THE (P,N) CHARGE-EXCHANGE REACTION IN INVERSE KINEMATICS AS A PROBE FOR ISOVECTOR GIANT RESONANCES IN EXOTIC NUCLEI By Samuel Israel Lipschutz A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Physics—Doctor of Philosophy 2018 ABSTRACT THE (P,N) CHARGE-EXCHANGE REACTION IN INVERSE KINEMATICS AS A PROBE FOR ISOVECTOR GIANT RESONANCES IN EXOTIC NUCLEI By Samuel Israel Lipschutz Charge-exchange experiments at intermediate (∼100 MeV/u) beam energies have long been a tool to study isovector giant resonances in nuclei. Since the 1970s, many probes and techniques have been developed to study and isolate different isovector giant resonances, which have illuminated the spin-isospin response of nuclei. However, these experiments have almost solely been restricted to stable nuclei. The advent of rare-isotope beam facilities has created increased interest in studying these resonances in radioactive nuclei. Such experiments are important for better constraining models that aim to describe the properties of nuclei and nuclear matter, with important applications in astro and neutrino physics. To take advantage of these rare-isotope facilities, new techniques in inverse kinematics need to be developed and validated. This thesis presents a new technique for studying isovector giant resonances on unstable nuclei in the ∆Tz = −1 direction, through the (p,n) reaction in inverse kinematics. The technique was validated through the measurement of the absolute differential cross section in the neutron-rich 16 C(p,n)16 N* reaction at 100 MeV/u up to ∼20 MeV of excitation energy. From the data, the Gamow-Teller (GT) strength distribution [B(GT)] and the spin-dipole (SD) differential cross section were extracted. The extracted B(GT) was compared with shell-model calculations in order to test their reliability in the neutron-rich, high-excitation energy regime. A total GT strength up to 20 MeV was extracted and compared to the model-independent sum rule, giving a quenching factor of 73.5 ±2.8(st) ± 15.5(sys)%. This experiment provided one of the first measurements of isovector strength in a neutron-rich nucleus and has paved the way for future exploration of isovector giant resonances in exotic nuclei. This thesis is dedicated to you, the reader, so that you may feel special. iv ACKNOWLEDGMENTS No PhD is completed in isolation, and certainly not this one. Without the help and guidance of countless people, my research could not have been completed, and this thesis could not have been written. Though I cannot thank everyone by name, know that I am grateful for all the help you have given me and for what it has allowed me to accomplish. I would like to give a special and wholehearted thanks to my advisor, Remco Zegers. Even in the depths of the great “factor of four” debate of 2017, you continued to provide the same excellent advice and mentorship that you have given me over the past six years. Though our opinions on Dutch licorice sharply diverge, it has not impacted your tireless commitment to see me through to the end. I would also like to thank the members of my committee for providing guidance and assistance throughout my time at the lab: Sean Liddick, Alex Brown, Michael Thoennessen, and Norman Birge. Shumpei Noji requires special acknowledgment. Without your steadfast assistance much of this thesis would not exists. Throughout the years you have gone above and beyond to help me many times. Thank you. By any measure, completing a PhD is a difficult task. Beyond the research and coursework challenges, maintaining the necessary perseverance and optimism that is needed to finish is an undertaking of its own. This was only possible with the help and friendship of my fellow students. Alanna, Bethany, Charlie, Chris, Eric, and Juan: you have been true friends since the beginning. Most importantly, I would like to thank my mother and father. Naturally, without you, I wouldn’t have gotten very far in life. It must be a distinct source of pride to have two sons who have squandered their twenties obtaining their PhDs. v Finally, I want to thank Jeny for supporting and encouraging me since the day we met. The help you have given me in just the last few days has been immeasurable. You have been there for me throughout the toughest parts of this process. I dont know how I could have done it without you. vi TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi Chapter 1 Introduction . . . . . . . . . . . . . . . . . 1.1 The Atomic Nucleus . . . . . . . . . . . . . . . . . 1.2 The Nuclear Shell Model . . . . . . . . . . . . . . . 1.3 Giant Resonances . . . . . . . . . . . . . . . . . . . 1.4 Nuclear Charge-Exchange Reactions . . . . . . . . . 1.5 Historical Development of Charge-Exchange Probes 1.6 The (p,n) Probe in Inverse Kinematics . . . . . . . 1.7 β Decay of 16 C . . . . . . . . . . . . . . . . . . . . 1.8 The Gamow-Teller Sum Rule . . . . . . . . . . . . 1.9 Quenching of Gamow-Teller Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 2 7 10 13 15 17 18 19 Chapter 2 Theoretical Techniques . . . 2.1 Distorted Wave Born Approximation 2.2 Effective Interaction . . . . . . . . . 2.3 Proportionality . . . . . . . . . . . . 2.4 Technical Details of Calculations . . 2.5 Shell Model Results . . . . . . . . . . 2.6 DWBA Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 22 28 29 30 33 36 Chapter 3 Experiment . . . . . . . . . . . . . . . . . 3.1 Experimental Method . . . . . . . . . . . . . . . . 3.2 Experimental Equipment . . . . . . . . . . . . . . 3.2.1 Beam Creation and Delivery . . . . . . . . 3.2.2 Ion Source . . . . . . . . . . . . . . . . . . 3.2.3 Cyclotrons . . . . . . . . . . . . . . . . . . 3.2.4 A1900 . . . . . . . . . . . . . . . . . . . . 3.2.5 S800 Spectrograph . . . . . . . . . . . . . 3.2.6 Focal Plane Detectors . . . . . . . . . . . 3.2.7 Liquid-Hydrogen Target . . . . . . . . . . 3.2.8 LENDA and VANDLE . . . . . . . . . . . 3.3 Digital Data Acquisition System . . . . . . . . . . 3.3.1 DDAS Overview . . . . . . . . . . . . . . 3.3.2 Digital Timing . . . . . . . . . . . . . . . 3.3.3 Intrinsic Timing Resolution of DDAS . . . 3.3.4 Timing Resolution of LENDA with DDAS 3.3.5 Energy Extraction Methods . . . . . . . . 3.3.6 KeVee to Tn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 39 44 44 44 45 46 47 48 50 50 52 52 53 54 58 60 63 vii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Implementation with the S800 . 3.4.1 Clock Synchronization . 3.4.2 Operation Infinity Clock 3.4.3 Trigger Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 64 65 67 Chapter 4 Data Analysis . . . . . . . . . . . . . . 4.1 Calibration . . . . . . . . . . . . . . . . . . . . 4.1.1 Neutron Detector Timing Calibrations . 4.1.2 LENDA and VANDLE Gain Calibrations 4.1.3 Particle Identification . . . . . . . . . . . 4.1.4 CRDC Calibration . . . . . . . . . . . . 4.2 Experimental Yield . . . . . . . . . . . . . . . . 4.3 Neutron Detector Efficiency and Acceptance . . 4.4 Background Subtraction . . . . . . . . . . . . . 4.5 Absolute Normalization Corrections . . . . . . . 4.5.1 Incident-Beam Measurement . . . . . . . 4.5.2 S800 Acceptance . . . . . . . . . . . . . 4.5.3 Target Thickness . . . . . . . . . . . . . 4.5.4 Target Density Variations . . . . . . . . 4.6 Cross Section Calculation . . . . . . . . . . . . 4.7 Systematic Error . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 70 70 75 77 81 85 89 96 102 102 103 108 111 112 116 Chapter 5 Results . . . . . . 5.1 Angular Distributions . . 5.2 Multipole Decomposition 5.3 B(GT) Extraction . . . . 5.4 Dipole Strength . . . . . . . . . . . . . . . Analysis . . . . . . . . . . Chapter 6 Discussion . . . . . 6.1 The (p,n) Probe in Inverse 6.2 Shell Model Comparison . 6.3 Quenching . . . . . . . . . 6.4 The First Two 1+ States . Chapter 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 121 122 127 130 . . . . . . . Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 134 135 136 138 Conclusion and Outlook . . . . . . . . . . . . . . . . . . . . . . . . 144 APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . Appendix A Derivation of the Gamow-Teller Sum Rule . . . . Appendix B Energy of a Compton Edge . . . . . . . . . . . . Appendix C The Number of Beam Particles in a Beam Bunch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146 147 150 152 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 viii LIST OF TABLES Table 1.1: Summary of the results from previous β-decay measurements of 16 C and from theoretical calculations. a taken from Ref. [73]. b taken from Ref. [71]. 18 Table 1.2: Quenching factors of GT strength for nuclei occupying the p, sd and pf shells. The factors presented are the quenching of the GT operator. To find the quenching of GT strength, the factor must be squared, so in the sd shell it is 0.762 = 0.58. . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 Table 2.1: Parameters for the optical potentials used in the 16 C(p,n)16 N cross-section calculations. All parameters are given in MeV and Fermi. V refers to a volume term and SO a spin-orbit term. The volume terms are WoodsSaxon potentials. The spin-orbit terms are derivatives of Woods-Saxons. The p and n refer to the potentials for the proton (entrance) and neutron (exit) channels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 Table 2.2: Isospin Clebsch-Gordan factors for the Ti = 2 → Tf = 1, 2, 3 transitions. The factors give in column 3 are normalized to the T − 1 transition. . . . Table 3.1: Table of the different decay channels accessible in the 16 C(p,n)16 N reaction for ranges of excitation energy in 16 N and their associated magnetic rigidities. The acceptance of each rigidity setting is ± %5. . . . . . . . . . Table 4.1: Table of the γ-ray calibration sources used for the neutron detector gain calibration. See text for details. . . . . . . . . . . . . . . . . . . . . . . . . 33 43 76 Table 4.2: Summary of corrections to the cross sections. Parameters with a range in values indicate that a run-by-run correction was performed. See text for details on how each correction was determined. . . . . . . . . . . . . . . . 114 Table 4.3: Summary of the source and estimated size of the major systematic uncertainties in the 16 C(p,n)16 N measurement. See text for details on each source of systematic uncertainty . . . . . . . . . . . . . . . . . . . . . . . 118 Table 6.1: Comparison of the quenching factors of GT strength for 16 C, various nearby nuclei, and adopted values for several nuclear shells. Quenching studies from charge-exchange reactions and β decay studies are presented. The p-shell factor comes from the empirical relation given in Ref. [117] with A = 16. See text for details. . . . . . . . . . . . . . . . . . . . . . . . . . . 137 ix Table 6.2: Charge-exchange differential cross section at zero momentum transfer and extracted B(GT) for the first two 1+ states in 16 N. A breakdown of the statistical and systematic error is shown. The systematic error for the B(GT) is increased by the uncertainty in σ̂GT . . . . . . . . . . . . . . . . 139 Table 6.3: Comparison of the B(GT) values from charge-exchange, β decay, and the shell model for the first two 1+ states in 16 N. . . . . . . . . . . . . . . . . 140 x LIST OF FIGURES Figure 1.1: Chart of the nuclides showing the number of neutrons on the x axis and the number of protons on the y axis. The black squares show the stable nuclei, the blue region shows nuclei that have been observed but are unstable, and the red region shows nuclei that are predicted to exist but have not been observed experimentally. The dashed lines show the nuclear magic numbers. Figure taken from Ref. [3]. . . . . . . . . . . . . . . . . . . . . 2 Figure 1.2: Neutron single-particle states in 208 Pb with three different potential models. The left shows a harmonic oscillator potential, the center shows a Woods-Saxon potential, and the right shows a Woods-Saxon plus spinorbit potential. The numbers in brackets give the maximum number of neutrons that each state can hold while the adjacent numbers give the running sum. Figure taken from Ref. [7]. . . . . . . . . . . . . . . . . . . 4 Figure 1.3: Schematic of the macroscopic depiction of various giant resonances. For each ∆L value, the different possible spin and isospin transfers are shown. Figure taken from Ref. [18]. . . . . . . . . . . . . . . . . . . . . . . . . . 8 Figure 1.4: Shell-model schematic showing the different 1p-1h contributions of the GT and SD giant resonances in the 16 C(p,n)16 N reaction. . . . . . . . . . . . 17 Figure 2.1: Experimental unit σ̂GT and σ̂F as a function of mass number. The dashed and dotted lines are the result of the parameterization of Taddeucci et al. Figure is reproduced from Ref. [56]. . . . . . . . . . . . . . . . . . . . . . 31 Figure 2.2: The square root of the ratio of the GT and Fermi unit cross sections as a function of incident proton energy. The data points come from reactions on 14 C at various energies. The dashed line is a linear fit to the data above 50 MeV. Figure taken from Ref. [56], see references therein. . . . . 32 Figure 2.3: Shell Model calculations of B(GT) for 16 C 0+ ; T = 2 →16 N 1+ ; T = 1, 2 transitions. The WBP (a) and WBT (b) interactions were used in the spsdpf model space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 Figure 2.4: Shell-model spectrum with the WBP interaction for the SD 16 N, presented as the peak cross-section value for each state. into final states of 16 N with Jfπ = 0− , 1− , 2− are shown. All T = 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 xi strength in Transitions states have . . . . . . . Figure 2.5: Scatter plot of the shell-model B(GT) for 0+ → 1+ transitions against the theoretical σ̂ determined from DWBA. The shell-model calculation was done with the WBT interaction in the spsdpf model space. The red dot indicates the 2nd shell-model state. In e10003 states with strengths below ∼ 0.05 could not be discerned. . . . . . . . . . . . . . . . . . . . . 37 Figure 2.6: The model angular distributions for ∆L = 0, 1, 2 that are used in the MDA analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 Figure 3.1: Kinematic correlations for the 16 C(p,n)16 N reaction at 100 MeV/u, in the plane of the laboratory neutron angle (θn ) and neutron kinetic energy (Tn ). The blue box indicates the approximate coverage of the neutron detectors in e10003. Solid lines indicate the excitation energies in 16 N. Dashed lines show the COM scattering angles. . . . . . . . . . . . . . . . 40 Figure 3.2: Top-view schematic of the experimental setup for the neutron detectors. 41 Figure 3.3: Picture of the complete setup for experiment e10003. In the bottom center of the picture, the south LENDA bars are visible. The north VANDLE bars can be seen in the top left. The liquid hydrogen target can be seen in the center of the photograph. . . . . . . . . . . . . . . . . . . . . . . . 42 Figure 3.4: Schematic showing the coupled cyclotrons and the A1900 fragment separator. Figure taken from Ref. [95]. . . . . . . . . . . . . . . . . . . . . . 45 Figure 3.5: Schematic of the S800 analysis line and spectrometer. Figure adapted from [94]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 Figure 3.6: Schematic of the two cathode readout drift chambers, with an example event trajectory passing through the detectors. The inset shows an example of a charge distribution detected by the pads. Figure originally taken from Ref. [101] and modified by Ref. [102] . . . . . . . . . . . . . . . . . 49 Figure 3.7: Pressure-temperature diagram for the liquid hydrogen target showing the phase transition from gas to liquid. The red line indicates the expected relationship taken from [104]. . . . . . . . . . . . . . . . . . . . . . . . . 51 Figure 3.8: Example of digital trace from LENDA (filled circles) with overlaid digital CFD (filled squares). The thick line is the third-order interpolation function calculated for this event. . . . . . . . . . . . . . . . . . . . . . . 55 Figure 3.9: Time-difference spectra from a split-signal electronics test. The time differences from the 3 different timing algorithms have been shifted to zero for comparison. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 xii Figure 3.10: Electronic timing resolution of DDAS as measured with a split signal from a LENDA bar. The upside down triangles utilize a linear-interpolation method while the filled circles use a cubic-interpolation method. See text for details. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 Figure 3.11: Measured timing resolution for LENDA readout with DDAS. Upside down triangles use a linear-interpolation method, while the filled circles use a cubic-interpolation method. . . . . . . . . . . . . . . . . . . . . . . . . . 59 Figure 3.12: A trace from a PMT of LENDA from a 22 Na source. The first integration region is used to determine the average baseline of the signal, while the second integrates the pulse and determines the energy of the event. . . . 60 Figure 3.13: Top panel shows a calibrated energy spectrum from 22 Na seen in a LENDA bar. Bottom panel shows a calibrated energy spectrum from a 241 Am source. 62 Figure 3.14: Scatter plot of light output against neutron energy from a 252 Cf source measured with LENDA and a EJ-301 liquid scintillator as the timing reference. The cuts on maximum light output are applied in this figure. . 63 Figure 3.15: Detailed diagram of the electronics setup in experiment e10003. The NSCL numbers refer to the NSCL E-Pool number for the module depicted. Care has been taken to draw the modules with enough detail to make them easily recognizable. The positions of the switches depicted in the first two modules on the left correspond to their actual setting in the experiment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 Figure 4.1: Two dimensional histogram showing the correlation between the neutron TOF, and dispersive angle (a) and position (b) in the S800 focal plane. The γ flash is seen prominently in the figures. The neutron events are to the right, but are cut off to clearly show the γ flash. . . . . . . . . . . . . 72 Figure 4.2: Neutron TOF plotted against the calibrated light output from the neutron detectors. The γ flash can be seen centered around 0. At low energies, a small correlation can be seen for the γ events. . . . . . . . . . . . . . . . 73 Figure 4.3: Fully Corrected neutron time of flight spectrum for the N15 data (3.1325 Tm setting). The dashed green line indicates the TOF cut between γ and neutron events. It corresponds to a neutron kinetic energy of ∼67 MeV, which is far above the energies used in the analysis. . . . . . . . . . . . . 74 Figure 4.4: Corrected neutron TOF spectra for the 14 N (2.9 Tm) (a), 16 N (3.38 Tm) (b), 15 C (3.6565 Tm) (c), and 14 C (3.38 Tm) (d) channels. The 15 C and 14 C data from the 3.52 Tm and 3.2828 Tm settings are not shown, but have similar TOF spectra. . . . . . . . . . . . . . . . . . . . . . . . . . . 75 xiii Figure 4.5: Example light output calibration for south LENDA bar 2. The data points are described in Table 4.1. The red line is a linear fit to the data. . . . . 77 Figure 4.6: Uncalibrated Particle Identification for the 3.1325 Tm setting. . . . . . . 78 Figure 4.7: Correlation between the ejectile TOF and the dispersive angle (a) and position (b) in the S800 focal plane for 15 N events. The artifacts seen in (b) are the result of bad events in the CRDCs, which are removed in subsequent steps of the analysis. . . . . . . . . . . . . . . . . . . . . . . . 79 Figure 4.8: Corrected PID spectra from the 3.1325 Tm (a), 2.9 Tm (b), 3.6565 Tm (c), and 3.38 Tm (d) settings. The main particles of interest in each setting are indicated. The remaining two settings 3.52 Tm and 3.2828 Tm are not shown, but are similar to the 3.6565 Tm and 3.38 Tm settings. . . . 80 Figure 4.9: The uncalibrated position (x-axis) and time (y-axis) measured by the CRDCs during the mask runs. . . . . . . . . . . . . . . . . . . . . . . . . 82 Figure 4.10: The pad distribution for CRDC 1 before (a) and after (b) the pad-by-pad gain matching. The data is from one experimental run of the 3.1325 Tm rigidity setting. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 Figure 4.11: Two dimensional histogram of lab neutron angle and kinetic energy measurements (a,c,e,g,i), and reconstructed excitation energy in 16 N and COM scattering angle (b,d,f,h,j). In (a,c,e,g,i) the solid lines indicate excitation energy in 16 N and the dashed lines indicate the COM scattering angle. . 87 Figure 4.12: Experimental yield for the 15 N channel (3.1325 Tm) in the Θcm = 4−6° bin. 89 Figure 4.13: A cartoon depicting the interactions of neutrons in a plastic scintillating material. In order for the neutron to be detected it must undergo a nuclear reaction with the hydrogen or carbon nuclei in the material. The angles of the scattering are exaggerated for artistic purposes. The dominate elastic scattering on hydrogen is shown at the top of the figure. Elastic (shown at the bottom of the figure) and inelastic (not shown) scattering on 12 C are also possible but highly quenched. See text for details. . . . . . . . . 91 Figure 4.14: Light output curves for BC-400 plastic scintillator. Each line shows the expected light output as a function of incident particle energy. Adapted from [114] see text for details. . . . . . . . . . . . . . . . . . . . . . . . . 92 xiv Figure 4.15: Two dimensional histogram of light output and incident neutron energy from the Geant4 simulation. Contributions from elastic scattering on hydrogen and 12 C are included in the simulation. Inelastic channels for carbon scattering are also included and create the few events above the elastic scattering trend line. These events amount to a tiny portion of the data (<0.05%). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 Figure 4.16: The black points show the measured intrinsic neutron detection efficiency for a single LENDA bar at a light output threshold of 30 keVee. The blue line is the result of the Geant4 simulation of the efficiency. Good agreement is seen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 Figure 4.17: Two dimensional histogram of Θcm and Ex in 16 N from the Geant4 simulation of the e10003 setup. . . . . . . . . . . . . . . . . . . . . . . . . . 94 Figure 4.18: Simulated neutron acceptance and efficiency in the Θcm =4-6 ° (a) and Θcm =14-16° (b) bin for the e10003 setup. . . . . . . . . . . . . . . . . . 95 Figure 4.19: Two dimensional histogram of lab neutron scattering angle and neutron kinetic energy for 15 N (a) and 11 B (b) gated events. The events from 11 B are used to model the background seen in the 15 N channels. . . . . . . . 98 Figure 4.20: Ratio of the 15 N and 11 B gated events integrated from 40 to 60 MeV in neutron kinetic energy as a function of the lab neutron scattering angle. The blue points show the region of charge-exchange reactions, while the black points indicate regions with only background. The red line is a linear fit to the black points in the range shown. . . . . . . . . . . . . . . 99 Figure 4.21: Experimental yields with overlaid background shape from the 11 B models. The most forward (Θcm =4-6°) and backward (Θcm =14-16°) COM angles are shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 Figure 4.22: The kinetic energy spectrum of 15 C events detected in the S800 Spectrograph. The black (blue) hashed region comes from the 3.52 (3.6565) Tm rigidity setting. The gray line is the full distribution constructed from the two settings. The thin blue line shows the simulated distribution from charge-exchange reactions. The low-energy tail of the distribution is energetically inaccessible to charge-exchange reactions and so must come from background processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 Figure 4.23: Experimental yield and background models for the 15 C and 14 C channels. The data comes from the 3.6565 Tm and 3.38 Tm rigidity settings respectively. Yield is primarily seen at backward angles. . . . . . . . . . 101 xv Figure 4.24: Histogram of the energy deposited in the object scintillator in the first run of the experiment. The peaks correspond to to single- and multi-particle hits in the scintillator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 Figure 4.25: A two-dimensional histogram of the kinetic energy and dispersive angle at the target for the 14 C channel as measured in the 3.2828 Tm rigidity setting. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 Figure 4.26: Full kinetic energy distribution for 14 C ejectiles reconstructed from 4 rigidity settings. The vertical line shaded regions indicate areas of incomplete acceptance. See text for details. . . . . . . . . . . . . . . . . . . . . . . . 105 Figure 4.27: Experimentally determined acceptance curve for 14 C in the 3.2828 Tm setting. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 Figure 4.28: Efficiency curves from the finite S800 acceptance for each rigidity setting used in the analysis. Lines of common color show settings centered around the same detected particle in the CE reaction. . . . . . . . . . . . . . . . 107 Figure 4.29: PID spectrum from the run where the 16 C was tuned into the focal plane of the S800. The beam consisted of 16 C and a small containment of 14 B. 109 Figure 4.30: Kinetic energy measured by the s800 Spectrograph for the two beam particles. The peak from 16 C has been scaled down for easy comparison with 14 B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 Figure 4.31: Relationship between energy loss difference and liquid hydrogen target thickness determined from simulation. The points are simulated values and the red line is a linear fit. . . . . . . . . . . . . . . . . . . . . . . . . 110 Figure 4.32: The density derived from the temperature measurement of the liquid hydrogen target for each experimental run. . . . . . . . . . . . . . . . . . . 111 Figure 4.33: Differential cross section for the 16 C(p,n)16 N reaction showing the contributions from each reaction channel. Panel (a) shows the cross section at forward angles in 250 keV bins. In (b) the same angle is shown with a coarser binning to highlight the channels at higher excitation energy, where the statistics is low. The most backward angle, with the poorest resolution, is presented in (c). The shaded bans in (b) and (c) indicate the systematic error estimation from the background subtraction for the 15 C and 14 C channels. The small vertical lines on the top of each figure indicate the threshold energies for the 15 N, 15 C, 14 C and 14 N channels. The color of the line matches the color of the data to indicate the channel. 115 xvi Figure 4.34: Total differential cross section from all decay channels in the 16 C(p,n)16 N reaction. The green band shows the systematic error including the systematic error from the background subtractions in the different channels and the systematic error in the neutron acceptance determination. . . . . . . 117 Figure 4.35: The total charge-exchange cross section from 0-20 MeV for each experimental run for the 15 N channel. The spread in the points gives an estimate of the uncertainty in the beam normalization. . . . . . . . . . . . . . . . 119 Figure 5.1: Differential cross section for the 16 C(p,n)16 N reaction as a function of excitation energy. Data is from Θcm =4-6° but smeared to match the resolution at Θcm =14-16°. . . . . . . . . . . . . . . . . . . . . . . . . . . 122 Figure 5.2: Measured COM angular distributions in 1 MeV bins from 0-20 MeV in 16 N. The black dots are the experimental data. The red, green, and blue lines are the ∆L=0,1,2 components of each fit determined from the MDA. The dotted line is the sum of the 3 MDA fit components. . . . . . . . . . 124 Figure 5.3: MDA results. The ∆L components of the cross section is shown as a function of excitation energy for each angular bin. At forward angles the spectrum is dominated by ∆L = 0 transitions, while at backward angles ∆L = 1 transitions appear at high excitation energies. . . . . . . . . . . 128 Figure 5.4: The differential cross section extrapolated to zero momentum transfer. The error bars show statistical errors. The blue peak shows the contribution from the IAS to the spectrum. See text for details. . . . . . . . . . . 129 Figure 5.5: Extracted GT strength distribution (a) and summed strength (b) in 16 N up to Ex =20 MeV. The green band shows the total systematic error in both plots. The data is compared with two shell interactions WBT and WBP, both calculated in the spsdpf model space. . . . . . . . . . . . . . 131 Figure 5.6: The black points show the dipole (∆L = 1) differential cross section determined from the MDA at 15°. The error bars give the statistical uncertainty from the MDA fits. The black histogram gives the smeared, DWBA cross sections from the 1,3 h̄ω shell-model calculation with the WBP interaction in the spsdpf model space. The blue and red histograms give the results of the calculations when pure 1 and 3 h̄ω contributions are calculated separately. Panel (a) Shows the result when the full MDA fit is used, while (b) excludes the ∆L=2 component. The 7-8 MeV bin was excluded from (a) since its error bar was too large to be meaningful and was consistent with zero. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 Figure 6.1: The amount of quenching between the experimental data and the shell model (WBP interaction) as a function of excitation energy. . . . . . . . 139 xvii Figure 6.2: A typical transition density with its 1p-1h components. This is the calculation of the transition to the third 1+ state in 16 N with the WBT interaction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 Figure 6.3: The transition densities and their 1p-1h components for the first two 1+ states in 16 N, calculated with the WBT interaction. The details of these states create an atypical transition density that has a significant effect on the unit cross section for these two states. . . . . . . . . . . . . . . . . . 142 Figure C.1: The figure shows all possible ways to distribute 2 particles into 3 buckets. Each row in the figure indicates a different arrangement of particles, giving a total of 6 possible configurations. . . . . . . . . . . . . . . . . . . . . . 153 xviii Chapter 1 Introduction 1.1 The Atomic Nucleus Nuclear physics is a unique field of study. While it has had significant impact on the world from its beginnings, a robust and a comprehensive predictive theory remains elusive. The most basic goal of nuclear science is to understand the forces that bind the atomic nucleus together and, with those forces, understand the many phenomena observed in nuclei and their interactions. Though they are not fundamental particles, the nuclear problem is typically defined in terms of the nucleon constituents—protons and neutrons—that interact via the strong nuclear force. Beginning with Rutherford’s discovery of the nucleus [1], the field has grown to discover more than 3000 different isotopes out of approximately 7000 predicted to exist [2]. These different nuclei are organized in a way similar to the periodic table of the elements, called the chart of the nuclides. This chart is shown in Fig. 1.1, where the number of neutrons is plotted on the horizontal axis and the number of protons is plotted on the vertical axis. The vertical and horizontal dashed lines in Fig. 1.1 indicated the “magic numbers” (2, 8, 20, 28, 50, 82, 126). The magic numbers and their meaning were a crucial discovery for nuclear physics. They correspond to numbers of nucleons that form and designate particularly stable or strongly bound nuclei. This is a phenomenon that is similar to the noble 1 Figure 1.1: Chart of the nuclides showing the number of neutrons on the x axis and the number of protons on the y axis. The black squares show the stable nuclei, the blue region shows nuclei that have been observed but are unstable, and the red region shows nuclei that are predicted to exist but have not been observed experimentally. The dashed lines show the nuclear magic numbers. Figure taken from Ref. [3]. gases of atomic physics, where a completely filled outer electron shell makes them inert. The magic numbers were first recognized by Maria Mayer in 1948 [4], and later further explained by Mayer [5], and Haxel, Jensen, and Suess in 1949 [6]. The insight made independently by Mayer and Jensen et al. was the necessity of a strong spin-orbit interaction to reproduce the observed magic numbers. Both Mayer and Jensen were awarded a portion of the Nobel Prize in Physics in 1963 for creating and explaining the nuclear shell model. 1.2 The Nuclear Shell Model The nuclear shell model organizes the protons and neutrons of a nucleus into shells that attempt to reproduce the observed magic numbers. Strictly speaking, the nucleus is a selfbound system of A nucleons, where the attractive nuclear potential between all the particles 2 binds the nucleus. The shell model, however, is a mean-field approximation that describes the motion of the individual nucleons within an average potential that models the total effect of the remaining A − 1 nucleons. This marks a significant difference from the atomic case. For an atom the negatively charged electrons are bound by the Coulomb potential well of the positively charged nucleus, and not by the collective effect of all other electrons in the system. There are two traditional forms of the potential used in the nuclear shell model: a harmonic-oscillator 1 V (r) = mω 2 r2 , 2 (1.1) or Woods-Saxon potential V (r) = V0 . 1 + exp[(r − R)/a] (1.2) For the harmonic-oscillator potential, m is the nucleon mass and ω is the oscillator frequency. For the Woods-Saxon potential, V0 is the depth, R is the radius, and a is the diffuseness of the potential well. In addition to the potential well described by V (r), a term to model the spin-orbit interaction between the nucleons is also included. For the spin-orbit potential, it is important to note that it must vanish at the center of the nucleus [7]. This motivates using a potential shape that is peaked at the surface. The derivative of a Wood-Saxon is typically used, Vso (r) = Vso 1 dfso (r) , r dr (1.3) where fso = 1 1 + exp[(r − Rso )/aso ] (1.4) These potentials are summarized in Fig. 1.2, where the neutron single-particle states of 3 10 Neutron Single-Particle Energies 0 [56] 168 N=6 [ 2] 168 3s [10] 166 2d [18] 156 1g [26] 138 0i [42] 112 N=5 single-particle energy (MeV) -10 -20 126 [ 6] 112 2p [14] 106 1f [22] 92 0h [30] 70 N=4 [ 4] 184 2d3 [16] 180 0j15 [ 8] 164 1g7 [ 2] 156 3s1 [ 6] 154 2d5 [12] 148 0i11 [10] 136 1g9 82 [ 2] 70 2s [10] 68 1d [18] 58 0g 50 [20] 40 N=3 [ 6] 40 1p [14] 34 0f [12] 20 N=2 [ 2] 20 1s -30 28 20 [10] 50 0g9 [ 2] 40 1p1 [ 4] 38 1p3 [ 6] 34 0f5 [ 8] 28 0f7 [ 6] 14 0d5 8 [ 2] 8 0p1 [ 4] 6 0p3 [ 6] 8 0p -40 [12] 82 0h11 [ 2] 70 2s1 [ 4] 68 1d3 [ 6] 64 1d5 [ 8] 58 0g7 [ 2] 20 1s1 [ 4] 18 0d3 [10] 18 0d [ 6] 8 N=1 [ 2] 126 2p1 [ 6] 124 1f5 [ 4] 118 2p3 [14] 114 0i13 [10] 100 0h9 [ 8] 90 1f7 2 [ 2] 2 N=0 Harmonic Oscillator Potential [ 2] 2 0s Woods-Saxon Potential [ 2] 2 0s1 Woods-Saxon Plus Spin-Orbit Potential -50 Figure 1.2: Neutron single-particle states in 208 Pb with three different potential models. The left shows a harmonic oscillator potential, the center shows a Woods-Saxon potential, and the right shows a Woods-Saxon plus spin-orbit potential. The numbers in brackets give the maximum number of neutrons that each state can hold while the adjacent numbers give the running sum. Figure taken from Ref. [7]. 4 208 Pb are shown. Though the harmonic oscillator clearly has the wrong asymptotic behavior (V(r→ ∞)→ ∞), it reproduces several of the observed magic numbers (2, 8, and 20). The Woods-Saxon potential, shown in the center of Fig. 1.2, breaks the the l degeneracy by bringing states with higher l values down in energy, but still fails to reproduce the observed magic numbers above 20. Adding the spin-orbit term completely breaks the degeneracy of the states and reproduces the magic numbers as Mayer and Jensen et al. first showed. The spin-orbit term in the nuclear case is of opposite sign to the atomic case, bringing higher J states lower in energy. This is shown as the right column in Fig. 1.2. This basic description of the shell model is referred to as the independent-particle model and works well for nuclei with one nucleon more or less than a magic number. In these circumstances the nucleus can be modeled well by an inert “core” and one orbiting nucleon. There the low-lying excited states of the nucleus can be understood as the valence nucleon simply moving into higher orbitals of the potential created by the core. However, in the more common case of several nucleons outside a closed core, the problem becomes more complex, since the interaction between the valence nucleons must be taken into account. This is done by solving the the Schrödinger equation, ĤΨ = EΨ, (1.5) where Ĥ is the Hamiltonian, Ψ is the wave function, and E is the energy. Unfortunately, we cannot write down the full expression for Ĥ, since we don’t know the true shape of the nuclear potential. Following the notation of [7], the Hamiltonian is often modeled as, H= n X  ! (Tk + Uk ) + k=0 n X k> r0 . 24 However, the expression for the scattering amplitude in Equation 2.11 contains the total wave function, which is the unknown we started with. To arrive at a useful result, the Born Approximation is made where the form of the solution in Equation 2.10 is recursively inserted into Equation 2.11. This gives a series solution, where each term contains higher and higher orders of the potential V . Taking just the first term, we have, f (~k, ~k 0 ) 2µ ≈− 4πh̄2 Z 0 0 e∓i~q ~r V (~r0 )d~r0 , (2.12) where ~q = ~k − ~k 0 is the momentum transferred. This gives a clear and explicit way to calculate the scattering amplitude. This is the first order Plane Wave Born Approximation (PWBA). This result is often written in terms of the T matrix, T = hφ | V | Ψi (2.13) T ≈ hφ | V | φi , where φ are plane waves. To derive the Distorted Wave Born Approximation (DWBA), it is useful to rewrite Equation 2.10 in an operator form known as the Lipmann-Schwinger equation, |Ψ± i = |φi + G± V |Ψ± i , (2.14) which, under the Born Approximation reduces to, |Ψ± i = |φi + G± V |φ± i . 25 (2.15) In nuclear reactions, the potential is often divided into two parts, V = U + W , where U is the part of the potential that describes elastic scattering for the system while W contains the residual interaction that is not included in U . U is chosen such that, |χi = |φi + GU ± U |φi , (2.16) where |χi is referred to as a distorted wave and GU ± is the Greens function for the problem containing only U . For charge-exchange calculations, W contains the interaction responsible for exciting the spin-isospin degrees of freedom in the target nucleus. The purpose of this separation becomes clear when you consider the total T matrix with these two potentials, T = hφ | W + U | Ψi . (2.17) For clarity, the full wave function is used in the above relation. It can be shown that, T = φ− U χ+ + χ− W Ψ (2.18) Since U only models the average potential of elastic scattering, it cannot connect the initial to the final states we are interested in for charge-exchange reactions. This leaves only the second term, T = χ− W Ψ , (2.19) where the effect of the U potential is incorporated into the determination of the distorted wave χ− . Reapplying the Born Approximation, we arrive at the T matrix for first order 26 DWBA, T DW BA = χ− W χ+ (2.20) So far the internal state of target and residual nucleus has been omitted from the derivation to focus on the details of the reaction theory. Equation 2.20 is rewritten in terms of a ~ form factor F (R), T DW BA = D ~ + χ− F (R) χ E , (2.21) ~ is the spatial distance between the center of the target nucleus and the position of where R the incident probe. The form factor is, ~ = F (R) Z drT drp ρp Vef f ρT , (2.22) where the residual interaction has been relabeled as Vef f . ρ are the transition densities of the target and projectile, ρ= X α OBT Dα hφfα ||σ̂τ̂± ||φiα i . (2.23) The one-body transition densities (OBTD) are calculated in the shell model between the initial and final states. The sum over α runs over the contributions from the different 1p1h combinations for single particle level transitions φiα to φfα . The single-particle wave functions are modeled with Woods-Saxon potentials. With the formalism above, the only remaining unknown is the form of the effective NN interaction that can meditate chargeexchange reactions. 27 2.2 Effective Interaction In this section, the phenomenological nucleon-nucleon interaction that is used to model charge-exchange reactions is discussed (Vef f ). The long-standing choice for interaction in the charge-exchange community has been the Love and Franey (LF) N-N interaction [85, 86]. The LF interaction is attractive for charge-exchange studies since it is directly parameterized in terms of the spin and isospin operators that are of interest for charge-exchange reactions. The potential between nucleons i and j is structured into central, spin-orbit, and tensor terms, ~ ·S ~ + V T (rij )Sij , Vij = V C (rij ) + V LS (rij )L (2.24) where Sij is the tensor operator Sij = 3 (σ̂i · rij )(σ̂j · rij ) 2 rij − σ̂i · σ̂j (2.25) The radial dependence of each of these terms is expanded as a sum of Yukawa potentials, V (r) = X Vi Y (r/Ri ), (2.26) i where Y (x) = e−x /x. The parameters Vi and the ranges Ri are fit to nucleon-nucleon scattering. The result of the work by Love and Franey gave effective nucleon-nucleon T matrices at a variety of incident beam energies. For the 16 C(p,n)16 N at 100 MeV/u, the LF interaction at 140 MeV was used. The 140 MeV interaction was chosen over the 100 MeV interaction because the parameters were better determined from the nucleon-nucleon scattering data [85, 86]. The GT component of the interaction does not change significantly 28 between these energies. However, the magnitude of the τ component of the NN interaction reduces with increasing energy. This is corrected for by using Equation 2.31. The interaction is folded over the transition densities of the target and projectile to arrive at the form factor given in Equation 2.23. 2.3 Proportionality This section provides a brief overview of proportionality between GT strength and chargeexchange cross section. The interested reader is referred to [56] for the complete derivation and a detailed discussion. Taddeucci et al. showed that the form of this relationship is a simple proportionality,   dσ = σ̂B(GT ), dΩ q→0 (2.27) where the q → 0 indicates that the cross section needs to be extrapolated to zero linear momentum transfer. The σ̂ is the proportionality constant of the expression and is called the unit cross section. Within DWBA and under the limits of high beam energy (>∼100 MeV/u) and the Eikonal approximation, Taddeucci et al. showed that the unit cross section factors into three components, σ̂ = KND |Jστ |2 . (2.28) K is the kinematic factor, Ei Ef kf . (h̄2 c2 π 2 )2 ki (2.29) ND is a distortion factor and Jστ is a volume integral of the relevant interaction (in this case, the Love Franey interaction). ND can be calculated as the ratio of distorted- and plane-wave cross sections (see Equation 2.23 of Ref. [56]). 29 While the factorization of the proportionality gives insight, the unit cross section is often determined empirically. When β-decay information for a low-lying state is available, σ̂ can be found by using the known B(GT) value and the measured charge-exchange cross section. When such information is not available, the unit cross section can sometimes be estimated by looking at compiled dependencies between σ̂ and mass number. Fig. 2.1 shows the mass dependence of the unit cross section over a range of nuclei for the (p,n) reaction at 120 MeV. The parameterization was determined by Taddeucci et al. [56]. The ratio of GT and Fermi unit cross sections was also investigated by Taddeucci et al. They defined, σ̂ (Ep , A) R2 = GT σ̂F (Ep , A) (2.30) The dependence of R2 on incident energy and mass number was investigated experimentally by measuring R2 for a set of targets at a variety of beam energies. A linear dependence on Ep was found, R(Ep ) = Ep . 55.0 ± 0.4 MeV (2.31) Fig. 2.2 shows this dependence for 14 C. The deviations below 50 MeV are below the incident energy region where the proportionality is applicable. 2.4 Technical Details of Calculations The cross-section calculations for the 16 C(p,n)16 N reaction at 100 MeV/u were completed using the DW81 code [87] (see Ref. [88] for a description of the original formalism). The theoretical GT strengths and one-body transition densities were calculated using OXBASH [8] in the spsdpf model space using either the WBT or WBP interactions. The optical 30 Figure 2.1: Experimental unit σ̂GT and σ̂F as a function of mass number. The dashed and dotted lines are the result of the parameterization of Taddeucci et al. Figure is reproduced from Ref. [56]. 31 Figure 2.2: The square root of the ratio of the GT and Fermi unit cross sections as a function of incident proton energy. The data points come from reactions on 14 C at various energies. The dashed line is a linear fit to the data above 50 MeV. Figure taken from Ref. [56], see references therein. potentials for the proton-16 C entrance channel and the neutron-16 N exit channel come from the parameterizations given in Refs. [89, 56]. The parameters used for the 16 C(p,n)16 N calculation are given in Table 2.1 for convenience. The LF interaction at 140 MeV was used for the N-N potential. The binding energies of the single-particle wave functions in the initial and final state were calculated using a Skyrme interaction. 32 Potential V r a W r a p-V 30.68 1.23 0.71 7.12 1.41 0.57 p-SO 3.54 1.00 0.65 -1.07 0.97 0.62 n-V 25.16 1.23 0.71 7.12 1.41 0.57 n-SO 4.01 1.00 0.65 -1.07 0.97 0.62 Table 2.1: Parameters for the optical potentials used in the 16 C(p,n)16 N cross-section calculations. All parameters are given in MeV and Fermi. V refers to a volume term and SO a spin-orbit term. The volume terms are Woods-Saxon potentials. The spin-orbit terms are derivatives of Woods-Saxons. The p and n refer to the potentials for the proton (entrance) and neutron (exit) channels. 2.5 Shell Model Results This section presents the results from the shell-model calculations performed with OXBASH. We start in the J π = 0+ ; T = 2 16 C ground state. In charge-exchange reactions (∆T = 1), the final state in 16 N can be Ti = 2 ⊗ ∆T = 1 → Tf = 1, 2, 3. If Ti is high, transitions to higher T states are suppressed by their Clebsch-Gordan coefficients. For the 16 C(p,n)16 N reaction, these factors are given in Table 2.2. Tf Coefficient Factor T+1 1 (T +1)(2T +1) 0.11 T 1 T +1 0.33 T-1 2T −1 2T +1 1 Table 2.2: Isospin Clebsch-Gordan factors for the Ti = 2 → Tf = 1, 2, 3 transitions. The factors give in column 3 are normalized to the T − 1 transition. In terms of angular momentum, the important transitions are Jfπ = 1+ (GT transitions), and Jfπ = 0− , 1− , 2− (dipole transitions). Fig. 2.3 shows the GT strength distributions 33 for 0+ → 1+ transitions calculated with OXBASH in the spsdpf model space, and the WBP and WBT interactions. Configurations in both the p and sd shells are included in the calculation. Transitions to T = 1 and T = 2 finals states are shown. Within this model space, the full strength of the sum rule is exhausted in the ∆Tz = −1 direction (the ∆Tz = +1 is Pauli blocked), so the sum of all calculated states is ∼121 . The T = 3 states contribute < 0.002 to the total strength and are neglected. Both interactions predict the majority of the GT strength to be below approximately 12 MeV, with strong states near 10 MeV. The Tf = 2 strength is a small component appearing at higher energies. The shell model does not give intrinsic widths to these states, which since they occur above particle separation thresholds are expected to be several MeV wide. Fig. 2.4 shows the more complicated case of spin-dipole transitions, calculated with the WBP interaction. Unlike the GT transitions, no simple proportionality between spin-dipole strength (B(SD)) and charge-exchange cross section has been established. The typical approach is to directly compare measured cross sections with theoretical cross section calculated in DWBA. Fig. 2.4 therefore shows the cross section value at the angle where it is maximized for each shell model state with Jfπ = 0− , 1− , 2− . The excitation energy of the states are aligned such that the first 2− state is at 0 MeV (the ground state of 16 N is 2− ). The shell model predicts a 0− ground state with slightly lower energy (∼300 keV), and so appears negative in Fig. 2.4. This difference is within the error of the calculation and below the resolution of the experiment. The shell model predicts several strong states at ∼5-10 MeV. Above ∼12 MeV, substantial strength comes form many states. The resonances is expected to extend above 20 MeV. 1 For 16 C, the GT sum rule is: 3(N − Z) = 3(10 − 6) = 12 34 3 WBP T=1 T=2 2.5 B(GT) 2 1.5 1 0.5 0 0 5 10 15 Ex [MeV] 20 25 30 25 30 (a) 3 WBT T=1 T=2 2.5 B(GT) 2 1.5 1 0.5 0 0 5 10 15 Ex [MeV] 20 (b) Figure 2.3: Shell Model calculations of B(GT) for 16 C 0+ ; T = 2 →16 N 1+ ; T = 1, 2 transitions. The WBP (a) and WBT (b) interactions were used in the spsdpf model space. 35 2 Jπ=0 Jπ=1 Jπ=2 - 1.8 [mb/sr] 1.6 1.4 max 1.2 1 dσ dΩ 0.8 0.6 0.4 0.2 0 0 5 10 15 20 25 Ex [MeV] Figure 2.4: Shell-model spectrum with the WBP interaction for the SD strength in 16 N, presented as the peak cross-section value for each state. Transitions into final states of 16 N with Jfπ = 0− , 1− , 2− are shown. All states have T = 1. 2.6 DWBA Results Within the shell-model and DWBA theory the proportionality between the charge-exchange cross section at vanishing momentum transfer and B(GT) can be explored. This is done by taking many shell-model states and calculating their differential cross sections at q = 0. Then, a theoretical proportionality constant (σ̂) is found for each state by dividing the cross section by the B(GT) value from the shell model. This analysis is presented in Fig. 2.5 A consistent proportionality is seen for the stronger states, with a gradual decrease in quality for the weakest transitions. The spreading in the weaker states comes from the interplay between the στ and the tensor-τ portion of the effective NN interaction. The tensor-τ portion mediates ∆L=2 ∆S=1 transitions that interfere with the desired GT transitions (see Chapter 6 for more details). Since these proportionality breaking effects occur in weak transitions, they don’t contribute much to the final GT spectrum. The scatter in the strong 36 20 18 σ [mb/sr] 16 14 12 10 8 6 4 2 0 10−5 10−4 10−3 10−2 B(GT) 10−1 1 Figure 2.5: Scatter plot of the shell-model B(GT) for 0+ → 1+ transitions against the theoretical σ̂ determined from DWBA. The shell-model calculation was done with the WBT interaction in the spsdpf model space. The red dot indicates the 2nd shell-model state. In e10003 states with strengths below ∼ 0.05 could not be discerned. states tends to be evenly distributed giving the average σ̂GT a small error. The red dot indicates the transition to the 2nd 1+ shell model state and is a clear outlier with a significantly higher predicted σ̂. The detailed properties of this state will be discussed in Chapter 6. In order to isolate the different orbital angular momentum transfers in the data, a multipole decomposition analysis (MDA) will have to be performed. Details on the procedure for the MDA is given in Section 5.2. Here the calculations of the characteristic angular distributions needed for the MDA are shown. Angular distributions with pure ∆L = 0 ∆S = 1 components from the 0+ → 1+ transitions were used as a model for the characteristic ∆L = 0 angular distribution. The calculations to 1− final states were used for the characteristic dipole angular distribution. Transitions to 2+ final states were used for the quadrupole component. For each multipole, the first 100 shell-model states were examined with the Q value in the reaction calculation set to zero. From these angular distributions, a representative 37 strong state was chosen as the model for each ∆L transfer. The deviation between these states was small compared to the overall systematic error of the measurement. These model angular distributions are shown in Fig. 2.6. 3 ∆L=0 ∆L=1 ∆L=2 dσ [arb] dΩ 2.5 2 1.5 1 0.5 0 0 5 10 15 Θcm [deg] 20 25 Figure 2.6: The model angular distributions for ∆L = 0, 1, 2 that are used in the MDA analysis. 38 Chapter 3 Experiment Nothing works! Everything is broken! Ancient Proverb The (p,n) charge-exchange experiment in inverse kinematics involves the detection of lowenergy neutrons in coincidence with fast residues in a magnetic spectrometer. The neutron laboratory scattering angle and neutron kinetic energy provide event-by-event kinematic information necessary to reconstruct the excitation energy in 16 N and the center-of-mass (COM) scattering angle. For experiment e10003, the neutrons were detected using two lowenergy neutron-detector arrays: LENDA [90, 91] and VANDLE [92, 93]. The fast residues of 16 N were detected by the S800 Spectrograph [94]. The experiment ran in March of 2015 with the goal of extracting the Gamow-Teller strength distribution up to 20 MeV of excitation energy in 16 N. Details on the experimental equipment and method will be given in the sections below. 3.1 Experimental Method Fig. 3.1 shows kinematic correlations between the laboratory neutron kinetic energy and scattering angle, and the excitation energy in 16 N and COM scattering angle for the 16 C(p,n)16 N reaction at 100 MeV/u. As Fig. 3.1 shows, to cover the excitation energy regime of interest (0-20 MeV) and the COM scattering angles of interest (0-15 degrees), a large range of labo39 ratory neutron scattering angles and low neutron energy must be measured. 30 15 20 EX= 0 MeV C(p,n)16N at100 MeV/u 16 Tn Lab [MeV] 18 16 14 200 12 10 150 8 6 100 4 2 0 0 θCM=50 20 40 60 80 100 Θ n Lab [Deg] 120 140 160 180 Figure 3.1: Kinematic correlations for the 16 C(p,n)16 N reaction at 100 MeV/u, in the plane of the laboratory neutron angle (θn ) and neutron kinetic energy (Tn ). The blue box indicates the approximate coverage of the neutron detectors in e10003. Solid lines indicate the excitation energies in 16 N. Dashed lines show the COM scattering angles. To this end, the centers of the LENDA and VANDLE bars were placed 1 m from the center of the target so as to cover laboratory scattering angles between approximately 40 and 140 degrees. A schematic of the setup can seen in Fig 3.2. Each array was split in half and arranged such that one section of LENDA was placed on the opposite side of the beam line from the corresponding section of VANDLE. The angles of the bars were further arranged such that the gap in angular coverage between two LENDA bars was filled by the VANDLE bars on the opposite side of the beam line. The arrangements of different sections of LENDA and VANDLE allowed both arrays to span the full laboratory angular range being measured, while also preventing any systematic bias between the two arrays from differences in the neutron background on either side of the beam line. The choice to arrange the angles 40 of the bars to fill the angular gaps in the detectors on the opposite side of the beam line allowed for continuous laboratory angular coverage and reduced systematic uncertainties in the estimation of geometrical acceptance. This is because regions of near zero acceptance, where systematic uncertainties due to slight deviations in detector placement are large, were avoided. Fig. 3.3 shows a picture of the complete experimental setup. North VANDLE North LENDA Beam South LENDA South VANDLE Figure 3.2: Top-view schematic of the experimental setup for the neutron detectors. 41 Figure 3.3: Picture of the complete setup for experiment e10003. In the bottom center of the picture, the south LENDA bars are visible. The north VANDLE bars can be seen in the top left. The liquid hydrogen target can be seen in the center of the photograph. 42 To isolate events that could be due to charge-exchange reactions, the ejectile was analyzed in the S800 Spectrograph. Since the experiment aimed to measure up to excitation energies of 20 MeV in 16 N, the 1n, 1p, 2n, 1p1n emission thresholds were exceeded. This caused several different nucleon decay channels to open (16 N, 15 N, 15 C, 14 C, 14 N), each of which needed to be measured to find the total charge-exchange cross section. The momentum spread of these residual ejectiles after the momentum kick from the decayed nucleons can be greater than the momentum acceptance of the S800 Spectrograph ( dp p = 5% [94]) . To account for this, seven different magnetic-rigidity settings were used to capture a large enough fraction of the momentum spectrum for each decay channel. The different decay channels and their associated rigidity settings used during the experiment are shown in Table 3.1. Ex in 16 N Reaction Detected Particle Set Centroid Rigidities 0 → 2.4 MeV 16 C(p,n)16 N 16 N 3.2828, 3.338 Tm >2.489 MeV 16 C(p,n)16 N → 15 N + n 15 N 2.98, 3.1325 Tm >11.478 MeV 16 C(p,n)16 N → 15 C + p 15 C 3.52, 3.6565 Tm >12.696 MeV 16 C(p,n)16 N → 14 C + p + n 14 C 3.2828, 3.38 Tm >13.322 MeV 16 C(p,n)16 N → 14 N + n + n 14 N 2.98, 2.9 Tm Table 3.1: Table of the different decay channels accessible in the 16 C(p,n)16 N reaction for ranges of excitation energy in 16 N and their associated magnetic rigidities. The acceptance of each rigidity setting is ± %5. With the kinematic information provided by the low-energy recoiled neutron and the momentum information of the fast ejectile, the differential cross section could be reconstructed up to 20 MeV of excitation energy in 16 N. 43 3.2 Experimental Equipment 3.2.1 Beam Creation and Delivery The 16 C(p,n)16 N reaction was performed with a beam of radioactive 16 C nuclei. Radioactive ion beams are created at the NSCL through the process of projectile fragmentation, where an accelerated stable ion beam is impinged on a thick 9 Be production target. The impinging ions are fragmented in flight by the target, creating a wide array of different residual isotopes, some of which are the unstable, exotic isotopes that are of interest for study. The specific isotopes of interest are then selected in flight by their magnetic rigidity using the A1900 fragment separator [95], possibly resulting in a nearly pure beam of the isotope of interest. This beam can be directed to one of several different experimental areas for study. 3.2.2 Ion Source At the NSCL, there are two options for ion sources: the Superconducting Source for Ions (SuSI) [96, 97] and the Advanced Room TEMperature Ion Source (ARTEMIS) [98]. Both take a stable element and ionize it using the Electron Cyclotron Resonance (ECR) method. In the ECR method, a plasma is confined by a magnetic field while microwave power is applied at the electron cyclotron frequency, ωc = eB , me (3.1) where e is the electron charge, me is the electron mass, and B is the applied magnetic field strength. The particles are ionized by collisions with the moving electrons. The ions are then extracted and injected into the K500 cyclotron for acceleration. For e10003, the SuSI 44 source was used to produce 18 O3+ ions. 3.2.3 Cyclotrons The coupled cyclotron facility at the NSCL [99, 100] houses two super-conducting cyclotrons, the K500 and the K1200, where the numeric designations refer to the maximum extraction energy achievable for protons. The cyclotrons accelerate charged ions by applying a strong radio-frequency (RF) electric field, while a magnetic field constrains the path of the particles to a circular orbit. Since the acceleration is more effective with higher charge states, a stripper foil designed to remove electrons from the ions is placed at the entrance to K1200. For light and medium-heavy nuclei, this allows for the creation of completely ionized beams. Fig. 3.4 shows a schematic of the two cyclotrons. Figure 3.4: Schematic showing the coupled cyclotrons and the A1900 fragment separator. Figure taken from Ref. [95]. For e10003, the K500 cyclotron accelerated a beam of 18 O3+ ions to an energy of 10.91 MeV/u. These ions were extracted and sent through the stripper foil at the entrance to the K1200 cyclotron, where the completely ionized 18 O nuclei were further accelerated to 120 MeV/u. The beam was then impinged on a 846 mg/cm2 target of 9 Be for fragmentation. 45 The resulting fragments then entered the A1900 fragment separator. 3.2.4 A1900 The A1900 fragment separator [95] consists of four dipole bending magnets and eight quadrupole triplet focusing magnets. The magnetic field of the dipole magnets serves to disperse the incoming particles according to their magnetic rigidity, Bρ = p/q, (3.2) where B is the magnetic field, ρ is the bending radius, q is the charge and p is the relativistic momentum. Together, Bρ is called the magnetic rigidity. Further selection is accomplished by placing a wedge at the intermediate image of the A1900 (see “image 2” in Fig. 3.4). The wedge causes species of different atomic numbers to lose different amounts of energy, allowing the particles to be separated by their new magnetic rigidities in the following two dipole magnets. For e10003, the magnets were tuned to select 16 C. The slits at the first image of the separator were set such that the momentum acceptance was dp p = 0.54%. An aluminum wedge of 240 mg/cm2 was placed at the intermediate image of the A1900 to further remove contaminants. This resulted in a 97.8% pure beam of 16 C, with an intensity of 3.0 · 104 pps pnA 18 0 as measured at the focal plane of the A1900. The largest remaining contaminant in the beam was 14 B. 46 3.2.5 S800 Spectrograph Following the selection of 16 C in the A1900, the beam was transported through the transfer line to the S3 experimental vault. The S3 vault consists of the S3 analysis line and the S800 Spectrograph as can be seen in Fig. 3.5. The spectrograph has a momentum acceptance of 5% and covers a solid angle of 20 msr. The analysis line can operate at maximum rigidity of 4.9 Tm and is limited to 4 Tm for the spectrograph itself. Since the reconstruction of the excitation energy and COM scattering angle for the (p,n) reaction in inverse kinematics comes entirely from the measurement of the recoiled neutron, high resolution of the ejectile was not needed. Therefore, S800 beam line was operated in achromatic beam-transport mode. The fast ejectile particles from the 16 C(p,n)16 N reaction were bent through the dipoles of the spectrograph into a series of focal-plane detectors. The purpose of the focal-plane detectors were to determine the hit position, angle, energy loss, and time-of-flight of the ions. From these parameters, the momentum, scattering angle, and particle identity (PID) can be determined. These focal-plane detectors are briefly described below. Figure 3.5: Schematic of the S800 analysis line and spectrometer. Figure adapted from [94]. 47 3.2.6 Focal Plane Detectors The first particle detectors at the S800 focal plane [101] are the Cathode Readout Drift Chambers (CRDCs). There are two CRDCs separated by 1073 mm allowing for two horizontal and vertical position measurements. Each CRDC is filled with a mixture of 20% isobutane and 80% carbon tetrafluoride. As the particle passes through the CRDCs, it ionizes the gas mixture, causing the freed electrons to drift towards an array of 224 cathode pads which lie along the dispersive (x) axis. As is illustrated in Fig. 3.6, the dispersive position can be determined by analyzing the image charge distribution on the pads. The non-dispersive (y) position is inferred from the drift time of the electrons in the detector. The drift time is determined from the difference between the S800 DAQ trigger time and the CRDC anode signal. Calibration details for the CRDCs are given in Chapter 4. From these two position measurements the dispersive and non-dispersive angles of the particle can be determined. This yields a total of four values: the dispersive position and angle (xfp and afp) and the non-dispersive position and angle (yfp and bfp). After the CRDCs, the particle passes through an ion chamber filled with a gas mixture (90% argon and 10% methane), where the energy loss of the beam is measured. As in the CRDCs the particle ionizes the gas mixture, which causes electrons to drift to one of 16 anodes, where the charge is read out. The amount of energy lost in the ion chamber is proportional to the square of the charge of the particle as given in the Bethe-Bloch formula [103],   4πe4 Z 2 2m0 v 2 dE = n z ln − ln 1 − − dx I m0 v 2 abs abs v2 c2  v2 − 2 c  , (3.3) where Z is the charge of the incident particle, nabs and zabs are the number density and atomic number of the absorbing material, m0 is the electron’s rest mass, I is the average ionization 48 Figure 3.6: Schematic of the two cathode readout drift chambers, with an example event trajectory passing through the detectors. The inset shows an example of a charge distribution detected by the pads. Figure originally taken from Ref. [101] and modified by Ref. [102] potential of the absorbing material respectively, and v is the velocity of the incident particle. Hence, by measuring the energy loss, the square of the atomic number can be determined after correcting for the difference in velocity across the focal plane. The final detector at the S800 focal plane was a 5-mm thick plastic scintillator, referred to as the E1 scintillator. It serves as the trigger for the S800 data acquisition and gives time and energy readouts from the PMTs attached at each end of the detector. The E1 scintillator measures the TOF of the outgoing ejectiles in the 16 C(p,n)16 N reaction with reference to the object scintillator described below. The TOF from the object scintillator to the focal plane is proportional to the mass number of the ejectile and is used with the energy-loss measurement to determine the particle identification at the S800 focal plane. The object scintillator is a thin plastic scintillator that is placed in the path of the beam in the object box of the S800 analysis line (see “object” in Fig. 3.5). For e10003, a 1127 µm 49 thick object scintillator was used. To ensure high-precision timing measurements could be made with these detectors and the neutron detectors, the signals from the object and focal plane scintillators were split and sent into both the S800 DAQ and the LENDA/VANDLE digital data acquisition system. 3.2.7 Liquid-Hydrogen Target The reaction target in e10003 was the Ursinus liquid-hydrogen target. It had an average thickness of approximately 66 mg/cm2 and a radius of 35 mm. The temperature was kept at approximately 19 K, at a pressure of just above 1 atm. Two 125 µm Kapton foils contain the liquid in the target. Fig. 3.7 shows the measured phase transition from gas to liquid, when the target was filled and cooled at the beginning of the experiment. The temperature and pressure were monitored throughout the experiment and any fluctuations of the density were corrected for on a run-by-run basis. Details on the target thickness determination are given in Section 4.5.3. 3.2.8 LENDA and VANDLE The Low Energy Neutron Detector Array (LENDA) [90, 91] is an array of 24 BC-408 plastic scintillator bars built by the charge-exchange group at the NSCL. Each bar has dimensions of 4.5 x 2.5 x 30 cm. It is designed to measure the kinetic energy of neutrons via the timeof-flight method with kinetic energies from 100 keV-10 MeV and with a neutron kineticenergy resolution of approximately 5%. A Hamamastu H6410 photomultiplier tube (PMT) is attached directly to each end of the LENDA bars to read out the scintillation light. At 1-m distance, the laboratory scattering-angular coverage of a LENDA bar is 2.58 degrees, and 50 1140 1120 P [Torr] 1100 1080 1060 1040 1020 1000 980 960 19 19.5 20 20.5 21 21.5 22 22.5 23 T [Kelvin] Figure 3.7: Pressure-temperature diagram for the liquid hydrogen target showing the phase transition from gas to liquid. The red line indicates the expected relationship taken from [104]. its vertical position resolution, determined from the time difference of the top and bottom PMTs, is 6 cm. VANDLE, the Versatile Array of Neutron Detectors at Low Energy, is a similar set of neutron TOF detectors from the University of Tennessee [92, 93]. VANDLE is comprised of three different sized detectors, 3 x 3 x 60 cm (small) 3 x 6 x 120 cm (medium) and 5 x 5 x 200 cm (large), allowing different experimental layouts to be optimized to the energy resolution and range needed. For e10003, 24 of the small sized plastic scintillator bars were used. Similar to LENDA, VANDLE is able to measure low-energy neutrons with comparable energy resolution and detection efficiency. Both sets of detectors had the signals from each PMT sent into the Pixie DDAS System. Details about the digital system and the performance of LENDA are given in Section 3.3. 51 3.3 Digital Data Acquisition System The data acquisition system (DAQ) for LENDA was upgraded prior to the running of this experiment in order to utilize a digital data acquisition system (DDAS). This upgrade effort was a significant portion of the work that makes up this thesis. Using a digital system allowed the VANDLE detectors to be easily integrated with the new DAQ, since adding more detectors only meant using more DDAS modules. The following sections give a brief overview of DDAS and describe the capabilities achieved for LENDA with DDAS. 3.3.1 DDAS Overview The fundamental function of DDAS is to produce a digitized waveform, which is referred to as the trace. Each signal from the detector system is sent through a Nyquist filter and into a high-frequency analog to digital converter (ADC), where the analogue voltage is sampled to produce a digital waveform [105]. LENDA and VANDLE were instrumented with XIA’s Digital Gamma Finder Pixie-16 boards [105]. For each of the 16 channels on the board, 250 megasamples per second (MSPS), 14-bit flash ADCs were used to perform the digitization of the incoming detector signals. After capturing the waveforms, 4 field-programmable gate arrays (FPGAs) implement onboard trace-processing routines to extract both time and energy information for each pulse. The Pixie-16 modules were placed in a PXI/PCI crate, which was controlled by a National Instruments MXI-4 crate controller connected with fiber-optic cable to a dedicated computer. 52 3.3.2 Digital Timing Unlike an analog system, where timing and energy measurements are performed by two independent circuits, DDAS extracts this information simultaneously from the trace. This is accomplished for both time and energy through digital filtering algorithms, either on the board or in off-line analysis. For the 250-MSPS digitizers (a period of 4 ns between clock ticks) used in e10003, some form of interpolation scheme is needed to achieve timing resolution below the sampling frequency. This can be accomplished by selecting an appropriate function for the detector signal of interest and fitting that function to the waveform to extract a sub-clock tick time for each pulse. Prior to e10003, this method was successfully implemented for VANDLE where a detector-limited timing resolution of 0.6 ns was achieved over the full energy range using the fit function, 4 f (t) = αe−(t−φ)/β (1 − e−(t−φ) /γ ). (3.4) With the highest energy signals, a timing resolution of 0.5 ns was achieved [92]. The fitting method requires significant computational post processing and the creation of a library of appropriate functions for each detector signal. An alternate choice—and the one that was utilized in e10003 for both LENDA and VANDLE—is to use a digital constant fraction discriminator (CFD) algorithm similar to the one implemented on the Pixie boards [105, 106]. Higher-order interpolations of the CFD are done off line in software that are unavailable on the board. The implementation for this work is as follows. The trace is processed through 53 a symmetric trapezoidal filter, F F [i] = i X j=i−(Tr −1) T r[j] − i−(Tr +Tg ) X T r[j] (3.5) j=i−(2Tr +Tg −1) where Tr is referred to as the rise time, Tg the gap time and T r the trace points. The filtered trace is used to form a digital CFD, CF D[i + D] = F F [i + D] − w F F [i] (3.6) where D is a user specified delay and w is a scaling factor [105, 107]. The units of Tr , Tg , and D are in clock ticks and w is a unitless scaling factor. The sub clock-tick timing information is then determined from a linear interpolation of the CFD’s zero crossing (this technique will be referred to as the Linear Algorithm). In Fig. 3.8, an example of this procedure is shown. The filled circles are a trace for a gamma ray from a 22 Na source generated by one photo multiplier tube (PMT) of a LENDA bar. The filled squares are its corresponding CFD signal with filter parameters Tr = 6, Tg = 2, D = 4, w = 5/8. 3.3.3 Intrinsic Timing Resolution of DDAS One of the significant challenges for high-precision timing with DDAS is the occurrence of non-linearities in the system’s response to fast scintillator pulses [92, 106]. This behavior makes it difficult to rely on a simple linear interpolation of the CFD’s zero crossing. This non-linearity is maximized when the arrival time of a signal places its zero crossing in between two clock ticks, which for the 250 MHz digitizers used here is a time difference of 2 ns. To explore this behavior for LENDA, a 22 Na source was placed on a LENDA bar where the 54 ADC Units 20000 Trace Interpolation 10000 CFD 0 -10000 -20000 -30000 40 50 60 70 80 90 100 Clock Ticks Figure 3.8: Example of digital trace from LENDA (filled circles) with overlaid digital CFD (filled squares). The thick line is the third-order interpolation function calculated for this event. signal from one PMT of the bar was split into DDAS with a 2 ns delay added to one copy of the signal. Signals from both the 511 and 1275 keV gamma rays from 22 Na were used in the following analysis. The CFD algorithm described above was performed on each signal with the filter set Tr = 2, Tg = 0, D = 4, and w = 1/8. This parameter set corresponds to the optimum choice for the Cubic Algorithm (described below) and is one example of the linear-interpolation algorithm failing in the presence of a non-linear CFD signal. Figure 3.9 shows the time difference between these two signals after correcting for the dependence of the time difference on signal amplitude (a phenomenon known as “walk”). The walk correction was done by plotting the pulse height against the time difference and correcting for any observed correlation of the time difference on the pulse height. The pulse heights in this data set ranged from approximately 4% to 58% of maximum ADC value. The linear interpolation with the above parameter set (labeled in the plot as “Unoptimized Linear”) 55 introduces systematic biases and fails to achieve good timing resolution. However, the shape of the digital CFD signal in the neighborhood of the zero crossing is a function of the CFD filter parameters. By performing an optimization of those parameters, significant improvement can be achieved in the performance of the linear-interpolation method. As the parameters for the CFD algorithm are discrete, a brute-force approach was used for the optimization. Each trace in the data was processed with every possible filter combination in the ranges Tr = 1 − 8, Tg = 0 − 3, D = 1 − 8, w = 1/8 − 7/8 (Tr , Tg , D were limited to integer values, while w was limited to steps of 1/8). The parameter set that achieved the best resolution was taken as the optimum set. The result of that optimization (Tr = 6, Tg = 2, D = 4, and w = 5/8) is shown in Fig. 3.9, where it is labeled as “Optimized Linear.” Significant improvement is made in the linear method after the optimization, but Counts as Fig. 3.9 shows, there is still some systematic bias in the distribution. 25000 Unoptimized Linear Algorithm Optimized Linear Algorithm 20000 Optimized Cubic Algorithm 15000 10000 5000 0 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 Time Difference [ns] Figure 3.9: Time-difference spectra from a split-signal electronics test. The time differences from the 3 different timing algorithms have been shifted to zero for comparison. Further improvements can be made through the use of a higher-order interpolation of the 56 CFD signal’s zero crossing. Here the two points before the zero crossing and the two points after are used to calculate a third-order polynomial interpolating function. An example of this is shown in Fig. 3.8 as a thick red line. Given the clock ticks and corresponding ADC values, the coefficient for the interpolating function are found by inverting the following equation,  3 (x1 )   (x2 )3    (x3 )3   (x4 )3     (x1 (x1 (x1  c3  y1          2 1 0     (x2 ) (x2 ) (x2 )   c2  y2    =           (x3 )2 (x3 )1 (x3 )0   c1  y3      c0 (x4 )2 (x4 )1 (x4 )0 y4 )2 )1 )0 (3.7) where x1,2 and y1,2 are the two points before the zero crossing and x3,4 and y3,4 are the two points after. The same optimization procedure as described above was used for this method (referred to as the Cubic Algorithm). The result of this optimization (Tr = 2, Tg = 0, D = 4, and w = 1/8) applied to the same data set as above is also shown in Fig. 3.9, labeled as “Optimized Cubic.” It is clear that the resolution has improved substantially using the cubic interpolation scheme and any systematic biases in the time difference spectrum have been removed. The intrinsic resolution results using the four points and a third-order polynomial were much better than the expected resolution of the detector (∼420 ps) Therefore, higher-order or more complicated interpolating schemes were not pursued. Fig. 3.10 shows the full-width at half-maximum (FWHM) electronic timing resolution of DDAS from the same data shown in Fig. 3.9. The width of the time difference distribution is plotted as a function of pulse height for both the optimized cubic- and linear-interpolation algorithms. Figure 3.10 shows that the Cubic Algorithm performs significantly better at low pulse amplitudes, while both techniques provide electronic resolution well below the expected 57 FWHM [ps] performance of a LENDA bar at an appreciable fraction of the ADCs range. 1400 Linear Algorithm 1200 Cubic Algorithm 1000 2ns Delay 800 600 400 200 0 0 0.1 0.2 0.3 0.4 0.5 0.6 ADC Fraction Figure 3.10: Electronic timing resolution of DDAS as measured with a split signal from a LENDA bar. The upside down triangles utilize a linear-interpolation method while the filled circles use a cubic-interpolation method. See text for details. 3.3.4 Timing Resolution of LENDA with DDAS The timing resolution of LENDA with DDAS was measured using the two correlated 511 keV gamma rays generated after the positron decay of 22 Na. Two LENDA bars were placed on opposing sides of the 22 Na source at a distance of 1 m. Events were recorded in a 4-way coincidence between the PMTs of the two bars. The time difference between the two bars was calculated as: 1 1 Tdiff = (T1,top + T1,bottom ) − (T2,top + T2,bottom ) 2 2 (3.8) where the numerical subscripts refer to bar 1 and bar 2. The width of the Tdiff distribution, √ corrected for walk and divided by 2, is taken as the timing resolution of a single LENDA 58 bar. In the same way described above, the parameters for this data set were optimized for the best resolution, yielding Tr = 2, Tg = 2, D = 4, and w = 1/8 for the linear algorithm and Tr = 2, Tg = 0, D = 2, and w = 1/8 for the cubic algorithm. Fig. 3.11 shows the measured timing resolution in FWHM of one LENDA bar as a function of pulse amplitude and energy in keVee for the linear and cubic timing algorithms discussed above. Unlike the data presented in Section 3.3.3, the signals in this measurement arrive randomly across all time differences, instead of always at a time difference of 2 ns. Because of the intrinsic resolution of the detector, the difference in achieved resolution between the cubic and linear algorithms is less than above. However, an improvement when using the cubic algorithm is still achieved, reaching about 400 ps at the highest energies. The saturation of the resolution at approximately 400 ps is a result of the intrinsic properties of the detector, and is not from FWHM [ps] the electronics. 0 1400 Energy [keVee] 50 100 150 200 250 300 350 400 450 Linear Algorithm 1200 Cubic Algorithm 1000 800 600 400 200 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 ADC Fraction Figure 3.11: Measured timing resolution for LENDA readout with DDAS. Upside down triangles use a linear-interpolation method, while the filled circles use a cubic-interpolation method. 59 3.3.5 Energy Extraction Methods There are many algorithms for extracting energy from digitized data, especially for highprecision spectroscopy (for example, see Ref. [108]). In the case of a plastic scintillator array, where signal-amplitude resolution is expected to be limited, relatively straightforward methods can be utilized. With the waveform saved, the energy of the event can be easily determined from the maximum trace value or, as is done for the e10003 analysis, by integrating the digitized pulse. In Fig. 3.12, a trace from one PMT in LENDA is shown, taken with a 22 Na source. The beginning 24 clock ticks of the trace is used to determine the average baseline of the signal, while a window around the maximum point in the trace is used to find ADC Units the pulse’s amplitude. 1600 1400 1200 1000 800 0 20 Baseline Window 40 60 Event Window 80 100 Clock Ticks Figure 3.12: A trace from a PMT of LENDA from a 22 Na source. The first integration region is used to determine the average baseline of the signal, while the second integrates the pulse and determines the energy of the event. To determine the resolution, an energy spectrum for 22 Na was measured using a LENDA bar. The source was placed in the center of the bar, and the energy of the event was taken 60 as the geometric average of the top and bottom PMT’s pulse integration values, EBar = p ET EB . (3.9) The top panel in Fig. 3.13 shows the calibrated energy spectrum in keVee obtained from this measurement. The two Compton edges from the 511 keV and 1274.5 keV gamma rays are seen at 341 and 1062 keVee, respectively. The Klein-Nishina formula convolved with a Gaussian [109] was used to fit the two Compton edges in order to extract a FWHM resolution of 23.5% at 341 keVee and 16% at 1062 keVee. Further, it was found that the position of the Compton edge was at 63.2 ± 1.9 % of the maximum yield of the Compton spectrum, where the error was determined from the fitting of the Compton edges. The bottom panel in Fig. 3.13 shows an energy spectrum from a similar measurement using a 241 Am source. Here, the photo peaks from the low-energy x-rays at 26.3 and 59.5 keVee can be seen. The photo peaks were fit with Gaussians to obtain a FWHM energy resolution of approximately 42.3%. 61 Counts Na Source 22 103 102 10 1 Counts 0 200 400 600 800 1000 1200 1400 16001800 2000 Light Output [keVee ] Am Source 241 1400 1200 1000 800 600 400 200 0 0 20 40 60 80 100 120 140 Light Output [keVee ] Figure 3.13: Top panel shows a calibrated energy spectrum from 22 Na seen in a LENDA bar. Bottom panel shows a calibrated energy spectrum from a 241 Am source. 62 3.3.6 KeVee to Tn The correspondence between keVee and neutron kinetic energy in keV can be measured using a 252 Cf fission source. This measurement was performed with an EJ-301 liquid scintillator placed next to the 252 CF source, which was positioned 1 m from a LENDA bar. The liquid scintillator served as the start signal for a neutron time-of-flight (TOF) measurement with the LENDA bar, where a 3-way coincidence between the two PMTs of the LENDA bar and the PMT of the liquid scintillator was required. Fig. 3.14 shows a result of this measurement, where light output in keVee from neutrons in LENDA versus their kinetic energy determined Light Output [keVee] from time-of-flight is presented. 4000 3500 3000 2500 2000 1500 1000 500 0 0 1 2 3 4 5 6 7 8 9 10 Tn [MeV] Figure 3.14: Scatter plot of light output against neutron energy from a 252 Cf source measured with LENDA and a EJ-301 liquid scintillator as the timing reference. The cuts on maximum light output are applied in this figure. For high neutron energies (>3 MeV), the relationship between maximum light output (full energy of the neutron deposited in the bar) and kinetic energy is linear. The relationship 63 approximately follows, Lmax (E) = 518.1 Tn − 499.5, Tn > 3 MeV (3.10a) where Lmax (E) is the maximum light output in keVee and Tn is the kinetic energy of the neutron in MeV. This relationship is non-linear at low neutron energies where it follows, Lmax (E) = 18.53 + 95.08 Tn + 81.58 Tn2 , Tn < 3 MeV. (3.10b) To remove background events, cuts on the light output of the neutron detectors following equations 3.10a and 3.10b were made in the analysis of e10003. 3.4 Implementation with the S800 Confident with the internal performance of DDAS, both LENDA and VANDLE were instrumented with the Pixie-16 250 MSPS modules for e10003. This section will discuss how the clock synchronization between the two systems was achieved and provide some details of the coincident trigger logic. 3.4.1 Clock Synchronization A 50 MHz clock was taken from the backplane of the DDAS crate, downscaled to 12.5 MHz and sent into a JTEC XLM72 module in the VME crate of the S800’s DAQ. Since this clock was from the backplane of the crate, it was a down-scaled version of the internal fast ADC clocks that perform the digitization of the detector signals and assign the events’ time stamps. To achieve the synchronization two specialized DDAS modules were used. The first 64 was inserted in the front of the crate and served as a breakout module for the clock. It is shown in the bottom of Fig. 3.15, labeled as clock breakout module. The second was a clock extraction module connected to the backplane of the crate behind slot 2. Slot 2 contained the master Pixie module, which distributed the clock to other Pixie modules over the backplane. The clock extraction module connected to the clock breakout module with a Cat5 cable, as indicated in Fig. 3.15. For each event, the scaler from the clock signal was latched and readout into the S800 data stream, allowing for a common time stamp between the two systems at a granularity of 12.5 MHz (80 ns intervals). This level of synchronization is only used for event correlation between the two systems, so it does not need the precision timing resolution achieved for the neutron TOF. Prior to the experiment, the setup was tested for stability and was found to remain synchronized for greater than 12 hours. Under the firmware available at the time of the e10003 (March 2015), the internal clocks of each channel would pick up a random 1 clock tick phase each time they were reset (the default of the readout system at the beginning of each run). Further, reseting the clocks in DDAS would mean that the clock in the S800 would have to be cleared and the two systems synchronized again. These issues motivated finding a way to reset these clocks as infrequently as possible. As will be discussed in Section 3.4.2, this complicated the setup. 3.4.2 Operation Infinity Clock The goal of Operation Infinity Clock was to allow the internal clocks of each DDAS channel to count continuously during the whole length of the experiment, without a reset at the beginning of each run. To achieve this, the DDAS readout program was edited such that the firmware call to reset the modules was only called once when the first run was started. The 65 S800 DAQ software was also modified to allow the scaler which keeps track of the DDAS clock, to remain uncleared between runs. Therefore, the clocks had to be reset only once at the beginning of the experiment and any time the DAQ crashed (which unfortunately occurred several times, requiring new timing offsets to be calculated for every channel). When the DDAS system was reset, it would take approximately 10 seconds before the clocks were cleared and the system entered a running state. During this time, it was necessary to ensure that the S800 system did not receive any triggers or clock cycles, so that only when the DDAS system reset would the S800 begin processing timestamps and events. This was accomplished using a specialized module that output a constant signal when DDAS was in a running state. This module is labeled “DDAS State Module” in the diagram show in Fig. 3.15. This module was connected to the master DDAS module in slot 2 of the crate by a cat5 cable. This signal was used as a veto for: 1. The clock signal from DDAS (which is output continuously) 2. The LENDA/VANDLE master trigger to the S800 3. The S800 DAQ When starting the first run, the veto of the clock was necessary since the S800 could not start counting the clock signals until DDAS reset its internal clocks and switched to a running state. The veto of the LENDA triggers and the S800 DAQ were necessary to ensure that no events happened in the S800 DAQ before the clock signal was ready. Reseting the two systems in this way synchronized the clocks and ensured that no events were triggered until both systems were ready. However, since the clocks needed to be free running after they were synchronized in the first run, the veto for the clock signal needed to be removed during the first run. This way, 66 the veto is not reapplied when the DDAS system stops taking data at the end of the run, and the two systems remain synchronized. To accomplish this, the veto cable for the clock was patched to the data-U and then back down to the experimental setup. This way, during the first run, the cable could be unplugged, thereby sending the clock signal to the S800 DAQ regardless of what state the DDAS crate was in. Experimental shift takers were given detailed instructions about the role of this “magic veto cable” to ensure the clock remained synchronized during the experiment. 3.4.3 Trigger Logic In order to select charge-exchange events from the 16 C(p,n)16 N reaction, both a neutron and a fast ejectile needed to be detected. To achieve this, the DAQ trigger logic was constructed to only record coincidence events between the two systems. For LENDA and VANDLE, the definition of a good trigger occurred whenever any bar had a signal in both the top and bottom PMT, within a 24-ns coincidence window. This initial check was conveniently performed internally by the FPGAs on the Pixie boards. The 24-ns coincidence window was chosen because it was the smallest value where the behavior of internal coincidence was reliable (at 8 and 16 ns, some coincidences would miss validation). The OR of all of the top-bottom coincidence in a given Pixie module was output from the Pixie breakout modules (this happend on the channel labeled o7 ). The breakout modules have lemo connectors for the available trigger inputs and outputs of the Pixie boards and are labeled as PIXIE 16 Breakout modules in Fig. 3.15. The outputs (o7 ) of all 8 Pixie modules were or-ed together to form the LENDA and VANDLE master trigger. This master trigger was sent up to the S800 DAQ where it was input as the secondary trigger to the S800 trigger box. If a coincidence was found, the S800 DAQ would send a 67 live trigger signal back down to LENDA and VANDLE. During the time that the initial neutron master trigger went up to the S800 DAQ and back down, the Pixie boards buffered the detector signals (the delay was approximately 1.2 µs). The live trigger was sent back into the boards external validation input (i4 ) on the breakout modules, which triggered an internal validation window of 700 ns. The size of this window was chosen to be long enough to validate low-energy neutron events, while short enough to keep the rate of random coincidence from overwhelming the DAQ system. Under these conditions, a raw coincident event rate of approximately 500 per second was written to disk. The S800 singles rate was approximately 10k. 68 DDAS State Module NSCL#3316 3 ns TTL or NIM Clock to S800 fast patch 1 on 20 ns cable A C X Y 2 ns A C X Y DDAS Trig to S800 fast Patch 2 on 20 ns cable A C X Y A C X Y A B C D X X A B C D + + S L A B C D X X A B C D + + S L A B C D X X A B C D + + S L A B C D X X A B C D + + S L A B Patch #108 DATA-U6 D X 1 ns Y B X Patch #107 DATA-U6 Y 1 ns D C X Y A C X Y B A D C X X 1 ns Y Y A D C X X 1 ns A B C D + + S L A B C D X X A B C D + + S L A B C D X X A B C D + + S L B 1 ns D 1 ns o7 B 1 ns D 1 ns Y A B C D X X A B C D + + o7 S L OUT 2ns o7 Out B 2 ns D 2 ns o7 o7 ~30 ns Y 2 ns D 2 ns 2 ns Obj->DDAS Obj->DDAS ~30 ns Comp Phillips 726 NSCL#2965 69 PIXIE-16 250MHz Clock breakout module To DDAS Clock mod on 5 ns cable 5 ns 1 2 3 4 5 6 7 x2 for FP signal DDAS breakout module i4: Validate DDAS 5 ns i4 To NSCL#2965 Mod 0 From NSCL#3261 Obj from patch 71 Norm 1 ns S800 Sync fast patch3 on 20 ns cable Out 2 ns Patch #110 DATA-U6 Magic VETO Patch #109 X B 2 ns ~8 ns PIXIE 16 Breakout Module Obj->DDAS Comp Y Y 4-Way Splitter NSCL#1166 Obj->DDAS o7 X X QUAD FAST AMP (Green face) NSCL#0679 IN S800 live trigger from fast patch 4 on 20 ns cable X Y FAN IN / FAN OUT MODULE (1X16) NSCL#3261 LVTTL o7 o7 4x4 Logic 21X6421 P-1 NSCL#0416 4x4 Logic 21X6421 P-1 NSCL#1341 TTL NIM NIM 2 ns B Y A B C D X X LeCroy 688AL NSCL#0439 o7 A B A B C Cat5 cable to DDAS Back plane (to 2nd port from top) x8 A B C S/N # 1047 | 1039 | 1043 | 1035 | 1040 | 1042 | 1036 | 1044 Figure 3.15: Detailed diagram of the electronics setup in experiment e10003. The NSCL numbers refer to the NSCL E-Pool number for the module depicted. Care has been taken to draw the modules with enough detail to make them easily recognizable. The positions of the switches depicted in the first two modules on the left correspond to their actual setting in the experiment. 69 Chapter 4 Data Analysis Have you tried looking at the traces? Remco Zegers The purpose of this chapter is to discuss how the raw data from e10003 was processed and dσ ). To accomplish this, a ROOT based analyfiltered to arrive at differential cross sections ( dΩ sis package was developed to unpack, build, and analyze the experimental data. Throughout the data, analysis the different “reaction channels” are discussed. This language refers to the different final ejectile particles that are possible in the 16 C(p,n)16 N reaction (see Table 3.1). In some of the reaction channels, there is more than one rigidity setting of the S800 spectrograph used in the analysis of that channel. These parts of the data are referred to as “settings” or “rigidity settings”. 4.1 Calibration This section describes the details of the calibrations procedures used in e10003 for the neutron detectors and the focal plane detectors of the S800 spectrograph. 4.1.1 Neutron Detector Timing Calibrations Since the basis of the (p,n) measurement is the TOF of the neutrons, the calibration of the timing signals from the neutron detectors is crucial. The first step in the calibration 70 procedure is to choose a start time for the neutron TOF. With the digital system, each signal in the setup has an absolute timestamp, which allows any signal to be chosen as the reference time for the TOF after the data taking. Originally, the in-beam scintillator located in the S800 object box was going to be used for this purpose. However, as the analysis progressed it became clear that this detector suffered significant radiation damage due to the high incident beam rate. This deteriorated the detector’s timing properties beyond what was acceptable to achieve the physics goals of the measurement. Instead, the timing of the beam-like ion measured by the E1 scintillator at the S800 focal plane was used as the reference time for the neutron TOF. This has the advantage of significantly lowering the incident particle rate (thousands instead of millions per second), preventing any radiation damage. However, using the S800 E1 scintillator introduces dependence of the neutron TOF on the dispersive angle and position in the focal plane. Figures 4.1 shows the correlation between neutron TOF, and the dispersive angle and the dispersive position. These events come from the experimental data, where 15 N has been selected in the S800 focal plane (see Section 4.1.3 for details about the particle identification). The dependencies can be corrected by looking at the events from the γ flash seen prominently in both figures. Since the γ events must all occur at the same time, the correlation can be fit and removed. After these corrections are performed the position of the γ flash is used to calibrate this time difference to reflect the actual TOF from the target to the neutron detectors. Since the distance from the center of the target to the bars is known to be 1 m, the time that the γ flash should appear can be easily calculated (1m/c=3.33-ns). 71 3 1000 afp [Deg] 2 800 1 0 600 −1 400 −2 200 −3 −4 −2 0 2 0 4 TOF [ns] (a) 300 250 250 200 xfp [mm] 200 150 150 100 100 50 0 50 −50 −100 −3 −2.5 −2 −1.5 −1 −0.5 0 0.5 1 0 TOF [ns] (b) Figure 4.1: Two dimensional histogram showing the correlation between the neutron TOF, and dispersive angle (a) and position (b) in the S800 focal plane. The γ flash is seen prominently in the figures. The neutron events are to the right, but are cut off to clearly show the γ flash. 72 In addition to the corrections for the focal-plane parameters, the neutron TOF must also be corrected for any walk (dependance of the TOF on the signal height in LENDA or VANDLE). Fig. 4.2 shows this dependence for events where 15 N has been selected in the S800 focal plane. 500 E [keVee] 400 300 200 100 0 −5 −4 −3 −2 −1 0 1 2 3 4 5 TOF [ns] Figure 4.2: Neutron TOF plotted against the calibrated light output from the neutron detectors. The γ flash can be seen centered around 0. At low energies, a small correlation can be seen for the γ events. Combining all of these effects provides the following expression for the neutron TOF, T OF = Tn − TE1 − a(af p) − b(xf p) − c − Tof f set , Lo (4.1) where Tn is the timestamp from LENDA or VANDLE, TE1 is the timestamp from the top PMT of the S800 focal-plane scintillator, af p is the dispersive angle in the focal plane, xf p is the dispersive position in the focal plane, Lo is the light output seen in LENDA or VANDLE in keVee, and a, b and c are correction coefficients. Fig. 4.3 shows the fully corrected TOF spectrum for the 15 N data. The γ events are seen 73 cleanly separated from the neutron events by TOF. A FWHM timing resolution of ∼700-ps is achieved in the experimental data. Similar calibration procedures were used for the other channels of interest 14 N, 15 C, 14 C, 16 N. Their TOF spectra are shown in Fig. 4.4. 50000 45000 γ 40000 N 15 35000 30000 25000 20000 15000 n 10000 5000 0 0 20 40 TOF [ns] 60 80 100 Figure 4.3: Fully Corrected neutron time of flight spectrum for the N15 data (3.1325 Tm setting). The dashed green line indicates the TOF cut between γ and neutron events. It corresponds to a neutron kinetic energy of ∼67 MeV, which is far above the energies used in the analysis. As can be seen in the Fig. 4.4, the γ flash for the 15 C channel has a long tail and what appears to be poor resolution. Since the target material is hydrogen, the γ rays must come from some excited state in the ejectile. However, below the neutron emission threshold, 15 C has only 1 known state at 740-keV. Since this state has a long lifetime of 2.61 ns, the γ flash appears smeared out. The 15 C can emit the γ ray later along its flight path, increasing the total distance it must travel to the neutron detectors. This effect was simulated and included in the timing calibration of the 15 C data. Fig. 4.4 also shows a sharp peak to the right of the γ flash in the 15 C and 14 C TOF spectra. These peaks come from a strong neutron-knockout background that dominates at 74 forward lab angles. Details of this background and its subtraction are given in Section 4.4. 4500 4000 3500 3000 2500 2000 1500 1000 500 0 0 20 40 60 TOF [ns] 80 200 180 160 140 120 100 80 60 40 20 0 100 0 (a) 14 N Events 14000 12000 10000 8000 6000 4000 2000 0 0 20 40 60 TOF [ns] 20 40 60 TOF [ns] 80 100 80 100 (b) 16 N Events 80 35000 30000 25000 20000 15000 10000 5000 0 100 (c) 15 C Events 0 20 40 60 TOF [ns] (d) 14 C Events Figure 4.4: Corrected neutron TOF spectra for the 14 N (2.9 Tm) (a), 16 N (3.38 Tm) (b), 15 C (3.6565 Tm) (c), and 14 C (3.38 Tm) (d) channels. The 15 C and 14 C data from the 3.52 Tm and 3.2828 Tm settings are not shown, but have similar TOF spectra. 4.1.2 LENDA and VANDLE Gain Calibrations Though the light output from the neutron detectors is not directly used in the determination of the excitation energy and COM scattering angle, it is necessary for setting the neutron detection threshold. The threshold affects the neutron detection efficiency and therefore the differential cross section. To accomplish this the light output for each bar is calibrated into 75 keVee using γ sources 241 Am, 22 Na and 137 Cs. The specific gamma rays used and their specifications are summarized in the Table 4.1. Source Eγ Type EγReduced 22 Na 1274.5 CE 1061.7 22 Na 511 CE 340.7 137 Cs 661.7 CE 477.4 241 Am 59.5 Ph 59.5 241 Am 26.3 Ph 26.3 Table 4.1: Table of the γ-ray calibration sources used for the neutron detector gain calibration. See text for details. Eγ is the true energy of the emitted gamma ray, CE refers to compton edge, Ph refers to photo-absorption and EγReduced is the expected energy of the gamma ray in the neutron detectors. For photo-absorption the expected energy is unchanged, but for compton edges the energy is reduced (see Appendix B for more details). The gain for the top and bottom PMT of each bar were matched in software using the compton edge of 137 Cs. The total energy of the bar was then defined as, E= p ET EB , (4.2) where ET,B are the gain matched energies for the top and bottom PMTs. The energy for the bar was then calibrated into keVee using the compton edges from 22 Na and the photo-peaks from 241 Am, E Cal = g · E + T h, 76 (4.3) where g is the gain and T h is the threshold from the fit. 1400 E [KeVee] 1200 1000 800 600 400 200 0 0 200 400 600 800 1000 1200 E [arb] Figure 4.5: Example light output calibration for south LENDA bar 2. The data points are described in Table 4.1. The red line is a linear fit to the data. Fig. 4.5 shows an example of the calibration curve from a LENDA bar (south LENDA 2) using the 4 points from 241 Am and 22 Na. The neutron detection threshold is set using this calibrated energy for each bar. The uncertainty in the calibrated energy from fitting uncertainties and drifts in the gain during the experiment are included in the final systematic error for the differential cross section. 4.1.3 Particle Identification Though the experimental conditions are tuned to explore charge-exchange reactions, many other processes can trigger the DAQ by producing a neutron and an ejectile in the 16 C + p reaction. To filter away the events from these processes the timing and energy loss information from the E1 scintillator and the ion chamber in the S800 are used to make a particle identification plot. This allows only the ejectile of interest (16 N, 15 N, 14 N, 15 C, 14 C) 77 to be included in the data analysis. Fig. 4.6 shows an uncalibrated particle identification plot for the 3.1325 Tm rigidity setting (the main setting for 15 N), where the TOF of the ion to the S800 focal has been shifted to near zero for convenience. 1600 1400 ∆ E [arb] 1200 1000 800 N 13 C 15 600 400 200 0 −30 −20 −10 0 TOF [ns] 10 20 30 Figure 4.6: Uncalibrated Particle Identification for the 3.1325 Tm setting. While even at this level separation between different particles can be seen, the resolution of both the TOF and energy loss signals can be improved by correcting for the dependence on dispersive position and angle in the focal plane. Fig. 4.7 shows these dependencies for the TOF when the 15 N events are selected in the uncalibrated PID spectrum shown in Fig. 4.6. Similar correlations and corrections exist for the energy loss measured in the ion chamber. After applying these corrections a particle identification plot is constructed for each rigidity setting. Fig. 4.8 shows a corrected PID spectrum from 4 of the rigidity settings measured in the experiment, each of which highlights an exit channel of interest. Elliptical gates are made on each particle extending out to 2 sigma in TOF and ∆E. The good events cut away in this process are included in the final normalization of the differential cross section. These gates are shown as black ellipses in Fig. 4.8. 78 3 afp [Deg] 2 1 0 −1 −2 −3 −3 −2 −1 0 1 2 3 TOF [ns] (a) 300 250 xfp [mm] 200 150 100 50 0 −50 −100 −150 −200 −4 −2 0 2 4 TOF [ns] (b) Figure 4.7: Correlation between the ejectile TOF and the dispersive angle (a) and position (b) in the S800 focal plane for 15 N events. The artifacts seen in (b) are the result of bad events in the CRDCs, which are removed in subsequent steps of the analysis. 79 (a) (b) (c) (d) Figure 4.8: Corrected PID spectra from the 3.1325 Tm (a), 2.9 Tm (b), 3.6565 Tm (c), and 3.38 Tm (d) settings. The main particles of interest in each setting are indicated. The remaining two settings 3.52 Tm and 3.2828 Tm are not shown, but are similar to the 3.6565 Tm and 3.38 Tm settings. 80 4.1.4 CRDC Calibration In order to find the particle’s position and angle from the pad distribution and electron drift time measured by the CRDCs, the detectors must be calibrated. To accomplish this a metal plate (a “mask”) with a series of known holes and strips is placed 8 cm upstream of each CRDC for a calibration measurement. The result of these calibration runs are shown in Fig. 4.9, where the mask hole and strip pattern is clearly visible. Taking the measured drift time and pad position of several of these holes a linear calibration of the form, X1,2 (mm) = sx1,2 (mm/pad) · X(pad) + cx1,2 (mm) y y Y1,2 (mm) = s1,2 (mm/ns) · Y (ns) + c1,2 (mm), (4.4) (4.5) x,y can be fit to the data. The s1,2 coefficients refer to the slope of the calibration for x and x,y y of CRDCs 1 and 2. Similarly C1,2 are the offsets from these calibrations. Since the x measurement comes from discrete PADs, the slope for that calibration is just the width of the pads, 2.54 mm/pad. The constant offset can be found by using any of the holes seen in Fig. 4.9. For the Y direction, 3 vertical points in the center of the mask pattern were used in the fit. To accurately determine the x position from the PAD distribution, the gains of the PADs must be matched and the resulting distribution fit on an event-by-event basis. If the gains of some PADs are substantially larger than the others, the fits can be biased towards those PADs, creating artifacts in the spectrum. Fig. 4.10 shows the effect of this calibration step for CRDC 1. Once the positions for both CRDCs are fully calibrated, the dispersive and non-dispersive 81 3000 40 2500 35 30 2000 TAC 25 1500 20 1000 15 500 10 5 0 0 50 100 150 200 250 300 0 Pad Number (a) CRDC 1 35 2500 30 2000 25 TAC 3000 20 1500 15 1000 10 500 5 0 0 50 100 150 PAD Number 200 250 300 0 (b) CRDC 2 Figure 4.9: The uncalibrated position (x-axis) and time (y-axis) measured by the CRDCs during the mask runs. 82 5000 4500 50 4000 E [arb] 3500 40 3000 30 2500 2000 20 1500 1000 10 500 0 0 20 40 60 80 100 120 140 160 180 200 220 0 PAD Number (a) 5000 4500 50 4000 E [arb] 3500 40 3000 30 2500 2000 20 1500 1000 10 500 0 0 20 40 60 80 100 120 140 160 180 200 220 0 PAD Number (b) Figure 4.10: The pad distribution for CRDC 1 before (a) and after (b) the pad-by-pad gain matching. The data is from one experimental run of the 3.1325 Tm rigidity setting. 83 angles can be calculated, af p = atan((X2 − X1 )/1073.0) (4.6) bf p = atan((Y2 − Y1 )/1073.0), (4.7) where 1073.0 is the distance between the CRDCs in mm. This defines 4 parameters for the particle’s track in the focal plane. However, determining the particle’s properties at the target position is often a more critical piece of information. Specifically for this measurement, the ejectiles momentum distribution is needed so that the cross section can be corrected for the momentum acceptance of the spectrometer, and the different rigidity settings can be stitched together. To accomplish this the transfer map of the spectrograph is calculated using the ion optics code COSY infinity [110]. With the inverse map (M −1 ) calculated by COSY the target parameters can be derived,     xf p ata         af p yta     .   = M −1      yf p  bta          dta bf p (4.8) The “ta” (as opposed to “fp”) designation is a common convention to indicate that these are the values at the target position. Similar to the focal plane parameters ata (bta) is the dispersive (non-dispersive) angle at the target, yta is the non-dispersive position, and dta is the deviation of the particles’ energy relative to the energy of the center track through the spectrometer (E0 ). The kinetic energy of the ejectile can be determined from dta easily through, 84 E = E0 (1 + dta). (4.9) From the energy and angles, the momentum vector p~ can be determined. The ejectiles kinetic energy is used to investigate the acceptance of the spectrometer (see Section 4.5.2 for more information). With the above calibrations completed, we have the ingredients necessary to calculate the COM scattering angle and excitation energy for each event. The following sections go through the major analysis steps needed to extract absolute differential cross sections from the data. 4.2 Experimental Yield In order to calculate the COM differential cross section as a function of excitation energy in 16 N, the lab measurements must be transformed into the correct variables. For the (p,n) reaction in inverse kinematics this is done with missing-mass spectroscopy, where the excitation energy of 16 N can be found with, Ex = q 2 + P 2 − 2P P (Mp + M16 C + T16 C − Mn − Tn )2 − (P16 n 16 C Cos(θn )) − M16 N . n C (4.10) Mx , Tx and Px refer to the rest mass, kinetic energy and momentum of particle x in the lab frame respectively. θn is the lab scattering angle of the neutron. The momentum of the neutron and 16 C are given by, Px = q 2Mx Tx + Tx2 . 85 (4.11) To find the center of mass scattering angle the following relation is used, Θcm = Sin−1 ! Pnlab Sin(θn ) , Pncm (4.12) where the Pncm is found by Lorentz boosting the neutron lab momentum vector into the COM frame. The above relations were used to find the excitation energy and COM scattering angle for each event. The yields in the lab and COM are presented in Fig. 4.11. The plots on the left show a two dimensional histogram of lab neutron kinetic energy (y-axis) and neutron scattering angle (x-axis) for the different decay channels. The figures on the right show the corresponding two dimensional histogram of COM scattering angle (y-axis) and excitation energy in 16 N (x-axis). On the left, the expected kinematic relationships for the 16 N, 15 N, and 14 N channels are clearly visible. However, for the 15 C and 14 C channels there is little apparent structure. Instead the spectra are overwhelmed with forward peaking background events. These events come from neutron-knockout reactions that have the same outgoing particles as true charge-exchange reactions. The neutrons from these events scatter off the downstream quadrapole magnet and hit the neutron detectors, creating false coincidences. Section 4.4 will discuss this background in more detail. The data were separated into two-degree bins in COM scattering angle and the excitation energy histogramed to form the experimental yield. As an illustrative example, Fig. 4.12 shows the experimental yield in the Θcm = 4 − 6° bin for the 15 N channel. The acceptance of the neutron detectors is clearly visible in the spectrum, where it creates artificial peak-like structures. Also evident in the spectrum are background events that need to be subtracted. The following sections will discuss the neutron acceptance corrections and the background 86 Θcm Tn [MeV] 30 20 10 0 MeV 20 18 16 103 14 20o 12 10 102 8 15o 6 10o 4 2 5o 10 0 40 50 60 70 80 90 100110120130140 θn 20 18 16 14 12 10 8 6 4 2 0 −20 102 10 1 −10 10 20 30 Ex [MeV] (b) 15 N 30 20 10 0 MeV 20 103 18 16 14 20o 12 102 10 8 15o 6 10o 4 10 2 5o 0 40 50 60 70 80 90 100110120130140 Θcm (a) 15 N Tn [MeV] 0 θn 20 18 16 14 12 10 8 6 4 2 0 −20 102 10 1 −10 0 10 20 30 Ex [MeV] (c) 15 C (d) 15 C Figure 4.11: Two dimensional histogram of lab neutron angle and kinetic energy measurements (a,c,e,g,i), and reconstructed excitation energy in 16 N and COM scattering angle (b,d,f,h,j). In (a,c,e,g,i) the solid lines indicate excitation energy in 16 N and the dashed lines indicate the COM scattering angle. 87 Figure 4.11: cont’d (f) 14 C 30 20 10 0 MeV 20 18 16 14 103 20o 12 10 8 15o 102 6 10o 4 5o 2 10 0 40 50 60 70 80 90 100110120130140 Θcm Tn [MeV] (e) 14 C θn 20 18 16 14 12 10 8 6 4 2 0 −20 102 10 −10 10 20 30 (h) 16 N 30 20 10 0 MeV 20 18 16 10 14 o 20 12 10 1 8 15o 6 10o 4 10−1 5o 2 0 40 50 60 70 80 90100110120130140 Θcm Tn [MeV] 0 Ex [MeV] (g) 16 N θn 20 18 16 14 12 10 8 6 4 2 0 −20 10 1 10−1 −10 0 10 Ex [MeV] 20 30 (j) 14 N 30 20 10 0 MeV 20 18 16 102 14 20o 12 10 8 15o 10 6 10o 4 5o 2 1 0 40 50 60 70 80 90 100110120130140 Θcm (i) 14 N Tn [MeV] 1 θn 88 20 18 16 14 12 10 8 6 4 2 0 −20 10 1 −10 0 10 Ex [MeV] 20 30 10−1 subtraction methods used for the different channels. 6000 5000 Yield 4000 3000 2000 1000 0 −5 0 5 10 15 20 25 30 Ex [MeV] Figure 4.12: Experimental yield for the 15 N channel (3.1325 Tm) in the Θcm = 4 − 6° bin. 4.3 Neutron Detector Efficiency and Acceptance Of significant importance to determining the absolute cross section for the 16 C(p,n)16 N reaction is the intrinsic neutron detector efficiency and geometrical acceptance of the LENDA and VANDLE bars. As was done in previous work with LENDA, the efficiency for each bar is determined through a Monte Carlo simulation that has been validated against experimental measurement [63, 64]. For e10003 this was done using the GEANT4 simulation toolkit [111] and its associated high-precision neutron physics library [112]. This is the same approach used by the VANDLE collaboration [93] and the European Low Energy Neutron Spectrometer (ELENS) [113]. In order to be detected by LENDA or VANDLE the neutron must first interact with the hydrogen or the carbon nuclei in the plastic scintillator. A cartoon of this process is 89 show in Fig. 4.13. In principle the neutron can interact with any combination of hydrogen and carbon nuclei, however the light output from interactions with the carbon is highly quenched [113, 93]. These carbon reactions can therefore contribute only a small amount to the detector’s response in the neutron energy region of interest (< 20 MeV). The elastic scattering on hydrogen dominates the spectrum. Fig. 4.14 shows the light output for plastic scintillator in keVee for protons, 12 C, and 4 He as a function of incident particle energy (adapted from [114]). The quenching of the carbon’s light output can be clearly seen–a 1 MeV proton will result in a 177.7 keVee of light output but a 1 MeV 12 C will make only 6.6 keVee. This effect pushes the carbon interactions below the light-output threshold for all but the most energetic carbon scatterings. The low-energy tail of the light-output function for protons was modified to better reproduce the measured neutron response from the 252 Cf data presented in Fig. 3.14. This is shown as the dotted blue line in Fig. 4.14. These light-output curves were used in the simulation of the neutron detection efficiency. Fig. 4.15 shows the result of the Geant4 simulation, where neutrons between 0 and 20 MeV were impinged on a single LENDA bar from a point like source. The light output was determined in the simulation by taking the energy loss from each simulation step inside the LENDA bar and converting it to a light output using the functions shown in Fig. 4.14. The light output of all the steps are added together to arrive at a total light output for the simulated event. In the previous work with LENDA [63, 64], MCNP was successfully used to simulate the response of the neutron detectors. Using Geant4 has allowed all aspects of the simulation to be done in one framework. Within the simulation the beam profile, the bulging of the target windows, the energy loss of the beam in the target, and the neutron detector response are 90 p H(n,n) n 12C(n,n) n 12C Figure 4.13: A cartoon depicting the interactions of neutrons in a plastic scintillating material. In order for the neutron to be detected it must undergo a nuclear reaction with the hydrogen or carbon nuclei in the material. The angles of the scattering are exaggerated for artistic purposes. The dominate elastic scattering on hydrogen is shown at the top of the figure. Elastic (shown at the bottom of the figure) and inelastic (not shown) scattering on 12 C are also possible but highly quenched. See text for details. 91 105 104 Modified Proton Light [keVee] Proton 3 10 2 10 12 C Alpha 10 1 10−1 10−2 −2 10 10−1 1 E [MeV] 10 Figure 4.14: Light output curves for BC-400 plastic scintillator. Each line shows the expected light output as a function of incident particle energy. Adapted from [114] see text for details. Figure 4.15: Two dimensional histogram of light output and incident neutron energy from the Geant4 simulation. Contributions from elastic scattering on hydrogen and 12 C are included in the simulation. Inelastic channels for carbon scattering are also included and create the few events above the elastic scattering trend line. These events amount to a tiny portion of the data (<0.05%). 92 all included at the same time. To obtain at a simulated efficiency for the LENDA bar, a light output threshold must be applied in the simulation to match the data. The threshold for the data was set at 30 keVee as determined for each bar following the calibration procedure outlined in Section 4.1.2. An effective threshold of 25 keVee and a scaling factor of 0.92 was applied to the simulation since it better reproduced the measured efficiency at low neutron energies. Fig. 4.16 shows the result of the Geant4 simulation with the measured efficiency overlaid at a threshold of 30 keVee. This efficiency was measured using a 252 Cf fission source as described in Ref. [90]. 0.5 0.45 Efficiency 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0 0.5 1 1.5 2 2.5 3 3.5 4 Tn [MeV] Figure 4.16: The black points show the measured intrinsic neutron detection efficiency for a single LENDA bar at a light output threshold of 30 keVee. The blue line is the result of the Geant4 simulation of the efficiency. Good agreement is seen. Confident with the simulations ability to reproduce the intrinsic neutron detector efficiency, the simulation was used to calculate the total neutron acceptance as a function of COM scattering angle and excitation energy for the full experimental setup. The beam was simulated through the bulged liquid hydrogen target, where it underwent a charge-exchange reaction. When the scattered neutron hit one of the detectors, the event was recorded. The 93 result of this simulation is shown in Fig. 4.17, where the neutron energy and angle have smeared to match the experimental resolution and transformed into the COM frame. 40 5000 35 Θ cm [Deg] 30 4000 25 20 3000 15 2000 10 1000 5 0 −10 −5 0 5 10 Ex in 15 16 20 N [MeV] 25 30 35 40 0 Figure 4.17: Two dimensional histogram of Θcm and Ex in 16 N from the Geant4 simulation of the e10003 setup. The total detection efficiency as a function of COM scattering angle and excitation energy is obtained by taking projections of Fig. 4.17 onto the x-axis in the same COM scattering angle binning used in the analysis and dividing by the total number of events simulated in that bin. Figs. 4.18 shows the simulated efficiency for the most forward (Θcm =4-6°) and most backward (Θcm =14-16°) angles used in the analysis. The sharp, periodic features seen in both figures are the result of measuring the lab scattering angle with discrete bars of differing geometrical acceptance–the VANDLE bars cover approximately twice the solid angle as the LENDA bars. These simulated curves were used to correct the experimental yields for the acceptance in each COM scattering angle bin used in the analysis. The simulation was also used to determine the experimental excitation-energy resolution of the full setup. This was done by simulating the response of the neutron detectors to 94 0.05 0.045 Efficiency 0.04 0.035 0.03 0.025 0.02 0.015 0.01 0.005 0 −5 0 5 10 15 Ex in 16 20 25 30 35 40 25 30 35 40 N [MeV] (a) 0.05 0.045 Efficiency 0.04 0.035 0.03 0.025 0.02 0.015 0.01 0.005 0 −5 0 5 10 15 20 Ex in 16N [MeV] (b) Figure 4.18: Simulated neutron acceptance and efficiency in the Θcm =4-6 ° (a) and Θcm =1416° (b) bin for the e10003 setup. 95 events with discrete excitation energies. As was seen in Ref. [64], the resolution worsened at large COM scattering angles. The resolution varied from σ(5°)FWHM ≈ 750 keV to σ(15°)FWHM ≈ 2 MeV. The following section describes the background subtraction used in the analysis. 4.4 Background Subtraction There are various possible background sources that can contribute to the charge-exchange spectrum. They can be categorized as: ˆ Foil background. There may be reactions off of the non-hydrogen nuclei in the Kapton foils that surround the liquid-hydrogen target. ˆ Random background. The TOF spectrum may have random coincidences between neutrons and uncorrelated beam bunches. ˆ Beam-induced background. Reactions on the hydrogen other than charge exchange that produce the correct ejectile of interest and a neutron. The foil induced background was studied by taking data with the empty target at the conclusion of the experiment. This was done in all of the rigidity settings used in the experiments. The contribution from these events was found to be vanishingly small (< %0.5) and could not be determined accurately with the statistic of the background runs. Instead the thickness of the target was increased to include the hydrogen nuclei in the Kapton foils and a systematic uncertainty of 0.5% was included. The random background was inspected by looking at events that occurred at negative TOFs. The rate of random coincidences was also found to be very small (< 1%). Similar to the foil background, the random background 96 shape could not be accurately determined and was included as a systematic error. The choice to use the S800 E1 scintillator for the neutron TOF made the random rate far less than in previous experiments when an in beam scintillator was used [63]. See Section 4.7 for a breakdown of the systematic errors. The vast majority of the background events came from knockout or fragmentation reactions on the hydrogen in the target that produced the correct ejectile and a neutron. These reaction types are associated with fast neutron emission at forward scattering angles. These neutrons can scatter off the surrounding materials and hit a LENDA or VANDLE bar, sometimes creating a false coincidence. In this circumstance the TOF of the neutron is not related to its energy and creates a complex background. Since modeling the different background processes and the scattering of the neutrons from different materials in the experimental vault was a difficult task, a background model built off the experimental data was used. Different techniques were used for the different reaction channels since the background shapes and intensity were different. For the 15 N and 14 N channels, a background model was based on events that created 11 B in the spectrometer. Events that create 11 B cannot be from charge-exchange reactions in the excitation energy range of interest of 0-20 MeV (the 11 B channel would open at 42 MeV). Therefore they should give a good representation of the neutrons from the possible background processes. A similar technique was used in the first LENDA (p,n) experiment [63, 64]. Fig. 4.19 shows the two dimensional histogram of neutron energy and scattering angle for 15 N gated events and 11 B gated events. Even at the preliminary level the shape of the 11 B spectrum seems to reproduce the shape of the background seen in the 15 N data. A more detailed look at the 11 B model for the background reveals that some rescaling is needed to characterize the background at higher excitation energy. Fig. 4.20 shows the 97 60 3 50 10 40 102 30 20 10 10 Tn [MeV] Tn [MeV] 60 103 50 40 102 30 20 10 10 0 1 40 50 60 70 80 90 100110120130140 0 1 40 50 60 70 80 90 100110120130140 θn θn (a) 15 N (b) 11 B Figure 4.19: Two dimensional histogram of lab neutron scattering angle and neutron kinetic energy for 15 N (a) and 11 B (b) gated events. The events from 11 B are used to model the background seen in the 15 N channels. ratio of the 15 N and 11 B data integrated between neutron kinetic energy of 40 and 60 MeV as a function of lab scattering angle. The blue points show the region of θn that contains charge-exchange events, while the black points contain only background. A linear function was fit to the black points to yield a rescaling function that models the background seen in the 15 N channel. Using this model yields the background shape in Fig. 4.21, where the forward (Θcm = 4−6°) and backward (Θcm = 14−16°) most angles are shown. The background is normalized to the data in the unphysical region below the channel’s threshold (2.5 MeV for 15 N). The data presented in Fig. 4.21 still contains the jagged acceptance of the neutron detectors and so contains periodic structures. Building the background model from the experimental data reduces systematic error because mistakes in the acceptance of the bars are the same between the data and the background model. A similar procedure was performed for the 14 N data. The 11 B data was again used as the background model, but with a different rescaling function. 98 3.5 3 Ratio 2.5 2 1.5 1 0.5 0 40 60 80 100 θn [Deg] 120 140 7 6 5 4 3 2 1 0 −5 0 Data Background arb arb Figure 4.20: Ratio of the 15 N and 11 B gated events integrated from 40 to 60 MeV in neutron kinetic energy as a function of the lab neutron scattering angle. The blue points show the region of charge-exchange reactions, while the black points indicate regions with only background. The red line is a linear fit to the black points in the range shown. 5 10 15 20 25 30 35 40 Ex [MeV] 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 −5 0 5 10 15 20 25 30 35 40 Ex [MeV] (a) 15 N at Θcm =4-6° (b) 15 N at Θcm =14-16° Figure 4.21: Experimental yields with overlaid background shape from the 11 B models. The most forward (Θcm =4-6°) and backward (Θcm =14-16°) COM angles are shown. 99 Characterizing the background for the 15 C and 14 C channels proved to be difficult due to the large amount of neutron-knockout and fragmentation reactions. After many iterations, events from the low-energy tail of the ejectiles kinetic energy distribution was used to estimate the shape of the background. Fig. 4.22 shows the kinetic energy distribution for the 15 C channel. The full distribution is reconstructed from two rigidity settings: 3.52 and 3.6565 Tm. The gray line represent the full distribution after combining the two settings with corrections for incomplete acceptance. The spectrum contains both charge-exchange and background events. However, charge-exchange events up to 20 MeV of excitation energy cannot create the events seen in the low-energy tail of the distribution. This makes these events a good candidate for a background model. −6 25 ×10 20 arb 15 10 5 0 1250 1300 1350 1400 1450 1500 1550 1600 T15C [MeV] Figure 4.22: The kinetic energy spectrum of 15 C events detected in the S800 Spectrograph. The black (blue) hashed region comes from the 3.52 (3.6565) Tm rigidity setting. The gray line is the full distribution constructed from the two settings. The thin blue line shows the simulated distribution from charge-exchange reactions. The low-energy tail of the distribution is energetically inaccessible to charge-exchange reactions and so must come from background processes. Fig. 4.23 shows the data and the above background model for the forward and backward 100 most scattering angles gated on the 15 C and 14 C channels. In both cases, the signal to background is clearly much lower than the 15 N case. Both channels have limited yield for the Θcm = 4 − 6° bin with increased strength appearing at backwards angles. The statistics for the 15 C background model are clearly limited. However, this technique was the best compromise between statistical and systematic error for this channel. 6 Data Background 4 arb arb 5 3 2 1 0 −5 0 5 10 15 20 25 30 35 40 0.5 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 −5 0 Ex [MeV] Ex [MeV] (b) 15 C Θcm = 14 − 16° Data Background arb arb (a) 15 C Θcm = 4 − 6° 18 16 14 12 10 8 6 4 2 0 −5 0 5 10 15 20 25 30 35 40 5 10 15 20 25 30 35 40 Ex [MeV] 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 −5 0 5 10 15 20 25 30 35 40 Ex [MeV] (c) 14 C Θcm = 4 − 6° (d) 14 C Θcm = 14 − 16° Figure 4.23: Experimental yield and background models for the 15 C and 14 C channels. The data comes from the 3.6565 Tm and 3.38 Tm rigidity settings respectively. Yield is primarily seen at backward angles. The 16 N channel does not suffer from a complex background because the PID cut on 16 N ensures that no breakup or fragmentation reaction occurred. A flat background subtraction 101 was done based on a linear fit to the data at higher excitation energies where there is no strength in the 16 N channel. 4.5 Absolute Normalization Corrections The goal of e10003 was to measure absolute cross sections in the 16 C(p,n)16 N reaction. This meant measuring the beam rate incident on the liquid-hydrogen target and determining the efficiencies and acceptances of each detector and analysis step. The neutron acceptance and efficiency calculation has already been discussed in its own section (see Section 4.3). This section discusses some of the remaining corrections that were used in final cross-section normalization. 4.5.1 Incident-Beam Measurement The scaler readout of the object scintillator was used to monitor in the incoming beam. The measured rate varied between 2 and 3 MHz throughout the experiment. To determine the total number of particles incident on the scintillator in each run, the averaged measured rates were averaged together and multiplied by the length of the run. This gave a run-byrun tracking of the number of incident beam particles during the experiment. To check the reliability of this measure, the correlation between the object scintillator rate and primary beam rate monitors like the A1900 Faraday bar were confirmed. However, at rates that are an appreciable fraction of the cyclotron’s RF frequency, the measured rate will diverge from the true incident rate. This occurs because the probability of finding more than 1 particle in a beam bunch becomes significant. If there are 2 or more beam particles in a beam bunch the object scintillator’s scaler will only count once and the 102 measurement will be incorrect. The measured rate by the object scintillator needs to be corrected for this effect. Appendix C gives a detailed derivation of this correction, which can be given as a linear function, Rcor = 5.6 · 10−8 Robject + 0.99, (4.13) where Robject is the raw scaler measurement. This effect was corrected on a run-by-run basis and increased the rate by approximately 10-15%. The energy loss spectrum (∆E) in the object scintillator was also measured. This is shown in Fig. 4.24, where events with 1 and 2 particles in the beam bunch can be seen. With the clear separation between these two cases, a separate determination of the incident rate can be made. As Appendix C shows the average incident beam rate (r) is related to the ratio of single hits (S1 ) and multi hits (S>1 ), and the RF rate (n), r S>1 = . S1 n (4.14) This technique confirmed the scaler measurement with the above correction to within 3%. 4.5.2 S800 Acceptance The S800 Spectrograph has a finite acceptance in both momentum and angle. To extract a properly normalized cross section this acceptance must be taken into account. Further, since the measurement of the cross section was broken across different, partially overlapping rigidity settings these acceptances are key to understanding how to merge the settings together. Here, the case of 14 C, where the largest correction is required, is discussed. The full 103 104 1 Hit Counts 103 2 Hits 102 3 Hits 10 1 0 5000 10000 15000 ∆ E [arb] 20000 25000 Figure 4.24: Histogram of the energy deposited in the object scintillator in the first run of the experiment. The peaks correspond to to single- and multi-particle hits in the scintillator. kinetic energy distribution of 14 C is spread across 4 rigidity settings: 3.1325 Tm, 3.2828 Tm, 3.28 Tm, and 3.52 Tm. Fig. 4.25 shows a two-dimensional histogram of the kinetic energy of the 14 C ejectiles and the dispersive angle at the target from the 3.2828 Tm setting. The acceptance cut of the S800 is clearly visible. There is complete acceptance in the center of the distribution, but the reduced acceptance in the dispersive angle removes events on both ends of the distribution. No acceptance cuts were seen in the non-dispersive direction. To correct for these acceptance effects, the full distribution for 14 C was reconstructed. Fig. 4.26 shows the full kinetic energy distribution of 14 C. The distribution from the 3.52 Tm setting was extended into the area of partial acceptance (< 1420 MeV) by correcting for the evens cut off due to the finite angular acceptance. The same procedure was applied to the low energy side of the 3.2828 Tm setting. These regions are shown in Fig. 4.26 with vertical line shading. Between the central two rigidity settings (3.2828 and 3.38 Tm), no correction for partial acceptance was needed because the areas of full acceptance in the two 104 5 104 4 ATA [Deg] 3 103 2 1 0 102 −1 −2 10 −3 −4 −5 1100 1150 1200 1250 1300 1350 1400 1450 1 T14C [MeV] Figure 4.25: A two-dimensional histogram of the kinetic energy and dispersive angle at the target for the 14 C channel as measured in the 3.2828 Tm rigidity setting. settings overlapped. −6 50 ×10 45 3.38 Tm 3.28 Tm 3.13 Tm 3.52 Tm 40 35 arb 30 25 20 15 10 5 0 1100 1150 1200 1250 1300 1350 1400 1450 1500 1550 1600 T14C [MeV] Figure 4.26: Full kinetic energy distribution for 14 C ejectiles reconstructed from 4 rigidity settings. The vertical line shaded regions indicate areas of incomplete acceptance. See text for details. With the full distribution reconstructed, acceptance curves can be constructed from the data by dividing the uncorrected distribution in each setting by the total. This gives the 105 acceptance of the S800 as a function of kinetic energy of the ejectile. For the 14 C channel only the two central settings (3.2828 and 3.38 Tm) were used to extract charge-exchange information. Fig. 4.27 shows the acceptance curve for the 3.2828 Tm setting obtained in this way. The acceptance on the low-energy side of the distribution has significant statistical fluctuations since it comes from an area of incomplete acceptance. 1.2 Acceptance 1 0.8 0.6 0.4 0.2 0 1100 1150 1200 1250 1300 1350 1400 T14C [MeV] Figure 4.27: Experimentally determined acceptance curve for 14 C in the 3.2828 Tm setting. The width of the ejectile’s kinetic energy distribution from charge-exchange reactions is a function of excitation energy in 16 N. The higher the excitation energy, the more energy is available for the decaying nucleon(s) to carry away—widening the ejectile’s kinetic energy distribution. To study this effect, the distribution from charge-exchange reactions needs to be known, but Fig. 4.26 gives the the distribution for background plus charge-exchange reactions. For the 15 N and 14 N channels, the charge-exchange distribution was estimated by gating around the kinematic relationships seen in Fig. 4.11. From the measured distribution a model for the neutron decay scheme was built into the Geant4 simulation to calculate excitation energy dependent acceptances for each channel. For the 15 C and 14 C channels this 106 was not possible since there were no clear kinematic relationships seen in the two channels. For these cases a decay scheme to the evenly spaced levels in the final nucleus was assumed. The systematic error associated with this procedure is discussed below. 1.2 Efficiency 1 0.8 3.1325 Tm (15N) 2.9 Tm (14N) 3.38 Tm (14C) 3.2828 Tm (14C) 3.6565 Tm (15C) 3.52 Tm (15C) 0.6 0.4 0.2 0 0 2 4 6 8 10 12 Ex [MeV] 14 16 18 20 Figure 4.28: Efficiency curves from the finite S800 acceptance for each rigidity setting used in the analysis. Lines of common color show settings centered around the same detected particle in the CE reaction. The results of these simulations are shown in Fig. 4.28, where the efficiency from the finite S800 acceptance is shown for each rigidity setting used in the analysis. The start of the lines for the higher lying channels indicates the threshold energy. Only for the 15 C and 14 C channels were there two different settings used in the analysis. For these channels the cross section was determined for each setting separately, then averaged together to form a final cross section. As Fig. 4.28 shows, the acceptance for most of the rigidity settings was quite high in the excitation-energy regime of interest. Therefore uncertainties in the procedure to determine these corrections will have a small effect on the final cross section. The exceptions to this were the corrections for the 3.2828 and 3.52 Tm settings for 14 C and 15 C respectively. Both 107 of these settings corresponded to the low-energy tail of their respective channels where it was clear that there was less charge-exchange reactions below 20 MeV of excitation energy. Since this approach corrects each setting into a total cross section, the two results can be checked against one another for consistency as a measure of the systematic error of the procedure. In both cases the results were statistically equivalent, and the systematic error was deemed smaller than the statistical error. 4.5.3 Target Thickness When filled, the windows of the liquid-hydrogen target bulged beyond the nominal 7 cm thickness of the frame. This bulging needs to be taken into account in the determination of target thickness so that the cross section can be properly determined. Originally the thickness was going to be estimated by looking at the energy loss of the 16 C beam going through the full and empty target into the S800 focal plane. This requires changing the rigidity of the spectrometer. However, during the analysis it became clear that the high rigidity of the 100 MeV/u 16 C beam put the spectrometer too close to its maximum operational rigidity and saturated the magnets (Bρ=3.94 (empty) and 3.9 (filled) Tm out of a maximum of 4 Tm). It became impossible to very accurately determine the momenta from the runs with different Bρ. To alleviate this issue the target thickness was determined by looking at the energy loss difference between the 16 C and the containment 14 B. With this approach the systematic error between the two settings from the saturation of the magnets was reduced. Fig. 4.29 shows the raw PID spectrum from the run where the beam was tunned through the spectrometer to the focal plane. Selecting events for 16 C and 14 B and reconstructing their energy yields Fig. 4.30, where the measured kinetic energy of the two particles after traveling through the target is shown. This gives an energy loss difference of 311.7± 0.3 108 2000 1800 ∆ E [arb] 1600 1400 1200 1000 800 B C 14 600 16 400 200 0 940 945 950 955 960 965 970 TOF [arb] 975 980 985 990 Figure 4.29: PID spectrum from the run where the 16 C was tuned into the focal plane of the S800. The beam consisted of 16 C and a small containment of 14 B. MeV. 120 100 B C 14 80 16 60 40 20 0 1000 1100 1200 1300 1400 1500 Kinetic Energy [MeV] 1600 1700 Figure 4.30: Kinetic energy measured by the s800 Spectrograph for the two beam particles. The peak from 16 C has been scaled down for easy comparison with 14 B. With the measured energy loss difference through the target between 16 C and 14 B, the Geant4 simulation was used to determine a relationship between the target thickness and 109 energy loss difference. This was done with the full simulation including the bulged surface of the target and the realistic beam profile determined from the data. A different simulation was run for both particles for a range of target bulge values. The result of these simulations is shown in Fig. 4.31, where a clear linear relationship is seen. From this relationship a half bulge amount of ∼1.0 mm was found, giving a total target thickness of 9.1 ± 0.3 mm. 313 ∆E [MeV] 312.5 312 311.5 311 310.5 310 0.6 0.8 1 1.2 1.4 1.6 Half Bulge [mm] Figure 4.31: Relationship between energy loss difference and liquid hydrogen target thickness determined from simulation. The points are simulated values and the red line is a linear fit. The curved surface of the target and the finite extent of the beam spot must also be taken into account. An average thickness over the width of the beam spot (∼3.9 mm FWHM) was calculated from the amount of bulging and the size of the beam. This reduced the effective target thickness by an amount substantially less than the 0.3 mm error bar. This corresponds to an aerial density of 65.8 ± 2.2 mg/cm2 . 110 4.5.4 Target Density Variations The temperature and pressure of the liquid hydrogen target were monitored throughout the experiment so that any changes in the density could be taken into account in the final cross section normalization. The density was determined from the temperature according to the following expression, 4 3 2 ρH f orLH = −0.0008T + 0.0672T − 2.1467T + 29.219T − 65.934, 2 (4.15) where T is the temperature in Kelvin [104]. Fig. 4.32 shows the density measurements averaged over each experimental run. The fluctuations are small (< 0.5%) and are corrected on a run-by-rub basis. 73.2 73.1 ρ [mg/cm3] 73 72.9 72.8 72.7 72.6 72.5 72.4 72.3 560 580 600 620 640 660 680 700 720 740 760 780 Run Number Figure 4.32: The density derived from the temperature measurement of the liquid hydrogen target for each experimental run. 111 4.6 Cross Section Calculation In addition to the corrections discussed in the above sections, there are other run-by-run corrections for various detectors and other overall normalization corrections. These factors are shown in Equation 4.16 and summarized in Table 4.2. With these many effects, the absolute differential cross section can be extracted. For e10003 the full definition for the differential cross section, containing all efficiency factors and corrections is, dσ Ns (1 + α) = [b/sr], dΩ ∆Ω10−24 ρ∗ daq ndectors tof pid crdc IC nhit s800 Trun trans Rcor NT (4.16) where Ns is the number of scattered particles from charge-exchange reactions in a given COM scattering angle bin, ∆Ω (sr) is the solid angle coverage of that bin, Trun (s) is the length of the run, ndetectors is the neutron detection efficiency, and Rcor (1/s) is the corrected incident beam rate described in Section 4.5.1. NT (1/cm2 ) is the number of target nuclei from the liquid hydrogen and the small contribution from the hydrogen in the cell’s Kapton foils given by,  NT =  ρH LH tcell NA 2 M MH + ρH kapton tf oil NA M MH  , (4.17) where tcell is the target thickness, tf oil is the thickness of the Kapton foils (250 µm), NA is Avogadro’s number, M MH is the molar mass of hydrogen, and ρH x is the density of hydrogen in either liquid hydrogen or kapton. The remaining factors, their descriptions, and their approximate values are summarized in Table 4.2. The following paragraph gives a brief explanation of how each of these corrections was obtained. The DAQ live time (daq ) was determined from the ratio of the S800’s live and raw clocks. Though the singles rate in the S800 focal plane was high (∼10k), the coincidence 112 condition with the neutron detectors brought live trigger rate down substantially—decreasing the overall dead time of the system. The PID TOF efficiency (tof ) adjusted for the number of events when the correct beam bucket could not be found in the digitized waveform of the object scintillator signal. This timing signal was needed for the PID, so events lacking the signal were discarded. The PID gate efficiency (pid ) was the 2σ radius of the PID gates. The CRDC efficiency (crdc ) was determined by comparing the number of hits in the IC to hits in the CRDCs for a given run. The IC pile up efficiency (IC ) was found by inspecting the number of events in the high-energy tail of the IC distribution for a given particle. It was found to be constant across different runs and detected particles. The single neutron hit efficiency (nhit ) was found by examining the neutron multiplicity spectrum. Only events in which 1 neutron was detected could be used in the analysis so higher multiplicities were discarded. The correction was found to be constant across the different rigidity settings. The beam transmission (trans ) from the object scintillator to the target position was measured at the beginning (77.5%) and end (82.3%) of the experiment by tuning the beam into the S800 focal plane. The increased transmission seen in the second measurement was applied to runs 631 and above. A jump in the cross section measurement was observed at run 631, motivating the use of the second transmission measurement. The object pile-up correction (α) was found by counting the number of multi-hit events in the object scintillator. Only events that deposited energy consistent with 1 particle in the object scintillator were used in the analysis. The timing resolution for the multi-particle hits was poor and so were discarded. 113 Factor Description Value daq DAQ live time 90-96% tof PID TOF efficiency 90-95% pid PID gate efficiency 95% crdc CRDC efficiency 50-70% IC IC pileup efficiency 98% nhit Single neutron hit efficiency 93% trans Beam transmission to target 77.5,82.3% α Object pileup correction 8-40% ρ∗ Target Density Correction ¡0.5% Table 4.2: Summary of corrections to the cross sections. Parameters with a range in values indicate that a run-by-run correction was performed. See text for details on how each correction was determined. Fig. 4.33 shows charge-exchange differential cross section calculated from Equation 4.16. The cross section is shown for the Θcm = 4−6° and Θcm = 14−16° bins, with the contribution from each decay channel shown. From the higher-resolution spectrum at forward angles, it is clear that the (p,n) reaction in inverse kinematics can give detailed structure information. The two known states from β decay at 3.3 and 4.3 MeV are resolved and can be identified. Alongside the low-lying structure information the response up to high excitation energy is mapped out. The error in the total cross section for excitation energies above ∼12 MeV increases with the opening of the other decay channels, where the statistical and systematic uncertainties are increased. 114 3 15 N N 14 N 14 C 15 C dσ/dΩ [mb sr-1] 2.5 16 2 1.5 1 0.5 0 0 5 10 15 20 Ex [MeV] (a) Θcm =4-6° 8 15 N N 14 N 14 C 15 C dσ/dΩ [mb sr-1] 7 16 6 5 4 3 2 1 0 0 5 10 15 20 Ex [MeV] (b) Θcm =4-6° Figure 4.33: Differential cross section for the 16 C(p,n)16 N reaction showing the contributions from each reaction channel. Panel (a) shows the cross section at forward angles in 250 keV bins. In (b) the same angle is shown with a coarser binning to highlight the channels at higher excitation energy, where the statistics is low. The most backward angle, with the poorest resolution, is presented in (c). The shaded bans in (b) and (c) indicate the systematic error estimation from the background subtraction for the 15 C and 14 C channels. The small vertical lines on the top of each figure indicate the threshold energies for the 15 N, 15 C, 14 C and 14 N channels. The color of the line matches the color of the data to indicate the channel. 115 Figure 4.33: cont’d (c) Θcm =14-16° dσ/dΩ [mb sr-1] 3 15 N N 14 N 14 C 15 C 16 2.5 2 1.5 1 0.5 0 0 5 10 15 20 Ex [MeV] Fig. 4.34 presents the total cross section for the Θcm = 4 − 6° and Θcm = 14 − 16° bins. The systematic error from the background subtraction and uncertainty in the neutron acceptance is shown as a green band. Section 4.7 gives more details about the systematic error estimation. 4.7 Systematic Error No absolute cross-section measurement is complete without an estimate of the systematic errors. Table 4.3 gives a breakdown of the major systematic errors in experiment e10003. Systematic error in the extracted differential cross section can either give an overall uncertainty in the normalization, or create a more complex uncertainty, like in the case of the background subtraction. The systematic error from the background subtraction is not included in Table 4.3 since it is treated as a function of COM scattering angle and excitation 116 3 dσ/dΩ [mb sr-1] 2.5 2 1.5 1 0.5 0 0 5 10 15 20 15 20 Ex [MeV] (a) Θcm =4-6° 2 dσ/dΩ [mb sr-1] 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 5 10 Ex [MeV] (b) Θcm =14-16° Figure 4.34: Total differential cross section from all decay channels in the 16 C(p,n)16 N reaction. The green band shows the systematic error including the systematic error from the background subtractions in the different channels and the systematic error in the neutron acceptance determination. 117 energy. Details of the estimation technique for each of the systematic errors are given in the following paragraphs. Source Estimate Incident rate determination 15.6% Background subtraction See text Neutron Acceptance 5% Target Thickness 3.3% Neutron Energy Threshold 1.8% Random Background < 1% Empty cell induced background < 0.5% Total 16.8% Table 4.3: Summary of the source and estimated size of the major systematic uncertainties in the 16 C(p,n)16 N measurement. See text for details on each source of systematic uncertainty The largest systematic error was in the incident beam rate normalization procedure. This error was estimated by looking at the total charge-exchange cross section from 0-20 MeV for each 15 N experimental run. If the normalization was perfect, each run should give statistically equivalent values. Therefore, the spread seen between the runs will give an estimate of the uncertainty in the incident beam rate measurement. These integrated cross sections are shown in Fig. 4.35. The rigidity setting appropriate for detecting 15 N (3.1325 Tm) was measured three times during the experiment, so the large gaps between the points correspond to the runs when the other rigidity settings were measured. From this plot a FWHM systematic error of 15.6% for the normalization procedure was estimated. Unlike the beam normalization, the background subtraction affects the shape of the extracted cross section. To estimate the uncertainty in the subtraction, the background 118 40 dσ Σ20 0 dΩ [mb/sr] 35 30 25 20 15 10 5 0 560 580 600 620 640 660 680 700 720 740 760 780 Run Number Figure 4.35: The total charge-exchange cross section from 0-20 MeV for each experimental run for the 15 N channel. The spread in the points gives an estimate of the uncertainty in the beam normalization. model was varied for each rigidity setting—giving a bin-by-bin estimate of the uncertainty. For the neutron-decay channels (15 N and 14 N), this involved varying the integration ranges used to find the scaling functions shown in Fig. 4.20. This indicated that the error from the background subtraction in the regions of interest was small (∼ 4%). For the proton-decay channels (15 C and 14 C), the uncertainty was higher. A similar procedure was followed to estimate the uncertainty where the different background models were compared. For 15 C the portion of the low-energy tail of ejectile kinetic energy distribution that was used to model the background was varied. For 14 C a separate background model using events that created 12 C was compared with the technique used above. This gave an uncertainty that varied from 10-30% depending on angle and excitation energy. The error in neutron acceptance comes from uncertainty in the exact positions of the neutron detectors. As Fig. 4.18 showed, the neutron acceptance is not a smooth function, so small changes to the angles of the neutron detectors can change the acceptance significantly. 119 During the analysis process small adjustments were made to the positions of the neutron bars to correct for any unphysical discontinuities seen in the excitation energy spectrum. The size of these adjustments gave a scale to vary the angles in the estimation of this systematic error. The Geant4 simulation was used to propagate the variations in detector angle to the simulated yield in each COM bin. This gave a relatively flat change of 5%, except for the bins in excitation energy from 9-11 MeV at forward angles. There the spectrum is particularly sensitive to the position of the bars because of the steepness of the kinematic lines. An additional 8% systematic error was added to these bins. The error in the target thickness determination was already discussed in Section 4.5.3. The error comes from the uncertainty in the energy loss difference between 16 C and 14 B. This error was estimated by comparing the shape of the measured distribution for the two nuclei. The error due to uncertainties in the neutron detection threshold were estimated in a similar way using the Geant4 simulation. The error in the threshold was determined by comparing the neutron detector gain calibrations before and after the experiment. The gain at the end of the experiment had decreased by approximately 9%, lowering the yield. This error was combined with any fitting uncertainties in the gain calibration and propagated through the full Geant4 simulation of the setup. This showed a 1.8% change to the neutron yield. The systematic error associated with the random and cell backgrounds were already discussed in Section 4.4 and are placed in Table 4.3 for convenience. As indicated in Table 4.3 these systematic errors combined to give a total systematic error of 16.8%. The final systematic error bands presented in chapter 5 are a combination of the systematic error from the background subtraction and the 16.8% error in the overall normalization. 120 Chapter 5 Results The trick is to just look at the question and write down the answer. Works every time This chapter gives the results of the 16 C(p,n)16 N reaction at 100 MeV/u. The differential cross section is used to form angular distributions, needed to perform a multiple decomposition analysis (MDA). The MDA isolates the different ∆L components of charge-exchange cross section so that GT strength and dipole cross sections can be extracted. 5.1 Angular Distributions dσ ) as a function of COM scattering angle) Angular distributions (differential cross section ( dΩ are formed from the experimental cross section data. To make these distributions for each excitation energy bin, the cross-section data was separated into 2 degree angular bins from 4 to 16 degrees in the COM. The bin from 0-2 degrees was below the neutron detection threshold. The bin from 2-4 degrees was excluded due to the large amount of low energy neutron background, which could not be subtracted reliably. Angles beyond 16 degrees in the COM were not used in the analysis. As discussed in Section 4.3, the excitation-energy resolution worsens at larger scattering angles. To correctly form angular distributions, the data at forward scattering angles must 121 be smeared to match the excitation-energy resolution at the most backward scattering angle used in the analysis. Fig. 5.1 shows an example of this smearing procedure. The figure shows the same data at Θcm = 4 − 6 ° as Fig. 4.34, but with additional smearing to match the excitation energy resolution for the Θcm =14-16° bin. 2 dσ/dΩ [mb sr-1] 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 5 10 15 20 Ex [MeV] Figure 5.1: Differential cross section for the 16 C(p,n)16 N reaction as a function of excitation energy. Data is from Θcm =4-6° but smeared to match the resolution at Θcm =14-16°. 5.2 Multipole Decomposition Analysis A multipole decomposition analysis is used to separate the different angular momentum transfer components of the cross section. The smeared angular distributions are fit with a linear combination of theoretical cross sections calculated in DWBA,         dσ dσ dσ dσ =a +b +c , dΩ total dΩ ∆L=0 dΩ ∆L=1 dΩ ∆L=2 122 (5.1) where a, b, and c are the fitting parameters. The fitting was done through χ2 minimization, where the error bars in the angular points came from a combination of the statical error and the systematic uncertainty from the neutron detector acceptance. For each ∆L component, the shape of the angular distribution is represented by a characteristic DWBA calculation. Section 2.4 gives details on the cross-section calculations and presents the characteristic angular distributions used in the MDA. The result of the MDA is shown in Fig. 5.2. Angular distributions in 1 MeV bins of excitation energy were prepared from the smeared excitation energy distributions and fit according to Equation 5.1. Good agreement between the fits and the data is seen. Of specific interest is the success of the fits in first bin (0-1 MeV), where a reduced χ2 of 1.09 is achieved. Between 0 and 1 MeV there are only expected to be 4 negative parity states at 0.0, 120, 298, and 397 keV. Starting from a 16 C ground state of 0+ , transitions to these states must all be ∆L=1. This is confirmed by the MDA. It is clear from the remaining fits that much of the data is strongly ∆L = 0 in nature. It is important to note that the ability of the MDA to distinguish between ∆L=1 and ∆L=2 is limited. Over the angular range used in the analysis the angular distributions for ∆L=1,2 transitions are similar. To explore this sensitivity, decompositions including only ∆L=0,1, and only ∆L=0,2 were compared with the full fit. In both cases, the portion of the distribution coming from GT type transitions did not change significantly. The fits without ∆L=1 were of a similar quality to the fits without ∆L = 2, yet the full fit strongly prefers the ∆L=1 component. Given this, all three components were used in the final analysis, but with the understanding that differentiating ∆L=1 and 2 is difficult. Fig. 5.3 shows the ∆L components of the cross section in each angular bin as determined by the MDA. As expected from the theoretical calculations, the dipole component peaks at the backward angles and appears at high excitation energy. In the 14-16° bin, there is 123 0.5 0.4 0.3 0.2 Data ∆ L=0 ∆ L=1 ∆ L=2 Total 01 to the number with just 1 particle S1 , H + H + . . .n−1 H0 S>1 = n−1 r−2 n−1 r−3 S1 n−1 Hr−1 = = −n−1 Hr +n−1 Hr −n−1 Hr−1 +n−1 Hr−1 +n−1 Hr−2 +n−1 Hr−3 + . . .n−1 H0 n−1 Hr−1 n Hr −n−1 Hr −n−1 Hr−1 , n−1 Hr−1 (C.9) where in the last step, the following was used, n Hr =n−1 Hr−1 +n−1 Hr−2 . . . +n−1 H0 . (C.10) Reducing the expression further, = n Hr −n−1 Hr −n−1 Hr−1 n−1 Hr−1 Hr H − n−1 − n−1 r−1 n−1 Hr−1 n−1 Hr−1 n−1 Hr−1 Hr n Hr = − n−1 −1 n−1 Hr−1 n−1 Hr−1 = n Hr (n + r − 1)!(r − 1)!(n − 2)! (n + r − 2)!(r − 1)!(n − 2)! = − −1 r!(n − 1)!(n + r − 3)! r!(n − 2)!(n + r − 3)! = = (C.11) (n + r − 1)(n + r − 2) (n + r − 2) − −1 r(n − 1) r r−1 . n−1 This yields a useful relation for incident beam rate, S>1 r−1 = , S1 n−1 155 (C.12) which for r >> 1 and n >> 1 reduces further to, S>1 r = . S1 n (C.13) Given the known number of buckets in a 1 second period (the RF rate) and the ratio of single particle to multi-particle buckets, the average incident particle rate can be determined. Correcting an In-Beam Scaler Measurement If the incoming beam rate is directly measured with an in-beam detector (a plastic scintillator for example), corrections will have to be made to the measured rate because of multi-particle buckets. The scaler will count only once if there is a bucket containing more than one beam particle, therefore the probabilities for each number of particles needs to be calculated. This can be done using Equation C.8, H P0 (n, r) = n−1 r−0 n Hr (n − 2 + r)!r!(n − 1)! = r!(n − 2)!(n + r − 1)! = (n − 1) , (n + r − 1) H P1 (n, r) = n−1 r−1 n Hr (n − 3 + r)!r!(n − 1)! = (r − 1)!(n − 2)!(n + r − 1)! = (C.14) r(n − 1) , (n + r − 1)(n + r − 2) 156 (C.15) and H P2 (n, r) = n−1 r−2 n Hr (n − 4 + r)!r!(n − 1)! = (r − 2)!(n − 2)!(n + r − 1)! = (C.16) r(r − 1)(n − 1) . (n + r − 1)(n + r − 2)(n + r − 3) Continuing the clear pattern, the following are presented for completeness: H P3 (n, r) = n−1 r−3 n Hr r(r − 1)(r − 2)(n − 1) = , (n + r − 1)(n + r − 2)(n + r − 3)(n + r − 4) H P4 (n, r) = n−1 r−4 n Hr r(r − 1)(r − 2)(r − 3)(n − 1) = , (n + r − 1)(n + r − 2)(n + r − 3)(n + r − 4)(n + r − 5) H P5 (n, r) = n−1 r−5 n Hr r(r − 1)(r − 2)(r − 3)(r − 4)(n − 1) = , (n + r − 1)(n + r − 2)(n + r − 3)(n + r − 4)(n + r − 5)(n + r − 6) (C.17) (C.18) (C.19) and H P6 (n, r) = n−1 r−6 n Hr = r(r − 1)(r − 2)(r − 3)(r − 4)(r − 5)(n − 1) . (n + r − 1)(n + r − 2)(n + r − 3)(n + r − 4)(n + r − 5)(n + r − 6)(n + r − 7) (C.20) 157 Taking the RF rate from e10003 of 21.8259 MHz and considering a one second period, the following table can be generated using Equations C.14 through C.20, RT N0 N1 N2 ... N6 Rm RT /Rm 5.00E+5 2.13E+07 4.78E+05 1.07E+04 ... 2.69E-03 4.89E+05 1.023 1.00E+6 2.09E+07 9.14E+05 4.01E+04 ... 1.48E-01 9.56E+05 1.046 1.50E+6 2.04E+07 1.31E+06 8.45E+04 ... 1.44E+00 1.40E+06 1.069 2.00E+6 2.00E+07 1.68E+06 1.41E+05 ... 6.99E+00 1.83E+06 1.092 2.50E+6 1.96E+07 2.01E+06 2.07E+05 ... 2.31E+01 2.24E+06 1.115 3.00E+6 1.92E+07 2.32E+06 2.80E+05 ... 5.97E+01 2.64E+06 1.137 3.50E+6 1.88E+07 2.60E+06 3.59E+05 ... 1.31E+02 3.02E+06 1.160 4.00E+6 1.84E+07 2.86E+06 4.42E+05 ... 2.55E+02 3.38E+06 1.183 4.50E+6 1.81E+07 3.09E+06 5.29E+05 ... 4.51E+02 3.73E+06 1.206 5.00E+6 1.78E+07 3.31E+06 6.17E+05 ... 7.45E+02 4.07E+06 1.229 5.50E+6 1.74E+07 3.51E+06 7.06E+05 ... 1.16E+03 4.39E+06 1.252 where RT is the true number of particles, Nx is the number of buckets containing x particles and Rm is the sum P6 x=1 Nx . An in beam detector will measure Rm since the scalar will only count once when there is one or more particles in the bucket. To correct for this the measured rate only needs to be multiplied by the correction factor seen in the last column of the table. As the true rate increases, the size of this effect becomes more significant. It is convenient to express this correction as a linear fit to the points given in the last column as a function of the measured rate in the object scintillator (Robject ), Rcor = 5.6 · 10−8 Robject + 0.99. 158 (C.21) This expression is used in the analysis of the experimental data. 159 BIBLIOGRAPHY 160 BIBLIOGRAPHY [1] E. Rutherford, “LXXIX. The scattering of α and β particles by matter and the structure of the atom,” The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, vol. 21, p. 669, 1911. [2] J. Erler, N. Birge, M. Kortelainen, W. Nazarewicz, E. Olsen, A. M. Perhac, and M. Stoitsov, “The limits of the nuclear landscape,” Nature, vol. 486, p. 509, 2012. [3] K. Gelbke et al., Isotope Science Facility at Michigan State University. 2006. [4] M. G. Mayer, “On Closed Shells in Nuclei,” Physical Review, vol. 74, p. 235, 1948. [5] E. Feenberg, K. C. Hammack, and M. G. Mayer, “On Closed Shells in Nuclei. II,” Physical Review, vol. 75, p. 1877, 1949. [6] O. Haxel, J. H. D. Jensen, and H. E. Suess, “On the “Magic numberNOTs” in Nuclear Structure,” Physical Review, vol. 75, p. 1766, 1949. [7] B. A. Brown, “Lecture Notes in Nuclear Structure Physics,” November 2005. [8] B. A. Brown, “Oxbash for Windows,” 2004. [9] B. A. Brown and W. D. M. Rae, “The Shell-Model Code NuShellX@MSU,” Nuclear Data Sheets, vol. 120, p. 115, 2014. [10] A. Gade and T. Glasmacher, “In-beam nuclear spectroscopy of bound states with fast exotic ion beams,” Progress in Particle and Nuclear Physics, vol. 60, p. 161, 2008. [11] T. Otsuka, R. Fujimoto, Y. Utsuno, B. A. Brown, M. Honma, and T. Mizusaki, “Magic numberNOTs in exotic nuclei and spin-isospin properties of the NN interaction,” Physical Review Letters, vol. 87, p. 082502, 2001. [12] T. Otsuka, T. Suzuki, R. Fujimoto, H. Grawe, and Y. Akaishi, “Evolution of nuclear shells due to the tensor force,” Physical Review Letters, vol. 95, p. 232502, 2005. [13] W. Bethe and W. Gentner, “Weitere Atomumwandlunger, durch γ-Struhlen,” Zeitschrift für Physik, vol. 19, p. 191, 1937. [14] M. Harakeh and A. V. D. Woude, Giant Resonances Fundemental High-Frequency Modes of Nuclear Excitation. Oxford University Press, 2001. 161 [15] G. C. Baldwin and G. S. Klaiber, “Photo-fission in heavy elements,” Physical Review, vol. 71, p. 3, 1947. [16] G. C. Baldwin and G. S. Klaiber, “X-ray yield curves for γ-n reactions,” Physical Review, vol. 73, p. 1156, 1948. [17] F. Buskirk, H.-D. Gräf, R. Pitthan, H. Theissen, O. Titze, and T. Walcher, “Evidence for E2 resonances at high excitation energies in 208 Pb,” Physics Letters B, vol. 42, p. 194, 1972. [18] S. Noji, A new spectroscopic tool by the radioactive-isotope-beam induced exothermic charge-exchange reaction. PhD thesis, University of Tokyo, 2012. [19] Y.-W. Lui, D. H. Youngblood, S. Shlomo, X. Chen, Y. Tokimoto, Krishichayan, M. Anders, and J. Button, “Isoscalar giant resonances in 48 Ca,” Physical Review C, vol. 83, p. 044327, 2011. [20] D. Youngblood, Y.-W. Lui, and H. Clark, “Giant resonances in 24 Mg,” Physical Review C, vol. 60, p. 014304, 1999. [21] Y. Tokimoto, Y. W. Lui, H. L. Clark, B. John, X. Chen, and D. H. Youngblood, “Giant resonances in 46,49 Ti,” Physical Review C, vol. 74, p. 044308, 2006. [22] Y. W. Lui, H. L. Clark, and D. H. Youngblood, “Giant resonances in 16 O,” Physical Review C, vol. 64, p. 064308, 2001. [23] Y.-W. Lui, P. Bogucki, J. D. Bronson, D. H. Youngblood, and U. Garg, “Giant resonaces in 112 Sn,” Physical Review C, vol. 30, p. 51, 1984. [24] M. Uchida, H. Sakaguchi, M. Itoh, M. Yosoi, T. Kawabata, Y. Yasuda, H. Takeda, T. Murakami, S. Terashima, S. Kishi, U. Garg, P. Boutachkov, M. Hedden, B. Kharraja, M. Koss, B. K. Nayak, S. Zhu, M. Fujiwara, H. Fujimura, H. P. Yoshida, K. Hara, H. Akimune, and M. N. Harakeh, “Systematics of the bimodal isoscalar giant dipole resonance,” Physical Review C, vol. 69, p. 051301, 2004. [25] T. Li, U. Garg, Y. Liu, R. Marks, B. K. Nayak, P. V. Madhusudhana Rao, M. Fujiwara, H. Hashimoto, K. Nakanishi, S. Okumura, M. Yosoi, M. Ichikawa, M. Itoh, R. Matsuo, T. Terazono, M. Uchida, Y. Iwao, T. Kawabata, T. Murakami, H. Sakaguchi, S. Terashima, Y. Yasuda, J. Zenihiro, H. Akimune, K. Kawase, and M. N. Harakeh, “Isoscalar giant resonances in the Sn nuclei and implications for the asymmetry term in the nuclear-matter incompressibility,” Physical Review C, vol. 81, p. 034309, 2010. [26] M. Itoh, H. Sakaguchi, M. Uchida, T. Ishikawa, T. Kawabata, T. Murakami, H. Takeda, T. Taki, S. Terashima, N. Tsukahara, Y. Yasuda, M. Yosoi, U. Garg, M. Hedden, B. Kharraja, M. Koss, B. K. Nayak, S. Zhu, H. Fujimura, M. Fujiwara, K. Hara, 162 H. P. Yoshida, H. Akimune, M. N. Harakeh, and M. Volkerts, “Systematic study of L ≤ 3 giant resonances in Sm isotopes via multipole decomposition analysis,” Physical Review C, vol. 68, p. 064602, 2003. [27] M. Uchida, H. Sakaguchi, M. Itoh, M. Yosoi, T. Kawabata, H. Takeda, Y. Yasuda, T. Murakami, T. Ishikawa, T. Taki, N. Tsukahara, S. Terashima, U. Garg, M. Hedden, B. Kharraja, M. Koss, B. K. Nayak, S. Zhu, M. Fujiwara, H. Fujimura, K. Hara, E. Obayashi, H. P. Yoshida, H. Akimune, M. N. Harakeh, and M. Volkerts, “Isoscalar giant dipole resonance in 208 Pb via inelastic α scattering at 400 MeV and nuclear incompressibility,” Physics Letters B, vol. 557, p. 12, 2003. [28] B. K. Nayak, U. Garg, M. Hedden, M. Koss, T. Li, Y. Liu, P. V. Madhusudhana Rao, S. Zhu, M. Itoh, H. Sakaguchi, H. Takeda, M. Uchida, Y. Yasuda, M. Yosoi, H. Fujimura, M. Fujiwara, K. Hara, T. Kawabata, H. Akimune, and M. N. Harakeh, ““Bi-modal” isoscalar giant dipole strength in 58 Ni,” Physics Letters B, vol. 637, p. 43, 2006. [29] M. Itoh, H. Sakaguchi, M. Uchida, T. Ishikawa, T. Kawabata, T. Murakami, H. Takeda, T. Taki, S. Terashima, N. Tsukahara, Y. Yasuda, M. Yosoi, U. Garg, M. Hedden, B. Kharraja, M. Koss, B. K. Nayak, S. Zhu, H. Fujimura, M. Fujiwara, K. Hara, H. P. Yoshida, H. Akimune, M. N. Harakeh, and M. Volkerts, “Compressional-mode giant resonances in deformed nuclei,” Physics Letters B, vol. 549, p. 58, 2002. [30] J. P. Blaizot, “Nuclear compressibilities,” Physics Reports, vol. 64, p. 171, 1980. [31] D. H. Youngblood, H. L. Clark, and Y.-W. Lui, “Incompressibility of Nuclear Matter from the Giant Monopole Resonance,” Physical Review Letters, vol. 82, p. 691, 1999. [32] D. H. Youngblood, H. L. Clark, and Y.-W. Lui, “Compressibility of Nuclear Matter from the Giant Monopole Resonance,” Nuclear Physics A, vol. 649, p. 49, 1999. [33] S. Nakayama, H. Akimune, Y. Arimoto, I. Daito, H. Fujimura, Y. Fujita, M. Fujiwara, K. Fushimi, H. Kohri, N. Koori, K. Takahisa, T. Takeuchi, A. Tamii, M. Tanaka, T. Yamagata, Y. Yamamoto, K. Yonehara, and H. Yoshida, “Isovector Electric Monopole Resonance in 60 Ni,” Physical Review Letters, vol. 83, p. 690, 1999. [34] R. G. T. Zegers, A. Van den Berg, S. Brandenburg, F. Fleurot, M. Fujiwara, J. Guillot, V. Hannen, M. Harakeh, H. Laurent, K. Van der Schaaf, S. Van der Werf, A. Willis, and H. Wilschut, “Search for Isovector Giant Monopole Resonances via the Pb(3 He,t p) Reaction,” Physical Review Letters, vol. 84, p. 3779, 2000. [35] M. Scott, R. G. T. Zegers, R. Almus, S. M. Austin, D. Bazin, B. A. Brown, C. Campbell, A. Gade, M. Bowry, S. Galès, U. Garg, M. N. Harakeh, E. Kwan, C. Langer, C. Loelius, S. Lipschutz, E. Litvinova, E. Lunderberg, C. Morse, S. Noji, G. Perdikakis, T. Redpath, C. Robin, H. Sakai, Y. Sasamoto, M. Sasano, C. Sullivan, J. A. Tostevin, 163 T. Uesaka, and D. Weisshaar, “Observation of the Isovector Giant Monopole Resonance via the 28 Si(10 Be, 10 B [1.74 MeV]) Reaction at 100 AMeV,” Physical Review Letters, vol. 118, p. 172501, 2017. [36] A. Erell, J. Alster, J. Lichtenstadt, M. A. Moinester, J. Bowman, M. Cooper, F. Irom, H. Matis, E. Piasetzky, U. Sennhause, and Q. Ingram, “Properties of the Isovector Monopole and Other Giant Resonances in Pion Charge Exchange,” Physical Review Letters, vol. 52, p. 2134, 1984. [37] K. Yako, H. Sagawa, and H. Sakai, “Neutron skin thickness of 90 Zr determined by charge exchange reactions,” Physical Review C, vol. 74, p. 051303(R), 2006. [38] A. Klimkiewicz, N. Paar, P. Adrich, M. Fallot, K. Boretzky, T. Aumann, D. CortinaGil, U. D. Pramanik, T. W. Elze, H. Emling, H. Geissel, M. Hellström, K. L. Jones, J. V. Kratz, R. Kulessa, C. Nociforo, R. Palit, H. Simon, G. Surówka, K. Sümmerer, D. Vretenar, and W. Waluś, “Nuclear symmetry energy and neutron skins derived from pygmy dipole resonances,” Physical Review C, vol. 76, p. 051603(R), 2007. [39] B. A. Brown, “Neutron radii in nuclei and the neutron equation of state,” Physical Review Letters, vol. 85, p. 5296, 2000. [40] S. Typel and B. A. Brown, “Neutron radii and the neutron equation of state in relativistic models,” Physical Review C, vol. 64, p. 027302, 2001. [41] S. Yoshida and H. Sagawa, “Neutron skin thickness and equation of state in asymmetric nuclear matter,” Physical Review C, vol. 69, p. 024318, 2004. [42] Y. Fujita, B. Rubio, and W. Gelletly, “Spin-isospin excitations probed by strong, weak and electro-magnetic interactions,” Progress in Particle and Nuclear Physics, vol. 66, p. 549, 2011. [43] D. H. Wilkinson, “Renormalization of the Axial-Vector Coupling Constant in Nuclear β Decay,” Physical Review C, vol. 7, p. 930, 1973. [44] M. P. Nakada, J. Anderson, C. Gardner, J. McClure, and C. Wong, “Neutron Spectrum from p+d Reaction,” Physical Review, vol. 110, p. 594, 1958. [45] J. Anderson and C. Wong, “Evidence For Charge Independence in Medium Weight Nuclei,” Physical Review Letters, vol. 7, p. 250, 1961. [46] P. H. Bowen, G. C. Cox, G. B. Huxtable, J. J. Thresher, and A. Langsford, “Neutrons Emitted at 0° From Nuclei Bombarded by 143 MeV Protons,” Nuclear Physics, vol. 30, p. 475, 1962. 164 [47] R. R. Doering and A. Galonsky, “Observation of Giant Gamow-Teller Strength in (p,n) Reactions,” Physical Review Letters, vol. 35, p. 1691, 1975. [48] C. Goodman, C. Goulding, M. Greenfield, J. Rapaport, D. Bainum, C. Foster, W. Love, and F. Petrovich, “Gamow-Teller Matrix Elements from 0° (p, n) Cross Sections,” Physical Review Letters, vol. 44, p. 1755, 1980. [49] J. Rapaport and D. Wang, “(~p,n) reaction on carbon isotopes at Ep =160 MeV,” Physical Review C, vol. 36, p. 500, 1987. [50] T. N. Taddeucci, J. Rapaport, and C. C. Fostster, “Spin excitations in 40 Ca(p,n),” Physical Review C, vol. 28, p. 2511, 1983. [51] B. D. Anderson, R. J. Mccarthy, M. Ahmad, A. Fazely, A. M. Kalenda, J. N. Knudson, J. Watson, R. Madey, and C. C. Foster, “Comparison of 0°(p,n) cross sections and B(M1) values for seperating current and spin contributions to isovector M1 transitions,” Physical Review C, vol. 26, p. 8, 1982. [52] X. Yang, L. Wang, J. Rapaport, C. D. Goodman, C. Foster, Y. Wang, W. Unkelbach, E. Sugarbaker, D. Marchlenski, S. de Lucia, B. Luther, J. L. Ullmann, A. G. Ling, B. K. Park, D. S. Sorenson, L. Rybarcyk, T. N. Taddeucci, C. R. Howell, and W. Tornow, “Dipole and spin-dipole resonances in charge-exchange reactions on 12 C,” Physical Review C, vol. 48, p. 1158, 1993. [53] L. Wang, X. Yang, J. Rapaport, C. D. Goodman, C. C. Foster, Y. Wang, R. A. Lindgren, E. Sugarbaker, D. Marchlenski, S. De Lucia, B. Luther, L. Rybarcyk, T. N. Taddeucci, and B. K. Park, “10 B(p,n)10 C reaction at 186 MeV,” Physical Review C, vol. 47, p. 2123, 1993. [54] J. Rapaport, T. N. Taddeucci, C. Gaarde, C. D. Goodman, C. C. Foster, C. A. Goulding, D. Horen, E. Sugarbaker, T. Masterson, and D. Lind, “12 C(p,n)12 N reaction at 120, 160, and 200 MeV,” Physical Review C, vol. 24, p. 335, 1981. [55] M. Bhattacharya, C. D. Goodman, R. S. Raghavan, M. Palarczyk, J. Rapaport, I. J. V. Heerden, and P. Zupranski, “Measurement of Gamow-Teller Strength for 176 Yb→176 Lu and the Efficiency of a Solar Neutrino Detector,” Physical Review Letters, vol. 85, p. 4446, 2000. [56] T. N. Taddeucci, C. A. Goulding, T. A. Carey, R. C. Byrd, C. D. Goodman, C. Gaarde, J. Larsen, D. Horen, J. Rapaport, and E. Sugarbaker, “The (p,n) reaction as a probe of beta decay strength,” Nuclear Physics A, vol. 469, p. 125, 1987. [57] K. P. Jackson, A. Celler, W. P. Alford, K. Raywood, R. Abegg, R. E. Azuma, C. K. Campbell, S. El-Kateb, D. Frekers, P. W. Green, O. Häusser, R. L. Helmer, R. S. Henderson, K. H. Hicks, R. Jeppesen, P. Lewis, C. A. Miller, A. Moalem, M. A. 165 Moinester, R. B. Schubank, G. G. Shute, B. M. Spicer, M. C. Vetterli, A. I. Yavin, and S. Yen, “The (n,p) reaction as a probe of Gamow-Teller strength,” Physics Letters B, vol. 201, p. 25, 1988. [58] E. R. Flynn and J. D. Garrett, “Excitation of the Parent Analogs of the 58 Ni Giant M1 Resonance by the Reaction 58 Ni(t,h)58 Co,” Physical Review Letters, vol. 29, p. 1748, 1972. [59] S. Rakers, F. Ellinghaus, R. Bassini, C. Bäumer, A. M. Van den Berg, D. Frekers, D. De Frenne, M. Hagemann, V. M. Hannen, M. N. Harakeh, M. Hartig, R. Henderson, J. Heyse, M. A. De Huu, E. Jacobs, M. Mielke, J. M. Schippers, R. Schmidt, S. Y. Van der Werf, and H. J. Wörtche, “Measuring the (d,2 He) reaction with the focal-plane detection system of the BBS magnetic spectrometer at AGOR,” Nuclear Instruments and Methods in Physics Research A, vol. 481, p. 253, 2002. [60] H. Okamura, S. Fujita, Y. Hara, K. Hatanaka, T. Ichihara, S. Ishida, K. Katoh, T. Niizeki, H. Ohnuma, H. Otsu, H. Sakai, N. Sakamoto, Y. Satou, T. Uesaka, T. Wakasa, and T. Yamashita, “Tensor analyzing power of the (d,2 He) reaction at 270 MeV,” Physics Letters B, vol. 345, p. 1, 1995. [61] T. Annakkage, J. Jgnecke, J. S. Winfield, G. R. A. Berg, J. A. Brown, G. Crawley, S. Danczyk, M. Fujiwara, D. J. Mercer, K. Pham, D. A. Roberts, and G. H. Yoo, “Isovector giant resonances in 6 He, 12 B, 90 Y, 120 In and 208 Tl observed in the (7 Li,7 Be) charge-exchange reaction,” Nuclear Physics A, vol. 648, p. 3, 1999. [62] M. Fujiwara, H. Akimune, I. Daito, and H. Ejiri, “Spin-Isospin Resonances in Nuclei,” Nuclear Physics A, vol. 599, p. 223, 1996. [63] M. Sasano, G. Perdikakis, R. G. T. Zegers, S. M. Austin, D. Bazin, B. A. Brown, C. Caesar, A. L. Cole, J. M. Deaven, N. Ferrante, C. J. Guess, G. W. Hitt, R. Meharchand, F. Montes, J. Palardy, A. Prinke, L. A. Riley, H. Sakai, M. Scott, A. Stolz, L. Valdez, and K. Yako, “Gamow-Teller Transition Strengths from 56 Ni,” Physical Review Letters, vol. 107, p. 202501, 2011. [64] M. Sasano, G. Perdikakis, R. G. T. Zegers, S. M. Austin, D. Bazin, B. A. Brown, C. Caesar, A. L. Cole, J. M. Deaven, N. Ferrante, C. J. Guess, G. W. Hitt, M. Honma, R. Meharchand, F. Montes, J. Palardy, A. Prinke, L. A. Riley, H. Sakai, M. Scott, A. Stolz, T. Suzuki, L. Valdez, and K. Yako, “Extraction of Gamow-Teller strength distributions from 56 Ni and 55 Co via the (p,n) reaction in inverse kinematics,” Physical Review C, vol. 86, p. 034324, 2012. [65] K. Langanke and G. Martı́nez-Pinedo, “Nuclear weak-interaction processes in stars,” Reviews of Modern Physics, vol. 75, p. 819, 2003. 166 [66] R. G. T. Zegers, R. Meharchand, Y. Shimbara, S. M. Austin, D. Bazin, B. A. Brown, C. A. Diget, A. Gade, C. J. Guess, M. Hausmann, G. W. Hitt, M. E. Howard, M. King, D. Miller, S. Noji, a. Signoracci, K. Starosta, C. Tur, C. Vaman, P. Voss, D. Weisshaar, and J. Yurkon, “34 P(7 Li,7 Be+γ) reaction at 100A MeV in inverse kinematics.,” Physical Review Letters, vol. 104, p. 212504, 2010. [67] K. Yako Unpublished. [68] Y. Satou, T. Nakamura, Y. Kondo, N. Matsui, Y. Hashimoto, T. Nakabayashi, T. Okumura, M. Shinohara, N. Fukuda, T. Sugimoto, H. Otsu, Y. Togano, T. Motobayashi, H. Sakurai, Y. Yanagisawa, N. Aoi, S. Takeuchi, T. Gomi, M. Ishihara, S. Kawai, H. J. Ong, T. K. Onishi, S. Shimoura, M. Tamaki, T. Kobayashi, Y. Matsuda, N. Endo, and M. Kitayama, “14 Be(p,n)14 B reaction at 69 MeV in inverse kinematics,” Physics Letters B, vol. 697, p. 459, 2011. [69] D. Alburger and D. Wilkinson, “Beta decay of 16 C and 17 N ,” Physical Review C, vol. 13, p. 835, 1976. [70] C. A. Gagliardi, G. Garvey, N. Jarmie, and R. Roberston, “0+ →0− beta decay of 16 C,” Physical Review C, vol. 27, p. 1353(R), 1983. [71] S. Grévy, N. L. Achouri, J. C. Angélique, C. Borcea, A. Buta, F. De Oliveira, M. Lewitowicz, E. Liénard, T. Martin, F. Negoita, N. A. Orr, J. Peter, S. Pietri, and C. Timis, “Observation of a new transition in the β-delayed neutron decay of 16 C,” Physical Review C, vol. 63, p. 037302, 2001. [72] K. Snover, E. Adelberger, P. G. Ikossi, and B. A. Brown, “Proton capture to excited states of 16 O M1, E1, and Gamow-Teller transitions and shell model calculations,” Physical Review C, vol. 27, p. 1837, 1983. [73] E. K. Warburton and B. A. Brown, “Effective interactions for the 0p1s0d nuclear shell-model space,” Physical Review C, vol. 46, p. 923, 1992. [74] G. F. Bertsch and H. Esbensen, “The (p,n) reaction and the nucleon-nucleon force,” Reports on Progress in Physics, vol. 50, p. 607, 1987. [75] A. Arima, “History of giant resonances and quenching,” Nuclear Physics A, vol. 649, p. 260, 1999. [76] B. A. Brown and B. H. Wildenthal, “Experimental and theoretical Gamow-Teller beta-decay observables for the sd-shell nuclei,” Atomic Data and Nuclear Data Tables, vol. 33, p. 347, 1985. [77] B. H. Wildenthal, M. S. Curtin, and B. A. Brown, “Predicted features of the beta decay of neutron-rich sd-shell nuclei,” Physical Review C, vol. 28, p. 1343, 1983. 167 [78] R. G. Zegers, H. Akimune, S. M. Austin, D. Bazin, A. M. D. Berg, G. P. Berg, B. A. Brown, J. Brown, A. L. Cole, I. Daito, Y. Fujita, M. Fujiwara, S. Galès, M. N. Harakeh, H. Hashimoto, R. Hayami, G. W. Hitt, M. E. Howard, M. Itoh, J. Jänecke, T. Kawabata, K. Kawase, M. Kinoshita, T. Nakamura, K. Nakanishi, S. Nakayama, S. Okumura, W. A. Richter, D. A. Roberts, B. M. Sherrill, Y. Shimbara, M. Steiner, M. Uchida, H. Ueno, T. Yamagata, and M. Yosoi, “The (t,3 He) and (3 He, t) reactions as probes of Gamow-Teller strength,” Physical Review C, vol. 74, p. 024309, 2006. [79] K. Yako, H. Sakai, M. B. Greenfield, K. Hatanaka, M. Hatano, J. Kamiya, H. Kato, Y. Kitamura, Y. Maeda, C. L. Morris, H. Okamura, J. Rapaport, T. Saito, Y. Sakemi, K. Sekiguchi, Y. Shimizu, K. Suda, A. Tamii, N. Uchigashima, and T. Wakasa, “Determination of the Gamow-Teller quenching factor from charge exchange reactions on 90 Zr,” Physics Letters B, vol. 615, p. 193, 2005. [80] K. J. Raywood, B. M. Spicer, S. Yen, S. A. Long, M. A. Moinester, R. Abegg, W. P. Alford, A. Celler, T. E. Drake, D. Frekers, P. E. Green, O. Häusser, R. L. Helmer, R. S. Henderson, K. H. Hicks, K. P. Jackson, R. G. Jeppesen, J. D. King, N. S. P. King, C. A. Miller, V. C. Officer, R. Schubank, G. G. Shute, M. Vetterli, J. Watson, and A. I. Yavin, “Spin-flip isovector giant resonances from the 90 Zr(n,p)90 Y reaction at 198 MeV,” Physical Review C, vol. 41, p. 2836, 1990. [81] T. Wakasa, H. Sakai, H. Okamura, H. Otsu, S. Fujita, S. Ishida, N. Sakamoto, T. Uesaka, Y. Satou, M. Greenfield, and K. Hatanaka, “Gamow-Teller strength of 90 Nb in the continuum studied via multipole decomposition analysis of the 90 Zr(p,n) reaction at 295 MeV,” Physical Review C, vol. 55, p. 2909, 1997. [82] H. Sakai and K. Yako, “Experimental determination of Gamow-Teller quenching value, Landau-Migdal parameter gN ∆ and pion condensation,” Nuclear Physics A, vol. 731, p. 94, 2004. [83] G. Martı́nez-Pinedo, A. Poves, E. Caurier, and A. P. Zuker, “Effective gA in the pf shell,” Physical Review C, vol. 53, p. R2602, 1996. [84] F. M. Nunes and I. J. Thompson, Nuclear Reactions for Astrophysics. Cambridge University Press, 2009. [85] W. G. Love and M. A. Franey, “Effective nucleon-nucleon interaction for scattering at intermediate energies,” Physical Review C, vol. 24, p. 1073, 1981. [86] M. A. Franey and W. G. Love, “Nucleon-nucleon t-matrix interaction for scattering at intermediate energies,” Physical Review C, vol. 31, p. 488, 1985. [87] R. Schaeffer, J. Raynal, and J. Comfort, “Program DW81,” unpublished, 1981. 168 [88] J. Raynal, “Multipole Expansion of a Two-Body Interaction in Helicity Formalism and its Applications to Nuclear structure and Nuclear Reaction Calculations,” Nuclear Physics A, vol. 97, p. 572, 1967. [89] P. Schwandt and M. D. Kaitchuck, “Analyzing power of proton-nucleus elastic scattering between 80 and 180 MeV,” Physical Review C, vol. 26, p. 55, 1982. [90] G. Perdikakis, M. Sasano, S. M. Austin, D. Bazin, C. Caesar, S. Cannon, J. M. Deaven, H. J. Doster, C. J. Guess, G. W. Hitt, J. Marks, R. Meharchand, D. T. Nguyen, D. Peterman, A. Prinke, M. Scott, Y. Shimbara, K. Thorne, L. Valdez, and R. G. T. Zegers, “LENDA: A low energy neutron detector array for experiments with radioactive beams in inverse kinematics,” Nuclear Instruments and Methods in Physics Research A, vol. 686, p. 117, 2012. [91] S. Lipschutz, R. Zegers, J. Hill, S. Liddick, S. Noji, C. Prokop, M. Scott, M. Solt, C. Sullivan, and J. Tompkins, “Digital data acquisition for the Low Energy Neutron Detector Array (LENDA),” Nuclear Instruments and Methods in Physics Research A, vol. 815, p. 1, 2016. [92] S. V. Paulauskas, M. Madurga, R. Grzywacz, D. Miller, S. Padgett, and H. Tan, “A digital data acquisition framework for the Versatile Array of Neutron Detectors at Low Energy (VANDLE),” Nuclear Instruments and Methods in Physics Research A, vol. 737, p. 22, 2014. [93] W. A. Peters, S. Ilyushkin, M. Madurga, C. Matei, S. V. Paulauskas, R. K. Grzywacz, D. W. Bardayan, C. R. Brune, J. Allen, J. M. Allen, Z. Bergstrom, J. Blackmon, N. T. Brewer, J. A. Cizewski, P. Copp, M. E. Howard, R. Ikeyama, R. L. Kozub, B. Manning, T. N. Massey, M. Matos, E. Merino, P. D. O’Malley, F. Raiola, C. S. Reingold, F. Sarazin, I. Spassova, S. Taylor, and D. Walter, “Performance of the Versatile Array of Neutron Detectors at Low Energy (VANDLE),” Nuclear Instruments and Methods in Physics Research A, vol. 836, p. 122, 2016. [94] D. Bazin, J. A. Caggiano, B. M. Sherrill, J. Yurkon, and A. Zeller, “The S800 spectrograph,” Nuclear Instruments and Methods in Physics Research B, vol. 204, p. 629, 2003. [95] D. J. Morrissey, B. M. Sherrill, M. Steiner, A. Stolz, and I. Wiedenhoever, “Commissioning the A1900 projectile fragment separator,” Nuclear Instruments and Methods in Physics Research B, vol. 204, p. 90, 2003. [96] P. A. Zavodszky, B. Arend, D. Cole, J. DeKamp, G. Machicoane, F. Marti, P. Miller, J. Moskalik, J. Ottarson, J. Vincent, and A. Zeller, “Design of SuSI - superconducting source for ions at NSCL/MSU - II. The conventional parts,” AIP Conference Proceedings, vol. 749, p. 131, 2005. 169 [97] P. A. Zavodszky, B. Arend, D. Cole, J. Dekamp, G. MacHicoane, F. Marti, P. Miller, J. Moskalik, J. Ottarson, J. Vincent, A. Zeller, and N. Y. Kazarinov, “Status report on the design and construction of the Superconducting Source for Ions at the National Superconducting Cyclotron Laboratory/Michigan State University,” Review of Scientific Instruments, vol. 77, p. 03A334, 2006. [98] G. Machicoane, D. Cole, J. Ottarson, J. Stetson, and P. Zavodszky, “ARTEMIS-B: A room-temperature test electron cyclotron resonance ion source for the National Superconducting Cyclotron Laboratory at Michigan State University,” Review of Scientific Instruments, vol. 77, p. 03A322, 2006. [99] X. Wu, H. Blosser, D. Johnson, F. Marti, and R. York, “The K500-to-K1200 coupling line for the coupled cyclotron facility at the NSCL,” Proceedings of the 1999 Particle Accelerator Conference, vol. 2, p. 1318 vol.2, 1999. [100] D. Morrissey, “The coupled cyclotron project at the NSCL,” Nuclear Physics A, vol. 616, p. 45, 1997. [101] J. Yurkon, D. Bazin, W. Benenson, D. J. Morrissey, B. M. Sherrill, D. Swan, and R. Swanson, “Focal plane detector for the S800 high-resolution spectrometer,” Nuclear Instruments and Methods in Physics Research A, vol. 422, p. 291, 1999. [102] G. W. Hitt, The 64 Zn(t,3 He) Charge-Exchange Reaction at 115 MeV Per Nucleon and Application to 64 Zn Stellar Electron-Capture. PhD thesis, Michigan State University, 2008. [103] G. F. Knoll, Radiation Detection and Measurement, vol. 3. John Wiley and Sons, 2010. [104] R. Zegers. private communication. [105] XIA, “Digital Gamma Finder (DGF) PIXIE-16 User’s Manual,” 2005. [106] C. Prokop, S. Liddick, N. Larson, S. Suchyta, and J. Tompkins, “Optimization of the National Superconducting Cyclotron Laboratory Digital Data Acquisition System for use with fast scintillator detectors,” Nuclear Instruments and Methods in Physics Research A, vol. 792, p. 81, 2015. [107] C. J. Prokop, S. N. Liddick, B. L. Abromeit, A. T. Chemey, N. R. Larson, S. Suchyta, and J. R. Tompkins, “Digital data acquisition system implementation at the National Superconducting Cyclotron Laboratory,” Nuclear Instruments and Methods in Physics Research A, vol. 741, p. 163, 2014. [108] H. Tan, M. Momayezi, A. Fallu-Labruyere, Y. X. Chu, and W. K. Warburton, “A fast digital filter algorithm for gamma-ray spectroscopy with double-exponential decaying scintillators,” IEEE Transactions on Nuclear Science, vol. 51, p. 1541, 2004. 170 [109] N. Kudomi, “Energy calibration of plastic scintillators for low energy electrons by using Compton scatterings of γ rays,” Nuclear Instruments and Methods in Physics Research A, vol. 430, p. 96, 1999. [110] M. Berz, K. Joh, J. A. Nolen, B. M. Sherrill, and A. F. Zeller, “Reconstructive correction of aberrations in nuclear particle spectrographs,” Physical Review C, vol. 47, p. 537, 1993. [111] S. Agostinelli et al., “GEANT4 - A simulation toolkit,” Nuclear Instruments and Methods in Physics Research A, vol. 506, p. 250, 2003. [112] E. Mendoza, D. Cano-Ott, T. Koi, and C. Guerrero, “New standard evaluated neutron cross section libraries for the GEANT4 code and first verification,” IEEE Transactions on Nuclear Science, vol. 61, p. 2357, 2014. [113] L. Stuhl, A. Krasznahorkay, M. Csatlós, A. Algora, J. Gulyás, G. Kalinka, J. Timár, N. Kalantar-Nayestanaki, C. Rigollet, S. Bagchi, and M. A. Najafi, “A neutron spectrometer for studying giant resonances with (p,n) reactions in inverse kinematics,” Nuclear Instruments and Methods in Physics Research A, vol. 736, p. 1, 2014. [114] Saint-Gobain, “Organic Scintillation Materials and Assemblies,” Catalogue, p. 16, 2015. [115] T. N. Taddeucci, C. D. Goodman, R. C. Byrd, T. A. Carey, D. J. Horen, J. Rapaport, and E. Sugarbaker, “A neutron polarimeter for (p,n) measurements at intermediate energies,” Nuclear Instruments and Methods in Physics Research A, vol. 241, p. 448, 1985. [116] B. D. Anderson, A. Fazely, R. J. McCarthy, P. C. Tandy, J. W. Watson, R. Madey, W. Bertozzi, T. N. Buti, J. M. Finn, J. Kelly, M. A. Kovash, B. Pugh, B. H. Wildenthal, and C. C. Foster, “Gamow-teller strength in the 18 O(p,n)18 F reaction at 135 MeV,” Physical Review C, vol. 27, p. 1387, 1983. [117] W. T. Chou, E. K. Warburton, and B. A. Brown, “Gamow-Teller beta-decay rates for A≤18 nuclei,” Physical Review C, vol. 47, p. 163, 1993. [118] S. M. Austin, N. Anantaraman, and W. G. Love, “Charge Exchange Reactions and the Efficiency of Solar Neutrino Detectors,” Physical Review Letters, vol. 73, p. 30, 1994. [119] K. Ikeda, “Collective Excitation of Unlike Pair States in Heavier Nuceli,” Progress of Theoretical Physics, vol. 31, p. 434, 1964. 171