7. trunk. 31+. ix $2.. 3! .3... 2m A». i... 1 Ha: . ”53% vv‘ . tunic... 1.5.1.? 34.3.. 31 s a. 5.1. «it! Patszgfiaz 31...}. ._ .5. . x 3.2.... is... nan“ an...“ .. 6.2,“. a»... 5‘ ‘I 1 flu; iva‘... Its .- r I. .97:- .. [Off] LlBRARY Michigan State University J This is to certify that the dissertation entitled INTERMEDIATE-ENERGY PROTON KNOCKOUT TO PROBE SINGLE-PARTICLE STRUCTURE AND NUCLEAR SPIN ALIGNMENT IN THE “ISLAND OF INVERSION” ISOTOPES 3"33MG presented by David Miller has been accepted towards fulfillment of the requirements for the Doctoral degree in Physics and Astronomy //flw/ , Major Professor’s Signature 7/16/05? Date MSU is an Affirmative Action/Equal Opportunity Employer 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 5108 K:IPrq/Aoc&Pres/ClRC/DateDueindd INTERMEDIATE-ENERGY PROTON KNOCKOUT TO PROBE SINGLE-PARTICLE STRUCTURE AND NUCLEAR SPIN ALIGNMENT IN THE “ISLAND OF INVERSION” ISOTOPES 31’33MG By David Miller A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Physics and Astronomy 2009 ABSTRACT INTERMEDIATE-ENERGY PROTON KNOCKOUT TO PROBE SINGLE-PARTICLE STRUCTURE AND NUCLEAR SPIN ALIGNMENT IN THE “ISLAND OF INVERSION” ISOTOPES 31,33MG By David Miller The “island of inversion” isotopes 31’3"ng were investigated through a proton-knockout reaction from 32"MAI at 90 MeV/u at National Superconducting Cyclotron Labora— tory (NSCL). Single-particle states, with no neutron excitations across the N = 20 shell gap from the sd-shell, were identified through the determination of the par- tial cross sections in the residue. The single-particle states identified lie above the ground state agreeing with the placement Of these magnesium isotopes in the island Of inversion. Nuclear spin alignment is evident following the reaction as determined by the an- gular distribution of emitted gamma rays. Angular-distribution analysis constrained by the outgoing residue longitudinal momentum allows for the determination of mul~ tipolarity when combined with linear polarization measurements. Incorporating mo- mentum constraints factors out possible systematic uncertainties of determining the in—beam gamma-ray detection efficiency. Development of digital data acquistion sys- terns provides enhanced capabilities for angular distribution and linear polarization measurements. Instrumentation developed for the array of gamma-ray detectors (SeGA) at NSCL and considerations for next-generation gamma-ray detectors are discussed. ACKNOWLEDGMENT I would like to personally thank the members of the committee for their guidance through my trek through the preparation Of my thesis: Bill Lynch, Stuart Tess- mer, Vladimir Zelevinsky, Carl Schmidt, and especially my thesis advisor, Krzysztof Starosta. Their insight has kept me focused on the task at hand through the toilsome years of graduate school. I am also grateful to the support staff that made my exper- iment run smoothly, and to my fellow graduate students who sat shifts to Obtain my data - especially to fellow member of Team Starosta, Phil. The analysis could not have been done without the work of other students who have went before me; Russ Terry’s thesis was particularly helpful in the writing process. I have enjoyed the camraderie and benefited from many of my peers. Working through hiccups in problem sets in the early years made things much more pleasant. I’ve had a host Of wonderful officemates to provide additional input about lab-related issues, as well as proactively distract me. My fellow housemates at the Neutron Society have made the last couple years here more cheerful than I ever expected, and the crowd across several departments that mobilizes for Sunday evenings has also (generally) made those Mondays just a bit more tolerable. I am also appreciative of the sanity checks that my “running” club GLH3 has given me and the support of a number of its members. I’m glad that I’ve left an impact on the club, and hope the friends I’ve coerced (HPP, SBH, UO, SSL, and others yet to be named) enjoy it in my absence. Meeting BS and FR in my first months here was comforting. The Fourth of July shindigs at CBS’s are memorable, and he has also been particularly supportive and entertaining. Without *HP, it would have been difficult to find the excitement Of running, to do something crazy like tackling a marathon, or to rediscover the joy Of bicycling (including all those wonderful DALMAC experiences). The physical activity that kept me sane through the years must also give credit to the Tuesday morning basketball gang, even though iii I rarely wanted to stir that early in the morning. My curiosity and thirst for knowledge is clearly accredited to my upbringing. I remember many of my early learning experiences being attributed to interactions with my siblings and parents. I am glad that their intellectualpursuits continue and especially thank the encouragement they have given me throughout the years. Of course, the National Science Foundation is highly important for the funding of the research; may their budget ever grow. I also should acknowledge Oak Ridge Associated Universities for giving me the most memorable experience Of my time at Michigan State by giving me an opportunity to be a participant in the Lindau meeting of Nobel Lauereates. I will remember that trip and the people I met there for years to come. iv TABLE OF CONTENTS List of Tables ................................. vii List Of Figures ................................ viii Introduction ...................................................... 1 1.1 The nucleus ................................ 1 1.2 Nuclear structure and shell model .................... 4 1.3 Shell model near the “Island of Inversion” ............... 8 1.4 Nucleon knockout reactions ....................... 10 Methods ......................................................... 15 2.1 Overview .................................. 15 2.2 Isotope production ............................ 16 2.2.1 Ion source ............................. 16 2.2.2 Coupled Cyclotron Facility .................... 17 2.2.3 A1900 ............................... 17 2.3 Particle detection ............................. 18 2.3.1 The S800 focal plane detectors .................. 20 2.3.2 Inverse mapping .......................... 22 2.4 Gamma-ray detection ........................... 23 2.4.1 Gamma-ray angular correlations ................. 25 2.4.2 Gamma ray linear polarization .................. 26 2.4.3 Segmented Germanium Array . - ................. 30 2.4.4 Relativistic kinematics ...................... 31 2.4.5 Velocity determination ...................... 32 Offline Detector Characterization ................................. 34 3.1 Energy calibration ............................ 34 3.2 Efficiency calibration ........................... 37 3.3 Polarization sensitivity determination .................. 39 3.3.1 Detector setup .......................... 40 3.3.2 Gamma-ray angular distribution ................. 42 3.3.3 Scattering asymmetry normalization .............. 45 3.3.4 Sensitivity determination ..................... 45 3.3.5 Figure-Of-merit discussion .................... 48 Proton knockout studies .......................................... 51 4.1 Electronics trigger ............................. 52 4.2 31Mg .................................... 53 4.2.1 Beam characteristics and particle ID .............. 53 4.2.2 Level scheme ........................... 56 4.2.3 Inclusive and partial cross sections ............... 61 4.2.4 Momentum distribution Of residues ............... 66 4.2.5 Angular distribution of gamma rays ............... 67 4.2.6 Linear polarization Of gamma rays ................ 70 4. 2. 7 Spin and parity assignments ................... 71 4.3 33Mg .................................... 74 4.3.1 Beam characteristics and particle ID .............. 74 4.3.2 Level scheme ........................... 75 4.3.3 Inclusive and partial cross sections ............... 80 4.3.4 Momentum distribution Of residues ............... 83 4.3.5 Angular distribution of gamma rays ............... 84 4.3.6 Linear polarization of gamma rays ................ 86 4.3.7 Spin and parity assignments ................... 87 5 Commentary ..................................................... 90 5.1 31Mg .................................... 90 5.1.1 Calculation Of momentum distribution ............. 90 5.1.2 Structure Of Observed excited states ............... 91 5.1.3 Spin alignment calculation and angular distribution ...... 94 5.2 33Mg .................................... 95 5.2.1 Calculation of the momentum distribution ........... 95 5.2.2 Structure Of populated states .................. 96 6 Outlook .......................................................... 102 6.1 Signal decomposition and gamma-ray tracking ............. 103 6.2 SeGA digital electronics ......................... 106 6.3 GRETIN A impact ............................ 112 7 Conclusions ...................................................... 116 A Gamma-ray angular distribution .................................. 118 B Relativistic Kinematics ........................................... 125 B.1 Lifetime .................................. 126 B2 Angle ................................... 126 B3 Energy ................................... 127 B4 Polarization ................................ 127 Bibliography ........................................................ 130 vi 3.1 4.1 4.2 4.3 4.4 4.5 6.1 A.1 LIST OF TABLES Comparison of SeGA to other detector arrays for polarization measure- ments. ................................... 49 Secondary beam characteristics. ..................... 52 Gamma-ray transition properties in 31Mg ................ 57 31Mg states with direct feeding from the proton knockout from 32Al. 65 Gamma-ray transition properties in 33Mg ................ 76 33Mg states with direct feeding from the proton knockout from 34Al. 82 Relevant quantities for angular distribution measurements for GRETA angles assuming ,6 = 0.4 .......................... 113 Maximum alignment coefficients for different types of alignment. . . . 120 vii 1.1 1.2 1.3 1.4 2.1 2.2 2.3 2.4 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 LIST OF FIGURES Chart of nuclides ............................. 3 Single-particle energies in spherical shell model ............. 6 Evolution Of the sd—p f shell gap ..................... 9 Cartoon Of a knockout reaction. ..................... 11 Schematic Of the S800 spectrograph. .................. 19 5800 Focal plane coordinate definition ................. 23 Polarization digaram ........................... 28 Depiction Of the segmentation of a SeGA detector ............ 31 226Ra energy spectrum for calibration .................. 35 Efficiency of SeGA ............................ 38 249Cf decay scheme ............................ 40 my coincidence setup ........................... 41 245Cm gamma ray energy spectrum following the a decay Of 249Cf . . 43 a2 coefficient fit for a-7 coincidences .................. 44 Scattering relative to reaction plane for a SeGA detector ....... 46 Geometric Compton scattering asymmetry for unpolarized gamma rays 47 Normalized Compton scattering asymmetry for polarized gamma rays 48 viii 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 4.16 4.17 4.18 4.19 4.20 4.21 4.22 Identification of incoming 32Al particles ................. 54 Identification of outgoing residues from incoming 32A] beam ..... 55 31Mg gamma-ray spectrum ........................ 56 Gamma-ray spectrum of 7—7 coincidences in 31Mg ........... 58 31Mg level scheme ............................. 60 Observed doublets in 31Mg ........................ 61 Acceptance cut for 31Mg isotopes .................... 63 Run-by-run inclusive cross section for 31Mg .............. 64 Longitudinal momentum distribution of 31Mg ............. 66 Angular distribution effects Of decays from 221 keV state in 31Mg . . 67 Automatic fitting of gamma-ray peaks in 31Mg ............. 69 Momentum-gated angular distribution of 623 keV gamma rays in 31Mg 70 Compton scattering asymmetry Of gamma rays in 31Mg ........ 72 ag-P diagram for gamma-ray decays in 31Mg .............. 73 Identification of incoming 34A] isotopes ................. 75 Identification Of residues resulting from fragmentation of 34Al beam . 76 Gamma-ray spectrum for 33Mg ..................... 77 Gamma-ray spectrum for 7-7 coincidences in 33Mg .......... 79 33Mg level scheme ............................. 80 Acceptance cut for 33Mg isotopes .................... 81 Run-by-run inclusive cross section for 33Mg .............. 82 Longitudinal momentum distribution for 33Mg residues ........ 83 ix 4.23 Momentum-gated angular distribution for 33Mg gamma rays ..... 4.24 Compton scattering asymmetry of gamma rays in 33Mg ........ 4.25 ag-P diagram for decays in 33Mg .................... 5.1 5.2 5.3 6.1 6.2 6.3 6.4 6.5 6.6 6.7 A.1 A.2 Effect Of quadrupole mixing on the angular distribution coefficient . . Convolution Of theoretical momentum distribution and experimental effects ................................... Nilsson diagram used for 31’33Mg .................... Induced signals recorded by digital data acquisition system ...... HPGe drift velocities ................. i .......... Timing characteristics for different channels Of DDAS ......... Energy-time matrix recorded by DDAS .. ................ Segmentation Of SeGA based on the drifting of electron holes to seg- ment electrodes .............................. Grid points for waveform sampling ................... Angular distribution sensitivity ..................... Angular distribution coefficients and linear polarization values for se- lected transitions ............................. Angular distribution coefficients and linear polarization values for dif- ferent prolate alignments ......................... 84 86 88 93 96 100 104 105 108 109 110 114 123 Chapter 1 Introduction 1. 1 The nucleus One of the aims of science is to understand the composition of matter which makes up our planet, solar system, and universe. With increasing knowledge, mankind has broken matter into its smaller building blocks. Molecules are composed Of atoms, and at the heart of every atom lies the nucleus. The properties of the nucleus are determined by its constituent parts which consist of chargeless neutrons and positively charged protons. These constituents are collectively known as nucleons. A nucleus is a quantum many—body system with a given number Z of protons and N of neutrons which characterize it as a particular isotope. A specific nucleus is Often referenced by 24x N where A is the total number of nucleons and X is the chemical symbol. The properties of nuclei are Of fundamental interest to understand the underlying forces which are responsible for binding the nucleons together. Nuclei are small dense Objects with sizes on the length scale Of femtometers (10"15 meters) and masses of approximately A GeV/c2. Each isotope has a binding energy B(N, Z) as a result of the forces interacting within the nucleus. In a simple model, these macroscopic properties can be calculated considering the nucleus as a liquid drop [1] including 1 volume and surface effects. Coulomb repulsion and strong pairing of the nucleons also must be treated tO reproduce the Observations. The strong pairing of nucleons in time—reversed orbits leads to an enhanced binding Of isotopes with an even number Of neutrons or protons. The forces in the nucleus conspire tO bind certain nuclei more than others. The nucleus could be more bound if a neutron changed to a proton or vice versa. This process is governed by fi-decay and causes some nuclei along a line with the same A to be stable. Other isobars having the same A decay towards this valley of sta- bility by sequential ,8 decays. For a given element, only a small number of nuclides are stable while a greater number are unstable. This is shown in the chart Of the nuclides (Fig. 1.1) which displays all the isotopes with different neutron and proton numbers. Heavier isotopes can also be unbound through other processes. These include a-decay where a doubly-charged 4He tunnels through the nuclear potential and escapes the nucleus as well as fission where the nucleus separates into two or more pieces. For nuclei with extreme proton-to-neutron ratios relative to the valley of the stability, individual nucleons can be unbound and emitted by the nucleus. The proton (or neutron) dripline is where it is no longer possibly to add an additional bound proton (or neutron), and the nucleus will decay by particle emission. In part, this can be characterized by the difference of binding energies for-neighboring iso— topes. These are called the proton separation energies (3,, = B (N, Z) — B (N, Z — 1)) and neutron separation energies (Sn = B (N, Z) - B (N — 1, Z )) However, the rate of charged particle decay can also be hindered by the probability of tunneling through the Coulomb barrier causing these to be particle-stable and proceed to the valley Of stability through beta decay. Beyond a naive macroscopic model, microscopic effects must be accounted for which impact the structure Of the observed nuclei. The nucleus is a strongly interact- ing system where several forces of the Standard Model converge. Theoretical models 2 100 - : y — I. ‘ll ,_/:’«‘“' 80 - - - - - _ g ' ‘,/A/ E I 3 .. , c 60 - - ' r-process - c ‘ . I 3 rp-process -' _ ‘.'----J-- o . I- . l O. 40 - , ~ - I ' 20 - _, i I Observed nuclei — I Stable nuclei 0 ' I I I I I l L" 0 20 40 60 80 100 120 140 1 60 180 Neutron number Figure 1.1: Chart of the nuclides with proton number Z on the ordinate and neutron number N on the abscissa. Stable isotopes are shown in black while unstable nuclei are shown in grey. Many of these unstable nuclei have been observed (light gray), but a large number of these nuclei are outside the current reach of experimental facilities with the predicted boundaries illustrated in dark gray. Magic numbers with increased binding energy are shown as dashed lines. Several astrophysical processes are also shown which involve nuclei far from the line of stability. must treat the strong, weak, and electromagnetic interactions tO describe the prOper- ties of nuclei. Since the strong force acts non-perturbatively, the exact mathematical form of the nucleon-nucleon interaction cannot be calculated explicitly. Instead, the form is determined in the context of several different approaches either theoretical or phenomenological [1]. Different approaches come to comparable results which agree well with nucleon-nucleon scattering data across a wide range of energies. The nu- 0—15 meters. cleon potential must be strong and short—ranged on the range of a few 1 It must also have a repulsive core at small distances. Two important factors which add much to the richness of nuclear structure are the presence of a strong spin-orbit interaction as well as a tensor interaction which allows mixing between states of dif- ferent orbital angular momenta. Early phenomenological models fit the interaction strength to scattering data with four parameters with quite a bit of a success as in the Hamada-Johnston potential [2]. Modern phenomenology has fine tuned to the po— tential to accommodate further data by adding some further parameters such as the Argonne v13 potential [3]. Ab initio approaches attempt to derive properties of nuclei by folding the nucleon—nucleon interaction over all the constituents of the nucleus. However, this becomes computationally intractable for nuclei with masses A Z, 10. For heavier nuclei, a shell model is Often employed which approximates the nucleon- nucleon interactions with a spherical potential well which is phenomenologically fit to reproduce key nuclear prOperties. Residual interactions are then considered between a selected subset of the nucleons. 1.2 Nuclear structure and shell model Nucleons have an intrinsic angular momentum, s = 1/2h, known as spin. Inside the nucleus, this spin couples to the orbital angular momentum, E, Of the nucleon to form the total angular momentum j. In the nuclear shell model, these nucleons fill 4 orbitals denoted by Tnéj where 7' describes the type of nucleon (u for neutrons and 7r for protons). These single-particle states are bunched in groups according to their principal quantum number n in a harmonic oscillator potential. All the nucleon single-particle states for a given nucleus are coupled to a final total angular momentum Of J (colloquially referred to as spin). Due to the strong pairing in nuclei, nuclei with an even number of protons and neutrons always have a ground state spin of 0+. For Odd-A nuclei, the nuclear spin is dominated by the effects Of the unpaired nucleon. Furthermore, the single-particle orbitals have a definite parity 7r = (—1)€ which describes the properties Of the wavefunction under spatial inversion. The parity of the nucleus is the simple product of all the individual nucleon parities. Measurement Of the parity gives access to determining the nature of the orbital angular momentum Of the nucleon. Each of these orbitals also has a corresponding single-particle energy within the mean field of the nucleus. Similar to electron orbitals in the noble gases, certain “magic” numbers of nucleons are considerably more stable [1]. This arises from a gap in the single-particle energies at the top of the Fermi surface. Isotopes which have a magic number Of both neutrons and protons are called doubly-magic, while those that have a magic number of one or the other are known as semi-magic. Magic nuclei tend to be more spherical, and for even-even nuclei possess a higher excitation of the first excited state (E (2?» and a smaller reduced transition probability (B (E2; 2iF —> 0?». These magic numbers (canonically 2, 8, 20, 28, 50, 82, and 126) create a clear division of the single-particle orbitals into different shells as shown in Fig. 1.2. With these magic numbers, there exists a handful of particle-bound doubly-magic nuclei (4He, 160, 40Ca, 48Ca, 56Ni, 78Ni, 100Sn, 132Sn, and 208Pb). These shells are referred to by the single-particle orbital angular momentum values included above the previous closed shell; i.e. s-shell, p—shell, and sd—shell for the shells up to 2, 8, and 20 single-particle orbitals, where s, p, d, f, g, . . . correspond to f’ = 0, 1, 2, 3,4, . . . 5 4 3s _ .7: 1 — -3SI/2 2 4 2d 2‘ I,’1g7/2 8 ’““- 2d5/2 6 Ig (’ 6 9 \ ‘199/2 10 2p I ’1f5/2 6 ‘ — - ~ I 3 T :4. ‘ P ‘ 291/2 2 1f <: ‘ 293/2 4 ‘1f7/2 8 28 _ , , -1d3/2 4 2 1d ’,:"“-281/2 2 IIIIII 1d5/2 6 1 1p ——————— 191/2 2 ..... 193/2 4 0 Is ------- 131/2 2 N nl nlj 2j+1 Figure 1.2: Shell model denoting single—particle energies up to the 1211/2 orbital be— fore (left) and after (right) including spin-orbit effects [4]. Circled magic numbers correspond to the gaps in the single-particle energies. respectively. Each orbital has 23' + 1 magnetic substates after which the orbital is filled. When all the orbitals in a shell are filled, that shell is closed, and the addition of a further nucleon requires significantly more energy due to the gap in the single— particle energies. Recent experiments with rare isotope beams have shown that the magic numbers which apply near stability can evolve as one proceeds towards more exotic isotopes [5]. These experiments indicate the weakening of the existing shell closures and the ap- pearance of new magic numbers [6,7]. Furthermore, the structure of the nuclei near the dripline can exhibit curious effects such as proton/ neutron skins or halo states such as in 11Li [8]. Since the nuclear potential is connected tO the spatial distribution Of the nucleons, these exotic structures cause a modification in the mean field in the nuclear medium. Due to the short range nature of the strong force, the mean nuclear potential can be approximated by a Woods-Saxon function: T—T‘ fws(7‘) = VII/(1 +exp< 0)) (1.1) with a depth V0, radius 1‘0 and diffuseness a. A large spin-orbit term must be included as well in order to reproduce the Observed shell gaps shown in Fig. 1.2. The Woods- Saxon potential, particularly the depth, is adjusted to reproduce the observed single- particle energies [5,9]. Where single-particle energies are not available, they can be extracted in a Hartree—Fock calculation with a self-consistent interaction based on effective energy—density functionals such as the Skyrme SkX interaction [10]. Once the properties of the single-particle orbits are derived, a residual two-body interaction remains in the Hamiltonian which is adjusted to reproduce other exper- imental observables. The nuclear wavefunctions are determined by diagonalization of this Hamiltonian. The residual two—body Off—diagonal matrix elements between 7 single-particle orbitals result in an additional correlation energy for a given eigen- state. To allow calculations to be computationally feasible, approximations about the single-particle states must be made. Generally, the nucleus is considered tO have an inert core where the orbitals are fully occupied and interactions are minimal such as in a nearby doubly-magic nucleus. The additional neutron and proton particles or holes are then built onto this inert core. Furthermore, the number of single-particle states must be truncated to a certain valence space. The two-body matrix elements are then calculated phenomenologically within the model space outside the inert core. However, the assumptions about the core and valence space depend on the nucleus of interest, so the derived effective interaction is only applicable to a certain region of the nuclear landscape [5]. In the present work, the region of interest is centered around Z = 12 and N = 20. TO describe states with normal single—particle ordering (as in Fig. 1.2), the Universal SD (USD) interaction [11] is used which assumes an inert core of 160 or 40Ca and includes the 1d5/2, 231/2 and 1d3/2 orbitals in its model space. 1.3 Shell model near the “Island Of Inversion” For the isotOpes with Z w 8 to Z x 13, the shell structure evolves considerably for neutron-rich nuclei. Here, the N = 20 shell closure is considerably weakening, while experiments show a pronounced N = 16 subshell closure [7] (see Fig. 1.3). With the weakening of the N = 20 shell gap, neutron excitations across the gap appear at relatively low excitation energies. The states built on these neutron excitations are called intruder states, and are referred to as npmh states by the number Of neutrons n excited into the fp-shell and remaining m holes in the sd-shell. TO treat these states properly, the 1 f7 /2 and 2193/2 orbitals must be included in calculations greatly increasing the model space. _ —- q- _ ’>‘~ 8 (a) }pf shell (b) a. 303. —— __—};;she" 2.0 E 14 16 I,” 3/2 a 16 w 4' ,x‘“ .® 1 (013 Ode/2 LU O '— 1S1/2 _" 1S1/2 _ C ( ) j<= l — 1/2 j): l + 1/2 proton neutron Figure 1.3: Effective single-particle energies for N = 16 isotones for silicon (a) and oxygen (b). Note the shift in the 0d3/2 orbital which is driven by the interaction shown in (0) between protons and neutrons with the same orbital angular momentum but opposite spins. This proceeds through the 07' Operator (d) (adapted from Otsuka et a1. [12]) For the nuclei with Z w 11 and N z 20, the correlation energy from the residual two-body interactions makes these intruder states the energetically preferred configu- rations. This region is known as the “island of inversion” [13] where the ground-state wavefunction is dominated by particle-hole excitations across the N = 20 shell gap. Experimental evidence Of the breakdown of the N = 20 shell closure is given by a small 21* energy in 32Mg [14] and an increase in collectivity [15,16]. Furthermore, ex- citation across the N = 20 shell gap are necessary to appropriately describe the mass systematics in the region [17]. To describe these cross-shell excitations, the T = 0 proton-neutron interaction is a major factor especially between the (15/2 protons and d3/2, f7/2 neutrons [13]. The mechanisms that influence this region are also believed to be important in other regions of the nuclear landscape where inversion has been 9 Observed, such as in 12Be [18,19], or is theoretically predicted to occur [13]. The boundaries of the island of inversion have been proposed by including the entire fp—shell into the model space [13] and deducing that increased pairing and proton-neutron interactions play a significant role in the region. Theoretically, the extent was originally suggested to lie between 10 S Z s 12 and 20 S N S 22. However, recent measurements Of the ground state spins of magnesium isotopes (Z = 12) have contributed information about the energetically favored configurations in this region [20—22]. With the additional information provided by data from neutron knockout [23] and Coulomb excitation [15,16], the isotones with N = 19 are also included in the island of inversion. Furthermore, recent results indicate intruder configurations in more neutron-rich 36Mg [24] and less neutron-rich 28Ne [16,25] extending the reach Of the island further than originally anticipated [13]. Determining the boundaries and mechanisms in the island of inversion is an important test Of theoretical models which predict properties far from stability. 1.4 Nucleon knockout reactions Nucleon knockout reactions are a powerful spectroscopic tool at intermediate ener- gies to study nuclei. After the removal of a nucleon from an incoming projectile, the longitudinal momentum distribution of the residue is sensitive to the orbital angular momentum Of the removed nucleon [26] as well as the distribution of magnetic sub- states in the residue. The partial cross sections to states populated in the residue also contain information about the wavefunction overlap between the neighboring nuclei. In the transitional region near the island of inversion, this can selectively probe states with intruder-like or normal configurations. In a typical neutron—knockout reaction (Fig. 1.4), a nucleus Of mass A impinges On a light nuclear target (such as 9Be or 12C) and the reaction residue Of mass A — 1 10 Knockout ' t'l PrOJec I e residue Gamma ray 9Be Target Figure 1.4: Cartoon of a knockout reaction. exit moving at nearly the same velocity of the Original projectile. The change in velocity is related to the momentum of the removed nucleon. The residue frequently leaves in an excited state which decays by gamma-ray emission. For energies above 50 MeV/ nucleon, the reaction mechanism is simplified and can be considered to proceed in a single step. If the removed nucleon is not observed, the reaction encompasses a number of breakup channels. The dominant processes are diffractive breakup and stripping. For the stripping reaction, the nucleon is absorbed by the target as opposed to diffractive breakup where the nucleon and core interact elastically in the nuclear mean field of the target. Knockout is predominately a peripheral processes which probes the nuclear surface removing the most weakly-bound valence nucleons. For a given reaction on a single scattering center, the cross section relates the rate of outgoing flux per unit surface area relative to the incoming flux of particles. Given a target of thickness T and density p with molar mass m z A - 103 mg, the number of scattering centers per unit area is pTNo / m where N0 is Avogadro’s number. The nominal thickness t is often quoted in units of mg/cm2 or t = pT. If there are N,- 11 incoming particles and N f outgoing residues, the total cross section is then a = ———. (1.2) Nuclear cross sections are typically quoted in units Of barns (10‘24 cm2) with knock- out reactions having cross sections around several millibarn. The cross section to individual states is sensitive to the structure Of the incoming and outgoing nuclei, including angular momentum selection rules. For the removal of a nucleon with total angular momentum j, the final states in the nucleus, (J 7') f must follow the selection rules: [Jr—Jl P<9). (2.8) The sensitivity depends on the incident energy and is limited by the difference in the scattering cross section. For a point-like polarimeter, the sensitivity Q0 calculated from the Klein-Nishina cross-section is: 1+0: W“) = m (2.9) E . . . . where a = T—n—Ez. For practical applications, the scattering asymmetry must address 6 the geometric asymmetries of the detector. The asymmetries are compared to those Observed from unpolarized gamma rays which are characterized by a parameter a. N unpolarized (0 I 45 = 900) (“E”) = NunpolarizedW: ¢ = 00) (2’10) _ N(6, a = 90°) — aN(9, a = 0°) ”4(9) " N(6,0 = 90°) + aN(6,0 = 0°) (2‘11) To determine the sensitivity of a detector, the analysis is done for gamma rays which have a well known polarization such as for pure dipole transitions which have an 29 angular distribution W(9) 2 1+ a2P2(cos 6). (2.12) For pure dipole transitions, the polarization distribution reduces to the form [53] 3 ' 29 P(6) = d: “2 8m , 2 (2.13) 2 + 20.2 - 3a2 Sln 0 where a2 is determined from analyzing the angular distribution Of gamma rays with (f 3, a “+” Sign for M1 transitions and a sign for E1 transitions. 2.4.3 Segmented Germanium Array The Segmented Germanium Array (SeGA) [54] serves as the workhorse for gamma ray spectroscopy at NSCL. It combines the superb energy resolution of germanium with a modest efficiency afforded by the array of detectors. In addition, angular res- olution smaller than the detector size can be achieved using the internal electronic segmentation of the individual crystals. Eighteen detectors of n—type high-purity ger- manium (HPGe) comprise SeGA. Each is a cylindrical single crystal of germanium cooled by liquid nitrogen to 100 K. The crystal is 8 cm long with a 7 cm diameter. A central core electrode is inserted along the central axis of the crystal. The outside surface is electronically segmented into 32 segments with 4 azimuthal divisions and 8 longitudinal slices (see Figure 2.4). A bias voltage applied to each detector causes electrons liberated by an incoming gamma ray to drift to the center Of the detec- tor. The corresponding holes generated drift outward and are collected by electrodes placed on the outer surface for each segment. The detectors are configurable into several standard arrays. For the present work, the “classic” SeGA configuration was used. This places the detectors in two rings sur- rounding the target position. Seven detectors are placed in the forward 37 degree ring, 30 Figure 2.4: Depiction Of the segmentation of a SeGA detector. and ten are placed in the backward 90 degree ring. The crystals are positioned such that the transverse segmentation lies along the 0 direction defined by the coordinate system with 2 along the beam axis. 2.4.4 Relativistic kinematics Experiments at National Superconducting Cyclotron Laboratory are typically per- formed with velocities from 30%-50% the speed Of light. At these velocities, rela- tivistic effects become considerable and must be accounted for in the analysis. In particular, the velocity of the particle, the energy of emitted radiation, and the angle of the radiation with respect to the nucleus’ velocity all become correlated as dis- cussed in detail in Appendix B. Particle tracking in the S800 spectrograph supplies the necessary information about the magnitude and direction of the residual particle’s velocity. To be able to precisely measure the gamma ray’s energy, the angle of emis- 31 sion with respect to the beam axis must be known. This is where the segmentation of the detectors is critical, allowing the determination Of the first interaction position within a detector based on the energy deposited in the individual segments. Gener- ally, the location Of the first interaction is taken as the midpoint of the segment where the most energy was deposited. This algorithm accurately identifies the segment with the first interaction point in most of the cases [55] especially when considering the possibility of having multiple interactions within a segment. However, the precision in the determination Of the interaction point is limited by the size of the segment. Furthermore, the exact location where the nucleus was at the time of 'y-ray emission is also unknown. Often this occurs within the target where the nucleus is actively slowing down leading to an additional effect related to the velocity uncertainty. These two effects are the major contributors to the limits Of the energy resolution obtainable in an experiment with intermediate energy projectiles. 2.4.5 Velocity determination SeGA was centered around the target position Of the S800 to detect the gamma rays emitted following the secondary reaction. Gamma-ray spectra were Doppler corrected based on the energies Of the fragments of interest at the middle of the target. The velocity of the fragments was determined in several ways which proved to be self- consistent. First, one can get the velocity based on the incoming particle energy and energy losses in the target from LISE++ [56] which uses a model based on the ATIMA 1.2 [57] characterization for heavy ion energy loss in matter. Alternatively, the velocity can be obtained on the downstream side of the target based on the de- viation from the central momentum in the S800 spectrograph and back trace to the center of the target using the energy losses Of LISE++. Energy losses based on GEANT4 also agree reasonably with those Obtained with LISE++. Furthermore, one could estimate the velocity based on the uncorrected DOppler-shifted energy de- 32 posited in the 90° ring Of SeGA which have their first interaction point within a different longitudinal segment and thus a different Opening angle. This constrains the velocity if the 7-ray energy and positioning of the detector are well known (see Eq. 38). Combining the systematic differences using these different methods, one Obtains a reasonable estimate of the uncertainty in the velocity. 33 Chapter 3 Offline Detector Characterization In preparation for the experiment with the beam, several source measurements were used to characterize the response Of the detector. These were necessary to establish the calibration parameters for energy and efficiency. Furthermore, the detector’s re- sponse to linear polarization also had to be investigated and its sensitivity quantified. 3. 1 Energy calibration To determine the energy Of the gamma rays detected in SeGA, one must know how the detector and the Analog to Digital Converters (ADCs) respond to gamma rays Of previously known energies. TO accomplish this task, several standard radioactive calibration sources were used. This allows one to characterize the response of the detector across a wide range of energies. Furthermore, the central contact of the de- tector and the 32 segments have to be calibrated separately as they behave differently to the photon’s energy deposit in the detector. For the central contact, the 13-bit channel output, 1:, from the ADC was assumed to be quadratically related to the energy Of the incoming photon, E, i.e. for some calibration coefficients a,, E = a0 + alt: + (I21:2 (3.1) 34 x103 103 > I .3 I :7 1’ 120 a E S : 100 o I O I 3 o i ; o i .1 H. 200 400 600 800 1000 1200 1400 1600 1800 2000 Energy(keV) Figure 3.1: Calibrated 22(’Ra energy spectrum of all 16 SeGA detectors summed. To determine these coefficients, a :226th radioactive source was placed in the target position of the array, and the gamma rays with known energies from 186 keV to 2448 keV were detected in the individual detectors. A Gaussian with a quadratic background was fit to each Of the photopeaks in the resulting gamma ray spectrum in order to extract the centroid channel of the ADC output. The calibration coefficients for each detector were chosen such that they minimized the chi squared comparing the calculated energies, E, in the detector and the known peak energies, E,. . 0'2 2 where a2 is the variation associated with the fitting of the peak. After calibration, the known 22°Ra spectrum is reproduced beautifully as seen in Fig. 3.1. Over the course of the experiment, calibrations parameters may shift slightly. To account 35 for this possibility, calibration data were taken at the beginning and the end of the experiment. In addition, an abrupt shift in the response Of several of the detectors was observed in the middle of the experiment. This was ameliorated by performing a linear correction to the calibration coefficients based on natural background sources in the experiment, in particular the electron-positron annihilation peak and peaks associated with natural potassium decay. These adjustments allowed for an accurate calibration for all the detectors across all the runs. The proper calibration of the energies deposited in individual segments is also im- portant as the energies of the segments is what the Doppler reconstruction algorithm relies on. The usual technique used is to gain match the segment energies such that it agrees with the central contact energy. One such calibration method is described in Hu et a1. [58] where an automated procedure is important as a typical SeGA configu- ration has ~ 500 segments which need calibration. One benefit to matching with the central contact energies is that all the data including events with incomplete 7-ray absorption can be used in the calibration. However, there are a sizable number of events where the measured energy deposited in the segments is deficient compared to the amount deposited in the central contact. These outliers can cause a simple x2-analysis to have small negative deviations from the expected results. To alleviate such deviations, an iterative approach was utilized to obtain the segment energy cal- ibration parameters. With each pass, events with large deviations between a single segment energy deposit and the central contact energy were excluded from the anal- ysis. The last pass rejected events which strayed more than 5 keV from the expected value from the central contact. At each pass, parameters were determined for each segment which minimized the X2 Of the segment energy relative to the central contact energy. The response of the segments is significantly less linear, so a quartic polyno- mial was used in the fit to the central contacts. There exists an electrical coupling in between the segments, however, which causes a correlation in the energies determined 36 from the segments. This can be accounted for by considering additional energy cali- bration parameters which depend on the coupling between two segments. For events which had interactions in two segments, the calibration included parameters from the cross-terms of the ADC channel outputs 2:1 and 3:2 up to the fourth power. This can be expanded to the events which interact more than twice in the detectors using the technique discussed in Venutrelli et al. [59]. 3.2 Efl'iciency calibration The efliciency of the detectors must also be accurately determined. This is espe- cially important for the measurement Of the partial cross section to the ground state following the knockout reaction. To investigate the 7-ray angular distribution, one also needs to know the relative efficiency Of the two rings of detectors in SeGA. The efficiency depends heavily on the energy of the incoming photon and this energy de- pendence needs to be well understood. The efficiency is also strongly influenced by the geometry of the array relative to the target especially during in-beam experi- ments where the energy and solid angle depend on the relativistic kinematics and in particular the angle Of the detectors with respect to the beam line. TO determine the efficiency, several radioactive sources were placed in the same position as the in-beam target prior to the experiment. For absolute efficiency mea- surement, the radioactivity of the source must be well-known. To. this end, a cali- brated source of 152Eu was used with an activity of 8.46 pCi measured on 1 May 1978. 152Eu has a well-known half-life Of 13.537 :1: .006 years. The intensity of gamma rays, 10, emitted can be calculated giving the efficiency, 5, of the germanium array. 152Eu only provides a good efficiency calibration for energies up to 1.5 MeV. To determine the efficiency above 1.5 MeV, a 56CO source was used which has high energy peaks up to 3.5 MeV. The intensity Of the cobalt source was not as well known, so a scaling 37 llr'Irrrrrtilrtrrtrrrrr'rritriII—rti p 8 p 8 TIIIfUIIIlIl—IrITIII'Illll‘ Absolute Efflclency c 2 IlllljlllllUlllljlllllll 0.02 I 0.01 ‘ A :fi 00 500 1000 1500 2000 2500 3000 3500 Energy (keV) Figure 3.2: SeGA efficiency in the 37° ring (crosses) and the 90° ring (triangles) with associated fits in the form of Eq. 3.3. factor was applied to agree with the efficiencies in the common energy range with 152Eu. The energy dependence was then fit as a curve to the discrete data points to get the continuous energy dependence of the efficiency. Several functional forms were tested producing comparable results. The efficiency was finally taken to be of the form «E; 60, E0. a) = eo(E — E0 + 502693)” (3.3) with the parameters 60, E0,a. determined by minimizing the chi-squared from the fit. For a 1 MeV photopeak emitted from a source at the target position, SeGA had a 2.2% efficiency. The efficiency was also determined for each ring in the array to understand the dependence of the efficiency on different angles of emission. The measured efficiency is displayed in Figure 3.2. The uncertainty in the efficiency was also determined across the range of energies. If the X2 is minimized at some value 38 x3, then the set of parameters included in the no error band are defined as [60, E0, 0. : X2(€0, E0,a) < X3 + 71.2}. (3.4) The minimum and maximum values Of the efficiency for this set of parameters are then determined for each given energy. Since the fit parameters are highly correlated, an evenly spaced mesh over the parameter space was taken with the variation along the principle axes Of the error matrix M. The error matrix for a set Of parameters {c} in defined as: 02x2({6}) (3.5) This allows for a more accurate sampling of the parameter space. However, this method always underestimates the width of an error band for a given energy. The granularity Of the mesh was made progressively finer until there was a negligible change in the width of the error band. The error in the intensity I from a source with N counts in the photOpeak is then accordingly rIIE.) = (/f~(E.)2 + f. 5000 Hz) would impact their performance. During the magnesium production runs, the outgoing residues were identified on an event-by-event basis using the energy lost in the ion chamber at the focal plane Of the 8800 as well as the time of flight from the object to the E 1 scintillator of the 8800. The time of flight was corrected tO improve the resolution based on the kinematic information provided by the CRDCs at the focal plane. The outgoing residues can then be clearly identified as shown in Fig. 4.2. The information provided by the CRDCs also allowed the accurate determination of the outgoing momentum vector of the nuclei relative to the central trajectory of the 8800. The mean velocity Of the residues in the middle of the target was determined to be )8 = 0.404 i- 0.002 by the methods discussed in Subsection 2.4.5. For residues with this velocity, the emitted gamma rays detected by SeGA are Doppler broadened 54 12601280130013201340136013801400142014401460 Time 01‘ flight (arb.) Figure 4.2: Outgoing residues for a reaction of an incoming 32Al beam identified by the 8800 spectrograph based on the energy loss in the ion chamber and their time of flight with selected isotopes of interest labeled. with a FWHM of 48 keV (2.7%) in the downstream ring and 38 keV (2.1%) in the upstream ring for a 1.8 MeV gamma ray. Using the additional information about 1312w outgoing particle momentum from the tracking through the 8800, the gamma-ray resolution can be further improved by about 10%. The major improvement was from determining the velocity of the individual particles relative to the central trajectory Of the 8800. Also, it is important to note that at this velocity the forward ring of SeGA is located at a center—of—mass angle of 54° where P2(cos 00m.) = 0.009(3). This limits any effects from the angular distribution in the forward ring allowing for accurate determination of 7-ray branching ratios from excited states. Higher orders of the angular distribution are assumed to be negligible. 55 Counts per 4 keV ’uI...l.ul.uIUJJ.ul 50100150200503.0035» '400 600 000 1000 1200 1400 1600 1000 2000 EMHKGV) 1III Counts per 4 keV §§§§ .‘t 100150200250300350 400 600 600 1000 1200 1400 1600 1000 2000 EnergyateV) Figure 4.3: Gamma-ray spectrum of 31Mg in the forward (top) and backward (bot- tom) rings Of SeGA; note the marked increase in resolution in the backward ring for the high energy 1.8 MeV gamma ray. 4.2.2 Level scheme Gamma rays emitted from 31Mg were detected in SeGA with the resulting spectrum shown in Fig. 4.3. Detailed information about the gamma rays observed is listed in Table 4.2 where the intensities have been determined using the Observed counts and the Doppler-corrected efficiency which accounts for both the Lorentz boost of the solid angle as well as the Doppler shift in the gamma ray’s energy. Most of these gamma rays were Observed in previous experiments [23, 65,66]. The energies of these transitions agree well with their formerly established values. Furthermore, the rela- tive intensities of gamma rays originating from an initial state are consistent with the 56 Table 4.2: Gamma-ray transitions observed in the current experiment with their intensities, angular distribution coefficients, and Compton scattering asymmetries. E,- (keV) El (keV) E, (keV) 7,81 0.2 A 221.1 50.5 171.1b 77 (4) 0.48 (10) 0.0 221.1b 24.2 (15) -O.8 (3) 673.2 221.1 452.6 (6) 13.2 (18) 50.5 623.3(5) 64(4) 0.02 (20) -004 (5) 0.0 673.2(7) 34.2 (23) 07(3) -0.08 (6) 944.5 50.5 894.4 (13) 9.0 (17) -04 (8) 1154.5 461 692.6(8) 16.8 (14) -04 (5) 007(9) 2014.5 944.5 1072.7(19)C 10.1 (16) 0.3(8) 461 1555.7 (22)0 24 (3) 0.5 (9) 221.1 1793.4 (18) 100 0.29 (18) unplaced 1104.0 (16) 13.0 (19) 1500.1 (24) 13(3) 1707(3) 15(8) 1936(4) 9(4) 1968(4) 17(5) aNormalized to 1793 keV transition Established values since measured values are Subject to effects due to the previously measured lifetime of the state [65] c Tentatively placed in the decay scheme based on energy balance branching ratios seen in previous experiments. This gives credence to the established placement of these gamma rays in the level scheme of 31Mg. However, as the popula- tion mechanism is different in this experiment than the neutron-knockout study [23] and fi-decay studies [65,66], several gamma rays were seen with a significantly higher intensity than in the previous experiments. Information available in the experiment allowed for the placement of one new level at 2015 keV. The placement of this level was established based on background- subtracted ,ny coincidences of the 1793 keV gamma ray with decays from the 221 keV state. The coincidence spectrum is shown in Fig. 4.4. The gates for background subtraction were taken below the 171 keV peak and above the 221 keV peak. In the coincidence spectrum, the 453 keV gamma ray which was previously established to feed the 221 keV state [66] is also readily apparent. Also visible is the Compton 57 Counts per 5 keV 200 400 600 800100012001400160018002000 Energy(keV) Figure 4.4: Background-subtracted gamma-ray spectrum coincident with the known ’7— Iay decays from the 221 keV state. edge of the 1.793 MeV gamma rays which interact once within the detector and tliLen Compton scatter outward without depositing their full energy. The Compton Scattered gamma rays have a minimum scattering energy according to Eq. 2.5 based On the domain of cos 0: E0 E - = ———————. 4.1 mm 1+2E0/mec2 ( ) In the rest frame, the 1.793 MeV gamma ray emitted towards the forward ring has a. Doppler-shifted energy of 2.422 MeV. This corresponds to a minimum energy of 0.231 MeV for the Compton scattered photon, resulting in 2.191 MeV deposited in the detector. When Doppler corrected to the center-of—mass frame, this energy deposit is 1.622 MeV which is where the Compton edge appears in the coincidence spectrum. No other gamma rays were found to be coincident with the decays from the 221 keV state. As the '7-7 coincidence efficiency scales like the singles efficiency squared, only the most intense transitions can provide '7-7 data due to the low efficiency of SeGA (e7 22% implying 67,7 :21 0.04%). Some gamma rays observed seem to have 58 energies which are fairly consistent with the difference in between established energy levels and have been tentatively placed in the decay scheme based on this information. The level scheme of 31Mg with the new information from the present experiment is displayed in Fig. 4.5 including comparisons to the predicted excited states from two theoretical calculations to be discussed in Section 5.1. Gamma rays which deexcite the 221 keV state have a low-energy tail in the Doppler-corrected peak shape (see Fig. 4.10) that is evidence of a long lifetime which is consistent with the previously measured half-life of 133(8) ps [65]. Given the residue velocity, this corresponds to a traversal distance of 1.76(11) cm in the setup. As the efficiency of SeGA depends directly on the position of the source of 'y-ray emission, this contributes an additional systematic uncertainty of about 10% in the intensity which is not quoted in Table 4.2. For the longer-lived 461 keV state, no deexciting 240 keV gamma ray was observed as the 10.5(8) ns half-life corresponds to a long mean flight path of about 3 m which would result in most decays occurring out of the field of view of SeGA. Several other transitions observed in Terry et al. [23] are also visible in the proton knockout experiment. Five of these remain unplaced in the level scheme which ac- count for a total intensity of 67(11)% of the 1.79 MeV gamma ray transition. Several of these are gamma ray doublets which appear above 1 MeV as shown in Fig. 4.6. There is no clear resolution of the doublet structure for the '7-ray peaks located at 1.1 and 1.9 MeV, but the width of the energy peaks suggest multiple gamma rays with similar energies. These peaks at 1.1 MeV and 1.9 MeV have a FWHM of 4.9% and 3.5% respectively which is considerably larger than the other gamma—ray peaks’ char- acteristic 2.1% relative width. This cannot be solely attributed to the lifetime of the states. The observed FWHM is significantly larger than the estimated 0.5% Doppler broadening resulting from variation of velocity in the target for short-lived states. Longer-lived states decay behind the target at the same post-target velocity with a 59 7/2‘ 7/2+ 1/2- 1p2h 7/2- 3p4h 5/2+ 7/2+ 2p3h WM) 5/2+ 293') 92—1441 1p2h 312- 192“ 3/2+ 293h 1/2+ 2p3h AMD+GCM 60 7/2+ 5/2+ 3/2” USD Figure 4.5: (left) The level scheme observed in 31Mg with suggested spins and par- ities including tentatively placed gammas (dashed). (middle) AMD+GCM calcula- tions from Kimura et al. [67] with the 0p1h states in bold. (right) The Ohw states calculated by the USD shell model interaction. Theoretical calculations have a shift in excitation energy to match the energy of the 3/2+ single-hole state determined by the experiment. See Sections 4.2 and 5.1 for details. 300 I I I T I I I I I I I rI I I l r I I l I I I l I I I 250 200 Counts / 4 keV 150 100 " 50 0800 1000 1200 1400 1600 1800 2000 2200 Energy (keV) Figure 4.6: Gamma-ray energy spectrum of 31Mg in the 90° ring of SeGA following Doppler reconstruction, showing several doublets at 1.1, 1.5, and 1.9 MeV. exponential position distribution downstream of the target. This causes a distinctive 7—ray peak shape (such as in Fig. 4.10) as well as ring-dependent energy shifts in the Doppler-corrected spectrum since the angle changes with the position of the decay. Neither of these effects are apparent, so the additional width indicates the existence of unobserved doublets. The energies of the doublet components were established by a fit which assumed a relative width consistent with that for well-resolved transition in the 'y-ray spectrum. 4.2.3 Inclusive and partial cross sections Determination of the knockout reaction cross section requires an accurate measure- ment of the number of 31Mg reaction residues, which in turn is related to the number of incoming 32Al beam particles. In the experiment, there is no direct accurate mea- surement of the incoming beam particles. However since most of the particles pass through the target without interacting, determining the number of these unreacted 61 particles after the target yields an accurate measurement of the incoming rate of 32Al impinging on the target. As the 31Mg reaction residues cannot be measured at the same time as the unreacted particles, two settings of the S800 spectrograph were used. In one setting, the 8800 Bp was chosen such that the unreacted particles were centered in the spectrograph. Since the unreacted particles have a small momentum spread and divergence, unreacted particles are detected at the focal plane of the spectrograph with nearly 100% efficiency. The number of particles detected at the focal plane, Ni, is related to the rate on the scintillator at the extended focal plane of the A1900, prp, and at the object scintillator, NOB J, via: N + N Ni = XFP 2 OBJ f, (42) with some normalization factor f which is related to the purity of the beam and the transmission through the 8800 analysis line. This normalization factor can then be used to determine N,- based on the scintillator rates when the S800 is placed on a different Bp setting. Two normalization runs were taken with the unreacted incoming 32Al particles delivered to the 8800 focal plane. The normalization factor f deviated only slightly in the two runs. For the remainder of the experiment, the S800 was tuned to maximize the accep- tance of the 31Mg reaction products while limiting contamination from the unreacted secondary beam. For a knockout reaction setting, the cross section can then be de- termined using the scintillator rates and the number of residues detected (N f) in the particle identification plot (Fig. 4.2) using Eq. 1.2. An additional correction had to be applied as the reacted particles have a larger momentum spread and suffer accep- tance cuts in the S800 so not all outgoing products are detected. Particles with a low momentum, relative to the central trajectory through the 8800, suffer a reduction in angular acceptance in the dispersive direction as displayed in Figure 4.7. To account 62 du (parts per thousand) Figure 4.7: The focal plane measurements of am and dta showing the acceptance cut for low momentum 31Mg residues from an incoming 32Al beam. for this, a systematic 2.5% correction to the cross section measurements has been made assuming that the angular distribution in the dispersive direction is symmetric with respect to the centroid of the distribution. The inclusive proton knockout cross section was determined to be 8.7(5) millibarn. Throughout the experiment, the efficiencies of the scintillators were monitored for any change due to irradiation damage which could degrade the particle detection efficiency of the S800 spectrograph or the stability of the normalization factor f. The stability of f is further validated by the consistency of the calculated inclusive cross-section on a run-by-run basis (see Fig. 4.8). The systematic uncertainty of 0.5 mb has been determined from the weighted standard deviation of the run—by-run cross sections. This uncertainty is the major constraint on the precision of the measurement, and is related to the consistency of the incoming beam normalization. The inclusive cross section includes the cross section to the ground state (0'0) as well as to excited states (aex). The partial cross sections to excited states were deter- 63 11°C i'r'r'rri'rfi'r'llll ' I ' fir r rI 10.5 - a 4 l 915% 8.5.: E ] iltlllfil l? E 10.0 - 5° 01 1 CO C 1 11 41111 i 8.0 - , 7.5-_ 7.0 - 1 l 31Mg Cross Section (mb) j 114111 6.5 s 60 I'I'I'I'I'IrTrI'I/ILI 'fi'l'l'l Runnumber Figure 4.8: Inclusive cross section of 32Al -—> 31Mg calculated on a run-by-run basis chronologically from left to right. The experiments with the 32Al incoming beam were performed in two sets of runs separated by several days corresponding to the break in the abscissa. mined by particle-gamma coincidences. The cross section to the ground state is the difference between these cross sections to excited states and the inclusive cross sec- tion. The direct population to each excited state I j was determined by the balancing of '7-ray intensities feeding (Iin) and depopulating (Iout) each energy level. Ij = [in _ 1011131 (43) where the intensities are determined from the number N of observed gamma rays as well as the absolute efficiency of SeGA at the given energy E, (i.e. [input = Nin,out/6(E7)). Due to the statistical nature of the sampling of the incoming and outgoing intensities from a level, negative cross sections consistent with zero can be obtained in the analysis. As a negative cross section is unphysical, these small 64 Table 4.3: 31Mg states with direct feeding from the proton knockout from 32Al. E (keV) J7r aexp (mb) 059 (mb) 031,025 (mb) 0.0 1/2+ 0.33 i 0.14 673.2 3/2+ 3.56 :l: 0.20 12.5 10.4 1154.5 0.53 :l: 0.13 2014.7 5/2+ 4.27 :l: 0.24 12.0 8.6 (2787.2)? 7/2+ unobserved 11.7 5.9 negative cross sections were set to be exactly zero. The remaining partial cross sections to excited states were normalized such that their sum was consistent with aex determined by all the assigned gamma ray transitions where 00 is held constant. Predominately, the knockout reaction directly fed states at 673 keV and at 2015 keV. All states fed directly and their partial cross sections are shown in Table 4.3 with the corresponding theoretical calculation of Sec. 5.1. U'nplaced transitions listed in Table 4.2 can modify I j when placed. However the intensity of each unplaced transition is fairly small; the most intense (1968 keV) of these unassigned gamma rays has a maximum contribution of 0.54:1:0.16 mb to a given state. Also, the unplaced transitions do not likely feed or decay from the same excited state. Fragmentation of the feeding intensities suggests that a sizable modification to any single partial cross section in Table 4.3 is unlikely, though possible. If all unplaced transitions feed or decay from the same excited state, this corresponds to a maximum correction of 2.1 mb which is not quoted in the uncertainties of Table 4.3. The tentatively placed decays from the 2015 keV state were included in the partial cross section and could account for an additional 1.1 mb adjustment to the partial cross section to that state. Future work with greater sensitivity is needed to clarify the placement of these transitions. Next generation gamma-ray detector arrays, such as GRETINA [68], with a higher coincidence efficiency would address this situation. 65 l llllllllllllllllIilllllllllllllllllllllT l l 1 .L 1 l l l l l l l 1 l l l I 12.2 12.4 12.6 12.8 13 13. pI (GeV/c) Figure 4.9: (Points) The measured momentum distribution of the 31Mg residues coincident with the 623 keV 'y-ray. (Line) Predicted momentum distribution based on the knockout of a d5/2 proton from 32Al to the 3/2+ 673 keV state in 31Mg including the effects from the loss of energy in the target as well as the incoming momentum width. 4.2.4 Momentum distribution of residues Coincident with the gamma rays detected by SeGA, the momentum distribution of the outgoing particles measured at the S800 focal plane contains information about the dynamics of the reaction. The distribution associated with the residues coincident with the 623 keV gamma ray is shown in Fig. 4.9. The contribution from the ’y-ray background was subtracted based on the momentum distribution coincident with gamma rays from the spectra above and below the transition of interest. The momentum distribution coincident with the 623 keV gamma ray is representative of all momentum distributions observed for the 31Mg residues after the knockout reaction, including those coincident with other gamma rays as well as those events where no gamma ray was detected. The shape of the distribution is an indicator of the orbital angular momentum of the removed nucleon and is consistent with 66 X .5 O u § '13 o 8 III—T1111llflllllllllllfl Intensity per 4 keV 8 § lllllllllllllllllllllllllll III I I- """._ l 1 l L L L I 1 L 1 l l 1 l l l 160 180 200 220 240 Energy (keV) l 1 1 l . l 1 120 140 Figure 4.10: Relative intensity of gamma rays emitted from the 221 keV state in 31Mg as determined by efficiency calibrations (Sec. 3.2) and proper Doppler reconstruction for the forward (dashed) and backward (solid) rings in SeGA. the knockout of an E = 2 proton for the populated states according to the reaction calculations discussed in Subsection 5.1.1. 4.2.5 Angular distribution of gamma rays In the examination of the decay from the excited states, evidence also exists for the presence of spin alignment along the outgoing direction of the knockout reaction residues (i.e. along the beam axis). In particular, the ratio of gamma rays detected in the forward and backward rings is substantially different for the two transitions which deexcite the 221 keV state as displayed in Fig. 4.10. The difference in the ratio between the transition to the ground state versus the transition to the 50 keV state is greater than 40. The angular distribution formula can be expanded in terms of 67 Legendre polynomials (see Eq. A.1). Since there is a noticeable difference in angular distribution, spin alignment must exist in the 221 keV state. From the alignment condition, w(m) = w(—m), a preference in the direction of 'y-ray emission limits the spin of the initial 221 keV state to J 2 3/ 2. However, previous lifetime mea- surements [65] and Weisskopf estimates of the electromagnetic transition rates [1] conclude that these are two pure dipole transitions. Given the known ground state spin of 1/2+ [22] and the selection rules for electromagnetic transitions (Eq. 2.2), the spin of the 221 keV state must be either 1 / 2 or 3/ 2. Combined with the requirements of spin alignment B ,\ 75 0, the 221 keV state must have a spin of 3/2. Further- more, since the spin alignment of the initial state must be the same for both the 171 keV and 221 keV transitions, any difference in their angular distributions must be attributed to differences in the final state. Particularly for pure dipole transitions, A2 < 0 for unstretched transitions (AJ = 0) and A2 > 0 for stretched transitions (AJ = i1) where A2 depends only on the properties of the gamma-ray transition (see Appendix A). This results in a sizable difference in the angular distribution for transitions with different AJ, similar to that observed in the 171 and 221 keV gamma rays. This suggests that the 50 keV state should be an unstretched dipole transition with J f = 3/ 2. This is also supported by the evidence that the 50 keV transition is known to be a dipole transition based on similar lifetime arguments [65]. While there are ample statistics for the angular distribution analysis for these low-lying transitions, the evidence may be affected by the long lifetime of the 221 keV state. Another case was examined which was free from such effects. Though the statistics are not as favorable, the 623 keV transition does not suffer from these issues which obfuscate the interpretation of the transitions from the 221 keV state. The 623 keV 7-ray decays from the 673 keV state which has a short lifetime. The intensity can also be accurately obtained in the spectral fit due to a low background and absence of contamination from the Compton edge of other 31Mg gamma rays. Figure 4.11 68 IIIIIIIlllllllllll'llIIII—r'l 888 Counts per 4 keV lllllllllllllllllllllllllllll4 q I 8 o 550 600 650 700 7 Energy (keV) Figure 4.11: Automatic fitting using ROOT of gamma—ray peaks in the 600 keV region, including determination of the background (thin line) as well as the residue between the experimental data (points) and the fit (thick line). shows the quality of the automatic fitting procedure. In addition, the 673 keV state is directly fed in the knockout reaction making it more amenable to reaction calculations as well as preventing effects from spin dealignment as discussed in App. A. The examination of the evolution of the spin alignment related to the outgoing momentum provides a test of the reaction mechanism. The angular distribution coefficient (12 was extracted from the experimental data over a range of cuts in the longitudinal momentum, as shown in Fig. 4.12. Given a prolate spin alignment in the most central part of the momentum distribution, the data agree with a transition with AJ = 0 including the dipole limitation imposed by the Weisskopf estimates. This suggests an assignment of J = 3/ 2 to the 673 keV state. It is important to note here that the reaction mechanism produces a strongly prolate spin alignment, B ,\ > 0, in the center of the momentum distribution which decreases as more of the central momentum is included. If this qualitatively holds, then the slope alone of the evolution of the 69 f :II'IIIIII—I—I—I—IIIII'IIII'rIII'IrII'II—I—TIIIII 0.2-l-———"-——_~~ _:l -o_ ' .... -o.2'— ' J A 1. -0.4 K V -0.6 .Ijjlti'lii llllllllllllllllllLLlllllLlllllllllll 30 4O 50 60 70 80 90 % central pII -0.8 M- OP Figure 4.12: The angular distribution coefficient a2 determined from the experimental data (triangles) and theory (solid line) as a function of increasingly broader longi- tudinal momentum gates for a 3/ 2—>3/ 2 623 keV pure dipole transition. Theoreti- cal curves are also shown for a 5/ 2—+3/ 2 transition for a dipole (long dashed) and quadrupole (short dashed) multipolarity. The abscissa describes the percentage of central momentum counts included. angular distribution is a good indicator of the sign of the a2 coefficient. This is independent of any systematic errors introduced by uncertainties in the determination of the efficiency of the gamma-ray detectors. Considering the whole momentum distribution includes this uncertainty in normalization of the detectors but also allows the use of all the statistics which makes it the most conclusive for some of the gamma- ray transitions. The 7—ray angular distribution coefficients extracted from the whole residue momentum distribution are quoted in Table 4.2. 4.2.6 Linear polarization of gamma rays With the presence of spin alignment in the system, gamma rays emitted from 31Mg should exhibit linear polarization. Determination of the sign of the linear polarization alone allows for the measurement of the parity of the excited states provided the par- 70 ity of one of the states in the transition is known. The sensitivity of SeGA to linear polarization was discussed in Sec. 3.3 by examining the Compton scattering within the crystal. Polarization measurements benefit from greater consistency in terms of the normalization of the scattering data since the normalization for a ring is only de- pendent on the scattering of unpolarized gamma rays in that ring whereas effects in both rings must be accounted for in angular distribution measurements. In addition, only a relative normalization of the two directions of scattering is needed. Further- more, angular distribution measurements depend on an accurate determination of the efficiency of the array which varies rapidly with energy (see Fig. 3.2) as compared to the Compton scattering asymmetry which varies more slowly (Fig. 3.8). In addition, the relativistic effects play less of a dramatic role. The polarization of gamma rays is also maximized at a center-of-mass angle of 90° which is close to the backward ring in the classic setup of SeGA which has 06,111, = 114° for residues moving at 6 = 0.4. Differences in the Compton scattering (Fig. 4.13) were clearly observed for the peaks in the 600 keV energy region. Note that the 7-ray background in the figure shows no presence of polarization which adds credibility to the in-beam normalization of the scattering. However, due to the low statistics, the linear polarization measurements have significant uncertainty as shown in Table 4.2. 4.2.7 Spin and parity assignments With an accurate measurement of the linear polarization and angular distribution coefficients, the multipolaritie's of the transitions can be determined. The linear po- larizations were extracted using the known 18% relative sensitivity of SeGA discussed in Section 3.3. To reduce the statistical uncertainty, the whole momentum distribu- tion of the reaction products was considered. Combining the 'y-ray multipolarity assignment with the selection rules for knockout reactions (Eq. 1.3) allows for firm spin and parity assignments. In particular, the selection rules imply the states fed 71 > I . . . . , 3 350 Q a 300 O. m 250 u... 5 200 8 150 100 :- 50 ': .a'; u: :"-:'_ WI. 750 800 Energy (keV) Figure 4.13: The Compton scattering of gamma rays relative to the reaction plane defined by the beam axis and the gamma ray propagation direction in 31Mg: (dashed, left axis) scattering perpendicular to the plane, (solid, right axis) scattering parallel to the plane, normalized (by a factor of 0.13) to the intrinsic scattering from a gamma ray emitted by a 152Eu source with no spin alignment directly in the production of 31Mg have a positive parity. The results discussed in Subsec. 4.2.5 allowed the spin assignment of the 673 keV state. Two of the gamma ray transitions from the 673 keV state have sufficient statis- tics to determine 'y-ray Compton scattering asymmetries and angular distributions as shown in Table 4.2. When including the full statistics across the momentum dis- tribution, the 673 keV gamma ray shows evidence of angular distribution. For this case, the linear polarization measurement is not consistent with zero and suggests a mixed M 1 transition with AJ = 1 or a stretched quadrupole transition to the ground state; the comparison to theoretical values is shown in Fig. 4.14 for a final spin of 1/2+. While the measurement is closer to the values for a stretched quadrupole, the momentum-constrained angular distribution analysis for the 623 keV transition (see Fig. 4.12) indicates an initial spin of 3/ 2 for this state. This supports the assignment of a mixed M 1 / E2 AJ = 1 transition with 6 z 2 for the 673 keV gamma ray. The 72 -1.‘- 7-1-11. 1 _1_ ., - ‘ —01.0—0.5 0.0 0.5 1.0 —01.0—0.5 0.0 0.5 1.0 Figure 4.14: (left) Experimental angular distribution coefficient and polarization for 673 keV gamma ray with 10 confidence interval compared to theoretical values of a decay to a final spin state of 1/2+ from an initial 66% prolate-aligned spin state of 3/ 2 (circles) or 5/ 2 (squares) for different multipolarities as indicated ~ lines of M 1 / E2 mixing are also shown; (right) Measurements for 693 keV gamma ray compared to values for a final spin state of 7/2— postulated by Mach et al. [65] assuming a 33% initial prolate alignment for an initial spin of 11/2 (squares) or 7/ 2 (stars). information for the 623 keV gamma ray is less conclusive when including the entire residue momentum distribution. However, the negative scattering asymmetry (and thus negative polarization) is more in agreement with a M 1 transition suggesting J7r = 3/2(+) for the final 50 keV state with the parity remaining tentative due to the substantial uncertainty in the Compton scattering asymmetry. The linear polarizations and angular distribution coefficients for the 693 keV tran— sition also deviate sizably from zero despite the significant uncertainties. Only the final spins which have expected values near the experimental data are shown for clarity in the exclusionary plot in Fig. 4.14 (see Appendix A for further information) which compares the measured angular distribution coefficients and linear polarizations to those calculated by transitions of different multipolarities and initial spins. Here, the spin and parity for the final state are not certain but have been previously postulated to be 7/2‘ [65]. Generally, a2 < 0 and P < 0 agrees best however with unstretched 73 1W 1 or stretched E2 transitions; however, an E 1 multipolarity is not too far removed from the experimental confidence interval. The normalization of the angular distribu- tion can add considerable systematic uncertainty to this analysis. Truly, one would want to investigate the evolution of the linear polarization with momentum gates similar to the angular momentum study. In the current experiment, the statistics are too prohibitive to do such an analysis. 4.3 33Mg 4.3.1 Beam characteristics and particle ID In the case of the 33Mg study, the production cross section for the 34Al secondary beam was significantly reduced. To achieve enough statistics within the allocated time, a thicker secondary reaction target of 2 mm beryllium (370 mg/cmz) was used at the target position of the 8800. The purity of the secondary beam was also signif- icantly lower (50%) with the major contaminants being 35Si and 37P. Similar to the study of 31Mg, the components of the secondary beam can be cleanly selected based on the time of flight from the A1900 to the object box of the S800 as displayed in Fig. 4.15. The reaction products shown in Fig. 4.16 were then identified in the S800 using the same method as in Section 4.2 where the unreacted secondary beam was removed in the S800 by Bp selection. Two normalization runs with the unreacted beam transmitted to the focal plane of the S800 were taken with a lower intensity to establish the properties of the incoming beam. For the 33Mg residues, the mean mid-target velocity was found to be [3 = 0.402 :1: 0.002. At this velocity, the forward ring of SeGA lies at a center-of-mass angle of 54° limiting the effects of angular distribution given P2 (cos 00m) = 0.011(3). With a thicker target than for the 31Mg experiment, the energy resolution of the gamma ray detection was worse. An in-beam gamma-ray resolution of 2.7% relative to the 7—ray 74 £1400r--................ S 01200 1000 800 600 400 llllllllllllllllllllllll 200 [IIIllIl'Ill'lllllllllIlllll l I 1 1 . I 900 950 1000 1050 oo 01 C Time (arb. units) Figure 4.15: Incoming 34Al particles identified by the time of flight from the ex- tended focal plane of the A1900 magnetic separator to the object box of the S800 spectrograph. energy was obtained for both rings. Corrections based on the S800 information allow a better determination of the velocity and improve the relative resolution to 2.5% in the backward ring. The improved resolution in the backward ring was used for determination of all energies of the gamma ray transitions observed in the experiment. 4.3.2 Level scheme The gamma ray spectrum resulting from the knockout reaction to 33Mg is displayed in Fig. 4.17. Information about the transitions observed in the experiment is summa- rized in Table 4.4. None of the transitions show any indication of lifetime effects as discussed in Subsec. 4.2.2 which suggests these states are all short-lived so the decay 75 AE (arb.) 1400 1420 1440 “me of flight (arb.) Figure 4.16: Outgoing residues for the reactions of an incoming 34Al beam identified by the S800 spectrograph based on the energy loss in the ion chamber and their time of flight with selected isotopes of interest labeled. Table 4.4: Gamma-ray transitions observed for 33Mg in the current experiment with their intensities normalized to the 483 keV transition as well as their angular distri- bution coefficients and Compton scattering asymmetries. E; (keV) a(keV) E7 7 0.2 A 484.1 0.0 483.1(9) 100 0.36 (17) 003(8) 705.0 484.1 219.7 (10) 1.8 (7) 0.0 704.4 (14) 6.7(12) 0.9 (8) 780.4 484.1 297.3(6) 57(3) 0.33 (21) 008(9) 0.0 780.7 (18) 15.2 (19) 1242.4 484.1 758.1 (16) 12.6 (18) 0.1(7) 0 0 1240 (3) 16 (3) unplaced 1069.6 (24) 5.0 (17) 1523(4) 12(3) 1856(5) 7(4) 1929(6) 7(3) 76 > I I I I I I I I I I I I I I I I I r I r £4 1. I I l l l I . N 3411 a .1 .235 3 0311. 2 1 211 [1’0 1 1 '1 111 'l I ‘ i 1 '10 ' 2" 1‘ 1. , - 11 1111111111, 200 400 600 800 1000 1200 1400 1600 1800 ' Energy (keV) > I I I I rI FI rI I I l I I I l I l I ' I I I I I r s . .. 1 N 33 . .. .9 5 §311 -: o 02 . I] I 75 211 1 1 1’1 1 ' '3 15 111' , 1 l 10 ' w '5 1 , 1 ., 1 ff “I‘ll-”Jud 1.1L] IIIHIU'J . 200 400 600 800 1000 1200 1400 1600 1800 Energy (keV) Figure 4.17: Gamma-ray spectrum of 33Mg detected in SeGA for the forward (top) and backward (bottom) rings. 77 occurs near the target (7' S, 10 ps). None of the levels established in previous exper- iments have any information about their lifetimes. The energies of these transitions in general agree with those observed in an earlier fi-decay experiment [69] as well as other recent measurements [70,71]. The ’y-ray branching ratios from the 1242.4 keV state also agree, within the uncertainties, with those in the fi-decay experiment. The decays from the 705 keV state warrant additional discussion. The weak 219.7 keV transition could only be clearly observed when summing the spectra from both rings of SeGA. Motivated by the previous branching ratio from the fl-decay experi- ment [69] and the number of counts observed for the 704 keV gamma ray, the expected counts in the two rings would be 192 d: 22 counts assuming isotropic angular distribu- tion. On top of the large background (S/N z 0.1), a small peak is indeed observed at the right energy with a total of 135 d: 54 counts which is within the margin of error of the previous experiment. However, the fi-decay experiment also tentatively placed a 546.2 keV transition emanating from the 705.0 keV state. This had a sig- nificantly higher intensity observed following fl-decay, and while approximately 300 counts would be expected in the 33Mg gamma-ray spectrum, no evidence of a peak is observed. A gamma ray consistent with the energy of 546.2 keV was also seen in the proton inelastic scattering experiment [71] without observing the other gamma rays associated with the 705 keV state in the fl-decay experiment. This suggests this gamma ray does not originate from the 705 keV state, but most likely feeds the ground state directly from an unassigned level of 546 keV. A new level could also be placed in the 33Mg level scheme based on the information provided by 7-7 coincidences. The 483 keV decay was seen to be in clear coincidence with the 297 keV gamma ray (see Fig. 4.18) allowing for the placement of a level at 780 keV. The coincidence is also seen taking the reverse gate on the 483 keV transition. Incidentally, a 780.7 keV gamma ray is also observed for the first time in the current experiment. This transition has an energy which agrees with the decay 78 I l I III I I I I I I l I I I I l I I I I Counts per 5 keV 88888 d O 0 -Im 0 500 1000 1500 2000 2500 Energy (keV) Figure 4.18: Background-subtracted gamma-ray spectrum coincident with the 297 keV gamma-ray transition in 33Mg. of the newly placed state to the ground state and has been placed in the level scheme accordingly. Theoretically, the established 758 keV decay from the 1242 keV state that feeds the 484 keV state should also be observed in the 7-7 coincidence measurement. However, given the intensity in the singles spectra and the efliciency of the detectors, the coincidence spectrum would only show approximately 23 counts for the 758 keV coincident decay which is within the background fluctuations. With the additional information from the 74y coincidences, the level scheme observed in this experiment can be found in Fig. 4.19. Several other transitions seen in the proton knockout experiment remain unplaced in the level scheme. These five transitions mentioned in Table 4.4 have a total intensity which is 31(6)% of the 483 keV gamma ray transition. These are mostly high energy transitions which cannot be placed due to the lack of coincidence information. 79 1242 (5/2—) 7’8 781 (5/2+) [705 l . ' 219 1.40 297 (5/2-) 1 w 1484 714 731 4 3 (313 r v v 0 Figure 4.19: 33Mg level scheme observed in the present experiment with suggested spin and parity assignments. 4.3.3 Inclusive and partial cross sections The inclusive cross section was determined similar to the method established in Sec- tion 4.2. However, during the experiment, a sudden change in the transmission was observed following a retuning of the primary beam. The normalization of the incom— ing beam was based on the rate on the object scintillator in order to provide the greatest consistency in the extracted cross section on on a run-by-run basis. The normalization factor was determined from two different runs with the S800 Bp tuned to select the unreacted incoming 34Al particles in the spectrograph. Fluctuations in this normalization add an additional 5% systematic uncertainty in the cross section measurement. The width of the momentum distribution once again led to a cut in the angular acceptance in the dispersive direction (Fig. 4.20) which accounts for another 2.5% of systematic uncertainty to the cross section measurements. The inclusive proton knockout cross section was determined to be 4.32 millibarns. For the 33Mg case, the additional systematic uncertainties from the normalization and acceptance determination are much more significant. The cross section on a run— 80 dtil (parts per thousand) -100 -80 ~60 -40 -2o 0 2o 40 60 80 100 au(mrad) Figure 4.20: Focal plane measurements of am and dta showing the acceptance cut for low momentum 33Mg residues from an incoming 34Al beam. by-run basis is shown in Fig. 4.21. The variation in the cross—section on a run—by-run basis adds a systematic uncertainty of 2.5% giving a total systematic uncertainty in the measurement of 0.4 millibarns. Here, the inclusive cross section is significantly decreased in comparison to the 31 Mg case. As the neutron separation threshold in 33Mg is 0.2 MeV lower, there is an increased possibility that highly—excited states populated in the reaction decay rapidly by neutron emission and thus do not make it to the focal plane of the S800 which could account for the decreased cross section. The cross sections to the excited states were extracted from the particle-gamma information along with the balancing of the gamma-ray intensities. The partial cross section to the ground state was determined by subtracting all the excited state par- tial cross sections from the inclusive cross section. The results are summarized in Table 4.5. There was some feeding to all the states which emitted gamma rays though most of the feeding populated the newly discovered 780.4 keV state. Unas- Slgned gamma transitions amount to a maximal adjustment to the cross section of 81 5.0 - I l I l I l I l I I I l I I q 4.9 - .4 4.8 - fl - 4.7 - -< A 4.6 - -1 .0 J «1 g 4-5" + -1 g 4.4 4 J :8 J J o 4'31 '3 ‘0 4.2 - — 0 4.0 _ .. a) . . 2 3.9 - .. co . , m 3.8 "' —( 3.7 - _ 3.6 - _. 3'5 I I —I— I I I I I I I I I I I 60 80 100 120 140 160 180 200 Run Number Figure 4.21: Inclusive cross section of 34Al —> 33Mg calculated on a run-by—run basis chronologically from left to right. Table 4.5: 33Mg states with direct feeding from the proton knockout from 34Al. E (keV) J7r aexp (mb) 0.0 (3/2+) 0.42 i 0.16 484.1 (3/2‘) 0.79 :l: 0.13 705.0 (5/2+) 0.24 i 0.04 ( ( 780.4 3/2") 20510.12 1242.4 1/2+) 0.811009 82 Counts per 61 MeV/c §§§§§§§ 0| O O ’ 13 13.2 13.4 13.6 13.8 pI (GeV/c) Figure 4.22: (Points) The measured momentum distribution coincident with the 297 keV gamma ray, (Line) Predicted momentum distribution based on the removal of a d5/2 proton from 34Al including effects from the loss of energy in the target as well as the incoming momentum spread. 0.88 :l: 0.17 mb. Though this likely will not highly impact any individual partial cross section, it removes certainty of the feeding of the weakly populated ground state and 705.0 keV state in the knockout reaction. 4.3.4 Momentum distribution of residues The longitudinal momentum distribution associated with the gamma rays is similar to the situation in 31Mg, which suggests the removal of a g = 2 proton from the incoming 34Al isotopes. However the experimental width is broadened due to the increased thickness of the target and cannot be well—described with typical input parameters for the reaction. The background-subtracted momentum distribution coincident with the observed 297 keV gamma ray is shown in Fig. 4.22. This is once again representative of the momentum distributions of all the states directly fed in the reaction, including the particle singles detected in the S800. 83 - I u q - q I I —1 q q I I —] 1 q u _. d1: 1 0.8 0.6 0.4 0.2 h——’——H h——>——« 02 04 -0.6 -0.8 n u IIIIIIIIIIIIIIII IIIIIII'III'IIIIII 4 —C}—< 1—0—1 -—>— F——{F——4 1—{3—1 —Cl—1 H—{}—H h—{}—d h——CF—« llllllllllllllll llllllllllllllllllf WW. 0 20 30 40 50 60 70 80 90 100 % central pll cut —L I Figure 4.23: Angular distribution coefficients extracted for 33Mg as a function of increasingly broader momentum gates for the 297 keV transition (squares) and the 483 keV transition (triangles). 4.3.5 Angular distribution of gamma rays The gamma rays emitted by the 33Mg nuclei were also examined for the presence of angular distribution. Only the two most intense gamma rays in the spectrum had enough statistics to perform a significant analysis. Both of the peaks were clearly separated enough to perform a systematic investigation based on the gating of the longitudinal momentum distribution shown in Fig. 4.23. Only the 297 keV shows strong evidence of variation in the momentum-constrained 7—ray angular distribution. It is important to reiterate here that based on a prolate alignment in the center of the momentum distribution the slope as a function of broader momentum gates is directly related to the Sign and magnitude of the angular distribution coefficient. This is independent of any systematic error in the absolute normalization. As the calculated angular distribution coefficient for the 483 keV transition is 84 relatively flat as more momentum is included, it is suggestive of an a2 coefficient close to zero. However, the coefficient extracted from the gamma-ray intensity and efficiency is significantly positive (20) which should also appear in the slope of the momentum-gated angular distribution analysis. This provides an estimate of the sys- tematic uncertainty in the absolute normalization of the gamma-ray intensities in the rings of roughly 10% based on the deviation from a2 z 0 and the difference between the P2(cos 6cm.) in the two rings. The uncertainty in the absolute normalization is avoided in the momentum-constrained analysis. For the low-energy 297 keV gamma ray with no evidence of lifetime effects, the Weisskopf estimates limit the transition to dipole multipolarity. The evolution of this transition clearly resembles the trend observed for the 623 keV transition in 31Mg discussed in Section 4.2. The trend combined with the Weisskopf estimates suggests that the 297 keV gamma ray corresponds to an unstretched dipole transition. On the other hand, the 483 keV transition shows no evidence of a variation in the momentum- gated angular distribution. This suggests one of two possibilities. The nuclear spin alignment has been reduced by the gamma rays which feed the state from above, or the properties of the transition are such that no spin alignment can be observed. Since the feeding states likely have significant prolate alignment in the most central part of the momentum distribution, some degree of alignment is expected as well in the fed 484 keV state. This seems to make the second interpretation more likely where the lack of angular distribution is a property of the 'y-ray transition. This means either the 484 keV state could be a J = 1 / 2 with no possible spin alignment or the transition is such that there is only a small effect in the angular distribution such as a AJ = 1 strongly mixed transition originating from a high spin (e.g. 7/ 2 —+ 5/2). Here the significant deformation of the 33Mg nucleus [70,71] can create a sizable enhancement of the E2 transition probability which would explain this strong mixing for the low lying 484 keV transition. 85 Counts per 4 keV 250 300 350 400 ‘450 500 550 600 Energy(keV) Figure 4.24: The Compton scattering of gamma rays relative to the reaction plane defined by the beam axis and the gamma ray propagation direction in 3P’Mg: (dashed, left axis) scattering perpendicular to the plane, (solid, right axis) scattering parallel to the plane, normalized (by a factor of 0.13) to the intrinsic scattering from a gamma ray emitted by a 152Eu source with no spin alignment. 4.3.6 Linear polarization of gamma rays Information was also extracted about the linear polarization of the gamma rays emit- ted by the reaction residues. Similar to the case of 31Mg, the backward ring lies at 0cm, = 114° given the measured velocity of the particles. This is close to where the linear polarization is maximized, however with the reduced statistics the evidence is less clear in the 33Mg case. Only the two low energy transitions have enough statistics to be sensitive to the *y-ray linear polarization. The resultant spectrum is shown in Fig. 4.24. Though the asymmetry looks quite large, this is reduced by the increased widths of the gamma rays which scatter parallel to the plane. The decreased res- olution of the parallel scattering 'y-rays is likely attributed to small inaccuracies in the Doppler reconstruction. Perpendicular scattering is completed confined within a longitudinal slice of the SeGA detector (see Fig. 3.7) which constrains the angle more 86 precisely than in the parallel scattering case. With the low statistics, the uncertainties in the measured asymmetries are sizable. For the 297 keV and 483 keV transitions, the Compton scattering asymmetries are A = —0.08 :l: 0.09 and A = —0.03 :1: 0.08 respectively. With significantly more statistics, one could investigate the evolution of the asymmetries as a function of longitudinal momentum; however that remains intractable here. 4.3.7 Spin and parity assignments Through analysis of the linear polarization and angular distribution of the gamma rays, the transition multipolarities can be established. With sufficient precision, some knowledge about the mixing ratios as well as the actual final and initial spins can be determined. However, often experiments must rely on prior knowledge about either the initial or final spin and parity of the transition. In 33Mg unfortunately, there are no firm previous assignments of spin and parity. Furthermore, the range of spins allowed to be populated by the knockout selection rules covers a wide range from 3 / 2" to 13/2— based on the removal of a d5/2 proton from the ground state of 34A1 with known J"r = 4‘ [72]. Given the uncertainties in the current experiment, strong support cannot be given to a given multipolarity; however, the polarization and angular distribution measurements suggest certain mulitpolarities, while excluding others. For the 484 keV transition which feeds the ground state, the positive a2 coeffi- cient supports a stretched dipole transition. The polarization measurement does not constrain this to electric or magnetic character, but would suggest a strong M 1 / E2 mixing as shown in Fig. 4.25 which compares the measured angular distribution coef- ficients and linear polarization with predictions for different multipolarities and initial Spins. Previous Coulomb excitation work [70] indeed suggests that this transition has a considerable E2 contribution with an initial spin of 7/2+ assuming a ground 87 1 O P I— T Y T I O i 4 0.5: 612 0.0: —0.5: —0.5 _1. 9-4-44-4—444-4444H -— _1. g» -01.0 -0.5 0.0 0.5 1.0 -01.0 Figure 4.25: (left) Experimental angular distribution coeflicient and polarization for 484 keV gamma ray with error bars and 10 confidence interval compared to theoretical values of a decay to a final spin state of 5 / 2+ (from 'Pritychenko et a1. [70]) from an initial 33% prolate—aligned spin state of 7/ 2 (empty circles) or 3/ 2 (filled circles) for different multipolarities as indicated — lines of M 1 / E2 mixing are also shown; (right) Measurements for 297 keV gamma ray compared to values for a final spin state of 7 / 2"" assuming a 66% initial prolate alignment for an initial spin of 9/ 2 (empty circles) or 5/ 2 (circles). state spin of 5 / 2+. The current experiment agrees with this interpretation, especially if one considers the possibility of a systematic uncertainty in the angular distribution measurement discussed in Subsec. 4.3.5. As the measurement only determines the spin and parity differences, this could also be interpreted as a 7/2‘ -+ 5/2" transi- tion which agrees better with the presence of direct feeding in the knockout reaction shown in Table 4.5. However, a possibility that this partial cross section could be affected by feeding from unplaced and/ or unresolved transitions cannot be excluded. For the 297 keV transition from the strongly populated 780 keV state, the data once again suggest a AJ = 1 transition which likely also has considerable mixing. Here the final spin state has been assumed to have J 7' = 7/2(_). From the calculated points in Fig. 4.25, this seems to support a similar situation as the 484 keV gamma MY; here with initial J7r = 9/2‘ favored slightly. Here, the knockout selection rules 88 clearly constrain the parity of the initial state. The spin assignment depends on the assignment of the lower-lying states, but a AJ = 1 transition remains the most likely. Further discussion regarding other possibilities considered in previous work will follow in Sec. 5.2. 89 Chapter 5 Commentary 5.1 31Mg 5.1.1 Calculation of momentum distribution The momentum distribution of the fragments following the proton removal shown in Figure 4.9 can be described well in a three-body reaction model based on the eikonal and sudden approximation [27,73]. For the standard shell ordering for the protons in the incoming 32A] nucleus with Z = 13, the valence proton would occupy the d5/2 orbital, which is consistent with the momentum distribution of the residues observed in the reactions. The calculations considered this valence d5/2 proton outside 0f the J7r core of the 31Mg residue in its excited state. The core and the valence PI‘Oton were coupled to the known ground state spin of 32Al [14], J7r = 1+. It is 1.mlDOrtant to note that the incoming 32Al fragments were assumed to impinge upon the Secondary fragmentation target while they were in their ground state. An isomeric State with J 7‘ = 4+ does exist in 32A1 with a half-life of 200 ns, which was previously Obse 1‘Ved by Robinson et al. [74]. However, the population of this isomer from the fragmentation reaction would have decayed considerably over the ~80 m flight path which corresponds to roughly four half-lives. Based on the selection rules (Eq. 1.3), 90 direct removal of a d5/2 proton from the 32Al nucleus would populate states with J 7' = (3/ 2,5/ 2, 7/ 2)+ in the residues. Population of the 1/2+ ground state is small since it could only proceed by the removal of a deeply bound 31/2 proton or a d3 /2 proton which is an unfilled orbital with standard shell ordering. 5.1.2 Structure of observed excited states The structure of the excited states was interpreted in the framework of several theo- retical models. A calculation using the USD shell model interaction was used for the analysis of the UpIh states. As the USD interaction is truncated above the N = 20 shell gap, its application to states with excitations across the shell gap is intractable. Since the ground state is dominated by a 2p3h configuration, the relative excitation energies of the 039111 states cannot be addressed by the calculations. To assess the intruder states, the technique using AMD+GCM [67] employing the Gogny DIS in- teraction [75] was used. The excitation energies in both calculations were shifted to reproduce the observed energy of the 0p1h 3 / 2;“ in Fig. 4.5. As the current work pre- dominately populated these OpIh states, the theoretical predictions related to these states are the main subject of the commentary. In the two calculations, there is a significant difference in the energy spacing of the 0p1h states due to the different approach used in determining the effective interaction. The levels at 673 keV and 2015 keV present themselves as candidates for the 017111 3/2+ and 5/2+ states respectively, which are the two lowest 0p1h states predicted theoretically [67]. The reaction preferentially populates these states since 32Al has been established to lie outside of the “island of inversion”. The ground state of 32Al is well described by the sd—shell with its 1+ ground state having a dominant configuration of 7761572 (El/d; /12 [76]. The theoretical partial cross sections to individual States shown in Table 4.3 have spectroscopic factors calculated from the sci-shell W8'erfunctions for the ground state of 32Al and for the excited 0p1h states 31Mg 91 labelled in Fig. 4.5. The spectroscopic factors for the 0p] h ground state in 32Al to any states in 31 Mg with population in the f p—shell is considered to be zero. The inclusive cross section of 8.7(5) mb has a reduction of spectroscopic strength of R3 = 046(2) which agrees well with previous trends for removal of a deeply-bound nucleon [77] characterized by the difference between proton and neutron separation energies in 32Al, AS = 11 MeV [78]. If the tentatively placed 'y-ray transitions discussed in Chapter 4 are disregarded, this would reduce the partial cross section to the 2015 keV state from 4.27 to 3.33 mb. This would agree better with the theoretical relative cross sections assuming a fixed reduction of spectroscopic strength. The cross section to the expected 0p1h 7/2+ state is calculated to be aspC2S = 5.9 mb. While the 2015 keV state could be attributed to this 7/2+ state, the data argue strongly against it. With a reduction factor consistent with that of the inclusive cross section, the expected experimental cross section would be 031,023 R3 = 2.7 mb which is significantly lower than the observed partial cross section. Furthermore, this would leave no evidence for the 5/2+ state which both calculations predict to lie lower in energy and which has a higher expected cross section from the reaction calculations. The 2015 keV state also predominately decays to a state which has J7r = 3/2(—) with the parity taken from the level ordering of Kimura et al. [67] which seems to agree with previous experiments [65, 66]. The multipolarity of the 7-ray transition from the 2015 keV state would then be significantly hindered based on the selection rules for a J7r = 7/2+ assignment. The Weisskopf estimate for a 7/2+ —> 3/2— M2 transition would be 170 ps which would exhibit lifetime effects similar to that seen in the lineshape for the 221 keV state which has a 133 ps half-life. Even considering Some enhancement of the M2 transition using the recommended upper limit for the tr anSition rate of 3 Weisskopf units [41], the spectral lineshape would still be affected by t he lifetime. These three factors argue against a J = 7/ 2 assignment for the 2015 keV State and support the interpretation of this state as the J = 5 / 2 single-hole state. 92 1.0; A 0.57 N . R 1 Q . Q 0.0]w 3/ 2 M 1 / E2 transition; (dashed) The angular distribution coefficient for an unmixed 7/ 2 —> 3/ 2 E2 transition. Angular distribution data would corroborate this assignment, but the experiment was not sensitive enough due to possible quadrupole mixing in the transition (see Fig. 5.1). Mixing ratios as small as 6 ~ 0.2 render sizable effects on the angular distribution as illustrated in Fig. 5.1. With the 2015 keV assigned the J = 5/2+ single-hole state, this leaves no firm observation of the 0p1h 7/2+ state. The USD calculation predicts the OpIh 7/2+ state to lie 2.11 MeV above the 3/2+ 0p1h state. This would place it at an excitation energy 2.78 MeV which would lie above the measured neutron Separation energy of 2.38 MeV [78]. This explains the non-observance of the 7/2+ as 1' t would decay by neutron emission, supporting the energy spacing between 0p1h States predicted by the USD interaction. However, unplaced 'y-ray transitions could have a significant intensity to suggest decay from a higher lying unassigned 7/2+ State - Placing these transitions in the level scheme would clarify the situation. The USD calculations are incapable of producing information about the states wlth intruder-like configurations. The other excited states were thus interpreted in 93 the framework of the AMD+GCM calculations [67]. Information from previous ex— periments [23,65,66, 79,80] also provide an important insight into the nature of these states. In particular, the neutron-knockout experiment [23] populated the 3/2‘ and ' 7/2‘ states located at 221 and 481 keV respectively. These are labeled as 1p2h states in the AMD+GCM calculation which agrees well with the removal of a f p— shell neutron from the known 2p2h ground state of 32Mg. These 1p2h states are populated in the present experiment through the decay of the Uplh states with no excitations across the shell gap. The other populated states in Fig. 4.5 are most likely higher lying states in the 2p3h rotational band based on the level ordering in the AMD+GCM calculation. There is no clear evidence for the population of the low-lying state (J7r = 3/2; ) of the highly deformed 3p4h rotational band which has a calculated deformation parameter ,8 ~ 0.6 [67]. This suggests that the 1155 keV state is more likely to correspond to the 7/2+ state in the 2p3h rotational band. 5.1.3 Spin alignment calculation and angular distribution The nuclear spin alignment along the beam axis was compared to calculations using the reaction theory discussed in Sec. 1.4. Following the knockout reaction, the par- tial cross section to each individual 31Mg magnetic substate m was calculated with the component of the total angular momentum along the beam axis J; = m. The Population of the substates was compared across the entire range of the longitudinal momentum. Incoming 32Al fragments were assumed to have no initial spin align- ment. A previous experiment with fragmentation showed no sizable alignment in the r‘381'd 11a] nucleus 12B after many nucleons were removed from the 22Ne projectile [81]. In tl'le fragmentation reaction which produced 32A], 16 nucleons were stripped off; this leaves little expectation for a non-zero initial spin alignment in the projectile. Fun; l71ermore, any spin alignment that would have been produced would have had tlme to relax during the time of flight to the secondary fragmentation target. 94 Alignment of the residues was calculated following the proton knockout reaction of the incoming 32Al fragments on a secondary reaction target which preferentially pop- ulated certain magnetic substates in 31Mg. The relative population of the magnetic substates defined the spin alignment according to Eq. A.3. Based on this alignment, the angular distribution of gamma rays relative to the alignment direction was calcu- lated. The predicted 'y-ray angular distribution for the 623 keV transition is consistent with the data from the experiment, as displayed in Fig. 4.12, with regards to a range of longitudinal momentum cuts. Since the gamma ray is emitted promptly (on the order of picoseconds) after the reaction, effects which dealign the spins are calculated to be negligible. In addition, higher orders of alignment were calculated to be negli- gible in the reaction, i.e. B4 << Bg. The spin alignment was also calculated for the higher-lying 0p1h states predicted by the shell model. This includes the 5 / 2+ state at 2015 keV which has a significant feeding in the reaction. For different spin states in the 31Mg nucleus, the theory predicts a comparable amount of spin alignment B A in the residue. At the peak of the momentum distribution, 82 z 0.5 is predicted. The angular distribution effect is significantly reduced to 82 a: 0.2 when considering the whole momentum distribution. This amount of alignment is sufficient to determine spins and parities accurately from the resulting 7-ray angular distribution and linear polarization, especially if an experimental setup that is more optimized for the task is utilized. 5.2 33Mg 5-2 - 1 Calculation of the momentum distribution The momentum distribution was calculated in a similar manner as in 31Mg. The Corn Darison to the measured experimental values is shown in Fig. 4.22. The width of the distribution was somewhat underpredicted using typical input parameters for the 95 I l I I I I II I r I I I I I F I l I 0.03 do/dpl (arb. units) 2 O b 9 ' a S 01’ IIII'IIII'IIIIIIIIIIIIIIIIIIIII 0.01 0.005 lllllllllllllllllllllllllllllll . I . 0 -400 -200 o 200 400 ApI (MeV/c) Figure 5.2: The theoretical momentum distribution of 33Mg after the knockout reac- tion relative to the central momentum before (thin) and after (thick) the convolution of experimental effects. calculation. The calculation was convoluted with the effects from the energy loss in the target and the momentum spread of the incoming beam. Due to the thicker target, the experimental effects modified the distribution as shown in Fig. 5.2. However, the effect of the target remained small compared to the momentum spread introduced by the knockout reaction. 5-2 -2 Structure of populated states The discussion of the structure of populated states in the present experiment relies on the interpretation of previous experiments. In particular, the spin assignment of the ground state in 33Mg is a subject of significant debate. Also, the extent of the isl . . . . . . . . . and of 1nvers10n and 1ts behav10r has undergone substantial reVISion Since it was 96 originally postulated and observed. The ground state spin and parity assignment in 33Mg has been previously probed using several different techniques. Beta-decay experiments [69] provided the first information about the configuration of the ground state. This experimental interpre- tation is dependent on the information about the parent’s ground state, which has been assumed based on systematics of Na isotopes. For the ground state of 33Na, the spin is determined by the odd unpaired d5/2 proton which occupies the Nilsson orbital with Q" = 3/2+ for prolate deformations using typical values of the nuclear potential [82]. For the ,B-decay experiment, the possibility of a 5/2+ ground state spin was also considered based on the proximity of energy levels using shell-model calculations [83] with a reduction of the gap between the 1f7/2 and 2193/2 neutron orbitals [84] resulting in a 0p0h ground state in 33Na. The 6 decay was seen to have sizable feeding of 202t10% to the 33Mg ground state suggesting a Gamow-Teller allowed transition where a d3 /2 neutron decays into a d5/2 proton populating positive- parity 1p1h states in 33Mg with spins from 1 / 2 to 7/ 2. The associated shell-model calculations favor a 3/2+ ground state for 33Mg. Meanwhile, a Coulomb excitation experiment [70] suggested that the 484 keV transition had E2 character with AJ = 1. When considering the observed feeding in the fi-decay experiment, the ground state spin had to be considered to be 5/2+ to explain the experimental data which was later corroborated by a proton inelastic scattering measurement [71]. However, a re- cent measurement [21] instead suggests that the ground state of 33Mg is dominated by a 2p2h configuration with J7r = 3/2— bringing into question the interpretation of the previous results. Further discussion follows the assumption that this recent measurement is accurate. The interpretation of the states pOpulated by the current knockout study is af- fected by the more complicated structure of the incoming 34A] isotope which lies on the border of the island of inversion. To reproduce the ground state spin and parity 97 of 4‘ measured using fi-N MR [72], the neutrons in the ground state must have a configuration which includes excitations across the N = 20 shell gap. In fact, the 2p2h component of the ground state needs to have a magnitude of at least 50% for the measured magnetic moment to explain the observations. In addition, the inter- pretation of Himpe et al. [72] needed to reduce the single-particle energy of the 113/2 orbital by 500 keV to describe the magnitude of the magnetic moment. This results in significant mixing of 7r(sd)up3/2 components in the ground state wave function. Coming from a 34Al nucleus where there are three active neutron orbitals (d3/2,f7/2, and (93/2), the neutron configuration in the resulting 33Mg nucleus after the knockout also should show a complicated structure of states. In fact, it is seen in Table 4.5 that the reaction feeds a number of low-lying states, which could each have their own unique neutron configuration. However, the spin and parity are still clearly limited by the knockout selection rules (Eq.1.3). The most strongly populated state at 780 keV has a significant cross section beyond the maximal uncertainty based on unassigned transitions. As the 2p2h con- figurations in 33Mg likely have a considerable number of low-lying states, the cross section to these states would be considerably fragmented, especially if contributions from the V193 /2 orbital are significant in the ground state of 34A]. This makes the 780 keV state the most likely candidate for the low-lying 0p0h in 33Mg which has negative parity based on the knockout selection rules. To make a clear spin assign- ment, one must utilize the information about the states which the 780 keV state feeds through gamma-ray transitions. A prompt 780 keV transition feeds the ground state of 33Mg which is known to have J7r = 3/2' from the fl-NMR experiment utilizing laser spectroscopy [21]. The Weisskopf transition rate for the 780 keV 'y ray limits the spin of the 780 keV state from 1 / 2 to 7/ 2. Knockout selection rules reject the spin assignment of 1 / 2. The 297 keV gamma ray to the 484 keV state is indicative of an unstretched dipole transition based on the momentum—constrained angular distri- 98 bution analysis in Subsec. 4.3.5. The previous Coulomb excitation work [70] requires that the 484 keV state decay by a AJ = 1 quadrupole transition to the ground state which is consistent with the current experiment as discussed in Subsec. 4.3.7. Given the measured 3/2‘ ground state, this suggests that both the 484 keV and 780 keV state have J 7r 2 5 / 2‘. This assignment agrees with the previous experimental data excluding that from B decay [69]. The fi-decay information however makes assump— tions based on the spin and parity of the parent 33N a. No direct measurements of 33Na have been performed to date. Also, recent measurements suggest a larger extent of the island of inversion [24] bringing into question the assumption that the parent has a 0p0h configuration as assumed in Numella et al. [69]. In a spherical shell model with regular ordering, the last neutron in 33Mg would fill the f7/2 single-particle orbital. A deformed shell model [85] tailored to reproduce the 2p2h 3/2‘ ground state of 33Mg shows that the unpaired valence nucleon would occupy the [330 1 / 2] Nilsson orbital for moderate deformations (6 S, 0.3) as seen in Fig. 5.3 with Q“ = 1/2‘. There is no clear evidence of this 1/2‘ state based on the knockout data. This certainly cannot be attributed to the 780 keV state as this is inconsistent With the observed angular distribution of the emitted 297 keV gamma rays as discussed in Subsec. 4.3.5 as well as being forbidden by the knockout selection rules. It is suggested that the 3/2— rotational state built upon this orbital could be the lowest lying in energy due to a large negative decoupling parameter as in the 1 p211 band in 31Mg calculated by Kimura et al. [67] This presents the low-lying rotational excitations with J7r = (3/ 2, 5/ 2)‘ as candidates for the 780 keV state. The momentum—constrained analysis for the 297 keV suggests a transition that is more likely unstretched as discussed above. This supports the assignment of 5 / 2‘ to this state assuming the final total angular momentum of 5/ 2 for the 484 keV state. The assignment of 3/2" cannot be ruled out and is consistent with the evidence in Subsec. 4.3.7. However, the magnitude and nature of the alignment in the wings of the 99 LLLJJLLIIlLLIlllllLLllllllllllLlllllllllllllllllI 6 _- Neutron one-particle levels In Woods-Saxon potential -_ [A ; vWS = — 40.0 MeV R = 3.990 fm (A = 31) a = 0.67 fm : i A 0" - . 4 E .’.,,-°’[303 7/2) E’ . " /"’ " E‘ 2 E —- q‘».:::i?emfi.o~:-: ;— . — O I': .... 43°12‘5ch :1— : . A -’ t.....oo. -‘ f ..... .‘~‘~ ...... .... 3211/2 .— > O _,/' .,.l__ l t a) o [— E . ‘~2[321 3/21 I V -2 M: c: - w [200 1/2) _ -4 .__ -6 :- -8 [2113/2] :- 3 [2201/2] I '10 IIIIIIIIIFTHIIIIIII llll'llllllllllIlllllllllllll -0.4 03 02 -0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 quadrupole-deformation parameter B Figure 5.3: Neutron one-particle levels calculated in Hamamoto [85] with the param- eters of the Woods-Saxon potential tailored to reproduce the ground state spins of 31Mg [22] and 33Mg [21]. 100 momentum distribution is not certain, so the statements deduced without momentum gating on the residues are considerably weaker. Since the measurement using the whole distribution is also subject to normalization uncertainties, it is possible that the a2 calculation is marginally high. A smaller a2 would bring the measurement in accordance with the momentum-constrained analysis suggesting a AJ = 0 mixed 1W 1 / E2 transition. The other states populated are likely dominated by 219211 configurations. In partic- ular, since the ground state and 484 keV state both show signs of direct feeding, they have a negative parity. This agrees with the interpretation of Yordanov et al. [21]. Furthermore, the authors suggest that the 705 keV state has an opposite parity from the ground state. While the direct feeding in the current work seems to discredit this, the uncertainties which remain due to unplaced transitions surely could reduce the partial cross section to this weakly populated state. The only other state observed in the present experiment is at 1242 keV with a sizable direct feeding in the reaction which excludes the assignment based on fi-decay of 1/2+; however no clear alternative is present from the analysis. 101 Chapter 6 Outlook The current work establishes the possibility of performing '7-ray angular distribution and linear polarization measurements with intermediate energy beams. Several op- portunities present themselves to improve the quality and consistency of the current work. To perform accurate measurements, experiments need to improve the reso- lution and sensitivity of the experimental setup and obtain higher statistics for the gamma rays of interest.. Fruition of these goals is within the reach of next-generation 'y-ray detector arrays such as GRETA [68] and AGATA [86,87]. In experiments using beams at intermediate energies, these improvements require a balance between different aspects of the array’s performance. For example, as detectors are placed further away from the target position, the angular resolution is improved. This allows for better Doppler correction of the energies and decreases the solid-angle attenuation factor for angular distribution and linear polarization effects. The trade-off lies in the decreased detection efficiency of the array as the detectors now cover a smaller solid angle. On the other hand, detectors could be moved closer to increase efliciency in exchange for a decrease in the energy resolution. Furthermore, due to the relativistic kinematics and angular distribution effects, the observed properties of the gamma rays in the lab reference frame depend on the angle 102 of emission relative to the beam axis. It is important to choose angles accordingly such that the sensitivity to properties of interest is maximized. This also must consider how the lab angles are related to the angle in the center-of-mass frame which varies with the velocity of the incoming beam. All these factors must be balanced accordingly when designing the detector setup, which is often limited by additional physical constraints. The next generation of ’7-ray spectroscopy will try to tackle these issues by im- plementing digital data acquisition for the arrays. This will allow for signal decom- position and 7-ray tracking in the detectors bringing about an improved interaction position resolution for the gamma ray. In turn, the detectors could have an in- creased energy resolution or could be brought closer to the target for an increased efficiency. While the next-generation arrays plan to implement digital acquisition, recently SeGA was outfitted with digital processing capabilities [88,89] to instrument the 18 central contact channels as well as the 576 segments, including interfacing this with the existing electronics for auxiliary detectors at N SCL. 6.1 Signal decomposition and gamma-ray tracking Digital data acquisition systems record the waveforms associated with the gamma- ray detection to memory and/ or disk of a computer allowing for further processing. Signal decomposition deconstructs the waveforms to determine precisely where the gamma-ray interacted within a given segment. The position determination is based on the Shockley-Ramo theorem [90] where the induced charge on the electrodes is related to the moving charges within the germanium semiconductor as a function of time. In particular, the electron—hole pairs generated when a gamma—ray interacts within the crystal induces charges on the segment where it interacts as well as neigh- boring segments. Here, the segment electrode which collects the drifting charges is identified as the segment where the interaction occurred, and the shape of the wave- 103 I I A 34 . E F E g 500 _. C4 . D4 r E4 __ 4 _ G4 r H4 3 g s E 'w—A-L w ‘A V—U: L‘— ‘fi if ‘m 8 i E E E I D F < 400 - r E :' ' E E v :A A l : AAAAAAAAAA .A A: AAAAAAAAA A o b r h r b b p p I P , .0 b h . 3600 A3- 33». 03. 03; 353. 63;. 63:, 113 :1: E v‘. -' . : : a A ‘—‘ i “‘r‘ *u i L, ‘9 - - L 4‘41- "1 ___ 3,... -,.I T‘ b b r '— —— I I ' b D 1 r - r E 4‘", I- :- L L L In - I < 1 : I : h P . 1 v p b p . ’A A A1 A IA ........ ’ I 'AA lAAAAl ’ ......... ’ LLA . I E I I f : I m h I > b I F . 6001 A2; 1321, 02; .021;52=_ F2_ 62_ 1-12 ’ t : It i I .‘ r E - A z : a . soo- erb' ‘33...” v i- ’ NW I I I I E . h I 1 b s 400: r r r r r " r r E E E E E E E E aooo-Ref- ALAAAA- AAAAAAAAA l AlAA A LA A 1AA A I A I A IA A I AAAAAAAA AI A_LAA I F I b I h . p b )1 . D b p t b 500; A1 L 31 ; c1 _ D1 1 E1 ._ F1 _ <31 _ H1 E F E I t E t F m' I p p - P . t . m 1r ’ ww“- ’“fi'f ‘_ ." : 3" - ’.. f'_.""'" ’ WWW . r p p 1,;- 1 '. :' b '- ..' r . t : : : 7 : =. .- r . ' _ 400: r r r r r r ' 1000 I I p h r p :_. 1 : ........... f ........... i-.21-1-11-. ........ 1.2: ........... t.--1 ....... f ........... AAln-AAILA 005 1.0 005 1.0 0.5 1.0 005 100 005 100 005 1.0 005 1.0 005 1.0 0'5 1'0 Time [us] Figure 6.1: Induced signals recorded in the 32 segments and central contact of a SeGA detector for several events where a 1332 keV gamma ray from 60Co interacts in single segment E2 at different positions. 7-7 coincidences are recorded in a CsF detector; this serves as a timing reference for the measurement [91]. forms induced in all of the segments carry information about the interaction position. The magnitude of these signals is related to the energy deposited by the gamma ray in the detector (see Fig. 6.1). To understand how the signal changes in time, one must be able to understand the motion of the charge carriers within the crystal. This requires both an accurate determination of the electric field in the detector as well as the mobility of the charge carriers within the crystal structure. The induced signals can then be calculated using weighting potentials 050 which consider the solution to Maxwell’s equations with boundary conditions such that only one electrode has a non-zero fixed potential. 104 E0014 'I'l'I'tI'TIIIIIIUIIII'l'Ii'l' ‘ E I...I'U'III'I'li'til'I'Ili'III'U‘ E o ' E o ‘ E 50.1 V > > 9 8 0.06 0.04 0.02 o 50010001500200025003000 o 50010001500200025003000 E(Vlm) E(V/m) Figure 6.2: Drift velocities as a function of electric field for (left) electrons and (right) holes along the [100] axis in germanium with corresponding model calculations for a HPGe detector [93]. The signal Q(t) induced in the electrode with a given weighting potential is then: Q(t) = -q¢0($(t)) (6-1) for a drifting charge q whose position $(t) as a function of time depends on the electric field in the device. The motion of the charge carrier depends on its mobility within the crystal which can be fundamentally different for the electrons and holes. This also depends on the orientation of the crystallographic axes. The mobility along one crystallographic axis can be related to the drift velocity for different electric fields in a relatively simple model [92] as shown in Fig. 6.2. Since this is anisotropic given the crystal structure of germanium, additional effects must be considered as discussed in further detail in Mihailescu et al. [94]. Once the motion of the charge carrier is 105 known, the resulting signals on the electrode for an interaction at a given position can be calculated. After the interaction positions are well known within the detector, ’y-ray trackng can be used to determine the ordering of these interactions. The ordering of interac- tions is vital because the first interaction point determines the angle in the Doppler correction formula; knowledge of this angle is the primary limit on the energy reso- lution for experiments with intermediate-energy beams. Combined with the second interaction point, this also determines the first scattering of the gamma ray which is intrinsically related to its linear polarization (see Subsec. 2.4.1). This tracking is gen- erally based on the Compton scattering formula (Eq. 2.5) by comparing the scattered experimental energies of the photon to the calculated energies determined by the scattering angle. Given a chosen ordering of interactions, the tracking gives a figure- of—merit based on the deviation in energy. Especially poor figure-of—merits could be an effect of incomplete absorption in the detector and can be rejected, reducing the signal-to-noise ratio in the spectrum. In the end, the ordering of interactions with the best figure-of-merit is taken to be the true ordering; though this is not guaranteed to be a unique solution [95]. Here, information about the cross-sections of the Compton and photoelectric effect can provide further information about the most likely order- ing of events. Further information is available about a number of algorithms in the literature [96—98], including methods of reconstructing gamma-ray interactions which occur across several detectors. 6.2 SeGA digital electronics While digital acquisition is an integral part of future 'y-ray detector arrays, signifi- cant improvement can be made through developing instrumentation to acquire digital data for current arrays. Such an implementation has been done with SeGA where 106 digital data acquisition modules were developed in collaboration with XIA LLC with the particular details of the Digital Data Acquisition System (DDAS) discussed in previous works [88, 89,91]. The system is comprised of 39 Pixie-16 modules with 100 MHz 12-bit digitizers arranged in 4 custom compact PCI/PXI crates each with its own dedicated computer for event processing (and eventually signal decomposition). With a conservative goal of improving the position resolution by a factor of two, the energy resolution could be improved two-fold or the efficiency of the array quadrupled (subject to experimental space constraints). For intermediate-energy beam experi- ments, one should note that the angle of gamma-ray emission relative to the beam axis is dominately determined by the longitudinal interaction position in a given de- tector. An increase by two in resolution should be feasible if one can discern if an event interacted in one longitudinal half or the other for a given event. Simple al- gorithms which only incorporate the integrated induced signals have been shown to generate this magnitude of improvement for a 662 keV gamma ray from a collimated 137Cs source [99]. One aspect that is especially important for the successful integration is the timing in the system. The waveforms acquired by DDAS should be able to be synchronized. Small constant offsets which can result from cable delays can be compensated for. To assess the constancy of this offset, tests were performed by giving DDAS a sinusoidal waveform from a high-precision signal generator. The relative phases between differ- ent channels gives information about the delay. If the frequency of the signal is well known, the X2 analysis to determine the phase is linearized [100] making the analysis for the ~600 channels significantly less computationally intensive. DDAS was shown to consistently capture synchronous waveforms with jitter less than 300 ps as shown in Fig. 6.3; note that this resolution is significantly below the 10 ns time step based on the 100 MHz clock of the Pixie-16 cards. Timing between detectors in SeGA and auxiliary detectors must also be established. For in—beam runs, a fast plastic scintilla- 107 Counts 1.— -111111 llllllllllllilLlLlllllil lllllll JLllLllllllllll IIHHIIILILIII —2 -1 o 1 2 o 0.2 0.4 0.6 Time [ns] Time [ns] Figure 6.3: (left) The mean delay of signals for segments in SeGA and (right) the standard deviation of the delay for events in a given channel [91]. tor detects the residue and produces the reference signal. This signal has a rapid rise time and decay time which is beyond the resolution of a 100 MHz ADC. To address this, each master trigger derived from the fast scintillator also creates a pulse with similar characteristics to that from a germanium detector to use for timing in DDAS. The time is then determined on an event—by—event basis using the information about the time to cross the constant fraction amplitudes of 10% and 20% from the energy sums calculated for each event on the Pixie-16 card. This results in a 10 ns timing resolution for the 1332 keV peak emitted from a 60C0 source in coincidence with the 1173 keV gamma rays detected in a fast CsF detector [91] as shown in Fig. 6.4. Once the waveforms can be properly synchronized and correlated with events in external detectors, the next step is to develop a. database of calculated waveforms to which experimental waveforms can be compared in order to determine the most likely 108 Counts l 2000 r 1 000 Counts 0 1000 2000 3000 T C Energy [keV] 3 O . ‘? . 500i 1 L Pl L l l 700 800 ‘ '960' 1 #1000 Time [ns] Figure 6.4: (lower left) Energy-time two-dimensional plot for a SeGA detector with coincident gamma rays in a CsF detector. (above) Time projections for all the events (shaded) as well as gated on the 1332 keV transition (clear). (right) Full energy projection [91]. 109 ‘1 0° 0 Longitudinal distance (mm) 0.1-1:010: O O 0 5 10 15, 20 25 Radial distance (mm) Figure 6.5: Segment division in a SeGA detector for a r-z plane based on the drifting of holes to a given electrode. The longitudinal distance is measured from the “A” segment (0 mm) to the “H” segment (80 mm). position of interaction. Algorithms designed for GRETINA were adapted to the SeGA geometry and utilized to investigate the sensitivity of the array. First, the division between segments was investigated based on the movement of the holes to a given segment electrode. The result of the calculation is shown in Fig. 6.5. The uniform 1 cm physical division of the electrodes is evident at the outermost radius as the holes drift to the nearby contact. Further away from the segment electrode though, the holes are subject to greater effects from the field. This especially is evident near the central contact which ends before the “A” segments. The fringe fields resulting from this create a much larger volume where interactions occur in what is identified as the “A” segment and decreased volume for “B” segment interactions. This change in volume should be accounted for when determining the interaction position within a segment based on the centroid of the spatial distribution. For lower energy gamma rays, this must be convoluted with the fact that gamma rays are also more likely to 110 8 d O Ill|l||leAJLILllIIIIIJIIll a 0| 0| O ------ ~ -~ l"'o""é""1'd"'1'5'"é'd"'2'5"'3'd”s§ Figure 6.6: Calculated grid points for an “A” segment of a SeGA detector in the 316-? plane (left) and :l:-y plane (right) which sample the segment volume to maximize the sensitivity to the changing shape of the waveform. interact in the detector on the side closer to the source. Considering the centroid from a naive uniform physical division clearly does not account for these effects. The database of calculated waveforms should maximize the position sensitivity for each individual calculation point. To achieve this, an iterative technique is applied to an originally uniform cylindrical grid. In a given iteration, the grid is adapted to make the x2 uniform between any two neighboring grid points. The adaptive grid thus takes finer samples where the waveform changes rapidly, for example near segment boundaries and electrodes. An example grid is shown in Fig. 6.6 for an “A” segment. A given waveform can then be compared to the database waveforms with a X2 analysis to assign the position of interactions. The signal calculations have been developed to a preliminary stage to demonstrate the sensitivity for SeGA detectors. To maximize the position resolution, other effects must be considered. The calculated signals need to be folded with the electronic response of the detector including the preamplifier as well as the cross talk (both dif- ferential and integral) between different electronics channels. Anisotropies in the drift velocity and charge trapping also are not accounted for in the current calculations. Source experiments to characterize the detectors of SeGA are necessary to constrain parameters such as the crystal orientation and the mobility of the charge carriers. 111 These topics will be addressed in future work. 6.3 GRETINA impact While DDAS will allow considerable improvements of the position resolution of SeGA, the detectors are not ideally designed for sensitivity of the produced waveforms to the interaction position. Next generation arrays such as GRETA [68] and its pro- totype GRETINA have incorporated this sensitivity directly into the design of the detectors, for example by minimizing cross talk between channels. The goal of the detector and electronics design was to achieve a position resolution of less than 2 mm RMS in three dimensions which has been recently demonstrated [101]. This is a five-fold improvement on the position resolution currently given by the segmentation of SeGA detectors. The possibilities with the GRETINA prototype alone are monu- mental, especially for experiments at fragmentation facilities. This superior position resolution will greatly enhance the sensitivity of the array to angular distribution and correlation effects. GRETA is proposed to cover the full 47r solid angle with 30 detector clusters each of which house 4 HPGe detectors in a common cryostat. The GRETINA pro— totype under current construction will have In coverage and will demonstrate all the necessary aspects of implementing the entire array. The physical design of GRETA places the 30 clusters in five angular rings (32°, 58°, 90°, 122°, and 148°) to cover the entire solid angle. GRETIN A will consists of 7 clusters which will be distributed among the forward angles to maximize the sensitivity of the intended experiment. The implications for 'y—ray angular distribution and linear polarization measurements are discussed here. The sensitivity to these effects depend on the values of the Legendre polynomials at the center-of-mass angle for a given velocity as well as the efficiency of the detectors 112 Table 6.1: Relevant quantities for angular distribution measurements for GRETA angles assuming 6 = 0.4. 2 Olab 00m. P2(COS 0om.) P2( )(COS 0om.) (imam/(1913b Emult 32° 47° 0.190 1.62 1.92 1.39 58° 81° -O.459 2.92 1.35 1.16 90° 114° -0.260 2.52 0.84 0.92 122° 140° 0.383 1.23 0.57 0.76 148° 159° 0.803 0.39 0.47 0.68 taking into account the effects of the Lorentz boost. The relevant quantities are _ expressed in Table 6.1. To maximize the sensitivity to angular distribution for a measurement in two rings, the detectors should be placed to maximize the difference in P2(cos 60m). At the same time, the efficiency should be optimized so the requisite statistics can be collected taking into account the forward boost in the solid angle for the Doppler effect as well as considering that gamma-rays emitted forward are Doppler shifted to higher energies and thus are detected with less efficiency. For angular distributions, the figure—of-merit, F, calculated from the statistical analysis is: 1 1 )-1/2, F = (P2031) - P2($2))(a + S (6.2) where 131,2 = cos 00,111,,13 and 61,2 is the efficiency of detectors at that angle. The largest difference in P2 is clearly between the two angles for the GRETINA setup in the forward hemisphere. From this alone, GRETIN A has at least a two-fold increase in angular distribution sensitivity compared to SeGA. This does not consider the four-fold efficiency increase, which will improve the sensitivity by another factor of two. Figure 6.7 shows the figure-of-merit for a two angle measurement with one of the angles fixed at 32°. This does not consider other effects which may be of experimental importance such as the energy resolution necessary to resolve the gamma transition as well as increased signal-to-noise for the forward angles. For reactions which only result in moderate spin alignment, a two-point measurement is sufficient as only a2 needs 113 . ’.‘. I ” a » ’I . 0.4 C,‘ ’ ‘ LL. : : 0.2; i f 0.1 f f 0.0 ”l r I i i L i i i i ' 1 I i i r l i “ 0 50 100 150 glab Figure 6.7: Angular distribution sensitivity for residues with 6 = 0.4 (solid) and ,8 = 0.525 (dashed) considering placement of a second ring with the first ring fixed at 32°. to be determined. Detectors should be focused in two rings to maximize the statistics and extract the most precise measurement. It is also useful to extract branching ratios for transitions in detectors which have angular distribution effects minimized. For the GRETA angles, effects from angular distribution are present in all rings for current typical intermediate beam energies. At the next generation Facility for Rare Isotope Beams (FRIB) [102] which will upgrade NSCL’s capabilities, primary beam energies of 200 MeV/u will be obtainable for elements as heavy as uranium. For these energies, velocities of nuclei near 6 = 0.525 will result in gamma-ray emission with minimal angular distribution effects in the 32° ring of GRETA while preserving significant sensitivity in other rings. At these higher beam energies, measurements at backward angles become increasingly important as shown in Fig. 6.7. GRETINA will also significantly improve the linear polarization sensitivity. This is a result of the array geometry as well as the abilities of ’y-ray tracking. The polarization is maximized at 90° in the center-of-mass frame. GRETINA’S detectors 114 at 58° are quite suitable for polarization measurements. Using segment information to determine the scattering angle alone, the sensitivity of the GRETIN A detectors would be comparable to SeGA (See Table 3.1). Better resolution of the scattering angle based on the better position resolution from signal decomposition increases the sensitivity by a factor of two based on simulations [96]. Greater benefit comes from the ability to track gamma rays through the detector. Current limitations on the sensitivity stem from limiting the analysis to gamma rays which only interact in two segments which reduces the detection efficiency markedly and does not consider the possibility of multiple interactions within a segment which could be resolved by signal decomposition. The gamma-ray tracking algorithms discussed by Schmid et al. [96] would deduce the correct ordering of the first and second interaction points ~80% of the time for a 2 mm position resolution. This increases the efficiency of detecting the relevant Compton scattering events by a factor of three. The combined improvements in sensitivity and detection efficiency will result in a figure-of—merit roughly 12 times better than SeGA. Furthermore, this does not consider the benefits from measuring the continuous distribution of the first scattering angle. With these improvements, angular distribution and linear polarization measure- ments will become accessible for many more transitions. .With the increased efficiency, one can even consider the possibility going further and consider 7—7 angular correla- tions to give more information. A sizable piece of this improvement is also available to detectors in SeGA by using the Digital Data Acquisition System. This Opens the opportunity to probe the spins and parities of excited states in many exotic nuclei. 115 " arr-j 1'. it. ._.-.. Chapter 7 Conclusions Knockout reactions are a powerful spectroscopic tool to explore the structure of exotic nuclei. In particular, proton knockout reactions on neutron-rich nuclei access even more rare isotopes. Furthermore, the nuclear spins of the reaction residues are aligned following the removal of the proton. This results in the angular distribution and linear polarization of the emitted gamma rays. Measurement of the distribution and polarization allows for spin and parity assignment of the populated states when combined with the information from the knockout selection rules or from previous experiments. For 31Mg, two levels were identified as the normal shell model states with no ex- citations across the N = 20 shell gap. The spacing between these levels agrees with those produced by a USD shell model calculation. Assignment of the spins of the populated states was possible by examining the angular distribution of the gamma rays in a momentum-constrained analysis which removes the possible systematic un- certainty related to the efficiency of the gamma-ray detector. Parities could also be tentatively assigned based on linear polarization measurements, knockout selec- tion rules, and Weisskopf estimates of transition rates. A quenching of spectroscopic strength was observed in the reaction which agrees well with trends for the removal 116 of a deeply—bound nucleon [77]. The lowest-lying state with normal shell ordering was also identified in the 33Mg nucleus after a proton was stripped from the 34A] nucleus. Due to the more compli- cated structure of the incoming 34A1, a fragmentation of strength was also identified to several 2p2h states in the residue. As no prior spins and parities are firmly es- tablished in 33Mg, absolute spin and parity assignments are not possible. However, information from the gamma-ray transitions suggests certain multipolarities for the decays from the low-lying states. A level scheme is proposed which agrees well with a recent measurement of the ground-state spin [69] as well as Coulomb excitation data [70]. The identification of the states with no excitations across the N = 20 shell gap contributes to the knowledge of the “island of inversion”. The qualitative location of these states in the level schemes are predicted well by calculations, but true quantita- tive agreement for all the states is diffith to reproduce. For 31’33Mg, the excitation energy of the lowest-lying state with a normal configuration above the inverted ground state remains relatively constant showing the similarity between these two isotopes. Similar investigations into more neutron-rich magnesium isotopes (as well as neon isotopes) would add additional understanding about this region and the interactions which are responsible for the inversion. Future improvements will make a wide swath of exotic nuclei accessible for explo— ration. As the shell structure evolves, methods of making model-independent spin and parity assignments are especially critical. Gamma-ray angular distribution and linear polarization measurements will be an essential part to probe this evolving structure. N ext-generation arrays of gamma ray detectors implementing digital data acquisition will have an increased sensitivity for making these measurements. The techniques established in the present work will allow future experiments to effectively harness this sensitivity to enrich our knowledge about the atomic nucleus. 117 Appendix A Gamma-ray angular distribution Many different styles of notation exist in the literature; here we rely on the definitions of Krane [103] as in Olliver et al. [43]. In this notation, the angular distribution is decomposed into a series of Legendre polynomials, PA, in the angle 6 with respect to the beam axis: 21’ me) = Z AA(J,-, Jf,6)B,\(J,-,0)PA(0036) (A.1) A20 for a gamma ray transition between an initial state with spin J; and a final state J f. This depends on a quantity B A that characterizes the initial spin orientation in the nucleus described by the parameter a (see Eq. A.4). A ,\ captures the characteristics of the transition including the mixing ratio 6 of electric and magnetic matrix elements. The expansion is taken up to order 1' = l + 1 where l is the minimum multipolarity of the transition dictated by the triangle relation. 1: [Ji—Jfl (A2) The transition probability for different multipolarities is roughly proportional to (Ero / hc)21+1. For typical gamma ray energies E and nuclear densities m z 1.2fm, Em << he, so higher orders of the angular distribution are negligible. Furthermore, 118 for spin orientation along an axis that is azimuthally symmetric, only even order Legendre polynomials contribute. The spin alignment depends on the reaction mechanism. Each reaction mechanism preferentially populates different magnetic substates in the reaction residues, where the magnetic substate m is the projection of the total nuclear spin, J, along the beam axis, i.e. m = J2. The reaction thus populates a distribution w(m) of the magnetic substates which is then normalized to unity. The coefficients BA depend on this substate distribution Ji J- J- A BA(J,-,a)=\/(2/\+1)(2J,-+1) Z (—1)Jz'+m ' ' w(m;a). (A.3) Typically, the substate distribution is taken to fit according to some model parame- terized by 0. In a given reaction, there is either a tendency to produce spins aligned with the quantization axis (prolate alignment, m = :ini) or perpendicular to it (oblate alignment, m = 0). A reasonable approximation to the substate distribution in the case of the oblate alignment is a Gaussian with zero mean as studied by Yamazaki [104] based upon work done by Diamond et al. [105]. With the proper normalization, the distribution has the form e—m2/202 w(m) = Z” e_m,2/202. (A.4) [— Note that this takes into account the azimuthal symmetry of the reaction, so w(m) = w(—-m). The Gaussian distribution has a width 0 which is connected to the amount of oblate alignment in the system. The maximal alignment happens when a —* 0 where only m = 0 is populated in the reaction. This results in a maximal alignment 119 Table A1: Maximum alignment coefficients for different types of alignment. prolate oblate (half-integer J) oblate (integer J) max _Bzmax(J,-) J,(2J-—1) 1 (2J-—1)(2J-+3) J-(J-+1) p20 (‘10— x/5 fifilflbfiifi —Z sz'(Ji+12) —\/(2Ji:1)z(2Jz'+3) for a given spin from Eq. A3 of J,‘ J; A 8313336010: x/(2/\ +1)(2J,-+1)(-—1)Jz' 0 0 0 (11.5) J,- J, /\ 3111300): \/(2,\ +1)(2J,-+1)(_1)Jz-+1/2 , —1/2 1/2 0 for integer and half—integer spins respectively. Similarly, for prolate alignment, the formula can be extended to favor the states m = :ini [44] with a Gaussian distribution about both. Accordingly, the substate distribution is e—(J.—Iml)2/2a2 w(m) = J (A6) 2' —(.I-—|m'|)2/2o-2 Zm’2—Ji 8 2 which leads to a maximal alignment coefficient of J ° J - /\ 33131541,) = \/(2A + 1)(2J,- + 1) ' . . (A.7) —J,- J,- 0 The amount of alignment is often cited in terms of the percentage of this maximum, Brele': 0‘) = Bg(J,-, a)/B§"ax(Ji). This is related to B2 since the higher order terms in the angular distribution have a considerably smaller effect on the experimental observables for the typical experiment. The reduced form of the maximum alignment coefficient is shown in Table A.1 where pm = B A / (2A + 1) is another coefficient of alignment commonly used in the literature. It is important to note that 82(Ji) > 0 120 for prolate alignment while Bg(J,;) < 0 for oblate alignment. Once there exists some degree of alignment in the nucleus, the gamma rays emit- ted will exhibit an angular distribution with respect to the alignment axis. The alignment condition automatically precludes angular distribution from J,- = 0 and J,- = 1 / 2 states for axially symmetric arrangements. The distribution of the gamma rays is related to the properties of the transition. This is described by the coefficient AA(J,-, Jf,6) which has the form 1 AA(J1', Jf, (5) =W(FA(L, L, Jf, Ji) + 26F/\(L, L +1,Jf,.]i) (A 8) +52FA(L + 1,L+1,Jf,Ji)) for a transition of a given multipolarity L expressed in terms of generalized F - coefficients. For a specific transition, magnetic and electric matrix elements can contribute to the transition. These must have the same change in parity though. For magnetic transitions, the parity change is An = (—1)L+1, and for electric transitions we have A7r = (—1)L. That is to say that a transition with a given character of multipolarity L mixes with the transition of opposite character with multipolarity L + 1. As higher multipolarity transitions are often greatly hindered due to the en- ergy dependence of the transition rate, higher order contributions to the mixing are ignored. The mixing is quantified by the mixing ratio 6 with = (Jf|T(7r’L)|J.-> (A9) defined in terms of the rates T for transitions with character 7r or opposite character 7r’. For a pure transition of multipolarity L (i.e. 6 = 0), the angular distribution coefficient in Eq. A.8 becomes A A = F ,\(L, L, J f, J,). The F coefficients are defined according to the generalized F coefficients [106] where only one gamma ray is observed 121 __ m, .. ; _1. L - . .1. [LL in“: 1-01.0 -0.5 00 0.5 1.0 —01.0 —0.5 0.0 0.5 1.0 Figure A.1: Angular distribution coefficients and linear polarizations at 60m, = 113° relevant to the current work for transitions with AJ = —2, —1,0, 1, 2 (filled squares, filled circles, stars, hollow circles, hollow squares respectively) given an initial 66% prolate alignment in a nucleus with initial spin 3/ 2 (left) or 5 / 2 (right). with relation to the alignment axis. These coefficients are then FA(L, L’, Jf, J,-) =(—1)Ji+Jf+1\/(2/\ +_1)(2L + 1)(2L’ + 1)(2J,- + 1) L L’ ,\ L L’ A (A.10) 1 —1 0 Jz' Ji Jf A table of F coefficients can be found in the compilation by Wapstra [107]. Angular distribution is also accompanied by linear polarization of the gamma rays. For mixed dipole transitions, the resulting polarization at a given angle is: 3 P(6) = i (30.2 sin2 6 + (14(251—sin2 6 — 1—2 sin2 26) — 8 sin2 6 X:w(m)(—1)J+m\/5(2J+1) J J 2 =1: 6 ) (A.11) m —m 0 5 35 —2- (2 + a2(2 —— 3sin2 6) + a4(2 — ~4— sin2 6 - 1—6 sin2 26)). Determination of a2 and P identifies the multipolarity of the transition. For similar nuclear alignments, pure transitions of the same multipolarity and AJ lie in a similar position in the 0.2-P plane irrespective of initial spin as shown in Fig. A.1 and often 122 1.0 0.5 0.0 —0.5 E2 7191.0 —0.5 0.0 0.5 1.0 Figure A.2: Angular distribution coefficients and linear polarizations for a mixed M 1 / E2 transition (3/ 2 —> 1 / 2) for different initial alignments from 10% (innermost) to 90% (outermost) prolate alignment. The curves represent mixed transitions with 6 E {—20, 20]. determination of solely the signs of a2 and P provides the necessary information about the multipolarity. The magnitude of these observables is also dependent on the amount of spin alignment in the initial nucleus as well as the mixing angle for the gamma-ray transition (see Fig. A.2). If several gamma rays are emitted in a cascade, their directions of propagation are also related. If only one of these is detected, then the correlation must be integrated over the possible directions of any intermediate unobserved gamma rays. In general, this has an effect of dealigning the nucleus. The reduction of the alignment coefficient 123 B A is given by the deorientation coefficient U A as UA(L1J11J2) + 52UA(L', J1, J2) U LL’6,J,J = A(7i 1 2) 141-62 (A.12) for an unobserved transition (J1 —> J2) with mixed multipolarity L, L’ with a known mixing ratio 6. The appropriate U A for the pure transitions are J J A U1 = (—1>JI+J2+L+V<2J1+1><2J2 + 1) 1 1 . (A.13) J2 J2 L The final alignment parameter for N unobserved intermediate transitions between levels J,- and Ji+1 would be B:\(JN+1a0) = U,\(L1, [4,1,51,11,10 - - - U,\(Ln,sz,6, JN.JN+1)BA(Ji,0), (A14) given the known multipolarities Li, L] and mixing ratios 6,- of the intermediate tran- sitions. 124 Appendix B Relativistic Kinematics The gamma rays emitted from a source with a relativistic velocity can display asym- metries in the lab frame where none exist in the intrinsic frame. Understanding the relativistic effects is necessary to make statements about the asymmetries in the in- trinsic frame from the lab measurements. If we consider the coordinate system with the particle moving with velocity 6c along the z-axis, the photon emitted in a di- rection 61ab frame can be related back to the intrinsic frame through the Lorentz transformation of the four-vectors. (7 0067) 010 0 AZ: (13.1) 010 \[3'700 7} where '7 = (1 — 52)_1/2. This influences important observables of the gamma ray in- cluding energy and angle of emission as well as the lifetimes of excited states in nuclei. However, the transformation of polarization is simply connected to the difference of the angle of emission in the two reference frames. 125 B.1 Lifetime The Lorentz transformation on the standard position four-vector at time t and loca- tion x :c“ = (ct, x) (B2) results in time dilation in the lab frame. Lifetimes measured in the lab frame are simply related through the '7 factor. 7cm : 7lab/'7 (BB) B.2 Angle The related between angles in the lab frame versus the center-of-mass frame can be determined by Lorentz boosting the four-frequency N” = 27rw(1, f1) (34) for a photon with frequency w propagating in the in direction. This gives the result: COS 6lab -' 5 6 = COS CM 1 — 5 cos 6131) . (3.5) sm 6131, sm 90M = 7(1 _ flcos 9151b) This transformation of angle also impacts the solid angle d9 = d(cos 6) effectively boosting the gamma rays forward in the lab frame. _ 2 dQCM = 1 (3 2 (B.6) dfllab (l3 COS 6lab - 1) 126 B.3 Energy The familiar relativistic Doppler shift equation can also be obtained by transformation of the four-frequency (Eq. 34) With the energy E proportional to the frequency, the intrinsic energy is Elab = Ecm'y(1 + ficos 6cm). (B.7) Utilizing the relation B.5, the intrinsic gamma ray energy can be expressed in terms of the lab measurements Ecm = V'Elabu — (3005 elab)- (138) BA Polarization The linear polarization of light could also be affected by the relativistic kinematics. Here, the measured quantity of interest is the asymmetry of Compton scattering. This is directly related to the magnitude of the electric fields E parallel and perpendicular to the scattering plane in the lab frame. 2 _ 2 _ En EL _ ——2—7 (13.9) To investigate the components of the electric field relative to the scattering plane, we must utilize the field-strength tensor containing the electric and magnetic fields, ( 0 Ex E, Ez ) —E 0 —B;; B 43, Bz 0 —B_,, (—E,,. —By BE 0 ) Applying the Lorentz transformation for a boost of velocity BC, we obtain the result for the transformed field-strength tensor F /\p = A; F WA]: which gives us the desired transformation of the electric field. I I ’72 I E = , E — B — ——- -E B.11 ,( a x ) 7 + law > < > However, the scattering angles are also transformed between the lab frame and the intrinsic frame. If we consider the scattering to occur in the xz-plane, then the parallel and perpendicular components for a photon scattering in the fi direction are related to the basis vectors in the intrinsic frame 13’ = sin 0’s; + cos 9’s; é[| = cos 6'6; — sin 6'6; (312) é’i = a; with similar definitions with unprimed quantities for the lab frame. We then can relate the perpendicular component of the electric field measured in the lab to that in the intrinsic frame. EL=E-éi=Ey=7(E,’/—BB;) (8.13) For the electromagnetic radiation of the photon, the magnetic field can be expressed in terms of the electric field and direction of propagation via B’ = fi’ x E’. This gives us the simple result El 2 E1 -7(1+ Bcos6’). (8.14) Likewise if we perform the transformation on the parallel component of the elec- tric field, we can obtain a similar relation. This requires the more involved task of 128 transforming the angles accordingly, but it can be shown that we achieve E“ = E"l .)(1+ flcos6'). (13.15) Combining equations B.14 and 8.15, the asymmetry of the electric field (B9) is invari- ant under Lorentz transformation. Thus the measured polarization in the lab frame is related to the intrinsic frame solely through the transformation of the according angle of emission, P(6) = P(6’), given a linear relationship between the asymmetry and the linear polarization. 129 BIBLIOGRAPHY [1] Aage Bohr and Ben R. Mottelson. Nuclear Structure, volume 1. World Scientific, Singapore, 3rd edition, 1998. [2] T. Hamada and I. D. Johnston. A potential model representation of two-nucleon data below 315 MeV. Nucl. Phys., 34:382—403, 1962. [3] R. B. Wiringa, V. G. J. Stoks, and R. Schiavilla. Accurate nucleon-nucleon potential with charge-independence breaking. Phys. Rev. C, 51:38—51, 1995. [4] M. G. Mayer and H. D. Jensen. Elementary Theory of Nuclear Shell Structure. John Wiley and Sons, inc., New York, 1955. [5] B. A. Brown. The nuclear shell model towards the drip lines. Prog. Part. Nucl. Phys., 47:517—599, 2001. [6] P. D. Cottle and K. W. Kemper. Evolution of shell closures in neutron-rich Z = 10—20 nuclei. Phys. Rev. C, 662061301, 2002. [7] R. V. F Janssens. Nuclear physics: Unexpected doubly magic nucleus. Nature, 459:1069—1070, 2009. [8] P. G. Hansen. Nuclear halos: Structure and reactions. Nucl. Phys. A, 588:c1 — 09, 1995. Proceedings of the Fifth International Symposium on Physics of Unstable Nuclei. [9] J. R. Terry, D. Bazin, B. A. Brown, J. Enders, T. Glasmacher, P. G. Hansen, B. M. Sherrill, and J. A. Tostevin. Absolute spectroscopic factors from neutron knockout on the halo nucleus 15C. Phys. Rev. C, 69:054306, 2004. [10] B. A. Brown. New Skyrme interaction for normal and exotic nuclei. Phys. Rev. C, 58:220—231, 1998. [11] RH. Wildenthal. Empirical strengths of spin operators in nuclei. Prog. Part. Nucl. Phys, 11:5, 1984. 130 [12] T. Otsuka, R. Fujimoto, Y. Utsuno, B. A. Brown, M. Honma, and T. Mizusaki. [13] [14] [15] [16] [17] [18] [19] [20] [21] Magic numbers in exotic nuclei and spin-isospin properties of the N N interac- tion. Phys. Rev. Lett., 872082502, 2001. E. K. Warburton, J. A. Becker, and B. A. Brown. Mass systematics for A = 29 — 44 nuclei: The deformed A ~ 32 region. Phys. Rev. C, 41:1147, 1990. D. Guillemaud-Mueller, C. Detraz, M. Langevin, F. Naulin, M. De Saint-Simon, C. Thibault, F. Touchard, and M. Epherre. fi-decay schemes of very neutron— rich sodium isotopes and their descendants. Nucl. Phys, A426z37—76, 1984. T. Motobayashi, Y. Ikeda, Y. Ando, K. Ieki, M. Inoue, N. Iwasa, T. Kikucht, M. Kurokawa, S. Moriya, S. Ogawa, H. Murakami, S. Shimoura, Y. Yanagi- sawa, T. Nakamura, Y. Watanabe, M. Ishihara, T. Teranishi, H. Okuno, and R. F. Casten. Large deformation of the very neutron-rich nucleus 32Mg from intermediate-energy Coulomb excitation. Phys. Lett. B, 34629—14, 1995. B. V. Pritychenko, T. Glasmacher, P. D. Cottle, M. Fauerbach, R. W. Ibbotson, K. W. Kemper, V. Maddalena, A. N avin, R. Ronningen, A. Sakharuk, H. Scheit, and V. G. Zelevinsky. Role of intruder configuration in 26’28Ne and 30’32Mg. Phys. Lett. B, 461:322—328, 1999. DJ. Viera, J. M. Wouters, K. Vaziri Jr., R. H. Kraus, H. Wollnik, G. W. Butler, F. K. Wohn, and A. H. Wapstra. Direct mass measurement of neutron-rich light nuclei near N = 20. Phys. Rev. Lett., 5723253, 1986. A. Navin, D. W. Anthony, T. Aumann, T. Baumann, D. Bazin andd Y. Blu— menfeld, B. A. Brown, T. Glasmacher, P. G. Hansen, R. W. Ibbotson, P. A. Lofy, V. Maddalena, K. Miller, T. Nakamura, B. V. Pritychenko, B. M. Sher- rill, E. Spears, M. Steiner, J. A. Tostevin, J. Yurkon, and A. Wagner. Direct evidence for the breakdown of the N = 8 shell closure in 12Be. Phys. Rev. Lett., 852266—269, 2000. S. D. Pain, W. N. Catford, N. A. Orr, J. C. Angélique, N. I. Ashwood, V. Bouchat, N. M. Clarke, N. Curtis, M. Freer, B. R. Fulton, F. Hanappe, M. Labiche, J. L. Lecouey, R. C. Lemmon, D. Mahboub, A. Ninane aond G. Normand, N. Soié, L. Suttge, C. N. Timis, J. A. Tostevin, J. S. Winfield, and V. Ziman. Structure of 12Be: Intruder d—wave strength at N = 8. Phys. Rev. Lett., 962032502, 2006. G. Neyens, M. Kowalska, D. Yordanov, K. Blaum, P. Himpe, P. Lievens, S. Mallion, R. Neugart, N. Vermeulen, Y. Utsuno, and T. Otsuka. Measurement of the spin and magnetic moment of 31Mg: Evidence for a strongly deformed intruder ground state. Phys. Rev. Lett., 942022501, 2005. D. T. Yordanov, M. Kowalska, K. Blaum, M. De Rydt, K. T. Flanagan, P. Lievens, R. Neugart, G. Neyens, and H. H. Stroke. Spin and magnetic 131 [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] moment of 33Mg: Evidence for a negative-parity intruder ground state. Phys. Rev. Lett., 992212501, 2007. M. Kowalska, D. T. Yordanov, K. Blaum, P. Himpe, P. Lievens, S. Mallion, R. Neugart, G. Neyens, and N. Vermeulen. Nuclear ground-state spins and magnetic moments of 27Mg, 29Mg, and 31Mg. Phys. Rev. C, 77:034307, 2008. J. R. Terry, B. A. Brown, C. M. Campbell, J. M. Cook, A .D. Davies, D.- C. Dinca, A. Gade, T. Glasmacher, P. G. Hansen, B. M. Sherrill, H. Zwahlen, D. Bazin, K. Yoneda, J. A. Tostevin, T. Otsuka, Y. Utsuno, and B. Pritychenko. Single-neutron knockout from intermediate energy beams of 30’32Mg: Mapping the transition into the “island of inversion”. Phys. Rev. C, 77:014316, 2008. A. Gade, P. Adrich, D. Bazin, M. D. Bowen, B. A. Brown, C. M. Camp- bell, J. M. Cook, S. Ettenauer, T. Glasmacher, K. W. Kemper, S. McDaniel, A. Obertelli, T. Otsuka, A. Ratkiewicz, K. Siwek, J. R. Terry, J. A. Tostevin, Y. Utsuno, and D. Weisshaar. Spectroscopy of 36Mg: Interplay of normal and intruder configurations at the neutron-rich boundary of the “island of inver- sion”. Phys. Rev. Lett., 992072502, 2007. H. Iwasaki, T. Motobayashi, H. Sakurai, K. Yoneda, T. Gomi, N. Aoi, N. Fukuda, Zs. Flp, U. Futakami, Z. Gacsi, Y. Higurashi, N. Imai, N. Iwasa, T. Kubo, M. Kunibu, M. Kurokawa, Z. Liu, T. Minemura, A. Saito, M. Serata, S. Shimoura, S. Takeuchi, Y.X. Watanabe, K. Yamada, Y. Yanagisawa, and M. Ishihara. Quadrupole collectivity of 28Ne and the boundary of the island of inversion. Phys. Lett. B, 6202118 — 124, 2005. P. G. Hansen and J. A. Tostevin. Direct reactions with exotic nuclei. Annual Review of Nuclear and Particle Science, 53:219—261, 2003. J .A. Tostevin. Single-nucleon knockout recations at fragmentation beam ener- gies. Nucl. Phys. A, 6822320c, 2001. P. G. Hansen and B. M. Sherrill. Reactions and single-particle structure of nuclei near the drip lines. Nucl. Phys. A, 693:133—168, 2001. A. S. Goldhaber. Statistical models of fragmentation processes. Phys. Lett. B, 532306—308, 1974. Peter A. Zavodsky, Hannu Koivisto, Dallas Cole, and Peter Miller. Emittance studies of ARTEMIS — the new ECR ion source for the Coupled Cyclotron F acil- ity at NSCL/MSU. In J .L. Duggan and IL. Morgan, editors, The CAARI 2666: Sixteenth international conference on the application of accelerators in research and industry, volume 576, pages 619—622. American Institute of Physics, 2001. X. Wu, H. Blosser, D. Johnson, F. Marti, and RC. York. The K500-to-K1200 coupling line for the Coupled Cylotron Facility at the N SCL. In Proceedings of the 1999 Particle Accelerator Conference, pages 1318—1320, New York, 1999. 132 [32] [33] [34] [35] [36] D. J. Morrissey, B. M. Sherrill, M. Steiner, A. Stolz, and I. Wiedenhoever. Commissioning the A1900 projectile fragment separator. Nucl. Instrum. Meth. B, 204290—96, 2003. W.-M. Yao, C. Amsler, D. Asner, R.M. Barnett, J. Beringer, P.R. Burchat, C.D. Carone, C. Caso, O. Dahl, G. D’Ambrosio, A. DeGouvea, M. Doser, S. Eidelman, J.L. Feng, T. Gherghetta, M. Goodman, C. Grab, D.E. Groom, A. Gurtu, K. Hagiwara, K.G. Hayes, J .J . Hernandez-Rey, K. Hikasa, H. Jawah- ery, C. Kolda, Kwon Y., M.L. Mangano, A.V. Manohar, A. Masoni, R. Miquel, K. M6nig, H. Murayama, K. Nakamura, S. Navas, K.A. Olive, L. Pape, C. Pa- trignani, A. Piepke, G. Punzi, G. Raffelt, J.G. Smith, M. Tanabashi, J. Tern- ing, N.A. T6rnqvist, T.G. Trippe, P. Vogel, T. Watari, C.G. Wohl, R.L. Workman, P.A. Zyla, B. Armstrong, G. Harper, V.S. Lugovsky, P. Schaffner, M. Artuso, K.S. Babu, H.R. Band, E. Barberio, M. Battaglia, H. Bichsel, O. Biebel, P. Bloch, E. Blucher, R.N. Cahn, D. Casper, A. Cattai, A. Cec— cucci, D. Chakraborty, R.S. Chivukula, G. Cowan, T. Damour, T. DeGrand, K. Desler, M.A. Dobbs, M. Drees, A. Edwards, D.A. Edwards, V.D. Elvira, J. Erler, V.V. Ezhela, W. Fetscher, B.D. Fields, B. Foster, D. Froidevaux, T.K. Gaisser, L. Garren, H.-J. Gerber, G. Gerbier, L. Gibbons, F.J. Gilman, G.F. Giudice, A.V. Gritsan, M. Griinewald, H.E. Haber, C. Hagmann, I. Hinchliffe, A. Héicker, P. Igo-Kemenes, J .D. Jackson, K.F. Johnson, D. Karlen, B. Kayser, D. Kirkby, S.R. Klein, K. Kleinknecht, I.G. Knowles, R.V. Kowalewski, P. Kre— itz, B. Krusche, Yu.V. Kuyanov, O. Lahav, P. Langacker, A. Liddle, Z. Ligeti, T.M. Liss, L. Littenberg, L. Liu, K.S. Lugovsky, S.B. Lugovsky, T. Mannel, D.M. Manley, W.J. Marciano, A.D. Martin, D. Milstead, M. Narain, P. Na- son, Y. Nir, J.A. Peacock, S.A. Prell, A. Quadt, S. Raby, B.N. Ratcliff, EA. Razuvaev, B. Renk, P. Richardson, S. Roesler, G. Rolandi, M.T. Ro- nan, L.J. Rosenberg, C.T. Sachrajda, S. Sarkar, M. Schmitt, O. Schneider, D. Scott, T. Sjiistrand, G.F. Smoot, P. Sokolsky, S. Spanier, H. Spieler, A. Stahl, T. Stanev, R.E. Streitmatter, T. Sumiyoshi, N.P. Tkachenko, G.H. Trilling, G. Valencia, K. van Bibber, M.G. Vincter, D.R. Ward, B.R. Webber, J.D. Wells, M. Whalley, L. Wolfenstein, J. Womersley, C.L. Woody, A. Yamamoto, O.V. Zenin, J. Zhang, and R.-Y. Zhu. Review of Particle Physics. J. Phys. 0., 33:1, 2006. D. Bazin, J. A. Caggiano, B. M. Sherrill, J. Yurkon, and A. Zeller. The 8800 spectrograph. Nucl. Instrum. Meth. B, 204:629-633, 2003. J. Yurkon, D. Bazin, W. Benenson, D. J. Morrissey, B. M. Sherrill, D. Swan, and R. Swanson. Focal plane detector for the 8800 high-resolution spectrometer. Nucl. Instrum. Meth. A, 4222291, 1999. C. Scheidenberger, Th. Stéihlker, W. E. Meyerhof, H. Geissel, P. H. Mokler, and B. Blank. Charge states of relativistic heavy ions in matter. Nucl. Instrum. Meth. A, 142:441—462, 1998. 133 [37] [38] [39] [40] [41] [42] [43] [44] Kyoko Makino and Martin Berz. Cosy infinity version 8. Nucl. Instrum. Meth. A, 427 2338—343, 1999. S. Agostinelli, J. Allison, K. Amako, J. Apostolakis, H. Araujo, P. Arce, M. Asai, D. Axen, S. Banerjee, G. Barrand, F. Behner, L. Bellagamba, J. Boudreau, L. Broglia, A. Brunengo, H. Burkhardt, S. Chauvie, J. Chuma, R. Chy— tracek, G. Cooperman, G. Cosmo, P. Degtyarenko, A. Dell’Acqua, G. Depaola, D. Dietrich, R. Enami, A. Feliciello, C. Ferguson, G. Folger, Foppiano. F ., A. Forti, S. Garelli, S. Giani, R. Giannitrapani, D. Gibin, J. J. Gomez Ca- denas, I. Gonzalez, G. Gracia Abril, G. Greeniaus, W. Greiner, V. Grichine, A. Grossheim, S. Guatelli, P. Gumplinger, R. Hamatsu, K. Hashimoto, H. Ha- sui, A. Heikkinen, A. Howard, V. Ivanchenko, A. Johnson, F. W. Jones, J. Kallenbach, N. Kanaya, M. Kawabata, Y. Kawabata, M. Kawaguti, S. Kel- ner, P. Kent, A. Kimura, T. Kodama, R. Kokoulin, M. Kossov, H. Kurashige, E. Lamanna, T. Lampién, V. Lara, V. Lefebure, F. Lei, M. Liendl, W. Lock- man, F. Longo, S. Magni, M. Maire, E. Medernach, K. Minamimoto, P. Mora de Freitas, Y. Morita, K. Murakami, M. Nagamatu, R. Nartallo, P. Nieminen, T. Nishimura, K. Ohtusbo, M. Okamura, S. O’Neale, Y. Oohata, K. Paech, J. Perl, A. Pfeiffer, M. G. Pia, F. Ranjard, A. Rybin, S. Sadilov, E. Di Salvo, G. Santin, T. Sasaki, N. Savvas, Y. Sawada, S. Scherer, S. Sei, V. Sirotenko, D. Smith, N. Starkov, H. Stoecker, J. Sulkimo, M. Takahata, S. Tanaka, E. Tch- erniaev, E. Safai Tehrani, M. Tropeano, P. Truscott, H. Uno, L. Urban, P. Ur- ban, M. Verderi, A. Walkden, W. Wander, H. Weber, V. P. Wellisch, T. Wenaus, D. C. Williams, D. Wright, T. Yamada, H. Yoshida, and D. Zschiesche. Geant4 - a simulation toolkit. Nucl. Instrum. Meth. A, 5062250—303, 2003. M.J. Berger, J.H. Hubbell, S.M. Chang, J.S. Coursey, R. Sukumar, and D.S. Zucker. XCOM: Photon Cross Section Database (version 1.3). National Insti- tute of Standards and Technology, Gaithersburg, MD, 2005. K. Alder and R. M. Steffen. Emission of gamma radiation and nuclear struc- ture. In J .D. Hamilton, editor, The electromagnetic intreaction in nuclear spec- troscopy, pages 39—54. North-Holland, Amsterdam, 1975. P. M. Endt. Strengths of gamma-ray transitions in a = 6-44 nuclei (iii). Atomic Data and Nuclear Data Tables, 2323 — 61, 1979. P. J. Twin. Gamma ray angular distribution and correlation measurements (ii) experimental methods following nuclear reactions. In J .D. Hamilton, editor, The electromagnetic interaction in nuclear spectroscopy, chapter 15, pages 701—733. North-Holland, Amsterdam, 1975. H. Olliver, T. Glasmacher, and Andrew E. Stuchbery. Angular distributions of 7 rays with intermediate-energy beams. Phys. Rev. 0, 682044312, 2003. Andrew E. Stuchbery. '7-ray angular distributions and correlations after projectile-fragmentation reactions. Nucl. Phys. A, 723:69—92, 2003. 134 [45] O. Klein and Y. Nishina. iiber die Streuung von Strahlung durch freie Elek- tronen nach der neun relativistischen Quantendynamik von Dirac. Z. Phys, 52:853, 1929. [46] Y. N ishina. Die Polarisation der Comptonstreuung nach der Diracschen Theorie des Elektrons. Z. Phys, 522869, 1929. [47] W. Heitler. The Quantum Theory of Radiation. Oxford University Press, Lon- don,1944. [48] L. S. Fagg and S. S. Hanna. Polarization measurements on nuclear gamma rays. Rev. Mod. Phys, 312711—758, 1959. [49] A. von der Werth, F. Becker, J. Eberth, S. Freund, U. Hermkens, T. Mylaeus, S. Skoda, H. G. Thomas, and W. Teichert. Two Compton polarimeter construc- tions for modern standard 'y-spectroscopy. Nucl. Instrum. Meth. A, 3572458 — 466, 1995. [50] G. Duchéne, F. A. Beck, P. J. Twin, G. de France, D. Curien, L. Han, C. W. Beausang, M. A. Bentley, P. J. Nolan, and J. Simpson. The clover: a new generation of composite Ge detectors. Nucl. Instrum. Meth. A, 432290 — 110, 1999. [51] B. Schlitt, U. Maier, H. Friedrichs, S. Albers, I. Bauske, P. von Brentano, R.D. Heil, R.-D. Herzberg, U. Kneissl, J. Margraf, H.H. Pitz, C. Wesselborg, and A. Zilges. A sectored Ge-Compton polarimeter for parity assignments in photon scattering experiments. Nucl. Instrum. Meth. A, 3372416 —— 426, 1994. [52] D. Miller, A. Chester, V. Moeller, K. Starosta, C. Vaman, and D. Weisshaar. Linear polarization sensitivity of SeGA detectors. Nucl. Instrum. Meth. A, 5812713—718, 2007. [53] G. D. Jones. Calibration of Compton polarimeters. Nucl. Instrum. Meth. A, 491:452—459, 2002. [54] W. F. Mueller, J. A. Church, T. Glasmacher, D. Gutknecht, G. Hackman, P.G. Hansen, Z. Hu, K. L. Miller, and C. Quirin. Thirty-two fold segmented germanium detectors to identify 7—rays from intermediate-energy exotic beams. Nucl. Instrum. Meth. A, 466:492—498, 2001. [55] O. Wieland, F. Camera, B. Million, A. Bracco, and J. van der Marel. Pulse distributions and tracking in segmented detectors. Nucl. Instrum. Meth. A, 4872441 — 449, 2002. [56] OR Tarasov and D. Bazin. LISE++: Radioactive beam production with in- flight separators. Nucl. Instrum. Meth. B, 26624657 - 4664, 2008. Proceedings of the XVth International Conference on Electromagnetic Isotope Separators and Techniques Related to their Applications. 135 [57] Jens Lindhard and Allan H. Sorensen. Relativistic theory of stopping for heavy ions. Phys. Rev. A, 53:2443—2456, 1996. [58] Z. Hu, T. Glasmacher, W.F. Mueller, and I. Wiedenhéiver. An automatic energy- calibration method for segmented germanium detectors. Nucl. Instrum. Meth. A, 482:715—719, 2002. [59] R. Venturelli, D. Bazzaco, M. Bellato, A. Pullia, and Th. Kréill. Fold-dependent energy correction in segmented germanium detectors. In LNL Annual Report 2062, INFN—LNL (REP), volume 198, pages 156—157. 2003. [60] F. Falk, A. Linnfors, B. Orre, and J .E. Thun. Alpha-gamma angular correlation and hyperfine interaction in the decay of ggng. Phys. Scripta, 1213-19, 1970. [61] BA. Logan, R.T. Jones, and A. Ljubiéic. A figure of merit for gamma-ray polarimeters. Nucl. Instrum. Meth., 1082603 - 604, 1973. [62] J. Simpson, P.A. Butler, and LP. Ekstrém. Application of a sectored Ge(Li) detector as a Compton polarimeter. Nucl. Instrum. Meth., 2042463 — 469, 1983. [63] K. Starosta, T. Morek, Ch. Droste, S. G. Rohoziriski, J. Srebrny, A. Wierzchucka, M. Bergstréim, B. Herskind, E. Melby, T. Czosnyka, and P. J. Napiorkowski. Experimental test of the polarization direction correlation method (PDCO). Nucl. Instrum. Meth. A, 423216 — 26, 1999. [64] M. Devlin, L. G. Sobotka, D. G. Sarantites, and D. R. LaFosse. Simulated response characteristics of Gammasphere. Nucl. Instrum. Meth. A, 3832506 — 512, 1996. [65] H. Mach, L. M. Fraile, O. Tengblad, R. Boutami, C. Jollet, W. A. Pléciennik, D. T. Yordanov, M. Stanoiu, M. J. G. Borge, P. A. Butler, J. Cederkall, Ph. Dessagne, B. Fogelberg, H. F ynbo, P. Hoff, A. Jokinen, A. Korgul, U. Késter, W. Kurcewicz, F. Marechal, T. Motobayashi, J. Mrazek, G. Neyens, T. Nilsson, S. Pedersen, A. Poves, B. Rubio, E. Ruchowska, and the ISOLDE Collaboration. New structure information on 3OMg, 31Mg, and 32Mg. Eur. Phys. A, 25:105— 109, 2005. [66] G. Klotz, P. Baumann, M. Bounajma, A. Huck, A. Knipper, G. Walter, G. Mar- guier, C. Richard-Serre, A. Poves, and J. Retamosa. Beta decay of 3132Na and 31Mg: Study of the N = 20 shell closure. Phys. Rev. C, 47:2502—2516, 1993. [67] M. Kimura. Intruder features of 31Mg and the coexistence of many—particle and many-hole states. Phys. Rev. C, 75:041302(R), 2007. [68] I.-Y. Lee, J.T. Anderson, J. Bercovitz, A. Biocca, J. Comins, M. Cromaz, M. Descovich, D. Doering, P. Fallon, J.M. Joseph, C. Larsen, C. Lionberger, T. Loew, A.O. Macchiavelli, J. Pavan, D. Peterson, D. Radford, B.A. Savnik, C. Timossi, S. Virostek, H. Yaver, and S. Zimmermann. Greta white paper for 136 [69] [70] [71] [72] [73] [74] [75] [76] [77] the 2007 nsac long range plan. Technical report, GRETIN A Physics Working Group, 2007. S. Nummela, F. Nowacki, P. Baumann, E. Caurier, J. Cederkall, S. Courtin, P. Dessagne, A. Jokinen, A. Knipper, G. Le Scornet, L. G. Lyapin, Ch. Miehé, M. Oinonen, E. Poirier, Z. Radivojevic, M. Ramdhane, W. H. Trzaska, G. Wal- ter, J. Aysté, and ISOLDE Collaboration. Intruder features in the island of inversion: The case of 33Mg. Phys. Rev. C, 642054313, 2001. B. V. Pritychenko, T. Glasmacher, P. D. Cottle, R. W. Ibbotson, K. W. Kemper, L. A. Riley, A. Sakharuk, H. Scheit, M. Steiner, and V. Zelevinsky. Structure of the ”island of inversion” nucleus 33Mg. Phys. Rev. C, 652061304, 2002. Z. Elekes, Zs. Dombradi, A. Saito, N. Aoi, H. Baba, K. Demichi, Zs. Fiiléip, J. Gibelin, T. Gomi, H. Hasegawa, N. Imai, M. Ishihara, H. Iwasaki, S. Kanno, S. Kawai, T. Kishida, T. Kubo, K. Kurita, Y. Matsuyama, S. Michimasa, T. Minemura, T. Motobayashi, M. Notani, T. Ohnishi, H. J. Ong, S. Ota, A. Ozawa, H. K. Sakai, H. Sakurai, S. Shimoura, E. Takeshita, S. Takeuchi, M. Tamaki, Y. Togano, K. Yamada, Y. Yanagisawa, and K. Yoneda. Proton inelastic scattering studies at the borders of the “island of inversion”: 30’31na and 3334Mg case. Phys. Rev. 0, 73044314, 2006. P. Himpe, G. Neyens, D.L. Balabanski, G. Bélier, J.M. Daugas, F. de Oliveira Santos, M. De Rydt, K.T. Flanagan, I. Matea, P. Morel, Yu.E. Penionzhkevich, L. Perrot, N.A. Smirnova, C. Stodel, J.C. Thomas, N. Ver- meulen, D.T. Yordanov, Y. Utsuno, and T. Otsuka. g factor of the exotic N = 21 isotope 34Al: probing the N = 20 and N = 28 shell gaps at the border of the ”island of inversion”. Phys. Lett. B, 6582203 — 208, 2008. J. Tostevin. personal communication. M. Robinson, P. Halse, W. 'Ii'inder, R. Anne, C. Borcea, M. Lewitowicz, S. Lukyanov, M. Mirea, Yu. Oganessian, N. A. Orr, Yu. Peniozhkevich, M. G. Saint-Laurent, and O. Tarasov. New isomer 32Alm. Phys. Rev. C, 532R1465— R1468, 1996. J. Dechargé and D. Gogny. Hartree—Fock—Bogolyubov calculation with the D1 effective interaction on spherical nuclei. Phys. Rev. C, 2121568—1593, 1980. H. Ueno, D. Kameda, G. Kijima, K. Asahi, A. Yoshimi, H. Miyoshi, G. Kato, D. Nagae, S. Emori, T. Haseyama, H. Watanabe, and M. Tsukui. Magnetic moments of ‘ngIN and 13.4119. Phys. Lett. B, 615:186—192, 2005. A. Gade, P. Adrich, D. Bazin, M. D. Bowen, C. M. Campbell, J. M. Cook, T. Glasmacher, P. G. Hansen, K. Hosier, S. McDaniel, D. McGlinchery, 137 A. Obertelli, K. Siwek, L. A. Riley, J. A. Tostevin, and D. Weissharr. Re- duction of spectroscopic strength: Weakly-bound and strongly-bound single- particle states studied using one-nucleon knockout reactions. Phys. Rev. C, 772044306, 2008. [78] G. Audi, A. H. Wapstra, and C. Thibault. 2003 atomic mass evaluation (ii). tables, graphs and references. Nucl. Phys. A, 729:337—676, 2003. [79] V. Tripathi, S. L. Tabor, P. Bender, C. R. Hoffman, Sangjin Lee, K. Pepper, M. Perry, P. F. Mantica, J. M. Cook, J. Pereira, J. S. Pinter, J. B. Stoker, D. Weisshaar, Y. Utsuno, and T. Otsuka. Excited intruder states in 32Mg. Phys. Rev. C, 772034310, 2008. [80] C. M. Mattoon, F. Sarazin, G. Hackman, E. S. Cunningham, R. A. E. Austin, G. C. Ball, R. S. Chakrawarthy, P. Finlay, P. E. Garrett, G. F. Grinyer, B. Hy- land, K. A. Koopmans, J. R. Leslie, A. A. Phillips, M. A. Schumaker, H. C. Scraggs, J. Schwarzenberg, M. B. Smith, C. E. Svensson, J. C. Waddington, P. M. Walker, B. Washbrook, and E. Zganjar. )6 decay of 32Na. Phys. Rev. C, 752017302, 2007. [81] G. Neyens, N. Coulier, S. Ternier, K. Vyvey, S. Michiels, R. Coussement, D. L. Balabanski, J. M. Casandjian, M. ‘Chartier, D. Cortina-Gil, M. Lewitowicz, W. Mittig, A. N. Ostrowski, P. Roussel-Chomaz, N. Alamanos, and A. Lépine— Szily. Nuclear spin alignment and static moments of light projectile fragments measured with the level mixing resonance (LMR) method. Phys. Lett. B, 393236—41, 1997. [82] Aage Bohr and Ben R. Mottelson. Nuclear Structure, volume 2. World Scientific, Singapore, 3rd edition, 1998. [83] J. Retamosa, E. Caurier, F. Nowacki, and A. Poves. Shell model study of the neutron-rich nuclei around N = 28. Phys. Rev. C, 55:1266—1274, 1997. [84] S. Nummela, P. Baumann, E. Caurier, P. Dessagne, A. Jokinen, A. Knipper, C. Le Scornet, C. Miehé, F. Nowacki, M. Oinonen, Z. Radivojevic, M. Ramd- hane, G. Walter, and J. Aystéi. Spectroscopy of 34"358i by B decay: sd— f p shell gap and single-particle states. Phys. Rev. C, 632044316, 2001. [85] Ikuko Hamamoto. Nilsson diagrams for light neutron-rich nuclei with weakly- bound neutrons. Phys. Rev. C, 762054319, 2007. [86] J. Simpson, J. Nyberg, W. Korten, et al. Agata technical design report. Tech- nical report, AGATA Collaboration, 2008. [87] Dino Bazzacco. The advanced gamma ray tracking array AGATA. Nucl. Phys. A, 7462248 — 254, 2004. Proceedings of the Sixth International Conference on Radioactive Nuclear Beams (RNB6). 138 [88] [89] [90] [91] [92] [93] [94] [95] [96] [97] [98] [99] W. Hennig, H. Tan, M. Walby, P. Grudberg, A. Fallu-Labruyere, W.K. Warbur- ton, C. Vaman, K. Starosta, and D. Miller. Clock and trigger synchronization of digital data acquisition modules. Nucl. Instrum. Meth. B, 26121000—1004, 2007. H. Tan, W. Hennig, M. Walby, A. Fallu-Labruyere, J. Harris, D. Breus, P. Grud— berg, W.K. Warburton, C. Vaman, T. Glasmacher, P. Mantica, D. Miller, K. Starosta, and P. Voss. Digital data acquisition modules for instrumenting large segmented germanium detector arrays. In Nuclear Science Symposium Conference Record, 2068, page 3196. IEEE, 2008. Zhong He. Review of the Shockley-Ramo theorem and its application in semi- conductor gamma—ray detectors. Nucl. Instrum. Meth. A, 463:250—267, 2001. K. Starosta, C. Vaman, D. Miller, P. Voss, H. Crawford, D. Bazin, T. Glas- macher, P. Mantica, H. Tan, W. Hennig, M. Walby, A. Fallu-Labruyere, J. Har- ris, D. Breus, P. Grudberg, and K. Warburton. Digital data acquisition system for experiments with segmented dectors at National Superconducting Cyclotron Laboratory. In preparation. M.A. Omar and L. Reggiani. Drift velocity and diffusivity of hot carriers in germanium: Model calculations. Solid-State Electronics, 30:1351-1354, 1987. I.-Y. Lee. GRETINA technical note: electron and hole drift velocity in Ge. Technical Report GRT—6-061112, GRETINA physics working group, 2006. L. Mihailescu, W. Gast, R. M. Lieder, H. Brands, and H. Jager. The influence of anisotropic electron drift velocity on the signal shapes of closed-end HPGe detectors. Nucl. Instrum. Meth. A, 4472350 — 360, 2000. NJ. Hammond, T. Duguet, and OJ. Lister. Ambiguity in gamma-ray tracking of “two-interaction” events. Nucl. Instrum. Meth. A, 5472535 — 540, 2005. G. J. Schmid, M. A. Deleplanque, I. Y. Lee, F. S. Stephens, K. Vetter, R. M. Clark, R. M. Diamond, P. Fallon, A. O. Macchiavelli, and R. W. MacLeod. A '7-ray tracking algorithm for the GRETA spectrometer. Nucl. Instrum. Meth. A, 430269 — 83, 1999. L. Milechina and B. Cederwall. Improvements in '7-ray reconstruction with positive sensitive Ge detectors using the backtracking method. Nucl. Instrum. Meth. A, 5082394 — 403, 2003. I. Piqueras, F. A. Beck, E. Pachoud, and G. Duchne. A probabilistic 'y-ray tracking method for germanium detectors. Nucl. Instrum. Meth. A, 5162122 — 133,2004. D.-C. Dinca. Study of the development of shell closures at N = 32,34 and approaches to sub-segment interaction-point determination in 32-fold segmented high-purity germanium detectors. PhD thesis, Michigan State University, 2005. 139 [100] T. Ahn. Digital signal processing for germanium detectors. Master’s thesis, SUNY, Stony Brook, 2004. [101] M. Descovich, I.Y. Lee, P. Fallon, M. Cromaz, A.O. Macchiavelli, D.C. Radford, K. Vetter, R.M. Clark, M.A. Deleplanque, F.S. Stephens, and D. Ward. In- beam measurement of the position resolution of a highly segmented coaxial germanium detector. Nucl. Instrum. Meth. A, 5532535 — 542, 2005. [102] F RIB: MSU/FRIB proposal. http://frib.msu.edu/about/msu-frib-proposal, 2009. [103] K.S. Krane. In N.J. Stone and H. Postma, editors, Low-temperature Nuclear Orientation, chapter 2. Elsevier Science, Amsterdam, 1986. [104] T. Yamazaki. Tables of coefficients for angular distribution of gamma rays from aligned nuclei. Nuclear Data Sheets. Section A, 321 — 23, 1967. [105] R. M. Diamond, E. Matthias, J. O. Newton, and F. S. Stephens. Nuclear alignment in heavy-ion reactions. Phys. Rev. Lett., 16:1205—1207, 1966. [106] K. Alder and A. Winther. Electromagnetic Excitation Theory of Coulomb Ea:- citation with Heavy Ions. North-Holland, Amsterdam, 1975. [107] AH. Wapstra. Nuclear spectroscopy tables. North-Holland, Amsterdam, 1959. 140