50% of the speed of light. Thus the 7-ray energy is Doppler shifted consid- erably, as can be seen from the relation between the photon energy in the projectile and the laboratory frame, EC", and Elab respectively7 Ecm = VEZabu — ,BCOS(010b)) . (2.27) 6,0,, is the angle between the direction of motion of the particle and the direction of the 7 ray, ,6 is the beam velocity in units of the speed of light, and 7 is the usual relativistic factor. Since, in an experiment, a single photon detector covers a range in solid angle, the width of the full-energy peak8 is increased. In addition the spread in beam velocities Afl, due to energy loss in the target and the intrinsic beam energy 7Since only one photon-emitting particle, the projectile, is considered here the subscript cm refers to the projectile rest frame and not to the center of mass of projectile and target. 8Events where the total photon energy was measured form the full-energy peak in the energy spectrum. 29 spread, will also contribute to the energy resolution. Using 2 2 AEfab = (on...) A32 + (81%) A62 (2.28) 83 80 it follows that the contribution to the energy resolution, due to the Doppler effect is ABS: 2 COS(0[ab) 2 2 2 ,5 sin (01“) 2 2 lab d (7) 1 — flcoswzab) 1 - 3 008(91ab) Figure 2.6 shows that these contributions are quite significant and that an accurate determination of the photon direction is very important. To the resolutions shown one should add the intrinsic energy resolution of the detector, which is for high-purity germanium detectors on the order or 0.1% and for N aI(Tl) detectors on the order of 7%. Figure 2.6 illustrates that the use of a NaI(Tl) detector is almost sufficient, since the resolution is dominated by the Doppler effect for an angular resolution of A0 = 10°. In practice, the assumption of Afl = 0.05 is quite high. Even though this value corresponds to a typical energy loss in the secondary target (see table 3.3), most of the particles considered in the present work decay outside the target and therefore have the same speed when emitting the photon, as discussed in the next paragraph. In order to correct for the Doppler Shift one needs to know not only where the photon interacts in the detector but also the position where the excited nucleus 7- decays. The transition rate for E2 transitions is given by (see [9])9 W(E2)=1.23-109-( E7 )5. (B(E2)) .3”, (2.30) MeV e2 f'm4 and the mean lifetime can be obtained as the inverse: T = W‘l. For 40Ar a lifetime of the first excited state of 1.2 ps is obtained, and for 408, which was measured as part of this thesis work, the lifetime is 22 ps. The path traveled by the particle in the laboratory is given by s = 7227. Hence, assuming 22 = 0.30, distances on the order of 0.1 mm to 1 mm are traversed between excitation and 7 decay. Therefore, 9E2 transitions are the only mode of 7 decay relevant to this work. 30 I l I | I |dE/d[3|*AB — — AB = 0.05 — S I N-POOO l I A r I be 10 — B=05 |dE/d9|*A9 a z 8 —- A9 = 100 —4 .2 6 I— _ a 4— — 8 2 - 5‘2 0 I I I I I 10 — [3:05 total ............. ON-h-OOO I l 3O 60 90 120 150 180 Olab (deg) Figure 2.6: Contributions to the photon energy resolution resulting from the energy spread of the beam (top) and from the detector’s angular resolution (middle). The quadratic sum of the two contributions is shown in the bottom. Also indicated is the intrinsic energy resolution of a NaI(Tl) detector. In practice, the assumption of Afl = 0.05 is quite high. Even though this value corresponds to a typical energy loss in the secondary target, most of the particles, considered in the present work, decay outside the target and therefore have the same speed when emitting the photon. Q 31 as a typical target thickness is on the order of 0.1 mm (20052-3 197Au target), most particles decay in the target or shortly after the target and the target position can be taken as the origin of the detected photons. But, since as outlined above, many nuclei decay outside the target (even if only shortly after it), as in the 408 case here of interest, one has to be careful about what to assume for the speed of the projectile, since the speed at the center of the target might not be the optimum value. As shown above, the accurate determination of the photon direction is very impor- tant in order to obtain good energy resolution. A discussion on the choice of detector is given in section 3.3, where the NSCL NaI(Tl) array is discussed. Cross Section In order to obtain the cross section one needs to know the number of beam particles that have been excited. This can be related to the number of photons emitted, since the nuclei de—excite, after Coulomb excitation, by emitting at least one photon. Other decay mechanisms, such as internal conversion or particle emission [15], are very unlikely for the nuclei studied here. Detection Efficiency. Unfortunately not all 7 rays that are emitted can be de- tected. A large number will escape, because they either do not hit the detector or they do not interact while passing through the detector. Therefore one has to know the 7-ray detection efficiency cm of the detector setup. In general 6 will be a function of the 7 ray energy and the photon direction. The source of the photons is assumed to be the center of the target. The efficiency for a certain direction and energy is defined as # of photons detected with direction 9 = = 2. 6 EU?” 9) # of photons emitted into direction {2 ’ ( 31) 32 where 0 stands for the direction defined by 9 and o in spherical coordinates. The detection efficiency, if the photons are emitted isotropically, is given by étot = /€dfl . (2.32) This is not the case for photons emitted after Coulomb excitation. However, the angu- lar distribution of the photons can be calculated, as shown in appendix B. The total efliciency is obtained by folding the detector efficiency with the calculated angular distribution W(Q) f we) - eIEIIm, 9) do 0 {{WM) d0 (2.33) 6tot(E7) : As indicated 6 depends in two ways on the photon direction S2; directly and from the dependence of the photon energy on Q due to the Doppler effect. The resulting total efficiency depends, of course, only on the photon energy. The total number of Coulomb excitations NCE is found by dividing the number of detected photons N, by the total detection efficiency etc, NCE = __ (2.34) The Coulomb excitation cross section a is defined as the ratio of the number of excitations NCE to the product of the number of incoming beam particles Nb and the number of target nuclei per unit area Nt 1 N, 1 = — . 2.35 NbNt 6tot NbNt ( ) O’ZNCE Therefore the cross section can be determined by measuring the number of emitted photons. 33 Chapter 3 Coulomb Excitation of 38’40’423 and 44,46 Ar 3.1 Introduction T wo of the basic properties of nuclei are their level structure and the transition matrix elements by which these levels are coupled. This information can be related to the shape of the nucleus. Nuclei that are spherical or stiff against deformation show large excitation energies and small coupling strengths. Nuclei that are deformed or soft against deformation have small excitation energies and larger coupling strengths. As pointed out in chapter 1, a large coupling strength indicates collective motion, since a single particle can not give rise to large matrix elements. A region in the chart of nuclei where little is known about the collectivity of nuclei is the region of the neutron-rich sulfur and argon isotopes. The structure of these nuclei is of particular interest as they are located between two major neutron shell gaps (N =20 and N =28), which are well known for the stable isotones. However, Sorlin et al. predicted in [16], based on fi-lifetime measurements, a rapid weakening of the N =28 shell gap below 48Ca. Shortly thereafter Werner et al. [17] published 34 similar conclusions based on relativistic mean field and Hartree—Fock calculations. In addition, the nuclei in this region play an important role in the nucleosynthesis of the heavy Ca, Ti, and Cr isotopes [16]. As discussed in chapters 1 and 2 a direct measurement of the excitation energy of the first excited J’r = 2+ state together with a transition strength measurement can give conclusive experimental evidence about the collectivity of these nuclei. The availability of a very high intensity 48Ca beam at the NSCL at MSU [18] makes the production of secondary beams of these nuclei possible through projectile fragmen- tation. The beam intensities (10—2000 particles/s) are sufficiently high to perform intermediate energy Coulomb excitation experiments. In a first experiment five iso- topes (38’40'428, 44’46Ar) were studied; the excitation energies of the first excited states and the corresponding B(EZT) values were determined. 3.2 Experimental Procedure 3.2.1 Secondary Beams Primary beams of 48Ca13+ and 40Ar12+ with energies up to E/A = 80 MeV and intensities as high as 5 pnA1 were produced with the NSCL room temperature electron cyclotron resonance (RTECR) ion source and the K1200 cyclotron. The 48Ca beam was produced using a new technique discussed in [18]. The secondary sulfur and argon beams were obtained via the fragmentation of the primary beams in a 379 mg/cm2 9Be primary target located at the mid-acceptance target position of the A1200 fragment separator [14]. The rates and purities of the secondary beams are listed in Table 3.3. 35 [ PMT ] NaI(T1) ] PMT ] TOF 5 30m ( Position.TOF Km {Beam PI Thin Plastic I PPAC PPAC 197Au Target PPAC Fast-ISlow- 2 Phoswich- (96 or 184 mg/cm ) Detector Scintillator Figure 3.1: Setup around the secondary target. 3.2.2 Experimental Setup Figure 3.1 shows the setup around the secondary target. The time of flight (TOF) between a thin plastic scintillator located after the A1200 focal plane and a parallel plate avalanche counter (PPAC) [19] located in front of the secondary target provided identification of the fragment before interacting in the secondary target as shown in figure 5.1. A fast—slow phoswich detector, which is described in section 3.2.3, was located at zero degrees with respect to the beam axis and allowed the identification of the beam particle after interacting in the target. The target thicknesses for the various beams are listed in table 3.3. Tracking detectors (PPACS) allowed particle tracking before and after the target (mainly used during beam tuning). Photons were detected in coincidence with beam particles by the NSCL NaI(Tl) array, which is described in section 3.3. Figure 3.2 shows the arrangement of the NaI(Tl) detectors in the experiment. 11 pnA (particle nA): 10-9 - —w1.6_,10_ P———a"j°‘“ z 6.2 ~ 1099—“;icle 36 NaI(Tl) Detectors Figure 3.2: Arrangement of the position sensitive NaI(Tl) detectors. 3.2.3 Particle Detection The forward particle detector serves several purposes: (a) It provides particle identifi- cation after the target on an event-by-event basis, (b) it ensures that only photons are analyzed that are emitted from beam particles scattered into laboratory angles of less than a certain maximum scattering angle. This angle is defined by the radius of the detector and the distance to the target (in this case 0:33“ 2 4.1°), (c) it provides the trigger (start detector), ((1) it counts the number of incoming beam particles (needed for the normalization of the cross section), and (e) it is used to measure photons in timed coincidence with beam particles, which strongly reduces the background, since it is possible to distinguish photons emitted from the target from photons emitted from the zero degree detector. The beam—particle detector used in this experiment consisted of a 0.6 mm thick 37 fast plastic scintillator (Bicron, BC400) followed by 10 cm of slow plastic scintillator material (Bicron, BC444). The detector is cylindrical and has a diameter of 101.6 mm (4”). A light guide followed the slow plastic and light was collected by two 5 cm diameter photomultiplier tubes (PMT). The thickness of the fast plastic was chosen so that the particles in this experiment lose about 20—30% of their energy in the fast plastic. The remaining energy is deposited in the slow plastic. The PMT signal, which was pr0portional to the produced scintillation light, was split into two streams and in one the full charge is collected (giving the total energy). In the other stream only the first pulse of charge (due to the fast plastic scintillator) was integrated (giving the energy loss in the fast plastic). This is illustrated in figure 3.3. The total signal is a sum of two components. By integrating the total charge (slow gate) the total energy is determined. An integration of the first part of the signal (fast gate) yields a signal roughly proportional to the energy loss in the fast plastic scintillator. This pulse-shape discrimination allowed the identification of the beam particle after the target. The detector had excellent Z resolution, but the neutron number was not completely resolved. Nevertheless, events from the breakup of the projectile could be rejected. Figure 3.4 shows a typical AE—E spectrum. The detector could tolerate rates as high as 5 - 104 particles/s. The detector was arranged so that fragments scattered into laboratory angles less than 0:23" = 4.1° were detected. For all beams, in this work, this opening angle corresponded to an impact parameter which is about 3—4 fm larger than the sum of the two nuclear radii, thereby ensuring the dominance of Coulomb excitation. 38 due to fast plastic I due to slow plastic Si gnalhei ght [_] 50ns fast gate ] 250ns J slow gate Figure 3.3: Illustration of pulse shape discrimination. The total signal is a superpo— sition of two components, a slow (light grey) and a fast (dark grey) contribution. Energy-loss (fast gate) . l 'l ... -..—... I Total Energy (slow gate) Figure 3.4: Energy-loss vs. total energy after the secondary Au target. The main contribution is the unreacted 44Ar beam (Z=18). If the beam particle collides with a gold nucleus fragmentation can occur and the resulting fragments are observed. The different Z—bands are indicated and clearly resolved. 39 3.3 The NSCL NaI(Tl) Array 3.3.1 Mechanical Setup and Principles of Operation Each N aI(Tl) crystal is cylindrical, 17.1 cm long and 5.0 cm in diameter and encapsu- lated in a 2 mm thick aluminum shield. Quartz windows, 0.5 cm thick, are attached at both ends. A 5 cm diameter photomultiplier tube (PMT) is optically coupled to each window. For mechanical stability the aluminum shield of the NaI(Tl) crystal is rigidly connected (epoxied) to a second aluminum pipe which holds the PMT and the connectors for high voltage (HV) input and signal output. In addition, each PMT is surrounded by a 0.5 mm thick magnetic shield. A total of 38 detectors are arranged in an aluminum frame in 3 concentric rings, of 11 (inner), 17 (middle) and 10 (outer) detectors, oriented co-axially to the particle beam axis around a 150 mm diameter beam pipe (figure 3.5). The radii of the three detector rings are 10.8 cm, 16.9 cm and 21.8 cm. As assembled the full array is 60 cm long, 51 cm wide and 55 cm high. When operated at its standard dedicated setup the array is surrounded by a 16 cm thick layer of lead to shield the scintillators from ambient background and 7 rays originating from the beam particles stopping in the fast-slow phoswich detector located at zero degrees with respect to the beam axis. The room background 7 rate is reduced, depending on the threshold settings, by about a factor of 100. When a “y ray interacts with the crystal a certain number of scintillation photons, proportional to the deposited energy, are produced, and about half propagate to each end of the crystal. Assuming an exponential attenuation, the light output on each side is described by the following formulas: E1 o< Ee_“(%+x) and E2 oc Ee-”(%_I) , (3.1) where a: is the distance of the interaction point from the center of the crystal, L is its total length, it describes the attenuation of the light. E is the total energy 40 \%§ a. \V :\/\%. '/ /%§\\\\\\\\l \’///////// Figure 3.5: Arrangement of the position sensitive NaI(Tl) detectors. The array is surrounded by a 16 cm thick layer of lead to shield from background radiation. deposited, and E1 and E2 are the measured signals. The omitted factor of propor- tionality includes the gain of the PMTs and amplifiers. Exponential attenuation of the scintillation light along each NaI(Tl) crystal can be achieved in various ways, e.g. by uniformly coating the crystal surface with a light absorbing substance or by diffusing the surface of the crystal [20, 21]. Similar detectors have successfully been used in the APEX experiment at Argonne National Laboratory [21]. It follows from equations 3.1 that one can recover the total energy and the interaction position through: E oc E1E2 and a: o< log(E1/E2) . (3.2) A more detailed derivation including the errors on these quantities is given in [20, 21]. There are several effects that change this idealized picture: The assumption of exponential attenuation of the scintillation light is only valid when the interaction point is close to the center of the detector, while at the edges solid angle effects (which lead to a reduction of the light absorption) are more important. This effect can be 41 I I I l I l I I I I 230 _ I I I I I _ 6 7T% 0'6 N...“ 60cc 1 W IDVI - . ... ...... 2m a 0.4 - f ------ 0... ‘3 15° ‘ - 800 """ t. 8 100 .- 0 2 - -- P U 50 _. . . o O T“ b. O T T I 600 A E 0'0 " § ”a. 800 1000 1200 1400 5 LL}— Energy (keV) m. :3 .,,__ as .0. -0.2 - a“... _ 400 E FWHM = 2.3 cm "'1: ..... g .... o .04 — . U ‘ ..... {M _ 200 -0.6 - at“. ”0.8M l 1 1 -~_ -_- 0 5 10 Position on detector (cm) Figure 3.6: Position calculated from the two PMT signals over the actual position of the collimated source. In the central region the relation is linear, while close to the edges of the detector the curve flattens because the assumption of exponential light attenuation is no longer valid. The open circles show a sample position spectrum with the source located at 8 cm. The dotted lines are cubic and Gaussian fits to the calibration curve and the position spectrum, respectively. The inset shows an energy spectrum with the source at the same position. The line is a double Gaussian fit with a quadratic background. corrected through a cubic position calibration which nicely describes the turnover of the position calibration curve close to the edges of the detector (figure 3.6). 3.4 Electronics Figure 3.7 shows the electronics diagram for the Coulomb excitation experiment. The Master Trigger was defined as a coincidence between a beam particle and a single N aI (Tl) PMT. In addition to particle-7 coincidences, particles without a coincidence 7 ray (particle singles) were also measured at a rate of every 500th incoming beam 42 :48 ES: >w3m u>5 OE LJ. 030 :52 8.32 .33:E:un_a 55.3.“. E3359 u at 5.2—.8530 umam 35164 n Dun. 355.3 amigo. 5:02:40 “are 3 M33}. u (mun. 5:355 3&5 3 DEC. u DOE 352.} £595 5:02.20 12.?0 a. as; H UQ< 5.5250 030 n 00 Figure 3.7: Electronics Diagram. 43 particle (down-scaled particles), which allowed the determination of the total number of incident beam particles. 3.4.1 Electronics for the NaI(Tl) Array A multi-channel high voltage (HV) power supply (LeCroy System 1440) provided typ- ically +1400 V to each PMT. Signals from the PMTs were fed into shaping amplifiers (custom built at NSCL/MSU) and the resulting signals were used for discrimination and energy measurement. The fast output (30 ns shaping time) of each shaper was fed into a constant fraction discriminator (LeCroy MSU 1806 CFD, LeCroy 3420) whose output signal was used for the generation of an event trigger, a scaler signal, and a timing signal. The event trigger required, in addition to a 7 signal, a beam particle signal in timed coincidence. After a trigger signal had been generated, all PMTS that belonged to a NaI(Tl) in which at least one PMT fired (according to the bit register, LeCroy 4448) were read out. The energy was obtained from the slow output of the shaping amplifier (5ps shaping time) by digitizing the resulting signals by 8-channel Silena peak sensing ADCs (Silena 4418/V). The trigger, whose signal is correlated in time with the beam particle signal, initiated a start signal for all time to digital converters (TDCs) which were stopped by individual PMT signals and therefore mea- sured the time between detection of the photon and the beam particle. (The TDCs consisted of a combination of time-to-FERA converters (LeCroy TFC 4303) followed by charge integrating ADCs (LeCroy FERA 43003). Between the START and STOP signals the former applied a constant voltage (-50 mV) on the input terminal of the latter, which resulted in a signal proportional to the time between the START and the STOP signal.) The TDC range was set to 200 us, which corresponded to the length of the coincidence window between photons and beam particles. The time spectra helped to distinguish between photons emitted from the target and photons 44 'fib; . . f...'. O . .'.'.'.'.V. ... .'.'. ' . . %% .. fifihfififififihfl”&%§¥§%¥% .. . ........0...9.....0 ... . ................... .. ... . ..............0... . 9. 0.. .................0. 0.... .0................ 9 ... ...... 0.............. .9 ......0.... ...99.. ............... ..r ........0 ......... 0. . ........ 0........ .....0... ..... w .......... .. .0... .0 .. .. ...... .. .. . . ... 9 ... ..... ... ’ JV? . . o... v . . I vvvv'vvvv v v vvvvv ..- .§UV¥VVVU¥”5 . %&%&%&%¥V%JV ......0... . . ...........4 ... . .. . .......... ... 0. ,. ..........< . .. ... .....0.. >9 ....... .0.4 ..............9 ... .......... ....4 ...... ............0.. §@5§UWW% w%&&&&&%~www¥w¥ ....... b..............4 ....... ...........0... ....... v..............< ......9 ............... .....00 >....9.........¢ ....... .9............. ......0 >.9......... ..c ......9 0......9....... ....... DOOOOOOOOOOOOOO‘ ....... ........... . . ....... ...........9...4 ..... . ............... .. .. >...........0..¢ ....... ............... fififififif ”UVVVVVVVVVVVVE 3 J m mmmmgmmnu ' .V. .‘ .... ....’...f‘ ibévwvvvvwb évvwb..dw%5%fifififlfi ....... ............... .....9..... .............. .9.9..........................9 .... . ......0 ..............9 .............. .0.............. ......................... .... ...................... .9 .... .................9.. . ..... ................ 0. 9 ... . ... ......90. . .4 ................0... 0 ......9.....0....... .0.........0... ... . .:.:.:.:.:.:.. 3:.}:.......:.:...:.:.:. .. ...QO‘ ......O...‘.. Figure 3.8: Lead housing and photon collimator for the position calibration of the NaI(Tl) detectors. A 0.5 MBq 60Co source was positioned at the top end of the opening. emitted by particles stopping in the zero degree detector. The event read-out time was about 300 ps. 3.4.2 Calibration and Gain Matching of the Detectors Before and after the experiment a position, energy, and efficiency calibration of the NSCL NaI(Tl) array was performed. For the position calibration a collimated 7—ray source was constructed, which consisted of a 8 mm wide and 15 cm long cylindrical opening in two lead bricks, and a 0.5 M Bq 60Co source, which was positioned at one end of the opening. Each NaI(Tl) detector was inserted into a lead housing (shown in figure 3.8) to shield it from ambient background and the collimated source was initially positioned so that the photons are directed at the center of the crystal (in longitudinal and transverse direction). Then a 45 first coarse gain adjustment was performed by tuning the PMT HV through a visual inspection of the PMT signal amplitudes. Afterwards the shaper gains were adjusted to the desired dynamic range, depending on the expected maximum 7-ray energy (4 MeV in this case), while, at the same time, keeping the total gain of two signal processing chains, belonging to the same detector, matched by centering the peak of the reconstructed position (as defined by equation 3.2) around zero, from which it follows that E1 = E2. The dynamic range was chosen large enough to account for the asymmetry in signal height when the 7 ray interacted close to the edges of the detector, i.e. since we wanted to detect photons in a range of up to 4 MeV we had to make sure that no component of the signal processing chain saturated if this photon deposited all its energy close to one PMT. Therefore the range of the shaper was set about a factor of 1.5 higher, which looks like a 6 MeV range for photons interacting in the center of the crystal. After each detector was matched it was moved in steps of 1—2 cm, according to a measuring tape attached to each detector, to different positions; the data was digitized and recorded on magnetic tape for later analysis. Figure 3.6 shows the reconstructed position, as defined in equation 3.2, versus the true position. One can see that in the central region of the detector the correlation is linear, while at the edges the assumption of equations 3.1 are no longer valid and the curve flattens. The data points were fitted with a third order polynomial, which nicely describes the turnover of the curve close to the edges of the detector. Photon sources of 88Y, 152Eu, and 228Th were used for an energy calibration. They provided calibration points from a few 100 keV to 2.6 MeV. Since the raw energy was not completely independent of position, a position dependent energy calibration was performed by applying position cuts to the energy spectra (10 cuts per detector). In between these slices we interpolated to obtain a smoothly varying calibration as a 46 I I I I I I 2500— a = —24.58 38y I J b = 2.03 228 2000 _. c = 4.57e—05 152Eu I Th_l % 5 1500 — _ >~ P." d.) l: 1000 —- _ m ISZEu Detector #03 500 _ (Central Slice) 0 l I I I I I 0 200 400 600 800 1000 1200 1400 Channel Figure 3.9: Sample energy calibration curve. A quadratic fit was used. The photon sources used for each point are indicated. Energy (MeV) S '3 g 9 Ln 40 50 60 70 80 40 50 60 70 80 Position (arb. units) Figure 3.10: Illustration of the position dependence: The left panel shows the uncal- ibrated energy over the position for an 88Y source. The right panel shows the same spectrum after a position dependent energy calibration. 47 function of position. Figure 3.9 shows the energy calibration curve for the central position slice of detector number 3. The effect of the position dependent energy calibration is illustrated in figure 3.10. Shown is the measured (reconstructed) energy over position for an 88Y source spectrum. One can see that after the calibration the measured energy is independent of position whereas some distortion is present in the uncalibrated spectrum. The sources of these distortions might have been non- uniformities in the NaI(Tl) crystal, which changed the light attenuation along the NaI(Tl) crystal from the ideal exponential behavior (see section 3.3). The detectors had an average position resolution of about 2 cm and an energy resolution of about 8% at 662 keV. The 2 cm position resolution translated into a contribution to the energy resolution, due to the Doppler effect, of 5% for the inner 11 detectors, whereas without the position information the energy resolution would have been much worse. This is illustrated in figure 3.11. The top panel shows an energy spectrum without Doppler corrections for the case of 40S on 197Au. A peak at 547.5 keV is visible, which corresponds to a transition in 197Au, which served as the (stationary) target nucleus. Also visible is a “bump” around 900 keV. The bottom panel shows the same spectrum, but each event has been Doppler corrected using the known photon interaction position in the NaI(Tl) detectors. Now one can clearly see a peak centered at 891 keV which corresponds to the 7 decay of the first excited state in the radioactive isotope 40S. The measured resolution is 8.7% and only slightly larger than the resolution obtained from a stationary source. From the given detector length and the radius of the inner ring of detectors it follows that for a source located at the center of the array the geometrical efficiency is close to 27r (assuming an azimuthal coverage of 80%). The photopeak efficiency for gamma radiation can be determined in two different ways. One method is to use photon sources that emit two or more 7 quanta in coincidence. Gating on one 48 I f I I I I FWMH = 9.0% 140 120 100 80 I3 = 0.0 547.5 keV 0 keV _ 40 20 é 80— Counts / (10 keV) O 400 600 800 1000 1200 1400 1600 Energy (keV) Figure 3.11: Typical spectrum illustrating the correction for the Doppler shift. The top panel shows the raw 7-ray spectrum, while the bottom panel shows the same spectrum corrected for the Doppler shift using the position information of the NaI(Tl) detectors. 49 2'0 I I I I F I I 85<01ab<95 15 (Sdetectors) _ S 3‘ 1.0 -- —1 fl .2 U E m 0.5 — — (,0 I l l l I I l 600 800 1000 1200 1400 1600 1800 2000 Energy (keV) Figure 3.12: Efficiency calibration. The used points are from gamma transitions resulting from the decay of 22Na, 88Y, and 152Eu. The coincidence method applied to the first two sources gave an absolute calibration, while 152Eu was used for a relative calibration. photopeak energy of the cascade in a specific detector and integrating the number of counts registered by all other detectors in the photopeak of the corresponding second 7 transition (and taking into account the coincidence ratio between the two transitions) leads to a value for the absolute photopeak efficiency of the detector array. The advantage of this method is that it is not necessary to have an absolute calibration of the photon source. The other method consists of using a calibrated photon source and applying corrections for the dead time of the data acquisition system. Both methods were compared and gave consistent results. Figure 3.12 shows efficiency calibration points from several sources. 22Na, 88Y provided absolute points through the coincidence method, whereas the points from 152Eu were used as a relative calibration. The functional form of the fit was 5 = A - E}? , where A and B are fit parameters with B yielding values around -1 (i.e. e or i). The photopeak efficiency for the inner 11 detectors, was about 10% at 890 keV and scaled 50 roughly with inverse energy in the region between 0.7 and 2 MeV. The exact values for the nuclei studied here are listed in table 3.3. 3.5 Analysis 3.5.1 Stability of Calibration During the course of the experiment unexpected energy shifts over time were ob- served; only two detectors had a stable energy and position calibration, determined by comparing source calibrations at the end and the beginning of the experiment. The most probable causes were gain shifts in the photomultiplier tubes, which have been used extensively. For subsequent experiments several PMTs were reglued and loose high voltage and signal leads were resoldered, which improved the gain stability. Nevertheless, the statistics in the two stable detectors were sufficient to estab- lish the energy of the first excited states in these nuclei. Since 7-ray peaks due to transitions in the projectile nucleus were observed in most detectors, an additional energy calibration was performed in order to move the peaks (corresponding to the first excited states) to the correct energies, as determined by the two stable detectors. In all, 8 detectors (all, except the most unstable) were used for the extraction of the Coulomb excitation cross section. 3.5.2 Angular Distributions In appendix B it is shown that the angular distribution of the de—excitation photons is given by (3.3) W(0) = z akacosw», I: even 51 I m a0 02 a4 bm,,,(fm) E/ A (MeV) 1 0 4—11; -1 0 E1 1 1 31; % 0 pure 2 0 L Q _ 12 2 1 :1. 1% 37 pure E2 2 2 El? _ g _ Z transrtlons 2 g -0535 0152 15 40 2 31; -0.663 -0.239 15 200 2 2% —0.405 -0.082 25 40 Table 3.1: Parameters a0, 02, 04 of the angular distribution for pure transitions and distributions in intermediate energy Coulomb excitation. The value for do ensures the proper normalization. with a.=: ulLlL’ I I 11: ° {if1.1.}Fk(LlL’vaf,If)\/2k+l6L6L,_ 0mg) 2 ,, /\ A k . g..(o(—) (Mm) Using equation B.14 one can also obtain the 0,, for pure EAm transitions. Listed in table 3.1 are the ak’s for pure E1 and E2 transitions together with the calculated coefficients for actual distributions. Figure 3.13 shows the angular distributions for pure E2 transitions and the calculated transitions from table 3.1. The similarity of the actual distributions with the 1:2, m=2 case shows that the main contribution to the excitation results from maximum angular momentum transfer along the beam direction, as shown analytically in [5]. The importance of considering the angular distribution is illustrated in figure 3.14. Shown are the ((1) integrated) angular distributions for isotropic 7 decay and 7 decay following Coulomb excitation of 40S on a 197Au target at E/A = 40 MeV. Here, the detector efficiency was approximated as a step function. In this case the measured cross section would have been overpredicted by 14% if the angular distribution were 52 Pure E2m Transitions l=2,m=:l'1 ‘97Au ( 408,40? ) '97Au ,/ I. \ E/A = 40 MeV E/A = 200 MeV E/A = 40 MeV bmin=15 fm bmm=15 fm bmin=25 fm Figure 3.13: The top row shows the angular distributions for pure E2 transitions. The quantization axis, which is identical to the beam direction, is going from the left to the right. The bottom row shows angular distributions for different beam energies and impact parameters. One can see that the main contribution results from maximum angular momentum transfer along the beam direction (similarity to the pure 1:2, m=2 case). 53 197Au ( 403’4OS* ) 197Au 012 I I I I ------ isbtropic 0.10 — — actual _ 0.08 0.06 W(0) sin(0) 0.04 0.02 0.00 i 0 30 60 90 120 150 180 90m (0) Figure 3.14: Angular Distribution of de-excitation photons. The area under the curve corresponds to the geometric efficiency. Neglecting the angular distribution of photons would result in a 14% larger cross section measurement. not considered. Figure 3.15 shows the angular distribution in the laboratory system. Folding with the Efficiency With the angular distribution normalized to unity / W(n) do = 1 (3.4) 41r the total efficiency is obtained as 5,0, = / W(0) {(E,(n), 0) do n = Z/W(o)e(E,(n),n)dn i9; 54 — Center of Mass ......... Laboratory ([3:0284) Figure 3.15: The angular distribution in the laboratory system and in the center of mass. 55 Ze(E.(0.-).0.) / W10) do. 91 In the last step it was assumed that the efficiency is constant in the range 52,-.2 During the source calibration, using either the coincidence method or a calibrated photon source, the following quantity is measured: ° 1 6880) = E/dflflifl. (3.5) n This is the efficiency integrated over Q (as indicated by the subscript g), with the weight (= %)corresponding to an isotropic distribution. (If N photons are emitted (iso) isotropically then en - N photons are detected within the solid-angle range {2.) Assuming that 6(0) is constant within 52 it follows that 47]- iso 6(0) = 5.], ). (3.6) Hence the total photo peak efficiency is given by 2:4 ISO) e,o,— _ 7r 5],], E)/( W(n (3.7) where 689?) (E7) is obtained from the efficiency calibration and W(Q) is obtained from formula 8.17. In this experiment the whole array was segmented (in software) into 7 regions (Shrug) defined by cuts in 0 (45—55°,. . . ,115—125°). For each of these regions the efficiency was experimentally determined and the photon angular distribution was calculated and integrated. Subsequently formula 3.7 was used to obtain the total efficiencies, which are listed in table 3.3 for all five isotopes. 2In the previous and following expressions Q is used to label both a range in solid angle, as the Q under the integral sign, and a direction in spherical coordinates Q = (I9, (15), as the Q in the argument of e() or W(). It should be clear from the context what is meant. 56 160 1 I I I I I I w 1 140 197Au (408,408“ ) 197Au _ E 120 — -— o 100 — I -— \ 80 — II — 8 G 60 — _ 5 4o — ‘ I a 20 II '11 .Ln ' ‘1 0 T 1 1 1 L I 400 600 800 1000 1200 1400 1600 1800 2000 Energy (keV) Figure 3.16: The observed photon spectrum for 408. A Gaussian fit with quadratic background was used to determine the peak area. 3.5.3 Photon Yields The number of emitted photons was obtained by integrating the observed 7-ray spec- tra. A sample spectrum is shown in figure 3.16. No cuts have been applied, besides the requirement of the correct isotopes before (AE—TOF) and after (AE—E) the tar- get. In contrast, the spectra shown in figure 3.18 have been obtained by requiring a photon multiplicity (fold) of one and a cut was made in the time spectra corre— sponding to photons emitted from the target. This condition could not be applied in the cross section determination, since the peaks due to the target and zero degree detector were not completely resolved. The spectrum in figure 3.16 is fitted well by a Gaussian peak with quadratic background. 3.5.4 Error Analysis The uncertainties on the extracted excitation energies and the measured B(E2) values are calculated by using standard error propagation [22]. The sources for systematic and statistical errors will be discussed for the example of 403. 57 Excitation Energy Statistical Errors. The largest contribution to the statistical error was the intrinsic resolution of the NaI(Tl) detectors, as discussed in sections 2.4.3 and 3.3. Other contributions were the uncertainty in the angle of the photon and the uncertainty in the velocity of the 7-emitting particle, due to energy loss in the target and the initial beam energy spread. Both of these contribute to the energy resolution via the Doppler shift. All of these uncertainties manifest themselves in the width of the measured peak in the energy spectrum. Systematic Errors. The only systematic error, besides the error in the energy calibration parameters, was the uncertainty in the target position with respect to the photon detectors, which was estimated to be 5 mm. This includes: . the position of target with respect to beam-line 0 the position of NaI(Tl) array with respect to beam-line 0 the position of NaI(Tl) detectors in the array (there is a 1—2 mm play) 0 the position of measuring tape on the detectors The resulting systematic error for the measured energy was 1.25%. In subsequent experiments this contribution was reduced through the use of a position calibration device that calibrates the detector position relative to a certain point on the beam line [23]. Therefore the last three points of the previous list can be replaced by the uncertainty in positioning of the calibration device, which is very small. Cross Section Statistical Errors. The only statistical error was due to the counting of photons in the photopeak. The numerical value was obtained by fitting the peak in the photon spectra (figure 3.16) and therefore includes the uncertainty in the background as well as the uncertainty in the peak. 58 Result Quantlty Contribution Magnitude on Quantity Energy statistical 0.4% calibration parameters 5 keV beam velocity 0.1% <0.01% target position 5 mm 1.25% total 1.4% Efficiency calibration parameters 8 % 0m... 0.1° 0.4 % beam velocity 0.1% <0.1% total 8 % Cross section statistical 5 % efficiency 8 % total 10 % B(E2T) am 01° 5 % excitation energy 1.4% 2.8 % beam velocity 0.1% 0.3% cross section 10 % 10 % total 12 % Table 3.2: Summary of uncertainties in the excitation energy, the detection efficiency, the excitation cross section, and the B(E2T) value for the 40S measurement. 59 Systematic Errors. The largest contribution to the uncertainty of the cross section came from uncertainties in the determination of the photon detection efficiency. The uncertainty in the efficiency includes contributions from the fit of the calibration curve, by far the largest contribution, and contributions due to the calculated angular distribution, which depends on dmax (see figure 3.13). In subsequent experiments the error due to the efficiency calibration was reduced by the use of calibrated photon sources. Here, one does not have to apply gates, which raises the number of counts by an order of magnitude, and thus reduces the statistical uncertainty. B(E2T) value Statistical Errors. The statistical error is the same as that of the cross section because of the linear relationship between excitation cross section and B(E2) value (equation 2.9). Systematic Errors. Besides the systematic errors from the cross section there were also uncertainties in the calculation of the factor of pr0portionality, the Winther and Alder functions GM” and g“. Contrary to the cross section the B(E2) value depends strongly on the maximum scattering angle and this uncertainty has to be considered. Table 3.2 lists the error sources and the contributions for the 40S case as an example. For the other cases the systematic errors are very similar and are not explicitly listed. The final results with total errors are listed in table 3.3. 3.6 Results and Discussion 3.6. 1 Observations Photons emitted from the fast moving fragments (11 a: 030) could be clearly dis- tinguished from photons emitted from the stationary target by their Doppler shifts. 60 Energy (keV) ‘2‘“: 1”: ,1...‘ J"; ‘;;‘;:. fit-n" 1.; 1“,." #1:)! ._ .. , _. . ~ hash-J" 40 80 120 160 40“ 80 120 160 Position (mm) Position (mm) Figure 3.17: Observed energies of 7-rays as a function of position without correction for Doppler shifting (left panel) and with the Doppler correction (right panel) for the 40S+”’7Au reaction. The target was located at 90 mm. The 7-rays near 547 keV are from the gold target, while those near 890 keV are from the (2fL —) 0;,_) transition in the projectile. Figure 3.17 shows the 7-ray energy spectrum as a function of position in the NaI(Tl) detectors for the 40S+197Au reaction. The left panel shows the 7-ray energies in the laboratory rest frame; i.e. before any Doppler shift adjustment. In this panel, the energy of the 547 keV (y —> g.s.) transition from the 197Au target is independent of position, while the energy observed for the 2f —) 03.3. 7-ray from the projectile 4oS depends on the position and, therefore, on the angle at which it was emitted. For forward angles a higher energy is observed, whereas for backward angles a lower energy is measured. The right panel shows the same energy spectrum, but each event has been corrected for the Doppler shift. Therefore this spectrum corresponds to the projectile rest frame (0 = 0.27c). In this panel, the energy of the 7-ray from the projectile is constant, while the energy of the target 7—ray now varies as a function of position. The Doppler-corrected, background-subtracted 7-ray spectra for all five nuclei 61 zoobfi'V'Il'YI'l'VYv' D 150 =0 p... O O l 0' O I Y t O ass , v=0.27c v=0.27c v=0.280 v=0.250 v=0.26c I x5 Counts / ( 1 0 keV) a: O a O I 0F 0 IVY—II" 20 . '- D 0 ‘1“ " F allllAlll‘A‘AA“ l L‘IIL‘LALALIA‘IAA ‘L‘L‘Al‘lll‘l ‘Ll‘L l... l I. “ I‘LL 500 1500 500 1500 500 1500 500 1500 500 1500 Energy (keV) Figure 3.18: Upper panels contain background subtracted photon spectra in the lab- oratory frame. The 547 keV (7 / 2+ —> 9.3.) transition in the gold target is visible as a peak, while the (2+ —+ gs.) transitions in each projectile are very broad. Lower panels contain Doppler-corrected, background-subtracted 7-ray spectra. studied here are shown in figure 3.18. All five spectra clearly show one photo- peak associated with each projectile. The measured energies of the 21* states and B (E2; 03.... —+ 21*) values are listed in Table 3.3. It should be noted that the B(E2 T) result obtained here for 388 is consistent with the lower limit set on the lifetime of the 2? state by Olness et al. [24]. In addition, the well-known energy of the 21* state of 388 [25, 26] was used to check the energy calibration procedure. N o excited states have been observed previously in 40"""S, but excited states have been reported for 44"“SAr. Crawley et al. [27] observed states in 44Ar using the 48Ca(3He,7Be) reaction and judged the 2;" state of 44Ar to lie at 1.61 MeV. The spectra in the study of Crawley et al. are quite difficult to interpret because the 62 background peaks are much larger than those from 44Ar. If the 1.144 MeV state, proposed here as the 2? state, was populated in the experiment of Crawley et 01., it would have been obscured by a peak corresponding to an excited state of 7Be. Mayer et al. [28] reported an energy of 1.55 MeV for the corresponding state in 46Ar from their work with the 48Ca(14C,1“O) reaction in agreement with the present work. 3.6.2 Comparison to Theory Self-consistent mean field techniques [17] predict permanent quadrupole deformations in 401428 of 62 ~ 0.25, only slightly smaller than those measured here (see Table 3.3). For 44’46Ar, Werner et al. [17] did not provide definitive predictions but instead showed that their two calculation techniques (Hartree-Fock+Skyrme and relativistic mean field) give very different answers for these two nuclei. The Hartree-Fock calcu- lations yield a significant prolate deformation (,32 = +0.17) for 44Ar and an oblate deformation (62 = —0.13) for 46Ar. On the other hand, the relativistic mean field calculations yield 52 = —0.13 for 44Ar and ,62 = 0.00 for 46Ar. The experimental B (E2; 0;“, -+ 2?) results agree better with the Hartree-Fock results, since a non-zero deformation is measured for 46Ar (52 = 018(2)) and a relatively large deformation is measured for 44Ar (32 = 024(2)). The effects of the N = 28 major shell gap persist in 46Ar because it is less deformed than 44Ar and its deformation and the energy of its first excited state are similar to 50Ti (E(2f) = 1554 keV, ,32 = 0.17, [29]). It would be of considerable interest to measure the 2;" state of 44S to see whether the N = 28 shell gap is still present even further from the line of stability. While the shapes of 40’428 can be understood with the mean field calculations of Werner et al. [17] which attempt to account for changes in single particle binding energies and residual interactions away from the line of stability, the data for all nuclei measured here except 46Ar can also be explained with shell model calculations 63 which use empirical interactions obtained from nuclei close to the stability line. These calculations were carried out in a model space in which the protons occupy the 0d5/2, 0d3/2 and 131/2 (3d) orbitals and the neutrons occupy the 0f7/2, 1123/2, 0f5/2 and 1121/2 (pf) orbitals. For many of the nuclei under consideration the dimension of the full 7r(sd)-1/(p f) model space is too large, and the calculations reported here have been truncated by leaving out the 0f5/2 and 1121/2 neutron orbitals. With this truncation the dimension for the 2+ state in 42S is 4335. For some nuclei such as 48Ca and 46Ar, this truncation can be compared to those performed in a model space which includes the 0f5/2 and 1121/2 orbitals, and the results for the orbital occupations and excitation energies of the 2+ states are found to be very similar. The Wildenthal sd— shell interaction [30], the recent F PD6 p f -shell interaction [31] and the WBMB sd — p f cross-shell interaction [32] was used. This latter cross-shell interaction successfully accounts for the properties of the N = 20 — 22 nuclei including the intruder state deformation in 32Mg [32]. The model space used and interactions are illustrated in figure 3.19. The B(E2) values were calculated using proton and neutron effective charges of 6,, = 1.356 and en :2: 0.658, respectively, which were chosen to reproduce the E2 transition strengths of the proton sd—shell transitions in 36S and 38Ar [25] and neutron p f—shell transitions in 48Ca [33]. In the top two panels of Fig. 3.20, the measured 62 values are compared to the results of the mean field calculations of Werner et al. and the present shell model calculations. The mean field calculations slightly underpredict the measured values for 40,428 and the shell model calculations slightly overpredict ,62 for these nuclei. However, the shell model calculation predicts that the 62 value of 46Ar is larger than that of 4‘l‘Ar, contrary to the downward trend in the data, which can be explained by the persistence of the N = 28 shell closure. The increase in B(E2) for the shell- model calculation is related to the crossing of the 0d3/2 and 131/2 proton orbitals 64 42 16 26 1cm >-< p—f shell P1/2 * Fm —0000— fm [ 00000000- :> sd-pf —0000— 20000: (13,2 s- -d shell 6+ Protons Neutrons Figure 3.19: Shell model space and interactions. The protons occupy the 7r(sd) shell and the neutrons the u(pf) shell. The model space was truncated in the calculations and the u0f 5 and lel orbitals have been left out. The arrows indicate the interac- tions used. The s- and2 p-shells below the sd- shell are not shown and are filled with 8 particles. 65 0'3? i f i r ‘: I E a a 1: iii 2 _ 0.2— — — — 6‘; t 8 1: a E : _ i f: 0 1:1 i 0-1: t r i 0.0- 4 — <> 1 9 r - - : m 1.5— -— — ' -— 2 _ _ - 2 v _ . A _ 1 >5 - - a an E 5 5 3:2. 4 2 1.0L— ” — — — m y— 2:; 7R. .1 l- - 388 408 428 “Ar “Ar Figure 3.20: The top two panels compare the experimental quadrupole deformation parameters [62] (solid points) to shell model calculations (stars) described in the text, relativistic mean field calculations (open diamonds) and Hartree-Fock (open squares) calculations. The bottom two panels compare the experimental excitation energies E (2+) (solid points) to the shell model calculations (stars). observed between 35K (which has a 3/2+ ground state [25]) and 37K (which has a 1/2+ ground state [25]). The bottom two panels of figure 3.20 show that the shell model calculations successfully reproduce the energies E (21*) in the nuclei reported here. The results presented demonstrate that a direct measurement of B(E 2; 0;, —> 21") is necessary to determine the nuclear collectivity, which is interpreted in terms of the deformation, and that the energies of the 2? states (without the B(E2) values) are not sufficient to deduce the deformation on the basis of systematics. For example, the global systematics of Raman et al. [34] give ,82 = 0.4 from the energies of the 2? states in 401428. The experimental ,62 deformations are significantly smaller. 66 [secondary beam 38 S 4”S 428 ‘4 Ar 46Ar Energy (MeV/nucleon) 39.2 39.5 40.6 33.5 35.2 Beam purity 0.99 0.65 0.55 0.99 0.99 Typical intensity on target (3") 50,000 17,000 1,800 50,000 27,000 Target thickness/(mg/cmz) 184.1 184.1 184.1 93.5 93.5 Energy loss in target (MeV/ nucleon) 9.1 8.4 7.9 5.1 4.9 Beam velocity (C) 0.269 0.271 0.276 0.254 0.262 023* 4.916 4.958 5.003 5.041 5.085 Efficiency (%) 4.66 7.24 7.28 5.36 3.75 Energy of first excited state (keV) 1286(19) 891(13) 890(15) 1144(17) 1554(26) 0(E2;03'.,_ -> 27,0131, 3 4.1°)(mb) 59(7) 94(9) 128(19) 81(9) 53(10) B(E2;03’.,. -> 21+) (e2fm4) 235(30) 334(36) 397(63) 345(41) 196(39) Iflzl 0.246(16) 0.284(16) 0.300(24) 0.241(14) 0.176(17) Table 3.3: Experimental parameters and results. The purity of the secondary beam is for reference only; the secondary fragments were positively identified on an event by event basis and only desired fragments were analyzed. The energy spread of the secondary beam was i3%. In summary, the energies and B(E2; 0:3. —> 2?) values of the 2? states of 381401428 and 44""‘Ar have been measured using intermediate-energy Coulomb excitation. The isotopes 40’428 are deformed, indicating the presence of a new region of deformed nuclei near N = 28. The data on the 21* state in 46Ar demonstrate that the N = 28 major shell gap persists at Z = 18. Both the mean field calculations and shell model calculations using empirical interactions can approximately reproduce the behavior of the 2? states of 401428. 67 Chapter 4 Direct Reactions In this chapter a short introduction to direct nuclear reactions is given. The main references are [35, 1, 36]. When a nucleon collides with a nucleus it can penetrate into the nucleus and form a compound system. These are compound nucleus reactions. On the other hand, nuclei have a relatively sharp surface, and it is likely that the incoming nucleon collides with a single nucleon, or a normal mode of nuclear motion, and the residual particle (which could be the same as the incoming particle as in elastic scattering or some other reaction product) escapes immediately. This kind of reaction is called a direct reaction because usually only a single (direct) interaction is involved in the scattering process. Since in a direct reaction no intermediate system is formed, the wavefunctions of the initial and final states overlap, and useful information on the initial configuration can be obtained from the final state. In order to extract nuclear structure information from direct reactions, the reaction dynamics has to be known. Cross sections depend on a nuclear matrix element, which contains both the wavefunctions of the nuclear states and the effective two—body interaction Veg mediating the transition between these states. Consequently, a priori 68 knowledge of the effective interaction is necessary in order to use direct reactions as a spectroscopic tool [37, 38]. Two approaches have been devised to obtain Veg. One approach is empirical and discussed briefly in section 4.1.2. The other is entirely theoretical and is outlined in section 4.1.3. Direct reactions can further be classified as inelastic scattering (e.g. (p, p’)) , as stripping (e.g. (d, p)) or pick-up (the inverse of stripping, e.g. (p, d)) reactions, and as knock-out (e.g. (p, p’ p”)) reactions. Each of these reactions has particular advantages for the study of certain aspects of nuclear structure. Stripping (or pick-up) as well as knock-out reactions are useful in studying the single particle nature of nuclei; in particular spectrosc0pic factors can be obtained. Inelastic scattering, on the other hand, is particularly effective in exciting collective states. One can imagine that the incoming projectile touches the nuclear surface and brings the drop (using the liquid drop model) into a state of oscillation, or if the nucleus is deformed, makes it rotate. The excitation cross section is then dependent on the degree of collectivity of the nucleus. A deformed nucleus (or a nucleus soft against vibration) is easy to excite, in contrast to a spherical (closed shell) nucleus. The results are somewhat model dependent, mainly because the projectile and the target interact strongly and perturbative methods can hardly be applied. In contrast, Coulomb excitation (see chapter 2) is well understood and largely model independent, giving very reliable results. Coulomb excitation is only sensitive to the protons in the nucleus, and a different experimental probe is needed to get information on the neutrons. Because of the Pauli principle, theplike-nucleon interaction is about 3 times weaker than the unlike- nucleon interaction in a nucleus. Therefore proton scattering (p, p’) in the energy range of E / Am 10— 50 MeV, which corresponds to typical kinetic energies of nucleons in a nucleus, is mostly sensitive to the neutrons [3]. Information on both the neutron 69 and the proton motion can be obtained by comparing the two experimental probes. Section 4.2.3 shows how the ratio of the neutron and proton matrix elements Mn/Mp can be extracted. 4.1 Elastic Scattering The simplest interaction between an incident particle and a target nucleus is elastic scattering. The particle’s direction of motion and / or state of polarization is changed, without loss of kinetic energy. 4.1.1 Optical Model Direct or shape elastic scattering occurs when the incident particle interacts with the nucleus as a whole. This interaction can be described fairly well by an average nucleon-nucleus interaction as single absorbing potential well: the Optical potential V(r) = (U + iW)f(r) . (4.1) U and W are the real and imaginary parts of the potential and f (r) is a Woods-Saxon form factor. This form is usually not sufficient to describe the elastic scattering data and surface W0 and spin-orbit V30 terms are introduced. V(’I‘) = V0(T) — V ' f(.’130) — Z {W ° f($w) - 4WDfid—:;f($0)} (4.2) h 2 1 d + (mwc) Vso(L ' 0');$f($30) I where —LZ‘€ 62 for r _>_ RC V00”) = z z e, 2 (4-3) fil’C—(3—fig) for r3120. Vc (r) is the Coulomb electrostatic potential of a uniformly charged sphere of radius RC = r0145. The other terms in equation 4.2 are the real, imaginary, and spin-orbit 70 parts of the optical potential. The form factors f (23,) are given in Woods-Saxon form 1 — ,A1 = _ with :13,- = r——:—: , (4.4) 1 + ext 61,- “1135) even though other functions can in principle be used. a,- is called the diffuseness parameter. Approximate values for the parameters are a. z 0.6 fm, r z 1.2 fm, V 2 50 MeV, W a: 10 MeV, and V30 z 8 MeV (see [36]). The real part of the optical potential describes the refraction of the incoming wave and the imaginary part takes into account all non-elastic processes through absorption. This potential is inserted into the Schrfidinger equation V29: + —(E — V)\p = 0, (4.5) whose solution is expressed in the form of an incoming plane wave plus an outgoing spherical wave \II o< e1“ + f(0) , (4.6) and hence the (elastic) differential cross section is given by do 2 E = If(9)I - (4-7) Practically one obtains the scattering amplitude f from the phase-shifts 6L “9) = i]; ZL:(2L + 1)(c2*'6L — 1)PL(cos(6)) . (4.8) Here, PL(a:) are the Legendre Polynomials. Calculations are done numerically. A very powerful computer code, which was used in this analysis, is ECIS [39]. 4.1.2 Effective Optical Potentials The usual procedure is to adjust the parameters of the optical potential (potential depths, radii, and diffusenesses), as given in equation 4.2, until the calculation agrees 71 strength V, W (MeV) radius r (fm) diffuseness 0 (fm) V 54.0 — 0.3218 + 0.4Z/A1/3 + 24(N — Z) /A 1.17 0.75 W 0.22E — 2.7 1.17 0.75 WD 11.8 — 0.25E +12.0(N — Z)/A 1.32 0.51 + 0.7(N — Z)/A V30 6.2 1.01 0.75 Table 4.1: Becchetti-Greenlees optical potential parameters [40] for elastic proton scattering. with the measurement. Often several sets of parameters reproduce the measured elastic cross section equally well, and the nuclear structure information extracted is model dependent. However, if data are available for neighboring nuclei, one can demand that the optical parameters vary slowly as one changes neutron number, proton number, and beam energy. A comprehensive analysis of a wide range of proton- scattering data for energies Ep below 50 MeV and A 240 was performed by Becchetti and Greenlees [40]. The optical parameters in their work were parameterized in terms of the neutron number N, proton number Z, and the incident lab energy of the proton E. The Optimum parameters found in [40] are given in table 4.1. The optical parameters obtained in this way are empirical and based on system- atics for stable isotopes. Thus it is not clear if these parameters are applicable when investigating nuclei with large neutron excess. Also, the lowest mass number A that was included in Becchetti and Greenlees’ fit around a proton laboratory energy of E, z 30 MeV was 56 which deviates considerably from the nuclei studied here. 4.1.3 Microscopic Optical Potentials A more fundamental approach is given by Jeukenne, Lejeune, and Mahaux (JLM) [41]. The authors derived a complex microscopic optical potential based solely on nuclear matter calculations. The derived potential can reproduce proton elastic scat- tering angular distributions, provided that the imaginary potential is adjusted by a 72 p. .. ..- I— .— p- n- I “‘6 P (9) Ill"! rIIlIIlI *1 val/gm .An Figure 4.1: This figure shows the fit by Becchetti and Greenlees for the proton scatter- ing cross sections on different targets around Ep 2 30 MeV. The lowest mass number included in this energy range was 56. 73 normalization Aw of about 0.8 [42]. The main idea in the J LM approach is to fold a more or less well known nucleon nucleon interaction 11 with calculated nuclear densities p resulting in a potential v0) = / p(r1)u(|r — r1|)dr'[3. (4.9) The subtleties lie in the determination of the different parts of the optical potential, such as the real volume and surface terms, and imaginary terms. (see [41]). With the cost of computing power becoming rapidly decreasing, realistic density calculations have become possible. Recent theoretical advances give promising results. For instance Kelley et al. [43] measured the elastic and inelastic proton scattering cross section on the radioactive isotope 38S. The cross sections were interpreted by fitting the data using the Becchetti-Greenlees optical parameters and applying the prescription by Bernstein [3] to obtain the ratio of neutron to proton matrix elements (see section 4.2.3). The resulting Mn/Mp is 2.06, which is unreasonably large (see section 4.2.3). A different analysis was performed on the same data by Alamanos et al. [44]. They used the JLM method with nuclear matter and transition densities obtained from shell model calculations. The resulting Mn/Mp is 1.58, which is much closer to the expected value of N /Z = 1.37. A possible source for the discrepancy of the two methods is that especially for the neutron rich isotopes the neutron and proton densities, and hence the corresponding potentials, have diflerent root mean square radii. This difference is not included in standard Woods-Saxon form factors, which are the same for both neutrons and protons. One might conclude that the use of an average phenomenological potential of Woods-Saxon form is inadequate for studies of neutron rich nuclei, since the micro- scopic approach seems to give more reliable results. 74 4.2 Inelastic Scattering Inelastic scattering occurs when the projectile interacts with the target leaving either the target or the projectile in an excited state. Energy is removed from the relative motion of the two particles and transformed into energy of the intrinsic motion of either particle. There are two simple pictures of this process. In terms of the single particle shell model one can think of a single nucleon lifted into a higher shell, whereas in the collective picture the scattered particle induces as surface vibration or a rotation if the nucleus is statically deformed. In this work we are interested in the second picture and the theoretical interpre- tation is done using the coupled channel formalism. 4.2.1 Coupled Channels The Schrédinger equation for the whole system of two colliding particles, represented by the wavefunction \P(r , If), is [36] (T- V036) +H(€))‘I’(r,€) = E‘1'(T,€)- (4-10) The (intrinsic) nuclear states x are defined by H(€)xa(§) = 6..)(..(€) , (4.11) where a labels the intrinsic states. The total wavefunction (including the relative motion) can then be written as a superposition of the intrinsic states $0.16) 2 Z ¢a(T)Xa(€) a (4.12) where the sum 62 goes over all states of the nucleus (discrete and continuum). The co- efficients of the expansion depend on the relative position of the two nuclei. Inserting 75 4.11 and 4.12 into equation 4.10 yields (T'— E + 6or )(War :0: Vaa’ wa’( T) a (413) with — —/5 as W o x. (ode (4.14) Equation 4.13 constitutes as set of coupled equations for the wavefunction of the elastic and all inelastic channels. In practice one can only include a few of the inelastic channels in the calculation and one compensates for the neglected channels by letting the interaction potential be complex. In our case, only two states have been included: the ground state and the first excited 2+ state. 4.2.2 Nuclear Deformation The interaction potential V(r, 6) depends on the character of the excited state. For example, a statically deformed nucleus can be described by a potential that depends on the orientation of the nucleus V = W? - R(9, 4)) and RU), a5) = 1200 + 5130099)) , (4-15) where 6 is the deformation parameter. Hence, 0 dV V = W? - R0) - flRoYz (0, (Md—7, - (4-16) The first term is the usual spherical optical potential, and the second term describes coupling between the incident and outgoing (inelastic) channels. 76 4.2.3 Mn/Mp In order to compare the neutron to the proton motion, one calculates Mn/Mp, which is the ratio of the neutron to proton matrix elements, which defined as Mn(p) = (JillzrfYAmilllle n(10) = / pygp)(7‘)7"\+2 dr, 0 where pig”) are the neutron (proton) transition densities. The B(EA) value is related to the proton matrix element through lMpl2 . 4. 2J5 + 1 ( 17) B(EA, J,- —+ J,) = Mn/Mp can be obtained by a comparison of a hadronic probe (in this case proton scattering) with an electromagnetic probe (Coulomb excitation) via a description by Bernstein et al. [3] Mn bp 6(7) 10’) bnN — = —— 1 —— — 1 M, 6,, < 6..., + 6,, 2 bp 6(1019’) 50> p’) N = _ ’ _ 1 1 b. ( 6.... + 6.... z _ N 500,1”) Z bp 600,100 — Z ( 5cm + an 5cm 1 . (4.18) Here, bum are the relative sensitivities of the (hadronic) probe to the neutrons (pro- tons) in the nucleus and 68m and 60ml) are the deformation lengths (6 = 6R) obtained via the electromagnetic probe and via (p, p’), respectively. The last equation can be interpreted in the following sense: The ratio of the matrix elements is equal to N /Z times a factor that is the ratio of the deformations measured via (p, p’) and Coulomb excitation. Because of the different sensitivities of the probes a correction term has to be added, which is given as the relative difference of the deformations measured with the two probes. In other words, if (p, p’) was a probe that is only sensitive to 77 the neutrons then bp = 0 and Mg = {Y 6(pm’) M,D Z 66m ' (4.19) From the collective model one expects a ratio Mn/Mp = N /Z , since the neutron and proton liquids move in the same way. In contrast, one also expects deviations from this simple picture, especially in the vicinity of closed shells. A nucleus with a single closed shell should not yield a value of Mn/Mp = N /Z , since e.g. the neutrons form a spherical closed shell and hence Mn = 0 [3]. However, core polarization restores the isoscalar character of the excitation to a large extent and brings Mn/Mp ~ N /Z , even for single closed shell nuclei [45]. Nevertheless a ratio Mn/Mp that deviates from N /Z is an indication of the importance of shell effects in the structure of these nuclei. 78 Chapter 5 Proton Scattering of 36’42’44Ar The availability of high-intensity radioactive beams creates new possibilities in nuclear structure studies through the use of direct reactions in inverse kinematics. Specifically the neutron-rich sulfur and argon isotopes have become accessible through projectile fragmentation of 48Ca, as described in section 3.1. Proton scattering in inverse kinematics was pioneered in an experiment by Kraus et al. [33], which measured the proton elastic and inelastic scattering on the fl-unstable isotope 56Ni. The low energy of the recoiling protons limits the target thickness to about 2—4 mg/cm2 and high beam intensities, of about 104 s”, are needed to perform the experiment in a reasonable amount of time (one or two days). Even with thin targets the energy resolution is only about 800 keV (FWHM). Since the nuclear states have to be resolved, an additional limit is put on the nuclei that can be studied. Usually in even-even nuclei the first excited 2+ state is well separated from the ground state and other excited states. Here, we focused on the even-even isotopes 42"MAL As a check the N = Z nucleus 36Ar was also investigated [46, 47]. The 36Ar nucleus has equal numbers of protons and neutrons (N = Z) and should yield a ratio Mn/Mp of unity by isospin symmetry [6]. The secondary beams of 42"MAr were produced as described in section 3.2.1. The 79 _ — — £2 “=- 3 T‘" - :3" ‘fl-f'm '_~_" ‘ ~ .7: ; .— >~ j a : an - — _ __ 8 _Contaminant 4 [Ll (less than 2% of total) j, . - .. — l .;l f a.» 1 Time of Flight Figure 5.1: The incoming particles were identified according to their time of flight and energy loss in the fast scintillator of the zero degree detector. The desired beam, in this case 42Ar, and the contaminants are completely separated. 36Ar beam was produced directly in the K1200 cyclotron, where the beam was in- jected from the Superconducting Electron Cyclotron Resonance ion source. The beam energy was adjusted with the help of a variable degrader and the A1200. The beam energies for all three isotopes were about E /A = 33 MeV. The exact values together with the target thicknesses and other experimental parameters are listed in table 5.2 on page 108. 5. 1 Experimental Setup The experimental setup in the 82 vault is schematically shown in figure 5.2. The incoming beam particles were tracked with two PPACs located 0.8 m and 1.8 m in front of the target. As in the Coulomb excitation experiment (chapter 3) the projectiles were identified according to their time of flight (TOF) as shown in figure 5.1. 80 — . — 5‘ ' T“WWW: (Energy. Position) TOP 5 ( Position,TOF AB E I i. :3. 7 W". 7 - I H ’ 7 30 m . Thin Plastics I PPAC PPAC Fast-ISlow- Scintillator Phoswich- Detector ” _ — SI - Telescope — (Energy, Position) Figure 5.2: Setup around the secondary target for the (p, p’) experiment. The target, a 2.7 mg/cm2 polypropylene (CH2),, foil, was located close to the center of the NSCL 32” scattering chamber. It was oriented at an angle of 60° with respect to the beam axis, thus providing an effective target thickness of 3.12 mg/cmz. The tilt of the target allows the protons, which are scattered toward angles of approxi- mately 80° in the laboratory system, to easily escape from the target. The kinematics of the reaction is discussed in section 5.1.4. The scattered beam particles were detected in a fast-slow phoswich detector and were identified as described in section 3.2.3, the only difference being that the de- tector had only one photo multiplier tube and had a smaller diameter of 3”, which corresponds to scattering angles of about 5°. Typical projectile scattering angles are on the order of 1° (refer to the lower panel in figure 5.13) and therefore all scattered particles are detected. 5. 1.1 Proton Detectors The recoiling protons were observed in a set of four telescopes, each of which consisted of a silicon strip detector (300 ,um) followed by two silicon PIN detectors (500 pm 81 each). The arrangement of the four telescopes is shown in figure 5.3. The telescopes were mounted at a distance from the target of about 23 cm. Telescopes 1 and 2 covered angles from 9,0,, = 67° — 79° and telescopes 3 and 4 covered angles from 6,01, = 70° — 82°. Position Sensitive Strip Detector The strip detectors were about 300 pm thick (see table 5.1) and had, as all other detectors in the telescope, an active area of 5 x 5 cm2. One side of the strip detector was segmented into 16 resistive strips, whose resistance was about 3 k0 over the strip length of 5 cm. Each of these strips was read out on both ends, and the position of the incident proton was reconstructed from the two signals assuming charge division. Thus the detector provided position information in 2 dimensions, where the direction perpendicular to the strips was given by the strip number itself. Within a strip the measured position resolution was about 0.5 mm for a 5 MeV signal. This should be compared to the strip width of about 3.1 mm. The telescopes were mounted so that lines of constant 0 were roughly perpendicular to the strips. Therefore a much better angular (0) resolution was achieved as compared to an orientation with the strips along the lines of constant 0 (better by a factor of 3.1/0.5 at 5 MeV). Angular resolution is crucial for the distinction of elastic from inelastic scattering. The unsegmented back side of the detector was used to obtain an energy signal. Figure 5.4 shows a schematic of the telescope and the working principle of the strip detector. Silicon PINs Since protons with an energy of more than about 6 MeV were not stopped in the 300 pm thick strip detector, it is followed by two 500 pm thick silicon PINS. These three silicon detectors st0pped protons up to an energy of about 14 MeV. The com- 82 Figure 5.3: Schematic 3D view of the setup around the secondary target for the (p, p’) experiment. The beam is going from left to right. The telescopes are labeled from one through four starting with the top right-hand telescope and counting clockwise. Telescopes one and two cover the same scattering angles, as do three and four. 83 Position Sensitive Si-Strip Detector with Resistive Readout non resistive O resistive : Energy r I 0—5—1 9 ' . . .’.—0 Y2 Y1 Y1 Y1+Y2 Position = 500 um Si PIN 500nm? SiPIN 30011111 f‘lLllllllLLllLlJ] Figure 5.4: Position sensitive silicon detector telescopes. The telesc0pe composition is shown in the bottom part and a schematic of the operating principle of the strip detector is shown in the top part. 84 Telescope Strip PIN 1 PIN 2 1 305 471 475 2 296 467 464 3 299 462 469 4 301 462 476 Table 5.1: Detector thicknesses in um. bination of the three signals allowed the identification of protons that had a higher energy and thus were not stopped. This was achieved by comparing the proton energy deduced from the measured energy loss in the first two detectors (E_from_E2), with the sum of all detectors (E3). If E_from_E2 was equal to E3 then the proton had been stopped in the third detector and the total proton energy was given by E3. If, on the other hand, E3 was not equal to E_from_E2 then the proton had passed through all silicon detectors and the proton energy was reconstructed from the energy loss in all three detectors, using the known relation between the proton range in silicon as a function of energy [48]. In order to reconstruct the proton energy from the energy loss the detector thicknesses had to be known. They are listed in table 5.1. Energy and Position Calibration All proton detectors were calibrated with a 228Th oz-source. Figure 5.5 shows the a—energy spectrum measured with one of the silicon PINs. Indicated in the plot are the a-energies and the corresponding parent nuclei. The peak centroids and errors were obtained by fitting the sum of a Gaussian and a skew Gaussian to the spectra. The skew Gaussian (an exponential tail convoluted with a Gaussian resolution) fits the low-energy tail that can be observed in figure 5.5. A sample calibration curve is shown in figure 5.6. The position along the strips of the strip detectors was calibrated using a mask containing a set of 1 mm wide slits separated by a center-to—center distance of 6 mm. 85 °o 4L9, \ 4% Q ° s" as" s 0;)! 6hr %\ «08) 59 56° bath ,9 {E 1; g (X - Energy (parent nucleus) 00 m N s, s a, *2 E a; E 2 2 ,0 8.78437 MeV(212Po) m [\ N 00 8 °° ‘>’ 38 m 3 o o. vi tr} 2 \o / 00 [x O V) q \0 Ln“... \ average = 6.06171MeV (weight=intensity of lines) Figure 5.5: A 228Th a-source was used for the energy calibration of the silicon detec- tors. The oz energies and the corresponding parent isotopes are indicated. 86 I l T l 9 — Backside of Strip Detector 1 _ a 8 — — E >. __ _ g 7 L5 6 — _ E: -0.0038(55) + 0.0067207(58)-channel 5 — l l | I ‘ 800 1000 1200 1400 Channel Number Figure 5.6: Energy calibration of strip detector backside for telescope 1. A linear fit is shown in addition to the data points. The error bars on the data points are smaller than the size plotted symbol. 87 O fifflf‘v" (Psi-«16M! «1 I-: 7 a": I C K J . ‘ . I ‘ ‘ I a. . .. " ' . J ' i t :- 1’ . . . . . . . s ‘. . . r . . . I . n‘ c a o ( I . . .0 ‘ ' .c O Perpendicular Position (mm) I . a I -10 — g: +- _20 ._ I __ 3 1‘ I ' It 3 5! -30 l l I l -20 O 20 Position Along the Strip (mm) Figure 5.7: Position calibration of the strip detector. The position perpendicular to the strips was randomized. One strip was not working properly and was excluded from the analysis. The distance between the 1 mm wide slits in the mask is 6 mm (center-to-center). A raw position was calculated from the two measurements on either end of a single strip as x = yA/(yA + 313), where y“; are the measured signals. The true position was obtained by fitting the raw positions, measured with the calibration mask, to the true position using a 3rd order polynomial. A calibrated 2D position spectrum, with the mask on the detector is shown in figure 5.7. Proton Identification Protons with an energy of more than 6 MeV and less than 14 MeV passed through the strip detector and were stopped in either the first or the second PIN. Figure 5.8 shows the energy loss in the individual detectors as a function of proton energy. This allows the identification of protons through the AE—E method illustrated in figure 5.9 (lower panels). Proton bands can be seen clearly. Protons with an energy of less than 88 N O t—d M UI Proton Energy (MeV) c: -I>- O\ 00 [\J Detector Signal (MeV) l \\ l \x I— \ \\\ — Strip \\ JIN 1 (4.92) \g‘ A PIN2(4.69) _ (2.29) \V V ‘V‘ ...... 7 __ \ _ ‘ \ — - Strip \ x g - - PIN 1 — — PIN 2 — l l l l l I l O 2 4 6 8 10 12 14 16 Detector Signal (MeV) I I j I I T I j T I I I I I I T I I I — — - Strip 4 - - PIN 1 ,’ — PIN 2 ,’ __ /\ ,' // \\ ’1 / \II / ,\ i— / l \ / ,’ ‘ \ __ / I / I’ / I ’111Lll’llllllllllll O 5 10 15 20 Proton Energy (MeV) Figure 5.8: Signals that are registered by the individual detectors in the telescope as a function of proton energy. As an example, a 12 MeV proton will lose 2.3 MeV in the strip detector, 4.9 MeV in the first PIN, and 4.7 MeV in the second PIN. 89 6 MeV were stopped in the strip detector and they could therefore not be identified by the previous method. However, their time of flight (TOF) to the strip detector was measured and different isotopes can be identified using a E—TOF matrix. This is Shown in the top right-hand panel of figure 5.9. The time resolution of 2 us did not allow a complete separation of the different isotopes, but the band structure can be observed. In order to remove all non-proton events from the analysis the identification of the scattered beam particle in the zero degree detector was required. As can be seen in the left-hand panel of figure 5.9, this condition alone removes almost all non-proton background. This can be understood by considering the possible reactions that lead to parti- cles other than protons being emitted into laboratory angles 0101, around 80°. One possibility is the occurrence of violent reactions on the carbon nuclei in the target. These reactions, however, lead to the breakup of the projectile and hence are rejected by the requirement of observing the beam particle after the target. N ucleon transfer reactions such as (p,d),(p,t), and (p,3’4He). can not produce the non-proton back- ground Since they are very forward focused because of the negative Q-value. (The total kinetic energy has decreased by lQl, which is on the order of several MeV.) For instance, the Q-value for the reaction p(3°Ar,35Ar)d is Q = -13 MeV and from equa- tion C.20 a maximum possible scattering angle for the deuteron in the laboratory of around 03;?" = 32° is inferred. Hence these particles can not hit the detectors that are mounted around 010;, = 80°. The groups of events marked A and B in the left-hand panels in figure 5.9, which are disconnected from the proton bands, are protons that miss the last PIN (B) or the first PIN (A). These protons enter the telesc0pe close to the edge of the detector and not perpendicular to the detector plane. Consequently, it is possible to traverse 90 the strip detector and miss the PIN, which is located about 5 mm behind the strip detector. The energy for these events has been reconstructed from the measured energy loss. 5.1.2 Beam Particle Tracking Beam particle tracking before the target is essential for an accurate determination of the scattering angle, which in turn is needed to distinguish elastic from inelastic scattering. The tracking detectors were calibrated using a hole mask with a grid of 2 mm diameter holes separated by a center-to—center distance of 1 cm. At the center of the mask was an additional fine grid with 1 mm diameter holes separated by 2.5 mm. A sample calibration Spectrum is shown in figure 5.10. New Tracking Method Unfortunately the in-beam position resolution of the PPACS was not aS good as expected. The reason was an (unexplained) strong rate dependence of the position Signal from one of the detectors. A test with an a-source, prior to the experiment, gave a resolution of about 2 mm (FWHM). With the beam, at a rate of about 30,000 particles/S, the resolution deteriorated to about 8 mm for PPAC 1 and to about 3.5 mm for PPAC 2. With these resolutions it would have been impossible to separate the first excited state from the ground state in telescopes 3 and 4, which depend more on good tracking than telescopes 1 and 2 because of the orientation of the target. The target was rotated, so that the lowest energy protons could traverse the smallest target thickness. Thus the normal to the target surface should point as much as possible towards telescopes 3 and 4 which will see the lowest energy protons, Since these telescopes are located at larger laboratory angles. This however means that a good position resolution of the tracking detectors iS crucial, Since a small change in the 91 Identification of incoming No identificaton of beam particle after target isotope after target 30 r I T I I I I I 25 __ E - TOF ..- * ' _, 9 0.) ‘ -.. ... E5 15 — . ., -- — g 10 _- _* ‘ -- — 0 -~ -- “ l ‘ . - . . . -. | 0 5 10 15 20 0 5 10 15 20 Time of Flight (ns) 30 I I l I I I I I I I 25 — AE (strip) - E -- 4He .. - 3 ~-:‘ 20 — -- H° 3.3., - 15 T -- -' . z. :14 g 10 — -- a 2 5 I- B ‘1‘»: -- n m annual-r“ “ g 0 l". i 4 + IL l “fi' i 'r i ‘r E; 25 _ AE (strip+pin A) - E -_ - 1' Q) 5 20 — -— 3 -I 2 H - 15 '_ J'- l H S. ' . ‘ .—l H (a... 10 _ -"' . ..‘: *‘.!I'. 'd 5 - 7 ,4} " . " I; ... ' _ ._ 0 .1! | I l | l . -L - -.'l l l l 0 5 10 15 20 25 0 5 10 15 20 25 30 Energy (MeV) Figure 5.9: Particle ID spectra in the telescopes. The identification of the incoming beam particle after the target, in the zero-degree detector AE—E spectrum, is sufficient to remove the non-proton background from the analysis. Particle ID spectra without any out are shown on the right-hand Side and bands from protons, deuterons, tritons, and even 3"’He can be seen. With the requirement of observing the beam particle after the target (left-hand Side) only the proton band remains, with very little background. See text for the groups of counts labeled A and B. 92 30 - I I .L t — E 20 —- II _ g 10 — a — .9 (L E -10 — ‘5 — E 20 > - _ Target Frame _ -30 — I I I I I J ~40 -20 0 20 40 Horizontal Position (mm) Figure 5.10: Calibrated PPAC position spectrum. The shadow of the target frame is visible. Broken ofl' drill bits can be seen in two of the small 1 mm holes close to the center of the target frame. The separation of the large (2 mm) holes is 1 cm and the 1 mm holes in the center are separated by 2.5 mm. The L—shaped pattern (the lower part is barely visible behind the target frame) served to check the correct orientation of the reconstructed position. The incident rate was about 600 36Ar/s at E/A=33.6 MeV. The applied voltage on the detector was 550 V and the gas pressure was 5 Torr (2,2,4 Trimethylpentane). 93 position on the target results in a large change in angle at which this point is viewed from the telescopes as illustrated in figure 5.11. On the other hand telescopes 1 and 2 “look” almost parallel onto the target surface and are therefore not so dependent on the particle tracking, even though they also benefit from good tracking results, Since the beam direction itself enters into the determination of the scattering angle. Since the usual method of calculating the beam particle trajectory (i.e. one cal- culates the trajectory assuming the beam particle crossed the two points that were measured by the tracking detectors) did not lead to resolved states in telescope 3 and 4 a novel technique was employed to obtain the trajectory, which gave satisfactory results. This technique, which uses information on the beam itself (i.e. the relation between the position of the particle and the slope of a particle trajectory), is described in appendix F. 5.1.3 Electronics The electronics for the strip detector constituted the largest part of the setup, since for each strip detector 32 channels (2 - 16 strips) had to be read out. In the electronics diagram (figure 5.12) the strips are labeled 1A through 16A and 1B through 16B, where A and B correspond to either side of one of the 16 strips. The signals from the strips were first amplified by charge sensitive preamps developed at Washington University and built by PICO systems (Kirkwood, MO). The preamps had a feed- back capacitance Cf of 1 pf (Von, oc Q/q, where Q is the released charged in the strip detector), which resulted in a high Signal gain. The preamp Signals were fed into Shaping amplifiers with a Shaping time of 300 ns. This small shaping time was necessary because, contrary to capacitive noise, the noise Spectrum of the resistive strip is high at low frequencies. The shaping time could not be lowered below that value, Since below that integration time, high-frequency pickup in the shapers be- 94 / low energy; *’ Short path sag, § high energy; ‘ long path Figure 5.11: Effect of target orientation and beam spot Size on the angular resolution a for the different telescopes. Telescopes 1 and 2 have a better resolution, Since they “look” almost parallel onto the target surface. The target angle was chosen so that the lowest energy protons have to traverse the smallest target thickness as indicated in the lower left-hand Side. The protons scattered towards telescopes 1 and 2 have to traverse a large amount of target material (as shown in the inset), but the proton energy is sufficiently high. Protons scattered towards telescopes 3 and 4 have a lower energy, but the target material is small enough to still allow a good energy resolution. The drawing is not to scale. 95 came the dominant source of noise. (The shapers have a high channel density of 16 channels per single width CAMAC module and therefore the circuits are not built for high-frequency applications.) This shaping time gave sufficient position resolution of about 0.5 mm for a 5 MeV signal. The Shaped Signals were digitized by 16—channel CAMAC peak sensing ADCs (P/S 7164H). The chosen Shaping time would have resulted in a poor energy resolution (about 200 keV), if the energy had been determined from the sum of the two Signals belonging to one strip. However, the energy Signal in the strip detector was obtained by feeding the Signal of the unsegmented back Side (hence no resistive noise) into a preamp followed by a shaper with a shaping time of about 5 us. The same setup was used for the PIN detectors. This resulted in a good energy resolution of about 50 keV (F WHM) for all Silicon detectors. The fast signals from the shapers for the strip backside and the PINS were input into constant fraction discriminators (Tennelec TC 455), whose output Signals were ORed (for each telescope separately). This telescope-OR had a width of 200 ns and if within that time a beam particle Signal was observed a master trigger was generated. For each telescope a coincidence bit was set and all detectors belonging to a telescope that triggered were read out. In addition to particle-proton coincidences ’downscaled’ (1 / 500) beam particle singles (i.e. events where only a beam particle was registered) were also measured for the normalization of the scattering cross section. The readout of the zero degree detector is essentially identical to that described in section 3.2.3. 5.1.4 Kinematic Reconstruction The purpose of the present experiment was to obtain the (p, p’) differential cross section do/dQ as a function of center of mass scattering angle 06",. In order to separate the elastic from the inelastic events the excitation energy of the particle, or 96 _ _ Eon—.2... in ._ R: “as E 358.5 >3: 9»: 02 5355:35 oma 9:33 H mm: ...HEEtSAQ ...:EE .5350 u EU E523 omen—.3 5:356 335 E @352 u «Emu 5:355 3&5 E uEF n be... 3583 4.85 552:6 1:35 a. ns.—:2 u UQ< 53550 850 n 00 V.M.N._ H.. . IIIIIIIIIIIIIIIIIIIII s2... n: in 030 8.82 is I'— uah _asm .— 55“ p m2 59;}: ...Su: ................................ _ Figure 5.12: Electronics Diagram. 97 equivalently its mass, after the scattering had to be determined. These two quantities can be obtained from the measured quantities 010,,(1H) and Elab(1H) through a Lorentz transformation. See appendix C and section 5.3.1 for details. Inverse Kinematics For a better understanding of the kinematics it is illustrative to look at the kinematic lines, which connect lines of constant Q-value in a plot that Shows energy versus scattering angle in the laboratory system. The kinematics in this experiment is different from normal kinematics where a heavy particle is bombarded with lighter one. Here, the light particle constitutes the target and one speaks of inverse kinematics. The kinematic lines for the p(3°Ar,36 Ar) reaction at E /A = 33.6 MeV are shown in solid in figure 5.13. The top part Shows the relation of lab energy to lab angle for the recoiling protons and the bottom part for the scattered beam particles. The lines of constant 06", for the protons (top part of figure), Shown dashed, are almost independent of 0,0,, in the range covered by the telescopes. Therefore a very good angular resolution was obtained in the center of mass, Since the energy in the laboratory system was measured with high precision. Figure 5.13 also illustrates the importance of good angular resolution, Since good energy resolution in the lab system is not sufficient to distinguish the excited state from the ground state. The bottom part of figure 5.13 Shows the kinematic lines of the scattered beam particles. The intrinsic beam energy Spread of the secondary beam, typically on the order of 1%, together with a high beam emittance make the use of a magnetic spec- trometer, such as the S800, to detect the scattered beam particles (without measuring protons) impossible, Since even with particle tracking the nuclear states could not be resolved. Another drawback is the background on the carbon atoms in the plastic 98 E,ab(36Ar) (MeV) l l l 36Ar(p,p') _. Telescopes l, 2 EM: = 33-6 MCV : = (E =1.97 MeV) C C C " Telescopes 3, 4 Blah (1H) (MeV) U! 50 60 70 80 90 100 1210 — p—s N O O I p—A H \O O I 1180 — l e =55° " l | "i l | 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 61ab(36Ar) (deg) Figure 5.13: Kinematics for the reaction 36Ar(p, p’) at E /A = 33.6 MeV. 99 target, which would be indistinguishable from the scattering on protons for small angles. In addition the recoil energy from the photon that is emitted after excitation further reduces the resolution. 5.2 Simulation A simulation program was written in DEC FORTRAN [49]. The program simulates the full detector response, the geometry of the setup, the beam emittance, as well as energy loss and angular straggling of the recoil protons in the target. A description of the program is given in appendix D. 5.2.1 Examples The simulation can be used to investigate the influence of certain effects, such as energy loss in the target or detector resolutions, on the final energy and angular resolution. In the following a few examples are given: Figure 5.14 shows the resulting Spectra for an ideal case, i.e. all detectors have perfect resolution and there is no straggling or energy loss in the target. The top two panels Show the kinematic lines, which are identical to the ones Shown in figure 5.13, except that here the detector acceptance restricts the scattering angle in the lab. The middle panels show the (reconstructed) excitation energy over the (reconstructed) center of mass scattering angle and the projection of these spectra on the energy axis is shown in the bottom two panels. In figure 5.15 all target effects have been included. The differences between the spectra for telescope 1 and telescope 3 are due to the target orientation. Figure 5.16 shows the result of the simulation when all target and detector eflects are included. One can see that the resolution after the inclusion of the detector effects is not much larger than that with only the target effects for telescopes 1 and 2. However, 100 the resolution in telescopes 3 and 4 shows a strong dependence on the resolution of the detectors, especially the position resolution of the tracking detectors. Figure 5.17 shows measured data and the agreement with figure 5.16 is quite good. 5.2.2 Efficiency The simulation program can also be used to calculate the proton detection efliciency (acceptance of the telesc0pes). Figure 5.18 shows the calculated efficiency for tele- scope 3. In the central region around ch = 25° the efficiency is nearly independent of the scattering angle 6m, and the value corresponds to the geometrical efficiency of the setup given by width of detector 27r(distance to target) ( 13 / 16) - 5 cm 27r - 23 cm 0.028 . 22 The factor of 13 / 16 was introduced, Since for this telescope 3 out of 16 strips were not working and hence the active width of the detector was (13/16) - 5 cm. For scattering angles close to the edges of the detectors the efficiency goes from its maximum to zero in a manner that can not be obtained through geometrical considerations, Since here the beam spot size and the angular spread of the beam are important. Most of the data points are not affected by the correction for the detector efliciency. 5.3 Analysis 5.3.1 Excitation Energy From the measured laboratory angle 6105 and the measured proton energy EM, the excitation energy was reconstructed using the equations given in appendix C. For each 101 91,1,(deg) Olab(deg) 70 75 8O 85 70 75 80 85 20 I I I I I I I I 9 15 *— -"- '— i g 10 — -~ —- LL] 5 __ -_ .4 0 e l l I I : I 'r 2 '— -"" —- "- § 1 - -- n S’ IV 0 _ -.. .- LL} -1 ._ -_ __ _2 I I I I I I J J o 10 20 30 4o 0 10 20 3o 40 50 9...,(deg) 9cm(deg) Telescope 1 Telescope 3 12x103 — __ ‘ 10 — __ — %, 0;... x 8 — __ — O : 6 — _ _ 4 - 2: -- - 5 2 — 7* n l . 0 I I I I I I -2 -1 0 l 2 -2 -1 O 1 2 3 Excitation Energy (MeV) Figure 5.14: Simulated Spectra under ideal conditions. The left-hand panels Show the spectra for telescope 1 and the right-hand panels Shows the spectra for telescope 3. The top row depicts the spectra measured in the laboratory system. The middle two spectra Show excitation energy plotted over the center of mass angle and the bottom panels Show excitation energy spectra (projection of the middle panels on the energy axis). The t0p row reproduces the kinematic lines shown in figure 5.13 102 6...,(deg) 91,,(deg) 70 75 80 85 70 75 80 85 -2 l l J l l l l l 0 10 20 30 40 0 10 20 30 40 50 06m(deg) 0cm(deg) Telescope 1 Telescope 3 250 — -- %200 — _ x 3150 .— __ .2100 — _ G 5 50 — J _. 0 l l -..4L_ , -2 -1 0 1 2 3 Excitation Energy (MeV) _Figure 5.15: Same as figure 5.14, except a 2.7 mg/cm2 polypropylene target was Included (energy loss and angular straggling). 103 e,,., E =1.167 (40) 0 (27° < 6cm< 33°) 54 E g 20 — — U ' I I O O O . . 0, I I . L‘ . . ,_,.,;:S.T...... 0 1 Excitation Energy (MeV) Figure 5.20: Shown are excitation energy spectra for 42Ar and 44Ar. The excita- tion energies are reproduced. The adopted values are 1.2082(3) MeV for 42Ar and 1.144(17) MeV for 44Ar, where the latter was measured in this work. 109 31'935 then inehr obser surer rneas At a beha toth cour exac addi CTOE tlie beai are the pro 83M the areas, the known target thickness, and total number of incoming beam particles were then used to calculate the differential cross section (see equation E.5) for elastic and inelastic scattering. Figure 5.21 shows the resulting cross sections for the known case of 36Ar. One observes that the cross section for elastic scattering does not follow the previous mea- surement by Kozub [46]. Similarly the inelastic cross section, which is compared to a measurement by Johnson and Griffiths [47], seems to fall below the older data points. At about 28—29° the measured cross sections drop by about a factor of 2. A Similar behavior can be seen by comparing the Becchetti-Greenlees optical model calculation to the 44Ar data which are shown in the lower part. The same is observed for 42Ar. Of course, one does not expect the Becchetti-Greenlees calculation to reproduce the data exactly, but especially the height of the maximum might be better reproduced. An additional reason to suspect a problem is the systematics of the other argon isotopes. Cross sections for 40Ar [50], 43Ar [51] do not Show such a drop in cross section and the height of the maximum around 66", z 40° is almost independent of isotope and beam energy. In the following different possible explanations for the drop in detection efficiency are presented: The same angle in the center of mass (for all isotopes) corresponds to the same proton energy in the laboratory, which gives some credence to an electronics problem that only occurs if the proton energy is above a certain value (which is around 8 MeV). This value can be obtained from figure 5.22 which shows the relation between the scattering angle in the center of mass and the laboratory proton energy. Figure 5.22 Shows that the relation is almost independent of excitation energy, which can also be seen in figure 5.13. Hence the elastic and inelastic cross sections dr0p at the same angle. However it also follows from figure 5.22 that this energy (8 MeV) does not correspond to any of the detector thresholds and it is therefore unclear what 110 1000 :_ I o I I I I I _E 6E o Kozub E 4- Ego a Johnson - _ if I . + '- A 2__ . o I thlS work (0gs ,2?) _ ‘0'] I B 100 _— o -: g 6; I i 0 !0I a CI 4: I i 0 I! f 2 E 2- {Ch I {iii - b I: in D f 1310:— HEI IIEEID D] Dc] ‘5 6E 36 [:1 D E 4: ElA=£MeV } Iiiiff I 2" l l | l l I ‘ 15 20 25 30 35 40 45 50 N I I TPITIWT on I I’I IIIII dO'ldQ (mb/sr) b) I 44 10 Ar f II BIA = 33.2 MeV I l l 20 25 30 35 40 45 Gem (deg) 00 Illll lllll p—d (It LII 0 Figure 5.21: Comparison of the measured cross section to previously published data. One can see that neither the elastic not the inelastic cross section agrees with the old data for angles larger than 28°. The lower graph Shows the large disagreement between the Becchetti-Greenlees calculation and the measured data for 06m > 28° for 44 AI‘ 111 Figux scatt the f Show ener; saw was Prol for 3 Can: the [Hi DO 25 I I I I — 13": 0.0 MeV / 20 - — - E’: 1.97 MeV , ’— 2 15 — , ’ 4 A , / 38° 5 10 _ _ 3 33° Lu 5 — 25° - 0 | | l O 10 20 3O 4O 50 Gem (deg) Figure 5.22: Relation between the proton laboratory energy and the center of mass scattering angle. The energies at which the proton is not stopped in the strip detector, the first PIN, and the second PIN are indicated. The corresponding angles are also shown. The insensitivity of the relation between EM, and 06m on the excitation energy E“ should be noted (the two curves are almost on top of each other). The same relationship is obtained for 42”MAr, since the beam energy per nucleon E/A was the same for all three isotopes. This, with the assumption that the electronics problem depended solely on the proton energy, justifies the use of the scaling function for all isotopes and all nuclear states. caused such a discrepancy. For each detector (strip, PIN 1, PIN 2) the time between the zero degree detector hit and the detector response (TOF) was measured. By inspecting the Energy-TOP matrix it was confirmed that the drop in cross section occurs at a proton (lab) energy away from any detector thresholds. (If, for example, the coincidence gate was not setup properly for the PINs then a problem should occur as soon as the proton energy is high enough to trigger the first PIN (6 MeV) and one should observe a drop in cross section at the corresponding angle, i.e. 25°.) Another explanation could be that at a certain scattering angle the beam particle misses the zero degree detector and therefore no coincidence could be registered. This possibility can be ruled out, because the radius of the zero degree detector is much 112 g0 thI that I Whlcl \\ most (8 .\I the ( Out be p for - Cur‘ obt exp the Fat larger than the maximum scattering angle. In addition, the spread in beam spot size and the angular spread of the incoming beam are larger than the deflection caused by the scattering and no sharp transition should be observed. Could it be that an object obstructs the flight path of the protons? This not possible, for several reasons. Firstly, one expects a smoothly vanishing cross section and not a step function with a non-zero cross section after the step. Secondly, any edge such as the target frame would degrade the resolution, since some protons will go through just a small part of that object. However the resolutions above and below that critical angle are the same. Lastly, the object should cover half the telescope, which is very unlikely to be missed during the setup or take down of the experiment. What could cause about every second event to be lost if Elab(1H) 2 8 MeV? The most likely explanation is that no coincidence gate was formed. The threshold energy (8 MeV) and the rate (about every second event) remain unclear. In conclusion, several possibilities have been investigated, but none can explain the observed behavior. Regardless of this, an experimental problem can not be ruled out and the evidence of the previous measurements and systematic trends can not be neglected. Therefore we decided to use the previous data on 3°Ar as a calibration for the detection efficiency. The resulting scaling curve is shown in figure 5.23. The curve was obtained by using the optical model parameters given by Kozub [46] to obtain the cross section for elastic scattering for the same angles as were used in this experiment. The calculation was used instead of the experimental values, because there are not corresponding experimental data points in Kozub’s work for all points in this work. The rather surprising result is shown in figure 5.23. Shown there is the ratio of the current measurement to Kozub’s data. The fitted curve is of the form Q—IUQ £1113 f(0) = we + 1121 -erf ( ) + 21240 + 111502 , (5.3) 113 1.2 _ I I I I I I _# ............ — Fit (with error lines) 1.0 ~— .. -I . Data Points a €0.23 _ .................. b! '2 0.6 - b ....... -c 0.4 — 0.2 — _. 36 00 AI I I I I I I 15 20 25 3O 35 4O 45 50 90m(deg) Figure 5.23: Ratio of the present data to the ones by Kozub. The fit and the corre- sponding error lines are explained in the text. where the w,- are fit parameters and erf() is the error function erf(:z:) = face—I2 dd: (5.4) —00 that allows for the smooth step in the fitted function. The error lines (dotted in the figure) were obtained via 02W» = 2 (1’1) (3) of. , (5.5) ik dw, dw,c where 03,, is the covariance matrix obtained through the fitting procedure. Only data points above 28° are affected by the scaling procedure. Data points below that angle are not scaled. The previously described scaling function has been applied to all isotopes (36’42’44Ar) and the resulting differential cross sections are shown in figures 5.24 (3°Ar) and 5.25 (42’44Ar). 114 I I 1000 — oblate ‘g g — — prolate i A —- - B2=0.256 - ‘53 2 - E 100 E s 4 , = B ‘ H _ '° -~.‘:. - r 10 """""" “l- - l “:- ° 36 \. E 4 Ar I L r s _ 10 20 3O 4O 50 Gem (deg) Figure 5.24: Different fits are obtained depending on the chosen shape of the nucleus. However, the measured data do not allow to distinguish between the two cases, even though it seems the oblate shape fits the inelastic data points better. Almost no difference is observed for the elastic channel. 5.3.3 Deformation Parameters and Mn/Mp The computer code ECIS [39] was used to extract deformation parameters from the data by fitting the elastic and inelastic cross sections, where V, W, and WD (refer to equation 4.2), in addition to the quadrupole deformation parameter ,62, have been allowed to vary during the fit. Other combinations of parameters have also been tried with essentially the same results. In each case the deformations of all potentials have been set to the same value and have also been kept fixed to one another during the fitting procedure. The starting point for all fits was the Becchetti-Greenlees prediction [40]. The results of the fit are displayed in the figures as solid lines. In the rotational model prolate and and oblate shapes fit the data equally well, as shown in figure 5.24. In the vibrational model the quadrupole deformation parameter fig corresponds to the phonon amplitude and as in the rotational model positive and negative values were tried and they also fit the data well. The obtained deformation 115 I I I I I I 1000:— . ——fit 3 E . --BG 5 c I _ e100.— ,. -x __ fl E \ /”! ‘5 V : x/i : g - I _ B 10:— i _= ‘3 E 5 :42 l I 1 ArI I I I I I 15 20 25 30 35 40 45 50 1000? $1005— 5 E Cl I "U B 10==—""1 .3 'D E .. E :44 I 1 ArI I I J I I 15 20 25 30 35 40 45 50 90m(deg) Figure 5.25: Measured cross sections and ECIS fits. Also shown are the optical model calculations for the elastic channel using the Becchetti-Greenlees optical potential parameters. The fit of the elastic data points for 44Ar is not as good as for the other two isotopes, but it has been verified that the extracted deformation fl; is quite independent of the actual Optical model parameters. 116 Figur was a samp One 5 same- Parai defor thou meal Show agre. ljelo. Calm p011) SeCti I. Were tab}, 0.5 r l I I I l -O- electromagnetic A (p,p')prolate ® (p,p') vibrational (neg) 0 4 § V (p,p') oblate — (p,p') literature ' l O (p,p') vibrational I (p,p') adopted here .... 03 '— Q I 201 _- <2“.I B a 0.2 — _ 0.1 —— A 0.0 I I I I I ArgonI 36 38 4O 42 44 46 Mass Number Figure 5.26: Deformation parameters obtained with the various fits: The value that was adopted in this work is the average of all four values. The assigned error bar is the sample variance of the data points about the average (a = s := \/ 171—1 201:,- — :E)2). One is not allowed to divide this value by 2 (=\/4), since the four methods use the same data and are therefore not independent. parameters are listed in table 5.3 and shown in figure 5.26. One can see that the deformations obtained using the various methods are grouped close together, even though their (10) error bars do not overlap. The value that was adopted by us is the mean of all four methods and is shown as a vertical bar. Previous measurements are shown as horizontal bars. One can see that the previous measurement for “At [47] agrees with the present result. ECIS fits have also been done, where only data points below 0cm 3 28° have been used with very similar results. The Becchetti-Greenlees calculations are also shown in figure 5.25 as dashed lines and they follow the data points well, especially for center of mass angles below the first minimum in the cross section. For larger angles, however, some deviation is apparent. With the help of equation 4.18 and the measured deformations, the Mn/Mp values were obtained and are listed together with other properties of the Argon isotopes in table 5.4. The interaction strengths of protons with protons (bp) and neutrons (bn) 117 labI [node last I‘ have inchn PI for 35. under Synnn not a] be no Derfln ln Slllglt A! II ”/1. rotational vibrational value adopted Isotope pos neg pos neg here 36AI 0.348(13) 0.406(15) 0.362(12) 0.318(10) 0.359(36) “Ar 0.290(13) 0.324(17) 0.308(14) 0.283(14) 0.301(18) “Ar 0.277(13) 0.315(8) 0.297(6) 0.263(6) 0.288(23) Table 5.3: Deformation parameters [32 obtained with the various ECIS fits for each model (rotational and vibrational) positive and negative figs have been used. The last column gives the adopted value. have been chosen to be 0.3 and 0.7 respectively [43, 52]. The quoted uncertainty includes a 1-0 error in bp/bn of 0.3. Figure 5.27 shows the obtained Mn/Mp values. One notices the very large value for 36Ar, indicating a large isovector contribution to the excitation. This is not understood, since the excitation of an N = Z nucleus should be isoscalar by isospin symmetry. Possibly, the prescription of Bernstein (equation 4.18) to obtain Mn/Mp is not applicable and, as in the 388 case (section 4.1.3), microscopic calculations might be necessary for a correct interpretation of the data. Such calculations are being performed now in collaboration with E. Bauge in France. In future experiments, measurements of the proton scattering cross section on the single-closed-shell nuclei 38Ar (N=20) and 4°Ar (N=28) would complete the set of Mn/Mp values for the neutron-rich argon isotopes. 118 FIgurt incon1 )leasu been p Figure 21150 in Neutron Number 1 8 20 22 24 26 28 2'0 I I I I I I 1.8 —- " Argon — 1.6 — _ Q 14 -— I _ z - I. 2: 1.2 — ] ] _ 2 1.0— .................................. .2 0.6 -— _ 0.4 I I L I I I 36 38 4O 42 44 46 Mass Number Figure 5.27: A comparison of the extracted Mn/Mp values. The value for 36Ar is incompatible with an isoscalar excitation that is expected for an N = Z nucleus. Measurements of 38Ar and 46Ar, which are to single-closed-shell nuclei, have not yet been performed. 2'0 I I I I I I 1.8 — I” o Argon m o Sulfur A 1.6 — _ s " -- a 1.4 — _ E 1.2— § 0 ] _ E 1.0 “" ---------------- - ------------- ... 0.8 L .3 1 0.6 l I I I I I 1 8 20 22 24 26 28 Neutron Number Figure 5.28: Same as figure 5.27, but here the data points for the sulfur isotopes are also included for comparison. 119 .mumumafieg cosmEuflmc @3855 23 so 828 9.8%: eon Ev FE monogflom 3.83 mmmmfi mu: .«0 33%: 8m + 3 326:8 855 =< .mmqouofl :owdfl mo $35963 me @9556 Jam. 23H. .9: I l 5%: 2392: 23” a @333 Es 3; 22:? 29.883 23:42 :39.” 26:5 :53: E: a: 28%? @383 page 8va A33 @384 E2 m3 $856.0 3 $83 :1 So $38.23 33% 6:3 @6383 E2. :s I I 8325 as $2 $85 a§§£.~ Es 2: 233?: 2836323 r: 8.0 8:82 336% $8.23 @253; Es as I I gang Gamma €98 @833 E: S2. .2212 2% am €533me E: 9625me 836$ 120 Chapter 6 Summary Two experiments have been performed to investigate the quadrupole collectivity of the neutron-rich argon and sulfur isotopes. In the first experiment the nuclei 38140328 and 44"“iAr were studied by Coulomb excitation on a 197Au target. A position-sensitive, high-efficiency NaI(Tl) array was constructed and was used to detect the de-excitation photons. The experiment al- lowed us to determine the energy of the first excited J’r = 2+ states in three nuclei (40’428, 44Ar), while a previously observed state in 4‘5.Ar was assigned definite spin. The excitation energy of the first excited state in 388 was reproduced. In addition, the 38.40.4123,“,46 Ar) were measured Coulomb excitation cross sections for all 5 isotopes ( and the reduced transition probabilities B(E2) were determined. The measurements revealed a new region of deformation around 40’428. The reduced transition probabil- ity for 4°Ar demonstrates that the N =28 shell closure persists at Z =18. The power of Coulomb excitation in the investigation of heavy sd-shell nuclei can be seen in figure 6.1. Shown there is the inverse quadrupole deformation parameter l/lfigl (the inverse is shown solely for illustrative purposes, since otherwise the large values in the foreground would block the view). The left-hand panel shows the available data before this Coulomb excitation experiment, while the right-hand panel shows all data 121 Figux R» U beflor sequc avail; UHSI quen‘ lI Huck in Ft (‘0er |I32I1 Q: 3’ i, C‘lCium 1441-30,, _ _~~ , ;[ -.-, \__/ Slug-0,, 20 ”we—~4— Inn” / 20 T-“““"‘-—-—~ 22 r“ xw— — , 22 ‘\““~ —_./ “4 n, - 2 “ w ~ -—4 2 ~~ n . Nguln’ll lgulnbif 28 ’ Neutn)n Igumbzef- 28 sues," Figure 6.1: Inverse quadrupole deformation parameter. (The inverse was taken solely for illustrative purposes.) The left-hand panel shows the data that was available before this thesis work. The result from this work (cross-hatched columns) and sub- sequent work (hatched columns) are shown on the right. available at the time of this writing. The cross hatched columns were measured in this Coulomb excitation experiment, and the hatched columns were measured subse- quently at the NSCL. In the second experiment the proton scattering cross sections on the unstable nuclei 42"‘4Ar were measured. The results are not completely understood and a group in France is working on microscopic nuclear matter calculations, which will help to correctly interpret the measured cross sections. 122 A.1 As SlI assun mt’ch: lS gin The I “here But} 1(3) Appendix A Coulomb Excitation A.1 Semi-Classical Approach and Perturbation Theory As shown in section 2.3 Coulomb excitation can be treated semi-classically. One assumes a classical Rutherford trajectory but treats the excitation process quantum- mechanically. The excitation cross section from an initial state ]i) to a final state If) is given by £13 (10 dfl — — PH). (A.1) d9) Ruth The Rutherford cross section is given by (3%)...=(%>28m”(3)’ with no given by equation A.15. Treating the electromagnetic interaction potential V(r(t)) as a time-dependent perturbation, P,.,, can be expressed to first order as 1 °° . PH, = Ia,_,,|2 with a,_,, = IE / e'“f“(f|V(r(t))|i) dt, (A.3) where (Ufi = (E f —E,-) /h. For sufficiently high energies or large impact parameters the Rutherford trajectory can be approximated as a straight line and the electromagnetic 123 pot Cha: wuh_ charg Ther and Thes hues “lth lOCatg the X Eben Xkfie potentials can be obtained through a Lorentz transformation of the field of a point charge at rest 4(17) = A(B)A'(1‘-"1(fi)flv')1 (Pt-4) with A’ (2’) being the 4-vector potential of a electric point charge at the origin A'(:I:’) = (A’, 26') = (0, 0, 0, z’ e—TZ) (A.5) 11(5) is the Lorentz transformation matrix that takes the stationary charge into a charge moving with velocity ,6 (see appendix C) 1 0 O 0 0 1 0 O 0 0 2'76 7 The resulting fields are the following Lienard-Wiechert expressions: (”1' = Zpe'y A.7 d) (,t) \/(b—:c)2+y2+’)r2(:1:—'ut)2 ( ) and v A(”)(r,t) = —C-¢(”)(r, t) . (A8) The superscript (p) indicates that this is the field generated by the projectile nucleus ’measured’ at position 1‘ and time t. The coordinate system that is used is oriented with the z-axis in the direction of motion of the projectile. The target nucleus is located at the origin and the projectile passes with impact parameter b (displaced in the x direction; i.e. 33,, = b, yp = 0, 2p 2 at ). The electric and magnetic fields are given by Ez—la—A- — V05 ° 6‘ . (A.9) B=V x A More explicitly: eZp'yb __ 7E0 (EJ. :)E:r : NICO (b2 + AWN”)g (1 + (:92) 124 Witl‘. follox with for e: for n for 1 and Th. E _ _ eZp'yvt __ 7E0 3 . 02+72v2t2fi (new (Bi =)B = fiEx=fi ‘32” = — ME“ ” (b2 + 72v2t2fi (1+(:)2)% 3,, = 0 with T 2 7i” and E0 = 93253. 7' defines the collision time. After expanding ¢ and A in multipole moments Winther and Alder [5] obtain the following excitation amplitudes Z 2 I M A— AM,- aH, = —z' hf; 26.1,.(5) (—)" K,(g(b))\/2/\TkA< f ”M“; “l > , (A.10) 7041 with (for p 2 0) 0°" (5) Z ”"4253” (Ii-1:35)? ((5) - 1)? ((A+1)(/\+#)P,i (c)_)1(/\—mu+1) u (c)) __ p _ 2A+1 H 2A+1 “1 v for electric excitations (7r 2: E) and GM“ (5) = wamlfiw (84:35) ((5) — 1)—%”P£(g) (All) for magnetic excitations (7r 2 M). Pf(:1:) are associated Legendre functions evaluated for a: > 1. For n < 0 the following relations can be used GEA—p = (—)”GE,\p and GMA—p = —(—)”GM,\” . (A.12) Even though [1 is listed as a summation index it is actually fixed by the nuclear matrix element to be and can be identified with the angular momentum transfer along the beam direction. The quantity 5 (b) is the adiabaticity parameter given by 7rao £(b) = ‘13—; (6+ 27) (A14) 125 where a0 is the half-distance of closest approach in head-on collisions, assuming the nuclei are point-like and if non-relativistic kinematics is used 2 7710112 00 (11.15) mo is the reduced mass of the two nuclei. The second term in the parentheses in equation (A.14) is a correction term that takes the recoil of the target nucleus into account. This expression is obtained by comparing the relativistic result with the low-energy Coulomb excitation result when large impact parameters are considered, therefore justifying the use of straight line trajectories [5]. A.1.1 Cross Sections The excitation cross section is obtained by integrating the product of the Rutherford cross section and the excitation probability. A sum over final and an average over initial magnetic substates must be performed, since they are not observed in the experiment. 00 1 i = 2 f d E i 2 min 2 .9# (€(bmin)) - a... (S) 2 he 62 «Au The function g“ is defined as 9u(§(b)) : 27’ (%)270Pdle#(€(/’))l2 b = QWflKu($)]2fEd$ E = «£2 [IK..+1<0I2—IK.(£)I?—2—,“KI+1(€IK..(€) . The K p are modified Bessel functions. The above expressions have been programmed using MATHEMATICA [53]. Given below is the MATHEMATICA code for the symbolic definitions (one can use them 126 for r 9qu A.1 The them The at [I SF‘SU and Iran. for symbolic calculations as well as for numerical ones) of the functions GEM, (5) and 9,,(5). The code is very compact, illustrating the power of MATHEMATICA. GEElam-.mu_,invbeta_]:= (-1)“mu * GE[1am,-mu,invbeta] /; (mu<0) GEElam-,mu_,invbeta_]:=( I‘(1am+mu) # Sqrt[16 PiJ/(lam*(2 1am +1)!!) * Sqrt[(1am-mu)!/(1am+mu)I] e SqrtEinvbeta‘2 - 1]‘-1 * ( (1am+1)(1am+mu)/(2 1am + 1 ) # LogondroPElam-l,mu,invbeta, LogondreTypo -> Complex 1 - 1am * (lam-mu+1)/(2 1am + 1 ) t LegendrePEIam+1,mu,invbeta, LegendreTypo->Complox ] ) I; ( (invbota > 1)&& (mu<=1am) it (mu>=0) ) ) gAlEmu-.xi-]:8 Hodulo[{mupos}, mupossAbs[mu]; Pi xi‘2 ( Abs[BesselK[mupoa+1,xi]]‘2 - AbsEBesselK[mupos,xi]]‘2 - 2 mupos/xi BessolKEmupos+1.xi] BesselKEmupos.xi]) J A.1.2 Relation Between Impact Parameter and Deflection Angle The impact parameter can be related to the scattering angle by calculating the mo- mentum transfer. In lowest order the momentum transfer to the target nucleus is t (00;) = f — (9%)rzo Zte dt. (A.16) —00 The subscript r = 0 means that we evaluate the field (generated by the projectile) at the original position of the target nucleus, which is the origin of the coordinate system. Evaluation of the integral results in __:Z£Zj£:p€2 ((1:0(m) = b’U ’ (A.17) where Zt,p are the proton numbers of target and projectile, b is the impact parameter, and v is the projectile velocity. Since the total momentum is conserved the momentum transfer to the projectile is the same (but opposite in sign) __.:2£Zt2:p62 by (A.18) pi”’(00) = —p§3)(00) 127 Hen whet moth of 1111 The I C46. The Mult The momentum in the z-direction is almost unchanged and given by the product of the relativistic factor 7, the (initial) velocity of the projectile v, and the projectile mass mp, pz 2 7mpv . (A.19) Hence, the deflection angle in the laboratory is given by x 2Z Z 2 z p where small angles have been assumed. Alternatively one can use the equation of motion in the center of mass system and calculate the scattering angle as a function of impact parameter assuming a relativistic Coulomb trajectory [54] mot}? ' 00 b = —cot(%06m) with an (A21) The relation between the laboratory and the center of mass angle is given by equation C.46. Both methods give identical results (within the approximations). A.2 Some Formulas In the following a list of useful equations is given. The relations will not be derived. A.2.1 Relations between 52, Q0, B(E2) . . . The nuclear radius for an axially symmetric nucleus is given by 12(0) = R0(1+ 52Y20(9, 4)) . (A22) Multipole moments are defined as QLM = i/p(r)rLYLM(6, 05) dr , (A.23) Q,,- = /p(r)(3:1:,-:I:j — 7260') dr ,and for L=0 and M22 (A24) 128 whr R0 i the The The Em “"11 all 167I' 3 Q20 = —5—4—7r_Z€ROfl2 ) (A.25) where 62 is the quadrupole deformation parameter, R0 is the nuclear radius (e.g. R0 = 1.2 fmA%), and Z is the nuclear proton number. In first order approximation, the deformation 52 is related to the B(E2) through 47r 1 = —- 2' + 2+ . .2 32 3(/B(E .0 —> )ZeRg (A 6) Electromagnetic transition probabilities are given by 87r(A +1) k2’\+1 . p E, ,- = B A; ,- k = — = — . W(7rA,J -—> Jf) A((2A+1)!l)2 h (7r J —) Jf) w1th it he (A27) The transition probability is related to the lifetime T and half-lifetime T% T w-1 Tl A 23 _ _ ln2 ° ( ' ) The B(7rA) values are defined as B(7I’A,0 —) A) = Z [(Jfo]M(7I’Au)]JiM,->]2 #11411 _ 2 — 2, + 1|-I Energy-weighted electromagnetic sum-rules from [5] 14.8yA—Z (32fm2 - MeV for A = ZB( 7r(A, i—>f) (Ef—E ,-) = 5.0 A(2A + 1) 21‘2”“2 e2fm2 - MeV for A 2 2, (A29) with R z 1.2A3fm. Single particle estimates (Weisskopf units) are defined as [9] Bu, (EA): -—1— E—H 2 —)2 R2’\e2 (A 30) A + 3 ’ ' and 2 Bw(MA) = E (L) 1229-9332,, . (A.31) 7r A+3 129 A.2.2 Constants he 2 197 MeV fm _ 1 ‘137 e2 = ahc = 1.44 MeV fm (I _ ehc _ 2M,,c2 MN 130 (A32) (A.33) (A34) (A35) Appendix B y-Ray Angular Distribution Following Relativistic Coulomb Excitation The excitation process does not populate the magnetic substates of the excited state evenly and hence the angular distribution of emitted photons will be anisotropic. In order to determine the detection efficiency the angular distribution of photons must be calculated and included in equation (2.33). In the following the 'y-ray angular distribution for intermediate energy Coulomb excitation is derived. 131 Per and by and “her with give: )Vhei Slatf Iff) : B.1 General Structure Perturbation theory is applied to the process All l1f> H7 IIfkaI III) Here HCE and H7 represent the interaction Hamiltonians for Coulomb excitation and radiative 'y-decay, respectively. The amplitudes for the two processes are given by at»; = (IfolHCElIiMi) (B-1) and ar—w = (IffofkalelIfo) , (132) where the final state consists not only of a nuclear state but also of a photon state with wave-vector k and polarization 0. The amplitude for the whole process is then given by “Hf! == “Hf—MU = 2 am; ar—w M! = ZaianHMflkalellfo) M! where the sum goes over all possible magnetic quantum numbers of the intermediate state.1 1One could view this process also as one where the state | f) = EM, a,-_,f|Ifo) decays to state lff) = [I ff M fka’), from which follows that ai—w = (ffIHvaI = Zai—U (IffofkalelIfo) (B3) MI 132 Fr) absolt ing or which be drt initial of the 0H); B.i B.2 1; I wh, ¢ tex From the transition amplitude one gets the transition probability by taking the absolute square, summing over final magnetic substates and polarizations and averag- ing over the initial magnetic substates. Since the final result is an angular distribution which will be normalized to unity at the end, constant factors in the equations will be dropped without special notice; e.g. the factor 5171:; from the averaging over the initial magnetic quantum number (M,) is not shown in B4. The angular dependence of the transition probability is due to the presence of the vector k in the amplitude a,_,ff. 0 and cp are the spherical coordinates of the vector k. WWW) = : [Oi—4M2 M,’,Mff,0 2 = Z :aHfofofkalelIfo) (B-4) M5,M”,a' M; = Z ”Midi—1;!(IffMflkaleIIfo) Mi’Mff'a M].M} >< (IffofkalelIfMj‘y B.2 Detailed Derivation B.2.1 What is (IffokalH,]Ifo)? H, is a scalar operator, a spherical tensor operator of rank zero. This is clear from rotational symmetry. Usually H, is expanded into a nuclear part and a (external) field part nuc ield H, = 203,, >056, ’ , (3.5) A.1: (nuc) A“ is a spherical tensor of rank A, which usually appears in where the operator O textbooks to derive transition probabilities. The difference is that O('"‘°) operates between two nuclear states and 0(fz'eld) operates between a one photon and a vacuum 133 state. The product of the two amplitudes can be related to the transition probabil- ity for one particular multipolarity A. This expansion of H, is not needed for this derivation. The only information needed is that H, is a scalar operator. In order to use this property H, needs to be sandwiched between angular momen- tum eigenstates. The nuclear part is already fine, but the photon part is given in plane wave eigenstates. Therefore we insert a complete set of photon angular momentum states: (IffofkalH7lIfol = Z(IffofkalIffofLMl(IffofLMlelIfol L,M : Z(kalLA/Il(IffA/IffLMlH'rlIfo) L,M = Z (kalLM)(IffofLMlL’M')(L’M’IH,|Ifo) L,M L’,M' 1 = ZIRUILMWHMHLMIIIMII L,M —’_——21, + 1° From the second to the third line the nuclear part in the first bracket was dropped. In the fourth line another complete set of states was inserted to express | ff) in terms of the total angular momentum. To get to the last line the Wigner-Eckart-theorem was used. (kalLM) is the overlap between a photon-state with wavevector k and polarization 0 with another photon state with sharp angular momentum L and projection M. We can rewrite |k0‘) in terms of [20) by using the rotation operator [k0) = B(z -> k)|z0). (B.6) Inserting a complete set of states and multiplying by (LMI from the left gives (LMlka) = Z (LM|R(z —> k)|L’M’)(L’M’|z0). (13.7) L’M' The rotation operator R(z —+ k) connects only states of the same angular momentum 134 and the 8111' where the 1 momentum the angular has to expaI Was d0ne e.1 Here 603 IS Combining Hence the and the sum reduces to a sum over M’ only. (LMlka) = :(LM]R(Z —> k)]LM')(LM'|z0) (B8) = §(D1T4M’(z ‘2 k))* k)[6EL + adML]. (B.11) Hence the following expression for the matrix-elements3 of H, are obtained (IffofkalelIfo> = 2V2L+1UIJMIILMIIIMf> L,M fo40(z —+ k) [6EL + 06ML], where the 6,1, are factors which are given explicitly in e.g. [56], but this information is not needed in this derivation. The 6,,L’s can be related to the mixing ratio for mixed multipole transitions. 2At this point we also have to introduce parity, which was not needed so far; in the previous formulas parity was implicitly contained in the summation over L since for gamma decay 7r and L are not independent as soon as the nuclear spins are given. 3See also [56] page 329 135 B.2.2 The Angular Distribution Putting the matrix elements of H, into the expression for the angular distribution B.4 one gets W(0,(p) = Z a,_,fa:_,f,\/(2L+1)(2L’+1) (B.12) ”li'Mff’a M,.M’f L,M.L',M’ X If! L I! If! I" 1! (_)(—L+Mf)+(—L’+M}) MflM—M, M,,M'—M; X D1160 DU); [dEL + UdML] [6EL’ + UdMUr, where the Clebsch-Gordan Coefficients have been replaced by expressions involving 3j-symbols (B.28). ngh D5,; can be evaluated as follows": L LI f _ M'—0’ L LI DMUDM’U — (—) DMaD—M’ = (—)M"" Z (2j+1)(,L4,’,,_IjJ) (Lj ”)19’23’ am’—0 = (-)M'“’Z(2J' + 1) (,1; #34,) (£31) Diho*(a,fl,7) , L ' L' L ' L' (_)M —UZ\/2j+1(M7:l-M’) (03_0_) }/jm‘(fiva) ],m Here, use has been made of the relations B23, B25, and B27. Using B.35 one can evaluate the sum over M ff which only occurs in the first two 3j-symbols in B.12 23(11): L’ 1; )(Iff L If) I LI L k : _ 2L+Iff+M —Mf _ k 2k () .23 )( +1)(M,_Mn) x I, I, k L’Lk M}—MfK. IfIfIff , °if no arguments are given we always assume the same Euler angles 136 with the fol II'(6. I Here. 6) staI Slimming 0“ TO gm this Stile-Chou 7‘11. I Th‘i‘refore with the following result: W(0,tp) = Z aiafa:_,f:\/(2L+1)(2L'+l) (2k+1) Mi,a,k,n I Mf'Mf L.1W,L’,M’ X L L’ k I, I, k LL’k M-M'K Mf—M}I€ IfIfJ x(_)(M;-L)+(M}—L’)+(2L+M'—M,)+k ,_ . L j L’ L j L, _ 1W (7 2 1 at ’ . XI) 2; 3+ )(Mm_M,) (00%) xmahah Here, 5A stands for [5B, + adMA]. The sum over M and M’ can now be performed: 2, L L’ k Lj L’ ZHM (M—M’n) (Mm—M’) M,M’ _ Z L L’ k Lj L’ ’ M—M’Is: Mm—M’ M,M’ L L, k L L, j (_)L+L’+j , M —M’n M —M’m _ L+L’+k = —2k_+—1—6kj6m" - II :M A v Summing over 0 yields: Lj L’ 2 (f3 5;) (”fly I jeven 2 6L 6L. = a={1.—1) 0 : jodd To get this result the symmetry properties of the 3j—symbol (B30) and the parity selection rules for EA and MA transitions were used.5 (—)L : for EA transitions ”1”? = (3.13) (—)L+1 : for MA transitions Therefore W(0, e) = )3 aisfaggm/(zL +1)(2L’+1)(2k +1) $23512 L = even 6 = 6 5e.g.: for mm = + and => L EL L 2 Odd 6L 2 ngL X I! If k LkL’ LL’k Mf—M}I€ 10—1 IfIfIff x (—)(Ml‘L)+(M}"L')+(2L-M;)+(L+L’+k)+lc X thc 6L 6U Z aiafa:_,f;\/(2L + 1)(2L’+1)(2k +1) I: even,n.M,’ MI,M}.L.L’ X I, I, k LkL’ LL’k Mf—M}I€ 10-1 IfIfIff X (-)M} Y1; 5L 5U 2 a. a, I, I, k H’ Hf’ Mf—M}rs: k even,x.M'-, MI,M}.L,L’ X (—)M}Fk(L, L’, I”, If) Y’s!“ 6;, 6L’ If particles are detected symmetrically around the z-axis one can integrate over ‘Pparticle which is equivalent to integrating, or averaging, over (p7. Using B.36 yields the final result: we) 2, a,- a? ' (—) ’ keven,M'-, —)f _)f (Mf —M}0 M,,M’,,L,L’ X Fk(L, LI, Iff,1f) V2k +1 Pk(COS(0)) 6L (5y :1 If If k M] Z alfifa’i—{f (M! —Mf O) ( ) k even,M,-. M,,L,L’ X Fk(L, LI, Iff,1f) V2k +1 Pk(COS(0)) 6L 6U I I k . 2 f f _ M, k even,Mi. Mf,L,L’ X Fk(L,L’, Iff,If) V216 +1 Pk(COS(9)) 6L 6y I I k kegml fl M,—M,0 () ( ) Mf,L,L’ X Fk(L, L], I”, If) V2k +1 Pk(COS(0)) (51461} 138 B.3 Angular Distribution Using the Winther and Alder Excitation Amplitudes Putting the excitation amplitudes from [5] into the expression for the angular distri- bution B. 14 yields6 W) = Z: k even,M‘-. &![.L,L’ x (If If k) (—)Mka(L, L',1,,,1,) \/2k+1Pk(cos(0)) 5,, 5”. The absolute square of the excitation amplitude can be written as follows u If /\ Ii 2 - K 6“.) (_)MI (—Mf —/1Mi) — C [1 If A Ii — EGAM(;)KII(€)(—) (—Mf "H Mi) C * 1.. p' If A Ii X GAp'(;) KM“) (_) (_Mf —#’Mi) I A I- I A 1' _ 2 f t f ‘ _ SIGMA: )QLIIK (5 )l (—M;-/1Mi) (-M, —#M.') The last step could be performed since u and ,u’ are fixed by the 3j-symbols. Hence c u M! [f A I,- 2 gGAu(;)Ku(€)(—) (—) (—Mf "#Mi) I, A 1,)(1, A 1,) W0 = G K () 2M 2| M— 5m (52» (am—m, -M,—uM.- MfHLL’ If If k _M : /—_ X ( ) fP‘k(L,L,Iff,1f) 2k+1Pk(COS(9)) 61,61}. M,-M,0 Using 835 one can evaluate the sum over M, which only occurs in the first two Z< I, A 1,)(1, A 1,) Mi —Mf—/.I.Mi "Mf—pMi _ 2I,+I,—M,—p _ k’ r If If I") (> a ) (2k+1)(_Mfo,c k’ ,n x A A k’ IfIfk’ p-pn: AA],- 6See [5] or appendix A for a definition of G “(5), K “(E) and £. 33' —symbols 139 W(9) = Z IGAL(; 62)| IKIIEU -)_Mf "XX- (2k'+1) k evenm A1f,L.L’ x If If k' A A k' IfIfk’ If If k —MfoIt p—pn A AI,- Mf—Mfo X (—)Mka(L, LI, Iff,1f) V2119 + 1 Pk(COS(0)) 6L 61,! Summing over the M f in the first and the last 3j-symbol (B.29) gives a 6H6“. Therefore c A A k W 9 = G — 2K 2 — “ - III :I AIIUIII II(€)|( ) (M10) L.L’ I I k X{111.}Fk(L,L,,Iff,If)\/2k+1Pk(COS(0))6L6LI . Here, we also indicated the impact parameter b in the formula, which was im- plicitly assumed before. The remaining step is to integrate this formula over impact parameters with a minimum of bmin, which is e.g. determined by 0mm from the ex- periment W(9) = / mm) b db. (3.15) bmin As in appendix A the integral over the square of the modified Bessel function / IKIIIoIZ’bdb (8.16) bmin can be expressed using the Winther and Alder function 9,,(5) 9,460)» = 27%} ) [pdle(€p())l = 27r/lKu(a:)|2:rd:c = «£2 [IKu+1(€)l2 — lKu(€)l- 2§KI+II mm 140 Thus W) = *2: IGAII(-:-)l2gu(€)(-)“(:_Au3) (8.17) L,LI X {111/{f } Fk(L, L’,Iff,If) V2k +1Pk(COS(6))6L 6L! . (13.18) Usually the angular distribution is expressed as we) = Z akPk(cos(0)) (13.19) I: even The following the MATHEMATICA code calculates the coefficients ak in front of the Legendre polynomials: 141 (s i gets excited to f and decays to ff StatesI 8 {{Ji,Pi},{Jf,Pf}.{Jff,Pff},delta} I9) SRules I { JiS[x-] > x[[1.1]]. (* initial spin *) PiSfx-] -> X[[1.2]]. (’ initial parity 1") JfS[x_] :> x[[2.1]], (* excited spin *) PfS[x_] '> x[[2,2]], (# excited parity t) JffS[x-] > x[[3,1]], (t final spin *) PffSEx-) :> x[[3,2]], (* final parity *) DeltaMixSEx-] :> x[[4]] (* mixing for simultaneous E2 and M1 decay *) } AngDis[K_,P_,T_,S_]:= Module[{LamHin,LamMax.ParityChange,kHax,tab.PureElamTransition}, CachinEK,P,T]; (* first do the kinematic calculation *) (a check for the involved transitions *) Lamflin 8 AbsEJfSESJ-JffSES]]S LamMin = RationalizeELamMin]; (t make sure we have an integer t) If[(Lamflin===0)&l(JfS[S]+JffS[S]>0),LamMin=1,,Print["siw"] ]; (t there is no H0 or 80 transition t) (e find out what multipolarities for deexcitation are possible t) If[PfS[S]=!=PffS[S],ParityChange=True,ParityChangeaFalse,Print["siw.ParityChange"]]; If[!IntegerQ[LamMin],Print["LamHin is no integer; LamHin 8 ",LamHin];Return[]]; (* we need to consider Elam and H(lam-1) transitions if the following is FALSE t) PureElamTransition 8 (EParityChange it EvenOELamMinJ) ll (ParityChange It OddQ[LamHin]); If[ PureElamTransition, LamMax=LamMin. (* Pure ELamMin transition t) LamHax=HinELamHin+1.JfSES]+JffS[S]], (t for Hlam transition consider next higher E transition if at all possilble 0) Print["siw AngularDistrib"]]; Print["LamMax ",LamMax]; Print["LamMin ",LamMin]; (* now we can determine the coeff. in front of the P_lm *) (* first we determine kMax #) kHax=HinE2JfSESJ,2LamMax]; tabSTable[Sum[ (* Print["11 ",11]; Print["12 ",12]; Print["mu ",mu]; f) If[11===LamHin,1,DeltaMixSES].Print["siw AngularDistrib 2"]] # If[12===LamHin.1,DeltaMixSES].Print[“siw AngularDistrib 3"]] t AbsEGEELamNEP].mu,1/BetaK[K]]]‘2 gAlEmu.XiHoK[K]] (-1)‘mu # ThreeJSymbol[{LamN[P].mu}.{LamN[P],-mu},{k,0}] # SixJSymbol[{JfS[S],JfSES].k}.{LamN[P].LamNEP].JiS[S]}] F[k,JffS[S].11,l2,JfS[S]] Sqrtf2k+1]. {11,LamHin,LamMax},{12,LamMin,LamMax}. (t sum over the possible deexcitation multipolarities *) {mu,-LamN[P],LamN[P]} (* sum over magnetic ON of excitation *) ] (* close Sum *) ,{k,0,kMax.2}]; (* close Table-loop *) Do[ Print["a",k," 8 ", N[ tab[[k/2+1]] / tab[[1]] ] J .{k,0,kMax,2}]; tab/tab[[1]] // Together ]//.SRules//.NKRules (* set the necessary attributes for AngDis *) SetAttributes[AngDis,HoldFirst] H[theta_,dis-List,K_:1]:= Module[{Hcm,Hcontracted,Uboost}, Hcm[t_]:= 1/(4Pi) PlusOO (Table[LegendreP[2k,Cos[t1],{k,0,Length[dis]-1}]'Times’dis); If[!List0[K], ReturnEHcmEthetaJJ J; Hcontracted[t_]:= HcmEt] t DomCmDomLab[t,BetaX[K]] //.Kfiu1es; Hboost[t_]:= Hcontracted[ThetaGm[t,BetaK[K]] J //.KRu1es; Hboost[theta] 142 BA Angular Momentum Re-coupling Coefficients, Rotation Matrices and Related Formulas The phase convention used for the Clebsch-Gordan coefficients is that of Condon and Shortly [57]. B.4.1 Angular Momentum Representation of the Rotation Matrix Definition DLmAmflfl) == (J'm'lR‘1(a,fi,7)ljm)= (J'mllem'> = >e""‘” (B36) Mm(67 ¢) : \J B.4.5 7 -— 7 Correlation Function Definition Fk(L,L’,Il,12) = (—)’1+’2‘1\/(2k+1)(212+1)(2L+1)(2L’+1) Lflk LUk X (1—10) {1212 11} (3'37) 145 Appendix C Relativistic Kinematics This chapter develops some formulas that are needed for computer calculations, which are used in chapter 5 to obtain the excitation energy and center of mass scattering angle from the measured particle energy and laboratory scattering angle. We will not have a single formula as a final result, but rather instructions on how to combine certain formulas to obtain what is desired. The formulas are derived and in the last section a summary is presented. C.1 Notation and Preliminaries We will use the Minkowski notation with x4 = iso and c = 1. To obtain the equations with the correct power of c at the right place, just follow these replacement rules: p——>cp , m—>mc2 ,and flag (C.l) Repeated indices are summed implicitly. If one knows a certain 4-vector in one system, one applies a Lorentz transformation to obtain this quantity in a different system. For practical purposes consider only transformations in the z-direction since one can obtain others by simply combining 146 Lorentz transformations and rotations. Assuming: (3 = ( 2.12, ) (02) then 12’ is obtained through p; = Aijpj , (summing over j is implicit) (C3) with A given by: 1 0 0 0 0 1 0 O A— 0 O 7 473 (0.4) 0 0 ivfl ”Y with fl = g and 1 7 = —— (05) (/(l -fi2) ' Consider the 4-momentum of a particle initially at rest p:(i(7)n):(z'1§ ). (0.6) The momentum after a boost in z—direction with velocity fl is given by 0 , 0 Wm iym p’ can also be written as1 0 , 0 p = I (0.8) p iE' Comparison of the two results yields the following very useful formulas: E = 7m and p = yflm = ,BE. (C.9) 1We will use p’ to label both the 4—vector and the magnitude of the total momentum. It should be clear what is meant from the context. 147 Taking the scalar product of p with itself yields p - p 2 —m2. Since 19 - p is a Lorentz scalar it has to be equal to p’ - p’ = p2 — E2. One obtains the well known result E2 = p2 + m2 (0.10) which can alternatively be obtained through the following E2 : 72777.2 2 m2 + (72 _1)m2 : m2 +72fi2m2 = m2 +p . The first line is an identity and going to the second line the definition of 7 (equation C.5) has been used. We see that A Operates on the particle states and not on the frame of reference. A boost by 3 is equivalent to choosing a new frame which moves with velocity —fi. C.2 Collision of Two Particles Consider a collision B (A, A’ )B’ , which means that B is the stationary target and A is the projectile. After the reaction we have the ejectile A’ and the recoil B’ . Figure Cl and table C.1 show the quantities involved: E denotes the total energy, m the rest mass, T is the kinetic energy defined as E — m, and p is the momentum. The subscripts cm and )0), refer to the center of mass and laboratory coordinate systems, respectively. To obtain an expression for the total energy in the center of mass system, consider the contraction of the total 4-momentum. This is a Lorentz scalar, meaning that this quantity does not change when going from one frame to another (pcm 19m 2 pic), 1010),) pcm ' pcm : _Egm _ _ 2 2 — plab ' plab “‘ plab _ Elab 148 Laboratory System Center of Mass System B(A,A')B' , A 3 fGIab A > . A 9cm 3 2/ <9. Figure C.1: Schematic drawing of a collision of two particles in the laboratory system and center of mass system. The arrows symbolize the momenta of the particles. . Coordinate system Quantlty laboratory center of mass 0 0 0 0 PA pAlab pAcm ZEAM, iEAcm 0 O 0 0 0 0 p3 = 0 chm —pAcm 2777.3 21?ch ZEch PA; , s1 (92); ) m; sm(0Ar ) PA! pAiab = 0 PA;,,, = O iEAiab p4“) cos(6A;ab) z'EAém PAgm cos(0A;m) iEAiab iEAgm pBIab Sin(gBIab) Pegab _ 0 —PA;,,, [’3' . — . ”38;... PHI... COSWBI...) ZEBém z.E‘Bfab 0 0 , 0 0 0 P = P = . pAlab pAlab ZEcm “EA,” + m3) z.(E‘lab) Table C.1: Momenta of particles before and after scattering in the center of mass frame and the laboratory frame (see also figure CI). p is the total momentum of particles A and B combined. 149 = pl2ab - (EX... + 21321:.me + mfg) = _(mil + ZEAlame + m2B) a where in the last step E2 = p2 + m2 was used. Therefore E02," 2 2&4“me + mg, + "123 . (C.11) In order to obtain an expression for the energies of the individual particles in the center of mass frame, we use the fact that in the center of mass frame the linear momenta of the two particles add up to zero (PAC... + chm = 0), therefore: 2 _ 2 pAcm _ chm 2 2 _ 2 _ 2 Ecm_mA — Ecm 7723. Using ECm = E Am + E30," to replace e.g. Eficm yields E3, = 213ch EC", — mfg + mi, (0.12) 01‘ Efm + m2 — m2 EB“, _—. 2E3 A . ((3.13) From E ch one can obtain flgm through: ’73 = Ech cm m8 _ —2 fich _ 1 — 78cm Since particle B is at rest in the laboratory system, the velocity of B in the center of mass )6ch is equal to the velocity of the center of mass in the laboratory 5ch Ban = fig... . (0.14) Other ways to obtain ficm are a E ficm : Pl b _____ pAlab : 5AM. Alab (0.15) Elab EA“), + m3 Ema), + m3 150 01‘ c _ _ _ __ _, ac 0.16 7 Ecm ( ) C.2.1 Relation Between 60m and 61a), From table C.1 pAga, sin(9A;,,) PAL... Sin(9A2m) 0 0 pA/ : and pAém : (018) ’a" 19%» cos(6A; ) pAIm cos(0A: ) 213% z'EAr p Aiab is also given by PAL”, 2 Acm—>lab PAgm 1 0 0 0 . PAL... Shim/1'...) _ 0 1 0 0 0 — 0 0 7 475 pAngOS(0A;m) 0 0 2'75 7 iEA’cm psin(6) _ 0 _ 7(pCOS(9)+flE) i7(fiPCOS(9)+E) A, lab Three equations result: PAgabSinWAgabl = PAgm Sin(9A;m) pAlab cos(0A;ab) = 7(pArcm cos(0A;m) + IBEA'cm) EAL“, = 7(52922," cos(0,,,m) + EAém) Dividing the first by the second equation yields sin(0,4rcm) [3E I 7(cos(6Arcm) + p—j‘m) sin(9A2m) : . (C.19) 7 (COS(9A;m) + [321' ) 151 tan(6 A105) If [3 > 3.4;," then there is a maximum scattering angle, which can be found by setting the derivative 00,120“ 80,427” equal to zero. The result is fiA’ t max 2 cm . 0.20 an( Alab) WWW) ( ) Unfortunately there is no closed expression for 0A2...) because there is no closed expression for E A?“ as a function of 0A2“. To solve numerically one uses the previous formula iteratively until agreement is reached. But the formulas are useful for the Special case where one particle is a photon. sin (0,211”) 7 (cos(6A;ab) — 3232—) lab tan(0Arcm) EAém : f)l(—.'BIDA;ab COS(0AIab) + EAIab) . Photons For photons the previous formulas can be simplified, since the photon (rest) mass is zero and it follows that E21) and 3:1. (C.21) Therefore: sin(l9)ab) t 60m ...( ) 7(008(01ab)-fl) tan(91ab) : sm(0m) 7(cos(06m) + 5) Ecm : 7Elab(1 — fiCOS(glab)) Elab = 7Ecm(1 + IBCOS(ocm)) C.2.2 Solid Angle Relation To find the relation between the solid angle subtended in the laboratory and the center of mass system use data), _ d(COS(0[ab)) do... ‘ d(cos(9m)) ' (0'22) 152 Figure C2: Relation between 2:,y, and 0. From figure C2 it follows that tan(6) = g (0.23) and comparison to equation (3.19 shows that :r and y are given as x = 7(COS(0A'°m)+fl,Zm) y = sin(0A:cm). In order to find cos(Hlab) we use 513 0 = —— . .24 cos( ) mi: +31 (C ) Therefore '7 (cos(QAICm) + flj ) COS(6(ab) = cm (0.25) 2 . \/Sln2(9Alcm) + ’72 (COS(0Alcm) + 32%;) With the abbreviations of (sin(0A'cm) —> s), (cos(OAtcm) —> c), and (fi—f— —> fl) it follows that ———:E:::Ezlab;; = 7(72(c + (1)2 + 82W - V(C + fl)(72(c + H)? + 32)“%(72(c + 3) _ c) 7(72(c + m2 + s?) — 70720: + I?)2 — c2 — cfl) (72(6 + fl)2 + 82)% 7(1 + Cfl) . (72(6 + fl)2 + 82)% After reversing the substitutions szab _ 7 (1 + cos(OAlcme, ) ((3.26) Photons For photons the last formula simplifies to dumb __ 7(1— CO5(9cm )5 ) (mam (720mm MW) +sin 5A.". 2‘33 PAM, = plab C43? 16cm = 2:: (C45) The first possibility is the easiest, but one might choose a different one depending on what other quantities are needed later on. C.4.5 Relation Between 60m and 610), tan ( 9A1“) = Sln(0Agm) , (C. 46) '7 (cos(OArcm) + L) 3A2... where 3 is the velocity of the center of mass system in the laboratory system. Photons Sln(01ab) tanwan) 7(C03(01ab) - fl) Sln(0cm) tan(9zab) 7(COS(0cm) + )8) Ecm = 7Elab(1 _ flCOS(61ab)) Elab : 7Ecm(l + BCOS(gcm)) C.4.6 Solid Angle Relation _ __5_ M- 2(1 COW/2%) dQcm — 2 . _ (72 (COW/“rm + mi...) + SmwA’mV): ((3.47) 156 Photons thab 1 dQcm _ 72(1+flc08(0¢m))2 Ecm 2 _ (Elab) 2 72(1— ,BCOS(0[ab))2 C.4.7 Invariant Mass 2 2 "'32 = (21333)) ‘ (Z 23:22,) - (C48) The excitation energy of particle A is obtained as E“ - m — (0) C A — A mA a ( '49) and the breakup energy is Eff = mA — 2mg,“ (0.50) 157 Appendix D Simulation of the Proton Scattering Experiment A simulation program was written in FORTRAN that simulates the full detector response, the geometry of the setup, the beam emittance, as well as energy loss and angular straggling of the recoil protons in the target. The code uses the N SCL histogrammer SMAUG for the generation of the spectra and since it is also linked to the standard analysis routines, actual data can be used as a starting point for the simulation. In this case actual tracking data are used in order to study the effect of the beam emittance on the resolution. D.1 Structure of Computer Code 0 Event Generation — scatter two particles in center of mass system; option to use cross section data or flat cross section — transformation to the lab system — depending on settings rotate coordinate system and/or move origin to match the incoming particle; the incoming particle is defined by actual measured PPAC data 158 — calculate and subtract the energy loss [48] in the target from the proton en- ergy; the scattering occurs at a random depth in the target; add straggling [58] to the proton direction; see figure 5.15; 0 Measurement — check which detectors are hit (that includes the particle and proton detec- tors) — add detector response (energy resolution, position resolution...) to “mea- sured” signals 0 Analysis — same (or similar) as analysis of real data 159 D.2 Description of Input File The following is a sample input file. It was used to create the spectra in figure 5.16. m_a : 33533.7840 m_b : 938.2723 m_apO : 33533.7840 m-bp0 : 938.2723 beam_energy : 1211.0400 de : 0.0200 use_cs : F ex_en_ap : 0.0000 target_thickness : 2.7000 axis_rotate : T target_intercept_ca1c : T do_angl_strag1 : T strag1_max_iter : 10 e_noise : 0.1000 pos_res_10mev : 1.0000 strip_x_randomize : T ppac1_resolution : 8.0000 ppac2_resolution : 3.0000 tof_distance : 40.0000 tof_resolution : 0.0000 hit_check_zero : T hit_check_te1 : T zero_radius : 38.1000 t_wethinkitis : 1211.0400 tracking_pos ° T tracking_angle T corr_for_target_th T tof_correction : F file_cs_gs_data : SYS$GAMMA3z[96035.SIMUL.36AR]cs_gs.data file_cs_ex_data : SYS$GAMMA3z[96035.SIMUL.36AR]cs_ex.data tot_cs_max : 300.000 All parameters are described in table D.1. If use_cs=T then two states are in- cluded, one with excitation energy (=Q-value) of zero and the other with excitation energy ex_en_ap. The cross section data files file_cs_gs_data and file_cd_ex_data are used for both states, respectively. If use_cs=F then only one state is used with a flat cross section and excitation energy of ex_en_ap. 160 Item Description m_a projectile mass (MeV) m_b target mass (MeV) m_apO ejectile mass (MeV) m_pr recoil mass (MeV) beam_energy beam energy (MeV) de beam energy spread use_cs if T then use cross section given in the files f ile_cs_gs_data for the elastic scattering and file_cs_ex_data for the ex- cited state ex_en_ap excitation energy of excited state; see text target_thickness axis_rotate target_intercept_calc do_angl_stragl stragl_max_iter e_noise pos_res_10mev strip_x_randomize ppac1_resolution ppac2_resolution tof_distance tof_resolution hit_check-zero hit_check_tel zero_radius t_wethinkitis tracking_pos tracking_ang1e corr_for_target_th tof_correction file-cs_gs_data file_cs_ex_data tot_cs_max target thickness (mg/cmz) if F then use z-axis as incoming direction; if T use direction supplied by PPAC data if F projectile hits target in the center (a: = 0, y = 0), if T see axis-rotate if T include angular straggling in target take a maximum of stragl_max_iter steps for the proton to travel out of the target noise in energy detectors (Silicon) (MeV) (FWHM) position resolution of strip detectors (mm) (FWHM) if T randomize position perpendicular to strip resolution of first PPAC (mm) (FWHM) resolution of second PPAC (mm) (FWHM) distance between time of flight detectors (m) resolution of time of flight detectors (ns) if T check if ejectile hits zero degree detector if T check if proton hit telescopes radius of zero degree detector (mm) beam energy used for analysis; can be different from beam_energy in order to include systematic errors if T use tracking to calculate position on target if T use tracking to calculate incoming beam direction; if F use z-axis if T correct measured proton energy for half of the target thickness if T use time of flight information to correct for beam energy spread; not implemented yet path to file containing the elastic cross section; use if use_cs kiT same as previous for inelastic scattering maximum possible cross section for correct calculation; if the cross section (from the previous two entries is larger than tot_cs_max then tot_cs_max will be used Table D1: Explanation of input file for simulation. 161 Appendix E Differential Cross Section The number of reactions N, observed in a detector is proportional to the number of incoming beam particles Nb and the number of target nuclei per unit area M. For a particle detector that has a perfect intrinsic detection efficiency (like the silicon detectors used in the proton scattering experiment, but unlike the photon detectors used in the Coulomb excitation experiment) the factor of proportionality a := N—C’fi; is given by pure geometrical considerations as d . 0 = /—31n(6)d¢d0, (13.1) where 3% is the differential cross section of the reaction to be observed. The integral goes over 0, which corresponds to the solid angle covered by the detector. Assuming a constant differential cross section within 9 d0 35 = a (/ sin(6) d¢ do) . (B.2) Q The integral over (25 can be replaced in the following way fsin(6) dab d9 = fsin(0) 2[1(8) qu d0 {2 o 0 = 27r [sin(0)e(0) . 0 162 6(0) is the detection efficiency as a function of angle 0, defined as the fraction of particles (all with scattering angle 0) that are actually detected. This efficiency can be obtained from geometrical considerations of the angular coverage or from a Monte Carlo simulation of the experimental setup. The differential cross section is obtained from the number or observed reactions N, in the following way: For a sufficiently small 0 integration range one can assume 6(0) (and also sin(0)) to be constant and obtains do _ a 0R- _ 27r 6(0) (cos(0zow) — cos(0high)) _ 1 1 ‘ " 27r 6(9) sin(0) A0 _ N, 1 1 — Nth 27r 6(0) sin(0) A0 ’ with N, : Number of reactions observed Nb : Total number of beam particles N, : Total number of target particles (1H per cm?) (B.3) 6(0) : efficiency for particle detection at angle 0 010w, 0mg,“ 0 : lower, higher limit and centroid of 0 bin In the proton scattering experiment N, is given by: Nt = NIH ‘1}; dt , (13.4) where N13 is the number of 1H atoms per molecule, (1, is the target thickness in units of mass per area, N A is the Avogadro number, and M is the molar mass (mass of N A molecules). For the polypropylene target (CH2) used in the (p, 19’) experiments NIH-:2 NA = 6.022 - 1023 mol—1 M =14i mol NA 23 1 1 N — = 2. .22-10 — “M 60 111011442— mol 163 = 8.603 - 1022 g-1 Therefore fl_ 1.850~106 .IYL 1 1 d9 — dt/(mgcm‘2) Nb 6(0) sin(0) A0 mb . (E.5) Equation 13.5 gives the differential cross section in 1:1} given the target thickness in mg cm”. 1 mb = 10‘27cm2 was used. 164 Appendix F A New Method for Particle Tracking F. 1 Traditional Tracking From the geometry of the setup (see figure F.1) it follows that the position on the target 3:, (the target is located at z, = 0) is given by xt = M, (F.1) 22-21 and the slope 9% is given by d1} _ $2 — 5131 (F2) d—Z — 22 — 21 , where 151(2) is the (measured) x-coordinate of the particle at z 2 21(2), which is the z-coordinate of detector 1(2). The error of the target position 0(a) is 02(xt) = ( Z2 0($1))2+( Z1 00102))2 , (F.3) 22—21 22—21 where 0(x1(2)) is the uncertainty in the measurement of the position 331(2). The error in the SIOpe of the particle trajectory (IQ?) is a2 (2%) = ( 1 )2(02(231) +0202». (F.4) 21—22 165 x Tracking 1 Tracking 2 Target ll 1 l . xl . .__>§_ Particle Uajectogy X2 1 21 x! i m dz I I O beam axis 1 i Z1 22 21:0 2 Figure F.1: Setup for particle tracking. Two tracking detectors determine the position of the beam particle 31,2 at two points along the beam line, from which the position on the target and the slope of the trajectory can be determined. The coordinates of the tracking detectors are 21 (detector 1) and z2 (detector 2). The target is located at zt == 0 F.2 New Method In the following a new algorithm for the determination of the target intercept and the slope of the trajectory is presented. The new method is superior to the traditional method, especially if the resolution of the tracking detectors is poor. F.2.1 Some Beam Physics The slope of a particle trajectory is not independent of the (transverse) position of the particle. This is illustrated in figure F.2. Shown is the phase-space—ellipse (PSE), i.e. the slope of the beam trajectory is plotted as a function of transverse position. In an ideal case the width (= 2 - W) is zero and the slope of the trajectory is a linear function of the particle position. Hence we can infer from one position measurement alone the trajectory of the particle. A prerequisite is, of course, a knowledge of the orientation of the PSE (i.e. the slope S). The width of the PSE introduces a large uncertainty, however. Assuming that the main beam direction is along the z-axis of the coordinate system, it follows that the slope of the trajectory if as a function of position a: is 166 X Figure F.2: Phase-space—ellipse (PSE) of the beam can be approximated by a straight line. The error that is introduced is W, the half-width of the PSE. The slope S is defined as S = £1393. 167 obtained through g:- = 3:1: . (F.5) It is important to not confuse the slope of the trajectory j—: with the slope of the PSE S. The latter changes as one goes along the beam-direction (S is negative before the focus, infinity at the focus, and positive after the focus of the beam). It is the same for all particles and a prOperty of the beam as a whole. In contrast fix; is different for each particle. For the same particle, however, it is independent of the position along the beam-line z. The position on the target is given by da: da: $t=$1+(Zt—Zl)d—z=$1—21'&;. (F.6) The uncertainty of the particle slope is 2 d5” 2 2 2 0' (IE) 2 S 0 (2:1) +W , (F.7) where the first term is due to the resolution of the detector and the second term is the half-width of the PSE (see figure F.2). The two contributions are independent and are therefore added in quadrature. The uncertainty in the position on the target is 02(1),) = 02(1‘1) + sz202(a:1)+ sz2 = (1 + 2? S2)2a2(a:1) + 2% W2. If the beam is focused at the target position the term 1 + z? 52 = O and the position resolution is equal to 2% W2, which yields essentially the beam spot size. The approx- imation of the PSE as a straight line results in a constant calculated position at the focus, and hence, the position resolution is equal to the half-width of the beam-spot. If the resolution of the tracking detectors results in a resolution at the target (using standard tracking), that is larger than the beam-spot size, then it is of course better 168 to assume a constant position on the target. But what happens if the beam-spot is on the order of the resolution? Then one should combine the three different methods (the usual one and the new method applied to both tracking detectors) to determine the target point 12¢ and the slope 4‘5. The new method relies on the independence dz of all three procedures, which is only almost true. For instance the error due to the width of the PSE is the same for both tracking detectors and should therefore be added linearly. Another advantage of the new method is that it can be applied, if the beam is not focused at the target, but instead slightly before or behind. The assumption of a constant target point would result in a much worse resolution. F.2.2 Determination of Position and Slope Assuming independence of the three measurements one finds the target position as the weighted average: 1,: (2%)“12 ”39:0 , (F.8) i 02W: and the slope of the trajectory as ‘1 dz“) _(277) 2—2) (F...) with uncertainties given by 1 —1 02(xt) 2 (E; m) (F.10) and 1 0264:): Z—(‘i-x—(T) . (Ru) 2' 0'2 The slope S and the half-width W of the PSE in terms of the beam emittance e, the beam-spot size at the focus D, the position of the focal point f at any position 2 169 l new method '— 0 1 l L l O 2 4 6 8 10 60(2) (mm) Figure F .3: Comparison of old and new method. Shown is the resolution of the target position o(:z:t) as a function of the resolution of the second tracking detector 0(x2). The positions of the tracking detector with respect to the target 21,2 are roughly 1 m and 2 m. The beam-spot size D is 10 mm and the emittance e is 0.1 mm. One can see that if the resolution of the second detector is larger than 2 mm the new method is better, and even for a resolution of 1 mm the weighted average of both methods results in an improvement. along the beam-line are [59]: (F.12) W = ((2%): (gig—”)2)—% . (F.13) Figure F .3 shows an example. The new resolution is considerably better as soon and as the resolution of the tracking detector is larger than 2 mm. In experiments where tracking is crucial not only the target position but also the slope of the trajectory is of importance and the effect of the different tracking methods can not easily be estimated. 170 F.2.3 Practical Method For practical purposes it is hard to determine the lepe of the PSE S for a given detector position and beam profile. A better procedure is to consider it a param- eter and optimize it until the best resolution is reached. This procedure has been applied here, and figure F.4 shows the effect on the energy resolution for the re- action 36Ar(p,p’)36Ar*. The resolutions of the tracking detectors are 8 mm and 3 mm (FWHM) and the new method results in a better resolution by almost a factor of 2. The position resolution of one detector was so poor because of a strong rate dependence. See section 5.1.2. 171 200 150 — _. 100— T 50— - m H s: 0 1 1 1 g 300 U 250 200 150 100 50 30 40 so 60 7o 30 9o Excitation energy (arb. units) Figure F .4: Measured excitation energy with angular cut in 66", = 25° — 30°. The top panel shows the spectrum obtained if normal tracking is used, whereas the bot- tom panel shows the same spectrum but the new tracking method is applied. The resolution improves by almost a factor of 2. 172 Appendix G Fitting of Spectra In order to fit a spectrum one needs to assign proper errors a,- to each point 2:.- (z' = 0,1 .. .). Since in each channel the number of points is “counted”, Poisson statistics applies and an estimate is to assign This is only an estimate, since the real error has to be taken from the parent distri- bution, which is not known. The estimate from equation G] is especially bad if the statistics is low. As a general rule, the number of counts per channel should be more than 10 [22]. For lower statistics it is necessary to take the error from the fitted curve (i.e. of = f(:1:,-), where f (2:) was fitted to the data points y,- as f (27,) = 3),), which should follow the parent distribution much closer. However, there is another reason to use the errors obtained from the fitted curve, which is explained in the following. Consider a set of data points {23, y,} and a function f (11:) is fitted to these points. The X2 of the fit is given through x2 = Z [511. — may] . (G2) 173 Writing f (x) as b . f (51:) with parameter b yields X2 = Z [313(11- — b - fun-M2] - (G-3) Since x2 has a minimum for the best fit it follows that —6—X—2 — _—2Z[‘1—0_2(yi bf($z))'f “330] = 0' (GA) X2 can be written as 1% = Xjfam—bM( »—b“3%m—wfla»]. i i The last term is proportional toQ2L 2'and is thus equal to zero. It follows that 6b x — —Z;§ y‘( (gr—bf (an). (6.5) Setting a? = y,- yields X2 = 2 y,- — Z bf(:1:,-) = Area(DATA) — Area(FIT) ((3.6) Thus the fit underestimates the area of the data by X2, which can be quite consider- able. However if one sets of = f (xi), i.e. the error is deduced from the fitted curve, then it follows from equation G.4 that f3__x 0: b:—2Z( y«—bf(a:,-)) , ((3.7) and thus 23y.- — b f(x,-)) = Area(DATA) — Area(FIT) = 0 . (G.8) Therefore the fitted function has the same area as the data. In practice, one starts out with errors derived from the data (0,2 = y.) and when the fit looks good one uses the errors derived from the fitted function (0-2 = f (11.3)) and fits again. 174 LIST OF REFERENCES [1] GR. Satchler, Introduction to Nuclear Reactions (Oxford University Press, 1990) [2] A. Bohr and BB. Mottelson, Nuclear Structure, Vol. 1 and 2, (World Scientific, 1998). [3] A.M. Bernstein et al., Comments Nucl. Part. Phys. 11, 203 (1983). [4] K. Alder et al., Rev. Mod. Phys. 28, 432 (1956). [5] A. Winther and K. Alder, Nucl. Phys. A319, 518 (1979). [6] CA. Bertulani and G. Baur,Physics Reports 163, 299 (1988). [7] B. Davids et al., Phys. Rev. Lett. 81, 2209 (1998). [8] C.J. Benesh et al., Phys. Rev. C40, 1198 (1989). [9] S.S.M. Wong, Introductory Nuclear Physics (Prentice Hall, 1990). [10] S. Wan et al., Z. Phys. A358, 213 (1997). [11] J .D. Jackson, Classical Electrodynamics (John Wiley & Sons, 1975). [12] H. Scheit et al., Phys. Rev. Lett. 77, 3967 (1996). [13] H. Geissel et al., Annu. Rev. Nucl. Part. Sci. 45, 163 (1995). 175 [14] B. M. Sherrill et al. , Nucl. Instrum. Methods B56, 1106 (1991). [15] M.J. Chromik et al. , Phys. Rev. C55, 1676 (1997). [16] O. Sorlin et al., Phys. Rev. C47, 2941 (1993). [17] TR. Werner et al. , Phys. Lett. B335, 259 (1994);Nucl. Phys. A597, 327 (1996). [18] R. Harkewicz, Rev. Sci. Instrum. 67, 2176 (1996). [19] D. Swan, et al., Nucl. Instrum. Methods A348, 314 (1994). [20] J.N. Carter et al., Nucl. Instrum. Methods 196, 477 (1982). [21] N. I. Kaloskamis et al. , Nucl. Instrum. Methods A330, 447 (1993). [22] RR. Bevington, D.K. Robinson, Data Reduction and Error Analysis for the Physical Sciences (WCB/McGraw-Hill, 1992). [23] H. Scheit et al., Nucl. Instrum. Methods A422, 124 (1998). [24] J.W. Olness et al. , Phys. Rev. C34, 2049 (1986). [25] PM. Endt, Nucl. Phys. A521, 1 (1990). [26] PM. Endt, Nucl. Phys. A633, 1 (1998). [27] GM. Crawley et al., Phys. Lett. B64, 143 (1976). [28] W. Mayer et al., Phys. Rev. C22, 2449 (1980). [29] S. Raman et al., At. Data Nucl. Data Tables 36, 1 (1987). [30] BA. Brown and EH. Wildenthal, Ann. Rev. Nucl. Sci. 38, 29 (1988). [31] W.A. Richter et al. , Nucl. Phys. A523, 325 (1991). [32] BK. Warburton, J.A. Becker and BA. Brown, Phys. Rev. C41, 1147 (1990). 176 [33] G. Kraus et al., Phys. Rev. Lett. 73, 1773 (1994). [34] S. Raman et al., Phys. Rev. C43, 556 (1991). [35] CR. Satchler, Direct Nuclear Reactions (Oxford University Press, 1983). [36] E. Gadioli and RE. Hodgson, Pre-Equilibrium Nuclear Reactions (Oxford Uni- versity Press, 1992). [37] D. Larson et al., Phys. Lett. B42, 153 (1972). [38] SM. Austin, An Empirical Efiective Interaction; in The (p,n) Reaction and the Nucleon-Nucleon Force (Plenum Publishing Corporation, 1980). [39] J. Raynal, Computer code ECIS88, unpublished. [40] F.D. Becchetti, Jr. and G.W. Greenlees, Phys. Rev. 182, 1190 (1969). [41] J.-P. Jeukenne et al., Phys. Rev. C16, 80 (1977). [42] MD. Cortina-Gil et al. , Phys. Lett. B401, 9 (1997). [43] J.H. Kelley et al. , Phys. Rev. C56, R1206 (1997). [44] N. Alamanos et al., Preprint , (1998). [45] VA. Madsen et al., Phys. Rev. C12, 1205 (1975). [46] R.L. Kozub, Phys. Rev. 172, 1078 (1968). [47] RR. Johnson and R.J. Griffiths, Nucl. Phys. A117, 273 (1968). [48] J.F. Janni, At. Data Nucl. Data Tables 27, 150 (1982). [49] DEC Fortran: Language Reference Manual (Digital Equipment Corporation, 1995) 177 [50] R. De Leo et al. , Phys. Rev. C31, 362 (1985). [51] Private communication with Francois Marechal. Data is not yet published. [52] M.A. Kennedy et al., Phys. Rev. C46, 1811 (1992). [53] S. Wolfram The Mathematica Book, (Cambridge University Press, 1996). [54] A.N.F. Aleixo and CA. Bertulani, Nucl. Phys. A505, 448 (1989). [55] Frauenfelder, H. and RM. Steffen, 1965, Alpha-, Beta- and Gamma-ray Spec- troscopic, Vol 2, ed. K. Siegbahn . [56] K. Alder and A. Winther, Electromagnetic Excitation (North-Holland, Ams- terdam, 1975). [57] EU. Condon, G.H. Shortly, The Theory of Atomic Spectra (Cambridge Uni- versity Press, 1935). [58] CR. Lynch and O.I. Dahl Nucl. Instrum. Methods 358, 6 (1991). [59] H. Wollnik, Optics of Charged Particles (Academic Press, 1987). [60] R. Anne et al., Z. Phys. A352, 397 (1995). [61] RF. Casten, Nuclear Structure from a Simple Perspective (Oxford University Press, 1990). [62] International School of Heavy Ion Physics; 4th Course: Exotic Nuclei Ed. R.A. Broglia and PC. Hansen (World Scientific, 1998). [63] M. Fauerbach et al., Phys. Rev. C56, R1 (1997). [64] RD. Gill, Gamma-Ray Angular Correlations (Academic Press, 1975). [65] T. Glasmacher et al., Phys. Lett. B 395, 163 (1997). 178 [66] K.L.G. Heyde, The Nuclear Shell Model (Springer-Verlag, 1994). [67] R. W. Ibbotson et al., Phys. Rev. Lett. 80, 2081 (1998). [68] IGOR Pro User’s Guide, IGOR Pro Programming and Reference Manual, (WaveMetrics Inc., 1996). [69] W.R. Leo, Techniques for Nuclear and Particle Physics Experiments (Springer- Verlag, 1994). [70] T. Motobayashi et al., Phys. Lett. B346, 9 (1995). [71] CM. Perey and EC. Perey, At. Data Nucl. Data Tables 17, 1 (1976). [72] F. Petrovich et al., Ann. Rev. Nucl. Part. Sci. 36, 29 (1986). [73] P. Ring, P. Schuck, The Nuclear Many-Body Problem (Springer-Verlag, 1980). [74] T. Suomijarvi et al. Nucl. Phys. A509, 369 (1990). 179