. .. q... .. I . .. . . z; i . I. 7 i . . ix!!! 1:: Ex. 5‘3 or .3 I: i129...» . Nautiila . :"mm tun-531w yv-uonu‘un .. n u m ‘V'glu‘lr :‘1. l .uev nun. uh LIBRARY Michigan State University 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 5l08 KIProj/AocaPrelelRC/DateDue.indd EVOLUTION OF NUCLEAR SHELL STRUCTURE: fi-DECAY AND ISOMERIC PROPERTIES OF NUCLEI IN AND NEAR THE fp SHELL By Heather Lynn Crawford A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemistry 2010 ABSTRACT EVOLUTION OF NUCLEAR SHELL STRUCTURE: fl DECAY AND ISOMERIC PROPERTIES OF NUCLEI IN AND NEAR THE fp SHELL By Heather Lynn Crawford One of the fundamental questions in nuclear structure science is how the nucleon single-particle energies evolve with changing proton-to-neutron ratio. The nucleon magic numbers, correctly described by the shell model near the valley of U stability, do not appear to be static across the nuclear landscape. The shifting energies of single-particle orbitals, resulting from variations in nucleon-nucleon interactions such as the tensor monopole interaction, lead to the erosion of some magic numbers, and the appearance of new subshell closures as the driplines are approached. Near the borders of, and within the f p shell, relatively low single-particle level densities lead to a number of distinct regions of changing shell structure. The low- energy structure of nuclei near N 232 and Z220 are stabilized by the presence of a subshell closure at N232, a result of an energy gap between the 1/2123 /2 level and the higher-lying 1/2p1/2 and V1f5/2 levels. An open question, however, is whether or not the continued upward shift of the 1/1 f5 /2 orbital in the Ca isotopes leads to another subshell closure at N234. At slightly higher masses, the migration of the 1499/2 orbital leads to the erosion of the expected N240 subshell closure, and the apparent development of a new region of deformation in the Cr, Mn and Fe isotopes. The question in this region is how quickly collectivity develops as a function of Z, as the 1499/2 orbital drops in energy with decreased occupancy of the proton 1f7/2 orbital below Z228. The U decay and isomeric properties of nuclei in and on the border of the f p shell have been studied at the National Superconducting Cyclotron Laboratory (NSCL) using the combined experimental set-up of the NSCL fl Counting System and 16 detectors from the Segmented Germanium Array. The nuclei studied, 53"54Ca, 54568c, 50K, and 61Cr, were produced in two separate experiments, through the fragmentation of a 76Ge primary beam by a 9Be target. Nuclei were implanted into a double—sided Si microstrip detector, and correlated with subsequent fl decays on an event-by-event basis. Detection of 7 rays in coincidence with the implant events permitted observation of [1.8 isomeric states, while those detected in coincidence with decay events permitted elucidation of the populated levels in daughter nuclei. The low-energy level structures of the neutron—rich 53"(’4’568c isotopes were inves- tigated and compared with the expectations of the extreme single-particle model, as well as with more advanced shell-model calculations using the GXPFl, GXPFIA and KBBG effective interactions. The results confirm the N232 subshell closure, but sug- gest a compression of the V2191 /2-V1 f5 /2 spacing relative to that assumed in current effective interactions, which may preclude formation of a N234 subshell closure in the Ca isotopes. The low-energy structure of 61Mn was probed through the ,8 decay of 61Cr. The structure was investigated for signs of developing deformation, but results suggest that the effects of the approaching deformation—driving 1/199/2 orbital are not yet significant at N236 for the Mn isotopes. The observation of an isomer in 50K permitted investigation of the changing shell structure just below the f p shell. The isomeric structure in 50K suggests that the inversion of the 7r1d3/2 and «281/2 orbitals which is known in 47K may persist above N > 28, which is at present unexpected in sd— f p cross—shell interactions. ACKNOWLEDGMENTS First, I want to thank my advisor, Paul Mantica. Thank you for providing the guidance, and having the patience to see me through my Ph.D. I know I sometimes made it more difficult than was necessary, but thanks for understanding that at times that was what I needed to do. You were always, as you say, my cheerleader, and I am grateful for that. Thank you for everything. I also want to thank Dave Morrissey, Michael Thoennessen, Alexandra Cade and Abby Bickley for serving on my guidance committee, and to acknowledge NSERC of Canada for funding support. Thank you to my fl group-mates, Jill, Josh, and most recently Andrew and Sophia. In particular, Jill and Josh, you led the way for me, and forced me to become a better person in order to meet the standards you set. Thank you for your friendship, and for making graduate school feel more like fun than work. Thanks also to Kei and Geoff, for always being sounding boards, and gamely answering all of my random questions. Nothing that I’ve accomplished would have been possible without my group. In my four years here, I’ve made friends I’ll have for life, better friends than I deserve. Thank you for all of the happy-hours, the softball (even if I was terrible!), and the random conversations over the last four years. In particular, I’d like to thank a few very good friends - you know who you are. Thank you for believing in me when I couldn’t seem to believe in myself, and for pointing out the pitfalls with my umbrella factory back-up plan. You cannot know what your friendship and support has meant to me. Finally, I have to thank my family. Mom, you wanted credit for a small piece of this work, but you deserve so much more than that. My family is my rock, and even 2301 miles away, I knew you were always right there with me. Thank you for always listening and for your blind faith in me. iv Contents 1.1.1 Evolution of nuclear shell structure ............... 1.2 Describing the structure of nuclei .................... 1.2.1 Structure near closed shells ................... 1.2.2 Collective structures and nuclear deformation ......... 1.3 Experimental observation of shell closures and nuclear structure . . . 1.3.1 E(2?) and B(E2 : 2?" ——> 01*) in even-even nuclei ........ 1.3.2 Tracking single-particle energies and the extreme single—particle model ............................... 1.4 Re—ordering of single-particle states in neutron-rich f p shell nuclei . . 1.4.1 N232 and N 234 subshell closures ............... 1.4.2 N240 shell closure and intrusion of the neutron 1g9/2 orbital 1.5 Motivation for the measurement ..................... 1.6 Organization of Dissertation ....................... 2 Experimental Techniques ...................... . . . 2.1 fl decay .................................. 2.1.1 fl-decay half-lives ......................... 2.1.2 ,B-delayed neutron emission ................... 2.1.3 Selection rules and log f t ..................... 2.2 7-ray decay ................................ 2.2.1 Selection rules and lifetimes ................... 2.2.2 Isomeric 7-ray transitions .................... 2.2.3 Internal conversion ........................ 2.3 Summary of the decay of exotic nuclei ................. 3 Experimental Setup ........................... 3.1 Isotope production and identification .................. 3.1.1 NSCL experiment 05101 ..................... 3.1.2 NSCL experiment 07509 ..................... 3.2 Detector apparatus for fl-decay experiments .............. 3.3 6 Counting System ............................ 20 22 22 25 31 32 37 39 46 46 49 50 52 52 3.3.1 Electronics ............................. 55 3.3.2 BCS calibrations ......................... 60 3.4 Segmented Germanium Array ...................... 62 3.4.1 Electronics ............................. 62 3.4.2 SeGA Calibrations ........................ 64 3.5 Data analysis: Correlations and data fitting .............. 69 3.5.1 Implantation-decay correlations ................. 70 3.5.2 Decay curve fitting methods ................... 76 4 Experimental Results .......................... 80 4.1 Low-energy structure of neutron-rich 218C isotopes ........... 81 4.1.1 6 decay of 53Ca to 53Sc ..................... 83 4.1.2 ,6 decay of 54Ca to 54Sc ..................... 86 4.1.3 B decay and isomeric structure of 54 Sc ............. 89 4.1.4 6 decay of 5680 .......................... 97 4.1.5 Isomerism in 5680 ......................... 104 4.2 Isomeric Structure of 50K ........................ 109 4.3 Structure of 61Mn from [3 decay of 61Cr ................ 113 5 Discussion ................................. 121 5.1 Interpretation of the low-energy levels in neutron-rich 2180 ...... 121 5.1.1 Low—energy structure of odd-A 53Sc ............... 122 5.1.2 Low—energy structure of even-A Sc isotopes ........... 123 5.2 Re-ordering of proton single-particle states in neutron-rich K isotopes 131 5.2.1 Systematics of the 19K isotopes around N228 ......... 131 5.2.2 Comparison to shell-model calculations using sd— f p cross-shell interactions ............................ 138 5.3 Onset of collectivity in the 25Mn isotopes ................ 139 5.3.1 Systematics in odd-A Mn .................... 140 5.3.2 Even-A Mn isotopes ....................... 143 5.3.3 Deformation in comparison to neighboring 24Cr and 25Fe iso— topic chains ............................ 145 6 Conclusions and Outlook ........................ 147 6.1 Conclusions ................................ 147 6.2 Outlook .................................. 149 6.2.1 Open questions in the f p shell .................. 149 6.2.2 The NSCL 6 Counting System with Digital Data Acquisition .’ 151 Bibliography ................................. 154 vi List of Tables 2.1 2.2 2.3 3.1 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 5.1 5.2 Classification of fl-decay transitions ................... 38 7-ray selection rules ............................ 40 Weisskopf single-particle lifetime estimates for 7 decay ........ 42 y-ray Transitions Used in the Recalibration of SeGAll ........ 67 fi-delayed 'y rays assigned to the decay of 5" Sc ............. 90 568C fi—delayed 'y rays ........................... 98 fl and fin contributions to the decay of 568C .............. 104 Prompt 568C 7 rays ............................ 105 Weisskopf half-life estimates for prompt 5680 7-ray transitions . . . . 109 'y rays observed in coincidence with 50K implantation events ..... 110 Weisskopf half-life estimates for prompt 50K 7-ray transitions . . . . 113 fl-delayed 7-raygtransitions in the decay of 61Cr ............ 116 Application of the Pandya transform relating neutron particle and hole states in 50’52Sc .............................. 126 Calculated transition probabilities for the 31” —+ 4? transition in 54Sc 130 vii List of Figures 1.1 1.2 1.3 1.4 1.5 1.6 1.7 Observables for atomic and nuclear shell closures ........... 2 Evolution of the nuclear shell model ................... 5 Single—particle states in Egnglg—g .................... 11 Illustration of parameters in j-j coupling ................ 15 Parabolic rule for p—n multiplets ..................... 16 Systematics of E (2?) across isotopic chains ............... 19 Monopole shift of the V1f5/2 orbital with increasing 7r1f7/2 occupancy 23 1.8 Systematics of E(2f) and B(E2 : 2? —+ 01+) across isotopic chains around N 232, N 234 ........................... 24 1.9 Systematics of E(2i’") around Z240, N254 and N240, Z228 ..... 27 1.10 Nilsson Diagram for 502N , Z 220 .................... 30 3.1 Schematic of the NSCL Coupled Cyclotron Facility .......... 47 3.2 Particle Identification Plots for NSCL Experiments 05101 and 07509 . 51 3.3 Schematic of the Beta Counting System Detectors ........... 53 3.4 Schematic of DSSD Electronics ..................... 55 3.5 Schematic of SSSD Electronics ...................... 57 3.6 Schematic of PIN Detector Electronics ................. 58 3.7 Schematic of Trigger Electronics ..................... 59 3.8 228Th and 908r DSSD Calibration Spectra for Illustrative Front and Back Strip ................................. 61 3.9 Schematic of the Geometric Layout of the SeGA Detectors ...... 63 3.10 Schematic of Electronics for SeGA Detectors .............. 64 viii 3.11 3.12 3.13 3.14 3.15 3.16 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 Residuals for Energy Calibration of Individual SeGA Detectors . . . . 66 Residuals for Total SeGA Array Energy Calibration .......... 66 Recalibration of SeGA 11 ........................ 68 SeGA Efficiency Calibration ....................... 69 Position dependence of rates in NSCL experiment 07509 ....... 74 Comparison of 55Ti ,B-delayed 7-ray spectra and half-life curves for analysis including pixels at the edge and over the full DSSD ..... 75 7-ray spectrum collected within 20 ,us following a 54Ca implantation . 82 fi-delayed 7-ray spectrum in the range 0-3 MeV following the decay of 53Ca .................................... 84 Decay curves for the B decay of 53Ca .................. 85 Decay scheme for decay of 53Ca to states in 53Sc ............ 86 fi-delayed 7-ray spectrum following the decay of 54Ca ......... 87 Decay curves for the decay of 54Ca ................... 88 Decay scheme for the decay of 54Ca to states in 54Sc ......... 89 fl-delayed 7-ray spectrum for the decay of 54Sc ............. 90 fl-decay curves for the decay of 5480, with and without the requirement of a coincident decay 7 ray ........................ 92 Decay Scheme for the decay of 54Sc to states in 54Ti ......... 93 77 coincidence spectra for fi-delayed 7 rays following the decay of 54Sc 94 Prompt 7-ray spectrum following implantations of 548C ........ 96 Decay curve for 110—keV isomeric transition in 5480 .......... 96 fi-delayed 7—ray spectrum following the decay of 56Sc ......... 98 7-gated decay curves for the decay of 5680 ............... 100 Decay scheme for 5680 to states in 56Ti ................. 101 77 coincidence spectra for the decay of 568C .............. 102 7-ray spectrum following implantation of 568C ............. 105 Decay curve for the 7-decaying isomeric state in 56Sc ......... 106 77 coincidences for prompt 7 rays in 568C ............... 107 ix 4.21 4.22 4.23 4.24 4.25 4.26 4.27 4.28 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 6.1 Decay curves for the 6 decay of 568C .................. 108 7—ray spectrum following 50K implantation events ........... 111 77 coincidence spectra for isomeric decay of 50K ............ 112 Proposed low-energy level scheme for 50K ............... 114 Decay curve for the decay of 61Cr to states in 61Mn .......... 115 B—delayed 7-ray spectrum following the decay of 6’Cr ......... 116 Decay scheme for the decay of 61Cr to states in 61Mn ......... 118 77 coincidence spectra for decay of 61Cr ................ 119 Level structure of 53Sc and schematic of coupling to states in 52Ca . . 123 Experimental level scheme for 50’52’54’5680 ............... 125 Shell model calculations for the even-A Sc isotopes using the GXPFl, GXPF 1A and KB3G effective interactions ............... 127 Energy separation of the 3 / 2+ and 1 /2+ states in the odd-A K isotopes from N220 to N230 ........................... 133 Known states in 48K, and their interpretation within the independent particle model framework ......................... 135 Low-energy structure of 50K known from decay of the 172-keV isomeric state .................................... 137 Comparison to shell model results for theodd-A 25Mn isotOpes . . . . 142 Low-energy level structure of the even-A 60’62’64Mn isotopes ..... 144 Sgn values for the isotopes surrounding Fe near N 240 ......... 145 Resolution of 228Th spectrum collected using DDAS vs. analog data acquisition ................................. 153 Chapter 1 Introduction 1.1 Nuclear shell structure The periodicity in the chemical nature of the elements has been represented on the periodic table for more than a century, and is now understood with the essential con- cept that the electrons surrounding the atomic nucleus are arranged in discrete energy levels, or shells. One of the most direct indicators of this electron shell structure in atoms can be seen by considering the first ionization energy, the energy required to remove an electron from the neutral atom, as a function of atomic number, Z. Shown. in the top panel of Fig. 1.1, the observed trend of ionization energy has discontinuities at specific values of Z. These discontinuities are understood as arising from the un- derlying electron shell structure. After filling a given energy orbital, the next electron is placed in a higher energy orbital, and the energy required to remove it decreases substantially. Analogous to the observed trend in atomic ionization energies, systematic ex- amination of many physical properties of nuclei has revealed periodic trends. Near the valley of stability, bulk properties related to the nuclear mass are suggestive of enhanced relative stability for nuclei with “magic” numbers of protons or neutrons, corresponding to 2, 8, 20, 28, 50, 82, and 126. One can consider, for instance, the one- 302 E 710 E3 , 18 34 320,1 54 86 ES 1 i l 8 15a 1 '5 10 s i .3 5 ,' . . .2 ‘NeiK. RbCF 0 20 40 60 80 100 Atomic Number (Z) 9 9? 20 0» 8§- 128 5 7E— i 50 82 ms: 6; ,1 l l 126 <1 5E— i\ . ll, 43” "Wit 3%- "1” “:2, .. 15+..11111411.1...111111111..11...1 0 20 40 60 80 100120140160 Neutron Number (N) Figure 1.1: (Top) Systematics of the first ionization energies of the atomic elements from hydrogen (Z21) to nobelium (Z2102). Data were taken from Ref. [1]. Discon— tinuities correspond to atomic electron shell closures. (Bottom) Systematics of the difference in neutron separation energy for even-even nuclei 9X N and their even-odd neighbors. Data were taken from Ref. [2]. Solid lines connect isotopic chains. Dis- continuities provide evidence for nuclear shell structure, representing closed neutron shells. neutron separation energy for a nucleus with mass number A, and atomic number Z, which is defined as follows: ‘n(N) = B(éXN)—B(A‘IZX~_1) = [MM-12m-» — M<§XN> + Mac? (1.1) Here, B (9X N) represents the nuclear binding energy, B(ZXN) = [ZMH + N412 — MVz‘XNHCQ, (1-2) where M H is the atomic mass of hydrogen, Mn is the neutron mass, M (QX N) is the atomic mass of the given nucleus, and contributions from the differences in the elec- tron binding energies are neglected. The neutron separation energy is analogous to the atomic ionization energy, representing the energy required to remove one neutron from the nucleus. The systematics of the neutron separation energy show periodicity suggesting an underlying shell structure of the nucleus. These shell effects are en- hanced by considering the difference in Sn values between even N nuclei and their N + 1 neighbors, defined as follows: A5,, = 5,,(N) — STAN +1) = [M(A"IZXN_1 + M(A+IZXN+1) — 2M(§XN)]C2. (1.3) The neutron separation energy differences, AS", are plotted in the bottom panel of Fig. 1.1 as a function of neutron number for the isotopes with even numbers of protons and neutrons up to fermium (Z2100). The discontinuities corresponding to neutron ‘magic numbers’ are apparent; an analogous examination of proton separation energy systematics similarly highlights the same ‘magic numbers’ for protons. Given the success in describing atomic structure in terms of electron shells, a simi- lar approach has been employed in nuclear physics to describe the structure of nuclei. However, in the atomic case, electrons move in an “external” field created by the well- understood Coulomb force. As such, the Schr6dinger equation for the atomic system can be solved exactly, and electron orbital energies accurately calculated, reproduc— ing the experimentally observed shell closures. The picture is more complicated in the nuclear case, as nucleons move within a field created by the surrounding nucleons, and the analytical form of the force involved, namely the strong force, is not defined at present [3]. The success of the nuclear shell model rests primarily with the choice of the po- tential assumed to confine the protons and neutrons within the nucleus. Historically, a starting point has been the harmonic oscillator potential. The energy levels aris- ing from the solution of the Schr6dinger equation assuming a harmonic oscillator potential are shown in Fig. 1.2(a). This simple potential reproduces the lowest shell closures, at 2, 8 and 20, however, the higher closures are not in agreement with obser- vation. A more realistic potential, bounded at r212, is considered to provide a better approximation. The Woods-Saxon potential has a shape intermediate between a square well and a harmonic oscillator, and takes the form __V0 VT 2 , U 1+exp1512—‘01 (1.4) where the parameter R is the mean nuclear radius, 0. represents the diffuseness of the nucleus, and V0 is the depth of the potential well, usually adjusted to a value of approximately 50 MeV. Solution of the Schr6dinger equation assuming this potential again reproduces the lowest shell closures at 2, 8 and 20 but still fails to accurately describe the higher lying magic numbers [see Fig. 1.2(b)]. The missing piece of the puzzle was put in place in the 1940’s, when the inclusion of a spin-orbit potential was shown to properly separate subshells and reproduce the experimentally observed magic numbers [5,6]. The spin-orbit potential takes the form: [V30 = V30(r)l-.' g, (1.5) where l represents the orbital angular momentum, and s? the intrinsic nucleon spin. The I: 5" term accounts for the interaction of the nucleon spin and its orbital motion, 2f ........ 1117972 1h =:::::j: 1h 35 ,,,,,,,,,,,,,,,,, 3511/12/2 33, 2d, 1g 2d -=:::::::ZZII:::: 333/2 .......... ‘lsifi 18—53:: ............. [Igg/Z ..____J40 2191/2 2p,1f 2p -====333.I.I£l-I-I-I- ''''' "fi 1f5/2 If =22: ~~~~~~~ \2p3/2 1f7/2 23,1d 23 ...... .. Ids/2 1d .............. 131572 11’ 1 ------------- 1P1/2 @ p @ """"""" 1P3/2 IS IS """""""""" 151/2 Harmonic Woods-Saxon Oscillator Woods-Saxon + Spin-Orbit (a) (b) (C) Figure 1.2: Evolution of the nuclear shell model to represent the experimentally ob— served nucleon “magic” numbers. The first approximation with a harmonic oscillator potential evolved to the more physical Woods-Saxon potential, and finally the in- clusion of a spin—orbit potential to reproduce the experimental observations. Figure adapted from Ref. [4]. and leads to a removal of the l-degeneracy of states, or a splitting of states with l > 0, based upon the total angular momentum (3’ 2 f’+ 5) projection. This reordering of the subshell states is shown in Fig. 1.2(c), and the shell closures that develop agree fully with observations of the magic numbers near the valley of 6 stability. The nuclear shell model has been further developed and extended to include more subtle aspects of nuclear structure. The basic tenets discussed here, which assumes a spherically symmetric potential, remains the starting point for understanding the level structures of spherical nuclei near closed shells. However, recent experimental evidence suggests that the shell structure of nuclei, i.e. the ordering and energy separation of single-particle states, is not static across the nuclear landscape, but rather evolves with varying proton-to—neutron ratios. The nature and origins of this re-ordering of single-particle states is the topic of the following section. 1.1.1 Evolution of nuclear shell structure Across the nuclear chart, as progressively more exotic nuclei are probed, there is increasing evidence for an evolution of shell structure. The ordering of single—particle levels as shown in Fig. 1.2(c) has been observed to change slightly with proton-to- neutron ratio, or isospin projection, T2 2 (N — Z ) / 2, which can change the observed nuclear properties. At Z24, in the relatively light isotopes, breakdown of the N28 shell closure has been observed in the neutron-rich nucleus liBe [7]. The last two neutrons in 12Be are expected to occupy the 1191/2 orbital. However, spectroscopic factors for the removal of a neutron from 1”Be provide direct evidence of significant sd—shell intruder contributions to the ground state wavefunction, and the disappearance of the N 28 shell closure in this isotope. At slightly heavier masses, the so—called ‘island of inversion’ at N 220 away from stable nuclei, also gives evidence of evolving shell structure. Nuclear ground states have been shown to be dominated by 1/(sd)‘2(fp)2 intruder configurations, suggesting the breakdown of the N 220 shell closure [8,9] for neutron—rich nuclei with 10 3 Z s 12 and 20 S N g 22. Shell evolution is also seen in the apparent erosion of the N 228 shell closure in the neutron rich 168 isotopes [10], and the development of a new subshell closure in neutron-rich 20Ca [11,12], 22Ti [13] and 24Cr [14] isotopes at N 232. At even higher masses, however, the shifting of single-particle levels is less likely to result in a change in the position of the major shell closures, due to the higher single-particle level densities. None the less, the evolution still occurs — in the region of N 250-82, and Z 240-50 there is a reordering of neutron single-particle levels as the 197/2 orbital rapidly drops in energy between 91Zr and 1318m [15,16]. These examples highlight the ubiquity of evolving shell structure over the nuclear landscape. It is clear from these instances that there are effects and interactions be— yond the scope of the simple shell model which can greatly alter single-particle level energies and ordering. Understanding the origins of and accounting for experimen- tally observed changes in shell structure are critical for the development of nuclear theory models — an understanding of the phenomena already observed will enhance the reliability of theoretical predictions. One nucleon-nucleon interaction, believed to be a possible cause of shell evolution, is discussed next. Proton-neutron monopole interaction One of the suggested reasons for the evolution of nuclear shell structure is the spin- isospin dependent part of the nucleon-nucleon interaction [17]. This is a tensor monopole force, which is additional to the spin-orbit interaction discussed previously, and can alter the effective nucleon single—particle energies (ESPEs). These ESPEs account for the mean effects from all other nucleons, on a nucleon in a given single-particle orbital. The single-particle energy of an orbital j is determined by its kinetic energy, and the interaction with the nucleons in the core that produce the mean field [18]. When nucleons are added to a different orbit j’, the single—particle energy of the original j orbital is altered. The nucleons in j’ and j will couple to form states with different total angular momentum, J, but in considering changing shell structure, it is the shift of single-particle orbitals irrespective of the details of the coupled J states that is interesting. Thus, the J dependence is averaged out, and only monopole effects are considered. The monopole component of a nucleon-nucleon residual interaction V for the effect of the nucleons in j' on the orbital j takes the form: 2 1 . ., . ./ r VT, = 2J( J + )< .7] IVIJJ >JF, (1.6) .7] 2J(2J +1) where < jj'IVIjj’ >JT is the matrix element for j and 3" coupling to a state with spin J, and isospin T. With this interaction, it is possible to evaluate, for example, the effect of nn(j’) neutrons in the orbital j’ on the single particle energy of the proton orbital j. The shift in the single-particle energy of the proton orbital j, denoted Aep(j), resulting from the presence of 12,, neutrons in the j’ orbital is given by [18]: AW) = g 13.370 + 1957117222). (1.7) With the inclusion of the tensor monopole shift, a new effective single-particle energy can be defined, which is dependent on the nucleon configuration, and thus changes with proton-neutron ratio. The same arguments hold for the inverse case, the effect of protons in a given orbital on neutron single-particle energies. It should be noted that protons (neutrons) in an orbital j will also affect the energy of protons (neutrons) in another orbital 3". However, only T 2 1 interactions, which are weaker than T 2 0 interactions [18], are relevant in proton-proton or neutron-neutron interactions. Thus, these effects are less significant than those arising from proton-neutron interactions. As discussed by Otsuka et al. [18], there is a general rule for the way in which the proton-neutron tensor monopole force affects the single-particle energies. An attrac- tive interaction is observed between opposite isospin particles (protons and neutrons) in j< 21—1/2 and j") 2 l’+1/2 orbits (and vice versa, j> 2 l+1/2 and j’< 2 l’—1/2 orbits), while the interaction between protons and neutrons in j< 2 l — 1 / 2 and j’< 2 l’ — 1 / 2 orbits (and similarly j> 2 l+ 1 / 2 and 3") 2 l’ + 1 / 2 orbits) is repulsive. The strength of the interaction is also maximized for I ~ 1’, so a stronger attractive interaction between protons and neutrons in spin-orbit coupling partner orbitals is expected [17]. Nearly all of the examples of evolving nuclear shell structure discussed previously can be partially explained in the framework of the proton—neutron tensor monopole interaction. The disappearing N 28 shell closure in 1gBe results from the low occu- pancy of the 7r1p3/2 orbital that causes the u1p1/2 orbital to migrate higher in energy, eroding the N 28 gap. The island-of-inversion nuclei can be largely explained by the decreasing attractive interaction between the u1d3/2 orbital and the 7rld5/2 orbital as protons are removed from the 7rld5/2 level. This reduction causes the V1d3/2 orbital to rise in energy, eroding the N 220 shell closure, and producing a N216 subshell closure for the lightest nuclei in the region. The region surrounding N 234, Z220 also experiences changes in neutron level ordering driven largely by a reduced monopole interaction between the 7r1f7/2 orbital and the u1f5/2 orbital, while the heavier nu- clei near N250 show evidence of shell evolution arising from the changing monopole interaction between the 2199/2 orbital and the spin-orbit partner 11197/2 orbital. 1.2 Describing the structure of nuclei While one of the goals of nuclear science is to create a single theoretical framework to adequately describe the entire nuclear landscape, the variation in nuclear structure across the chart of the nuclides is significant. The nuclear shell model provides a basis for understanding the overall structure of all nuclei, but important simplifications can be made in certain regions of the nuclear chart. The most basic shell model interpre- tation, the independent particle model, provides an excellent qualitative description of the low-lying states in nuclei one nucleon from closed shells [19]. A few nucleons from a closed shell, this model can be extended, and residual nucleon-nucleon inter- actions taken into account, to successfully describe the low-energy structure of nuclei. However, for nuclei far from closed shells, these types of microscopic shell model cal- culations become intractable. Geometric models provide a better description of the growing collective behavior of mid-shell nuclei, as macroscopic deformations play an increasingly important role in determining the relevant nuclear structure. 1.2.1 Structure near closed shells The independent particle model The so—called “independent particle model” describes a nucleus in terms of non— interacting particles orbiting within a central spherically symmetric potential, which is itself produced by all the nucleons. Within this model, nucleons are assumed to sequentially fill the single-particle levels generated by this nuclear potential, in much the same way that electrons are assumed to fill atomic energy levels. Also in anal- ogy to the atomic picture of valence, in which filled electron shells do not contribute to the chemical properties of an element, filled nucleon orbitals are assumed to be inert within the structure of the nucleus. A closed shell will always contribute a to- tal angular momentum and parity, J 7' 2 0+ in the nuclear case, and so the ground state angular momentum of a nucleus with only one valence particle in a shell n1 j is determined entirely by the single valence nucleon as J 2 j , with parity 7r2(-1)‘ [19]. The independent particle model provides a framework for predicting the ground state spin and parity of nuclei one nucleon removed from a closed shell, and has also proven successful in predicting the low-lying excited states in these nuclei. Low- energy levels in such nuclei are formed by simply elevating the valence particle from the ground state orbital to a higher single-particle state. An example of the success of this approach can be found in the structure of 2gng127, as shown in Fig. 1.3. With one valence neutron above the N 2126 shell closure, the ground state in 209Pb is of» served to have J “29/ 2+, as expected with a (12299/2)1 configuration. The other six levels corresponding to promotion of the single valence neutron to the other single- particle levels above the N 2126 shell closure are all evident in the low-energy struc- ture [see Fig. 1.3(b)]. The neutron spectroscopic factor is near unity for each of these seven states. The spectroscopic factor is a measure of the contribution of a given single-particle configuration to a nuclear state, as determined by comparison of ex- perimental reaction cross-sections with calculations assuming pure single-particle con- 10 ll 8, 8 3d 5 4s??? /2g7/2 E ___._...=3/2+ 1.09 1:71:15/2’ :395/2 3; _____7/2+ 1.05 89/2" ‘ ll11/2 "4 + m2. ——-————-1/2 0.98 -. a ”1113/2 +§ ___5/2+ 0.98 1 .45 15/2- 0.58 ~ .. N 1h ml’ ——11/2+ 0.86 11/2‘""".""’f - . , ~ - 0- ———9/2+ 0.83 2 9 7‘ V 821319127 (a) (b) Figure 1.3: (a) Schematic representation of the nucleon configuration for the one- valence neutron nucleus 209Pb. (b) Experimentally observed levels of 209Pb below 3 MeV [2]. Black levels correspond to single-particle states arising from occupation of the seven single-particle levels above the N 2126 shell gap. The purity of these single particle configurations is supported by the near-unity spectroscopic factors. Grey levels correspond to levels with configurations beyond the scope of the independent particle model. figurations [20]. Thus, spectroscopic factors near 1 indicate the relative purity of the single-particle states in 209Pb. Strictly speaking, the independent particle model is applicable only to nuclei with one valence particle (or hole) outside of a closed shell. However, the fundamental ideas of this model can be extended to nuclei with a few valence particles (or holes) outside of a closed shell inert core. The low-energy structures of such nuclei can be understood by including a residual nucleon-nucleon interaction between valence nucleons above the inert core. 11 Few nucleon configurations A few nucleons outside of a shell closure, the independent particle model can be extended to provide an adequate description of the low-energy states within a nucleus. The simplified idea of the nucleus as consisting of valence nucleons occupying single- particle states above an inert core of closed shells provides a basis for interpreting the structure solely in terms of the interactions among valence particles. However, with more than one valence particle, it is important to consider the residual nucleon- nucleon interaction, which largely determines the observed structure of few-nucleon systems. Two nucleons occupying orbitals with quantum number nllljl and 7121ng can couple, in the most general case, to produce a multiplet of states with [j] — jg] S J g I j1 + jg]. If the nucleons are identical and occupy the same orbital, the requirement of an antisymmetric wavefunction limits J to values from 0 to (2j-1), and without a residual nucleon-nucleon interaction, all J states in a multiplet are expected to be degenerate in energy. Under the effect of a residual interaction, V12, this degeneracy is broken, and each state J experiences an energy shift AE(j1j2J), given by [19]: . . 1 . . . . A13(31J2J) = \/—2J;+_1: < J1J2J||V12||J192J >, ' (1-8) where the double-bar indicates summation over all M states. The exact form of the residual nucleon-nucleon interaction, V12, is not known, although it is expected to have a strong attractive short-range component, as well as a long-range component which is of particular importance in producing collective nuclear properties [19]. One example of a commonly used interaction is the so—called ‘pairing plus quadrupole’ interaction, which takes the following form: Vppq 2 Vim-7. + nr¥r3 P2(cos 9), (1.9) 12 where 0 represent the angle between the radius vectors to each particle (F1 and F2), and n represents the strength of the quadrupole interaction. me-r represents the short-range attractive pairing interaction, while the quadrupole component is a first approximation to the long-range component. An even simpler approximation for the residual nucleon-nucleon interaction is a simple delta-function interaction, which only represents the short-range interaction and has been shown to reproduce many of the results from more realistic interactions. The relative spacing of states within proton- proton (p-p) or neutron-neutron (n-n) multiplets, for example, can be well-described by assuming a delta-function residual interaction. Taking the form of the residual interaction to be V12 2 V06(F1 — F2), where F1 and F2 are the radial coordinates of the two nucleons, and V0 is the strength of the interaction, Eqn. 1.8 becomes [19]: 2 . . . . J1 J2 J _ AE(]1]2J) 2 —V0FR(2]1+1)(2]2 +1) 1 1 1f 11 + l2 — J even 2 ‘2‘ 0 2 0 if l1 + l2 — J odd, (1.10) where F3 is a function of the radial coordinates (i.e. 7111111212) only. Thus, only half of the states in a given p—p or n—n multiplet will be shifted under the influence of a delta-function residual interaction. Further information can be gained by considering that the isospin portion of the nuclear wavefunction for a p-p/n-n configuration is symmetric (T21), which means that the Pauli principle requires the wavefunction to be antisymmetric in either spin or space. A symmetric spin configuration 821 thus requires the spatial wavefunction to be antisymmetric. However, a delta-function interaction between two nucleons vanishes if F1 2 F2, and so will not alter the energy of states with S21, and will only affect S20 states. Given that only half of the states of a p-p or n—n multiplet will be affected by a delta-function residual interaction, as indicated from Eqn. 1.10, a general rule arises describing the splitting of energy levels in p—p and n-n multiplets. If the parity is positive, even J value are lowered in energy, 13 with the extreme J value (Jmin 2 Ijl — jg] or Jmam 2 | j1 + jgl) lowered the most, while for negative parity, odd J values are lowered. If jl 2 jg, as in the case of even- even nuclei, the parity of the expected states is positive, and this interaction results in a dramatic lowering of the 0+ state. This interaction is, in fact, the underlying explanation in this model for the fact that even-even nuclei have 0+ ground states [19]. The residual interaction in p-p/n-n multiplets is restricted to T21 interactions, which simplifies the possible effects of the residual interaction on the relative spacing of states in a multiplet. Proton-neutron (p—n) multiplets however can also include T20 interactions, which have been observed to be stronger than the T21 interac- tions [19]. Thus, a different approach is more appropriate for describing the energy splitting of states within a p—n multiplet under the effect of the nucleon-nucleon resid- ual interaction. Insight into the effect of a residual interaction on the relative ordering of low-lying levels in few nucleon systems can be gained by considering a multipole decomposition of the interaction. Any interaction that depends on the separation be— tween two particles, as the residual interaction is expected to, can be expanded in terms of multipoles [19]. Such a decomposition allows, among other important re- sults, the derivation of the parabolic rule [21] to describe the energies of states in two particle p—n multiplets. Within a decomposition, each multipole, It has an angular dependence, Pk(cos 6), where Pk(x) are the Legendre polynomials, and 0 is the angle between the angular momentum vectors for each particle, as shown in Fig. 1.4, and given by the following: J(J+1)—j1(j1 +1) —2202 +1). 2¢jirj1+1>1202 +1) (1.11) cos0 2 The monopole component, k20, is constant, and thus affects all states equally, sim- ply giving an overall energy shift to the entire multiplet of states. This monopole interaction is in fact the same effect discussed previously in Section 1.1.1. The quadrupole component, [£22, is proportional to P2(cos 9) 2 915(3 cos2 0), and is 14 Figure 1.4: Schematic representation of the coupling of two nucleon angular momenta (jl, jg) values in a geometric representation to a final J, and the definition of the angle 0 between them. maximized for 0 2 0° and 180°, which corresponds to the alignment or antialignment of the angular momentum vectors of the two constituent particles. Thus, the attractive particle-particle quadrupole interaction lowers the energies of extreme J values the most, and raises intermediate spins. The J dependence on the energy splitting between states in a given p—n multiplet is given by the following expression [19], which arises from the Pg(cos 0) proportionality of the 1:22 multipole combined with Eqn. 1.11 for the angle 6: [J(J + 1) moi +1) — 1202 +1112 42101 +1)j2(j2 + 1) oc A[J(J +1)]2 + BJ(J + 1) + C, (1.12) 13501227 ) 0‘ where A, B and C are independent of J. This relationship takes the form of a parabola in J (J + 1), and despite the omission of higher multipoles of the residual interaction, has been shown to provide a good first approximation to experimental data. Figure 1.5 illustrates the parabolic rule applied to the proton-neutron (Half/21112], /2) multiplet in 122Sb, and the proton-neutron hole (7r f71/21/ f7712) multiplet in 48Sc. The multiplet 15 J J 0.1??4§§Z$2 012.311??? 0’ (a) 2’ 0’) 485C (“fi/2Vf2i2) Relative Excitation Energy (keV) Relative Excitation Energy (MeV) 1 1 ’2sz (“87/2Vh11/2) 0_ , -300» . . . . 1 . . . . . . 0153045607590 0102030405060 J(J+1) J(J+1) Figure 1.5: Examples of the qualitative success of the parabolic rule describing the level ordering of p—n multiplets. (a) illustrates the p—n particle—particle parabola in 122Sb [22], while (b) illustrates the p-n particle-hole parabola in 48Se [23]. Dots rep— resent experimentally measured energies, while the solid line is a best fit parabola. under consideration in the case of 488C is actually a particle-hole p—n multiplet. The residual interaction in this case is repulsive, and thus the extreme J values corre- sponding to high p—n overlap are pushed higher in emery, while the intermediate spins are lowered. Moving beyond the case of two valence particles, within the single—particle model it can be assumed that nucleons in a partially filled single-particle orbital will pair to .17'20+ where possible [19]. Under this assumption the low-lying states of any nucleus near a closed shell can be interpreted, in the independent particle model, as arising from either a single valence nucleon configuration, or a p—n multiplet outside a core of fully paired nucleons. 1.2.2 Collective structures and nuclear deformation The description of nuclei that are more than a few nucleons away from closed shell con- figurations becomes progressively less practical purely within the shell model frame- work. These calculations rapidly become intractable for mid—shell nuclei due to the 16 large numbers of valence nucleons. Additionally, the simplifications made by assum- ing inert closed shell cores become less appropriate. Significantly more complex ex- citations develop in these nuclei, which are better described in a more macroscopic framework [19]. Collective interactions between nucleons typically lead to quadrupole (oblate or prolate) deformation of the overall shapes of mid-shell nuclei. The extent of defor- mation is denoted by a quadrupole deformation parameter, fig. fig 2 0 in the case of spherical nuclei, while this deformation parameter takes on positive values for pro- late deformations, and negative values for oblate deformation. Models which attempt to describe deformation of the nuclear shape, such as the Nilsson model, provide a framework in which to understand the structures of such deformed mid-shell nuclei. The Nilsson model is, in many ways, similar to the spherical shell model discussed earlier, but follows the variation of the energies of the spherical shell-model orbitals in a deformed nuclear potential - it is the single-particle model applicable to deformed nuclei [19]. While the previous sections have discussed a theoretical framework for under- standing the structure of nuclei near shell closures, the shell-model framework re— quires knowledge of the location of shell closures, and the ordering of single-particle states. The following section discusses some of the experimental observables which can provide such information. 1.3 Experimental observation of shell closures and nuclear structure One of the most critical pieces of information needed to allow development of nuclear structure models and theory is the location of shell and subshell closures across the nuclear landscape. As discussed previously, the location of shell closures is not static across the nuclear chart, and an ongoing experimental focus in nuclear physics is to 17 locate these shell closures in previously unknown nuclei, and track the migration of single-particle orbitals which lead to this dynamic shell structure. A number of experimental observables are sensitive to the presence of shell closures across the nuclear landscape. Nuclear masses and binding energies provide the most fundamental mapping of structural evolution across the nuclear landscape. Changes in the trends of neutron-separation energies across isotopic chains, for example, can indicate the presence of neutron shell closures, or the development of deformation, depending on the nature of the observed change. Ground state nuclear moments are also good indicators of shell closures. The nuclear quadrupole moment provides a measure of the nuclear charge distribution, and thus the nuclear shape. Near closed shells, nuclei are expected to be spherical, and ground state quadrupole moments should be small. Away from shell closures, nuclei may be axially deformed, and have large magnitude quadrupole moments. Another indicator for shell closures comes from the excited states of nuclei. Certain characteristics of excited state configurations are expected near closed shells, and such states can be used to identify shell closures, when compared with the systematic behavior in neighboring nuclei. The properties of the first 2+ excited state in even-even nuclei, for instance, are sensitive to the presence of shell closures, as is discussed in the following section. 1.3.1 E 2+ and B E2 : 2+ —> 0+ in even-even nuclei 1 1 1 Across the nuclear landscape, there are certain pervasive structural patterns that emerge — the near-spherical shape of nuclei near closed shells and the dominance of quadrupole collective nucleon excitations near midshell, for instance. Thus, experi- mental quantities that are related to the amount of quadrupole collectivity can be good probes of nucleon shell closures. One such measure of quadrupole collectivity in even-even nuclei is the energy of the first 2+ excited state, E(2f). The first 2+ excited state in an even-even nucleus is understood as arising from the breaking of a pair of nucleons and excitation of one nucleon of the pair into a higher-lying level. At closed 18 \l E01) (MeV) @ ”it. 0 ' 100120 140 Neutron Number (N) Figure 1.6: Behavior of E (21") across isotopic chains. Data were taken from Ref. [25]. Solid lines connect isotopic chains. The trend of increased values of E (2?) near neu- tron closed shells is very clear; a similar trend is seen for E (2?) along isotonic chains for closed proton shells. The same magic numbers seen in the variation of the neutron separation energies in Fig. 1.1 are apparent. shells there is an observed peak in E (2]) relative to neighboring nuclei. This reflects the fact that the energy required to break a pair of nucleons and promote one across a shell gap would be significant, while away from the shell closures, multi-state mix- ing lowers the collective quadrupole states [24]. Macroscopically the curvature of the potential energy surface is lessened, deformation requires less energy, and the energy of the first 2+ state is reduced. Thus, the value of E(2f) in even-even nuclei, relative to the neighboring even-even nuclei is a good indicator of nuclear shell structure and shell closures. Figure 1.6 illustrates the systematic variation of E(2f') for even-even nuclei across the nuclear chart, and the appearance of the characteristic peaks in this quantity at closed shells. Based on the observed trends in the energy of the first 2+ state in even-even nuclei, Grodzins [26,27] derived an empirical relation some time ago between E(2f) 19 and the quadrupole deformation parameter fig: 1225 f) — —— MeV. (1.13) E(2 _ 147/332 At closed shells, E (21*) values are high, and according to Eqn. 1.13, fig is small, indi- cating the near-spherical nature of magic nuclei. Moving away from closed shells, fig increases as E (21*) drops, representative of the increasing collectivity and developing deformation of mid—shell nuclei. A related indicator of quadrupole collectivity, and hence shell closures, in the nuclear chart is the reduced transition matrix element for the E2 transition between the first 2+ excited state and the 0+ ground state in even-even nuclei, B(E2 : 21+ —1 0;”). With an inverse relation to the energy of the first excited 2+ state according to the Grodzins rule [26], the B(E2 : 2f —» 0?) value peaks near midshell, and reaches minimum values near closed shells. These two quantities, E(2f) and B(E2 : 2f —» 01+) are among the most readily accessible in nuclei far from stability. Trends observed in these two independently measured values have served as indicators for shell closures at the far reaches of the nuclear chart (e.g. [13,14,28]). 1.3.2 Tracking single-particle energies and the extreme single- particle model Tracking the evolution of nucleon single-particle energies requires more information than simply the location of shell gaps in the nuclear chart. It is desirable to have a measure of the changing single-particle energies between closed shells. This involves not only identifying states within a nucleus, but also determining the nucleon configu- rations that contribute to these states. Elucidation of the contributing configurations permits a better appreciation of the shifting single-particle energies that lead to the observed nuclear structure. One method to obtain information on the wavefunctions contributing to nuclear 20 states is analysis of nucleon transfer reactions. Stripping reactions and the inverse, nucleon pickup reactions, are direct reactions which are constrained to the nuclear surface, and thus excite specific and limited nucleon degrees of freedom. The cross- sections for such reactions are dominated by the wavefunctions of the transferred nu- cleons [29]. This property makes single-particle transfer reactions uniquely powerful in characterizing the single-particle character of a nuclear state. Cross-section measure- ments can be compared to expectations assuming single-particle configurations, and experimental spectroscopic factors can be extracted [30]. These spectroscopic factors provide a measure of the single—particle character of nuclear states, as was mentioned above in the case of 209Pb (see Fig. 1.3). While other techniques, including fi decay, do not provide access directly to the wavefunctions of nuclear states, information regarding parentage of states can be ac— cessed indirectly by taking advantage of the predictions of the single-particle shell model. Within the framework of this model the spin and parity of the ground state should be representative of the single-particle orbital occupied by the last unpaired nucleon(s). The extreme single-particle model can be extended and predicts certain multiplet configurations corresponding to specific combinations of single-particle lev- els for nuclei with two unpaired nucleons. The spin, parity and relative energy of low-lying nuclear states can provide guidance for assignment of states as probable members of specific multiplets, and thus insight into the configuration leading to the states. While these inferences are not as robust as the spectroscopic factors from transfer reactions, such interpretations can serve as first steps in understanding the underlying nuclear structure, and relative single-particle energies. It is possible, in this way, to qualitatively track the relative ordering of single-particle orbitals. Lid- dick et al. [31] relied on such qualitative observations in the region of N 234, and gained insight into the 1r1f7/2-1/1 f5/2 monopole migration. 21 1.4 Re-ordering of single-particle states in neutron-rich f p shell nuclei The previous sections have discussed the general usefulness of the shell model in describing the structure of the atomic nucleus, as well as the nucleon-nucleon forces which are believed to drive the evolution of this structure across the nuclear chart. Specific regions of evolving structure within the f p shell are highlighted in this section. 1.4.1 N 232 and N 234 subshell closures As discussed in Section 1.1.1, the proton-neutron tensor monopole interaction has a strong influence on evolving shell structure across the nuclear landscape. One region mentioned previously is that near Z 220 and N 234. The monopole interaction in this region is that between the 1f5/2 (l— 1 / 2) neutron orbital and the 1f7/2 (l +1 / 2) proton orbital. The low-energy structures of neutron-rich nuclei with neutrons occupying the 1f7/2, 2113/2, 2111/2 and 1f5/2 orbitals are strongly influenced by the occupation of the proton 1f7/2 orbital, as illustrated in Figure 1.7. With increasing 7r1 f7 /2 occupancy in moving from 20Ca to 28N i, the energy of the 5 / 2" level, which has a large contribution from the u1f5/2 orbital, drops, moving much closer to the 3/2— (V2193 /2) ground state. Along with the sizeable spin-orbit splitting between the V2113 /2 and V2111 /2 states, the monopole migration of the u1f5/2 orbital suggests that, for nuclei with low occu- pancy of the 7rlf7/2 orbital, e.g., the 20Ca, ggTi and g4Cr isotopes, there is a sufficient gap in energy between the V2p3/2 and V2191 /2 states to produce an N 232 subshell closure. This expectation has been borne out experimentally in the 20C& [11, 12], ggTi [13] and g4Cr [14] isotopes in the systematic variation of the E(2[’) as a func- tion of neutron number, shown in Fig. 1.8(a). The rise in E(2f) at N 232 relative to the nearest even-even neighbors for the even-even Ca, Ti and Cr isotopes is clear. Further support for the N 232 subshell closure below Z 228 comes from the small B(E2 : 2]" 20;”) transition probabilities measured for 33Ti3g [28] and ggCI‘gg [32] [see 22 9 4.01 0.) a 3.5' _‘ g6 30 ‘12 2.5- ; 7/2‘ m ‘— I c, 2.0 —‘ 1, '. .2 ‘s “ ' B 1.5‘ \‘_ :“‘—- . —o - .8 1.0. _“ s -- ‘ "— 1/2 iii ‘~ ,7— 5/2' 0.51 ‘—-~ . _' OJ - _ - - o - - — 3/2' 49 51 - 53 55 57 - Figure 1.7: Evolution of the low-energy states in the N229 isotones from Z 220 to Z 228. The monopole shift of the u1f5/2 orbital with increasing occupation of the 7r1 f7 /2 orbital is suggested by the decreasing energy of the 5/2‘ state along the chain. Fig. 1.8(b)]. The high-spin structures of 50’52’54Ti [13] are also indicative of a sizeable N 232 gap. The separation between the 6? and the closely space 81*, 9? and 101’“ levels in 54Ti has been interpreted as an indication of the substantial cost to promote a V2113 /2 neutron across the N 232 gap to either of the V2191 /2 or V1 f5 /2 orbitals [13]. The experimentally-confirmed N 232 subshell closure has been well-reproduced by shell-model calculations in the f p shell using several effective interactions [33—35]. Be- yond N 232, the GXPFl interaction [33] further predicts a subshell closure at N 234 for the gOCa and ggTi isotopes, arising from the continued upward monopole shift of the V1 f5 /2 orbital [33]. Experimental evidence, however, is inconsistent with the prediction of N 234 as a subshell closure in the ggTi isotopes. Measurement of E(2‘1f) for 56Ti34 by Liddick et al. [36] placed the first 2‘L level at 1127 keV, nearly 400 keV drop lower than the E(2'1f)21495 keV observed for 54Ti3g [13] [see Fig. 1.8(b)]. This result, at odds with the expectation of a N 234 subshell closure for the ggTi isotopes, spurred the development of the new GXPF 1A effective shell model interaction [34]. To improve the description of the low-energy structure of 56Ti, five of the two—body 23 9 4000- - £4: (a) ’ " 20(23— ".Z - Ti — :N/ 3500 22 m 24Cr — 3000- 2500- g—Zflca 2000- 1500 —. _:'.j—‘_ .——. __ — ‘9: 3' ‘1— Ti 1000 4 ‘ .'—: 9—" .——~. 22 C ~ .._3 — 24 r 500 22 24 26 28 30 32 34 Neutron Number ’7 30 ' D. (b) zoCa° E 25 . I: \ 22T1 . f; " o' \ CI‘ 0 E 'r 24 T 20. 1| 3‘ i‘ +—-« , , . . N II ‘ ‘\ 6i 15‘ '0’ ‘1 .4, '. 24Cr % ’ ‘I ' ' f I “ " .1 "..--| i. ’ ‘ s“ "o . 10 i ‘\‘ “\ x, “f L a" 22T1 5- ’ 1' ' 0 ~ i . . T “O zoca 22 24 26 28 30 32 34 Neutron Number Figure 1.8: Behavior of (a) E(2if) and (b) B(E2 : 21* —> 01*) across the 20Ca, ggTi and g4Cr isotopic chains in the region of N 232,34. 24 matrix elements were altered to reduce the strength of the monopole interaction in nu- clei with N >32. These modifications produced an effective interaction that provides improved agreement with experimental data for isotopes in the region, and continues to predict a N 234 subshell closure for the goCa isotopes, which have no protons in the 7rlf7/2 orbital and thus a maximum separation between the 1/1f5/2 and 7rl f7 /2 [34]. The most direct method of addressing the possible N 234 subshell closure in the Ca isotopes would be to measure E (21*) for 33Ca34 directly. Low production yields have kept such a measurement elusively out of reach at the present time. 1.4.2 N 240 shell closure and intrusion of the neutron lgg/g orbital According to the shell model orbital ordering shown in Fig.1.2(c), the 199/2 single- particle orbital is well-separated from the 197/2, 2d5/2, and other higher-energy or— bitals, giving rise to a well-established N ,Z 250 shell closure. A less pronounced shell closure at N ,Z 240, arising from an energy separation between the 199/2 orbital and the lower-lying f p shell levels is also expected within the simple spherical shell-model framework. Near N ~Z however, there is little evidence of the suggested magicity at N or Z240. The self-conjugate nucleus 282nm, for example, shows no evidence of magicity, as it has a rotational-like yrast structure [37], and a large quadrupole deformation, fig20.39 [38], suggestive of a highly collective structure. This structure is similar to that of EggSr38, which again has a rotational-like yrast structure, and a large quadru- pole deformation, with fig >0.4 [39]. The dominance of a deformed structure rather than the expected spherical ‘doubly-magic’ structure of §8Zr40 is explained by the shape-polarizing effect of a well-stabilized deformed shell gap at N, Z238 [40]. At large prolate deformation (fig 2 0.4) a significant gap forms at N238 between Nils- son orbitals arising from the 199/2 orbital. Energetically, unless stabilized by a large 25 spherical shell gap (for instance at N, Z 2 28 or 50), it is preferential for a nucleus to assume a quadrupole deformation and occupy the local energy minimum at nu- cleon number 38. These deformed shell gaps strongly stabilize the highly deformed N 2Z 238 76$r core, and lead to a situation in which the 80Zr deformed configura- tion is more stable than the spherical configuration, even with a subshell closure at N , Z 240. The stability of the deformed 768r core is also believed to be responsible for the particle stability of the proton-rich nucleus §3Y38 [41]. Except for the collectivity at N N Z, the expected shell gap at nucleon number 40 is evident for both protons and neutrons in the neutron-rich nuclei. Figure 1.9(a) shows E(2'1”) for the neutron-rich 4OZr isotopes relative to their nearest neighbor even- even isotones over a range of neutron numbers from N 250 to N258. 90Zr50, 96Zr56 and 98Zr58 all show evidence of an increase in E(2f’) relative to the nearest isotonic even-even neighbours, supporting the presence of a Z240 subshell gap. However, this quasi-magic gap is eroded with the addition of neutrons beyond N258, due to the attractive monopole interaction between 111 97 /2 neutrons and 7r199/2 protons [15]. The bulk of evidence for the expected N 240 subshell gap comes from studies of ggNim and its nearest—neighboring isotopes. As is shown in Figure 1.9(b), the energy of the first 2+ state in 68Ni is 2.033 MeV [45], significantly higher than in the neighboring 66Ni and 70N i isotopes, providing an experimental signature indicative of a subshell closure at neutron number N 240. Further evidence for the N 240 subshell closure in the Ni isotopes comes from the E(4f)/E(2f) ratio, which reaches a local minimum at 68Ni, and the low transition probability [46,47] for excitation to the first excited 2+ state. While in the Ni isotopes there are numerous indicators for the goodness of the N 240 subshell closure, with the addition or removal of protons away from the magic number Z228, the N 240 closure is rapidly washed out. Figure 1.9(b) shows the systematics of E(2'1") for the even-Z g4Cr through 32Ge isotopic chains surrounding N 240. The E(2f) values for the g4Cr, ggFe, 30Zn and 32Ge isotopes all decrease with 26 E01) (keV) B 8 Z LIL A 1000 - 1'- 34 36 38 40 42 44 46 (Se) (Kr) (Sr) (Zr) (Mo) (Ru) (Pd) Proton Number 9 Ge (2:32) 5‘: (b) — Zn (2:30) +2 2000 - —" Ni (2:28) .—, 8 —— Fe (2:26) . m — Cr (2:24) . .' —-—Ni —’ 1000-—,==:——’—:—‘: — __ :3 _~. =, s‘ —‘ ‘:‘— Zn ‘ ’-——- Ge — Cr Fe 30 32 34 36 38 4O 42 Neutron Number Figure 1.9: Systematic variation of E(2i’") for the (a) isotonic chains in the region of 2:.40, and (b) neutron-rich g4Cr, ggFe, 28Ni, 302n and 32Ge isotopic chains in the regron around N 240. Data were taken from Refs. [14,25,42—44]. 27 increasing neutron number, even through N 240. Deformation of the ground state is well-known in the case of the higher-Z 3gGe isotopes. As discussed by Leenhardt et al. [48], systematics of B(E2: 2720?) support the trends seen in E(2'1"), and suggest that the gem isotopes are a transitional region between the spherical 28Nl isotopes and the well-deformed ggGe isotopes. Collectivity has also been observed to develop near N 240 for the isotopes below Z 228. Hannawald et al. [42] identified the first 2+ state in 66Fe40 at E(2f)2573 keV following fi decay from 66Mn. This low excitation energy was taken to be an indication of collectivity near the ground state, and a quadrupole deformation parameter of fig20.26 was obtained from the Grodzins relation [26]. A recent measurement has extended the data for the 26Fe isotopes past N 240, identifying the first excited state in 68Fe4g at 517 keV in a 2p knockout reaction [43]. This is the most neutron-rich Fe isotope studied, and shows that the decreasing trend of E(2f) for the 25Fe isotopes continues past N 240. The systematic decrease of E(2f) is even more pronounced in the g4Cr isotopes [see Fig. 1.9(b)]. The low E(21+) values in 60Cr36 and 62Cr3g are indicative of defor- mation, leading to deduced values of fig20.27 and 0.31 respectively [44]. Deformation is also suggested by the reported large quadrupole deformation lengths for 60’62Cr obtained in the analysis of proton inelastic scattering [49,50]. While data are not yet available for the low-energy structure of 64Cr40, the trend of lowering E(2f) values is predicted [51] to continue at and beyond N 240, as observed already in the 26Fe isotOpes. Understanding the origin of collectivity, which apparently develops in the region 0f N =40 in nuclei above and below Z 228, requires a knowledge of the interactions at Work. The strength of the N 240 subshell closure is dependent on the energy separation between the 1f5/2,2p1/2 neutron orbitals and the higher-lying 199/2 orbital. The strong repulsive monopole p-n interaction between protons in the 7rlf7/2 (l + 1 / 2) orbit and neutrons in the 1/1 99 /2 (l + 1 / 2) orbit decreases with the removal of 7rlf7/2 28 protons (Z <28), as does the attractive 7rlf7/2 (l + 1 / 2) - u1f5/2 (l — 1 / 2) interaction, which causes a narrowing of the V199/2-V1f5/2 energy separation. This effectively weakens the N 240 subshell closure, which may only manifest itself in 68Ni due to reinforcement by the spherical Z 228 closed shell [52]. In addition to the weakening of the spherical N 240 subshell closure by the monopole interaction, the first two Nilsson substates of the V1 99 /2 orbital are steeply downsloping in energy with increasing quadrupole deformation, as is shown in Fig. 1.10. The reduced energy gap at N 240 is insufficient to maintain the nucleus in a spherical configuration [48], as occupation of the deformed levels is energetically preferential, and collective deformed ground states develop, further eroding the N 240 subshell closure. Thus, the 199/2 neutron orbital plays a strong role in determining collectiv- ity near N 240, as its approach to the Fermi surface permits development of ground states with significant quadrupole deformation. The decrease in energy of 9/ 2+ states in odd-A Cr isotopes approaching N 240 [53—55] offers further evidence of the 199/2 orbital approaching the Fermi surface. Considering Figure 1.9(b) in more detail, the decrease in E(2if) suggests an onset of deformation in the g4Cr isotopes beginning at N 236, while in the 26Fe isotopes, such evidence for deformation first appears at N 238. The point at which collectivity sets in in this region will add to the understanding the effect of the neutron 199/2 orbital on the structure of nuclei at the extremes of the f p shell. The neutron-rich 24Cr and 26Fe isotopes have been investigated, but scant data are available for the neighbouring odd-Z g5Mn isotopes near to N 240. An understanding of the onset of deformation for the g5Mn isotopes would contribute significantly to the understanding of the evolving structure in the region near N 240. 29 7/2[404] 5/2[413] s " .64 ‘74 ../2 a 2 4 5 i391 1/2 440 1ng2 9’ 7/21413l 31- 93 ‘ 5/2 422 112199} 4 mm 5.0 413 - . Figure 1.10: Nilsson diagram for the neutron orbitals between 20 and 50. The Nilsson substates arising from the 199/2 orbital are highlighted in bold red, and with increasing deformation, the lowest two substates are observed to be steeply downsloping. Figure adapted from Ref. [2]. 30 1.5 Motivation for the measurement The motivation for the present study was to more fully understand the evolution of the single particle energies in the neutron—rich nuclei of the f p shell. Within the f p shell, and at its upper and lower borders, the evidence of evolving single-particle level ordering has been described. Further experimental data are required to constrain nuclear theory, and help in the development of cross-shell interactions, which are believed to be critical for the description of nuclei on the border of major shells. Near N 232,34, fi decay of the 53’33Ca isotopes was investigated to gain insight 53’3‘[Sc isotopes. Isomeric decay 0f into the low-energy structures of the daughter states in 541563c provided additional information for these isotopes. The odd-Z 218C nuclei are nuclei with a few nucleons more than a closed shell, and, as discussed previously, the low-energy level schemes can be described in terms of proton-neutron multiplets. Within this simple framework, insight into the neutron configurations in the Sc isotopes can be gained. An improved understanding of the situation in the 218C isotopes may shed light on the case of the 20C& isotopes, and the open question of a N 234 subshell closure at 54Ca. At higher Z, the onset of deformation for nuclei with Z <28 approaching N 240 is also interesting. As discussed, 26Fe and g4Cr isotopes have been well studied, but the location of the onset of deformation for the intermediate g5Mn isotopes is not clear. While extensive level schemes up to high—spin have been deduced for the lighter 57-ggMn isotopes [56], no level structures are available for Mn isotopes with N >38. In-beam 7 rays have been recently observed for neutron-rich 59“‘63Mn produced by multinucleon transfer between a 70Zn projectile and a 238U target [57]. For the most neutron-rich 63Mn only a single 7 ray was assigned. Five 7 rays were assigned to 62Mn, completing the low-energy level scheme proposed previously from fi decay [58], a Structure which is complicated by the presence of two fi-decaying states. 7 rays have been assigned from both fi decay [59] and in-beam work [57] in the case of 31 61Mn. Transitions identified in the former fi decay by Sorlin et al. were not placed in a level scheme, while only yrast states have been identified in the latter in-beam work. The present work studied the fi decay of 61Cr to states in 61Mn to produce a more complete level scheme for this nucleus. Comparison with shell model calculations in different valence spaces were found to provide insight into the onset of collectivity, and the monopole migration of the deformation inducing V199 /2 orbital with decreasing occupancy of the 7r1f7/2 orbital. The low-energy isomeric structure of §3K31 was also studied, and provided insight into the evolution of the proton single particle 1d3/2 and 131/2 states just below the f p shell. The insight gained in this region of the nuclear chart does not clarify current open questions, but rather adds to mounting evidence for structural evolution outside the predictions of current theoretical treatments. 1.6 Organization of Dissertation An introduction to the concepts of nuclear shell structure and the need for studying neutron-rich nuclei was presented in this chapter, including the motivation for the present measurement. The principles of the measurement techniques of fi and 7 decay are presented in Chapter 2, as well as an explanation of how they pertain to nuclear structure. Details of the experimental setup are presented in Chapter 3, including a description of the detectors. The results for the low-energy level structures of 50K, 53’54’5630 and 61Mn are presented in Chapter 4, and the implications of these results discussed in Chapter 5. Chapter 6 includes a summary of this work, and a discussion 0f the possibilities for future studies using both fi decay and other complementary techniques. 32 Chapter 2 Experimental Techniques fi-delayed 7—ray spectroscopy is a powerful method to probe the structure of the most exotic nuclei, including short-lived nuclei near the driplines. A highly sensitive technique, fi decay itself can provide valuable information on half-lives, decay Q values, and branching information, even in cases with very few events. The branching of the fi-decay process populates states in a daughter nucleus selectively, based on the relative angular momenta and parities of the parent and daughter states. The excited states populated in a daughter nucleus most often de—excite by the emission of 7 rays. The-energy of the emitted 7 rays provides information on the excitation energy of the states populated in the fi decay, another key observable in nuclear structure studies. This chapter details the relevant processes of fi and 7 decay, and outlines how their combined use permits access to details of the structure of exotic nuclei. 2. 1 fi decay fl decay, in the most qualitative sense, describes the process of the conversion of a proton in the nucleus into a neutron, or vice versa. During this process, the atomic number (Z) and neutron number (N) of the nucleus is changed, but the mass number (A) remains the same. Thus, fi decay moves nuclei along isobaric chains towards 33 stability. One of the earliest observed radioactive decay modes, fi decay is governed by the weak force, and encompasses three distinct decay processes: fi‘ : g‘XN 22313;“ 0‘ +v+ Q}, (21) 6+ : g‘XN 224’ 1,7,, + 8+ + z/ + and (2.2) Electron Capture : Z AXN + e‘ 494 Y1?“ + V + Qfl, (2.3) where e’ is an orbital electron, fii is a beta particle, l/ is a neutrino, '17 is an anti- neutrino, and Q [3 represents the characteristic amount of energy released for a given fi decay process [3]. The energy released, Q73, must be greater than zero for fi decay to occur. On the neutron—rich side of stability fi" decays occur, driving nuclei towards stability along an isobaric chain by converting neutrons to protons within the nucleus. On the neutron-deficient side of the line of stability, both fi+ and electron capture processes can occur, though fi+ is only observed in nuclei with a sufficiently large Q ,3, due to a 1.022 MeV threshold for the creation of an e—-fi+ pair. Electron capture is the process in which an orbital electron is captured by the nu- cleus, and combines with a bound proton to achieve the transformation of a nucleus éXN into Z—f’YNH' The Q3 for electron capture is given by QEC 2 [m(/Z’XN) — m(Z_’14YN+1)]c2 — Bn, where Bn is the binding energy of the captured n—shell elec- tron [3]. Only one particle, a neutrino, is emitted in the electron captures process. Due to the extremely small neutrino mass, and the fact that neutrinos have no charge, it is nearly impossible to observe this particle directly. Thus, electron capture processes are typically observed by monitoring secondary radiation. The captured electron is 111081? often an inner shell (i.e. K or L shell) electron, and its capture during the de- cay process creates a vacancy, which is filled by electrons from higher shells. These downward transitions result in emission of characteristic X rays or Auger electrons, Which can provide a signature for the electron capture process. The other possibility for neutron-deficient nuclei to move towards stability along 34 an isobaric chain is fi+ decay. The Q 5 value for fi+ decay is given by Q W” 2 [m(’§XN) — m(z_14YN+1) — 2me]c2, and thus there is an energy threshold for fi+ de- cay, requiring an atomic mass difference of at least two electron rest masses, 1.022 MeV, between the parent and daughter species. As was the case in electron capture, a neu- trino is emitted during the decay, but a second particle, a fi+, is also emitted. It is very diflicult to detect the neutrino, but it is possible to detect the fi+, either directly, through its interaction with atomic electrons in a detection medium, or by detection of the two 511-keV 7 rays which result when the positron thermalizes and annihilates with an electron. On the neutron-rich side of stability, the relevant fi—decay process is fi‘ decay. The Q [3 value for this decay is the simplest case, given by the difference in the atomic masses of the parent and daughter, Q ,6— 2 [m(’§X N) — m( Z +’1’YN_1)]C2. As was the case in fi+ decay, this process involves emission of two particles, an anti-neutrino and a high-energy electron or fi‘ particle. The fi— can be detected directly through its interaction with atomic electrons in the detection medium. It is this decay mode that is of interest in the present study of neutron-rich nuclei. 2.1.1 fi-decay half-lives One of the most fundamental quantities describing a fi-decay process is the rate at which the decay occurs. fi decay is governed by first-order rate law, and the number of nuclei, N, in a sample at a time t is given by: N(t) = Noe-M, (2.4) Where No is the number of nuclei originally in the sample, and A is the decay constant Characteristic of the decay. Physically, A represents the probability per unit time for the decay of a given nucleus [3]. The decay constant, A, is closely related to the more 35 often quoted half—life, T1/2 as: T,,=_(_> (2.5) The half—life corresponds to the average time required for half of the nuclei in a sample to decay. Governed by the weak interaction, the half-life of fi-decay processes ranges from a few milliseconds to seconds or even longer. The longest-lived fi—decaying nuclei have half-lives in excess of 1018 years. The half-life of a fi-decay process depends strongly on the energy of the decay processes, and differences (i.e. spin and parity) between the parent initial state and the final state in the daughter. The dependence on the initial and final state structures can be qualitatively understood, and relates to the selection rules for fi decay, which are discussed in Section 2.1.3. The next section addresses the process of fi-delayed neutron emission, a decay mode which can be a significant part of the decay of neutron-rich nuclei. 2.1.2 fi-delayed neutron emission Neutron emission is a decay process involving direct emission of a neutron from a nucleus according to the following equation: ’2sz —>”‘IZ XN—l + n + Q... (2.6) where 62,, represents the mass-energy difference between the initial and final states. When the daughter nucleus, A-IZXN_1, is left in its ground state, Qn 2 -Sn, where Sn is the neutron separation energy, a measure of how much energy is required to remove a neutron from a given nucleus, as discussed in Chapter 1 (see Eqn. 1.1). In general the neutron separation energy becomes lower with the addition of neutrons along an .iSOtOpic chain. Thus, for the most neutron-rich nuclei, neutron separation energies are (mite low, and in fact when the separation energy reaches zero, the nucleus is said to be at the neutron dripline. Not being subject to the Coulomb force, and thus without a C'Oulomb barrier, the last neutron is simply not confined within a nucleus at the 36 neutron dripline. While the ground state of a fi-decay daughter nucleus may be bound with respect to neutron emission (i.e. Sn > 0), there are many cases in fi‘ decay in which the decay populates states in the daughter which are above the neutron separation energy, or neutron threshold. When this occurs, the emission of a neutron from the excited state of the daughter can proceed rapidly, as neutron emission is mediated by the strong force and competes favorably with 7-ray emission mediated by the electromagnetic interaction. Following neutron emission, excited states in the fin daughter may be populated, and further de—excite through emission of 7 radiation characteristic of that nucleus. It is possible to identify fi-delayed neutron emission in the decay of neutron-rich nuclei through the detection of these characteristic 7 rays. 2.1.3 Selection rules and log ft fi“ decay involves the emission of two particles, a fi‘, or high energy electron, and an anti-neutrino. These emitted particles are leptons with intrinsic spin S21 / 2, and can carry additional orbital angular momentum. If the spins of the two emitted particles are antiparallel, they are coupled to S g20, and the system undergoes what is called Fermi decay [3]. The other possibility, that the spins of the electron and neutrino are aligned parallel, coupling to S [321, is called Gamow-Teller decay. The orientation of the electron and anti-neutrino spins provides the first classification of fi decay — whether the electron and/or neutrino carry additional angular momentum further divides the fi-decay process. The so—called allowed fi decay describes the situation in which the electron and neutrino carry no orbital angular momentum (120). With Sfi20, an allowed Fermi decay has no change in the nuclear spin, AJ 2 [Ji - J f] 2 0. Additionally, with l20, there is no change in parity between the initial and final states, 7r 2 (-l)’ in this Case. $321 for allowed Gamow-Teller decays, and one unit of angular momentum is Ca«I‘ried by the electron and neutrino, even with zero additional units of orbital 37 Table 2.1: fi decay selection rules for allowed and forbidden transitions. Table adapted from Ref. [4]. Transition Type Arr Al Al logft Superallowed No 0 0 2.9 — 3.7 Allowed No 0 0, 1 4.4-6.0 First forbidden Yes 1 0, 1, 2 6 - 10 Second forbidden No 2 1, 2, 3 10 - 13 Third forbidden Yes 3 2 3 4 2 15 ’ 3 angular momentum (I20). Allowed Gamow-Teller decays thus have J; 2 if + 1, and so [Jf — 1] 2 J,- 2 [Jf + 1], or AJ20,1. Again, with l 2 0, there is no change in parity between final and initial states in allowed Gamow—Teller decay. Since AJ20 is allowed in both cases, most transitions are mixed Fermi and Gamow-Teller decays, except when Ji 2 J f 2 0, in which case only Fermi decay is possible [3]. This special case is called a “super-allowed” decay. Allowed fi-decay transitions are among the fastest, although other transitions in which the electron/neutrino carry non-zero orbital angular momentum (l > 0) are possible. These so—called ‘forbidden’ transitions are not completely forbidden as the name suggests, but are slower and thus much less likely to occur than allowed transitions. The degree of forbiddenness increases with l - first forbidden decays have the electron and neutrino emitted with 121, while l22 corresponds to second forbidden decays, etc. Table 2.1 summarizes the selection rules for fi decay up to third forbidden decays. The comparative half-life, or ft value provides a method for comparing the fi- decay probabilities in different nuclei. The dependence of the fi-decay rate on specific nuclear properties such as the Coulomb interaction based on the atomic number 0f the daughter, Z, and the maximum energy of the fi transition are incorporated into the value of f, the Fermi integral [3]. With these well-understood dependencies r emOved, differences in ft, or the more commonly reported log f t, represent differences betvWeen the initial and final nuclear states. Values for the Fermi integral are tabulated 38 for ranges of Z and the maximum fi energy, Emu, but approximate values can be obtained using the empirical expression for log f [3‘ [60]: log f5_ 2 4.010g Emaa: + 0.78 + 0.02Z — 0.005(Z — 1) log Emax. (2.7) The value of t in log f t is the partial half-life for the decay to a specific state in the daughter nucleus. If the parent nucleus has a total halflife T3301, then the partial half-life for decay populating a specific state i in the daughter is given by [3]: total partial,i _ 1/2 T1/2 _ BR,’ (2.8) where BB,- is the branching ratio to the state i in the daughter. 2.2 7-ray decay Following fi decay, the daughter nucleus may be left in an excited state. Internal transition or colloquially, 7—ray decay, is the process by which an excited nucleus can release excess energy in the form of electromagnetic radiation, or a photon. From a schematic perspective, the emitted electromagnetic radiation connects two states within the nucleus, each of which have definite spin, J, and parity, 7r. The emitted photon must connect the states in energy, and conserve both spin and parity. This results in strict selection rules for 7-ray transitions. 2.2.1 Selection rules and lifetimes Photons have no intrinsic spin, and thus the orbital angular momentum, A, carried by a photon is constrained by the spins of the initial (J,) and final (J f) nuclear states as: [Ji—Jfl SA S (Ji+Jf)h. (2.9) 39 Table 2.2: 7-ray decay selection rules and multipolarities up to A23. Table adapted from Ref. [4]. Radiation Type Name A2AJ Arr E1 Electric dipole 1 Yes Ml Magnetic dipole 1 No E2 Electric quadrupole 2 N 0 M2 Magnetic quadrupole 2 Yes E3 Electric octupole 3 Yes M3 Magnetic octupole 3 No While there may be a number of possible A values, or multipolarities for 7—rays con- necting two nuclear states, the transition rate is strongly dependent on A, and the lowest possible 7-ray multipolarity will be favored. However, the required conserva- tion of parity plays a significant role in determining the type of 7-ray transition which occurs. Each nuclear state corresponds to a distinct distribution of matter and charge, and thus an electromagnetic transition connecting two nuclear states necessarily cor- responds to a change in the distribution of protons and neutrons within that nu- cleus [4]. A shift in the charge distribution produces an electric field, while a shift in the distribution of current gives rise to a magnetic field. Thus, de—exciting 7-ray transitions can be classified as either electric or magnetic in nature. Electric transi- tions are characterized by a parity dependence on A of A7r2(-1)A, while for magnetic transitions, the parity change is A7r2(-1))‘+1. Combined with the angular momentum requirements, the resulting 7-ray selection rules are summarized in Table 2.2. Based upon the type and multipolarity of a de—exciting 7-ray transitions, estimates can be made for the expected lifetime of the parent state. The decay constant for the emission of a photon from a well—defined single state with excess energy is given by what is known as Fermi’s golden rule [4]. The transition rate, T(EA) or T(M A), for 40 a 7-ray transition with energy E, can be written as follows [4]: _ 87r(A+1) E 2"“1 T(EA) _ amA[(2/\+1)H]2(fi) EB(EA), (2.10) _ h 2 87r(A+1) E 2"“1 TM“) — Ghee—777,5) A[(2A+1)11]2('EZ') 718(EA)’ (2'11) where a is the fine structure constant, 5 is Plank’s constant divided by 27r, c is the speed of light, mp is the mass of the proton and B(EA) and B(MA) are the reduced electric and magnetic transition probabilities respectively, which contain all of the nuclear wavefunction information. To obtain useful estimates of the transition rates for 7-ray transitions, estimates of the reduced transition probabilities are required. However, these values depend strongly on nuclear models. A commonly-used value was derived for the extreme single-particle model, discussed earlier in Section 1.2.1. Assuming that the transition involves the change of a single particle in a uniform density nucleus, then the so-called Weisskopf single-particle estimates for the reduced transition probabilities are given by the following [4]: 2 BW(EA) = (1:72” (£3) A2V3e2(fm)2*, (212) 10 _ 3 2 _ h 2 _ BW(MA) = #12)” 2(fi‘3’) A<2A W3 (51;) (rm)2A 2. (2.13) Combination of the Weisskopf single-particle transition probability estimates with the expressions for the 7-ray transition rates yields the so-called Weisskopf estimates for 7-ray decay transition rates. The Weisskopf estimates for the lifetimes of certain types of 7-ray transitions are given in Table 2.3. 41 Table 2.3: Weisskopf single-particle lifetimes estimates for electric and magnetic 7 transitions up to multipolarity 3 [4]. A represents the mass number of the nucleus, while E7 is the 7 energy in MeV. Radiation Type Weisskopf Estimate for Transition Lifetime (3‘1) E1 1.0 x 1014/12/3Ei,3 M1 5.6 x 1013153 E2 7.3 x 107A4/3E2 M2 3.5 x 107A2/3E§ E3 3421215; M3 16/14/313; 2.2.2 Isomeric 7-ray transitions The possibility exists in nuclei for excited states to be relatively long-lived, or isomeric. Isomeric states are those which have unexpectedly long lifetimes compared to other nearby states in the nucleus. While there is no clear lifetime cut-off for classifying a nuclear state as isomeric, for example states with lifetimes > 10‘9 s can be considered isomeric, particularly in comparison to the typical lifetime for nuclear states of order 10‘15 3. When low-lying states in a nucleus have very different angular momenta and there are no intermediate states, the 7-ray decay from the higher-lying state will be hindered due to the large change in angular momentum, a result predicted even by the Weisskopf single-particle estimates. This leads to long-lived nuclear excited states, or isomeric states. When this long lifetime arises due to a large change in angular momentum between the excited state and lower-lying states, this is known as spin isomerism. Also evident from the expressions for the Weisskopf single-particle lifetime estimates is a strong 7—ray energy dependence. Low-energy 7-ray transitions may be long-lived, particularly in heavy nuclei with large mass number. There are other sources of hindered 7-ray transition probabilities which can lead to nuclear isomerism. As an example, in highly-deformed nuclei, the projection of the total angular momentum onto the nuclear symmetry axis, K, is an approximately 42 conserved quantum number. As such, selection rules emerge for transitions allowed based on AK, and K isomerism is possible when transitions are inhibited by a large AK [19]. Aside from large changes in quantum numbers hindering 7-ray transitions, isomers can arise when transitions to lower states involve a significant change in nucleon con- figuration, even without a large change in angular momentum. The overlap between initial and final states can be poor, and thus hinder the transition. These are called structural isomers and are analogous to the molecular isomers in chemical systems. As such, the presence (or absence) of nuclear isomers can provide critical insight into nuclear structure. 2.2.3 Internal conversion A competing process to 7-ray decay, internal conversion, occurs when an excited nucleus de-excites by interacting electromagnetically with an orbital electron, and ejecting it [3]. This process involves no direct electromagnetic radiation (no photons are involved), and the ejected electrons will have discrete energies, related to the transition energy according to: EIC = Etransition _ electron binding energy- (2-14) Since orbital electrons occupy discrete shells, within a conversion electron spectrum, separate lines will appear corresponding to ejection of K, L, M, etc. shell electrons. To quantify the process of internal conversion, and its competition with 7-ray decay, an internal conversion coefficient (a) is defined as follows: a _ number of internal conversion decays _ A12 (2 15) _ number of 7-ray decays — A7 ’ ' As there are contributions to the internal conversion process from electrons in a 43 number of electron shells, the total probability of decay, atotal is given by summing a for all possible atomic electrons. However, atom) is usually dominated by contributions from ejection of inner (i.e. K and L) shell electrons. Compilations for internal conversion coeflicients are available [61], which provide the internal conversion coefficient for each atomic electron shell, as well as the total value. Internal conversion is expected to compete most strongly with 7-ray decay for heavy elements (high Z), and low transition energies (low E) [3]. 2.3 Summary of the decay of exotic nuclei Governed by the selection rules presented in Table 2.1, the allowed fi decay of a state in the parent nucleus selectively populates states in the daughter according to AJ 2 0,1. Usually the fi-decaying state is the ground state, though isomeric states may also undergo fi decay. If the spin and parity of the decaying state is known, the selectivity of the fi-decay process immediately provides limitations for the possible J7r values of the states populated in the decay. Following the fi-decay process, excited states populated in the daughter will de- excite by 7-ray emission to the ground state, or if the populated state is above the neutron separation threshold for the nucleus, by neutron emission to states in the fin daughter. fi-delayed neutron emission can be identified and quantified by moni- toring the 7—rays emitted when states in the fin daughter de—excite, as discussed in Section 2.1.3. Emitted 7 rays provide information not only on the energy of states in the daughter, but can also limit spin and parity values, when combined with other information. The most intense 7-ray transitions will, in general, correspond to low multipolarity transitions, usually limiting non-isomeric transitions to AJ 2 1, 2. This requirement can further constrain possible J ’r values for states populated in the de- cay based on the observed 7-ray transitions connecting to other states. Considering energetics, 7 rays may de—excite a state directly to the nuclear ground state, in which 44 case the 7-ray energy is equal to the excitation energy of the state. Alternatively, 7 rays may occur in cascade, through an intermediate state. Level energies can be determined by looking for 7 energy sums in the case of 7-cascades. Often, there is more than one possible cascade, and identifying 7-ray transition sums can help in building up nuclear level schemes. 7-7 coincidence measurements also provide invalu- able information in determining 7-ray cascades, while 7-ray transition intensities can additionally constrain the possibilities. Prompt, or isomeric 7-rays can be similarly used to determine low-energy level schemes of parent nuclides. Summing relationships and 7-ray intensities provide valu- able input for assembling level structures from prompt 7-ray data. The very powerful techniques of fi-delayed 7-ray spectroscopy and prompt 7—ray spectroscopy were used to study the low-energy level structures of neutron-rich nuclei in the fp shell in the present work. The next chapter will detail the experimental set-up used for the present study. 45 Chapter 3 Experimental Setup The motivation for nuclear structure studies of exotic neutron-rich radioactive iso- topes was presented in Chapter 1, focusing on important cases in the f p shell. The second chapter presented the fundamentals of fi and 7 decay, discussing the selec- tivity of these processes and the utility of fi-delayed and prompt 7-ray spectroscopy. This chapter focuses on the more practical aspects of the experimental set-up used in two fi-decay experiments that studied neutron-rich f p shell nuclei at the National Superconducting Cyclotron Laboratory (NSCL) at Michigan State University. Pro- duction and separation of the short-lived isotopes of interest using fragmentation and in-flight separation is discussed. The details of the experimental end-station detectors, including the fi counting system (BCS) and Segmented Germanium Array (SeGA) are presented, including a description of the readout electronics and calibrations per- formed for each detector system. Details of the procedure for implantation-decay correlations and elements of data analysis are also included. 3.1 Isotope production and identification Only a handful of very long-lived radioisotopes are naturally abundant on Earth. Thus, to study unstable nuclei, it is necessary to produce these species in an artificial 46 RT-ECR final focus A1900 \ stripping _ foil . " , image 2 . _ , (dispersive 7"}?49'5-‘83/71 " “ = " ' lane .7 1%,, ‘3 -- p ) 41' fi'KIZOO target Figure 3.1: Schematic of the NSCL Coupled Cyclotron Facility, showing the K500 and K1200 cyclotrons and the A1900 fragment separator. way. At NSCL, secondary beams of exotic rare isotopes are produced using the pro— jectile fragmentation technique. The fragmentation process involves the interaction of a medium-energy primary beam (100—160 MeV / nucleon) on a stationary target. The primary beam is a stable isotope that is first ionized in one of the NSCL’s electron cyclotron resonance (ECR) ion sources. The primary beam of ions is then accelerated by the coupled K500 and K1200 cyclotrons, shown schematically in Fig. 3.1. After extraction out of the K1200 cyclotron, the primary beam has an energy in the range 100 to 160 MeV/ nucleon. The beam is then impinged on a production target. Typi- cally the production targets are thin 9Be foils with thicknesses of order hundreds of mg/cm2. Beryllium is used because of its high number density and robust physical properties. Fragmentation of the primary beam in the production target, in which a range of nucleons is abraded from the heavy beam, results in the production of nuclei between the A and Z of the primary beam and hydrogen. As fragmentation reactions are mostly peripheral reactions, there is a relatively small momentum transfer [4], and the surviving larger-mass fragments emerge from the thin production target with energies close to that of the primary beam. These secondary beam components are forward focused and continue into the A1900 fragment separator [62]. 47 Given the large variety of ions produced in the fragmentation process, and the low production rates for the most exotic nuclei, it is critical to efliciently separate the ions of interest from all others produced. This is accomplished at NSCL using the A1900 fragment separator [62]. The A1900 separator is a large acceptance, achromatic, high resolution separator. Fragments entering the separator after the production target are separated by a combination of magnetic selection and energy loss (Bp—AE—Bp). Ions of a single magnetic rigidity, Bp 2 mv/ q, are selected in the first stage of the separator. Since ions emerge from the production target with approximately the same velocity (7)), this first separation selects ions with nearyly the same mass / charge (m/ q) ratio. At the second intermediate image of the separator, the dispersive image (image 2, or I2), maximum horizontal dispersion of the beam is present and the ions pass through a wedgeshaped energy degrader. The ions entering the wedge have the same Bp, but experience an energy loss proportional to Z2 in the wedge material, and thus exit the wedge degrader with different energies, and different momenta. A second B p separation stage in the second half of the separator then provides isotopic resolution [63]. The rate and purity of the secondary beam are determined by the composition and thickness of the production target and wedge degrader, as well as the momen- tum acceptance settings. Variably-adjustable slits are located at each of the three intermediate images of the separator, and can be narrowed to limit the momentum acceptance of the separator below the maximum Ap/p ~5%. The A1900 is a versatile device, and can be tuned to accommodate the needs of a wide variety of experiments. Many experiments require a high level of purity in the secondary beam, while others have more stringent requirements in terms of the overall fragment rate. The two experiments discussed in this dissertation made use of so-called “cocktail beams”, with multiple component nuclei, as the overall beam rate, and not beam purity, was most important. Regardless of the specific requirements of a given experiment, a critical experi- 48 mental necessity is particle identification. Nearly all experiments at NSCL use the combination of energy loss in a thin detector and time—of—flight (TOF) to identify secondary beam particles. The general requirement for particle identification is thus two detectors separated by some distance over which to measure a time of flight. The BCS set-up measures energy loss in a silicon PIN detector, which is a part of the BCS detector telescope, described in detail in the following section. The time-of-flight mea- surement with the BCS uses the signal from the energy-loss PIN as a start, and either the cyclotron RF signal, or a signal from a plastic scintillator at image 2 of the A1900, as a stop signal. The plastic scintillator can also be used to provide an event-by-event momentum correction to the time-of-flight for those experiments in which the A1900 is operated with a large momentum acceptance. The 12 scintillator is read out by two photomultiplier tubes, one on each end, permitting determination of the approximate position of fragments in the scintillator. The position information is used to correct time-of-flight measurements for the differing path lengths and velocities of ions when a large momentum acceptance is used. Such time-of-flight corrections were necessary for both experiments discussed in this dissertation. This section has described in general terms the production, separation and iden- tification of exotic nuclei at NSCL. The following subsections present the specific details for production of the cocktail beams used in NSCL experiments 05101 and 07509, the experiments discussed in this dissertation. 3.1.1 NSCL experiment 05101 The main focus of NSCL experiment 05101 was the study of neutron-rich gOCa iso— topes, and settings of the A1900 magnetic rigidities were optimized for 54Ca and 54K production. Some data, however, were taken with an incorrect target setting, which resulted in a secondary beam of isotopes in the region of 60V. The fragmentation pro- cess used to produce these neutron-rich nuclei was the interaction of a 76Ge beam with a Be target. The primary beam of 76Ge was ionized to 12+ in the room—temperature 49 ECR ion source and accelerated to 11.6 MeV/ nucleon in the K500 cyclotron. Follow- ing foil stripping to a charge state of 30+, the 7’6Ge primary beam was accelerated to a full energy of 130 MeV/ nucleon by the K1200 and impinged on a 47-mg/cm2 9Be target, located at the object position of the A1900 separator. Secondary fragments were separated through the A1900 with magnetic rigidity settings of Bp24.403 Tm before the wedge, and Bp24.134 Tm following a 300 mg/cm2 Al wedge located at the intermediate image of the separator. The full momentum acceptance of the A1900 (Ap/p~5%) was used for fragment collection. Given the large momentum acceptance used in the experiment, the plastic scintillator at the dispersive image of the A1900 (12) was used for event-by-event momentum correction, and as the stop for the time— of—flight measurement. The start of the TOF measurement was provided by the signal of the most upstream PIN detector in the BCS silicon telesc0pe, PIN01, described in the next section. The secondary cocktail beam consisted of ~10 isotopes, including 60V, delivered to the experimental end-station of the BCS + SeGA located in the N3 vault. The particle identification plot for this setting is shown in Fig. 3.2(a). 3.1.2 NSCL experiment 07509 The isotopes of interest in NSCL experiment 07509 were those centered only around 54Ca. The primary beam of 7"’Ge was accelerated to 130 MeV/ A in the same man- ner as in NSCL experiment 05101, which was detailed in section 3.1.1. The beam was impinged on a 352-mg/cm2 thick Be production target, and the projectile-like fragments entered into the A1900 fragment separator. The first half of the separator was set to a magnetic rigidity of 4.103 Tm. A thin 45 mg/cm2 Al wedge was placed at the A1900 intermediate image to provide differential energy loss. The second half of the separator was set to a magnetic rigidity of 4.030 Tm to optimize production and separation of 54Ca. The A1900 was operated with a maximum momentum accep- tance of ~5%. In total 24 isotopes, including 54Ca, were directed to the experimental end station set up in the S2 vault. As was the case in experiment 05101, the TOF DJLHUJUJ NAOOO COCO (21) Mb) 000 CO Energy loss (arb. units) N O\ O 850 .~ 800 : ; _ 750 \‘I O O 650 ‘:-- .‘ :';‘ (b) 550/ , 50095:: " ,, -' 33:;- 4502 ‘35 660 700 740 780 820 Time of flight (arb. units) Energy loss (arb. units) 8 O Figure 3.2: Particle identification plots for the relevant experimental settings in NSCL experiment (a) 05101, and (b) 07509. In both cases, time of flight measurements were started by the signal from the most upstream PIN detector of the BCS, and stopped by a signal from the 12 scintillator. Energy loss was measured in the upstream PIN detector of the BCS. The particle groups were identified by their subsequent decay. 51 measurement was started by the signal from the upstream PIN detector of the BCS, PIN01, and stopped by the delayed signal from the 12 scintillator. The 12 scintilla- tor was also used to provide event-by-event momentum correction for the TOF. The particle identification plot for experiment 07509 is shown in Fig. 3.2(b). 3.2 Detector apparatus for fi-decay experiments The experimental apparatus for both experiments discussed here consisted of the NSCL fi counting system (BCS), and detectors from the SeGA. Secondary beam fragments were implanted into the detectors of the BCS, which provided triggers for both implantation and fi-decay events. 7 rays were detected in the SeGA detectors, and 7-ray events were recorded for both implantation and decay processes. The fol- lowing sections describe the end-station systems in detail for NSCL experiments 07509 and 05101. The end-station set-ups for these two experiments were nearly identical. However, the minor differences between the set-ups will be noted when present. 3.3 fi Counting System The NSCL fi counting system (BCS) [64] consists of a stack ’of Si detectors. A schematic of the BCS detectors as set-up in NSCL experiment 07509 is shown in Fig. 3.3. The centerpiece of the system is a highly-segmented double-sided Si mi- crostrip detector (DSSD), into which the secondary beam is implanted. The high segmentation of the implantation detector permits event-by-event correlation of frag- ment implantations with their subsequent fi decays. The detectors of this system as set-up for NSCL experiment 07509 are described in this section. The readout elec- tronics, calibration procedures and calibration results are also described. The central detector of the BCS was a 4 cm by 4 cm, 40x40 DSSD, manufactured by Micron Semiconductors. The segmentation of the detector with 40x1 mm strips 52 PINOI PIN03 SSSDl SSSD3 SSSDS [ PIIIIOZ SSSD2 SSSD4 SSSD6 l w ‘1 y l [ [l I z l [ l 20 mm 1 7mm ‘ ) . - \\\ 12 mm 1 mm - \‘\ . 7 Beta Counting System / ’ \\ ,,,,, ,.,.. ,»/.I - 1.l..L.l33_ Figure 3.3: Schematic diagram of the silicon detectors which constitute the NSCL fi Counting System. The arrangement of detectors for experiments 07509 and 05101 were as pictured; for details regarding detector thicknesses, see the text. Figure adapted from Ref. [65]. 53 on the front and 40x1 mm perpendicular strips on the back resulted in electrical segmentation into 1600 individual pixels, each of which act as an individual detec- tor. Fragments were continuously implanted in this detector, and implantations were correlated with decays on an event-by—event basis. The correlation requirements are described in detail in Section 3.5 at the end of this chapter. A DSSD detector thick- ness of 995 mm was used in NSCL experiment 07509, while the implantation detector used in experiment 05101 had a nominal thickness of 979 ,um. Downstream of the implantation detector were six 5 cm by 5 cm, 16-strip single- sided silicon strip detectors (SSSD3), which constituted the fi calorimeter. Segmented in only one direction, the orientation of strips in the SSSDs was alternated between the horizontal (51:) and vertical (y) directions for adjacent detectors, as is shown in Fig. 3.3. This arrangement of detectors allowed for two-dimensional position information for any particles moving downstream of the DSSD. The thicknesses of SSSDl through SSSD6 for NSCL experiment 07509 were 975 pm, 981 pm, 977 pm, 989 pm, 988 ,um, and 985 pm respectively. For experiment 05101, the thicknesses were 981 pm, 977 pm, 975 pm, 988 pm, 989 pm, and 985 pm for SSSDl-SSSD6 respectively. A stack of three silicon PIN detectors were placed upstream of the DSSD, which acted as active energy degraders and provided the energy loss and timing information required for particle identification. The thicknesses of the three PIN detectors were 297 pm for PIN01, 297 pm for PIN02, and 488 pm for PIN 03, during NSCL experiment 07509. Experiment 05101 had thicknesses of 991, 997 and 309 pm for PIN01, PIN02, and PIN03, respectively. A passive aluminum energy degrader was placed upstream of the PIN detectors to ensure fragments were stopped near the center of the DSSD implantation detector. A 7.5 cmx7.5 cm Al degrader foil was held on a rotating mount, that was used to adjust the angle, and thus the effective thickness, of the foil with respect to the beam. The Al degrader thickness for experiment 07509 was 4 mm, while experiment 05101 only required a degrader with a thickness of 2.3 mm. 54 MCS CPA 16 Preamp Grounding” Non-inverting Lgvv Impedance+CAEN V785 Board (Inverting) Gam Matching ADC for FRONT (BACK) .Hish Gain P' s t 5153820 Fast 1C0 ys ems 812w CAEN V785 3 _ Sh./ Disc. .- D . . . chler (ECL t)Isc (Nmmveflmg) (1:3? ADC Fast disc. '16—channel OR CAEN V977 F I /O t Dela 9»- ’ an n u ’+ Y Coinc. Reg. ‘ OR (of all FRONT, all BACK strips) Figure 3.4: Schematic diagram of the signal processing electronics for the DSSD of the BCS. 3.3. 1 Electronics A schematic diagram of the signal processing electronics for the DSSD is shown in Figure 3.4. Signals for the front and back segments were carried from the detector within the vacuum chamber to feedthroughs on a backplate by separate ungrounded 50-pin ribbon cables. Outside the vacuum, 50-pin ribbon cables carried signals for the front and back of the DSSD to a grounding board. On the ground board, signals from each strip were paired with ground signals, and were grouped in units of 16 to match the input channel density of the conventional analog electronics. The grouping performed on the ground board created blocks with F1-16, F17-32, F33-40, B1-16, Bl7—32 and B33-40, where F and B represent front and back strips respectively, and numbers refer to the strip number of the detector. One of the challenges of using a single detector to detect both implantation events and fi decays is the dynamic range required to measure both types of events. Ion implantation deposited energies of order several GeVs, while fi decays deposited much 55 lower energies, of order a few hundred keVs. The BCS overcomes this challenge by using dual-gain preamplifiers from Multi Channel Systems (MCS). The 16—channel preamps have a high-gain amplification for the lower-energy fi signals of 2 V/pC, and a low-gain amplification for the high-energy implantation signals of 0.1 V/pC. The 50 Q low-gain signals from the MCS preamps were impedance matched to the 1 k5) inputs of 32-channel CAEN 785 VME analog-to—digital converters (ADCs). The high- gain signals from the preamplifiers required further signal processing to amplify the small signal. The high-gain outputs of the MCS preamps were sent into PicoSystems programmable CAMAC shaper/discriminator modules. These l6—channel modules have programmable discriminator thresholds and gains, and analog energy and digital timing outputs for each channel. A logical OR of the discriminator output of all 16 channels is also available from each module. Analog energy signals were sent to VME CAEN 785 ADCs, where each block of 16 high-gain signals and the corresponding low— gain signals were digitized in a single ADC module. The digital timing outputs were sent to SIS 3820 VME scaler modules for rate monitoring on a channel-by—channel basis. The logical OR of the 16 channels was used to define the master trigger logic, as discussed at the end of this section, and also sent to a VME CAEN V977 coincidence register for sparse readout. A given 32-channel ADC contained both high- and low- gain signals from one block of 16 DSSD channels. The coincidence register reduced the event size by allowing restriction of the ADC digitization to those modules that had data above threshold. Shown in Fig. 3.5 is a schematic of the electronics for the six SSSDs downstream of the implantation DSSD. Signals from each of these 16—strip detectors were carried, with paired grounds, by 34-pin ribbon cables from the detector to the backplate feedthroughs, and from the backplate to MCS preamps identical to those used with the DSSD. However, in the case of the SSSDs, only high-gain signals were processed. Similar to the DSSD electronics, the SSSD signals were amplified and discriminated using PicoSystems shaper/ discriminator modules. Analog energy signals were sent to 56 SSSD MCS CPA 16Hggh Pico Systems 319W CAEN V785 Preamp Gam Sh./ Disc. (Lmear ADC vFast disc. out) l6-channel OR CAEN V977 Fa In/O t Dela +— Ln ’1’» y Coinc. Reg. OR (of 16 strips for a given SSSD) Figure 3.5: Schematic diagram of the signal processing electronics for the SSSD5 of the BCS. VME CAEN 785 ADCs, with two SSSDS worth of signals in a single ADC, in pairs of SSSDs 1 and 2, 3 and 4, and 5 and 6. The individual channel discriminator outputs were not monitored in sealers. However, the OR of the 16 discriminator. outputs was sent to the VME CAEN V977 coincidence register allowing sparse readout of the ADCs, as was the case in the DSSD electronics. Additionally, for NSCL experiment 07509, a logical OR was taken of the ORs of the discriminator signals from SSSD5 and SSSD6. This signal was used to veto the master gate trigger against light particles that did not stop in the DSSD, and increase the data acquisition live time. Such a veto was not required in NSCL experiment 05101. The electronics for all three upstream PIN detectors were essentially identical. The signal from each detector was sent through a four-channel Tennelec TC178 preampli- fier, and then a Tennelec TC248 shaping amplifier. The slow unipolar output from the shaping amplifier was then sent to a VME CAEN 785 ADC, while the fast timing output was analyzed by a Tennelec TC455 constant fraction discriminator (CFD). One of the CFD outputs from PIN01 was sent to a coincidence register. The set bit for PIN01 was used to control the readout of the ADC containing the signals from all three PIN detectors, as well as signals derived from the PIN detectors, including those from time-to-amplitude converters (TACs) used to measure time-of-flight. Tim- ing signals from the CFDs of all three PIN detectors were distributed to VME SIS 57 Tennelec Tennelec 810 C _ AEN V785 TC178 Quad."" TC248 (Unipd’lar) ADClO Preamp Amplifier VEast utput Tennelec TC 4 5 5 S§S3820 Quad.CFD ca er i 200 ns CAEN V977 PINOI Delay Coinc. Reg. only i CAEN V775 Common TDC (“09) HESS? V Figure 3.6: Schematic diagram of the signal processing electronics for the three PIN detectors used in NSCL experiments 07509. 3820 scaler modules for rate monitoring, and to time—to—digital converters (TDCs) as a stop signal, where the module was gated with a common start derived from the master trigger of data acquisition. Figure 3.6 is a schematic of the electronics for the PIN detectors. The master-gate trigger (MG) for data readout was derived using the discriminator signals from the implantation detector, the DSSD. The MG for experiment 07509 was generated in hardware by taking a logical OR of all of the front strips logically ANDed with a logical OR of all of the back strips of the DSSD. The MG was vetoed by the logical OR of signals from SSSD5 and SSSD6. The MG was more complicated in the case of NSCL experiment 05101. The MG was generated by a logical OR of: (1) DSSD front AND back, as described for 07509, (2) signal above threshold in any of the 16 strips of the most upstream SSSDOl, and (3) signal above threshold in the most upstream PIN01. A master gate live (MGLive) signal was derived in both experiments from MG by requiring a hardware logical AND with a signal from the data acquisition computer indicating not-busy. The MGLive signal was used to trigger data acquisition, as well as to generate digitization gates for all ADCs, TDCs, and 58 MG NIM-ECL 8183820 Fan In/ Out Converter Sealer CAEN V977 Coinc. Reg. Gate CAEN V775 TDC Common Start DSSD CAEN V785 Gate/ Delay DSSD MGLive f Generator ADCs Gate Fan III/Out [ SSSD CAEN V785 Gate/ Delay SSSD Generator ADCs Gate ADClO CAEN V785 1 ! V L— Gate/ Delay * ADClO Generator Gate NIM-ECL 8183820 [Fan 31/03] Converter PH Sealer NSCL Data Acquisition Control Figure 3.7: Schematic diagram of the trigger electronics for data acquisition of NSCL experiment 07509. the coincidence register. A schematic of the trigger electronics for 07509 is shown in Fig. 3.7. A critical requirement for fi-decay experiments with continuous implantation is the ability to correlate implantations and their subsequent decays, and have a measure of the time elapsed between the events. To this end, each event included a timestarnp generated by a SIS3820 sealer module specially configured to act as a 50—MHz clock. The 32-bit clock data word, which had a resolution of 20 ns per clock tick, was scaled into a 24-bit parameter in software. This clock parameter, with a resolution of 5.12 [1.8 per scaled tick, was sufficient for the measurement of fi-decay half-lives in the ms regime. 59 3.3.2 BCS calibrations Each of the silicon detectors of the Beta Counting System was calibrated for energy using a 228Th (1 source. An a spectrum was collected from the 228Th source for each detector, containing sufficient statistics (~100 counts/peak) to allow position determination of the four main (1 peaks below 8 MeV, in all strips including those at the edge of the detector. Each strip was gain matched by adjusting a slope parameter to place the lowest-energy (5.4 MeV) a peak at’channel 350 in a 9—bit histogram in each strip of the detector. Calibrated spectra included, in addition to gain-matching, adjustment by an offset determined by the pedestal of the ADC. A simple shift was applied to each strip to align the first data channel for each strip at zero. Histograms of the raw energy parameter for a representative strip (strip 20) on the front and back of the DSSD are shown in Figures 3.8(a) and (b), respectively. Calibrated spectra for the same detector strips following the gain-matching and offset adjustment, are presented in Figures 3.8(c) and (d). A 90Sr source was used to set energy thresholds for each silicon detector. The fi decay of 90Sr has an end-point energy of 546 keV, but the daughter 90Y, which also undergoes fi decay, has an end-point energy of 2.280 MeV [2]. The signal induced in a 1 mm thick Si detector for such fi energies is a AE signal of order hundreds of keV. Thus, the thresholds of detectors deep within the stack of the BCS detectors could be checked even after other detectors were in place. Hardware thresholds were set for each BCS detector (the DSSD and SSSDs) by comparing the analog output triggered by the CFD output for the same detector channel on a digital oscilloscope. Spectra were then collected for each strip using the 90Sr source, and software thresholds were established based upon the position of the noise signal. The calibrated 90Sr spectra for DSSD front strip 20 and back strip 20 are shown in Figures 3.8(c) and (f), respectively. The arrow in the figures indicates the position of the software threshold, which was applied when determining the signals contributing to sums, etc., in data analysis. 60 431000.2(a)F2ront.20.hienergy £31000 (b) Back. 20. hienergy 8 800; 28Th 3 800 22 23Th 0 600; c) 600- 400% 400- 200;— W 200 G0“‘10072‘01'1‘7300‘400 500 00”‘100“200 300 400 500 Channel Channel £1000:(C)Front.20.hiec @1000: (d) Back.20.hiecal 8 800: 228111 8 300: 228Th c) 600; U 600:— 400; 400;— 200; J] J L 2002— [M 00‘ 100 200 300400 500 00400200300400 500 hannel Channel «EZSOOE (e) Front.20.hiecal «'é’ 2500 (f) Back. 20. hiecal : 905 :3 2000 .— 903r O E O E . 01500; g 1000;— i f software 0E 503E...- .thrsshsld ....... 0 50 100 150 200 250 0 50 100 150 200 250 Channel Channel Figure 3.8: Calibration spectra for a representative front and back DSSD strip from NSCL experiment 07509. (a) and (b) are the raw ADC spectra for a 228Th source on the front and back of the detector respectively, while (c) and (d) are the same spectra following gain-matching and offset adjustment. (e) and (f) are calibrated 90Sr spectra for the front and back strips of the DSSD respectively, from which software thresholds were set. 61 3.4 Segmented Germanium Array Detectors from SeGA [66] were used for monitoring prompt and fi-delayed ’y rays during NSCL experiments 07509 and 05101. Each SeGA detector consists of a single cylindrically symmetric, n-type high-purity Ge crystal, with a diameter of 7 cm, and a length of 8 cm. The core electrode lies along the central axis of the detector. The outside surface of each crystal is electrically segmented into 4 quadrants azimuthally, and 8 x 10 mm thick slices longitudinally. The high segmentation of the SeGA de- tectors is useful for in—beam '7-ray spectroscopy experiments with fast beams, where a Doppler correction to the '7-ray detected energy is critical for achieving the best possible resolution [66]. However for the case of fl-delayed 'y-ray spectroscopy, the nuclei are stopped before ’y-ray emission. Therefore determination of the position of interaction from the segmentation of SeGA is not required, and only the central con- tact signal for each SeGA detector was read out during NSCL experiments 07509 and 05101. A total of 16 SeGA detectors were mounted in a close-packed geometry surround- ing the beam pipe containing the BCS detectors, as shown in Fig. 3.9. The detectors were arranged in two concentric circles of eight detectors each, a configuration known as ,B-SeGA or barrel-SeGA. The crystal of each detector was oriented with its long axis parallel to the beam pipe, face to face with a detector in the opposite ring. The array was positioned such that the implantation detector of the BCS defined the plane between the two rings of detectors. The electronics for the SeGA detectors, and details of the energy and efficiency calibrations for the array are described in the next sections. 3.4.1 Electronics A schematic of the SeGA electronics is presented in Figure 3.10. 'I\1vo identical output signals are available from each SeGA detector from the internal preamplifier attached 62 Secondary Beam Figure 3.9: Geometric arrangement of the 16 SeGA detectors surrounding the beam pipe containing the BCS silicon detectors. to the central contact signal. These two identical signals were distributed through different pathways to provide both timing and energy signals. One signal was sent to an Ortec 863 timing fraction analyzer (TFA) for fast shaping, and then into a Tennelec TC455 constant fraction discriminator. One of the CFD outputs was delayed, and used as the stop signal for a Philips CAMAC 7186H TDC that was started by the MGLive signal. A second CFD output was sent to-a VME SIS3820 sealer module for rate monitoring. The other central contact preamplifier signal was processed by an Ortec 572 shaping amplifier and the unipolar output was digitized by a four—channel Ortec AD413A CAMAC ADC. The ADC was gated by a signal derived from the MGLive trigger. The width of this gate was set to 20 us to permit the detection of isomers with half-lives in the ps range following implantation events. The time between implantations and photon emission was measured by a time-to-amplitude converter (TAC), which was started on the MGLive signal, and stopped by an OR of all of the SeGA CFD outputs. 63 S 8G A Ortec Ortec 572 _ AD413A Gated by Central —~{ Preamp 1, M GLive l Amplifier Quad 8k Contact v ADC PhlllpS Start Ortec 863 _ T311651? 1 7186H signal Quad TFA f ' TDC from _ SIS3820 ' Sealer Figure 3.10: Schematic diagram of the signal processing electronics for SeGA detec- tors. 3.4.2 SeGA Calibrations Energy calibrations were completed for the entire array of SeGA detectors before, during, and after NSCL experiment 07509. TWO calibration sources were used to cali- brate the full 4—MeV range of the y-ray histogram: a 56C0 source, for the high-energy calibration, and a NIST-Standard Reference Material (SRM) source, containing the radionuclides 154’155Eu and 125Sb, suitable from 40 keV to 1500 keV. Energy calibra— tion data were collected by triggering the data acquisition system on a logical OR of the CF D outputs for each of the 16 SeGA detectors. Calibration sources were placed outside of the beampipe and moved during the run to ensure that each detector accu— mulated sufficient statistics, ~50 counts in the 2.6 MeV peak of 56Co, to perform the calibration. The calibration spectra were analyzed using the Oak Ridge Data Analysis and Manipulation Module (DAMM) to determine the centroid and associated error for the location of each ’y-ray transition, based on a Gaussian fit. Energy calibration parameters were obtained from a third-order polynomial fit to the centroid for each peak as a function of the known y-ray transition energy. A third-order polynomial was required to account for both non-linearity of the detector response at low energies, and a previously observed [67] non-linearity in the response of the Ortec AD413 ADCs between low (<1.5 MeV) and high energies. Application of calibration parameters in 64 ‘ 'th. ‘4- software provided calibrated energy histograms for each SeGA detector. The energy resolution of all detectors was 5 3.5 keV full-width at half-maximum (FWHM) for the 1.3 MeV transition in 60Co. During experiment 07509, one of the detectors of SeGA exhibited poor gain stability, even shifting gain during a single hour-long run. Due to the instability of this detector, it was omitted from all analysis, leaving a 15-detector array for this experiment. All 16 detectors of the array were included in analysis for the case of experiment 05101. The residuals from the energy calibration applied to each SeGA detector are pre- sented in Figure 3.11. The residual is defined as the difference between the known 7-ray transition energy and the value deduced from the calibrated spectrum. Vertical error bars are the error in the centroid resulting from the fit of a Gaussian to a given peak in the calibrated energy Spectrum. The residuals for the summed spectrum of all 15 SeGA detectors is given in Fig. 3.12. A conservative systematic error of :l:0.25 keV was attributed to the energy calibration of the SeGA detectors. This error encom- passes the complete distribution of the residuals, and corresponds to approximately twice the root-mean square residual value. The errors quoted for 'y-ray energies in the present work were determined by adding, in quadrature, this error from the en- ergy calibration (i025 keV), and the error in the centroid of the peak locations as determined using DAMM. One of the SeGA detectors, SeGAll, had to be recalibrated due to a gain shift during the course of experiment 07509. Well-known 'y-ray transitions from nuclides in the cocktail beam made a recalibration possible. The 'y—ray transitions used to recal— ibrate SeGAll during experiment 07509 are summarized in Table 3.1. The residuals of the fit to these known transitions before the new energy calibration was imple- mented is depicted in Fig. 3.13(a), while the residuals with the corrected calibration are shown in (b). The recalibration did not alter the 21:0.25 keV systematic error for the full array attributed to the energy calibration. The gains of all other detectors were checked for gain shifts by monitoring the location of well-known transitions in 65 S‘ 0.6 SeGAOI 2- SeGAOS .. SeGA09" SeGA13- U _ . l l l T . i i ,: l 1 j :3 n - ' - 1 l . ' V l , L , 1 ' , l ' - .v ' g V ' 1 1.. . 1 "HT 1.. 1 30.6 . . ‘ 1 , ‘ ‘ . ‘ E | . . ‘ E '5‘ 0.6_ SeGAOZ ; SeGA06 » SeGAIOj SeGA14: é 0;Omitted£rom_r . mi +_.1i'-. . ’ l . ‘u .. 17,. Q - ' ’7‘? ' I " i : analySIS l E , E E 0.6_ SeGA03 7: 1 SeGA07 _j SeGAll :: SeGAlSE u ‘. + in '. t '7. .-. A , l7“ “hi. i l U 1 1. ' 1 1 ' r ‘ .. ' T . _ 1 0.6: SeGA04 _ } SeGA08‘ _: SeGAlZ; SeGA16E ni' 'lnl l l -—».' .l L. 7". .'+ + A‘.* ..L1l A l, u_. ’1 _ .,. q; |-. {11. 1, 41-6“Ass-niinnnl-n--n;nwwn.1; 0 1000 2000 J 1000 2000 3 1000 2000 ) 1000 2000 Energy (keV) Figure 3.11: Residual plots for the 15 individual SeGA detectors included in the analysis of data from experiment 07509. The residual is defined as Elfinm — E2”. :l:0.25 keV 500 1000 1500 2000 2500 Energy (keV) Figure 3.12: Residual plot for the sum of all 15 SeGA detectors used in experiment 07509. A systematic error of i025 keV (marked by the shaded bands) was attributed to the SeGA energy calibration. Table 3.1: Known 'y-ray transitions used to perform a re-calibration of SeGAll during experiment 07509. Source of Transition E7 (keV) 54Sc isomer decay 110:1:1.0 [68] 57T1 decay 174.8104 [69] 57v decay 267.8103 [70] 55T1 decay 3234104 [71] 56V decay 668.4:le.3 [70] 55Ti decay 672.5i0.4 [71] 57v decay 692.4104 [70] 58V decay 879.8104 [70] 5481: decay 10012105 [31] 56V decay 10061103 [70] 58v decay 10564104 [70] “Se decay 14948108 [31] 57Ti decay 18615104 [69] 58V decay 22175104 [70] the data, but no other corrections were necessary. Absolute peak detection efficiencies were deduced for each SeGA detector by plac- ing the SRM standard source at the position of the DSSD within the vacuum chamber. Two calibration data sets were collected in parallel for each detector, one set using the N SCL data acquisition system (DAQ) triggered on the output of the CF D for the SeGA detector of interest, and the other using a PC—based multi-channel analyzer (MCA). The PC-based MCA collected data independent of the DAQ trigger, and recorded data for a dead-time adjusted total of 3600 3, thus assuring a true measure- ment of the absolute detector efficiency. The DAQ system was known to have a trigger rate dependence, but ambiguity in the ADC dead-time correction made absolute ef- ficiency determination less certain. Thus, the two independent systems were used to cross-check one another and ensure a good determination of the 7-ray efficiency of the array. Peak efficiencies were deduced by comparison of the observed emission rates as detected by SeGA with the known activity of the isotopes in the SRM source. A rel- ative efficiency measurement for photon energies above 1.5 MeV was also completed using a 56C0 'y-ray source. Data was taken using only the DAQ system, triggered by 67 (a) Original Calibration g—e fITTlllllllllllllllll lllllllll an Residual (keV) O [1 1111 llllLLlJlllllllllll llllllLLlllllleLlllAl (b) Re- fit Calibration [ l] 1[Hl[l[ [ 400 800 i200 160011’2000‘l‘2‘400 Energy (keV) 0.6 0.4 0.2 Residual (keV) -0.2 -0.4 -O.6 r—.—_—. [[llllTTlllll llllllllllllllrglflllllTl—rl on Figure 3.13: Residuals for the energy calibration of SeGAll during gain-shifted runs using (a) the original energy calibration, and (b) a re-fit energy calibration based upon known 'y-ray transitions within the data. 68 log [Peak Efiic1ency] b c': \O \) ‘ O y = -0.14x5 + 1.31x4 - 3.82x3 X + 0.33x2 + 13.99x - 16.12 -1.5 ‘9 '1. i t g = 71.5 2.0 2.5 3.0 3.5 log [Energy (keV)] Figure 3.14: Peak detection efficiency curve for N SCL experiment 07509. Data points represent measured efficiences using the NSCL data acquisition system. Data were fitted with a fifth-order polynomial, denoted by the solid line. a. logical OR of the individual SeGA detector CFD outputs for the relative efficiency measurement. Peak efficiencies were deduced from the 56Co data by assuming the same absolute efficiency for the 873.2 keV transition in 154Eu from the SRM source and the 846.7 keV transition in 56Co, and normalizing the other transitions in 56Co to this value based on known relative intensities. The resulting efl'iciency curve for SeGA for experiment 07509 is shown in Figure 3.14, and was fitted with a fifth-order polynomial. The peak efficiency for the array corresponds to 6.6 % at 1 MeV and 3 ‘70 at 3 MeV. Similar peak efficiencies were obtained, and an analogous efficiency curve produced for NSCL experiment 05101. 3.5 Data analysis: Correlations and data fitting The previous sections of this chapter provided details of the physical experimental set-up used in NSCL experiments 05101 and 07509, including the beam delivery, detector systems, and electronics. This section describes the other critical component of correlation experiments with continuously implanted beams: the logic and methods 69 used to build up correlations from the data and to extract the values of interest, including half-lives and decay branching ratios, in a robust manner. 3.5.1 Implantation-decay correlations The method of continuous implantation fi-decay spectroscopy relies on the ability to correlate individual implantation events with their subsequent 6 decays. The method makes use of both position and time information to perform the correlations. Each event was evaluated in software to determine the nature of the event. Raw signals from each detector were calibrated and compared to software-defined upper and lower thresholds. Software flags were set to indicate valid data for those detectors in which signals satisfied the threshold requirements. Additionally, for the segmented detectors in which multiple strips may have had valid signals for a given event, the energies observed in each strip were evaluated to localize the event to a single strip. The spatial position, or pixel (56,31) of an event in the DSSD was defined to be the front and back strip with the largest recorded energy signal. Implantation events were identified as those with a valid signal in each of the upstream PIN detectors (PINl, PIN2 and PIN3) as well in the low-gain histograms of both a front and back strip of the DSSD. Implantation events additionally required the absence of a valid signal in SSSDl. When these conditions were met for a given event, the event was positively identified as an implantation and the AE, time—of-fiight and time—stamp information was stored within a two-dimensional array in software, in the array element corresponding to the (113,31) position of the pixel of the DSSD to which the implantation had been localized. Decay events were identified as those events which had no valid signal in any of the upstream PIN detectors but had a valid signal in the high-gain histograms of both a front and back strip of the DSSD. Once an event was assigned as a decay, a series of checks were made based on time and position to correlate the decay event with a previous implantation. The (11:, y) position of the decay event was checked to 70 see the last time that an implantation event occurred at the same geometric location within the DSSD. Only identical (as,y) coordinates were considered in the case of experiment 07509 (implantation rates of ~100 Hz). When rates were sufficiently low, as in the case of the other experiment 05101 (average implantation rate ~30 Hz), the correlation position was extended to also allow correlation of a decay with an implantation in any of the eight nearest neighbor pixels (mil, yil). If an implantation event had occurred, an additional check was made to ensure that there had not been back-to-back implantations in that pixel within a short time frame, which could cause ambiguity in assigning a decay to one implantation over the other. Once the integrity of the implantation was verified, the time between the decay event and the implantation was checked to ensure that the event had occurred within a pre—defined correlation time window for each isotope, established as some multiple of the expected decay half-life. With the geometric and temporal conditions met, the decay event was considered to be positively correlated with a previous implantation. Decay curves were constructed for all isotopes by histogramming the time differ— ence between the implanted ion of interest and the correlated fl decay. Decay curves were fitted as described in Section 3.5.2 and the half-lives deduced for each isotope. Additionally, by integrating the decay curve for the parent isotope contribution, the total number of detected 6 decays for the parent was deduced. Combined with the total number of implanted ions from the particle identification spectrum, the aver- age ,B-detection efficiency of the DSSD was calculated for each isotOpe. The average ,B-detection efficiency for all isotopes studied in NSCL experiment 07509 using a single-pixel correlation was found to be 11.4i0.4%. The basic requirements for correlation of an implantation and its subsequent decay are well established. However, there are a number of variables which can be adjusted to optimize the correlation efficiency based on the experimental conditions and expected properties of the decay. The correlations are determined entirely by the constraints set by the geometry and timing in analysis. The following section describes the adjustable 71 variables in the correlation procedure, and the choices made in the present analysis based on implantation rates. Implantation rates and correlation limits As discussed in the previous section, the two time frames critical to the correlation procedure are the minimum time between back-to—back implantations in a pixel, and the correlation time window, corresponding to the maximum elapsed time over which an implantation would be correlated with a decay. These time windows were adjusted in the data analysis on an isotope-by-isotope basis to optimize the correlations, using the expected half-life of the parent isotope. The correlation time windows were chosen to be approximately ten half-lives of the parent nuclide, permitting a good determi- nation of the parent, subsequent generations, and the background contributions to a decay curve. The minimum time between back—to—back implantations was set to match the correlation time. To minimize random correlations certain conditions related to the rate of events in the detector must be met. The implantation rate must be sufficiently low that, on average, the time between back-to—back implantations in a pixel is greater than the length of the correlation time window. The overall implantation rate for experiment 07509 was ~100 Hz. However, these implantation events were not distributed evenly over the detector surface. The beam profile was such that pixels near the center of the detector experienced a higher implantation rate than those at the edge of the detector. An average of less than 5 s separated implantations in the most central pixels of the implantation DSSD. The average time between implantations for each pixel of the surface of the DSSD is shown in Fig. 3.15(a). However, for most nuclei included in experiment 07509, even the highest implantation rate allows a sufficiently long correlation time window. A more significant problem was posed by the build- up of activity from subsequent generations in pixels with higher implantation rates. With an average of four 6 decays to reach stability from each implanted fragment, 72 the accumulated activity in the detector creates a background of random correlations in all spectra. In Figure 3.15(b) is shown the average number of observed fl-decay events occuring in a one-second time interval in each pixel of the detector. Assuming the average ,B-detection efficiency of 11.4:t0.4%, the average rate of ,6 decays is more than 3 / s in the most central pixels. The higher background rates increase the chance of random correlations, and both *y-ray spectra and half-life curves will have pixel- dependent background contributions. The non-uniformity of the beam profile was used to provide an alternate analysis of decays which required a longer correlation time window. A cut was made in position based upon the time between back-to—back implantations, and only a subset of pixels corresponding to those with >100 s between back-to-back implantations was used for the primary analysis of isotopes with longer half—lives. The pixels satisfying the cut are shown in Fig. 3.15(c). The back-to-back implantation rate requirement was used in software to limit analysis to the N800 pixels, or half of the detector surface on the edge of the detector with lower rates. Imposing this restriction did not affect the average fi-detection efficiency for a single-pixel correlation, but did result in a reduction of statistics by a factor of ten. The benefits of making a restriction on the back-to-back implantation rate are best illustrated in an example. The 6 decay of 55Ti is reported to have a half-life of 1.3:l:0.1 s, and 9 7—ray transitions have been previously reported [71]. 55Ti was a component of the cocktail beam in NSCL experiment 07509, and its decay was examined for both the subset of pixels of the DSSD shown in Fig. 3.15(c), and all pixels, with a correlation time window of 5 s. The resulting fi-delayed 'y-ray spectra are presented in Figs. 3.16(a) and (b) respectively. The known transitions in 55Ti and its daughter 55V are marked by the filled circles. The background contribution to the spectrum of Fig. 3.16(a) is reduced in the spectrum of Fig. 3.16(b), while the known transitions are apparent in both spectra. The half—life curve is also affected by background, as illustrated in Figs. 3.16(c) and (d) The decay curve generated from 73 v—tr—A O 031D- (a) Time between implantations (s) <5. Time between implantatlons (s) 102 101 DSSD Back Strip Figure 3.15: Rate profiles over the surface of the implantation DSSD. (a) shows the average time, in seconds, between implantations for each pixel of the DSSD. (b) is the average number of decay events in a given pixel within a 1 second time interval. (c) illustrates the pixels of the DSSD in which the time between implantations is 2100 s. This rate requirement, which encompasses 1/2 of the DSSD pixels, was applied in analysis to select a subset of pixels. 74 E 2000:; (a) All pilJSCIS a F (C) T1/2([3) = 2.2:h0.l ms N 1600:- x 8 .. All pixels a 1200: $104. ‘5 ° ° :3 : 8 800?” g U 403: 8 55TI 3, gogw)..Edge pixilf ‘ " ‘é’ (:d)T1/2(B)- — 1. 310. 2ms N 605— 0 x 8 Edge pixels \ . N :2 g t: 402- 4.: 8 7 ‘3 55Ti 20; 0100 bk d U 0: U C' ............................ g..- 0 400 800 1200 1600 2000 0‘” i000 2000 3000,4000 5000 Energy (keV) Time (ms) Figure 3.16: (a) and (b): fi-delayed 'y-ray spectra following the decay of 55Ti for data taken from (a) all pixels of the DSSD, and (b) only pixels near the edge of the DSSD, in which the time between implantations was 2100 s. (c) and (d): B-decay curves derived from data using (c) all pixels of the DSSD, and (d) pixels near the edge of the DSSD. all data in Fig. 3.16(c) was fitted with an exponential decay for the parent, the growth and decay of the daughter 55V, and a constant background. The resulting fit gave a half-life a factor of two longer than the known value. The decay curve resulting from data in the edge pixels only [see Fig. 3.16(d)] was fitted in the same manner and the deduced half-life is in good agreement with literature [71]. Primary analysis for data from NSCL experiment 07509 was restricted to the subset of pixels with low implantation rates as described above when the required correlation time window was greater than 1 s. This includes the analysis of the ,6 decays of 53Ca and 5‘I'Sc, discussed in the next chapter. The restriction to a subset of DSSD pixels allowed a clean analysis of the decay properties of these longer-lived isotopes. However, it would be preferable in future experiments to further defocus the beam over the detector surface to reduce the implantation rate in the central detector pixels. The loss in statistics due to fragments missing the detector would likely be less than that arising from limiting the analysis region. The implantation 75 rate was considerably lower for NSCL experiment 05101; the total implantation rate was ~30 Hz. It was therefore not necessary to limit the analysis to any specific region of the DSSD. 3.5.2 Decay curve fitting methods One of the key values that can be extracted from fi-decay spectroscopy data is the half-life of implanted nuclei. A critical step in extracting a half-life from continu- ous implantation data is the correct correlation of implantation events with their subsequent decays, which was described in the previous section. Equally important, however, is the method used to fit the decay curve data with functions describing not only the decay of the parent nucleus of interest, but also contributions from decay of the daughter isotope(s) and randomly correlated background events, which must be included due to the <100% fl-detection efficiency of the DSSD. Typically, fitting techniques require the minimization of a “goodness of fit” pa- rameter, which often takes the form of a X2 value. A commonly used parameter is the Gaussian X2» defined as follows [72]: N 2 2 (311 - 3011(3)) XGaussian = Z 02 a (3.1) i z where y,- represents the 1th data point, with associated variance 01-, and y f,t(a) is the fit to the data point, which is a function of the set of fit parameters, a. The best fit parameters are determined by minimizing the value of x2 iteratively, by considering the numerical derivatives of x2 with respect to each of the fit parameters, a. Data from low count-rate experiments follow Poisson statistics [4], and the use of X2Gaussian is not appropriate. An alternative method for nuclear counting experi- ments with low statistics is to use the maximum likelihood of the Poisson probability 76 distribution. The likelihood function (L) for Poisson statistics is given by [73]: N N [Jyze—u £=Hamw=fl W, (M) t Z where Pp(y,-, p) is the probability of measuring a value of y,, when the true mean of the distribution is given by p, for N events. Taking the natural logarithm of Eqn. 3.2, the log-likelihood function is given by: N ln£ = Emmy-p—lnyil]. (3.3) i=1 With some rearrangement of Eqn 3.3, a version of the X2 statistic derived from maximizing the log-likelihood function for Poisson distributed data can be expressed as: N 2&1; = ZZlyfit — yi 1n(yfit) + [M31101 (3.4) 2 where llfit is the fit to the data point yi. Minimization of this xi L correctly determines the fit parameters for Poisson distributed data, irrespective of the procedure used to histogram the event data. Results are discussed in the next chapter which include half-lives deduced using two distinct methods: a curve fitting method based within the ROOT data analysis framework [74], which optimized a fit to the binned decay curve data, and a maximum likelihood approach, where each instance of a correlated decay event is evaluated and associated with a specific member of a decay chain, or to background. Both of these methods used the Poisson statistical approach and are outlined in more detail in the following sections. 77 Functional curve fit with ROOT The majority of cases under consideration in the present study had good statistics, and the half-lives were deduced from the decay curve data using a traditional curve-fitting approach. A ROOT-based fitting program was written that provided a graphical user interface and the option to customize the fitting function based on the properties of a given decay. Fit functions were based on the Bateman equations for nuclear decay processes including parent, daughter, and grand-daughter decays as required, with the option of including [B-delayed neutron branching and the associated fin daughter decays. Two fitting algorithms were available with the ROOT package: minimization of a Gaussian X2, defined as in Eqn. 3.1, and maximization of the Poisson likelihood function by minimization of the xi L statistic as defined in the previous section. The latter method was applied for all decay curve fits discussed in Chapter 4. Maximum likelihood analysis The Maximum Likelihood (MLH) fitting method was based upon the maximization of a custom log-likelihood function defined for a given data set. As applied to the analysis of nuclear decay half-lives, the MLH method provides a mathematically cor- rect description of the probability of observing chains of decay events, and thus makes use of all available information, on an event-by-event basis, to deduce the unknown decay properties. The MLH method and its application to low-statistics fl-decay data are described in detail elsewhere [65]. No B decays in the present work were analyzed using the MLH method to deduce a half-life value. However, the method was adapted to treat low-statistics isomer 'y-ray decay data for 50K, discussed in the next chapter. The nature of 'y-ray decay, with discrete energy transitions, allowed simplification of the MLH method as compared to the treatment of 6 decay. As it is possible to discriminate 'y-ray decays based on energy, there are essentially only two options for the identity of a decay event — a real parent 'y-ray decay, or a random background event. Contributions from the decays 78 of subsequent generations in the likelihood function can thus be excluded. Gating on a specific 7-ray energy also significantly reduces the background rates associated with any given isomer decay event. The elimination of contributions from subsequent decays and background greatly simplifies the MLH method, reducing the Poisson log-likelihood maximization technique to treatment of individual data points. 79 Chapter 4 Experimental Results The previous chapters have outlined the techniques of B decay and fi-delayed and prompt 'y-ray spectroscopy, as well as the details of the experimental set-up for NSCL experiments 05101 and 07509. Results from these experiments are presented in this chapter. Prompt 'y-ray information provided information on as isomers for nuclei implanted in both experiments. fi-decay half-lives and ,B-delayed y-ray spectra were also obtained from implantation-decay correlations as described in the previous chapter. 77 matrices were constructed for events with 'y multiplicities of two or higher, allowing analysis of '7 coincidence information. These data were used to establish level schemes. Combined, the data obtained permitted considerable expansion of the knowledge of the neutron- rich f p shell nuclei. Experiment 07509 included over 20 neutron—rich nuclei in the region surrounding 540a, and provided new information on the quantum structure of the neutron-rich 53’54’3?Sc isotopes, both through the decay of the parent 53’33Ca nuclides, as well as via prompt ’y-ray emission from isomeric states in the even-A 54'56Sc isotopes them- selves. Prompt 7 rays from §8K were also observed in experiment 07509, providing information on the low—energy structure of this nuclide. New data for the low-energy structure of neutron-rich géMn was obtained from experiment 05101, via 6 decay from 80 the parent giCr isotope. This chapter presents results, including fl-decay half-lives, fi- delayed 'y-ray information and prompt ’y-ray information, for 50K, neutron-rich 21Sc and 2008. isotopes from N =32 to N =35, and 61Cr. 4.1 Low-energy structure of neutron-rich 218C isotopes The presence of subshell gap at N :32, the result of a sizeable energy spacing between the V2123 /2 single-particle orbital and the higher-lying V2p1/2 and u1f5/2 orbitals, has been observed experimentally in the 20C&, 22Ti and 24Cr isotopes. Additional to this subshell closure, the possibility exists in the 20Ca isotopes for a N :34 subshell closure, resulting from the continued upward shift of the V1f5/2 orbital attributed to a reduced monopole interaction. Confirmation of the presence or absence of a N =34 subshell closure in the Ca isotopes has been the focus of a number of experimental efforts in the f p shell. The primary goal of experiment 05101 was to investigate the 5 decay of 54K to states in 54Ca, with the hopes of populating the first 2+ excited state in 54Ca. Observation of the ’y—ray de-exciting the 2] state in 54Ca would have provided a direct measurement of the excitation energy of the state, and when considered in the systematics of the 20Ca isotopes, potentially the first evidence for a N =34 subshell closure. However, only 73 54K implantations were observed after one week of running time. The production yields proved prohibitively low for a successful ,8—7 measurement during the allotted time. Experiment 07509 was also motivated by the possibility of observing the ’y-ray decay of the 2] state in 54Ca and measuring E(2f). Within the data of experiment 05101, 3 counts were observed at 327 keV in the ’y-ray spectrum collected within 15 ,us following 54Ca implantations. Could an isomeric state be present in 54Ca? The possibility for such an isomer arises when low-energy 4' and 5' negative parity states, an outcome of the coupling of the ”2121/2 and 1x1 99 /2 orbitals, are considered. 81 _110 keV (54Sc contamination) 00 1000 2000 3000 4000 Energy (keV) Figure 4.1: 'y-ray spectrum collected within 20 [IS following a 54Ca implantation. No evidence of an isomer decay is observed. The peak at 110 keV is due to random correlations with the 'y ray in the isomer decay of 5430 The isomeric decay from a 5" state in 54Ca would be expected follow the sequence 5‘ —> 3] -—+ 2] —> 0]". Thus, the goal of experiment 07509 was to investigate the possible isomeric decay of 54Ca, and measure the energies of the ’y—ray transitions involved, in particular that of the 21L ——1 0:“ transition, expected to be 22 MeV. The 'y-ray spectrum collected within 20 us following a 54Ca implantation in experiment 07509 is shown in Fig. 4.1. No evidence of an isomer was observed. While the structure of 54Ca has proven difficult to probe directly at present, the possibility of an N =34 subshell closure at Z =20 can be investigated indirectly by considering the level structures of neighboring isotopes, e.g. the 21Sc isotopes. The low-energy structure of the Sc isotopes should be described by the coupling of the single valence f7/2 proton to the valence neutrons. Thus, the low-energy levels in the 218C isotopes surrounding N :34 should be sensitive to the neutron single-particle energy spacings, and may shed light on the possible N =34 subshell closure in the Ca isotopes. Four neutron-rich 21Sc isotopes from N =32 to N =35 were included in the cocktail beam delivered to the NSCL 6 counting system in experiment 07509, as well as three 20Ca isotopes from N :32 to N =34. The 20Ca isotopes provided access, via 82 their 6 decays and 6—delayed y-ray emissions, to the low-energy structure of their daughter 2180 isotopes. Prompt 'y-ray emission in 54Sc complemented the decay data, providing additional insight into the low-energy structure of this isotope. Prompt 7 rays provided the only access into the low-lying levels of 568C, for which the parent 56Ca was not produced with sufficient yield to permit 6—7 analysis. The results for the low-energy structures of the Sc isotopes, determined both from decay of the Ca isotopes and prompt 'y-ray emission from the Sc isotopes themselves, are presented in the following sections. 4.1.1 6 decay of 53Ca to 53Sc The 6—delayed ”y-ray spectrum for 53Ca in the range 0 to 3 MeV for decay events within 5 3 following an implantation event is shown in Fig. 4.2. Due to the length of the correlation time window used in the analysis of the 53Ca decay, this spectrum includes only decays correlated with implantations isolated in pixels towards the edge of the implantation detector, as discussed in Section 3.5.1. One transition is apparent in this spectrum at 2109.0:l:0.3 keV, and has been assigned to the 6 decay of 53Ca. This newly assigned transition has recently been confirmed in one-proton knockout from 54Ti [75]. The decay curve constructed from 53Ca—correlated 6 decays with the additional requirement of a coincident 2109—keV 7 ray is presented in Fig. 4.3(a). Extraction of the half—life from the full data without a 7-ray coincidence requirement is complicated by uncertain quantities, including daughter half-lives and neutron branching ratios, which must be included in the fitting function. It is therefore desirable to additionally gate the decay curve on a known *y ray, to remove contributions from both background and subsequent decays in the decay chain. The newly observed 2109—keV "y ray could be used in the case of 53Ca, and the full statistics of the DSSD could be used in this case — the additional 2109-keV 7-ray coincidence requirement reduced background events, allowing extraction of a clean decay curve. The 'y-gated decay curve was 83 VI 2109 (40 Counts {3 keV N II I IIII IIIT ITrr] ‘ l [[1 1 . ~ [llllllll 00 500 1000 1500 2000 2500 3000 Energy (keV) Figure 4.2: 6—delayed 'y-ray spectrum in the energy range 03 MeV following the decay of 53Ca. The spectrum is limited to include ’y rays coincident only with 6 decays in the subset of pixels near the edge of the DSSD. The transition assigned to the 6 decay of 53Ca is marked by its energy in keV. t—d fitted with a single exponential decay and a constant background, yielding a half-life value of 461i90 ms. This value is higher than the value of 230i60 ms previously reported by Mantica et al. [76]. However, as noted in Ref. [76], the possibility of a second 6—decaying state in 53Ca, which could account for the discrepancy between the present result and previous measurements, cannot be excluded. A fit to the decay curve without the additional 7 requirement is required to determine the total number of observed 53Ca 6 decays. The fit to the un-gated decay curve, considering only decay correlated with implantations towards the edge of the implantation detector, and fixing the half-life at 461 ms, is presented in Fig. 4.3(b). An absolute 'y intensity of 56:l:l2% was determined for the 2109—keV transition from a Gaussian fit to the peak in Fig. 4.2, the absolute efficiency of the 15-detector geometry of SeGA (4321:0270), and the number of 530a 6 decays (252) that were correlated with fragment implantations, as determined from the fit to the decay curve of Fig. 4.3(b). The absolute intensity of the 2109-keV transition suggests that the majority of 6 intensity from the decay of 53Ca, excluding 6n contributions, proceeds through an excited state depopulated by this transition. The 2109-keV transition is thus proposed to directly feed the 53Sc ground state, and the associated 2109-keV 84 (a) T1/2([3-y) = 461190 ms Counts/ 250 ms All I”! Counts/200 ms 10”” 1000 2000 3000 4000" 5000 Time (ms) Figure 4.3: 6-decay curves for the decay of 53Ca. (a) Decay curve constructed consid- ering 6-decay events correlated with 53 Ca implantations over the entire surface of the DSSD, with the additional requirement of a coincident 2109-keV decay 'y ray. These data were fitted with a single exponential decay and a constant background. (b) De- cay curve constructed considering only data in the subset of pixels near the edge of the DSSD, with a correlation time of 5 s. Data were fitted with an exponential decay of the parent, with the half-life fixed at 461 ms, growth and decay of the 6 and 6n daughters, assuming a 40% neutron-branching ratio [77] and a constant background. 85 (1/2)_ 0 53Ca T1/2 = 461 i' 90 ms Q; = 9.2 i 0.3 MeV Sn : 53 i 0.4 MCV £4. ‘5‘,“ ,__.,_.,§.‘._, '1': \\ 3(+) 0 9” 523C 1‘3 (%) 10g fi‘ Q0] /2- \ 56i12 4.5:0.2(3 ) "’ 2109'” (7/2)- H 0 53Sc Figure 4.4: Decay scheme for the decay of 53Ca to states in 5380. The number in square brackets following the 'y-ray energy is the absolute 7—ray intensity. Apparent log f t values were calculated based on the apparent fl-feeding values, the deduced half- life and the excitation energy of the state [79]. The decay Q value, Q3, and neutron separation energy, Sn, were deduced from data in Ref. [80]. state is tentatively assigned spin and parity (3 / 2)‘, on the basis of selection rules for allowed [3 decay from the presumed (1 / 2)“ 530a ground state [76,78]. The proposed level scheme for 5380 following the ,8 decay of 53Ca is shown in Fig. 4.4, including the apparent fl-feeding branches and log f t values. 4.1.2 6 decay of 540a to 54Sc The fl-delayed 7-ray spectrum following the decay of 54 Ca in the range 0 to 2 MeV is shown in Fig. 4.5. With the short correlation time of 1 3, data from the entire surface of the DSSD could be included in the analysis. One transition at 247.3i0.3 keV has been assigned to the 6 decay of 540a. The absolute intensity of this single transition was deduced to be (65i9)%. This transition corresponds to the 7 ray previously observed by Mantica et al. [76] at 246.9i0.4 keV with an absolute intensity of (97:1:32)%. The 247—keV transition is placed as a ground—state transition and the 247-keV state 86 360:— 247 N50:— >, E 54°? 83°;— §Cr 20g- 10:— 0 500 1000 1500 2000 Energy (keV) Figure 4.5: fi-delayed 'y-ray spectrum following the decay of 540a to states in 54Sc, in the range of 0-2 MeV. Observed transitions assigned to the decay of 54Ca are marked by their energy in keV, while transitions in the 6 decay of the daughter 54‘Sc are marked by a filled circle. assigned J 7r:(1)+ based on the apparent allowed 6 decay to this level from the 0+ 540a ground state. As discussed by Mantica et al. [76], the 247-keV state populated in the decay of 540a de-excites promptly (within a few us). A prompt 247-keV transition from a 1“” state to the 54Sc ground state precludes a J "24+ assignment for the 54Sc ground state, as Weisskopf estimates for a 247-keV M3 transition would suggest a half-life of order days. The observed fl-feeding pattern to states in 54Ti from 54Sc had previously limited the ground state spin and parity to (3,4)+ [31]. The present result suggests a 3+ assignment for the ground state spin and parity of 543C — a 247- keV E2 transition is expected to have a half-life in the ns range, consistent with the observed prompt nature of the transition. The unobserved ,8 intensity of 35d:9% can be attributed to a possible fin branch in the 540a decay, as direct feeding to the (3)+ ground state of 54Sc is unlikely from the 0'Jr ground state of 54Ca. The decay curve derived from 54Ca-correlated fl decays within 1 s of an implan- tation event is shown in Fig. 4.6(a). The decay curve of Fig. 4.6(a) was fitted with a single exponential decay for the parent 540a, exponential growth and decay of the [6 daughter isotope, 54Sc (T1 /2 = 526 :l: 15 ms, see Section 4.1.3), and a constant background component. The deduced half-life for the decay of 540a was 101:l:9 ms, 87 : (a) T1/2(B) = 104m ms Counts! 25 ms 53 \ \ [‘1‘ ll /2([3'Y) = 107:]:14 ms 111A L “nil.” . (b) T1 Counts/25 ms 8 l Til—ILL y—n ' I LII A 0 E —/ "." Q llllllltlllll lLlLlll‘ l .111 I’LILJJ 0 200 * 400 600 so ‘iboo Time (ms) Figure 4.6: fi-decay curves for the decay of 54Ca. (a) Decay curve constructed from 54Ca—correlated B—decay events over the entire DSSD surface, using a correlation time of 1 5. These data were fitted with an exponential decay of the parent, growth and decay of the 6 daughter, and a constant background. (b) The decay curve for 540a, with the additional requirement of a 247-keV decay 'y ray coincident with the 6 decay. The curve was fitted with a single exponential decay and a constant background. within 10 of the value of 86i7 ms reported by Mantica et al. [76]. The difference in the deduced half-life value is a result of the new, longer half-life for the 54Sc daughter nucleus (see Section 4.1.3). A fit to the present data using the previous 54Sc half-life of 360i60 ms [31] used by Mantica et al. gave a half—life value of 88i13 ms. Shown in Fig. 4.6(b) is the decay curve obtained by requiring a coincidence with the 247—keV B-delayed 'y ray observed in the 540a decay. The 247-keV *y-gated decay curve yields a half-life value of 107i14 ms, again consistent with both the un—gated decay curve ' and the previous value reported by Mantica et al. [76]. The proposed decay scheme for levels in 548C populated following the 6 decay of 540a is shown in Fig. 4.7. 88 54 Tl/2= 107i14 ms Ca \zf 10.3:03 MeV E4444 ~- 444 44 Bu: (3549)% Sn: 4,740.5 MeV *" 6:8 (7/2)-53 0 9‘03” Sc 1,, (%) logfl‘ «(4’ + “lab, 65:9 45:02 (I) 2473 O)" o 54SC Figure 4.7: Decay scheme for 54Ca. The energy of each state is given in keV, and the number in square brackets following the 'y-ray energy is the absolute 7—ray intensity. Qg and Sn were deduced from data in Ref. [80]. 4.1.3 3 decay and isomeric structure of 54Sc Analysis of the 6 decay of 54Se used a correlation time of 5 s, with primary analysis restricted to the subset of pixels towards the edge of the implantation DSSD, as discussed in Section 3.5.1. The fi-delayed ’y-ray spectrum for 54Sc in the range 0 to 3 MeV is presented in Fig. 4.8. Twelve transitions were identified in this spectrum. The four lowest energy transitions were eliminated as candidates for transitions in the decay of 54Sc on the basis of their apparent half—lives, and likely belong to the decay of the daughter nucleus, 54Ti. 7*)! analysis indicates that these four transitions are in coincidence with one another, and the lowest energy transition at 108 keV corresponds to a known transition in 54V [68]. The remaining eight transitions are assigned to the 6 decay of 54Sc, and are summarized in Table 4.1. The transitions observed at 1002, 1021 and 1495 keV were observed previously in 3 decay [31] and deep-inelastic studies [81]. The decay curve for 54Sc—correlated decay events including decay events over the 89 \l C rérlrlr . A l 495 Counts/keV O Au: O [I 1111 IIITFII NW CO p—u O O o 500 1000 1500 2000 2500 3000 Energy (keV) Figure 4.8: fl-delayed 7-ray spectrum following the decay of 54Sc, for 'y-ray transitions in the energy range 0-3 MeV. Transitions in the decay of 54Sc are marked by their energies in keV, while transitions belonging to the decay of the daughter are marked by a filled circle. Transitions marked by shaded triangles were used to gate the fl-decay curve of Fig. 4.9(a). Table 4.1: Energies and absolute intensities of fi-delayed 'y rays assigned to the 6 decay of 54Se. The half-life deduced from a 7—gated decay curve, considering data over the entire surface of the DSSD, is also included for all transitions in which statistics were sufficient to fit a curve. E, (keV) Isbsolute (%) Initial State (keV) Final State (keV) Half-life (ms) 484.6 :l: 0.4 4 :l: 1 3000 2516 564i99 840.5 :l: 0.4 7 i 2 3338 2497 620i104 1002.4 :t 0.3 40 i 4 2497 1495.0 522:1:28 1020.8 :l: 0.4 9 :l: 2 2517 1495.0 556i53 1495.0 :l: 0.3 79 :l: 5 1495.0 0 525i19 1504.0 :l: 0.3 2 :l: 1 3000 1495 521i113 1965.7 :h 0.4 7 :l: 1 3460 1495 4582!:100 2517.5 :t 0.3 5 :l: l 2517 0 - 90 entire surface of the DSSD, with the additional requirement of a coincident 1002, 1021, 1495, 1504 or 1966-keV 7 transition, is presented in Fig. 4.9(a). A fit to these data using a single exponential decay and constant background yielded a half-life for 54Sc of 526:1:15 ms, longer than the value 360i60 ms determined by Liddick et al. [31]. One possible explanation for the discrepancy between the present half-life requiring a coincident decay 7 ray, and the previous value would be the presence of a second ,B-decaying state in 54Sc. However, as noted by Liddick et al. [31], the ordering of low- energy states predicted by the shell model is not supportive of a ,B-decaying isomer. Additionally, decay curves gated on individual 7 transitions yielded half-life values in good agreement with one another (see the last column of Table 4.1, which is not indicative of an isomer. Another possible explanation for the discrepancy between the half-lives is the deconvolution of the previously observed fl-decay curve. A half-life of 2.1:l:1.0 s, used in the present analysis, was determined in the 07509 data set for the daughter 54Ti and is in agreement with, though slightly longer than the literature value [82]. Additionally, the unobserved fl intensity in the present work is suggestive of a neutron-branching contribution of ~20%. Inclusion of these factors, not considered by Liddick et al. [31], would result in an underestimation of the parent half-life in the previous work. The proposed levels in 54Ti populated by the 6 decay of 54Sc are shown in Fig. 4.10. Placement of the 1002-, 1021— and 1495-keV transitions follows the assignments made in previous work [31,81], and is confirmed by both the absolute '7-ray intensities of the transitions, and observed '77 coincidences. The new transitions with energies 484, 840, 1505 and 1965 keV were placed on the basis of the observed 77 coincidences, as shown in Fig. 4.11. Energy-sum relationships were used to place both the 1505-keV cross-over transition connecting the states with energies 3000 and 1495 keV, and the 2517-keV transition between the 2517-keV level and the ground state. The apparent fl feeding and log ft values deduced from the absolute '7 intensities are included in Fig. 4.10. J7r for the states with energies 1495, 2496 and 2517 keV are 91 (a) T1/2(B-y) = 5265:15 ms H I" I I T11 Counts/ 100 ms ES ., . .... fl §. . III! 1 VIII Counts/100 ms t—O ..§ 0“" 1000 2000 3000 4000 5000 Time (ms) Figure 4.9: Decay curves for the H decay of 543C to states in 54Ti using a correlation time of 5 s. (a) Decay curve, considering data over the entire surface of the DSSD, with the additional requirement of a coincident 1002, 1021, 1495, 1504, 1966 or 2518- keV 7 ray (marked by the shaded triangles in Fig. 4.8). These data were fitted with a single exponential decay and constant background. (b) 5480 fl-decay curve constructed considering fl decays localized to the subset of pixels on the edge of the DSSD. The decay curve was fitted with a function including a single exponential decay, growth and decay of the daughter, ,Bn daughter and granddaughters, and constant background, assuming a neutron branching ratio of 20%. The half-life of the parent 54Sc was fixed at 526 ms, as determined from the 7-gated decay curve. 92 XS?‘ (4.5)+ 0” 110 (3)+ l 0 Tl/Z = 526 i 15 ms 54SC QB: 11.4:0.4Mev M44974 8,, = 6.8 : 0.2 MeV 9,031; X/ ’ In (‘4') 1°31? gs“ XS -°‘\ «bx/‘10 3460 7:1 57:01 ‘9 193: 79\t.\"" x 3337 7:2 58:02 6,39 49°‘—\6x>\\x%\ x} a 6:1 59:01 ‘* ‘6‘“ «6,901,901,; 3000 + - Q- - 10:2 58:01 :+ v 15‘ (0’2 9°" /—-25‘7 33:4 53:01 fl w 0“” a 66"“. 21:7 57:02 (2" ‘ " v ‘99 1495.0 0+ 11 ll 0 54Ti Figure 4.10: Proposed decay scheme for 54Sc to states in 54Ti. The energy of each state is given in keV. The number in brackets following the 7—ray transition energy is the absolute 7—ray intensity. Apparent log f t values were deduced from the apparent 6 branches, the 54Sc half-life, decay Q value, and the excitation energy of the populated states [79]. 93 $2000 * (e) 1020-keV gate 1495[14] 516003 $1200 5 800 4 E 1002[48] Q 5: \ £1 4; 1020[7] 315 1965[13] 5 33 g - 1020[12] \ 8 2; 010 / 111 1111 “5 o | Ill L [11111111 0 E g: (C) 840-keV gate 1495[10] 2 33g) 1505-keV gat Q 5: Q 2; a 4: 1002[7] a» : 1495[4] ‘ : G » 1 E3, 8 » U 2: i U 1 1 o, Mull“ JJ Illlllll.“ 9 lll 3205(6) 1002-keV gate 1495150] % 6:01) 1965-keV gate 1495113] N16; :1 5: \ ; \ i":12? :3 4, g 3; 8 3“ U 4: 340m] U 2— 1 0 1.11.11 ‘1, lJLlrlllLJlJlflLhmHm o 400 800 1200 1600 2000 400 800 1200 1600 2000 Energy (keV) Energy (keV) Figure 4.11: (a) 'I‘wo-dimensional 'y'y coincidence matrix for 'y rays observed in co- incidence with 54Sc~correlated decay events over the entire surface of the DSSD. 7 coincidence spectra gated on the (b) 485 keV, (c) 840 keV, (d) 1002 keV, (e) 1021 keV, (f) 1495 keV, (g) 1504 keV and (h) 1965-keV transitions in coincidence with a 54Sc-correlated fi—decay event. The number in square brackets following the energy is the number of counts in the peak. 94 adopted from previous works [31,81]. Apparent allowed feeding to both the 2+ and 4+ states suggest a J 7' assignment of 3+ for the 54Sc ground state, in agreement with the assignment made based on the fi decay of 540a into 54Sc, discussed in Section 4.1.2. No direct feeding is therefore expected to the 0+ 54Ti ground state. The missing fl intensity can be accounted for by a fin branch of 16:1:9% in the decay. An allowed ,8 decay from the (3)+ 54Sc ground state limits the spin of the newly-placed states with energies 3000, 3337 and 3460 keV to J 7' 2 2+, 3+ or 4+. The '7-ray spectrum collected within a 20-ps window following a 54Sc implantation event is shown in Fig. 4.12. Observation of a 110-keV isomeric transition in 54Sc was first reported by Grzywacz et al. [68], where the 110-keV transition was assigned E2 multipolarity, based on Weisskopf estimates. As shown in Fig. 4.12, the 110—keV isomeric transition was observed with high statistics and an improved half-life for this isomeric state was deduced. The prompt y-ray energy is shown as a function of time after implantation of 54Sc fragments into the DSSD in Fig. 4.13(a). The 110—keV transition is visible, with an intensity decaying with time. Fig. 4.13(b) is a projection of the data in Fig. 4.13(a) onto the time axis, showing the time dependency of the 110—keV transition. This decay curve was fitted with a single exponential function with a half-life for this isomeric state of 2.77:1:002 ,us. The present half-life value agrees with measurement of Grzywacz et al. [68], but significantly improves the precision. The improved precision of the half-life for the 110-keV isomeric transition in 54Sc allowed confirmation, through comparison with Weisskopf estimates, of the E2 mu]- tipolarity assignment for this transition. The (3)+ spin and parity assignment for the 54Sc ground state, and an E2 isomeric transition limits J7r of the 110-keV state to J = 1, 5. A spin of J =1 is excluded by the non-observation of feeding to this state by the fl decay of 54Ca. Thus, under the assumption of an E2 multipolarity for the transition, the 110—keV isomeric state in 54’Sc is tentatively assigned a spin and parity of J'”=(5)+, as was originally suggested by Grzywacz et al. [68]. However, this assign- ment assumes a single-particle nature for the final and initial states. The situation 95 20000 1 10 16000 Counts/keV 1 2000 ll TllllllilTIlllllllIlll 8000 4000 llllhlllllllilll‘ C.) mI I I I I I I . I I A . I 4 l—L—L—L—l—n 50 100 150 200 250 300 Energy (keV) Figure 4.12: 7-ray spectrum in the range of 0—300 keV collected within the first 20 us following a 54Sc implantation event. The 7—ray transition is marked by its energy in keV. 200.82 $1805 100 a 16 ' 0 314 33 12 60 LS 100 >- 30 40 6 4 7 20 20 0 I I I I I I I I 0 a l (b) T172 (54Sc-110kev v) :33 {- =2.77:0.0211s g t .. 1 “1000 E U WWW; 123456789 Time(us) Figure 4.13: Time dependence of the 110-keV isomeric transition in 5480. (a) The prompt 'y-ray energy plotted as a function of time following a 54Sc implantation. (b) A projection of (a) onto the time axis. The data were fitted with a single exponential decay to deduce a half-life for the isomeric state in 54Sc. 96 is less certain when more complicated wavefunctions involving configuration mixing are considered. As will be discussed further in Section 5.1.2, a 4+ spin and parity assignment for the 110—keV isomeric state may also be possible. 4.1.4 ,8 decay of 568C The analysis of the 8 decay of 5680 only required a correlation time of 1 s, and the full DSSD implantation detector could be used in the analysis. The 8—delayed 'y-ray spectrum for 56Sc in the range 0 to 2 MeV is shown in Fig. 4.14. Nine transitions in this spectrum have been assigned to the decay of 56Sc, and are summarized in Table 4.2. Two additional transitions were assigned to the decay of the granddaughter nuclide, 56V (T1 )2 = 216 :l: 4 ms [70]) - no transitions are known in the decay of the daughter 56Ti (T1)2 = 200 : 5 ms [71]). The transitions with energies 592, 690, 751, 1129 and 1161 keV agree well with '7 rays previously observed in both ,8 decay [31] and deep inelastic in-beam experiments [81]. In Ref. [31], it was noted that the transition at 592.3:t0.5 keV was within the error of a known transition in 55Ti [83], and the possibility of 8—delayed neutron emission was suggested. In addition to the 592—keV transition, another known transition in 55Ti, at 1204 keV, is present in the 8-delayed y—ray spectrum, confirming ,8-delayed neutron emission in the decay of 5680. Liddick et al. [31] proposed two 8-decaying states in 56Sc: a low-spin state with T1/2 = 35:1:5 ms, and a higher-spin state with slightly longer T1/2 = 60:1: 7. Given the different spin values, the two ,8-decaying states were observed to populate different levels in the 56Ti daughter. 'y-gated decay curves were generated for each ’y—ray tran- sition identified in Fig. 4.14 and fitted with a single exponential decay plus constant _ background. The y-gated decay curves are presented in Figure 4.15, and the resultant half-life values are summarized in Table 4.2. The values deduced from decay curves gated on the 690 and 1161-keV transitions were 73d:10 ms and 78:1:9 ms respectively, consistent within 10 of the previous half-life determination for the higher-spin isomer, 602t7 ms [31]. The transitions assigned to the 8n decay also have half-lives consistent 97 3602- en:— 1129 840:: 690 1161 U E 59%: K 30: 751 _ 1467 20; 1/204]1495 10: (17'12 11.61611.1111..1.31‘.. ‘. . 800 1200 1600 2000 Energy (keV) c? Figure 4.14: 8-delayed 'y-ray spectrum in the range of 0-2 MeV following the decay of 56Sc to states in 56Ti. Transitions marked by their energies in keV are those assigned to the decay of 56Sc. Transitions marked by a shaded square are attributed to the 8 decay of the grand-daughter nuclide, 56V. Table 4.2: Energies and relative intensities of 8-delayed 'y rays observed following a 56Sc—correlated decay event and assigned to the decay of 56Sc. Observed half-lives for 'y-gated decay curves are also included. E7 13””th Decay Initial State Final State T1/2 (keV) (%) Mode (keV) (keV) (ms) 591.7 i 0.3 14 i 2 8n 591.7 0 78 :1: 25 689.6 :1: 0.3 18 :1: 2 8 2979 2289 73 :1: 10 750.9 :1: 0.4 8 :1: 2 8 1880 1128.7 24 :1: 7 1128.7 i 0.3 48 :1: 4 8 1128.7 0 51 :t 6 1160.6 :1: 0.3 30 :1: 3 8 2289 1128.7 78 i 9 1203.5 :1: 0.3 8 :1: 1 8n 1795 591.7 68 :1: 19 1466.8 i 0.3 6 :1: 1 8 - - 60 :1: 13 1494.8 :1: 0.3 3 :1: 1 8 4474 2979 150 :t 44 1711.6 :1: 0.3 3 :1: 1 8 - - 29 :1: 10 98 with decay from the higher—spin state, as do the transitions with energies 1467 and 1495 keV. The half-life of the lower—spin 8-decaying state in 56Sc was previously deduced to be 35 :1: 5 ms from a two-component fit to the decay curve gated on the 1129- keV 7 transition, which dep0pu1ates a state fed directly and /or indirectly by both 8—decaying states in 56Sc. 'I\1vo newly-identified 7—ray transitions with energies 751 and 1712 keV were found to have half—lives similar to that extracted previously for the lower-spin isomer. These two 7 rays apparently depopulate states fed exclusively by the lower-spin isomer. The weighted average of the half-lives for the 751- and 1712-keV transitions yields a half-life for the lower-spin isomer of 26 :1: 6 ms. Levels in 56Ti and 55Ti populated in the decay of the two 8—decaying states in 5680 are shown in Fig. 4.16. The states in 55Ti are known from the 55Sc 8 decay in NSCL experiment 07509 [84], previous 8-decay studies [31], and deep inelastic work [83]. The three states at 1129, 2290 and 2980 keV in 56Sc were previously identified and tentatively assigned spin and parities [31,81], which are shown in Fig. 4.16. 77 coincidences [see Fig. 4.17(a),(c),(d)] also confirm the placement of the 690-, 1129— and 1160—keV 7 transitions associated with these three levels. The observed 77 coincidences between the 1129— and 751-keV transitions [see Fig. 4.17(b),(c)] suggest the placement of an additional state populated by the lower-spin 8—decaying state at 1880 keV. 77 coincidences between the 1495— and 690-keV transitions [see Fig. 4.17(a),(e)] suggest that the higher-spin 8—decaying state populates a level at 4474 keV, in addition to populating the two states at 2290 and 2980 keV. Only the 1467- and 1712-keV 7 transitions remain unplaced in the present level scheme for 56Ti. Absolute 7-ray intensities were determined for the 8—delayed 7-ray transitions in the 5680 decay by comparison of the number of observed 7 rays, adjusted for the abso- lute efficiency of SeGA, with the number of observed 56Sc decay events. Typically, the number of parent decays is obtained from the decay curve fit, as previously described in Section 3.5.1. However, a precise value for 56Sc could not be obtained using this 99 a — (a) T1/2(Bn-y)=78:25 ms £10 (1) Tl/2(Bn-y)=68:19 ms 3 10. 592 keV g 1204 keV >1 : g ‘a’ 2 a a [L U 1 _______ 1 U : a (b)T1/2(B-v)= 73:10 ms a } (g)T1 /2(B-y) 60:13 ms 1‘3 690 keV n10.1467kev E 10: ] Q 2 s ] . a _ 8 ' ] 0 1 I. q u g a (c) T172191) 24:7ms a... (h)T172(B-v)=150t44ms 3 10 751keV 8 : 1495 keV \ 7 v—I :3 . U (”100... . 1..Ln.11 .L.1. 1.1.l1. .1..1 m ’ . 1 ”4.1...“ 1...._L..._._.. E 8 (1d) T1/2([3- Y)— "' 51:1:6 ms 810: (1)T1/21(‘B'Y)= 29:1:10 ms 3 : 1129 keV g} 17121tev ,7, 10. y, 1: E a :3 . [L :1 O O U l;'-1\L U "’ 7T" ' ‘ ” 800 1000 f; - (c) T1/2(B-y) = 81:9 ms Tune (ms) Q 10E 1160 keV 3 I § . O U 1: c" 206‘ 400 606‘” 800 ”1000 Time (ms) Figure 4.15: Decay curves for 56Se using a correlation time of 1 s, and data over the entire surface of the DSSD, with the additional requirement of a coincident decay 7- ray transition. Decay curves are gated on the (a) 592-keV, (b) 690—keV, (c) 751-keV, (d) 1129-keV, (e) 1160-keV, (f) 1204—keV, (g) 1467-keV, (h) 1495-keV and (i) 1712-keV 7 rays in the decay. All curves were fitted with a single exponential decay (dotted line) and a constant background component (dot-dashed line). 100 .89 >62 game 9 ate. a .888 85838 Essen 9: eee Se: 38?: a :88 as ea Q@ .396 “:8on 23 mm H: a cavemen ea 38 2X: 43 Became .opflm mafiaomvin Eamémawfi 2: m0 35% 2E. .Esonx you mm mopgm mammaompfi 93 23 mo 8%“me 9,338 23. .moammofim .878 830er 2% mm 8%“QO 43.7% wafiovfi on» $23028 5955: 23. 823983 e 9538 2: v.8 8&8qu 48.1. otoaofl 2: @5828 383.83 5 £883 6:8 .095 we 430% n 23 386:8 vmuflsaoa Fonda 3 £32 98 53 mt. em cosine: otofiofl 9: mo havoc b 23 mamaoz£ amen 8% 683% $32 wmmoaem 6:6 mama Fem PE 1 O _ +O O +O NUTWM W 2a: 3.9: +6 5%: 3.9 78 32 ence. an 2.6% e. New 1 .o 43.00 92.. +8 $2 4» .038 g _ film 660 N12 any: 4.x. Onox \& 4x «Ham :8 o 1327/ . as «a: N en a . /H/\xnc /&0x 06/ \A/ \AAc 4X .0X ./nc \x& 6 101 DJ §(a) 690-keV gal—“7 1160[7] 112915] 149512] 111111 11111. (b) 751-keV gate 1-1111 690[5] (c) 1129-keV gate 7/51[2] fi60[3] 111111 1| 1. [1| Counts/2 keV YYTH'YITY YYITNITIT] m:1171 112913] Counts/ 2 keV ‘7’ (HO I-I Counts/2 keV ."f ...‘." A I IIIIIY TIT IYY E _. 690[9] (d)11601kev gate ”3i 112913] \ : 82; K :3 _ O : ("1? 3 ; (e)1495-keV gate a 690[2] a .- | *a :1 O U 00 800““‘1‘2'00d‘i600d‘2000 Energy (keV) b . J; Figure 4.17: 77 coincidence spectra gated on the (a) 690-keV, (b) 751-keV, (c) 1129— keV, (d) 1160-keV, and (e) 1495-keV transitions in coincidence with 56Sc-correlated 8-decay events. These coincident spectra include data within a 1 s correlation time, for implants collected over the entire surface of the DSSD. The numbers in square brackets following the energies are the number of counts in the peaks. 102 approach. The presence of the two 8-decaying states in this case results in a decay curve fit with a large number of free parameters. Instead, the number of 5680 decays was determined using the total number of implantations, taken from the particle iden- tification, and the average 8-detection efficiency of 11.4:t0.4%. The absolute 7-ray intensities are included in Table 4.2, and in the decay scheme of Fig. 4.16. Apparent 8 branches were deduced from the absolute 7—ray intensities. Decay from the higher- spin isomer apparently p0pulates both the 4+ and 6+ states in 56Ti. This suggests that the higher-spin isomer has J 7r25'", in contradiction to the previous assignment of (6,7)+ by Liddick et al. [31]. However, the large Q5 value for the decay allows the (4)+ state at 2289 keV to be populated by cascades from above, resulting in the observed intensity difference between the 690- and 1161—keV transitions. Given this possibility, the higher-spin 8—decaying state has been tentatively assigned as (5,6)+. The state with an energy of 4474 keV, also populated by the higher-spin 8—decaying state, has J7r limited to 4+, 5+, 6+ or 7+, depending on the spin and parity of the higher-spin 8—decaying state in 56Sc. Decay from the lower-spin 8—decaying state in 568C apparently directly feeds the first excited 2+ state in 56Ti, which limits the spin and parity of this 8—decaying state to 1+, 2+ or 3+. The J 7' for the 56Sc ground state can be further restricted if there is direct 8 feeding to the 56Ti ground state. The absolute intensity of the 592-keV tran- sition in the 8n daughter 55Ti suggests a lower-limit for the neutron branching ratio of 14i2%. Even under the assumption that the two unplaced 7 transitions directly populate the 56Ti ground state, there is 29i7% of the 8 intensity unaccounted for. Two scenarios can account for the missing intensity: either direct feeding to the 56Ti ground state or 8n decay populating the ground state of 55Ti directly. Each possibility was investigated by reanalyzing the 8 decay of 5680 with the longer correlation time of 5 s. With a longer correlation time, it was possible to compare the intensities of 7—ray transitions in the decay of the 8n daughter 55Ti (672.5 keV, Igbsome = 44:1:4%), and the 88 granddaughter 56V (668.4 keV, 181’”th = 26 :1: 2%). Within a 5 s correlation 103 Table 4.3: Determination of the 8 and 8n contributions to the decay of 5680, based on the observation of 7-ray transitions in subsequent decays within the 8 and 8n decay chains. 5680 Decay Signature E7 Igbsomte Counts (5 8 Number of ‘70 of 56Sc Decay Decay (keV) (%) Correlation Observed Decays Mode Time Decays 5 56V4560. 6684:03 26:2 94:12 362:54 71:14 8n 55Ti—>55V 672.5:0.4 44:4 64:10 145:26 29:6 time, >90% of the 55Ti (T1 )2 = 1.3 :t 0.1 s [71]) populated in 8n decays would have decayed, while 56Ti (T1 )2 = 200 :1: 5 ms [71]) populated in the 8 decay of 568C would have decayed into 56V (T 1 )2 = 216:1:4 ms [70]), which would also have decayed. Thus, the ratio of 8 to 8n decays in 568C can be determined by the ratio of the intensities of the 668- and 673-keV transitions, which are detected with nearly the same efficiency in SeGA. The results of the 8n/ 88 analysis are summarized in Table 4.3. Approximately 70% of the decays of 568C populate states in 56Ti. Thus, the missing 8—intensity cannot be fully accounted for by 8n decay to the ground state of 55Ti, and there is apparent direct population of the 0+ 56Ti ground state. The higher-spin (5,6)+ 8- decaying state cannot populate the 56Ti ground state directly, and so the lower-spin isomer must decay directly to the 56Ti ground state. The apparent feeding of the 0+ 56Ti ground state and the first 2+ excited state thus suggests a spin and parity for the lower—spin 8-decaying state of J“=1+. Direct feeding from the (1)+ 8-decaying state limits J7r of the state at 1880 keV to (0,1,2)+. 4.1.5 Isomerism in 568C The 7-ray spectrum collected within a 20—118 time window following the implantation of 56 Sc is presented in Fig. 4.18. Five transitions were observed, and are summarized in Table 4.4. Those transitions with energies 140, 188, and 587 keV are in agreement with the three transitions previously reported by Liddick et al. [31]. 104 @140 \120— Irlj k Counts 40 20 IIIIfiTIII III III III 48 140 188 11 587 727 ”1'00 200300400 500 600 700 800 Energy (keV) Figure 4.18: The prompt 7-ray spectrum collected within the first 20 [is following a 5680 implantation. The 7 rays assigned to the isomer decay in 56Sc are marked by their energies in keV. The additional peak visible in the spectrum at 110 keV is contamination from the isomer decay of 54So. Table 4.4: Energies and relative intensities of isomeric 7 rays observed in the 20 ,us window following a 568C implantation event. Relative 7 intensities are corrected for internal conversion [61], based on the transition multipolarity and energy. Transition multipolarities were assigned based on Weisskopf half-life estimates. E7 [relative Transition Total Internal Initial Final (keV) (%) Multipolarity Conversion State State Coefficient (atotal) (keV) (keV) 47.7 : 0.3 70 : 19 M1 (1.16:0.02)x10~7 775 727 140.5 : 0.3 61 : 7 M1 (66:01) x 10-3 727 587 187.8 : 0.3 61 : 8 E2 (2.63:0.04)x10'-2 775 587 587.2 : 0.3 100 : 12 M1 (2.34:0.04)x10-4 587 0 727.1 : 0.3 32 : 5 E2 (2.45:0.04)x10-4 727 0 105 keV) 00 O O :71! A a: V ( O\ O O ‘IIHII l 1 l Jlfl'lllllllllLlll'lllllllllllllllllllll'l ‘ 11.5‘L‘211i215‘”3.0 Time (118) E (b) T1/2 (56Sc-isomeric y) : + 3 10E 568C 290 _ 30 ns 23 Z S : bkgd O U 1:— . i no” u on 0 00 «no 1;.'_ :.]“l—L‘- fl . I . I . . 1 . I 2 3 4 5 . 6 Tlme (118) Figure 4.19: Time evolution of the prompt 7-ray transitions in 56So. (a) Prompt 7-ray energy plotted as a function of the time elapsed following a 563C implantation event. (b) Projection of (a) onto the time axis, gated on the five prompt 7—ray transitions. The data were fitted with a single exponential decay and a constant background. The observed 7-ray energ is presented as a function of the time elapsed between a 5630 implantation and prompt 7-ray emission in Fig. 4.19(a). The five transitions with energies 48, 140, 188, 587 and 727 keV appear to decay with the same half-life. Figure 4.19(b) contains a projection of the data of Fig. 4.19(a), gated on the five isomeric transitions, onto the time axis. The resulting decay curve was fitted with a single exponential decay and a constant background with a half-life for the isomeric state in 568C of 290i30 us. The low-energy structure of 56Sc populated by isomeric decay, presented in Fig. 4.16, was established based on the observed 77 coincidences (see Fig. 4.20), and relative 106 3 8 (a) 48-keV gate %10§:(d)587-keV gate 140 22 i4 6 1401211 / <4 s;— 48[15] W20“ 1 +03 587[14] :3 6: S4 5 o o 45— U 72713] 2 \ U 2;— O ‘ ' ‘ JM'HM'” O0 100 200 300 400 $10 48 [20] (b) 140-keV gate Energy (keV) fi 3 587 [19] > 33 6 <1) f(e) 727-keV gate G M . :3 4 \ r 48121 o 59 :/ U 2 c: 1: 3E :3 ~ 'I I .1 11.11.11. 0 ’— g 2 (c)188-keV gate U E M 587 [17] '. ...... E 4 C0 20 400 60 800 g Energy (keV) 0 2 U 0 111.1. 0 200 400 600 800 Energy (keV) Figure 4.20: 77 coincidence spectra for the (a) 48-keV, (b) 140-keV, (c) 188-keV, (d) 587-keV and (e) 727-keV transitions in the 20—113 time window following a 5680 implantation. The number of coincident counts are given by the numbers in square brackets following the transition energies. transition intensities corrected for internal conversion. Energy-sum relationships to the cross-over transitions with energies 188 and 727 keV provided further support for the proposed level scheme. A novel approach was taken to determine which of the two 8—decaying states in 56Sc is populated by the isomeric decay. A half-life curve was constructed in- cluding only 8—decay events correlated with 56Sc implantations that were in turn coincident with one of the five prompt 7 rays. The resulting decay curve is pre- sented in Fig. 4.21. A fit to these data with a single exponential decay for the 56Sc parent, growth and decay of the daughter (56Ti, T1/2 = 200 :1: 5 ms [71]) and grand- 107 «0100:- E : T (isomer 56SC 1m lant-B) 8 1/2 3101’s 2 - 1 I 1 e 1 :1 10 . 8 1":‘l‘\ ] ] l ] ‘\ I». 1 ‘1 J '1 X I l, / l ——11— . .- 1 I i—f—‘—‘. . 1 . I I II-I I H I rrnilT—.L_111HH] ‘ n U' "n 1 4 .. + e _ 6..." 568cm 1 ”9%“ 56Ti1:\ .111111Hn1.l1.i.....l....l.:. l ....... L1... 200 400 600 800 1000 OS- Figure 4.21: Decay curve for 56Sc-correlated 8 decays considering only 56Sc implan- tations coincident with identified isomeric 7—ray transitions. The decay curve was fitted with a single exponential decay, growth and decay of the 8-decay daughter and granddaughter, and a constant background. daughter(56V, T1/2 = 216 .1: 4 ms [70]), and constant background yielded a half-life of 30:1:5 ms, consistent with decay of the lower-spin 8-decaying state, previously assigned J“: (1)+ (see Section 4. 1. 4). Knowledge of the quantum numbers of the lowest-energy state of the isomer ‘band’ permits tentative J 7' assignments for the remaining low-lying levels populated in the isomeric decay of the isomeric state in 5680. The 188-keV transition, which directly depopulates the isomeric level, has E2 multipolarity, based on Weisskopf estimates (see Table 4.5). The 48—keV transition is assigned multipolarity M 1, on the basis of the observed favourable competition with the 188-keV transition, and comparison with Weisskopf estimates. It follows that the 140-keV transition must also have M1 multipolarity. The observed intensities of the 727-keV and 140-keV transitions depopulating the isomeric state then suggest the former transition to be E2, which according to Weisskopf estimates, will proceed at a comparable rate to the 140—keV M1 transition. Finally, the remaining 587-keV transi- tion must have M1 multipolarity. These assignments lead to the tentatively assigned 7' = (2)+, (3)+ and (4)+ values for the states in 5680 at 587, 727 and 775 keV excitation energy above the base state, assuming that the base state has J 7’ —.(1)+ 108 Table 4.5: Weisskopf half-life estimates for the five prompt 7-ray transitions observed in 5630. E7 Weisskopf single particle half-lives (keV E1 M1 E2 M2 E3 M3 47.7 4.7 ps 114 ps 209 ,us 5.9 ms 4.0 hours 1.5x106 years 140.5 184 is 4.5 ps 942 ns 27 118 7.6 s 92 years 187.8 77 fs 1.9 ps 221 ns 6.3 ,us 990 ms 6.7 years 587.2 2.5 fs 61 fs 738 ps 21 ns 339 118 2.1 hours 727.1 1.3 fs 32.2 fs 254 ps 7.2 ns 76 us 1084 s 4.2 Isomeric Structure of 5OK The neutron-rich 19K isotopes with N Z 20 are situated on the border of two major shells, with protons occupying the high-lying sd—shell single-particle orbitals, and neutrons filling low-lying fp—shell orbitals. The low-energy levels in the K isotopes are thus sensitive to the relative spacing of the proton single-particle orbitals below the Z =20 shell closure. Recent evidence [85,86] suggests an evolution of the energies of the 7rld3/2 and 71281/2 states for the K isotopes with N >28 outside of the predictions of the most current sd- f p cross-shell effective interactions. The low-energy levels in §8K31 are thus interesting to better understand the evolution of the proton single— particle states below Z =20 with increasing neutron excess. Isomerism in 50K was first reported by Lewitowicz et al. [87], with more details included in the work of Daugas [88]. The prompt 7-ray spectrum for 50K implanta- tions collected in this work is presented in Fig. 4.22(a). Three transitions are apparent in this spectrum, at 43, 128 and 172 keV, and their details are summarized in Ta— ble 4.6. The observed transitions are in agreement with those previously observed by Daugas [88]. However, the additional two transitions with energies 70:1:1 keV and 101:1:1 keV also reported in Ref [88] are not apparent in the 7—ray spectrum depicted in Fig. 4.22(a). Fragment-77 coincidences, shown in Fig. 4.23, indicate a coincidence relationship between the 43— and 128-keV transitions, however, the ordering of these transitions is not uniquely determined. The third transition at 172 keV in the spec- 109 Table 4.6: Energies, relative intensities and half-lives for the 7 rays observed within the 20 us following a 50K implantation event. Relative intensities are corrected for internal conversion [61], based on the transition multipolarity and energy. Transition multipolarities were assigned based on comparison to Weisskopf half-life estimates as discussed in the text. Half-lives for individual 7-ray transitions were determined using the MLH method. A weighted average of the MLH values is also presented, as well as the results of a fit to the data for all three transitions using the ROOT fitting program. E7 [felative Transition Total Internal Number of MLH (keV) (‘70) Multipolarity Conversion Counts for Half-life Coefficient (atom) T1/2 (ns) 42.8:0.3 84:22 M1 0107:0003 4 65733 128.4:03 75:15 M1 00059:00002 21 135% 172.2:03 100:19 E2 00287:00008 11 2293,35 MLH Weighted average: 154%: ns ROOT fit: 158i27 ns trum in Fig. 4.22(a) is placed as the cross-over transition, based on the energy-sum relationship. The time evolution of the 50K 7 rays is given in Figure 4.22(b). The three tran- sitions observed in Fig. 4.22(a) are apparent, though statistics are low. The number of counts are further reduced when considering only 7 rays with decay times greater than 1 ps, after the background of prompt :1:—rays which occurs upon fragment im- plantation into the DSSD. The 7-ray spectrum with this cut on the decay time is shown in Fig. 422(0), and the low statistics are evident. The total number of counts for each of the three transitions is summarized in Table 4.6. An MLH analysis was used to extract the isomer half-life information due to the low statistics for the observed 7 rays. The likelihood function (see Section 3.5) was evaluated for each event, and maximized to yield a half-life value for each of the three 7 rays assigned to the decay of the 5OK isomeric state. The resultant half-lives are listed in Table 4.6. The half-lives are consistent with one another, and a weighted 4+3 average yields a value of 15 _ ‘3 ns for the isomeric state with energy 172 keV in 50K. 110 > : 33100: (a) 172 N I 128 \ g 80_— : 43 8 60— U : A O 1 20 lllllLllllllllllllllllllilll1 1 1 1 1 1 0 50 100 150 200 250 300 350 400 450 500 Energy (keV) 9250-5 . . . - ' g; (b) 200:; LII t... 1111-1 9.5.1. 10011-1135" ’ - 50 —" ' Hf ' 1000 TAC Time (us) > Z 128 $4) 105 (C) E : ‘5 83' 172 O .. C) 6:- 4} 2} 43 I co {)1 O 100 150 200 250 Energy (keV) Figure 4.22: (a) 7-ray spectrum observed within the 20 ,us window following a 50K implantation event, in the energy range of 0—500 keV. The transitions assigned to the isomeric decay of 50K are marked by their energies in keV. (b) Time evolution of the 7 rays observed following 50K implantations. The three observed transitions are marked by arrows. (c) 7-ray energy spectrum of (a) with the requirement of a TAC time greater than 1.0 ps, beyond the x—ray flash. The three counts with energy 199 keV in this spectrum are background. 111 :5 6» (a) 43-keV gate 5 E 4 .128 [14] 5 3 c U 2 ] 1 7 > 0 3‘9 6 43 [12] (b) 128-keV gate N 5 E 4 S 3 0 U 2 If Not (c) 172-keV gate ‘ > U #4 N \ 83 gl 0 . U , 1i ‘ ‘1; 00 100200 300 400 500 Energy (keV) Figure 4.23: 77 coincidence spectra gated on the (a) 43—keV, (b) 128-keV and (c) 172- keV 7 rays observed within 20 us following a 50K implantation event. The numbers in square brackets following the energy of a peak are the number of counts in that peak. 112 Table 4.7: Weisskopf half-life estimates for the three prompt 7-ray transitions observed - 50 in K. E7 Weisskopf single particle half-lives (keV E1 M1 E2 M2 E3 M3 42.8 6.5 ps 158 ps 359 as 10 ms 8.6 hours 4.1x106 years 128.4 241 fs 5.9 ps 1.5 11s 42 as 14.2 s 206 years 172.2 100 fs 2.4 ps 340 ns 9.6 as 1.8 s 14.7 years The results of a decay curve fit, using the ROOT program with Poisson analysis, to the summed data for all three transitions yields a value of 158:1:27 ns, consistent with the average MLH result. Both half-life values agree with the value of 125:1:40 ns reported by Daugas [88]. Comparison of the deduced lifetime with the Weisskopf estimates (see Table 4.7) suggests an E2 multipolarity for the 172-keV transition. The competition of the 43-128 keV 7-ray cascade with the direct ground state transition from the isomeric 172-keV state suggests that the multipolarities of the 43— and 128-keV transitions are M1. The low-energy level scheme for 50K is presented in Fig. 4.24. Spins and parities for the low-energy states populated in the prompt 7-ray decay of 50K are discussed in detail in Chapter 5. 4.3 Structure of 61Mn from 8 decay of 61Cr Below the Z :28 shell closure and approaching N :40, there is evidence in both the 24Cr and 26Fe isotopic chains of increasing collectivity, which develops below Z =28 with the approach of the 1g9/2 neutron orbital to the Fermi surface. While the onset of collectivity has been investigated in the 24Cr and 26Fe isotopic chains, data in the intermediate 25Mn isotopes is less abundant. The level structure of 61Mn will extend the information in the Mn isotopic chain, and provide insight into the development of collective behavior for Z =25. Implanted 61Cr fragments were correlated with their subsequent 8 decays by re- 113 69’ x, \‘ <11” «999 (6,9679 x\s\ (3)- «f\‘ 611- 3’1 ’1718 tax/0' (2)- 0“" 128 4 (D- 1 11 0 Figure 4.24: Proposed low-energy level scheme for 50K. The numbers in square brack- ets following the 7-ray transition energies are the relative transition intensities. The order of the two 7-ray cascade is not uniquely determined — the tentative 2‘ state may alternately be located at an excitation energy of 42.8 keV. quiring the presence of a high-energy implantation event in a single pixel of the DSSD, followed by a low-energy 8 event in the same or any of the eight neighbouring pixels, using a correlation time of 3 s. The decay curve for 61Cr-correlated 8 decays given in Fig. 4.25 was fitted with a single exponential decay combined with an exponen- tial growth and decay of the short-lived daughter, 61Mn, whose half-life was taken to be 670i40 ms from Ref. [89]. A constant background was also included as a free parameter in the fit. A half-life of 233:1:11 ms was deduced for the ground-state 8 decay of 61Cr. This new value compares favorably with the previous measurement by Sorlin et al. [59] of 251i22 ms, but is more than 10 shorter than the 270:1:20 ms value reported earlier by Ameil et al. [90]. The 8—delayed 7—ray spectrum from the decay of 61Cr is shown in Fig. 4.26. The spectrum covers the energy range 0 to 2.5 MeV. The seven transitions assigned to the decay of 61Cr are listed in Table 4.8. The peaks observed with energies of 207 and 629 keV in Fig. 4.26 are known transitions in the decay of the 61 Mn daughter [91]. The 1028- and 1205-keV transitions from the decay of the grand-daughter 61Fe [92] are 114 Counts/ 30 ms 8 10 ...1... .I‘..1....1....1...‘. 0 500 1000 1500 2000 2500 3000 Time(ms) Figure 4.25: Decay curve for the 8 decay of 61Or to 61Mn, constructed from decays correlated with 61Cr implantations using a 3 s correlation time. The data were fitted with a single exponential decay for the parent, growth and decay of the daughter, and a constant background. also seen in Fig. 4.26. The peaks at 355, 535, 1142 and 1861 keV correspond in energy to the transitions previously reported by Sorlin et al. [59]. However, the transition observed in that work at 1134 keV was not apparent in the present work, see Fig. 4.26. The three additional 7-ray transitions at 157, 1497, and 2378 keV can be assigned for the first time to the decay of 61Cr. It is worth noting that two transitions with energies 155 and 355 keV had previously been assigned to the 8 decay of 62Cr [58]. Observation of these transitions in Fig. 4.26 suggests that they are associated with levels in 61Mn, and their observation in Ref. [58] may be evidence for 8—delayed neutron decay of the 62Cr ground state. The in-beam 7-ray spectra for 61Mn and 62Mn obtained by Valiente—Dobon et al. [57] include transitions with energies of 157 and 155 keV, respectively, supporting the present assignment of the 157-keV 7-ray transition in Fig. 4.26 to the decay of 61Cr. The proposed decay scheme for levels in 61M11 populated following the 8 decay of 61Cr is presented in Fig. 4.27. The 8-decay Q value was taken from Ref. [80]. Absolute 7-ray intensities were deduced from the number of observed 61Cr 7 rays, the 7-ray peak efficiency, and the number of 61Cr implants correlated with 8 decays, as derived from the fit of the decay curve in Fig. 4.25. The two-7 cascade involving 115 8 _ [[IIIIIITIIIIIIIFF \ Counts / keV 3 535 500 I“ l , If 1 . ‘. ] ‘ ‘ ‘ Idllldl 1‘! I thin 11 l 1000 1142 1861 1497 2378 1500 , ['2000 2500 Energy (keV) Figure 4.26: 8—delayed 7-ray spectrum for the 8 decay of 61Cr, in the emery range of 0-2.5 MeV. Transitions marked by their energies in keV are assigned to the decay of 61Cr. Transitions marked by a filled circle are known transitions in the daughter 61Mn, while transitions marked by a square belong to the decay of the 61Fe granddaughter. Table 4.8: Energies and absolute intensities of the 8-delayed 7 rays assigned to the decay of 61Cr. The initial and final states for those transitions placed in the proposed 61Mn level scheme are also indicated. E7 I absolute Einitial Efinal Coincident (keV) (‘70) keV keV 7 rays (keV) 157.2:l:0.5 9i2 157 0 354.8:1:0.4 1621:? 1497 1142 535, 1142 534.6:1:0.5 5:121 2032 1497 355 1142.2:1:0.4 21:1:2 1142 0 355 l497.3:l:0.5 9:1:2 1497 0 1860.8:1:0.4 20i2 1861 0 2378.2:l:0.4 11:1:1 2378 0 116 the 1142- and 355-keV transitions was confirmed by 77 coincidence relationships (see Fig. 4.28). However, the ordering of the two transitions is not uniquely determined. The arrangement shown in Fig. 4.27 was based on the absolute intensities of the 1142- and 355-keV transitions. No evidence was found for 7 rays in coincidence with the 157-keV ground-state transition within the statistical uncertainty of the measurement. Two counts in the 355-keV coincidence spectrum of Fig. 4.28(b) suggested placement of the 535-keV line in cascade from a higher—lying level at 2032 keV. This tentative placement is represented in Fig. 4.27 by a dashed line. The apparent 8-decay feeding to levels in 61Mn was deduced from the absolute 7-ray intensities. The deduced values, along with apparent log f t values, are also given in Fig. 4.27. The observed 8 branches all have apparent log f t values between 4 and 6, consistent with allowed transitions. The 1142-keV level was found to have a small 8 branching with a large error, and might not be directly populated by the 8 decay of 61Cr. A ground-state spin and parity assignment of 5/2‘ was assigned to 61Mn by Runte et al. [91], based on the systematic trends of ground-state J7r values for the less neutron-rich, odd-A Mn isotopes. The first excited state at 157 keV has been tentatively assigned J ”=7/ 2' from the in—beam 7-ray results. Allowed 8 feeding to the lowest two states would then limit the J7r quantum numbers of the 61Cr ground state to values of 5 / 2“ or 7/2‘. Gaudefroy et al. [58] tentatively assigned the ground state of 61Cr to be J 7r=5/ 2“ on the basis of the observed 8-decay properties of the progenitor, 61V, and such an assignment is adopted in Fig. 4.27. It is likely that the 1142—keV state has low spin, either J"=1/2‘ or J"=3/2'. The level at 1142 keV is apparently weakly fed by 8 decay, was not identified in the yrast structure of 61Mn [57], and does not depopulate to the 7/2- level at 157 keV within the statistical uncertainty of this measurement. The 1497-keV level is tentatively assigned a J7r of 3/2“ or 5/2- because of the apparent allowed 8-decay branch and the favorable competition between the two depopulating 7 rays that suggests that both transitions 117 (5/2—) 0 T1/2=233i11 ms 61 Cl“ QB=9.4:04 MeV g ,3/ Q ‘7.” $7 Ip(%) logft ge Q «R? 43" 11:1 50:01 0?. \ 2378.2 .8 4'7 e’ 9 8° 3" 5+1 54+02 ————— —-.-—-°;7 —————— 2032 I . \ \ 1 a? 4‘7 49' 20:2 48:01 ;‘ pvg 1860.8 1 0' . : "9 3 4? . 5 4°? «5' (3/2'5/2‘) . 3: ‘9 4,? 20:5 49:02 ' ! $._1497 / ‘11 _ 3’ 5:4 56:04 (1/2,3/2-) "' 1142.2 s 1L Q, S? “)7 7/2') .5; 9:2 5.6:01 ( ~ 157.2 30:15 51:02 (5/2‘) ' 0 61Mn Figure 4.27: Decay scheme for the 8 decay of 61Cr to states in 61Mn. The number in brackets following the 7—ray energy is the absolute intensity. Qfi was determined from data in Ref. [80]. 118 Counts/ keV w A l T OHM w uh I 535[2] Counts/ keV hot—N l b.) I 355[4] Counts/ keV linumuluuln (a) 157 keV (b) 355 keV 1 142[3] (c) 1142 keV l 500 N c°~1 1000 1500 2000 Energy (keV) Figure 4.28: wry-coincidence spectra for the decay of 61Cr, gated on the (a) 157-keV, (b) 355-keV and (c) 1142-keV 7-ray transitions. These coincident spectra include data within a 3 s correlation time, for implants collected over the entire surface of the DSSD. In square brackets following the energy for each peak is the number of counts in that peak. 119 have M1 multipolarity. The proposed levels at excitation energies of 1861 and 2378 keV are apparently fed by allowed fl decay as well, and can take J 7’ values in a range from 3/2— to 7/2". 120 Chapter 5 Discussion 5.1 Interpretation of the low-energy levels in neutron-rich 218C One of the more interesting questions with regard to nuclear structure in the f p shell is that of the possible N :34 subshell closure in the 200a isotopes. The maximum upward monopole shift of the V1f5/2 orbital with the completely empty 7r1 f7 /2 or- bital in the Ca isotopes may result in a significant gap between the 1/1f5/2 orbital and the lower-lying 112121 /2 orbital, creating a subshell gap at N =34 similar to that previously observed at N :32. As discussed previously in Section 4.1, a number of experiments have been performed to probe the structure of 54Ca directly, with the hope of determining whether or not a subshell closure exists at N =34 for the Z =20 Ca isotopes. However, these experiments have proven difficult, and to date, no direct measurement has been successful. Having one proton in addition to the Ca isotopes, the low-energy structure of the 218C isotopes may provide insight into the neutron single-particle energy spacing near N =34, and thus into the possible N =34 subshell closure in Ca. The low-energy structures of 53’54’568c have been determined in the present work, by considering both the B decay of the parent nuclei 53154(3a, as well as the isomeric decay of 5415680. 121 The low—lying levels of these 2180 isotopes will be discussed in the context of the extreme single-particle model, as a coupling of the odd 1 f7 /2 proton to the valence neutron configurations in the 20Ca core isotopes. The success or failure of simple coupling schemes to explain the observed low-energy structure of the Sc isotopes can also provide indirect indications of the presence or absence of subshell closures in the Ca core nuclei. Below, the structures of the neutron-rich Sc isotopes are discussed in the frame- work of the simple coupling of the valence proton and neutrons. The experimental levels are then compared to the results of more realistic shell-model calculations. 5.1.1 Low-energy structure of odd-A 53Sc The low-energy structure of 53Sc can be interpreted as a weak coupling of the valence 7r f7 /2 proton to states in 52Ca, which, assuming a robust N =32 subshell closure, can be viewed as a doubly magic core [12]. Coupling of the valence proton to the first 2+ state in 52Ca should produce a quintet of states in 5380 with J 7‘ values ranging from 3/2- to 11/2‘, at energies centered around ~2.6 MeV. Studies of 53Sc produced in deep inelastic reactions [93] have populated 9/2" and 11/2" states near this energy, which are likely members of the 7r f7 /2®2+ multiplet (see Fig. 5.1). The fl decay of 53Ca was observed to populate only one state in 53Se. The population of a single state in the decay of 53Ca decay is expected for allowed fl decay from the presumed 1/2" 530a ground state, since the decay should only populate the 3/2" state of the 7r f7 /2®2+ multiplet. Thus, the single new level at 2109 keV is a likely candidate for the 3 / 2" member of the 7r f7 /2®2+ multiplet. The three known levels in 53Sc, shown in Fig. 5.1, reside at above 2 MeV, as expected within the weak coupling framework. While not all states of the expected quintet have been identified, this simple scheme describes the three known levels well. The apparent success of the weak coupling description for 5380 provides support for the robust nature of the N =32 subshell closure in the 20Ca isotopes. 122 5/2 28 g 2 + (ll/22‘ 2617 1 / :___1 ( ) 2562 (9/2)_ 2283 2% a; “f7/2®52Ca(2+) (3/2)- 2109 E/ : 7%? lélgi 0+ 0 (7/2)- 0 7,2- 0 EXP EXP GXPFIA 52Ca 53sC 53SC Figure 5.1: The comparison of the known levels in 53Sc [75,93] to states in 52Ca suggests that 53Sc is accurately described by weak coupling of the odd 7rf7/2 proton to the 5203. core. The experimental states are also in good agreement with predictions of shell-model calculations using the GXPFlA effective interaction [94]. A more advanced approach to describe the low-energy structure of 53Sc comes from considering more realistic wavefunctions, that allow configuration mixing. Shell- model calculations using the GXPFlA effective interaction [34] were performed for 53Sc [94], and are presented in Fig. 5.1. The calculated levels are in good agreement with the three known excited states, although there are significantly more calculated levels. Additional experimental work is required to identify the remaining levels and complete the expected 7r f7 /2®2+ multiplet in 53Sc. 5.1.2 Low-energy structure of even-A Sc isotopes The structure of the even—A (odd-odd) 218C isotopes can be described by coupling the valence 7rf7/2 proton particle with p3 /2, 191/2, f5/2, and at higher energies, 99/2, va- lence neutron configurations. The success of such a description is shown most directly by considering the cases of 508C and 52Sc, where this simple interpretation explains many aspects of the known low-energy structures. 508C has one proton and one neutron outside of the doubly magic 48Ca core, and the lowest energy levels can be described by coupling a valence 7r f7 /2 proton and 1423/2 123 neutron. The low-energy structure, known from fl decay [95], transfer reactions [96], and charge exchange reactions [97] is summarized in Fig. 5.2(a). The four states lowest in energy with J 7' = 2+ to 5+ arise from the configuration (71' f7 /2)1®(l/p3 /2)1. The particle-particle coupling rules [21] predict the arrangement of the states of this multiplet in a downward-opening parabola, which agrees with the experimental energy level ordering. Promotion of the valence neutron in 5080 to the 1421/2 level gives the configuration (7r f7 /2)1®(1/p1 /2)1, producing only two states with J 7r=3+, 47L. 'IWo 3Jr states are identified in the level scheme of 508C above 2 MeV, and either is a candidate for the 3+ member of this doublet. The energy spacing between multiplets arising from configurations involving the V123 /2 and 1411/2 states depends on the up3/2-1/p1/2 single-particle spacing, and thus the Vp3/2-l/p1/2 spin-orbit splitting. The apparent separation of these two multiplets in 5080 is on the order of 2 MeV, a gap sufficient to account for an established N =32 subshell closure in 2180. The structure of 52Sc can be similarly described by the coupling of a valence 7r f7 /2 proton particle and 1403/2 neutron hole, taking 520a as the inert core [93]. States in this odd-odd nucleus have been identified in 5 decay studies [11], in—beam 7—ray spec- troscopy following secondary fragmentation [98] and deep-inelastic work [93,99]. The known levels are presented in Fig. 5.2(b). Shell model calculations for 5280 using the GXPFI, GXPFlA and KB3G effective interactions are presented in Fig. 5.3(a). As discussed in Refs. [93,99], the lowest-lying quartet of states in 528C can be associated with the configuration (7r f7 /2)1®(Vp3 /2)—1. The four states with JW=2+-5+, should form an upward-opening parabola, since the coupling involves a proton particle and neutron hole. The lowest 3+ and 4+ states in this parabola are expected to be close in energy. The energy separation between these states has not been established exper- imentally, but is likely to be small [99]. The Pandya transform [100] can be applied to relate the particle-particle (7r f7 /2)1®(I/p3 /2)1 states of 5080 to the particle-hole (7r f7 /2)1®(up3 /2)_1 states in 5280. The Pandya relationship assumes the validity of j j coupling and is based on the simple form of the coefficients of fractional parent- 124 (1)+ 2614 (1)+ 2746 (1) \-’2527 (3)+ 2331 (6)+ 2332+x (3)+ /‘\2222 (1)+ 1848 (4)+ 1654+x 1 + 1 4 + (2)+/—\587 (5)+ 212 (3)+ 328 (4)+ +X (1)+ 24 (5’6” x g2)+ F, 252 x (4gp —— (5)+ 0 (3)+ 0 ($43—ng Q)+ 0 503C29 523C31 54SC33 563C35 (a) (b) (C) (d) V1f5/2 V2P1/2 —— —‘°—‘ —‘°_ v2p3/2—oooo— —ooo<>—- —-oooo—— —oooo-— Figure 5.2: Experimentally known levels in the even-A (a) 50Sc29, (b) 528cm, (c) 54SC33, and (d) 563035 isotopes. Also included are the expected ground-state neutron configurations in the f p—shell orbitals above the N =28 shell closure. 125 Table 5.1: Results of the application of the Pandya transform to the low-lying 7rf71/2 <8) upfi/2 neutron-particle states in 50Sc to predict the 1r f.;/12 <8) "pl/2 neutron-hole states - 52 1n Sc. 5030 (”f%/2VP§/2) 5280 (”[71/2Vpg/12) J 7' Experimental J 7' Calculated Experimental A(Ecalc — Earp) Energy (keV) Energy (keV) Energy (keV) (keV) 3+ 0 34 26 0 26 2+ 257 4+ 0 x -x 3+ 328 5+ 229 212+x l7-x 4+ 756 2+ 880 675 105 age (the probability of a given J state arising from a specific nucleon configuration) relating one- and two-hole states [100]. The results of this transform applied to 508C to the corresponding states in 5280 are presented in Table 5.1. For a small energy separation between the 3+ and 4+ states (x) in 52Sc, the transform provides excellent agreement, suggesting that the configurations of these low-lying levels in both nuclei are fairly pure. Calculations using the GXPFl effective interaction [12] support the relative purity of the four lowest-energy levels (see Fig. 5.3). At higher excitation energies in 52SC, the coupling of the 7rf7/2 proton to an ex- cited 1421/2 neutron to give a pair of states with J "2.3+, 4+ accounts for the observed 4+ state in 52Sc at ~1.7 MeV. The 6+ state at ~2.3 MeV in 528C has been shown in shell-model calculations to also have parentage that includes this configuration, al— though the extent of this contribution varies between shell-model interactions. Calcu- lations using the GXPFIA interaction [99] suggest that the primary parentage of the 2.3—MeV 6+ state is (7rf7/2)1®(1/(p§/2p[/2)), while the (7rf7/2)1®(V(p§/2f51/2)) con- figuration is predicted to dominate the first 8+ state in 52Se, experimentally observed at ~3.6 MeV [99] [see Fig. 5.3(a)]. Experimentally, the energy separation between states dominated by the (Wf7/2)1®(Vp3/2)_1 and (7r f7 /2)1<§§>(1/(p;23 /2p[ /2)) configura- tions appears to be ~2 MeV, again corresponding to the spin-orbit splitting between the 1423/2 and 1421/2 single-particle states, in agreement with the results for 508C. As 126 28:088.: 858% an. 089 Be .3. <3me ..3 E88 2: we... 823830 38. :3. £5 mmaopofl omen 3 was omvm 3V .ommm 93 “$295 23 3 >22 m .333 £32 :Boaxizgcofitmaxm 2.: mo :oESQSOO ”mat. 8:me 3 3V 3 mm mm Hm omen omen ommm .082: 5.3on 588i BE 20391 3.3wa :Exoi. mxm 1029+... fimmxwv 558$ mxm . o I I... I I I I I I |I+m o +3 [in I+m o +3 In \ In If... 0 +5 Hm 3 [. +3... In. S... a: I... I... I... . ..s +~ IL 3 .+_ wIINI I+o .IIIL+o [+0 :3 +6 [Hm 3 +2 +5 [Rm 3 I+~ +N +~ who is IL+ N. +v fl H Iim I+v [am it I: I+v Ii. l+ I J...m II“... I... .. Hm I“... I... 1.3% I+m I.L +0 + I... a ll H f I: L: L: 521.: I+v + I? I+o +~ +N .3 M I... .II... I... I. T. .L... N o + I+N I+ \H \+m Ik+ qua L3 If J+~ +m +0 [+0 1 . 1 II. II»... +. .2. Ir... . :3 ..1... m. I... I... 1.. I: I... a I+m + + I. I L. I j m I+m i + +o in I HM it I H... + _ .. I... I... . w 127 noted by Fornal et al. [99], the apparent separation between the 1421/2 and V f5 /2 single-particle states, as inferred from the energy spacing between the 6+ and 8‘L states in 528C, is slightly smaller than that assumed in the GXPFIA interaction. The low-energy structure of 54Sc, with two additional neutrons, should be sig- nificantly different. The 1423 /2 orbital is now fully occupied, and the 54Sc structure at low energy should reflect coupling of the valence f7/2 proton particle to either a 121/2 or f5/2 neutron, again taking 52Ca as an inert core. The 1421/2 single-particle level is expected to lie below the V f5 /2 state in the 218C isotopes, based on the ob- servations in 52Se. Assuming this ordering, the (7rf7/2)1®(Vp1/2)1 coupling yields a doublet of states with J 7r23+, 4+, which is expected to be lowest in energy in 54Sc. Shell model results [31,94] predict the 3+ and 4+ states of this configuration to be close in energy, and thus the spin and parity of the 54Sc ground state is not clear from a theory perspective. The ground state of 54Sc has been tentatively assigned J 7r23+ in the present work (see Sections 4.1.2 and 4.1.3). Considering the situation purely in terms of the proton-neutron coupling, it may be natural to assume that the isomeric 110-keV level corresponds to the 4+ state arising from the (7r f7 /2)1®(up1 /2)1 config— uration. However, within the simple coupling framework it is difficult to understand why the in-multiplet M1 transition is hindered strongly enough to yield the experi- mental isomer half-life of 2.8 us. In this case, the two other states identified in 54Sc, the isomeric (5)+ level and higher-lying (1)+ state, may arise from the configuration (7rf7/2)1®(1/f5/2)1. This coupling results in a sextet of states ranging from J"=1+ to 6+, and the particle-particle coupling would produce a downward-facing parabola. There are however, inconsistencies in the level ordering compared with the spin and parity assignments made in 54Se. The 1+ state lies above the possible 5+ state in energy in the level scheme as presented in Fig. 5.2(c), a result which is inconsistent with the expectations of the particle-particle parabola, and calculations using the GXPFI, GXPF 1A and KB3G effective interaction [see Fig. 5.3(b)]. An alternative J 7' assignment for the states in 54Sc with a 4+ ground state and 6+ llO—keV isomeric 128 level is a possibility, but the placement of the 6+ level below the 1+ level is also unex— pected would also contradict the shell-model predictions. The apparent experimental level ordering could be better explained if the (7r f7/2)1®(1/p3 /2)_1 configuration was the origin of the 3+ and 5+ states. However, as demonstrated for the lighter Sc iso- topes, the Vp3/2-l/p1/2 spin-orbit splitting is a fairly constant N2 MeV, so that the V f7 /2-up3 /2 multiplet should reside at ~2 MeV in 543C. An alternative to the 5+ assignment for the 110-keV isomeric state is a 4+ assign- ment. While not expected based on comparison of the deduced isomer lifetime with Weisskopf single-particle estimates, which favor an E2 multipolarity for the 110—keV transition, a 4+ spin and parity cannot be excluded for the 110—keV state. Calculations using the GXPFl [33], GXPFIA [34] and KB3G [35] interactions all predict a doublet of states near the 54 Sc ground state with J 7r23+ and 4+. The transition probabilities, B(M 1 : 3+ —> 4+) and B(E2 : 3+ —> 4+) calculated with these three interactions are shown in Table 5.2, along with the calculated partial half-lives. In all cases, the M1 component would dominate the transition. However, the B(Ml : 3+ —> 4+) val- ues are small, giving rise to half-lives of order nanoseconds or longer — much slower than the picoseconds expectation from the single-particle Weisskopf estimate. The B(Ml : 3+ —+ 4+) value is very sensitive to the degree of mixing in the wavefunc- tions for the 3+ and 4+ states [94]. With increased configuration mixing, there is increased suppression of the B (M 1 : 3+ -—> 4+) value, hindering the transition. A 4+ spin and parity assignment is in better agreement with the expectation for the level ordering in 54Sc. However, more work is required to make a firm spin assignment for the llO—keV isomeric state in 54Sc. The 1+ state at energy 247 keV in 54Sc can only be explained by the (7r f7 /2)1®(u f5 /2)1 configuration. Even if this is the only member of the multiplet identified, the energy separatiOn between the ground state with (7r f7 /2)1®(up1 /2)1, and the excited state with (7r f7 /2)1®(V f5 /2)1, appears to be small. Shell model results using the GXPFI and GXPFlA effective interactions [31,94] predict a upl/g-I/f5/2 separation of more 129 Table 5.2: Calculated transition probabilities and half—lives for the transition between the 3+ and 4+ states in 54Sc, using the GXPFl, GXPFIA and KBBG shell-model effective interactions. The Weisskopf single-particle estimates for the half-life of a 110—keV transition are also included. GXPFI GXPF 1A KB3G Single-Particle B(Ml : 3+ —+ 4+) (11%,) 0.02611 0.00369 0.00080 T1/2(M1)(s) 8.10x10“10 1.56x10-8 8.94x10‘7 9.30x10-12 B(E2 : 3+ —» 4+) (e2fm4) 5.07 6.24 6.96 T1/2(E2)(s) 4.05x10-6 1.72x10-5 1.03><10-3 2.89x10—6 than 1 MeV, which manifests itself in 54Sc as an expanded low-energy level structure, seen in Fig. 5.3(b). The compression in the experimentally established structure of 54Se suggests a Vp1/2-Vf5/2 single-particle energy separation smaller than that as- sumed in the GXPFI and GXPFIA effective interactions, and is inconsistent with a robust N :34 subshell closure. Another rapid structure change should be evident in the heavier 56Sc nucleus, since the addition of two neutrons to 54Sc will fill the 1411/2 single-particle orbital. The unpaired valence neutron will then occupy the V f5 /2 state. Under this assump- tion, the lowest energy states in 568C would have a (7r f7 /2)1®(V f5 /2)1 configuration. The resulting sextet of states with J"=1+ to 6+ should be arranged in a downward facing parabola, with the 4Jr state at the vertex. The challenge in comparing these ex- pectations in 568C to the new data is that the absolute energies of the two fi-decaying states are not known. The states expected at low energy in 568C with J”:1+, 2+ and (5,6)+ are likely attributed to the (7r f7/2)1®(1/ f5 /2)1 configuration. The 3+ and 4+ states cannot be trivially assigned the same configuration, as they may also arise from the (7r f7 /2)1®(Vp1 /2)1 coupling, depending on the relative position of the 1421/2 and uf5/2 single-particle orbitals. In fact, significant mixing would be expected be- tween the 3+/4+ states in the 7r f7 /2-u f5 /2 and 7r f7 /2-Vp1 /2 multiplets if there is only a small energy gap between the uf5/2 and 1491/2 effective single-particle energies, as suggested in 54Se. This mixing can also be seen in the calculated levels [94], even with 130 a large Up1/2-l/f5/2 gap. 5.2 Re-ordering of proton single-particle states in neutron-rich K isotopes The neutron—rich 19K isotopes with N 2 20 sit on the border of two major shells, with protons occupying the nearly full sd shell below Z =20, and neutrons occupying the next major shell, the f p shell. As such, shell-model descriptions of the neutron- rich K isotopes remain challenging, as cross-shell interactions are required to describe the interaction of valence sd—shell protons and f p—shell valence neutrons. Interactions which account for excitations of nucleons across the shell gaps are even more chal— lenging than those within a given shell. However, new interactions do exist which attempt to accurately reproduce the cross-shell interactions between the sd and fp shells [101—103]. The results of shell-model calculations using such interactions will be discussed for the region surrounding 50K in the following sections. However, the structure of the neutron-rich K isotopes will first be described in terms of coupling of the valence nucleons, and the expected shifts in nucleon single-particle energies resulting from the monopole interaction. The systematics of the neutron-rich 19K iso- topes surrounding the N =28 shell closure are examined within this framework in the next section. 5.2.1 Systematics of the 19K isotopes around N :28 As discussed in Chapter 1, the tensor monopole interaction between protons and neutrons is attractive between j< (l — 1 / 2) and 3') (l + 1 / 2) orbital pairs and repul- sive for j<-j< or j>-j> orbital pairs. The strength of the interaction is maximized between single-particle orbitals with similar I values [17]. Considering the effect of 1f7/2 (1+1 / 2) neutrons on the proton sd—shell orbitals, it is apparent that the tensor 131 min-La!" monopole interaction with the 1d3/2 (l-1/2) proton orbital will be attractive, while the interaction will be repulsive for the 1d5/2 (1+1 / 2) proton orbital. The 2.31/2 pro- ton orbital should exhibit no tensor monopole shift, since there are no partner orbitals for [20. Thus, based solely on the energy shifts expected due to the tensor monopole interaction, the 1d3/2 proton orbital is expected to decrease in energy relative to the 231/2 orbital with increasing neutron 1f7/2 occupancy. This expectation appears to be borne out in the experimental data for the even-N K isotopes between N =20 and N =28. The first 3/2+ state in the odd-A K isotopes can be associated with the --1 1d; /12 proton-hole configuration, while the first 1/2+ state is associated with the 231/2 proton-hole (equivalently, the 23b2 proton-particle) configuration. This relationship has been confirmed in the case of 39K in the analysis of proton pick-up reactions with 40Ca [104]. Thus, the relative energy of the 3/21” and 1/2] experimental levels is directly related to the separation between the proton 1d3/2 and 281/2 single—particle orbitals. A change in the energy separation of the 1d3/2 and 2.91/2 proton-hole states in the odd-A K isotopes with increasing 1/1f7/2 occupancy is apparent in Fig. 5.4. The 71281712 state, which is. initially well—separated from the 7r1d; /12 ground state in 39K, rapidly drops in energy relative to the «1d; /12 state as the neutron 1f7/2 or- bital is filled. In fact, in 47K28, with a full 111 f7 /2 orbital, the proton 1d3/2 orbital has dropped below the 7r2sl/2 orbital, and 47K has been experimentally confirmed to have a ground state J 7' of 1/2+ [105], corresponding to a 231712 proton hole configuration. Moving past the N =28 shell closure, neutrons begin to occupy the 2123/2 orbital. The tensor monopole interaction between the 2123/2 neutron orbital and the 1d3/2 proton orbital, while attractive, is weaker than the V1f7/2-7I'1d3/2 tensor interaction. Again, the «231/2 orbital should exhibit no tensor monopole shift. However, a second component to the monopole interaction, known as the central force [106], plays a significant role in the evolution of the relative spacing of the 231/2 and 1d3/2 proton orbitals in this region [107]. The central force is an attractive interaction between pro- tons and neutrons, which does not have the spin dependence of the tensor monopole 132 — N r—IUINUI .3 QUI I 9 {J1 I l 20 22 24 26 218 30 32 Neutron Number B(I/zt) - E(3/2‘{) (MeV) Figure 5.4: Systematic behavior of the separation between the 3/2+ and 1/2+ states in the odd-A potassium isotopes between N :20 and N =30. The observed energy separation is directly related to the separation between the 1613/2 and 231/2 proton orbitals in the K isotopes. Experimental data points are indicated by filled circles, while predictions using the SDPF-U effective interaction [101] are represented by the open squares. force. The central force does have an I dependence, and like the tensor force it is max- imized for similar 1. Below N =28, the central force between the u1f7/2 and 1r1d3/2 orbitals is stronger than that between the u1f7/2 and 7r2sl/2. The central force con- tribution, in that case, reinforced the tensor monopole interaction and lowered the 7r1d3/2 state relative to the 7r2sl/2 state. However, above N =28, the central force between the u2p3/2 and W281/2 orbitals is expected to be stronger than the combined tensor monopole and central forces between the u2p3/2 and 7T1d3 /2 orbitals. Thus, with the addition of neutrons to the 2193/2 orbital, the W281/2 orbital is expected to drop in energy relative to the 7r1d3/2 orbital, and the inversion of the proton orbitals, which was observed between 45K26 and 47K28, reverses. The open question is how quickly this reversal occurs, which is related to the strength of the monopole inter- action between the neutron and proton orbitals in the fp and 3d shells respectively. An examination of the experimentally known low-energy structures of the K isotopes beyond N =28 can provide first insight into this question. States in the nucleus 48K, which has one neutron beyond N =28, are known from deep-inelastic work [108] and recent multinucleon transfer reactions [85] and are shown in Fig. 5.5. The ground state of this nucleus was initially assigned a spin and parity 133 of 2‘ on the basis of the expectations from the shell model and observed fl feeding to states in 48Ca [109]. However, Broda et al. [85] reassigned the ground state of 48K as J ”:1- following analysis of multi-nucleon transfer and deep-inelastic reaction data. Such an assignment is still consistent with 48K fi-decay results since direct fl decay to 3' states in 48Ca is not observed. Taking the J7r re—assignments from Ref. [85], the ground state and first excited state at energy 143 keV in 48K can be interpreted as arising from the coupling of the 2123/2 neutron with the 231/2 proton hole, which gives rise to a 1‘ state (the ground state) and a 2‘ state (the 143-keV state). The higher-energy states in 48K at 279 keV and 728 keV should correspond to the 2‘ and 3" states of the (V2p3 /2)1®(7r1d3 /2)"1 multiplet. The 0' and 1" members of this multiplet are unknown at this time. The expectation of an upward-opening parabola for the particle-hole coupling would place the O“ and 1‘ levels above the 2‘ state at 279 keV. The separation between the 1613/2 and 231/2 proton hole states in 48K appears to be of order a few hundred keV, consistent with a narrowing of the 7r1d3 /2- «231/2 gap with the addition of a single 2123/2 neutron. This interpretation suggests ' that the «131/2 level remains above the 1d3/2 level in 48K29. The addition of two neutrons in the K isotopes beyond N =28 corresponds to the odd-A nucleus 49K. Recent data [86] for 49K from multinucleon transfer suggest that the proton 251/2 orbital still remains above the 1d3/2 orbital, as indicated in the systematics shown in Fig. 5.4. In this nucleus, the 7r1d3/2-7r2sl/2 energy separation has narrowed further to only 92 keV. Moving further to 50K, one can explore the relative position of the 7r1d3/2 and «231/2 orbitals with three neutrons now in the 2123/2 orbital. The low-energy levels in 50K from the present work are presented in Fig. 5.6. Assuming that the proton 231/2 orbital has dropped further in energy and is again below the 7r1d3/2 orbital, the ground state configuration in 50K would arise from the (V2193 /2)-1®(7r1d3 /2)—1 configuration. This configuration would account for the 0" ground state as previously assigned based on the 18 decay of 50K to states in 50Ca [110]. However, it is difficult 134 (7*) 3586 (5+) 2177 _ - (712d )'1®(v2p )1 (3 ) 728 3/2 0_ 3/2 _. 3- 1_ (23 279_ 2- 143' _1 1 (1') 0_*—' (“151/2) ®("Zips/2) 48K Figure 5.5: Known states in 48K [85,108], and the possible origin of the low-lying states in the simple framework of the independent particle model. 135 to understand the presence of the observed isomer in 50K if the low-lying states all arise from the (V2p3/2)“1®(7T1d3/2)—1 configuration. As discussed in Section 4.2, the deduced half-life of 15473?3 ns for the 50K isomeric state at 172 keV is consistent with an E2 multipolarity for the 172-keV transition. The observed high level of competi- tion between the direct 172-keV ground state transition and the two ’7—ray cascade suggests that the first M1 transition of the cascade must be hindered. Such hindrance is unexpected for an in-multiplet M1 transition. An alternative is that the ground state in 50K has the (V2123 /2)'1®(7r2sl /2)_1 configuration. The ground state would be expected then to have a spin and parity of 1‘, with the first excited state being the J "=2” state of the 281/2 proton—hole con- figuration. The isomeric state could then be explained as the (112])3 /2)_1®(7r1d3 /2)—1 configuration. The isomeric state would most likely correspond to the 3‘ state of this multiplet. The M1 transition connecting the 3‘ isomeric state and the lower-lying 2‘ state would be hindered as a result of the forbidden nature of a 7r1d3/2-7r2sl/2 transition. Such a hindrance would explain the observed competition between the di- rect 172-keV E2 ground-state transition, and the two 'y-ray cascade. Thus, a J ":1— ground state in 50K, arising from a 231/2 proton hole configuration, seems to provide a-more satisfactory understanding for the observed isomerism in 50K than any state arising from a (1d3/2)—l configuration. Such a result would suggest that, even with three 2123/2 neutrons, the 2231/2 orbital remains above the 7r1d3/2 orbital. The rela- tive energy of states arising from the two proton-hole configurations would suggest a 7r1d3/2-7r231/2 gap comparable to that in 48K, of order a few hundred keV. However, to resolve the uncertainty in this situation, a direct measurement of the ground state spin and parity is necessary. 136 37— 671 3:—— 335 l 312 \ \ _ .38; ‘ ,, (fl2d3/2)—1®(V2p3/2)_1 ._ (3) N, 53’ “$8 171.8 ’ (2)‘ ~33" 128.4 (fl251/2)—1®(V2p3/2)—1 .4 2——— 89 _ (I). 0 0-___0 EXP SDPF-U 50K Figure 5.6: Proposed low-energy level scheme for 50K, and the possible origin of the low-lying states in the simple framework of the independent particle model. The energy of the suggested 2‘ state is not uniquely determined, and may alternately be located at energy 42.8 keV. The states below 1 MeV, calculated using the SDPF- U [101] effective interaction are also shown. 137 5.2.2 Comparison to shell-model calculations using sd— f p cross-shell interactions There has been significant work in recent years on the deveIOpment of shell-model effective interactions for calculations in the sd— f p valence space, due primarily to the experimental observation of a weakening N =28 shell closure [111,112]. These new ef- fective interactions consist of three basic components: (a) a proton-proton interaction in the sd shell, (b) a neutron-neutron interaction in the f p shell, and (c) a cross-shell proton-neutron interaction. Generally, standard interactions are used for the intra- shell parts, such as the USD interaction for the sd shell [113], and the KB [114] or GXPFIA [34] interactions in the f p shell. The most uncertain part of the 361- f p shell model calculations is the cross-shell component. Utsuno et al. [102] have recently developed a new sd— f p shell interaction that makes use of the USD interaction in the sd shell, and the GXPFIB interaction (a slight modification of the GXPFIA interaction with a better 1p3/2-1p1/2 energy gap around the Ca isotopes [102]) in the f p shell. The cross-shell part of the interaction is built up from three components: a central force, a spin-orbit force and a tensor force. Results for the K isotopes using this effective interaction have been found to be consistent with experiment through 47K [102], correctly predicting the inversion of the «231/2 and 7rld3/2 single-particle orbitals. Another recent sd— f p shell effective interaction has been developed by Nowacki and Poves [101]. Their SDPF-U interaction is a refinement of the previous SDPF-NR effective interaction [115], which makes use of the USD interaction in the 3d shell, the KB interaction in the f p shell, and the G matrix calculations of Kahana, Lee and Scott for the cross-shell interaction. The monopole part of the cross-shell interaction was empirically modified to produce the correct evolution of effective single-particle states by fitting to a number of reference states across the region of interest. Results from this interaction for the 19K isotopes are in good agreement with experiment up 138 to and including N =28, correctly predicting the inversion of the 7r2sl/2 and 7r1d3/2 single particle orbitals in 47K [101]. However, immediately beyond N =28, the SDPF- U effective interaction predicts a 3/2+ ground state in 49K30, in contradiction with recent experimental results [86]. The separation between the 71'281/2 and 7r1d3/2 sin- gle particle orbitals predicted by the SDPF-U interaction for N > 30 continues to increase [101] and is inconsistent with the suggested (V2p3 /2)‘1®(7r2s3 /2)"1 ground state configuration in 50K. Shell-model calculations for 50K were performed using the SDPF-U effective interaction [94] and are presented in Fig. 5.6. The apparent discrepancy between theory and experiment may be related to the 2p3/2-1d3/2 and 2123/2-131/2 monopole strengths. These monopole interactions largely determine the evolution of the «231/2-7r1d3/2 single-particle energy spacing with the filling of the V2193 /2 orbital beyond N =28. Additionally, the 2p3/2-1d3/2 and 2123 /2-281 /2 monopole strengths were changed between the SDPF-U interaction and the previous SDPF-NR interaction, which had correctly reproduced the 1/2+ 49K ground state [101]. Addi- tional data will be required to better constrain the cross-shell effective interactions, and improve the predictive power of sd—fp shell interactions. 5.3 Onset of collectivity in the 25Mn isotopes Approaching N =40 below the Z228 shell closure, an increase in collectivity has been inferred at low energy for the neutron-rich 24Cr and 25Fe nuclei. The apparent onset of deformation in these isotopes has been associated with the presence of the neutron 199/2 single-particle orbital near the Fermi surface as N =40 is approached. The observed structural changes in this region are again attributed in part to the tensor monopole interaction. The repulsive interaction between 1 f7 /2 (l+ 1 / 2) protons and 199/2 (1+ 1 / 2) neutrons is reduced with the removal of 1f7/2 protons below Z 228, and the 199/2 orbital drops in energy towards the Fermi surface. The role of the V1 99 /2 orbital in driving deformation, as a function of nuclear charge, will be critical 139 to assessing the persistence of the N =50 shell gap in 7BM and other neutron—rich nuclei in this region of the nuclear chart. As noted in Chapter 1, the systematic variation in E(2'1f) values [F ig. 1.9(b)] sug- gest that collectivity sets in already at N = 36 for the Cr isotopes, while first evidence for such effects in the Fe isotopes occurs at 64Fe, which has N = 38. Therefore, the low-energy structure of 61Mn, with N = 36, might exhibit features at low energy sug- gestive of a change in collectivity when compared to other odd-A Mn isotopes nearer to stability. The 6 decay of 61Cr to states in géMn36 was studied in the present work. The newly identified states extend the systematics of the odd-A Mn isotopes through N :36, and can be investigated for signs of developing collectivity. 5.3.1 Systematics in odd-A Mn The systematic variation of the known energy levels of the odd-A Mn isotopes with A = 57 — 63 is shown in Fig. 5.7. The levels below 2 MeV for 57Mn were taken from H decay of 57Cr [116], in-beam y-ray spectrOSCOpy [117], and the (d,3He) transfer reaction [118]. [3 decay populates seven excited states below 2 MeV, as well as the 57Mn ground state. The states directly fed by 16 decay are all assumed to have negative parity, since the 57Cr parent has ground state J 7' = 3/2‘. The negative parity yrast states observed in 57Mn were populated by a heavy-ion induced fusion-evaporation reaction, that provided tentative spin assignments of the levels up to J = 25 / 2. The (d,3He) transfer reaction provided unambiguous J 7’ assignments to the negative parity states below 2 MeV at 84, 851, and 1837 keV, based on angular distribution data. Positive parity was deduced for the excited states with energies 1753 and 1965 keV. Based on the spectroscopic strengths for proton pickup [118], the first 7/2“, 1/2+, and 3/2+ levels in 57Mn were deduced to carry ~ 40% of the summed single—particle strength from shell model expectations. Valiente—Dobon et al. [57] studied neutron-rich Mn isotopes by in-beam *y-ray spectroscopy following deep inelastic collisions of a 70Zn beam on a 238U target. 140 Negative-parity yrast states were identified up to J = 15/ 2 in 59Mn34 and J = 11/2 in 61Mn36. Only a single transition, with an energy of 248 keV, was assigned to the structure of 63Mn38. The non-yrast levels in 59’6an shown in Fig. 5.7 were identified in fl-decay work. The B decay of 59Or to levels in 59Mn was most recently reported by Liddick et al. [69]. The ground state spin and parity of the parent 59Cr, tentatively assigned to be J 7‘ = 1/2“, restricts the range of states in 59Mn accessible by allowed [3 decay to those with J7r = 1/2-,3/2_. The ground state of 59Mn is known to have J7r = 5/2— [119]. The non-yrast J7r values for 59Mn were inferred‘from the fl- and *y—decay patterns. The present work has expanded the knowledge of states in 61Mn, identifying three new states, in addition to the three previously known yrast states [57]. The negative parity yrast structures below 2 MeV in odd-A 57’59’61Mn exhibit little variation as a function of neutron number. These levels are consistent with shell model calculations reported in Ref. [57] employing the GXPFIA, KBBG and f pg effective interactions. The authors of Ref. [57] do note, however, that the proper ordering of the 9 / 2“ and 11/2" levels at N 1 MeV is reproduced for all three isotopes only by the f pg interaction. The experimental energy gap between the ground state, with J7r = 5/2", and the 7/2" first excited state exhibits a regular increase from 57Mn (83 keV) to 65Mn (272 keV) [58]. Both the GXPFIA and fpg shell model interactions reproduce the observed trends well. The shell model calculations in Ref. [57] were extended to include non-yrast states below 2 MeV. The calculations were performed with the code ANTOINE [120] using the GXPFlA [34] and fpg [46] effective interactions. The full fp model space was utilized for the calculations with the GXPFIA interaction. The valence space for the calculations with the cross shell interaction f pg was limited to 2 neutron excitations from the fp shell to the 1g9/2 orbital. As compared to Ref. [57], truncation of the basis states was necessary to reach convergence for the higher density of states at lower spin for the most neutron-rich Mn isotopes considered here. However, in those 141 A v %2— 3 .......... 5 §-—-/3 3 —' 3"“ .......... l 66 7-~-- 7 33 ..__—- — 5-—-— _5 I: _ 5"“‘/ w..- 7.-...-,._/ 7—-—*"‘—'/ 5 —‘ 5 — Expt GXPFIA fpg Expt GXPFIA fpg 57Mn32 59M1134 L 7 Energy (MeV) N __ 7—-——\ 7 4- —— _\___ 5 5 Expt GXPFIA fpg Expt GXPFIA fpg 61Mn36 63Mn38 Figure 5.7: Systematic variation of the low-energy levels of the neutron-rich, odd-A 25Mn isotopes, along with the results of shell model calculations with the GXPFIA and f pg interactions. Only levels below 2 MeV are presented. In addition, the shell model results shown are limited to the three lowest energy levels calculated for J 7' = 1/2" — 7/2‘, and the yrast levels for J7r = 9/2" and 11/2". Levels with assumed negative parity are shown as solid lines, and those with positive parity are presented as dotted lines. Spins are given as 21 and, in nearly all cases, are tentative as discussed in the text. 142 cases where the calculations were completed with both 2 and 6 neutron excitations allowed from the f p shell, similar results were obtained. The low density of levels below 1 MeV is persistent in 57’59’61Mn, and agrees well with the shell model results with the GXPF 1A effective interaction. The only excited state in this energy range is the 7/2‘ level, which is expected to carry a reduced 1f7/2 single-particle component with the addition of neutrons [118]. The low-spin level density below 1 MeV is shown to increase at 61Mn in the shell model results with the fpg interaction. Here, the influence of the 199/2 neutron orbital becomes apparent at N = 36. However, this is not borne out in the observed level structure. The regular behavior of the low-energy levels of the odd-A 25Mn isotopes through N = 36 follows the similar trend observed in the even-even 26Fe isotopes, where the possible onset of collectivity at low energy is not evident until N = 38 is reached. 5.3.2 Even-A Mn isotopes Although no new data for the even-A Mn isotopes are reported in this work, the systematic variation of these states provides additional support for the delayed onset of collectivity in the Mn isotopes. Only a few excited states in the odd-odd 60’62Mn were established based on fl-decay studies [58,121]. The low-energy levels in these isotopes are shown in Figure 5.8. Two ,B-decaying states are known in both isotopes, but the location and ordering of the proposed 1+ and 4+ states has not been estab- lished in 62Mn37 [58]. Good agreement was noted between shell model results with the GXPFI interaction and both the low-energy structure and fi-decay properties of 60Mn35 [121]. The low-energy structure of 62Mn built on the 1+ fi-decaying state shows marked similarity to that in 60Mn, albeit the excited 2+ and 1+ states are both shifted ~ 100 keV lower in energy with the addition of two neutrons. Although Valiente-Dobon et al. reported a number of in-beam 'y rays associated with the de- population of yrast levels in both 60Mn and 62Mn, no level structures were proposed due to the lack of coincidence data [57]. Again, the systematic variation of the (few) 143 E 2 ’- ,gs __ CI LT-l 1 _ 1+ 1+ 2+..— 4+ (04'2“) if 2“— ._ + 1+ + 0 1 mm 1 60MI135 64Mn39 Figure 5.8: Experimentally known low—energy level structures for the even—A Mn isotopes from N235 through N239. The location and ordering of the proposed 1+ and (3+, 4+) states in 52Mn is unknown — for the purposes of the figure, the 1+ states were aligned for all isotopes. excited levels established in the odd-odd Mn isotopes through N 2 37 does not sug- gest a sudden onset of collectivity, in line with the trend established for the yrast states of even-even 26Fe isotopes. Beyond N237, low-energy levels in 64Mn are known from the 6 decay of 64Or [122]. A low-lying 2‘ state has been identified at 40 keV in 64Mn, as shown in Fig. 5.8. Identification of such an unnatural parity state is an indication for the approach of the 11ng /2 level close to the Fermi surface in this isotope. This structural signature provides the first indication of the approaching deformation-inducing 1g9/2 orbital in the Mn chain. Additional spectroscopic measurements in this isotopic chain are required to more fully characterize the onset of deformation in the Mn isotopes. 144 30 32 34 36 38 40 42 44 46 Neutron Number Figure 5.9: Two neutron separation energies in the N 240 region. Hollow symbols indicate masses taken from the 2003 Atomic Mass Evaluation [80], while filled symbols are masses measured recently using the time—of-flight method at NSCL [123]. Figure was adapted from Ref. [123]. 5.3.3 Deformation in comparison to neighboring 24Cr and 26Fe isotopic chains The low density of states in 61Mn has been shown to agree with the results of shell model calculations in the f p model space without the need to invoke neutron ex- citations into the 199/2 orbital. There is no compelling evidence for the onset of collectivity in the 25Mn isotopes through N237. However, at N239, there is evi- dence of the influence of the 1g9/2 neutron orbital, in the form of a 2‘ unnatural parity state at low energy in 64Mn. These observations are consistent with the be- havior of the neutron-rich Z +1 26Fe isotopes, where no indication of collectivity was observed until N 238. While spectroscopic information is not available beyond N239 in the Mn isotopes, nuclear mass systematics in the region surrounding N 240 have recently been extended in the Cr, Mn and Fe isotopes [123]. The systematics of the mass observables, such as the two—neutron separation energies (82"), are sensitive to deformation effects, as discussed in Chapter 1. Two-neutron separation energies are shown in Figure 5.9 for the isotopes surrounding N 240 and Z225. 145 As discussed by Estradé [123], a reduction in the slope of the Sgn values as a func- tion of neutron number has been observed to coincide with the onset of deformation in the Fe isotopes. The reduced slope corresponds to extra neutron binding as defor- mation sets in. The reduction in the slope begins in the 25Mn isotopes at N237, in agreement with the spectroscopic observations to date. A change in the slope of the two neutron separation energies for the Fe isotopes is observed at N 240, coinciding with the onset of deformation in the Z226 isotopic chain. The Cr isotopes present a less clear case. A subtle change in the slope begins at N 236, in agreement with the earlier onset of deformation in the even-even 24Cr isotopes evident in the systematic variation of the 2? states. Thus, mass systematics appear to agree with spectroscopic observations for the onset of deformation in the Cr, Mn and Fe isotopes. Collectivity, resulting from the lowering of the [1199/2 orbital closer to the Fermi surface, occurs earliest in the 24Cr isotopes, while the onset of deformation in the 25Mn isotopic chain occurs later, more like the situation in the 26Fe isotopes. With the removal of additional 1f7/2 protons below 24Cr, in the 22Ti and 20Ca isotopes, the V199 /2 orbital is expected to be lowered further, and collectivity may develop even sooner. The lowering of the V1 99 /2 may also lead to a new “island of inversion” surrounding 3%Ti40, as first suggested by Brown [124], and discussed by Tarasov et al. in light of enhanced production cross sections in the region near 62Ti [125]. 146 Chapter 6 Conclusions and Outlook 6. 1 Conclusions The low—energy level structures of nuclei within and near the fp shell have been investigated using fi-delayed and prompt 7-ray spectroscopy, to find evidence for changing nuclear shell structure. The low-energy levels of neutron—rich 535436180 were investigated through the fi decay of the parent nuclides 53'54Ca, as well as through prompt isomeric 7-ray emission from 541568c. These nuclei reside in the vicinity of the known N 232 subshell closure, a result of the large gap between the u2p3/2 orbital and the higher-lying V2121 /2 and V1f5/2 orbitals. The low-energy structures of the Sc isotopes were analyzed in the extreme single-particle model framework, as a coupling of the valence 1f7/2 proton to states in the corresponding 20Ca core. Such a description worked well for 53Se, where the valence proton weakly couples to states in doubly-magic 33Ca32. The success of this description provided further confirmation for the validity of the N 232 subshell closure in the Ca isotopes. Interpretation of the low-energy levels in the even- A Sc isotopes as resulting from the coupling of a 1f7/2 proton and valence neutrons to a Ca core provided information on the separation between the V2p1/2 and u1f5/2 orbitals. The energy separation between these orbitals is proposed by some shell- 147 model interactions to be sufficient to produce a new subshell closure at N 234 in the Ca isotopes. While more work is required to complete the relevant multiplets of states in 541568c, early indications suggest that the low-energy levels in the Sc isotopes are compressed relative to shell-model predictions and that a significant N 234 subshell gap does not exist between the 121/2 and f5/2 neutron levels in the 218C isotopes. More information is required for the 20Ca isotopes, specifically at 54Ca to determine conclusively whether the removal of the final f7/2 proton is sufficient to create a substantial subshell closure at N 234. The low-energy isomeric structure in §8K, with protons in the sd shell and neu- tron occupying the f p shell, was investigated using prompt 'y-ray spectroscopy. The observed isomeric structure of 50K was suggestive of a re—ordering of the proton sin- gle particle states below the Z 220 shell closure. The isomeric structure was best explained with a V2123 /2 <8) 71231/2 configuration for the ground state in 50K, which gives rise to a 1‘ ground state spin and parity. This is inconsistent with previous assignment of 0‘ for the ground state of 50K, based on fl-decay measurements [110]. The predictions of the most recent sd- f p cross-shell effective interactions, the SDPF— U interaction from Nowacki and Poves [101], and the newest interaction from Utsuno et al. [102], are also inconsistent with the suggested 1‘ ground state in 50K. Theo- retical work is required to reconcile the predictions of these cross-shell interactions with the recent experimental results in the K isotopes. Additional spectroscopic in- formation is also required for the K isotopes beyond N 228 to provide the constraints necessary to more fully understand the monopole interactions involved at work, and the evolution of the proton single-particle energies below Z 220 in the region. The 6 decay of 61Cr was studied to extract details on non-yrast excited states in the daughter nucleus ggMn36. This nucleus is situated near N 240, in a region where evidence of collectivity has been observed in the neutron-rich 24Cr and 25Fe isotopes, with an apparent onset at N 236 and N 238 respectively. The structure of the neutron-rich Mn isotopes is important to determine the onset of deformation in 148 the Z225 isotopic chain, and improve the understanding of the role of the V199 /2 orbital in enhancing deformation. The dependence of single-particle energy shifts on the proton number Z in this region can also be probed through study of the low- energy levels in the Mn isotopes. The low—energy level scheme for 61Mn, deduced following 6 decay of 61Cr, features five new excited states above a l-MeV excitation energy. However, the low-energy structure below 1 MeV did not resemble that from the shell model results that consider the 199/2 orbital. The low density of states below 1 MeV is a consistent feature observed in the odd—A Mn isotopes with A 2 57, 59, 61, and follows the results of shell model calculations restricted to the f p model space. There is also little variation in the yrast structures of these odd-A Mn isotopes up to J 2 15 / 2, as reported by Valiente—Dobon et al. [57]. No compelling evidence was found for the onset of collectivity in the low-energy structures of the Mn isotopes through N 2 37. The only evidence for the onset of collectivity in the Mn isotopes comes at N 239, where a unnatural parity states suggests the approach of the V1 99 /2 orbital close to the Fermi surface. These observations are consistent with the behavior of the even-even, Z +1 26Fe isotopes, but not with that of the even-even Z —1 24Cr isotopes, where possible onset of collectivity at low energy has been suggested from the E(2f) energy at N 2 36. Little data are available for both yrast and non-yrast levels in the 25Mn isotopes beyond N 2 36. Such data will be critical to evaluate the role of the neutron 199/2 orbital in defining the structural properties of the neutron-rich Mn isotopes. 6.2 Outlook 6.2.1 Open questions in the fp shell While the results presented here have provided information on the nuclear structure in and around the f p shell, a number of questions remain. Further experiments are necessary in these regions may serve to provide the information critical to clarify the 149 situation. The present work has provided information on the low-energy level structures of the neutron rich 53:54:568c isotopes in the region surrounding N 234. However, the actual parentage of the low-energy states in 54 Sc and 56Sc needs to be confirmed. Firm assignments for the spin and parity of these low-energy states, and determination of the purity of the single-particle configurations would serve to more fully characterize these nuclei. A measurement of the angular distribution, or internal conversion of the isomeric 110-keV transition in 54Sc would provide the required information to confirm the multipolarity of the transition, and establish the spin and parity of the 110—keV state. Neutron knock-out reactions into both 54Sc and 568C would access the neutron spectroscopic factors for the low-energy states in these nuclei. Spectroscopic factors will be critical in providing the final confirmation for the parentage of the low-energy states in 541568c, and providing the most robust check for shell-model calculations. However, due to the close energy spacing in the low-lying levels of 54,5630, these experiments may benefit from the improved resolution expected using next—generation 'y-ray detector arrays, with sub—segment position resolution. The direct study of the low energy states of neutron-rich 54’56Ca isotopes is also critical, but will likely have to wait for next generation accelerator facilities due to the low production cross-sections for these isotopes. Evidence for the inversion of the 2d3/2 and 1.51/2 proton orbitals in the potas- sium isotopes beyond N 228 is based on indirect measurements. The present data in 50K is suggestive of a 1‘ ground state, arising from a configuration involving a 231/2 proton hole. The tentative assignment suggests that the 71’231/2 orbital remains above the 7rd3/2 past N 228. However, absolute confirmation of the ground state spin could be obtained through a measurement of the ground state hyperfine splitting in 4950K. Measurement of the ground-state g-factors in the K isotopes beyond N 228 would also provide valuable information. Such a measurement using fi-N MR and laser spectroscopy has been proposed at ISOLDE, and will provide information regarding 150 the ground-state configurations in the K isotopes from N 229 to N 232 [126]. Proton transfer reactions would provide the spectroscopic factors necessary to disentangle the proton single-particle contributions to excited states in 49’50K. Knowledge of the parentage of these states would further clarify the relative spacing of the 7r1d3/2 and «231/2 orbitals, and the appropriate strength of the cross-shell proton-neutron tensor interactions. There is evidence of a new region of deformation developing in the 24 Cr, 25Mn and 26Fe isotopes near N 240, as the neutron 199/2 orbital drops in energy with increased occupation of the proton 197/2 orbital. The present results extended the knowledge in the Mn isotopic chain at N 236. Within this region, further fl-decay studies of the Cr isotopes beyond N 236 into the Mn daughters would provide an important extension of the knowledge in this isotopic chain. It appears that collectivity in the Mn isotopes does not occur until N 238, as was the case in the Fe isotopic chain. Determination of the level densities in the odd-A 63’65Mn, which is achievable through 6 decay, could provide the experimental signature for developing collectivity in the Mn isotopes. thure facilities will also prove critical in this region, in pushing spectroscopy of isotopes in the Cr, Mn and Fe chains to larger neutron excess. 6.2.2 The NSCL 6 Counting System with Digital Data Acquisition The NSCL 6 Counting System provides a means for determining critical fl-decay properties needed to test nuclear structure models far from the valley of stability. The work presented in this dissertation is an example of the value both isomeric and fl-delayed 7-ray spectroscopy with the combination of the BCS + SeGA can play in characterizing the evolution of nuclear shell structure as a function of isospin. However, planned improvements to the BCS will serve to expand its capabilities. A digital data acquisition system (DDAS) has been developed at NSCL based 151 on the Pixie-16 Digital Gamma Finder (DGF) modules by XIA LLC [127]. When used with segmented germanium detectors such as SeGA, one of the greatest assets of digital data acquisition is the opportunity for pulse-shape analysis of digitized signals. Pulse shape analysis permits the localization of 'y-ray interaction points to sub-segment resolution, which in turn permits more accurate Doppler corrections, critical for in-beam 7 spectroscopy. DDAS is being adapted for use with the BCS. Preliminary testing of the BCS dual- gain preamplifiers with DDAS has shown that the system is compatible with the fast (~300 ns rise time) output signals of these preamplifiers. With correct settings of the DDAS energy filter parameters, the resolution achieved with DDAS was comparable to, or slightly better than that achieved with the standard BCS analog electronics. Figure 6.1 illustrates the 228Th spectra obtained for a single strip in a 1 mm SSSD using both DDAS and the standard BCS analog electronics. The main advantage in using digital data acquisition for the BCS is the improved detection efficiency. The application of DDAS with the BCS silicon detectors is expected to reduce noise levels. This noise reduction will increase the detection efficiency of the BCS, since more of the low-energy 6 particles will produce true triggers. A gain in efficiency will also be realized due to the zero dead-time option of DDAS — data storage buffers permit readout of the modules with no dead—time losses. In addition to the gains in efficiency, the ability to capture and record waveforms will expand the possibilities for the use of the BCS in charged-particle decay studies, where short-lived (~ as) charged particle decays may be characterized via detailed waveform analysis [128]. 152 3 o ’83 : a") “a 35 - Analog — 2500 4: it: — DDAS : a 712/ 30 L 2000 m E 25 3 § 0 8.78 MeV I 0 U _. U 200 I] I i 1500 150 —§ 1000 ~ A FWHM _ 100 ~ 95 keV : —_ 500 50 J _ o 1 1 1 1 I . ._ - —J1 1 I 1 14 1 l 14_1 1 0 Figure 6.1: 228Th a spectrum collected in a single strip of a 1 mm thick SSSD using DDAS (red line) and the standard analog BCS electronics. The resolution achieved in DDAS is comparable to that obtained with the analog electronics. 153 Bibliography [1] DR. Lide. CRC Handbook of Chemistry and Physics. CRC Press, Inc., Florida, 1996. [2] RB. Firestone. Table of Isotopes, edited by Virginia S. Shirley (Wiley, New York, 1996). [3] KS. Krane. Introductory Nuclear Physics. John Wiley &. Sons, Inc., New York, 1988. [4] W. Loveland, D.J. Morrissey, and GT. Seaborg. Modem Nuclear Chemistry. John Wiley & Sons, Inc., New Jersey, 2006. [5] O. Haxel, J.H.D. Jensen, and HE. Suess. Phys. Rev., 75:1766, 1949. [6] MG. Mayer. Phys. Rev, 7521969, 1949. [7] A. Navin, D.W. Anthony, T. Aumann, T. Baumann, D. Bazin, Y. Blumenfeld, B.A. Brown, T. Glasmacher, P.G. Hansen, R.W. Ibbotson, P.A. Lofy, V. Mad— dalena, K. Miller, T. Nakamura, B.V. Pritychenko, B.M. Sherrill, E. Spears, M. Steiner, J .A. Tostevin, J. Yurkon, and A. Wagner. Phys. Rev. Lett., 852266, 2000. [8] 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. Phys. Rev. Lett., 99:072502, 2007. [9] E.K. Warburton, J.A. Becker, and BA. Brown. Phys. Rev. C, 41:1147, 1990. [10] L. Gaudefroy, O. Sorlin, D. Beaumel, Y. Blumenfeld, Z. Dombradi, S. Fortier, S. Franchoo, M. Gélin, J. Gibelin, S. Grévy, F. Hammache, F. Ibrahim, K.W. Kemper, K.-L. Kratz, S.M. Lukyanov, C. Monrozeau, L. Nalpas, F. Nowacki, A.N. Ostrowski, T. Otsuka, Yu.-E. Penionzhkevich, J. Piekarewicz, E.C. Pollacco, P. Roussel—Chomaz, E. Rich, J.A. Scarpaci, M.G. St. Laurent, D. Sohler, M. Stanoiu, T. Suzuki, E. Tryggestad, and D. Verney. Phys. Rev. Lett., 97:092501, 2006. [11] A. Huck, G. Klotz, A. Knipper, C. Miehé, C. Richard-Serre, G. Walter, A. Poves, H.L. Ravn, and G. Marguier. Phys. Rev. C, 3122226, 1985. 154 [12] A. Gade, R.V.F. Janssens, D. Bazin, R. Broda, B.A. Brown, C.M. Campbell, M.P. Carpenter, J .M. Cook, A.N. Deacon, D.-C. Dinca, B. Fornal, S.J. Freeman, T. Glasmacher, P.G. Hansen, B.P. Kay, P.F. Mantica, W.F. Mueller, J .R. Terry, J .A. Tostevin, and S. Zhu. Phys. Rev. C, 742021302(R), 2006. [13] R.V.F. Janssens, B. Fornal, P.F. Mantica, B.A. Brown, R. Broda, P. Bhat— tacharyya, M.P. Carpenter, M. Cinausero, P.J. Daly, A.D. Davies, T. Glasmacher, Z.W. Grabowski, D.E. Groh, M. Honma, F.G. Kondev, W. Krolas, T. Lauritsen, S.N. Liddick, S. Lunardi, N. Marginean, T. Mizusaki, D.J. Morrissey, A.C. Mor- ton, W.F. Mueller, T. Otsuka, T. Pawlat, D. Seweryniak, H. Schatz, A. Stolz, S.L. Tabor, C.A. Ur, G. Viesti, I. Wiedenhéiver, and J. Wrzesir’lski. Phys. Lett. B, 546255, 2002. [14] J.I. Prisciandaro, P.F. Mantica, B.A. Brown, D.W. Anthony, M.W. Cooper, A. Garcia, D.E. Groh, A. Komives, W. Kumarasiri, P.A. Lofy, A.M. Oros- Peusquens, S.L. Tabor, and M. Wiedeking. Phys. Lett. B, 510217, 2001. [15] P. Federman and S. Pittel. Phys. Lett., 69B2385, 1977. [16] RF. Casten. AIP Conf. Proc., 238267, 1991. [17] T. Otsuka, R. Fugimoto, Y. Utsuno, B.A. Brown, M. Honma, and T. Mizusaki. Phys. Rev. Lett., 872082502, 2001. [18] T. Otsuka, T. Suzuki, R. Fujimoto, H. Grawe, and Y. Akaishi. Phys. Rev. Lett., 952232502. 2005. [19] RF. Casten. Nuclear Structure from a Simple Perspective. Oxford University Press, Inc., New York, 1990. [20] CA. Bertulani and P. Danielewicz. Introduction to Nuclear Reactions. Institute of Physics Publishing, London, 2004. ' [21] V. Paar. Nucl. Phys, A331216, 1979. [22] V.L. Alexeev, B.A. Emelianov, A.I. Egorov, L.P. Kabina, D.M. Kaminker, Yu.L. Khazov, I.A. Kondurov, E.K. Leushkin, Yu.E. Loginov, V.V. Martynov, V.L. Rumiantsev, S.L. Sakharov, P.A. Sushkov, H.G. B6rner, W.F. Davidson, J .A. Pinston, and K. Schreckenbach. Nucl. Phys, A297 2373, 1978. [23] T.W. Burrows. Nucl. Data Sheets, 10721747, 2006. [24] RF. Casten, J. Zhang, N.V. Zamfir, M.A. Caprio, H. Zhang, and F. Du. Phys. Lett. B, 481:1, 2000. [25] S. Raman, C.H. Malarkey, W.T. Milner, C.W. Nestor Jr., and RH. Stelson. Atom. Data and Nucl. Data Tables, 3621, 1987. [26] L. Grodzins. Phys. Lett., 2:88, 1962. 155 [27] F .S. Stephens, R.M. Diamond, J.R. Leigh, T. Kammuri, and K. Nakai. Phys. Rev. Lett., 292438, 1972. [28] D.-C. Dinca, R.V.F. Janssens, A. Gade, D. Bazin, R. Broda, B.A. Brown, C.M. Campbell, M.P. Carpenter, P. Chowdhury, J .M. Cook, A.N. Deacon, B. For— nal, S.J. Freeman, T. Glasmacher, M. Honma, F.G. Kondev, J.-L. Lecouey, S.N. Liddick, P.F. Mantica, W.F. Mueller, H. Olliver, T. Otsuka, J.R. Terry, B.A. Tomlin, and K. Yoneda. Phys. Rev. C, 712041302(R), 2005. [29] N.K. Glendenning. Direct Nuclear Reactions. World Scientific Publishing Co. Pte. Ltd., Singapore, 2004. [30] I.J. Thompson and F .M. Nunes. Nuclear Reactions for Astrophysics. Cambridge University Press, New York, 2009. [31] SN. Liddick, P.F. Mantica, R. Broda, B.A. Brown, M.P. Carpenter, A.D. Davies, B. Fornal, T. Glasmacher, D.E. Groh, M. Honma, M. Horoi, R.V.F. Janssens, T. Mizusaki, D.J. Morrissey, A.C. Morton, W.F. Mueller, T. Otsuka, J. Pavan, H. Schatz, A. Stolz, S.L. Tabor, B.E. Tomlin, and M. Wiedeking. Phys. Rev. C, 702064303, 2004. [32] A. Biirger, T.R. Saito, H. Grawe, H. Hiibel, P. Reiter, J. Gerl, M. Gorska, H.J. Wollersheim, A. Al-Khatib, A. Banu, T. Beck, F. Becker, P. Bednar- czyk, G. Benzoni, A. Bracco, S. Brambilla, P. Bringel, F. Camera, E. Clement, P. Doornenbal, H. Geissel, A. Gérgen, J. Grebosz, G. Hammond, M. Hellstriim, M. Honma, M. Kavatsyuk, O. Kavatsyuk, M. Kmiecik, I. Kojouharov, W. Korten, N. Kurz, R. Lozeva, A. Maj, S. Mandal, B. Million, S. Muralithar, A. Neusser, F. Nowacki, T. Otsuka, Zs. Podolyak, N. Saito, A.K. Singh, H. Weick, C. Whel— don, O. Wieland, M. Winkler, and the RISING Collaboration. Phys. Lett. B, 622229, 2005. . [33] M. Honma, T. Otsuka, B.A. Brown and T. Mizusaki. Phys. Rev. C, 65:061301(R), 2002. [34] M. Honma, T. Otsuka, B.A. Brown, and T. Mizusaki. Eur. Phys. J. A, 252499, 2005. [35] A. Poves, J. Sa’lnchez-Solano, E. Caurier, and F. Nowacki. Nucl. Phys. A, 6942157, 2001. [36] SN. Liddick, P.F. Mantica, R.V.F. Janssens, R. Broda, B.A. Brown, M.P. Car- penter, B. Fornal, M. Honma, T. Mizusaki, A.C. Morton, W.F. Mueller, T. Ot— suka, J. Pavan, A. Stolz, S.L. Tabor, B.E. Tomlin, and M. Wiedeking. Phys. Rev. Lett., 922072502, 2004. [37] SM. Fischer, C.J. Lister, D.P. Balamuth, R. Bauer, J .A. Becker, L.A. Bernstein, M.P. Carpenter, J. Durell, N. Fotiades, S.J. Freeman, P.E. Garrett, P.A. Haus- laden, R.V.F. Janssens, D. Jenkins, M. Leddy, J. Ressler, J. Schwartz, D. Svel- nys, D.G. Sarantites, D. Seweryniak, B.J. Varley, and R. Wyss. Phys. Rev. Lett., ' 872132501, 2001. 156 [38] [39] C.J. Lister, M. Campbell, A.A. Chishti, W. Gelletly, L. Goettig, R. Moscrop, B.J. Varley, A.N. James, T. Morrison, H.G. Price, J. Simpson, K. Connel, and O. Skeppstedt. Phys. Rev. Lett., 5921270, 1987. OJ. Lister, P.J. Ennis, A.A. Chishti, B.J. Varley, W. Gelletly, H.G. Price, and AN. James. Phys. Rev. C, 422R1191, 1990. [40] B. Blank and RH. Regan. Nucl. Phys. News, 10220, 2000. I41] I42] [43] [44] I45] I46] [47] Z. Janas, C. Chandler, B. Blank, P.H. Regan, A.M. Bruce, W.N. Catford, N. Curtis, S. Czajkowski, Ph. Dessagne, A. Fleury, W. Gelletly, J. Giov- inazzo, R. Grzywacz, M. Lewitowicz, C. Longour, C. Marchand, Ch. Miehé, N.A. Orr, R.D. Page, C.J. Pearson, M.S. Pravikoff, A.T. Reed, M.G. Saint- Laurent, J .A. Sheikh, S.M. Vincent, R. Wadsworth, D.D. Warner, and J .8. Win- field. Phys. Rev. Lett., 822295, 1999. M. Hannawald, T. Kautzsch, A. W6hr, W.B. Walters, K.-L. Kratz, V.N. Fe— doseyev, V.I. Mishin, W. B6hmer, B. Pfeiffer, V. Sebastian, Y. Jading, U. Kéister, J. Lettry, H.L. Ravn, and the ISOLDE Collaboration. Phys. Rev. Lett., 8221391, 1999. P. Adrich, A.M. Amthor, D. Bazin, M.D. Bowen, B.A. Brown, C.M. Camp— bell, J .M. Cook, A. Gade, D. Galaviz, T. Glasmacher, S. McDaniel, D. Miller, A. Obertelli, Y. Shimbara, K.P. Siwek, J .A. Tostevin, and D. Weisshaar. Phys. Rev. C, 772054306, 2008. O. Sorlin, C. Donzaud, F. Nowacki, J.C. Angelique, F. Azaiez, C. Bourgeois, V. Chiste, Z. Dlouhy, S. Grévy, D. Guillemaud—Mueller, F. Ibrahim, K.—L. Kratz, M. Lewitowicz, S.M. Lukyanov, J. Mrasek, Yu.-E. Penionzhkevich, F. de Oliveira Santos, B. Pfeiffer, F. Pougheon, A. Poves, M.G. Saint-Laurent, and M. Stanoiu. Eur. Phys. J. A, 16:55, 2003. R. Broda, B. Fornal, W. Krolas, T. Pawlat, D. Bazzacco, S. Lunardi, C. Rossi- Alvarez, R. Menegazzo, G. de Angelis, P. Bednarczyk, J. Rico, D. De Acflna, P.J. Daly, R.H. Mayer, M. Sferrazza, H. Grawe, K.H. Meier, and R. Schubart. Phys. Rev. Lett, 742868, 1995. O. Sorlin, S. Leenhardt, C. Donzaud, J. Duprat, F. Azaiez, F. Nowacki, H. Grawe, Zs. Dombradi, F. Amorini, A. Astier, D. Baiborodin, M. Belleguic, C. Borcea, C. Bourgeois, D.M. Cullen, Z. Dlouhy, E. Dragulescu, M. Gorska, S. Grévy, D. Guillemaud—Mueller, G. Hagemann, B. Herskind, J. Kiener, R. Lemmon, M. Lewitowicz, S.M. Lukyanov, P. Mayet, F. de Oliveira Santos, D. Pantalica, Yu.-E. Penionzhkevich, F. Pougheon, A. Poves, N. Redon, M.G. Saint-Laurent, J .A. Scarpaci, G. Sletten, M. Stanoiu, O. Tarasov, and Ch. Theisen. Phys. Rev. Lett, 882092501, 2002. N. Bree, I. Stefanescu, P.A. Butler, J. Cederkiill, T. Davinson, P. Delahaye, J. Eberth, D. Fedorov, V.N. Fedosseev, L.M. Fraile, S. Franchoo, G. Georgiev, K. Gladnishki, M. Huyse, O. Ivanov, J. Iwanichi, J. Jolie, U. K6ster, Th. Kréll, 157 I48] [49] [50] [51] [52] [53] I54] [55] R. Kriicken, B.A. Marsh, 0. Niedermaier, P. Reiter, H. Scheit, D. Schwalm, T. Sieber, J. Van de Walle, P. Van Duppen, N. Warr, D. Weisshaar, F. Wenan- der and S. Zemlyanoy. Phys. Rev. C, 782047301, 2008. S. Leenhardt, O. Sorlin, M.G. Porquet, F. Azaiez, J .C. Angelique, M. Belleguic, C. Borcea, C. Bourgeois, J.M. Daugas, C. Donzaud, I. Deloncle, J. Duprat, A. Gillibert, S. Grévy, D. Guillemaud—Mueller, J. Kiener, M. Lewitowicz, S.M. Lukyanov, F. Marie, N.A. Orr, Yu.—E. Penionzhkevich, F. de Oliveira San- tos, F. Pougheon, M.G. Saint-Laurent, W. Shuying, Yu. Sobolev, and J .8. Win- field. Eur. Phys. J. A, 1421, 2002. N. Aoi, H. Suzuki, E. Takeshita, S. Takeuchi, S. Ota, H. Baba, S. Bishop, T. Fukui, Y. Hashimoto, H.J. Ong, E. Ideguchi, K. Ieki, N. Imai, H. Iwaski, S. Kanno, Y. Kondo, T. Kubo, K. Kurita, K. Kusaka, T. Minemura, T. Moto- bayashi, T. Nakabayashi, T. Nakamura, T. Nakao, M. Niikura, T. Okumura, T.K. Ohniski, H. Sakurai, S. Shimoura, R. Sugo, D. Suzuki, M.K. Suzuki, M. Tamaki, K. Tanaka, Y. Togano, and K. Yamada. Nucl. Phys, A8052400c, 2008. N. Aoi, E. Takeshita, H. Suzuki, S. Takeuchi, S. Ota, H. Baba, S. Bishop, T. Fukui, Y. Hashimoto, H.J. Ong, E. Ideguchi, K. Ieki, N. Imai, M. Ishihara, H. Iwasaki, S. Kanno, Y. Kondo, T. Kubo, K. Kurita, K. Kusaka, T. Minemura, T. Motobayashi, T. Nakabayashi, T. Nakamura, T. Nakao, M. Niikura, T. Oku- mura, T.K. Ohnishi, H. Sakurai, S. Shimoura, R. Sugo, D. Suzuki, M.K. Suzuki, M. Tamaki, K. Tanaka, Y. Togano, and K. Yamada. Phys. Rev. Lett., 1022012502, 2009. K. Kaneko, Y. Sun, M. Hasegawa, and T. Mizusaki. Phys. Rev. C, 782064312, 2008. L. Gaudefroy, A. Obertelli, S. Péru, N. Pillet, S. Hilaire, J.-P. Delaroche, M. Girod, and J. Libert. Phys. Rev. C, 802064313, 2009. DE. Appelbe, C.J. Barton, M.H. Muikku, J. Simpson, D.D. Warner, C.W. Beau- sang, M.A. Caprio, J.R. Cooper, J.R. Novak, N.V. Zamfir, R.A.E. Austin, J .A. Cameron, C. Malcolmson, J.C. Waddington, and F .R. Xu. Phys. Rev. C, 672034309, 2003. AN. Deacon, S.J. Freeman, R.V.F. Janssens, F.R. Xu, M.P. Carpenter, I.R. Calderin, P. Chowdhury, N .J . Hammond, T. Lauritsen, C.J. Lister, D. Sew- eryniak, J .F. Smith, S.L. Tabor, B.J. Varley, and S. Zhu. Phys. Lett. B, 6222151, 2005. S.J. Heeman, R.V.F. Janssens, B.A. Brown, M.P. Carpenter, S.M. Fischer, N.J. Hammond, M. Honma, T. Lauritsen, C.J. Lister, T.L. Khoo, G. Mukherjee, D. Seweryniak, J .F. Smith, B.J. Varley, M. Whitehead, and S. Zhu. Phys. Rev. C, 692064301, 2004. 158 [56] D. Steppenback, A.N. Deacon, S.J. Freeman, R.V.F. Janssens, S. Zhu, M.P. Car- penter, P. Chowdhury, M. Honma, T. Lauritsen, C.J. Lister, D. Seweryniak, J.F. Smith, S.L. Tabor, and B.J. Varley. Phys. Rev. C, 812014305, 2010. [57] J.J. Valiente-Dobon, S.M. Lenzi, S.J. Freeman, S. Lunardi, J .F . Smith, A. Got— tardo, F. Della Vedova, E. Farnea, A. Gadea, D.R. Napoli, M. Axiotis, S. Ay— din, D. Bazzacco, P.G. Bizzeti, A.M. Bizzeti-Sona, G. Benzoni, D. Bucurescu, L. Corradi, A.N. Deacon, G. de Angelis, E. Fioretto, B. Guiot, M. Ionescu-Bujor, A. Iordachescu, S. Leoni, N. Marginean, R. Marginean, P. Mason, R. Menegazzo, D. Mengoni, B. Million, G. Montagnoli, R. Orlandi, F. Recchia, E. Sahin, F. Scar- lassara, RP. Singh, A.M. Stefanini, D. Steppenbeck, S. Szilner, C.A. Ur, B.J. Var- ley, and O. Wieland. Phys. Rev. C, 782024302, 2008. [58] L. Gaudefroy, O. Sorlin, C. Donzaud, J .C. Angelique, F. Azaiez, C. Bourgeois, V. Chiste, Z. Dlouhy, S. Grévy, D. Guillemaud—Mueller, F. Ibrahim, K.-L. Kratz, M. Lewitowicz, S.M. Lukyanov, I. Matea, J. Mrasek, F. Nowacki, F. de Oliveira— Santos, Yu.-E. Penionzhkevich, B. Pfeiffer, F. Pougheon, M.G. Saint-Laurent, and M. Stanoiu. Eur. Phys. J. A, 23:41, 2005. [59] O. Sorlin, C. Donzaud, L. Axelsson, M. Belleguic, R. Béraud, C. Borcea, G. Can- chel, E. Chabanat, J .M. Daugas, A. Emsallem, M. Girod, D. Guillemaud- Mueller, K.-L. Kratz, S. Leenhardt, M. Lewitowicz, C. Longour, M.J. Lopez, F. de Oliveira Santos, L. Petizon, B. Pfeiffer, F. Pougheon, M.G. Saint-Laurent, and J .E. Sauvestre. Nucl. Phys, A6692351, 2000. [60] G. Friedlander, J .W. Kennedy, E.S. Macias, and J.M. Miller. Nuclear and Ro- diochemistry, 3rd Edition. John Wiley & Sons, Inc., New York, 1981. [61] T. Kibédi, T.W. Burrows, M.B. Trzhaskovskaya, P.M. Davidson, and CW. Nestor Jr.. Nucl. Instrum. Methods Phys. Res. A, 5892202, 2008.; Internal ' conversion calculator available online at: http://physics.anu.edu/nuclear/bricc/. [62] D.J. Morrissey, B.M. Sherrill, M. Steiner, A. Stolz, and I. Wiedenhéiver. Nucl. Instrum. Methods Phys. Res. B, 204290, 2003. [63] A. Stolz, T. Baumann, T.N. Ginter, D.J. Morrissey, M. Portillo, B.M. Sherrill, M. Steiner, and J.W. Stetson. Nucl. Instrum. Methods Phys. Res. B, 2412858, 2005. [64] J .I. Prisciandaro, A.C. Morton, and RF. Mantica. Nucl. Instrum. Methods Phys. Res. A, 5052140, 2003. [65] J.B. Stoker. 6-Decay Half—life of the rp-Process Waiting—Point Nuclide 84Mo Ph.D. thesis, Michigan State University, 2009. [66] W.F. Mueller, J.A. Church, T. Glasmacher, D. Gutknecht, G. Hackman, P.G. Hansen, Z. Hu, K.L. Miller, and P. Quirin. Nucl. Instrum. Methods Phys. Res. A, 4662492, 2001. 159 [67] SN. Liddick. Beta-Decay Studies of Neutron—Rich Nuclides and the Possibility of an N234 Subshell Closure Ph.D. thesis, Michigan State University, 2004. [68] R. Grzywacz, R. Béraud, C. Borcea, A. Emsallem, M. Glogowski, H. Grawe, D. Guillemaud-Mueller, M. Hjorth-Jensen, M. Houry, M. Lewitowicz, A.C. Mueller, A. Nowak, A. Plochocki, M. Pfiitzner, K. Rykaczewski, M.G. Saint- Laurent, J .E. Sauvestre, M. Schaefer, O. Sorlin, J. Szerypo, W. Trinder, S. Vi- teritti, and J. Winfield. Phys. Rev. Lett., 812766, 1998. [69] SN. Liddick, P.F. Mantica, R. Broda, B.A. Brown, M.P. Carpenter, A.D. Davies, B. Fornal, M. Horoi, R.V.F. Janssens, A.C. Morton, W.F. Mueller, J. Pavan, H. Schatz, A. Stolz, S.L. Tabor, B.E. Tomlin, and M. Wiedeking. Phys. Rev. C, 722054321, 2005. [70] RF. Mantica, A.C. Morton, B.A. Brown, A.D. Davies, T. Glasmacher, D.E. Groh, S.N. Liddick, D.J. Morrissey, W.F. Mueller, H. Schatz, A. Stolz, S.L. Tabor, M. Honma, M. Horoi, and T. Otsuka. Phys. Rev. C, 672014311, 2003. [71] PF. Mantica, B.A. Brown, A.D. Davies, T. Glasmacher, D.E. Groh, M. Horoi, S.N. Liddick, D.J. Morrissey, A.C. Morton, W.F. Mueller, H. Schatz, A. Stolz, and S.L. Tabor. Phys. Rev. C, 682044311, 2003. [72] RR. Bevington, and D.K. Robinson Data Reduction and Error Analysis for the Physical Sciences 2nd ed.. McGraw-Hill, Boston, 1992. [73] G.F. Grinyer. High Precision Measurements of 26Na 6‘ Decay Masters thesis, The University of Guelph, 2004. [74] R. Brun and F. Rademakers. ROOT: A Data Analysis Framework. Available online at: http://root.cern.ch. [75] S. McDaniel, A. Gade, R.V.F. Janssens, D. Bazin, B.A. Brown, C.M. Camp- bell, M.P. Carpenter, J.M. Cook, A.N. Deacon, D.-C. Dinca, S.J. Freeman, T. Glasmacher, P.G. Hansen, B.P. Kay, P.F. Mantica, W.F. Mueller, J .R. Terry, J .A. Tostevin, and S. Zhu. Phys. Rev. C, 812024301, 2010. [76] RF. Mantica, R. Broda, H.L. Crawford, A. Damaske, B. Fornal, A.A. Hecht, C. Hoffman, M. Horoi, N. Hoteling, R.V.F. Janssens, J. Pereira, J.S. Pinter, J.B. Stoker, S.L. Tabor, T. Sumikama, W.B. Walters, X. Wang, and S. Zhu. Phys. Rev. C, 772014313, 2008. [77] M. Langevin, C. Détraz, D. Guillemaud-Mueller, A.C. Mueller, C. Thibault, F. Touchard, G. Klotz, C. Miehé, G. Walter, M. Epherre, and C. Richard-Serre. Phys. Lett., 130B2251, 1983. [78] L. Coraggio, A. Covello, A. Gargano, and N. Itaco. Phys. Rev. C, 802044311, 2009. 160 [79] [80] [81] [82] I83] [84] [85] [86] [87] National Nuclear Data Center LOGF T web program. Available online at http: / / www.nndc.bnl. gov / logft / . G. Audi, A.H. Wapstra, and C. T hibault. Nucl. Phys, A7292337, 2003. B. Fornal, S. Zhu, R.V.F. Janssens, M. Honma, R. Broda, P.F. Mantica, B.A. Brown, M.P. Carpenter, P.J. Daly, S.J. Freeman, Z.W. Grabowski, N.J. Hammond, F.G. Kondev, W. Krolas, T. Lauritsen, S.N. Liddick, C.J. Lister, E.F. Moore, T. Otsuka, T. Pawlat, D. Seweryniak, B.E. Tomlin, and J. Wrze- sir’lski. Phys. Rev. 0, 702064304, 2004. T. D6rfler, W.-D. Schmidt-Ott, T. Hild, T. Mehren, W. B6hmer, P. Méiller, B. Pfeiffer, T. Rauscher, K.-L. Kratz, O. Sorlin, V. Borrel, S. Grévy, D. Guillemaud-Mueller, A.C. Mueller, F. Pougheon, R. Anne, M. Lewitowicz, A. Ostrowsky, M. Robinson, and M.G. Saint-Laurent. Phys. Rev. C, 5422894, 1996. S. Zhu, R.V.F. Janssens, B. Fornal, S.J. Freeman, M. Honma, R. Broda, M.P. Carpenter, A.N. Deacon, B.P. Kay, F.G. Kondev, W. Krélas, J. Kozemczak, A. Larabee, T. Lauritsen, S.N. Liddick, C.J. Lister, P.F. Mantica, T. Otsuka, T. Pawlat, A. Robinson, D. Seweryniak, J .F. Smith, D. Steppenbeck, B.E. Tom- lin, J. Wrzesil’lski, and X. Wang. Phys. Lett. B, 6502135, 2007. H.L. Crawford, R.V.F. Janssens, P.F. Mantica, J.S. Berryman, R. Broda, M.P. Carpenter, N. Cieplicka, B. Fornal, G.F. Grinyer, N. Hoteling, B.P. Kay, T. Lauritsen, K. Minamisono, I. Stefanescu, J.B. Stoker, W.B. Walters, and S. Zhu. T o be published. R. Broda, W.Kr(’)Ias, B. Fornal, A. Gadea, N. Marginean, L. Corradi, A.M. Ste- fanini, J. Wrzesir’lski, T. Pawlat, S. Beghini, G. De Angelis, M. Carpenter, F. Della Vedova, E. Farnea, E. Fioretto, B. Guiot, R.V.F. Janssens,. S. Lunardi, P. Mantica, P. Mason, G. Montagnoli, D.R. Napoli, R. Orlandi, I. Pokrovskiy, E. Sahin, F. Scarlassara, S. Szilner, C.A. Ur, M. Trotta, J .J . Valiente-Dobon, and S. Zhu. Annual Report 2006 INFN-LNL-217, ISSN 1828-8545 (Legnaro, 2007). R. Broda, J. Wrzesir’lski, A. Gadea, N. Marginean, B. Fornal, L. Corradi, A. Ste- fanini, W. Krolas, T. Pawlat, B. Szpak, S. Lunardi, J .J . Valiente-Dobon, D. Men- goni, E. Farnea, S. Beghini, M.P. Carpenter, G. De Angelis, F. Della Vedova, E. Fioretto, B. Guiot, R.V.F. Janssens, P.F. Mantica, P. Mason, G. Montag- noli, D.R. Napoli, R. Orlandi, I. Pokrovskiy, E. Sahin, F. Scarlassara, S. Szilner, C.A. Ur, M. Trotta, and S. Zhu. To be published. M. Lewitowicz, J .M. Daugas, R. Grzywacz, L. Achouri, J .C. Angelique, D. Bai- borodin, R. Bentida, R. Béraud, C. Bingham, C. Borcea, W. Catford, A. Em— sallem, G. de France, M. Glogowski, H. Grawe, D. Guillemaud-Mueller, M. Houry, S. Hurskanen, K.L. Jones, R.C. Lemmon, A.C. Mueller, A. Nowak, F. de Oliveira- Santos, A. Plochocki, M. Pfiitzner, P.H. Regan, K. Rykaczewski, M.G. Saint- Laurent, J .E. Sauvestre, M. Sawicka, M. Schaefer, G. Sletten, O. Sorlin, 161 M. Stanoiu, J. Szerypo, W. Trinder, S. Viteritti, and J. Winfield. Nucl. Phys, A6542687c, 1999. [88] J .M. Daugas. Ph.D. thesis, Caen University, 1999. [89] MR. Bhat. Nucl. Data Sheets, 882417, 1999. [90] F. Ameil, M. Bernas, P. Armbruster, S. Czajkowski, Ph. Dessagne, H. Geissel, E. Hanelt, C. Kozhuharov, C. Miehe, C. Donzaud, A. Grewe, A. Heinz, Z. Janas, M. de Jong, W. Schwab, and S. Steinhauser. Eur. Phys. J. A, 12275, 1998. [91] E. Runte, K.—L. Gippert, W.-D. Schmidt-Ott, P. Tidemand—Petersson, L. Ziegeler, R. Kirchner, O. Klepper, P.O. Larsson, E. Roeckl, D. Schardt, N. Kaf- frell, P. Peuser, M. Bernas, Ph. Dessagne, M. Langevin, and K. Rykaczewski. Nucl. Phys, A4412237, 1985. [92] J. Bron, H.W. Jongsma, and H. Verheul. Phys. Rev. C, 112966, 1975. [93] S. Bhattacharyya, M. Rejmund, A. Navin, E. Caurier, F. Nowacki, A. Poves, R. Chapman, D. O’Donnell, M. Gelin, A. Hodsdon, X. Liang, W. Mittig, G. Mukherjee, F. Rejmund, M. Rousseau, P. Roussel-Chomaz, K.-M. Spohr, and Ch. Theisen. Phys. Rev. C, 792014313, 2009. [94] B.A. Brown, private communication. [95] DE. Alburger, E.K. Warburton, and B.A. Brown. Phys. Rev. C, 3021005, 1984. [96] U. F ister, R. Jahn, P. von Neumann-Cosel, P. Schenk, T.K. 'Ifelle, D. Wenzel, and U. Wienands. Nucl. Phys, A5692421, 1994. [97] F. Ajzenberg—Selove, R.E. Brown, E.R. Flynn, and J.W. Sunier. Phys. Rev. C, 322756, 1985. [98] A. Gade, R.V.F. Janssens, D. Bazin, B.A. Brown, C.M. Campbell, M.P. Car— penter, J.M. Cook, A.N. Deacon, D.—C. Dinca, S.J. Freeman, T. Glasmacher, B.P. Kay, P.F. Mantica, W.F. Mueller, J.R. Terry, and S. Zhu. Phys. Rev. C, 732037309, 2006. [99] B. Fornal, R.V.F. Janssens, R. Broda, N. Marginean, S. Beghini, L. Cor- radi, M.P. Carpenter, G. De Angelis, F. Della Vedova, E. Farnea, E. Fioretto, A. Gadea, B. Guiot, M. Honma, W. Krolas, T. Lauritsen, S. Lunardi, P.F. Man— tica, P. Mason, G. Montagnoli, D.R. Napoli, T. Otsuka, T. Pawlat, G. Pollarolo, F. Scarlassara, A.M. Stefanini, D. Seweryniak, S. Szilner, C.A. Ur, M. 'Ii‘otta, J .J . Valiente-Dobon, J. Wrzesil’lski, and S. Zhu. Phys. Rev. C, 772014304, 2008. [100] SP. Pandya. Phys. Rev, 1032956, 1956. [101] F. Nowacki and A. Poves. Phys. Rev. C, 792014310, 2009. [102] Y. Utsuno, T. Otsuka, B.A. Brown, M. Honma, and T. Mizusaki. AIP Conf. Proc., 1120281, 2009. 162 [103] Y. Utsuno, T. Otsuka, T. Mizusaki, and M. Honma. Eur. Phys. J. Special Topics, 1502187, 2007. [104] S. Hinds and R. Middleton. Nucl. Phys, 842651, 1966. [105] F. Touchard, P. Guimbal, S Biittgenback, R. Klapisch, M. De Saint Simon, J .M. Serre, C. Thibault, H.T. Duong, P. Juncar, S. Liberman, J. Pinard, and J.L. Vialle. Phys. Lett., 108B2169, 1982. [106] T. Otsuka, T. Suzuki, M. Honma, Y. Utsuno, N. Tsunoda, K. Tsukiyama, and M. Hjorth—Jensen. Phys. Rev. Lett., 1042012501, 2010. [107] A. Gade, B.A. Brown, D. Bazin, C.M. Campbell, J.A. Church, D.G. Dinca, J. Enders, T. Glasmacher, M. Horoi, Z. Hu, K.W. Kemper, W.F. Mueller, T. Otsuka, L.A. Riley, B.T. Roeder, T. Suzuki, J .R. Terry, K.L. Yurkewicz, and H. Zwahlen. Phys. Rev. C, 742034322, 2006. [108] T. Ishii, M. Asai, A. Makishima, M. Matsuda, S. Ichikawa, T. Kohno, and M. Ogawa. CNS-REP, 64:27, 2004. [109] LG. Multhauf, K.G. Tirsell, S. Raman, and J .B. McGrory. Phys. Lett., 578244, 1975. [110] E.K. Warburton. Phys. Rev. C, 4421024, 1991. [111] O. Sorlin, D. Guillemaud—Mueller, R. Anne, L. Axelsson, D. Bazin, W. Bdhmer, V. Borrel, Y. Jading, H. Keller, K.-L. Kratz, M. Lewitowica, S.M. Lukyanov, T. Mehren, A.C. Mueller, Yu.E. Penionzhkevich, F. Pougheon, M.G. Saint- Laurent, V.S. Salamatin, S. Shoedder, and A. W6hr. Nucl. Phys. A, 5832763, 1995. [112] S. Grévy, J .C. Angelique, P. Baumann, C. Borcea, A. Buta, G. Canchel, W.N. Catford, S. Courtin, J .M. Daugas, F. de Oliveira, P. Dessagne, Z. Dlouhy, A. Knipper, K.L. Kratz, F.R. Lecolley, J.L. Lecouey, G. Lehrsenneau, M. Le- witowicz, E. Liénard, S. Lukyanov, F. Maréchal, C. Miehé, J. Mrazek, F. Ne- goita, N.A. Orr, D. Pantelica, Y. Penionzhkevich, J. Péter, B. Pfeiffer, S. Pietri, E. Poirier, O. Sorlin, M. Stanoiu, I. Stefan, C. Stodel, and C. Timis. Phys. Lett. B, 5942252, 2004. [113] B.A. Brown and EH. Wildenthal. Ann. Rev. Nucl. Part. Sci., 38:29, 1988. [114] T.T.S. Kuo and GE. Brown. Nucl. Phys, A1142241, 1968. [115] S. Nummela, P. Baumann, E. Caurier, P. Dessagne, A. Jokinen, A. Knipper, G. Le Scornet, C. Miehé, F. Nowacki, M. Oinonen, Z. Radivojevic, M. Ramdhane, G. Walter, J. Ayst6, and the ISOLDE Collaboration. Phys. Rev. C, 632044316, 2001.] [116] ON. Davids, D.F. Geesaman, S.L. Tabor, M.J. Murphy, E.B. Norman, and RC. Pardo. Phys. Rev. C, 1721815, 1978. 163 [117] DE. Appelbe, R.A.E. Austin, G.C. Ball, J.A. Cameron, B. Djerroud, T.E. Drake, S. Flibotte, C.D. O’Leary, A. Melarangi, C.E. Svensson, J.C. Waddington, and D. Ward. Eur. Phys. J. A, 82153, 2000. [118] N.G. Puttaswamy, W. Oelert, A. Djaloeis, C. Mayer-Béricke, P. Turek, P.W.M. Glaudemans, B.C. Metsch, K. Heyde, M. Waroquier, P. van Isaker, G. Wenes, V. Lopac, and V. Paar. Nucl. Phys, A4012269, 1983. [119] M. Oinonen, U. K6ster, J. Aysté, V. Fedoseyev, V. Mishin, J. Huikari, A. Joki- nen, A. Nieminen, K. Perajarvi, A. Knipper, G. Walter, and the ISOLDE Col- laboration. Eur. Phys. J. A, 102123, 2001. [120] E. Caurier, Code ANTOINE, Strasbourg, 1989; E. Caurier and F. Nowacki. Acta Phys. POL, 302705, 1999. [121] SN. Liddick, P.F. Mantica, B.A. Brown, M.P. Carpenter, A.D. Davies, M. Horoi, R.V.F. Janssens, A.C. Morton, W.F. Mueller, J. Pavan, H. Schatz, A. Stolz, S.L. Tabor, B.E. Tomlin, and M. Wiedeking. Phys. Rev. C, 732044322, 2006. [122] L. Gaudefroy. Study of the N 228 shell closure in the nucleus ‘llgArgg with neu- tron transfer reaction: applications to astrophysics and 67 spectroscopy of neu- tron rich nuclei around N 232/34 and N 240 Ph.D. thesis, De L’Universite Paris XI Orsay, 2005. [123] A. Estradé. Time-of—flight mass measurements of neutron rich isotopes at the NSCL Ph.D. thesis, Michigan State University, 2010. [124] B.A. Brown. Prog. Part. Nucl. Physics, 472517, 2001. [125] 0B. Tarasov, D.J. Morrissey, A.M. Amthor, T. Baumann, D. Bazin, A. Gade, T.N. Ginter, M. Hausmann, N. Inabe, T. Kubo, A. Nettleton, J. Pereira, M. Por- tillo, B.M. Sherrill, A. Stolz, and M. Thoennessen. Phys. Rev. Lett., 1022142501, 2009. [126] G. Neyens, M. Bissell, M De Rydt, P. Lievens, C. Geppert, J. Krae— mer, A. Krieger, and R. Neugart. Ground-state properties of K- isotopes from laser and 6-NMR spectroscopy. Available online at http: / / greybook.cern.ch / programmes / experiments / IS484.html. [127] K. Starosta, C. Vaman, D. Miller, P. Voss, D. Bazin, T. Glasmacher, H. Craw- ford, P. Mantica, H. Tan, W. Hennig, M. Walby, A. Fallu—Labruyere, J. Harris, D. Breus, P. Grudberg, and WK. Warburton. Nucl. Instrum. Methods Phys. Res. A, 6102700, 2009. [128] R. Grzywacz. Nucl. Instrum. Methods Phys. Res. B, 2042649, 2003. 164