VALIDATION OF THE β-OSLO METHOD; AN INDIRECT METHOD FOR CONSTRAINING NEUTRON-CAPTURE CROSS SECTIONS By Katherine Louise Childers A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Chemistry - Doctor of Philosophy 2021 ABSTRACT VALIDATION OF THE β-OSLO METHOD; AN INDIRECT METHOD FOR CONSTRAINING NEUTRON-CAPTURE CROSS SECTIONS By Katherine Louise Childers One of the prevalent questions in nuclear science is the origin of the elements. There are two stellar nucleosynthesis processes considered to be responsible for the production of the majority of the abundances of the elements heavier than iron; the slow neutron- capture process (s-process) and the rapid neutron-capture process (r-process). Both of these processes are characterized by the successive capture of neutrons on nuclei, with the major differences between the processes being the timescale over which the processes occur and the host environment. The s-process occurs in low neutron-density environments, such as low- to intermediate-mass stars, and proceeds slowly along the valley of stability. Since the nuclei involved are close to stability, the reactions involved are amenable to direct measurements. The r-process progresses through an explosive event with high neutron densities which drives material far from stability. The recent observation of a neutron star merger event by LIGO and Virgo and the subsequent electromagnetic follow up has demonstrated that an r-process event can occur in these rare events, but it has not ruled out other potential astrophysical sites. To better understand and model the r-process, several nuclear properties are needed for a large number of nuclei, including neutron-capture cross sections. R-process nuclei are not viable for direct measurement of neutron-capture cross sections since the nuclei involved are far from stability, and thus have short half-lives. Therefore, several indirect measurement techniques have been developed to provide experimental constraints on neutron-capture cross sections. One such method is the β-Oslo method, which uses β decay to populate highly excited states of a nucleus. The resulting de-excitation via the emission of γ rays is used to extract statistical nuclear properties of the daughter nucleus. These properties are then used as input in a reaction model to constrain the neutron-capture cross section. The β- Oslo method can provide a large number of constrained neutron-capture cross sections far from stability, but it is necessary to validate the method using a direct neutron capture measurement. This work will present a validation of the β-Oslo method in the A = 80 mass region with the 82Se(n,γ)83Se reaction. The nuclide 83Se can be accessed through the β-decay of 83As, which was studied at the National Superconducting Cyclotron Laboratory with the total absorption spectrometer, SuN. Using the β-Oslo method, the cross section of the 82Se(n,γ)83Se reaction was constrained. A direct measurement of the 82Se(n,γ)83Se reaction was performed with the Detector for Advanced Neutron Capture Experiments and the cross section obtained from the direct measurement is compared to the cross section determined using the β-Oslo method. The results are in good agreement, validating the β-Oslo method as a viable method for constraining neutron-capture cross sections. To Mom and Dad: Jo and Malcolm Childers iv ACKNOWLEDGMENTS It has definitely been a journey to get to this point, and I have met so many amazing people along the way. It is hard to express in words how much I appreciate all of you, but I will try! First off, I would like to thank my advisor, Sean Liddick. Thank you for giving me so many opportunities to learn, through my research, the many experiments, and all of the conferences and schools you encouraged me to go to. I appreciate your patience and all of the knowledge you have passed on. (Also, thanks for all the fajitas.) I would also like to thank my committee members: Paul Mantica, Dave Morrissey, and Artemis Spyrou. Thank you all for your guidance and support. I have always felt that I could come talk to any of you for advice or help. A huge thank you goes to Aaron Couture. I dont think I would have been able to get through the selenium analysis without you. Thank you for putting up with me for so long (probably a little longer than expected...) and thank you for all the support and advice. I am really lucky to have had the opportunity to work with you, and I have learned so much. I would also like the thank the rest of the DANCE team: Chris Prokop, Cathleen Fry, Shea Mosby, and John Ullmann. Being able to run an experiment out at LANL was an awesome opportunity and I appreciate all of your support so much. I would be remiss to not give a shout out to the Chemistry and Physics faculty at Otterbein University, especially Robin Grote, Joan Esson, Carrigan Hayes and Jerry Jenkins. You guys were supportive throughout undergrad and taught me the skills I needed to be successful in graduate school and become the scientist I am today. I have many fond memories of all of you and my time at Otterbein. v Graduate school would not be possible without the support and encouragement of friends, and I have made some wonderful friends during my time here. Becky Lewis, thank you for welcoming me in when I started grad school and showing me the ropes, and for always cheering me on, even from afar. Ben Crider, thank you for your help (and infinite patience) with analysis and coding. I have learned so much from you. Andrea Richard, a simple thank you isn’t enough. You have been there for me through some tough times, and you always know what to say to support and encourage me. Thank you for all of your mentorship, and most of all, your friendship. Thank you to the many other friends I have made: Kyle Brown, Stephanie Lyons, Mallory Blyth, Caley Harris, Alicia Palmisano, Aaron Chester, and many more. I appreciate you guys so much and will cherish the many memories. Lastly, thank you to my family. Aunt Cathy, thank you for always being so excited and encouraging me to go for what I wanted. The time I have spent with you has been a big part of my life. Those of you that know me know that I of course need to mention the smallest member of my family, my guinea pig Revan. You won’t be able to read this but thank you for being the best little buddy and roommate. Most importantly, thank you to my parents, Jo and Malcolm Childers. I am where I’m at today because of you. Thank you for your endless support, for cheering on every accomplishment, no matter how big. I love you both so much. vi TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Stellar nucleosynthesis of heavy elements . . . . . . . . . . . . . . . . . . . . 1.2 The rapid neutron-capture process . . . . . . . . . . . . . . . . . . . . . . . 1.3 Uncertainties of nuclear properties . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Hauser-Feshbach model of neutron capture . . . . . . . . . . . . . . . . . . . 1.4.1 Optical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4.2 Nuclear level density models . . . . . . . . . . . . . . . . . . . . . . . 1.4.3 . . . . . . . . . . . . . . . . . . . . . 1.5 Indirect techniques for constraining (n,γ) cross sections . . . . . . . . . . . . 1.6 Current status of β-Oslo validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.7 Dissertation Outline γ-ray strength function models Chapter 2 Study of 83As β decay at the NSCL . . . . . . . . . . . . . . . . . 2.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.1 The Summing NaI (SuN) detector . . . . . . . . . . . . . . . . . . . . 2.1.2 Double-sided silicon strip detector . . . . . . . . . . . . . . . . . . . . Isolation of 83As β decay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Chapter 3 β-Oslo Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 Population of highly excited states in 83Se . . . . . . . . . . . . . . . . . . . 3.2 Unfolding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Extraction of primary γ rays . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Extraction of the functional forms of the NLD and γSF . . . . . . . . . . . . 3.5 Normalization of the NLD and γSF . . . . . . . . . . . . . . . . . . . . . . . 3.5.1 NLD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.2 Limited spin population . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.3 γSF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6 Constraint of 82Se(n,γ)83Se via the β-Oslo Method . . . . . . . . . . . . . . Chapter 4 Direct Measurement of 82Se(n,γ)83Se . . . . . . . . . . . . . . . . 4.1 DANCE: The Detector for Advanced Neutron Capture Experiments . . . . . 4.2 Targets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Calibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Neutron Fluence Characterization . . . . . . . . . . . . . . . . . . . . . . . . 4.4.1 Normalization to 197Au . . . . . . . . . . . . . . . . . . . . . . . . . 82Se Data Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.1 Background Subtraction . . . . . . . . . . . . . . . . . . . . . . . . . 4.5 vii 1 1 3 5 7 8 10 14 17 19 19 21 21 22 26 28 48 49 51 54 55 57 58 59 60 62 66 66 71 72 73 76 79 80 4.5.1.1 Scattering Background . . . . . . . . . . . . . . . . . . . . . 4.5.1.2 Target contaminants . . . . . . . . . . . . . . . . . . . . . . 4.5.2 Efficiency Determination . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.3 Cross Section Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 83 87 98 Chapter 5 Results and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . 101 5.1 Comparison of β-Oslo Cross Section to Directly Measured Cross Section . . . 101 5.2 Conclusions and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 viii LIST OF TABLES Table 2.1: Voltage applied to each PMT and scaling factors used for gain matching for SuN. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 Table 2.2: Scale factor and intercept used for the calibration of SuN’s segments. . . . 27 Table 2.3: Scaling factors used for gain matching the front and back strips of the DSSD. 31 Table 2.4: Half-lives of implanted isotopes and their decay products. . . . . . . . . . 32 Table 3.1: GDR parameters determined from the fit of a GLO function to photoab- . . . . . . . . . . . . . . . . . . . . . . . . . . sorption cross section data. 63 Table 3.2: NLD and γSF models available in TALYS. . . . . . . . . . . . . . . . . . 65 Table 4.1: Composition of the enriched 82Se target and the Q-value of neutron capture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . on each isotope. 71 Table 4.2: ESum gate used for each isotope identified in the 82Se target data. . . . . 86 Table 4.3: Efficiencies determined by DICEBOX and GEANT4 simulations (M = [2,5], 2−, 3 2−. 100000 γ-ray cascades 5.0 MeV > ESum > 5.9 MeV) for J π = 1 were simulated for 30 realizations of artificial nuclei. 2+, 1 . . . . . . . . . . . . 100 ix LIST OF FIGURES Figure 1.1: A schematic of the paths of both the s-process and r-process and the . . subsequent decay modes. From [2], c(cid:13) 2006 John Wiley & Sons, Inc. Figure 1.2: Calculated residual abundances produced through the r-process, as a func- tion of mass number. Reprinted from [3], with permission from Elsevier. Figure 1.3: Residual solar r-process abundance pattern (black dots) compared to mod- eled r-process abundance patterns using three different mass models. The lighter bands represent the uncertainty of the calculation with the rms error compared to known masses. The darker bands represent the uncer- tainty when the rms error is artificially reduced to 100 keV. Reprinted from [8], with permission from Elsevier. . . . . . . . . . . . . . . . . . . . Figure 1.4: (a) Comparison of the residual solar r-process abundance pattern to abun- dance patterns modeled for the same three mass models as in Figure 1.3. Here the uncertainty taken into account is that of the β-decay half-lives. (b) Same as (a) but with the uncertainty of neutron-capture rates studied instead of the β-decay half-lives. Reprinted from [8], with permission from Elsevier. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 1.5: Comparison of the 50Ti(n, γ)51Ti cross section constrained using the β- Oslo and Oslo methods (grey crosses) to experimentally determined cross sections (points). Reprinted figure with permission from [68] Copyright 2019 by the American Physical Society. . . . . . . . . . . . . . . . . . . . Figure 2.1: A representation of the layout of the CCF (comprised of the K500 and . . . . . . . . . . K1200 cyclotrons) and the A1900 Fragment Separator. Figure 2.2: A picture of the e14505 setup in the S2 vault. SuN is shown on the right in the large blue box, the DSSD and veto and located in the center of SuN (indicated by the purple box), and the PIN detectors are located upstream of SuN in the red box. . . . . . . . . . . . . . . . . . . . . . . . 2 4 7 8 20 22 22 Figure 2.3: A schematic of the SuN detector. . . . . . . . . . . . . . . . . . . . . . . 23 Figure 2.4: An example of TAS for 60Co. On the left is the decay scheme of 60Co. On the right is a figure from [71], showing the 60Co gamma spectra obtained with (a) a single segment of SuN, (b) the top half of SuN, and (c) the total summation of all segments. Reprinted from [71], with permission from Elsevier. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 x Figure 2.5: A picture of the DSSD used for experiment e14505. . . . . . . . . . . . . 28 Figure 2.6: Gain matched energy spectra for the DSSD low gain setting. The front (top) and back (bottom) strip numbers are shown as a function of channel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . number. Figure 2.7: Gain matched energy spectra for the DSSD hi gain setting, using a 228Th source. The front (top) and back (bottom) strip numbers are shown as a . . . . . . . . . . . . . . . . . . . . . . . . . function of channel number. Figure 2.8: Particle Identification plot. PIN Energy is from the first PIN detector located upstream of SuN. TOF is determined from the timing between the PIN detector and the focal plane scintillator located in the A1900 fragment separator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 30 32 Figure 2.9: SuN total absorption spectrum as a function of time. . . . . . . . . . . . 33 Figure 2.10: (Top) Implantation rate of 83As ions in the DSSD. (Bottom) Implantation rate of 84Se ions in the DSSD. . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.11: Complete level scheme of the decay of 84Se into 84Br from ENSDF. Ab- solute γ-ray intensities per 100 decays listed next to γ-ray energies. [73] . Figure 2.12: (Top) TAS spectrum as a function of time gated on the 408 keV peak corresponding to 84Br (only first three beam cycles shown) with the de- cay profile modeled via the Bateman equations overlaid. (Bottom) TAS spectrum as a function of time gated on the 2053 keV peak that has con- tributions from the decay chain of 83As → 83Se → 83Br (only first three beam cycles shown) with the decay profile of each decay, as well as the . . . . . . . . . . . . total, modeled via the Bateman equations overlaid. Figure 2.13: Activity profiles of all decaying species modeled with the Bateman equa- tions, shown over the full time range. . . . . . . . . . . . . . . . . . . . . 34 35 36 37 Figure 2.14: Regions of production and decay during the pulsed beam setting. . . . . 38 Figure 2.15: Plot of the long decay period during which the beam was not being deliv- . . . . . . . . . . . . . . . . . . . . ered to the experimental end station. Figure 2.16: Excitation energy (Ex) as a function of γ-ray energy (Eγ), ungated, with . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . all events. Figure 2.17: Ex, Eγ matrix with 50 keV binning, for a time gate of 26000 seconds to . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . the end of the run. 39 40 40 xi listed next to γ-ray energies. [74] Figure 2.18: Decay scheme of 83Br → 83Kr. Absolute γ-ray intensities per 100 decays . . . . . . . . . . . . . . . . . . . . . . Figure 2.19: Ex, Eγ matrix with 50 keV binning, for the decay of 83Br → 83Kr simu- lated in Geant4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.20: Ex, Eγ matrix with 50 keV binning, for the decay of 83Se → 83Br. . . . . Figure 2.21: Decay scheme of 83Se → 83Br with selected levels and γ-ray transitions. 41 41 42 Absolute γ-ray intensities per 100 decays listed next to γ-ray energies. [74] 42 Figure 2.22: Ex, Eγ matrix with 50 keV binning, for the decay of 84Se → 84Br simulated in Geant4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . of 83Se → 83Br and 84Br → 84Kr. Figure 2.23: Ex, Eγ matrix with 50 keV binning, for the decays of the isomeric state . . . . . . . . . . . . . . . . . . . . . Figure 2.24: Decay scheme of the isomeric state of 83Se → 83Br with selected levels and γ-ray transitions. Absolute γ-ray intensities per 100 decays listed next to . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . γ-ray energies. [74] Figure 2.25: Decay scheme of 84Br → 84Kr with selected levels and γ-ray transitions. 43 43 44 Absolute γ-ray intensities per 100 decays listed next to γ-ray energies. [73] 45 Figure 2.26: Ex, Eγ matrix with 50 keV binning, for the decay of 83As → 83Se. . . . 45 trum of 83Se. Figure 2.27: (Top) γ-ray energy spectrum of 83Se. (Bottom) Excitation energy spec- . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.28: Decay scheme of 83As → 83Se with selected levels and γ-ray transitions. γ-ray intensities relative to the 734.9 keV transition listed next to γ-ray energies. [74] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.1: Flow chart of Oslo method. Bold items are programs used for the respec- . . . . . . . . . . . . . . . tive steps in the flow chart. Figure from [75]. Figure 3.2: Schematic of the population of states in 83Se through the β decay of 83As . . . . . . . . . . . . . . . . . . . . . . . . and neutron capture on 82Se. 46 47 50 51 Figure 3.3: Unfolded Ex, Eγ matrix for 83Se, with 50 keV binning. . . . . . . . . . . 54 Figure 3.4: Primary Ex, Eγ matrix for 83Se, with 200 keV binning. . . . . . . . . . . 56 xii Figure 3.5: Nuclear level density for 83Se, with the experimental data (black circles) as well as the upper and lower limits (blue band). Known levels are indicated by the solid black line, the level density at the neutron separation energy, ρ(Sn), is indicated by the white square, and the Constant Temperature . . . . . . . . . . model extrapolation is shown as the dashed black line. Figure 3.6: Percentage of the total level density of 83Se determined from Ref. [37] as a function of the spin. Spin range highlighted in blue represents the spin range populated following an allowed β decay and one dipole transition. Figure 3.7: Gamma strength function for 83Se (black circles) with the upper and lower limits based on systematic uncertainty are indicated by the blue band. Photoabsorption cross section data from [81] (green squares, blue triangles, purple circles) and the corresponding fit of the GLO function to the data and an average fit are also shown. . . . . . . . . . . . . . . . . . Figure 3.8: A plot of χ2 values versus a range of Γγ values, obtained from fitting the average GDR resonance parameters to the experimental γSF to determine the best Γγ for the data set. . . . . . . . . . . . . . . . . . . . . . . . . . 60 61 62 63 Figure 3.9: Cross section of the 82Se(n,γ)83Se reaction calculated using TALYS (black). The upper and lower limits (light blue) are based on the uncertainty of the experimental NLD and γSF. The grey band shows the range of cross sections resulting from combinations of all available NLD and γSF models. 64 Figure 4.1: Schematic representation of the DANCE detector geometry, including the . . . . . . . . . . . . . . . . . . . . . . . . . beam-line, and LiH sphere. Figure 4.2: Example of signals from a γ ray and an alpha particle, from a single BaF2 crystal. c(cid:13) 2006 IEEE Reprinted, with permission, from [94]. . . . . . . . Figure 4.3: Short integral of the detector signal versus the long integral of detector signal from the data taken with the 82Se target. Gate used to separate alpha signals shown in red. . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.4: Photographs of 82Se target and target holder from a top down view (left) . . . . . . . . . . . . . . . . . . . . . . . . . . . and a side view (right). Figure 4.5: (Left) Representation of fit used for calibration with a 22Na source. (Right) Calibrated 22Na spectrum. . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.6: Alpha decay spectra of one crystal and representation of the fit used for . . . . . . . . . . . . . . . . . . . . . . . . crystal-by-crystal calibration. 67 68 69 72 73 74 xiii Figure 4.7: (a) Neutron yield measured by the 6Li beam monitor. (b) ENDF/B- VII.1 evaluated cross section of 6Li(n, α)3H used in the neutron fluence determination. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.8: (a) Neutron yield measured by the gas-filled 235U fission chamber. (b) ENDF/B-VII.1 evaluated cross section of 235U(n, f ) used in the neutron fluence determination. . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.9: Neutron fluence at the beam monitor position as a function of neutron energy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.10: 2D plot of ESum as a function of En for the 197Au data, over the 4.89 eV resonance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.11: Figure depicting the subtraction of a linear background from the time- of-flight spectrum of neutron capture on 197Au. The blue histogram is the data pre-subtraction, and the magenta histogram is the data after subtraction of the linear background, shown in red. . . . . . . . . . . . . Figure 4.12: Diagnostic plot showing the ESum projections of the different components of the background subtraction on the 4.89 eV 197Au resonance. The total ESum shape (red) corresponds to the data before the background subtrac- tion was performed. The magenta ESum shape corresponds to the sum of the background regions on either side of the resonance. The blue ESum shape is the background subtracted capture ESum. . . . . . . . . . . . . Figure 4.13: Cross section of neutron capture on 197Au, in the region of the 4.89 eV resonance (black squares). The scaled ENDF/B-VII.1 cross section is shown in red [96]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.14: Counts as a function of the sum of energy deposited in DANCE (ESum) and neutron energy (En) for the 82Se data. . . . . . . . . . . . . . . . . . Figure 4.15: Counts as a function of the sum of energy deposited in DANCE (ESum) and neutron energy (En) for the 208Pb data. . . . . . . . . . . . . . . . . Figure 4.16: Example scattering background subtraction for a single En bin on a 82Se resonance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.17: Counts as a function of the sum of energy deposited in DANCE (ESum) and neutron energy (En) for the 82Se data, after the scattering background was subtracted. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.18: Yield of neutron capture on 82Se, gated on an ESum range of 5.0 - 5.9 . . MeV. Contributions of resonances from other Se isotopes are labeled. 75 75 76 78 79 80 81 82 84 84 85 88 xiv Figure 4.19: ESum projection of (a) 77Se (b) 74Se (c) 78Se and (d) 80Se . . . . . . . . 88 Figure 4.20: Spectra of 74Se (with 77Se contributions removed) gated on an ESum range of 7.7 - 8.6 MeV. . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.21: Yield of neutron capture on 82Se, gated on an ESum range of 5.0 - 5.9 . . . . . . . MeV, after the contribution from 74Se has been subtracted. Figure 4.22: Spectra of 78Se (with 77Se and 74Se contributions removed) gated on an . . . . . . . . . . . . . . . . . . . . . . . . ESum range of 6.8 - 7.7 MeV. Figure 4.23: Yield of neutron capture on 82Se, gated on an ESum range of 5.0 - 5.9 MeV, after the contributions from 74Se and 74Se have been subtracted. . Figure 4.24: Spectra of 80Se (with 77Se, 74Se, and 78Se contributions removed) gated . . . . . . . . . . . . . . . . . . . . on an ESum range of 6.4 - 6.8 MeV. 89 89 90 90 91 Figure 4.25: Yield of neutron capture on 82Se, gated on an ESum range of 5.0 - 5.9 MeV, after the contributions from 74Se, 78Se and 80Se have been subtracted. 91 Figure 4.26: Energy of individual γ rays, gated on multiplicity 2, for several strong reso- nances in the 82Se data compared to simulated data for J π = 1 Points correspond to the average across all realizations, while error bars correspond to the standard deviation of the bin content for each realiza- tion. The integral of each plot is normalized to unity for a true comparison. 94 2+, 1 2−, 3 2−. Figure 4.27: Energy of individual γ rays, gated on multiplicity 3, for several strong reso- nances in the 82Se data compared to simulated data for J π = 1 Points correspond to the average across all realizations, while error bars correspond to the standard deviation of the bin content for each realiza- tion. The integral of each plot is normalized to unity for a true comparison. 95 2+, 1 2−, 3 2−. Figure 4.28: Energy of individual γ rays, gated on multiplicity 4, for several strong reso- nances in the 82Se data compared to simulated data for J π = 1 Points correspond to the average across all realizations, while error bars correspond to the standard deviation of the bin content for each realiza- tion. The integral of each plot is normalized to unity for a true comparison. 96 2+, 1 2−, 3 2−. Figure 4.29: Energy of individual γ rays, gated on multiplicity 5, for several strong reso- nances in the 82Se data compared to simulated data for J π = 1 Points correspond to the average across all realizations, while error bars correspond to the standard deviation of the bin content for each realiza- tion. The integral of each plot is normalized to unity for a true comparison. 97 2+, 1 2−, 3 2−. xv Figure 4.30: Simulation efficiencies and their uncertainties as a function of the lower ESum gate boundary, for M = [2,5] and an upper ESum gate boundary of 5.9 MeV. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.31: Cross section of the 82Se(n,γ)83Se reaction measured with DANCE. . . . 98 99 Figure 5.1: Comparison of the directly measured cross section of 82Se(n,γ)83Se (black) to the neutron capture cross section determined via the β-Oslo method (black line). The blue lines indicate the upper and lower limits of the cross section determined via the β-Oslo method through a systematic study of the uncertainty. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 xvi Chapter 1 Introduction 1.1 Stellar nucleosynthesis of heavy elements The origin of the elements remains one of the mysteries that scientists are still working to understand today. It is known that the majority of the elements on Earth were not created terrestrially and instead were present when the Earth was formed. So where did they come from? The lightest nuclei (hydrogen, helium, and trace amounts of lithium) were created through primordial nucleosynthesis during the Big Bang. After the Big Bang, as galaxies and the earliest stars, or protostars, began forming, conditions became right for fusion reactions to begin, starting a chain of stellar nucleosynthesis pathways. These fusion reactions are responsible for the production of most of the nuclei up to A ∼ 60. However, around A ∼ 60, there is a peak in the nuclear binding energy of nucleons as a function of mass, and combined with the large Coulomb barrier between charged particles, these fusion reactions are no longer energetically favored. A reaction that is energetically favorable is neutron capture, since there is no Coulomb barrier hindering the addition of a neutron to nuclei. Neutron capture is instead influenced by the free neutron density in the stellar environment. There are two main neutron capture processes, the slow neutron-capture process (s-process), and the rapid neutron-capture process (r-process). The s-process is characterized by a low neutron density (∼ 1010/cm3 ) environment [1], which leads to a relatively long time between 1 neutron captures, hundreds or thousands of years, compared to the relatively quicker β- decay half lives. This keeps the s-process close to the valley of stability. A schematic of the s-process pathway is shown in Figure 1.1. Since the nuclei involved in the s-process are close to stability, many of the properties of these nuclei have been experimentally determined, and therefore, through the use of astrophysical models, the basic mechanisms and astrophysical sites of the s-process are fairly well constrained. Currently, theoretical studies support low to intermediate mass asymptotic giant branch (AGB) stars as one of the sites of the s- process [1]. Unlike the s-process, there is much that is still unknown about the r-process, which will be discussed in more detail in the following section. Figure 1.1: A schematic of the paths of both the s-process and r-process and the subsequent decay modes. From [2], c(cid:13) 2006 John Wiley & Sons, Inc. 2 1.2 The rapid neutron-capture process The r-process occurs in stellar environments with a much higher neutron density (∼ 1020/cm3) than the s-process, and because of this, the rate of neutron capture is also much higher. In most cases, the time scale of neutron capture reactions will be much less than β-decay half-lives (on the order of fractions of a second), leading to many rapid, sequential neu- tron captures before β decaying, creating a zig-zag pattern progressing up the chart of the nuclides. This drives the r-process farther from stability, as shown in Figure 1.1, to the neutron-rich region of the chart of the nuclides. Eventually, the neutron density will drop, neutron-capture reactions will cease, and the neutron-rich nuclei produced will β decay back to stability, populating the nuclei in between. The observed solar abundance of heavy nuclei produced is assumed to be a combination of both the s-process and the r-process. (As a note: The solar abundances are considered here as the default abundances of elements in the universe, or the cosmic abundances.) Since the s-process is better understood and easier to model, the solar abundance pattern resulting from the r-process is determined by subtracting the contribution of the s-process from the total observed solar abundance. The resulting r-process residual abundance pattern is shown in Figure 1.2. The three main peaks that are visible in the r-process abundance, at A = 80, 135, and 195, correspond to the shell closures at the magic neutron numbers of 50, 82, and 126 (see Figure 1.1). Nuclei with these magic numbers will have additional stability, and will therefore accumulate as the reaction flow passes through them, leading to an increase in their abundance. Due to the exotic nature of the nuclei produced along the r-process path, there is uncer- tainty in the properties of many of these nuclei, as well as the general astrophysical conditions 3 Figure 1.2: Calculated residual abundances produced through the r-process, as a function of mass number. Reprinted from [3], with permission from Elsevier. that drive the r-process. In order to define the astrophysical conditions of the r-process, the astrophysical locations must be determined. While there have been many locations proposed, the most likely candidates are supernovae and neutron star mergers [1, 3]. Unfortunately, there has been no evidence of r-process nuclei observed in supernova events so far, and even the most promising models have issues, such as not producing a successful explosion or the heaviest nuclei, that prevent a conclusive result [1]. Neutron-star mergers were originally put forth as a potential r-process site, as this site has the high neutron density needed to support the r-process. However, until recently, a neutron-star merger had never been observed, and thus were not considered to be as likely as supernova. In August of 2017, the Advanced LIGO and Advanced Virgo gravitational-wave detectors made the first observation of a neu- tron star merger, GW170817 [4]. The subsequent transient kilonova event that was observed had characteristics consistent with what would be expected of an r-process environment and 4 the production of lanthanide elements [5–7], supporting neutron-star mergers as a site of the r-process. There are still many questions remaining about astrophysical conditions and other possible astrophysical sites have been hypothesized [8]. Every astrophysical site is characterized by distinct conditions, including the temperature, neutron density, and initial composition. These different conditions will produce a unique abundance pattern, which can be modeled and compared to the observed solar r-process abundance pattern. Comparisons have proven to be difficult, however, due to the uncertainties the nuclear properties that are needed for these models, namely, the masses, β-decay half lives, and neutron-capture cross sections. 1.3 Uncertainties of nuclear properties In order to determine the impact that the uncertainties of nuclear properties have on mod- eled abundance patterns, a study employing Monte Carlo simulations was performed by Mumpower et al. [8]. The masses, β-decay half-lives, and neutron-capture cross sections over a large mass range were varied using a probability distribution based on the current uncertainties of each property, and for each set of inputs, a simulated r-process abundance pattern was generated. Shown in Figure 1.3 are abundance patterns calculated using three different mass models (HFB-17 [9], DZ [10], and FRDM1995 [11]). Based on a comparison to measured masses, the uncertainty of the masses calculated via the models was determined to be 500 keV rms. The uncertainty in the calculated abundance pattern, based on the propagation of the mass uncertainty, can be large, causing difficulty in comparing to the known abundance pattern. If an artificial reduction of the uncertainty is performed, to 100 keV rms (indicated by the dark band in Figure 1.3), then the calculated abundance pat- 5 tern become more distinct, but in some areas the uncertainty still remains large. However, without this reduction, the light band represents the current predictive power of these mass models. For β-decay half lives, current uncertainties were approximated to be a factor of 10 based on a comparison of known experimental half-lives to a global QRPA calculation of half- lives [8]. For neutron-capture cross sections, a comparison of experimental values to those calculated using three different models (TALYS [12], NONSMOKER [13], and CIGAR [14]) led to the determination of a current uncertainty of a factor of 1000. Shown in Figure 1.4 are abundance patterns calculated with the uncertainties associated with β-decay half lives and neutron-capture rates. Again, these uncertainty bands are large. With the uncertainty of the masses, β-decay half-lives, and neutron-capture cross sections combined, it becomes clear that the current theoretical models used to predict these inputs lead to error bars that are too large to distinguish between different astrophysical models. Therefore, the reduction of uncertainty introduced into the calculation of abundance patterns from nuclear properties of heavy nuclei is of great importance. There have been advances in the measurements of both masses ( see References [15–18] for a selection of recent work) and β-decay half-lives (see References [19–22]) of nuclei that are critical for r-process simulations. However, the un- certainty in neutron-capture cross sections for most r-process nuclei remain large due to the difficulty of directly measuring these reactions. A direct measurement of a neutron-capture reaction requires a target of the nucleus of interest, which, in the case of r-process nuclei far from stability, is not feasible due to their short half lives. This has led to the development of indirect methods of constraining neutron-capture cross sections using experimental infor- mation [23]. Many of these indirect methods (as well as current theoretical cross-section calculations) rely on the Hauser-Feshbach statistical model of neutron capture. 6 Figure 1.3: Residual solar r-process abundance pattern (black dots) compared to modeled r-process abundance patterns using three different mass models. The lighter bands represent the uncertainty of the calculation with the rms error compared to known masses. The darker bands represent the uncertainty when the rms error is artificially reduced to 100 keV. Reprinted from [8], with permission from Elsevier. 1.4 Hauser-Feshbach model of neutron capture In lieu of directly measuring neutron-capture reactions, the cross section can be calculated using the Hauser-Feshbach model of neutron capture [24]. The Hauser-Feshbach model describes neutron capture as a two-step process; first, a compound nucleus is formed, and second, the compound nucleus decays via the emission of γ rays. It is assumed that the second step is independent of the first step [25], and therefore, the subsequent decay of the compound nucleus is independent of the method through which the nucleus was formed. Instead, the decay is governed by the statistical properties of the nucleus. The Hauser- Feshbach model relies on information about the interaction between the neutron and the target nucleus (optical model potential - discussed in Section 1.4.1), as well as information 7 Figure 1.4: (a) Comparison of the residual solar r-process abundance pattern to abundance patterns modeled for the same three mass models as in Figure 1.3. Here the uncertainty taken into account is that of the β-decay half-lives. (b) Same as (a) but with the uncertainty of neutron-capture rates studied instead of the β-decay half-lives. Reprinted from [8], with permission from Elsevier. describing statistical properties of the compound nucleus formed during the reaction (nuclear level density - Section 1.4.2 and γ-ray strength function - Section 1.4.3). In many cases, especially for r-process nuclei, most of this information is not known, so theoretical models are required to obtain the necessary information. Validating indirect means of inferring neutron-capture rates, which is the subject of this thesis, is necessary to provide confidence when extrapolating theoretical models to unknown nuclei. 1.4.1 Optical model An optical model potential is used to describe the interaction that takes place between the target nucleus and the incoming neutron. The central assumption of the optical model 8 is that the complicated interaction can be represented by a single mean field potential. There are two main types of optical model potentials - a phenomenological potential, and a microscopic potential. The optical model potential most often used for r-process calculations is that of Koning and Delaroche [26], which is a phenomenological optical model. A suitable analytical form for the potential is chosen (often the Woods-Saxon potential), and parameters are determined via fitting to experimental data, where available. If experimental data are not available, a global parameterization is used. In a microscopic optical model potential, the strength and shape of the nuclear potential is calculated by folding the nucleon-nucleon effective interaction with the nuclear density distribution. A semi-microscopic (in between phenomenological and microscopic) optical model potential that is available and commonly used is the Jeukenne-Lejeune-Mahaux (JLM) model [27]. This potential usually results in a calculated neutron-capture cross section that is within 20% of what would be obtained using the Koning-Delaroche model. The most important quality (particularly for r-process nuclei) of an optical model potential is that it is able to reliably predict nuclear properties that are unable to be measured experimentally. Fortunately, for most r-process nuclei, the uncertainty propagated from the optical model potential is the smallest source of uncertainty in the calculated cross section. Most of the uncertainty in the results of the Hauser-Feshbach calculation comes from the models used for calculating the properties that will be discussed next - the nuclear level density and the γ-ray strength function. 9 1.4.2 Nuclear level density models The nuclear level density (NLD) is the number of energy levels available at a given excitation energy (Ex), spin (J), and parity (π). It can be represented by the function: ρ(Ex, J, π) = ∆N (Ex, J, π)/∆Ex, (1.1) where ∆N (Ex, J, π) is the number of levels with the energy bin ∆Ex. A level spacing parameter, D(Ex, J, π), can be determined as the inverse of the level density, D(Ex, J, π) = 1/ρ(Ex, J, π). (1.2) The total level density within an excitation energy bin can then be calculated by summing over all spins and parities. In cases where only a subset of levels are populated, it is important to know the distribution of the spin and parity amongst the states in the nucleus. The spin and parity distributions are commonly assumed to be uncorrelated. The spin distribution, s(Ex, J), is dependent on excitation energy and spin, and following Ericson [28], and can be approximated as: s(Ex, J) (cid:39) 2J + 1 2σ2(Ex) − (J+1/2)2 2σ2(Ex) . e (1.3) The σ2 parameter in Eq. 1.3 is called the spin cut-off parameter. This expression is de- rived assuming that many particles and holes are excited to higher-lying states, and has been found to be an appropriate approximation for J ≤ 30 [28]. In a study done by von Egidy and Bucurescu [29], a set of 310 nuclei were fit with Eq. 1.4 to obtain an empirical 10 parameterization of σ2 based on its mass dependence. σ2(Ex) = p1Ap2(E − 0.5P ´a)p3 (1.4) From the fit of the above equation, it was determined that p1 = 0.391, p2 = 0.675, and p3 = 0.312. P ´a is the deuteron pairing energy, which can be calculated using mass excess values M (A, Z) via: P ´a = 1 2 [M (A + 2, Z + 1) − 2M (A, Z) + M (A − 2, Z − 1)]. (1.5) The fit was repeated separately for sets of only the even-even, odd, and odd-odd nuclei and the parameters determined from the fit were found to be essentially the same as the parameters listed above, confirming this parameterization to be independent of nuclear shell structure differences. It is worth noting that the nuclei used in this study were all stable or close to stability. There is a trend in experimentally determined spin distribution favoring lower spins at lower excitation energies and the distribution shifting to higher spins as the excitation energy increases. It is assumed that this trend will hold for nuclei further from stability, but as of now, the spin distribution is not considered to be well known for nuclei at the far reaches of the nuclear chart. The two parities (positive and negative) are usually assumed to be distributed equally across all nuclear states, regardless of energy. It has been shown through several experimental and theoretical studies that even if there is a deviation from an equal number of positive and negative parities, it does not have a large effect of the determination of the level density [23]. 11 The first nuclear level density model developed was the Fermi gas (FG) model, which described the nucleus as a gas of nucleons arranged in single-particle orbits. These nucleons behave in accordance with the Pauli exclusion principle, which states that they must each have a unique set of quantum numbers. In the lowest energy configuration of a nucleus, the nucleons will fill up the lowest-energy single-particle states. When energy is added to the nucleus, nucleons are excited to higher energy single particle states. The number of available configurations and the density of the single particle states will both increase with an increasing amount of energy available to the system. Using this description, Bethe determined a level density function [30], ρ(Ex) = aEx √ πe2 5/4 12a1/2E x (1.6) where Ex is the excitation energy, and a is the level density parameter, determined using the single particle level density parameters of protons (gp) and neutrons (gn), √ π 6 a = (gp + gn). (1.7) Since nucleons do interact, a modification was made to the FG model, and led to the Back- shifted Fermi Gas (BSFG) model [31]: ρ(Ex) = √ πe2 a(Ex−∆) 12a1/2(Ex − ∆)5/4 . (1.8) The BSFG model includes a shift parameter, ∆, which is intended to take into account the effect of the interactions of nucleons, based on the separation energy the interacting nucleons must overcome [32,33]. ∆ can be calculated from the pairing energy of protons and neutrons, 12 but along with the level density parameter a, it can be adjusted to match an experimental nuclear level density [12]. Another commonly used level density model is the Constant Temperature (CT) model [34], which describes the level density by the function, ρ(Ex) = e(Ex−E0)/T 1 T (1.9) where Ex is the excitation energy and E0 and T are free parameters that are related to an energy shift and constant temperature, respectively, that can be found by fitting to experimental data. The CT model was found to reproduce known nuclear level densities are lower excitation energies, while the FG and BSFG models were found to reproduce level densities are higher energies better [34, 35]. For this reason, the CT and BSFG models are often used together to describe a full range of excitation energy. While the nuclear level density models described above do a reasonable job at reproducing experimental data, they are unable to describe some of the finer details in the level density. When used to extrapolate to nuclei far from the valley of stability, the nuclear level density models become more unreliable, with no experimental data to help constrain the model, which leads to a larger uncertainty in the calculated cross section. Ideally, a microscopic model, based on fundamental interactions and first principles, would be used to calculate level densities. However, the limits of computational power limit the applicability of microscopic approaches to determining the nuclear level density. A recently developed approach by Goriely et al. has been used to calculate level densities for a large range of nuclei (up to 150 MeV and J=30) using a Hartree-Fock basis [36]. Several other versions of this approach have also been developed; the spin- and parity-dependent combined Hartree-Fock-Bogoliubov and 13 combinatorial method [37], and the temperature-dependent Hartree-Fock-Bogolyubov-plus- combinatorial method [38]. To date, the efficacy of microscopic methods for level densities of nuclei far from stability has been difficult to determine due to the scarcity of data available for comparison. 1.4.3 γ-ray strength function models The γ-ray transmission coefficient T (Eγ) represents the probability of a γ ray of energy Eγ being emitted from a nucleus. T (Eγ) is related to the γ-ray strength function (γSF), f (Eγ), through the expression TXL(Eγ) = 2πE (2L+1) γ fXL(Eγ), (1.10) where X denotes the type of electromagnetic radiation the transition is (E for electric and M for magnetic), and L denotes the angular momentum. In general, the strength function can be defined in terms of the average partial radiative width, ΓXL, of a γ ray of energy Eγ from an initial energy level Ei, with a spin Ji and parity πi, and the level density at that energy, ρ(Ei, Ji, πi) [39]: ←− f XL(Ei, Ji, πi, Eγ) = (cid:104)ΓXL(Ei, Ji, πi, Eγ)(cid:105)ρ(Ei, Ji, πi) (2L+1) E γ . (1.11) Through the principle of detailed balance, this “downward” strength function (designated by the left arrow symbol) describing γ decay, can be connected to the “upward” strength function (designated by the right arrow symbol), which describes the reverse process of photoabsorption. This “upward” strength function can be described in terms of the average 14 photoabsorption cross section, σ(Eγ) [39]: −→ f XL(Ef , Jf , πf , Eγ) = 1 (cid:104)σXL(Ef , Jf , πf , Eγ)(cid:105) (2L + 1)(π¯hc)2 (2L+1) γ E , (1.12) where Ef is the final state after absorption, with a spin Jf and parity πf . One feature of the photoabsorption cross section that has been observed to be constant throughout the nuclear chart is a broad E1 resonance, called the Giant Dipole Resonance (GDR). Through the Brink hypothesis [40,41], which states that the shape of the photoabsorption cross section is independent of the initial excitation energy of the nucleus, it can be assumed that the shape of the E1 cross section at an excited state will be the same as at the ground state. Through this assumption and the connection of the photoabsorption cross section to the γSF, the γSF can be described using the shape of the lower energy tail of the GDR. Therefore, the GDR can be used to help constrain models of the γSF, since there is generally more data available on the GDR than on the γSF. In order to describe the shape of the GDR, and in turn the γSF, several phenomenological models have been developed. One of the first functions to be used to fit the GDR is the Standard Lorentzian (SLO) function, which was developed by Brink and Axel [40, 41]. The SLO function is characterized by the energy (EXL), strength (σXL), and width (ΓXL) of the GDR: fXL(Eγ) = KXL where KXL is defined as: (E2 σXLEγΓ2 γ − E2 XL XL)2 + E2 γΓ2 XL , (1.13) KXL = 1 (2L + 1)π2¯h2c2 . (1.14) 15 This function was found to fit the GDR well and was thus used to represent the E1 strength, considered the dominant contribution to the GDR, as well as contributions from M1 and E2 transitions. Kopecky and Uhl introduced a temperature dependence to the SLO, in order to account for a discrepancy between the SLO function and GDR data at lower energies, that ultimately, can impact calculations of the neutron-capture cross section [42]. This new function, the Generalized Lorentzian, is predominantly used to represent the E1 component of the GDR and is also characterized by the energy (EE1), strength (σE1), and width (ΓE1) of the GDR: (cid:32) fE1(Eγ, T ) = KE1 where (E2 γ − E2 Eγ ˜ΓE1(Eγ) E1)2 + E2 γ ˜ΓE1(Eγ)2 (cid:33) + 0.7ΓE14π2T 2 E3 E1 σE1ΓE1, (1.15) The temperature, T , can be approximated by T ≈(cid:112)Ex/a, and is connected to the Fermi E2 E1 (1.16) . ˜ΓE1(Eγ) = ΓE1 E2 γ + 4π2T 2 Gas nuclear temperature, but currently it is often used as a free parameter when fitting data. This function differs from the SLO function most significantly at lower γ-ray energies, but it has been found to correct discrepancies in cross section calculations using the SLO function. Recently, a surprising feature of the γSF was discovered - an enhancement in the strength at low γ-ray energies (below 4 MeV) and high excitation energies. This feature has been called the “upbend”, and was first seen in Fe isotopes in 2004 [43]. Since then, the upbend has been observed in many other nuclei, including Mo [44, 45], Sc [46, 47], La [48], and Sm [49]. One of the Mo studies used a different technique to determine the γSF [45], thus confirming the upbend as being a feature of the γSF, and not the method used. The upbend has been found to impact the calculation of neutron-capture cross sections, and therefore, 16 it is important to have a good understanding of the characteristics [50]. In a study of 56Fe, the upbend was found to be predominantly dipole in nature [51], and while it is not clear if the upbend is electric or magnetic, it has been shown to have a bias towards a magnetic character [52], and the nuclear shell model also predicts a magnetic character [53]. Like the NLD, it would be best to use microscopic models of the γSF to determine information about the finer details of the γSF and to be able to predict the shape of the γSF for many nuclei, particularly those far from stability. While there has been much research dedicated to microscopic modeling of the γSF, two publications in particular have reported the results of large-scale calculations of γSF parameters. One of the publications used the the quasi-particle random-phase approximation (QRPA) model incorporated into Hartree-Fock models [54] and the other used the QRPA model incorporated into Hartree- Fock-Bogoliubov [55] models. 1.5 Indirect techniques for constraining (n,γ) cross sec- tions The majority of the uncertainty in Hauser-Feshbach calculations of neutron-capture cross sections comes from the unreliability of the nuclear level density and γ-ray strength function models described above. To reduce the uncertainty of these calculations, the nuclear prop- erties needed for neutron-capture cross section calculations need to be constrained based on experimental data. Since neutron-capture reactions on nuclei far from stability are un- able to be measured directly at this time, indirect methods have been developed. Four of the major indirect techniques being used today are: the surrogate method [56], the γ-ray strength function method [57, 58], the Oslo method [59–62], and the β-Oslo method [63]. 17 The surrogate method utilizes charged particle reactions to populate the compound nucleus that would be formed in the neutron-capture reaction of interest. The surrogate reactions can be performed with both traditional stable-beam experiments, or in inverse kinematics experiments. In normal kinematics, the targets available for measurement are restricted to those that are stable or near stability. In inverse kinematics, though, the major limitation is the ability of current facilities to deliver radioactive beams with a sufficient rate for the measurement of interest. A recent validation of the surrogate method, in inverse kinematics, for constraining neutron-capture cross sections cites a beam rate of at least 104 particles per second is needed for successful measurements [64]. The γ-ray strength function method utilizes photodisintigration reactions (γ, n) to study the γ-ray strength function of the com- pound nucleus of interest. The goal is to infer the cross section of the inverse (n, γ) reaction. As with other measurements, the γ-ray strength function method is also limited by the sta- bility of the target needed for the reaction being studied. The Oslo method also utilizes charged particle reactions to study the compound nucleus of the neutron-capture reaction of interest, and is therefore also restricted by the stability of the targets needed. It has been used extensively for nuclei along the valley of stability to extract nuclear level densities and γ-ray strength functions, and has been shown to reliably reproduce known cross sections of neutron-capture reactions. A method based on the Oslo method, the β-Oslo method, has recently been developed. Instead of charged particle reactions, the β-Oslo method utilizes β decay to populate the compound nucleus, which gives it the advantage of being applicable to neutron-rich nuclei that are further from the valley of stability. As the subject of this work, the β-Oslo method will be covered in more detail in Chapter 3. 18 1.6 Current status of β-Oslo validation While the traditional Oslo method has been validated extensively [65–67], the β-Oslo method has not been validated as extensively. To date, there is only one comparison of an (n, γ) cross section constrained via the β-Oslo method to one determined through a direct mea- surement [68]. The cross section of 50Ti(n, γ)51Ti was constrained by both the β-Oslo method (via the measurement of 51Sc→51Ti) and the Oslo method (via the measurement of 50Ti(d, p)51Ti), and these results were then compared to existing data of a direct measure- ment of 50Ti(n, γ)51Ti, as shown in Figure 1.5. All three cross sections were found to be in good agreement, and it was concluded that the β-Oslo method is indeed a valid method for constraining neutron-capture cross sections. However, since 50Ti is a lighter mass nucleus, it has a lower level density, and it cannot be assumed that neutron capture on 50Ti proceeds through a compound system. Therefore, a validation needs to be performed in a heavier mass region, with a nucleus that has a higher level density and is, in principle, a more statistical system. 1.7 Dissertation Outline In this dissertation, the process of validating the β-Oslo method in a high mass region is detailed. Chapter 2 covers the details of the β-decay experiment performed at the Na- tional Superconducting Cyclotron Laboratory and the steps of the data analysis to isolate the β decay of 83As. Chapter 3 provides the details of the β-Oslo method, along with the results of the β-Oslo method applied to the β decay of 83As, to constrain the cross sec- tion of 82Se(n, γ)83Se. Chapter 4 covers the experimental details of a direct measure of 82Se(n, γ)83Se performed at Los Alamos National Laboratory, along with the subsequent 19 Figure 1.5: Comparison of the 50Ti(n, γ)51Ti cross section constrained using the β-Oslo and Oslo methods (grey crosses) to experimentally determined cross sections (points). Reprinted figure with permission from [68] Copyright 2019 by the American Physical Society. data analysis. Chapter 5 shows a comparison of the directly measure neutron-capture cross section to the cross section obtain via the β-Oslo method, and provides a final conclusion and an outlook of future work. 20 Chapter 2 Study of 83As β decay at the NSCL 2.1 Experimental Setup An experiment measuring the β decay of 83As was performed at the National Superconduct- ing Cyclotron Laboratory (NSCL) at Michigan State University in East Lansing, Michigan. A primary beam of 86Kr at 140 MeV/µ was produced by the Coupled Cyclotron Facility (CCF) and impinged on a 188 mg/cm2 9Be target, producing a cocktail beam with multiple fragments. A schematic of the front end of the CCF is shown in Fig. 2.1. The fragments were separated in the A1900 fragment separator [69,70] with a 0.5% momentum acceptance, producing a beam centered on 83As that was delivered to the end station in the S2 vault. At the end station, ions were implanted in a Si double sided strip detector (DSSD) located in the center of the Summing NaI (SuN) detector [71], used to detect β-delayed γ rays. In or- der to veto light ions, which could increase the background, a Si surface detector was placed behind the DSSD. Ions were identified on an event-by-event basis based on a measurement of energy loss in two Si PIN detectors placed upstream of SuN, and the time-of-flight (TOF) measured between a thin scintillator at the focal plane of the A1900 and one of the PIN detectors. A picture of the end station setup is shown in Fig. 2.2. 21 Figure 2.1: A representation of the layout of the CCF (comprised of the K500 and K1200 cyclotrons) and the A1900 Fragment Separator. Figure 2.2: A picture of the e14505 setup in the S2 vault. SuN is shown on the right in the large blue box, the DSSD and veto and located in the center of SuN (indicated by the purple box), and the PIN detectors are located upstream of SuN in the red box. 2.1.1 The Summing NaI (SuN) detector SuN is a total absorption spectrometer (TAS), which employs the γ-summing technique [71]. A cylindrical NaI(Tl) scintillator detector, SuN is 16 inches long and 16 inches in diameter 22 with a 1.8 inch borehole down the center. The SuN detector is depicted in Fig. 2.3. SuN has eight optically-isolated segments of NaI (separated by aluminum), which are each read out by three photomultiplier tubes (PMTs). The signals from each of the PMTs in a segment are added together to obtain the total signal observed in that segment. The total signal from all of the segments are then summed in order to obtain a summed γ-ray spectrum. SuN has a high efficiency for detecting γ rays (85(2)% for the 662 keV transition from the decay of 137Cs) and has nearly 4π coverage. Figure 2.3: A schematic of the SuN detector. There are several methods through which γ rays can interact with the NaI crystals of SuN. The three main interactions are photoelectric absorption, Compton scattering, and pair production. For low energy γ rays (< 1 MeV) the dominant interaction is photoelectric absorption, in which the γ ray is absorbed by an atom. The absorbing atom will then emit a photoelectron. The interaction of the photoelectron with the NaI will then create scintillation light that will be detected by the PMTs. The process of Compton scattering takes place when an incident γ ray interacts with an electron in the detector material. A portion of the energy of the γ ray is transferred to the electron (called the recoil electron), leading to a detection of only the energy transferred to the electron. Compton scattering is the most likely process for mid-energy γ rays. Pair production is a possible process for γ rays 23 above 1.022 MeV, but will most likely happen above 5 MeV. Pair production is the creation of a electron-positron pair through the interaction of the incident γ ray with the Coulomb field of the nucleus. The positron will annihilate, producing two characteristic 511-keV γ rays. One or both of these γ rays have the possibility of interacting with SuN. The γ-summing technique is used to determine the excitation energy of a nucleus after β decay. While SuN can be used in order to obtain the energy of the individual γ rays being emitted in a cascade, the signals from all eight segments can be summed to obtain the full energy deposited from all γ rays in a cascade, which corresponds to energy of the excited state(s) populated through β decay. It is worth noting that there is the chance for multiple γ rays depositing in a single crystal, leading to additional peaks that could appear in the single segment spectrum. It is also possible for some γ rays to escape the volume of the detector, leading to an incomplete recorded energy, which would appear in the spectrum as peaks below the energy of the excited state fed through β decay. The γ-summing technique can be illustrated with an example of 60Co, as shown in Figure 2.4. The β decay of 60Co populates a 2505-keV state in 60Ni, which then decays through two characteristic γ rays (1173 keV and 1332 keV). The γ ray energy spectrum from a single segment of SuN, shown in panel (a) of Figure 2.4, would exhibit two peaks attributed to the full energy of the 1173- and 1332-keV γ-ray transitions in 60Ni. There would also be a small peak at 2505 keV corresponding to the summed energy of the 1173- and 1332-keV γ rays, from instances where both γ rays are simultaneously deposited in the single segment. The spectrum of the top half of SuN is shown in panel (b), where again there are two peaks corresponding to the two γ rays transitions in 60Ni, as well as a larger peak at 2505 keV (relative to the spectrum of a single segment) since the chance of both γ rays depositing in the segments comprising the top half of SuN is higher than the chance of both depositing 24 in only a single segment. The spectrum from the total summation of the signal from all segments (panel (c) of Figure 2.4) shows a “sum peak” at 2505 keV, corresponding to the 2505 keV state populated in the β decay of 60Co. Figure 2.4: An example of TAS for 60Co. On the left is the decay scheme of 60Co. On the right is a figure from [71], showing the 60Co gamma spectra obtained with (a) a single segment of SuN, (b) the top half of SuN, and (c) the total summation of all segments. Reprinted from [71], with permission from Elsevier. SuN’s PMTs were gain matched using the 1460.8 keV γ ray from 40K, which is natu- rally occurring and can be observed in background data. The first step in gain matching is adjusting the high voltage applied to the PMTs, such that the 40K peak appears in ap- proximately the same channel for each PMT. The second step involves the application of a multiplication factor applied in software to precisely center the centroid of the 40K peak in the same channel number for each PMT. The voltages and multiplication factors are listed in Table 2.1. After being gain matched, the energy spectrum of each of the three PMTs of 25 a segment can be summed together in software to determine the segment signal. All eight segments were then calibrated with a linear energy calibration using a 60Co source and a 137Cs source. The parameters determined from the linear calibration are shown in Table 2.2. Table 2.1: Voltage applied to each PMT and scaling factors used for gain matching for SuN. PMT Number Voltage (+V) 0 1 2 3 4 5 6 7 8 9 10 11 730 740 780 774 771 764 794 803 812 838 824 848 Gain Matching Factor 1.0920 1.0000 1.1099 1.1302 1.0000 1.1195 1.1148 1.0000 1.1316 1.0949 1.0000 1.0993 PMT Number Voltage (+V) 12 13 14 15 16 17 18 19 20 21 22 23 835 820 821 831 865 824 889 853 831 909 930 892 Gain Matching Factor 1.0866 1.0000 1.0995 1.0959 1.0000 1.1033 1.1467 1.0000 1.1131 1.0796 1.0000 1.1026 2.1.2 Double-sided silicon strip detector The beam, centered on 83As, delivered to the experimental end station was implanted into a 2.54 x 2.54 cm2 and 1 mm thick double sided silicon detector (DSSD) located in the center of the borehole of SuN. The DSSD consists of 16 electrically-segmented strips on the front, 26 Table 2.2: Scale factor and intercept used for the calibration of SuN’s segments. Segment Number Calibration Scale Calibration Intercept 0 1 2 3 4 5 6 7 0.4193 0.4195 0.4113 0.4260 0.4191 0.4216 0.4046 0.4172 -16.02 -15.95 -14.95 -16.50 -16.41 -17.93 -16.20 - 14.74 and another set of 16 perpendicular strips on the back. These strips provide the ability to define pixels in software. A pixel is defined as the intersection of a front strip and a back strip. Together, these pixels create a grid-like pattern that allows for the determination of the location of events. When an event occurs, the location is defined by the pixel with the highest deposited energy recorded. A picture of the DSSD is shown in Figure 2.5. Two gain ranges were used to capture the full energy of ion implantations (on the order of 1000s of MeV) and the energy loss of the β-decay electrons (keV to MeV range). Both gain ranges needed to be adjusted to ensure that all front and back strips have the same response. The low gain, used for detecting ion implantations, is gain matched by applying a multiplication factor in software to shift the centroid of the maximum energy peak of implanted ions of each strip so the peak was centered in the same channel for all front or back strips. The low gain matched energy spectrum of both the front strips and the back strips is shown in Figure 2.6. The high gain, used for detecting β-decay electrons, was gain matched using the most intense alpha peak from a 228Th source. Similarly to the low gain, a multiplication 27 factor was applied in software to shift the centroid of the peak in each DSSD strip so that it was centered in the same channel for all strips. The high gain matched energy spectrum of both the front and back strips is shown in Figure 2.7. All scaling factors used for both the high gain and low gain matching are given in Table 2.3. Figure 2.5: A picture of the DSSD used for experiment e14505. 2.2 Isolation of 83As β decay Ions of 83As, 84Se and 85Se were delivered to the experimental end station. These three isotopes are labeled in the particle identification (PID) plot, which is a plot of particle time- of-flight versus energy loss as described in Section 2.1, shown in Figure 2.8. Conventionally, the β decay of the isotope of interest would be identified by correlating to implanted ions, which are identified via the PID. In software, a correlation timing window is defined. Once a decay is observed, the correlation is performed by opening the timing window and looking backwards in time to identify an ion that arrived within the specified timing window. The correlated ion must also have arrived in the same DSSD pixel the β decay is observed in. The correlation timing window is generally chosen based on the half-life of the isotope of interest, and needs to be long enough that the decay will be seen, but not too long that there is a chance for another ion to be implanted in the same time window. In the present 28 Figure 2.6: Gain matched energy spectra for the DSSD low gain setting. The front (top) and back (bottom) strip numbers are shown as a function of channel number. experiment, the implanted isotopes and their decay products have half-lives on the order of seconds to hours (summarized in Table 2.4), which would require long correlation time windows. Along with a high implantation rate, the preferred method of correlation would be too difficult to successfully identify the decay of 83As. Due to this, another method of isolating the β decay of 83As needed to be used. The beam delivered to the experimental end station was pulsed in an eight minutes on, eight minutes off cycle for seven hours, after which the beam was turned off for eight hours. The pulsed setting was chosen based on the half-lives of the species involved, to obtain a 29 Figure 2.7: Gain matched energy spectra for the DSSD hi gain setting, using a 228Th source. The front (top) and back (bottom) strip numbers are shown as a function of channel number. cleaner spectrum associated with the decay of individual isotopes. Shown in Figure 2.9 is the total absorption (or summed) spectra corresponding to the decay of all isotopes and their daughters and granddaughters as a function of time. There are several strongly populated states in the decay of some of the isotopes that appear here as the horizontal bands. Longer lived isotopes can be identified based on the bands extending into the long beam-off period starting around 25000 seconds. The projection of any excitation energy region onto the x-axis would yield the decay profile of that excitation energy which may have a contribution from multiple decaying species. 30 Table 2.3: Scaling factors used for gain matching the front and back strips of the DSSD. Front Strip Number 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Low Gain Matching High Gain Matching Factor 0.990 Factor 0.680 1.075 0.998 1.130 1.027 1.004 1.203 0.973 0.973 0.998 0.995 1.042 0.990 1.036 0.726 1.054 0.643 0.681 0.698 0.685 0.650 0.658 0.687 0.682 0.654 0.669 0.644 0.658 0.710 0.656 0.642 Back Strip Number 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Low Gain Matching High Gain Matching Factor 0.978 Factor 0.639 0.939 0.935 0.953 0.965 1.036 0.974 0.914 0.953 1.005 0.949 0.995 1.010 0.950 0.974 0.939 0.657 0.644 0.670 0.671 0.645 0.680 0.673 0.688 0.657 0.649 0.663 0.678 0.621 0.729 0.674 Using the half-lives and implantation rates of all present species (shown as a function of time for 83As and 84Se in Figure 2.10), their individual contributions to this total decay profile were modeled using the Bateman equations, which describe the growth and decay of activities of a decay chain [72]. The contribution of 85Se to the total decay profile was determined to be on the order of 1%, and is therefore ignored in the subsequent analysis. The general solution of the Bateman equations for a decay chain Q1 → Q2 → Q3 → ... → 31 Figure 2.8: Particle Identification plot. PIN Energy is from the first PIN detector located upstream of SuN. TOF is determined from the timing between the PIN detector and the focal plane scintillator located in the A1900 fragment separator. Table 2.4: Half-lives of implanted isotopes and their decay products. Parent 83As 84Se 85Se t1/2 13.4 s 3.26 m 32.9 s Daughter t1/2 Granddaughter 83Se 84Br 85Br 22.25 m, 70.1 s 31.76 m 2.90 m 83Br 84Kr 85Kr t1/2 2.374 h stable 4.480 h Qn, where there is also an external source of production, Si, is shown in Equation 2.1. n(cid:88) i=1 Qn(t) = n−1(cid:89) j=i λj,j+1 × n(cid:88) j=i  Qi(0)e (cid:81)n p=i p(cid:54)=j −λj t (λp − λj) (cid:81)n Si(1 − e p=i p(cid:54)=j −λj t) (λp − λj) + kj   (2.1) Here, Qn is the number of atoms of a species present at time t, λn is the decay constant for species n, and λn,n+1 is the partial decay constant, which takes into account the branching ratio, as shown in Equation 2.2. λn,n+1 = BRn,n+1 × λn (2.2) 32 Figure 2.9: SuN total absorption spectrum as a function of time. To ensure the activity profiles produced using the Bateman equations were accurate representations of the experimental data, energy cuts were taken for excitation energy peaks in the 2-dimensional spectrum, shown in Figure 2.9, that could be attributed to the decay of a single isotope or a decay chain. Each cut was projected onto the x-axis to obtain the decay profile associated with that isotope or decay chain. This was then compared to the decay profile produced using the Bateman equations. The first activity profile compared was one attributed to the 408.2 keV level in 84Br, which is populated through the β decay of 84Se and decays directly to the ground state. The decay scheme of 84Br is shown in Figure 2.11. The Bateman equations for the mass 84 chain were scaled according to the efficiency of SuN (70% at 400 keV) and β-feeding values (100% for the 408.2-keV energy level). The calculated activity profile was overlaid on the activity profile produced by gating on the 408.2-keV peak 33 Figure 2.10: (Top) Implantation rate of 83As ions in the DSSD. (Bottom) Implantation rate of 84Se ions in the DSSD. in the TAS as a function of time spectrum. The result is shown in the top panel of Fig. 2.12. The second activity profile reproduced was one attributed to two decays within the same decay chain. The peak at approximately 2050 keV in the excitation energy spectrum has contributions from the β decay of 83As populating a 2076.84-keV level in 83Se and the β decay of 83Se populating a 2051.45-keV level in 83Br. Due to the poor energy resolution of SuN, these levels could not be resolved. Both states de-excite via cascades of multiple γ rays to the ground state. The detector efficiency, β-feeding values, and relative γ-ray intensities (obtained from ENSDF [73]) of each cascade were used to scale the calculated decay profiles. 34 The calculated decay profiles were compared to the activity profile obtained by gating on the 2050-keV peak in the TAS as a function of time spectrum. The result is shown in the bottom panel of Fig. 2.12. Figure 2.11: Complete level scheme of the decay of 84Se into 84Br from ENSDF. Absolute γ-ray intensities per 100 decays listed next to γ-ray energies. [73] Once the calculated decay profiles were confirmed to represent the actual decay profiles, all decay profiles contributing to the total activity were calculated via the Bateman equa- tions. All activities over the full time scale are shown in Fig. 2.13. Regions of the pulsed beam setting and the long decay period are shown in Figures 2.14 and 2.15, respectively. The decay profiles of individual isotopes were then used to identify time regions in which selected isotopes could be isolated in order to subtract them from the total activity profile. The subtractions are done with 2D matrices of the excitation energy (Ex) as a function of individual γ-ray energies (Eγ), since this is the starting point for the β-Oslo method (which will be discussed in the next chapter). The total Ex, Eγ matrix is shown in Figure 2.16. From approximately 26000 seconds to the end of the run, the only activities expected to be present are from 83Se → 83Br and 83Br → 83Kr, as shown in Figure 2.15. The 2D Ex, Eγ matrix gated on this time region is shown in Fig. 2.17. Based on the results of the Bateman equations, there should be more 83Br than 83Se decay at later times. However, only 1.4% 35 Figure 2.12: (Top) TAS spectrum as a function of time gated on the 408 keV peak corre- sponding to 84Br (only first three beam cycles shown) with the decay profile modeled via the Bateman equations overlaid. (Bottom) TAS spectrum as a function of time gated on the 2053 keV peak that has contributions from the decay chain of 83As → 83Se → 83Br (only first three beam cycles shown) with the decay profile of each decay, as well as the total, modeled via the Bateman equations overlaid. of the β decays from 83Br into 83Kr populate observable excited states above the threshold of SuN (see level scheme in Fig. 2.18), leading to a 2D Ex, Eγ matrix that has comparable contributions from the decays of 83Br and 83Se. In order to independently isolate these two decays, the decay of 83Br → 83Kr was simulated in Geant4 and is shown in Figure 2.19. The 36 Figure 2.13: Activity profiles of all decaying species modeled with the Bateman equations, shown over the full time range. number of decays simulated was determined by integrating the calculated activity profile from 26000 seconds to the end of the run. The 83Br → 83Kr matrix was subtracted from the matrix shown in Fig. 2.17 to isolate a matrix of only 83Se → 83Br, shown in Fig. 2.20. A decay scheme of 83Se → 83Br with selected levels and γ-ray transitions is shown in Fig. 2.21. Here, the majority of the β feeding goes to a series of levels between 2600 and 2800 keV and eventually decays through cascades that have several lower energy, high intensity γ rays. This can be seen in the the Ex, Eγ matrix as well, providing confirmation that 83Br was isolated. The simple decay of 84Se → 84Br (decay scheme shown in Fig. 2.11), consists of a single excited state fed in β decay that results in a single γ-ray transition to the ground state. The decay of 84Se was 37 Figure 2.14: Regions of production and decay during the pulsed beam setting. also simulated using Geant4. The resulting simulated Ex, Eγ matrix is shown in Fig. 2.22. The two remaining decays, an isomeric state of 83Se (Qβ− = 3673(5) keV) decaying to 83Br and the decay of 84Br (Qβ− = 4660(3) keV) to 84Kr could not be independently isolated nor could they reliably be simulated, due to their larger Qβ− values. After each beam on period, 83As (t1/2 = 13.4 s) decayed away, and the Ex, Eγ spectrum from 102 seconds after the beginning of the beam off period to 19 seconds before the beginning of the next beam on cycle, was obtained. This time period included contributions from the decay of 83Se, 83mSe, 83Br, 84Se, and 84Br, as can be seen in Figure 2.14. The isolated spectra of the decays of 83Se, 83Br, and 84Se were all scaled and removed, leading to a 2D Ex, Eγ matrix with the decays of 83mSe and 84Br in a 9 to 1 ratio. This matrix is shown in Fig. 2.23. The decay schemes of 83mSe → 83Br and 84Br → 84Kr are shown in Figures 2.24 and 2.25. Many of the strongly fed levels and high intensity γ rays can be observed in the 2D matrix. 38 Figure 2.15: Plot of the long decay period during which the beam was not being delivered to the experimental end station. Each of the isolated matrices were then scaled and subtracted from the portion of the beam-on periods with the same ratio of 83mSe and 84Br decays to obtain the final, isolated 83As → 83Se matrix, shown in Fig. 2.26. This 2D matrix is used as the starting point for the β-Oslo method, which will discussed in the following chapter. The spectra of γ-ray energy (Eγ) and excitation energy (Ex) measured in SuN for the β decay of 83As to 83Se (projections from the Ex, Eγ matrix) are shown in Figure 2.27. A decay scheme with select levels and transitions of 83As → 83Se is shown in Fig. 2.28. Both spectra are consistent with literature of known levels and γ-ray transitions of 83Se. In the Eγ spectra, the strongest γ-ray transition with energy 734.9 keV is evident, as well as the second strongest transition with energy 1113.4 keV (Iγ = 36.1(11)% relative to the 734.9 keV transition [74]). While there are no β-feeding intensities reported in literature, a strongly fed level with energy 39 Figure 2.16: Excitation energy (Ex) as a function of γ-ray energy (Eγ), ungated, with all events. Figure 2.17: Ex, Eγ matrix with 50 keV binning, for a time gate of 26000 seconds to the end of the run. 1062.89(7) keV is clear, as well as strongly fed levels in the energy range of approximately 1800 - 2200 keV and 2500 - 3100 keV. 40 (keV)γE01000200030004000500060007000 (keV)xE01000200030004000500060007000110210310 Figure 2.18: Decay scheme of 83Br → 83Kr. Absolute γ-ray intensities per 100 decays listed next to γ-ray energies. [74] Figure 2.19: Ex, Eγ matrix with 50 keV binning, for the decay of 83Br → 83Kr simulated in Geant4. 41 (keV)γE0100200300400500600700800 (keV)xE0100200300400500600700800110210310410 Figure 2.20: Ex, Eγ matrix with 50 keV binning, for the decay of 83Se → 83Br. Figure 2.21: Decay scheme of 83Se → 83Br with selected levels and γ-ray transitions. Abso- lute γ-ray intensities per 100 decays listed next to γ-ray energies. [74] 42 (keV)γE01000200030004000500060007000 (keV)xE01000200030004000500060007000110210310 Figure 2.22: Ex, Eγ matrix with 50 keV binning, for the decay of 84Se → 84Br simulated in Geant4. Figure 2.23: Ex, Eγ matrix with 50 keV binning, for the decays of the isomeric state of 83Se → 83Br and 84Br → 84Kr. 43 (keV)γE0200400600800100012001400 (keV)xE0200400600800100012001400110210310410 (keV)γE01000200030004000500060007000 (keV)xE01000200030004000500060007000110210310410 Figure 2.24: Decay scheme of the isomeric state of 83Se → 83Br with selected levels and γ-ray transitions. Absolute γ-ray intensities per 100 decays listed next to γ-ray energies. [74] 44 Figure 2.25: Decay scheme of 84Br → 84Kr with selected levels and γ-ray transitions. Ab- solute γ-ray intensities per 100 decays listed next to γ-ray energies. [73] Figure 2.26: Ex, Eγ matrix with 50 keV binning, for the decay of 83As → 83Se. 45 (keV)γE01000200030004000500060007000 (keV)xE01000200030004000500060007000110210310410 Figure 2.27: (Top) γ-ray energy spectrum of 83Se. (Bottom) Excitation energy spectrum of 83Se. 46 Figure 2.28: Decay scheme of 83As → 83Se with selected levels and γ-ray transitions. γ-ray intensities relative to the 734.9 keV transition listed next to γ-ray energies. [74] 47 Chapter 3 β-Oslo Method The β-Oslo method [63] is based on the established Oslo method [59–62] for extracting the nuclear level density (NLD) and γ-ray strength function (γSF) of a nucleus. These sta- tistical properties are extracted from data obtained by populating highly excited states in the nucleus of interest and observing the subsequent emission of γ rays as the nucleus de- excites. While the Oslo method uses charged particle reactions to accomplish the population of highly-excited states in the nucleus, the β-Oslo method utilizes β decay of neutron-rich nu- clei. The use of β decay gives the β-Oslo method the advantage to experimentally constrain the NLD and γSF of nuclei further from stability than the Oslo method or other reaction- based techniques. However, this method has limitations based on β-decay selection rules and is restricted to nuclei that have large β-decay Q values and a high NLD at the neutron separation energy (Sn), to ensure that the statistical region of the nucleus is being popu- lated. The experimental information extracted from the application of the β-Oslo method is then used to constrain a neutron-capture cross section using the Hauser-Feshbach model for neutron capture as is explained in the Introduction. The β-Oslo method is comprised of four main steps: 1. Unfolding the spectra of γ rays observed, for each excitation energy [59]. 2. Isolation of the primary γ rays. These are the first γ rays emitted from each excited state in cascades to the ground state [60]. 48 3. Extraction of the functional forms of the NLD and the transmission coefficient (T (Eγ)) (See Section 3.4 for conversion of T (Eγ) to γSF) [61]. 4. Normalization of the NLD and γSF [61, 62]. Each of these four steps will be covered in more detail in the following sections, along with the results of the β-Oslo method being utilized to constrain the 82Se(n,γ)83Se cross section. The level density and γ-ray strength function of 83Se are determined from the β decay of 83As → 83Se (see Chapter 2 for experimental details). A flow chart of the various steps, programs, and inputs that are used for a β-Oslo analysis is shown in Figure 3.1. The MAtrix MAnipulation (MAMA) program [76] is used for the first two steps of the analysis: the unfolding of the spectra, and the isolation of primary γ rays. A suite of other programs (shown in bold in Figure 3.1) are used for determination of several input parameters, and for the remaining steps of the Oslo analysis. 3.1 Population of highly excited states in 83Se The starting point of the β-Oslo method is a 2D matrix of the excitation levels and γ rays emitted in the nucleus of interest (the product of the neutron-capture reaction being constrained). For the case of constraining the cross section of 82Se(n,γ)83Se, the highly excited states of 83Se were populated via the β decay of 83As. This reaction is shown schematically, along with the corresponding neutron-capture reaction, in Figure 3.2. To ensure that the statistical region of the nucleus is being populated, a Qβ− value above 4 MeV is preferred [77]; 83As has a Qβ− value of 5.671 MeV. It is also preferred that the Qβ− value be close to the neutron separation energy (83Se Sn = 5.818 MeV), to cover the majority of the energy range of γ rays emitted after neutron capture. 49 Figure 3.1: Flow chart of Oslo method. Bold items are programs used for the respective steps in the flow chart. Figure from [75]. 50 Root HistogramMAMA MatrixUnfold Raw MatrixGenerate ResponseFunctionMAMAExtract PrimaryTransitionsFill and RemoveNegative CountsFill and RemoveNegative CountsExtract Level Density andGamma TransmissionCoefficient RhosigchiExperimental lower    threshold Spin Cutoff Parameter Fermi Gas model parameters (a and E1) Constant Temperature model parameter (T) Robin Normalize Level Density(and obtain slope fortransmission coefficient) Counting Known low-energy levelsMass number and protonnumber Average s-wave levelspacing at neutronseparation energy D0Convert TransmissionCoefficient to GammaStrength Function andNormalize Normalization Neutron separation energy Neutron separation energyAverage s-wave levelspacing at neutronseparation energy D0Average total radiativeresonance width  ΓγSpin of target nucleusStarting value for level density parameter a γ     lower limitEγEx      upper and lowerlimitsLevel Density at Neutron Separation Energy D2rho ρ()Sn Figure 3.2: Schematic of the population of states in 83Se through the β decay of 83As and neutron capture on 82Se. 3.2 Unfolding As previously discussed in Chapter 2, the spectra of γ-ray energies deposited in SuN is obtained via the summation of the spectra of each individual segment. While SuN has a high efficiency for detecting the total energy of a single γ ray within the full detector volume (above 80% for a 1 MeV γ ray), the efficiency of an individual segment detecting the full energy of a γ ray is much lower, around 40%. This is mainly due to the various interactions through which the energy of γ rays is deposited in the segments of SuN. While the full energy of a γ ray interacting through photoelectric absorption will be deposited in one interaction, there is a chance that some γ rays will lose energy through pair production or be Compton scattered to another segment, in which case only part of the full energy would be deposited in a given segment. These types of γ ray interactions lead to incomplete energy sums in the γ-ray energy spectra, which is what the method of unfolding addresses. 51 In order to determine the full γ-ray energies expected from the raw data, an iterative procedure was used to unfold the response function of the detector [59]. This response function, which represents the response of the detector to a range of γ-ray energies, was simulated using Geant4 [78]. The goal of this method is to determine a “unfolded” γ-ray spectrum that, when combined with the response function, will match the experimental data. The “folded” spectrum f can be represented by: f = Ru (3.1) where f is the folded spectrum, R is the response function matrix, and u is the unfolded spectrum. The iteration method is applied as follows: 1. An initial trial unfolded function is defined as: u0 = r (3.2) where r is the observed experimental γ-ray spectra. 2. The first folded spectrum, f 0, is calculated using the response function matrix R and the first trial unfolded function u0: f 0 = Ru0 (3.3) 3. The next trial unfolded function, u1, is determined by applying the difference of the folded spectrum, f 0, and the observed spectrum r to the original trial unfolded function 52 u0 as a correction factor: u1 = u0 + (r − f 0) (3.4) 4. The new trial unfolded function is folded to determine the next folded spectrum, f 1, which is then used to calculate the next trial unfolded function: f 1 = Ru1 u2 = u1 + (r − f 1) (3.5) (3.6) Step 4 is repeated until the folded spectrum matches the experimental spectrum within uncertainties (f i ≈ r). Typically, around 30 iterations are performed. The unfolding procedure used in the traditional Oslo and β-Oslo methods previously focused only on unfolding the Eγ axis of the 2D matrix shown in Figure 2.26. A recent development to the β-Oslo method is to also unfold the Ex axis [79]. Since the determination of the excitation energy is directly linked to the measurement of γ-ray energies, the excitation energy is also effected by incomplete summing of γ-ray cascades. There is also the chance of additional background from β-decay electrons also interacting with SuN. The response function is dependent on the initial excitation energy (Ex), the γ-ray multiplicity (Mγ) at the excitation energy, and the Qβ value of the decay being measured. Response functions covering the possible combinations of Ex, Mγ, and Qβ values have been simulated in Geant4. The rest of the unfolding proceeds in the same method described above. The unfolded 2D matrix for 83Se obtained by applying this method to both the Eγ and Ex axis of the original raw 2D matrix for 83Se is shown in Figure 3.3. 53 Figure 3.3: Unfolded Ex, Eγ matrix for 83Se, with 50 keV binning. 3.3 Extraction of primary γ rays To extract the functional forms of the nuclear level density and γ-ray strength function, the Ex, Eγ matrix needs to contain only the primary, or first, γ rays emitted in cascades from each excited state. The first generation method was developed to extract these primary γ rays for each excitation energy bin [60]. The first generation method is based on the assumption that the γ-ray emission from any excited state is independent of how that excited state was populated. For each excitation energy bin i (here, 200 keV wide) of the unfolded matrix, the γ-ray spectrum, fi, contains γ rays from all cascades that are possible from excited states within that excitation energy bin. Therefore, γ-ray spectra fj = [2,5] and 5.0 MeV > ESum > 5.9 MeV, efficiencies were calculated, and are shown in Table 4.3. A global average was taken over all simulations to obtain an average efficiency of 0.422 ± 0.038. A global average is used to represent all contributing neutron capture resonances. 93 Figure 4.26: Energy of individual γ rays, gated on multiplicity 2, for several strong resonances in the 82Se data compared to simulated data for J π = 1 the average across all realizations, while error bars correspond to the standard deviation of the bin content for each realization. The integral of each plot is normalized to unity for a true comparison. 2−. Points correspond to 2+, 1 2−, 3 94 Figure 4.27: Energy of individual γ rays, gated on multiplicity 3, for several strong resonances in the 82Se data compared to simulated data for J π = 1 the average across all realizations, while error bars correspond to the standard deviation of the bin content for each realization. The integral of each plot is normalized to unity for a true comparison. 2−. Points correspond to 2+, 1 2−, 3 95 Figure 4.28: Energy of individual γ rays, gated on multiplicity 4, for several strong resonances in the 82Se data compared to simulated data for J π = 1 the average across all realizations, while error bars correspond to the standard deviation of the bin content for each realization. The integral of each plot is normalized to unity for a true comparison. 2−. Points correspond to 2+, 1 2−, 3 96 Figure 4.29: Energy of individual γ rays, gated on multiplicity 5, for several strong resonances in the 82Se data compared to simulated data for J π = 1 the average across all realizations, while error bars correspond to the standard deviation of the bin content for each realization. The integral of each plot is normalized to unity for a true comparison. 2−. Points correspond to 2+, 1 2−, 3 97 Figure 4.30: Simulation efficiencies and their uncertainties as a function of the lower ESum gate boundary, for M = [2,5] and an upper ESum gate boundary of 5.9 MeV. 4.5.3 Cross Section Results With the yield of neutron capture on 82Se determined and the efficiency of γ rays detected within the ESum and multiplicity cuts calculated, the cross section of 82Se(n,γ)83Se was determined using Equation 4.6. The results are shown in Figure 4.31. The statistical un- certainty shown in Figure 4.31 includes the uncertainty of the measured yield of neutron capture propagated through all background subtractions, as well as the uncertainty of the beam-monitor yields propagated through the determination of the neutron fluence. In ad- dition to the statistical uncertainty shown, there is a systematic uncertainty of 4.25%. The systemic uncertainty includes the 4% uncertainty from the neutron-fluence normalization and the 0.47% uncertainty from the efficiency, added in quadrature. The measured cross section was converted to a Maxwellian-Averaged Cross Section (MACS) using Equation 4.12, where µ is the reduced mass, σ(En) is the cross section at 98 neutron energy En, and δ(En) is the bin width of the bin centered on En. At a temperature (kT ) of 30 keV the MACS was determined to be 5.67 ± 4.72 mb. The experimentally deter- mined MACS is in agreement with the recommended theoretical value of 9 ± 8 mb at 30 keV from the Karlsruhe Astrophysical Database of Nucleosynthesis in Stars (KADoNiS) [98]. σM ACS(kT ) = 2√ π σ(En)Ene En kT δ(En) (4.12) (cid:16) µ (cid:17)2 1M eV(cid:88) kT En=10eV Figure 4.31: Cross section of the 82Se(n,γ)83Se reaction measured with DANCE. 99 Table 4.3: Efficiencies determined by DICEBOX and GEANT4 simulations (M = [2,5], 5.0 MeV > ESum > 5.9 MeV) for J π = 1 30 realizations of artificial nuclei. 2−. 100000 γ-ray cascades were simulated for 2+, 1 2−, 3 Jπ = 1 2− Jπ = 3 2− 0.392 0.389 0.394 0.390 0.391 0.390 0.390 0.389 0.392 0.391 0.390 0.389 0.386 0.390 0.387 0.392 0.389 0.391 0.388 0.395 0.395 0.392 0.391 0.386 0.392 0.393 0.394 0.392 0.394 0.390 Realization 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 2+ Jπ = 1 0.409 0.468 0.413 0.408 0.406 0.406 0.405 0.406 0.408 0.409 0.407 0.410 0.407 0.407 0.408 0.404 0.407 0.405 0.404 0.406 0.409 0.411 0.411 0.411 0.413 0.412 0.406 0.411 0.413 0.408 0.405 0.469 0.466 0.470 0.466 0.467 0.466 0.466 0.468 0.466 0.467 0.466 0.469 0.470 0.467 0.468 0.469 0.465 0.467 0.472 0.467 0.467 0.471 0.465 0.468 0.465 0.467 0.466 0.469 0.466 100 Chapter 5 Results and Conclusion 5.1 Comparison of β-Oslo Cross Section to Directly Measured Cross Section The cross section of the 82Se(n,γ)83Se reaction was determined via a direct method and an indirect method. As discussed in Chapter 3, the β-Oslo method was utilized to determine the 82Se(n,γ)83Se cross section through a study of 83As β decay. As discussed in Chapter 4, a direct measurement of neutron capture on 82Se has been measured using the Detector for Advanced Neutron Capture Experiments, DANCE. These two cross section determinations were compared to validate the β-Oslo method. The results are shown in Figure 5.1. The β- Oslo cross section is systematically higher compared to the directly measured cross section. The discrepancy is likely due to the β-Oslo cross section missing the resonance behavior, which is not unsurprising, since TALYS calculates an average cross section. There is better agreement at higher energies (above approximately 40 keV). 5.2 Conclusions and Outlook The r-process is important to understanding the production of heavy elements, but there is still a lack of knowledge about the r-process, including the astrophysical sites where it takes 101 Figure 5.1: Comparison of the directly measured cross section of 82Se(n,γ)83Se (black) to the neutron capture cross section determined via the β-Oslo method (black line). The blue lines indicate the upper and lower limits of the cross section determined via the β-Oslo method through a systematic study of the uncertainty. place. Models of the r-process can be used to produce abundance patterns with the aim of pinpointing the astrophysical site, but such models require information on nuclear properties of the nuclei involved, including neutron-capture cross sections. The cross sections of the neutron-capture reactions that drive the r-process are a critical piece of information, but presently there is a lack of directly measured neutron-capture cross sections in the neutron- rich region where the r-process takes place. Results from theoretical models are used for cases where experimental data does not exist. The lack of knowledge on the accuracy of these theoretical calculations leads to an inability to reproduce r-process abundance patterns to the degree of certainty needed for successful comparisons to known solar abundance patterns. The β-Oslo method has been developed to constrain neutron-capture cross sections for β- unstable nuclei not amenable to direct measurement, using experimental data from β-decay 102 studies. The nuclear level density and γ-ray strength function can be extracted from the experimental β-decay data, and used to perform a Hauser-Feschbach calculation of the cross section. A previous study showed that the cross section of 50Ti(n, γ)51Ti, obtained via the β-Oslo method, had excellent agreement with a directly measured cross section of 50Ti(n, γ)51Ti. Since 51Ti is a relatively low mass nucleus with a low level density, a comparison of a higher mass nucleus (and therefore a higher level density) is necessary to validate that the efficacy of the β-Oslo method and the assumptions made in the analysis will hold for higher level densities. The present work has determined a cross section for 82Se(n, γ)83Se from both a direct measurement of neutron capture on 82Se, and the application of the β-Oslo method on the β decay of 83As to 83Se. The two cross sections are in reasonable agreement. It is recommended to investigate the slight overestimation of the indirectly determined cross section in the future to determine whether the overestimation is a systematic deviation or an artifact of the analysis performed. It is also recommended to perform a validation in a heavier mass region for a nucleus with a higher level density than. Both 50Ti and 82Se are located near shell closures, so it would be advantageous to perform a validation with a nucleus that somewhere between shell closures. Overall, the β-Oslo method has been shown to be a viable method for constraining neutron-capture cross sections. 103 BIBLIOGRAPHY 104 BIBLIOGRAPHY [1] Christopher Sneden, John J Cowan, and Roberto Gallino. Neutron-Capture Elements in the Early Galaxy. [2] Walter Loveland, David J. Morriessey, and Glenn T. Seaborg. Modern Nuclear Chem- istry. John Wiley & Sons, Inc., 2006. [3] M. Arnould, S. Goriely, and K. Takahashi. The r-process of stellar nucleosynthesis: Astrophysics and nuclear physics achievements and mysteries. Physics Reports, 450:97– 213, 2007. [4] LIGO Scientific Collaboration and Virgo Collaboration. GW170817: Observation of Gravitational Waves from a Binary Neutron Star Inspiral. Phys. Rev. Lett., 119(16):30– 33, 2017. [5] B. D. Metzger, G. Mart´ınez-Pinedo, S. Darbha, E. Quataert, A. Arcones, D. Kasen, R. Thomas, P. Nugent, I. V. Panov, and N. T. Zinner. Electromagnetic counterparts of compact object mergers powered by the radioactive decay of r-process nuclei. Monthly Notices of the Royal Astronomical Society, 406(4):2650–2662, 8 2010. [6] Daniel Kasen, Brian Metzger, Jennifer Barnes, Eliot Quataert, and Enrico Ramirez- Ruiz. Origin of the heavy elements in binary neutron-star mergers from a gravitational- wave event. Nature, 551:80–84, 2017. [7] M R Drout, A L Piro, B J Shappee, C D Kilpatrick, J D Simon, C Contreras, D A Coulter, R J Foley, M R Siebert, N Morrell, K Boutsia, F Di Mille, S Holoien, D Kasen, J A Kollmeier, B F Madore, A J Monson, A Murguia-Berthier, Y.-C Pan, J X Prochaska, E Ramirez-Ruiz, A Rest, C Adams, K Alatalo, E Ba˜nados, J Baughman, T C Beers, R A Bernstein, T Bitsakis, A Campillay, T T Hansen, C R Higgs, A P Ji, G Maravelias, J L Marshall, C Moni Bidin, J L Prieto, K C Rasmussen, C Rojas-Bravo, A L Strom, N Ulloa, J Vargas-Gonz´alez, Z Wan, and D D Whitten. Light curves of the neutron star merger GW170817/SSS17a: Implications for r-process nucleosynthesis. Science, 358:1570–1574, 2017. [8] M. R. Mumpower, R. Surman, G. C. McLaughlin, and A. Aprahamian. The impact of individual nuclear properties on r-process nucleosynthesis. Progress in Particle and Nuclear Physics, 86:86–126, 2016. [9] S Goriely, N Chamel, and J M Pearson. Skyrme-Hartree-Fock-Bogoliubov Nuclear Mass Formulas: Crossing the 0.6 MeV Accuracy Threshold with Microscopically Deduced Pairing. Phys. Rev. Lett., 102:152503, 2009. 105 [10] J Duflo and A. P. Zuker. Microscopic mass formulas. Phys. Rev. C, 52(1), 1995. [11] P. Moller, J. R. Nix, W. D. Myers, and W. J. Swiatecki. Nuclear Ground-State Masses and Deformations. Atomic Data and Nuclear Data Tables, 59:185–381, 1995. [12] Arjan Koning, Stephane Hilaire, and Stephane Goriely. TALYS User Manual (v. 1.95). Technical report. [13] Thomas Rauscher and Friedrich-Karl Thielemann. Astrophysical Reaction Rates From Statistical Model Calculations. Atomic Data and Nuclear Data Tables, 75(1-2):1–351, 2000. [14] M. Beard, E. Uberseder, R. Crowter, and M. Wiescher. Comparison of statistical model calculations for stable isotope neutron capture. Phys. Rev. C, 90:034619–undefined, 2014. [15] G. Savard, R. C. Pardo, S. Baker, C. N. Davids, A. Levand, D. Peterson, D. G. Phillips, T. Sun, R. Vondrasek, B. J. Zabransky, and G. P. Zinkann. CARIBU: A new facility for the study of neutron-rich isotopes. Hyperfine Interactions, 199(1):301–309, 7 2011. [16] J Van Schelt, D Lascar, G Savard, J A Clark, P F Bertone, S Caldwell, A Chaudhuri, A F Levand, G Li, G E Morgan, R Orford, R E Segel, K S Sharma, and M G Sternberg. First Results from the CARIBU Facility: Mass Measurements on the r-Process Path. Phys. Rev. Lett., 111:–undefined, 2013. [17] J Hakala, J Dobaczewski, D Gorelov, T Eronen, A Jokinen, A Kankainen, V S Kolhinen, M Kortelainen, I D Moore, H Penttil¨a, S Rinta-Antila, J Rissanen, A Saastamoinen, V Sonnenschein, and J A Yst¨o. Precision Mass Measurements beyond 132 Sn: Anoma- lous Behavior of Odd-Even Staggering of Binding Energies. [18] M Matoˇs, A Estrade, M Amthor, A Aprahamian, D Bazin, A Becerril, T Elliot, D Galaviz, A Gade, S Gupta, G Lorusso, F Montes, J Pereira, M Portillo, A M Rogers, H Schatz, D Shapira, E Smith, A Stolz, and M Wallace. TOF-Bρ mass measurements of very exotic nuclides for astrophysical calculations at the NSCL. Journal of Physics G: Nuclear and Particle Physics, 35:6, 2008. [19] G Lorusso, S Nishimura, Z Y Xu, A Jungclaus, Y Shimizu, G S Simpson, P.-A S¨oderstr¨om, H Watanabe, F Browne, P Doornenbal, G Gey, H S Jung, B Meyer, T Sumikama, J Taprogge, Zs Vajta, J Wu, H Baba, G Benzoni, K Y Chae, F C L Crespi, N Fukuda, R Gernh¨auser, N Inabe, T Isobe, T Kajino, D Kameda, G D Kim, Y.-K Kim, I Kojouharov, F G Kondev, T Kubo, N Kurz, Y K Kwon, G J Lane, Z Li, A Montaner-Piz´a, K Moschner, F Naqvi, M Niikura, H Nishibata, A Odahara, R Or- landi, Z Patel, Zs Podoly´ak, H Sakurai, H Schaffner, P Schury, S Shibagaki, K Steiger, H Suzuki, H Takeda, A Wendt, A Yagi, and K Yoshinaga. β-Decay Half-Lives of 110 106 Neutron-Rich Nuclei across the N 82 Shell Gap: Implications for the Mechanism and Universality of the Astrophysical r Process. Phys. Rev. Lett., 114, 2015. [20] M Quinn, A Aprahamian, J Pereira, R Surman, O Arndt, T Baumann, A Becerril, T Elliot, A Estrade, D Galaviz, T Ginter, M Hausmann, S Hennrich, R Kessler, K.-L Kratz, G Lorusso, P F Mantica, M Matos, F Montes, B Pfeiffer, M Portillo, H Schatz, F Schertz, L Schnorrenberger, E Smith, A Stolz, W B Walters, and A W¨ohr. β decay of nuclei around 90Se: Search for signatures of a N = 56 subshell closure relevant to the r process. Phys. Rev. C, 85:35807, 2012. [21] A I Morales, J Benlliure, T Kurtuki´an-Nieto, K.-H Schmidt, S Verma, P H Regan, Z Podoly´ak, M G´orska, S Pietri, R Kumar, E Casarejos, N Al-Dahan, A Algora, N Alkhomashi, H ´Alvarez-Pol, G Benzoni, A Blazhev, P Boutachkov, A M Bruce, L S C´aceres, I J Cullen, A M Denis Bacelar, P Doornenbal, M E Est´evez-Aguado, G Farrelly, Y Fujita, A B Garnsworthy, W Gelletly, J Gerl, J Grebosz, R Hoischen, I Kojouharov, N Kurz, S Lalkovski, Z Liu, C Mihai, F Molina, D M¨ucher, B Rubio, H Shaffner, S J Steer, A Tamii, S Tashenov, J J Valiente-Dob´on, P M Walker, H J Wollersheim, and P J Woods. Half-Life Systematics across the N 126 Shell Closure: Role of First-Forbidden Transitions in the β Decay of Heavy Neutron-Rich Nuclei. Phys. Rev. Lett., 113, 2014. [22] T. Kurtukian-Nieto, J. Benlliure, K. H. Schmidt, L. Audouin, F. Becker, B. Blank, I. N. Borzov, E. Casarejos, F. Farget, M. Fern´andez-Ord´o˜nez, J. Giovinazzo, D. Henzlova, B. Jurado, K. Langanke, G. Mart´ınez-Pinedo, J. Pereira, and O. Yordanov. Beta-decay half-lives of new neutron-rich isotopes of Re, Os and Ir approaching the r-process path near N = 126. Eur. Phys. J. A, 50(9):1–8, 2014. [23] A. C. Larsen, A. Spyrou, S. N. Liddick, and M. Guttormsen. Novel techniques for constraining neutron-capture rates relevant for r-process heavy-element nucleosynthesis. Progress in Particle and Nuclear Physics, 2019. [24] Walter Hauser and Herman Feshbach. The inelastic scattering of neutrons. Physical Review, 87(2):366–373, 1952. [25] N Bohr, R Peierls, and G Placzek. Nuclear Reactions in the Continuous Energy Region. Nature, 144:200–201, 1939. [26] A J Koning and J P Delaroche. Local and global nucleon optical models from 1 keV to 200 MeV. Nucl. Phys. A, 713:231–310, 2003. [27] J.-P Jeukenne, A Lejeune, and C Mahaux˜. Microscopic calculation of the symme- try and Coulomb components of the complex optical-model potential. Phys. Rev. C, 15(1):10–29, 1977. [28] Torleif Ericson. The statistical model and nuclear level densities. Advances in Physics, 9(36):425–511, 1960. 107 [29] T Von Egidy and D Bucurescu. Experimental energy-dependent nuclear spin distribu- tions. Phys. Rev. C, 80:54310, 2009. [30] H. A. Bethe. Nuclear Physics B. Nuclear Dynamics, Theoretical. Rev. Mod. Phys., 9:69, 1937. [31] W Dilg, W Schantl, H Vonach, and M Uhl. Level density parameters for the back-shifted fermi gas model in the mass range 40 ¡ A ¡ 250. Nuclear Physics A, 217:269–298, 1973. [32] A. G. W. Cameron. Nuclear Level Spacings. Can. J. Phys., 36:1040–1057, 1958. [33] T. D. Newton. Shell Effects on the Spacing of Nuclear Levels. Can. J. Phys, 34:804–829, 1956. [34] A Gilbert and A G W Cameron. A composite nuclear level density formula with shell corrections. Can. J. Phys., 43:1446, 1965. [35] A. Gilbert, F. S. Chen, and A. G. W. Cameron. Level Densities in Lighter Nuclei. Can. J. Phys, 43:1248–1258, 1965. [36] S. Goriely, F. Tondeur, and J. M. Pearson. a HartreeFock Nuclear Mass Table. Atomic Data and Nuclear Data Tables, 77(2):311–381, 2001. [37] S Goriely, S Hilaire, and A J Koning. Improved microscopic nuclear level densities within the Hartree-Fock-Bogoliubov plus combinatorial method. Phys. Rev. C, 78:64307, 2008. [38] S Hilaire, M Girod, S Goriely, and A J Koning. Temperature-dependent combinatorial level densities with the D1M Gogny force. Phys. Rev. C, 86:064317, 2012. [39] M Krticka and F Bev´a. What do we really know about photon strength functions? Re- lated content Scissors mode of Gd nuclei measured, with the DANCE detector. Journal of Physics G, 35:014025, 2007. [40] D M Brink. Individual Particle and Collective Aspects of the Nuclear Photoeffect. Nuclear Physics, 4:215–220, 1957. [41] Peter Axel. Electric Dipole Ground-State Transition Width Strength Function and 7-Mev Photon Interactions. Phys. Rev., 126(2), 1962. [42] J Kopecky and M Uhl. Test of gamma-ray strength functions in nuclear reaction model calculations. Phys. Rev. C, 41(5), 1990. [43] A Voinov, E Algin, U Agvaanluvsan, T Belgya, R Chankova, M Guttormsen, G E Mitchell, J Rekstad, A Schiller, and S Siem. Large Enhancement of Radiative Strength for Soft Transitions in the Quasicontinuum. Phys. Rev. Lett, 93(14), 2004. 108 [44] M. Guttormsen, R. Chankova, U. Agvaanluvsan, E. Algin, L. A. Bernstein, F. Inge- bretsen, T. L¨onnroth, S. Messelt, G. E. Mitchell, J. Rekstad, A. Schiller, S. Siem, A. C. Sunde, A. Voinov, and S. Ødeg˚ard. Radiative strength functions in 93-98Mo. Phys. Rev. C, 71(4):044307–undefined, 2005. [45] M Wiedeking, L A Bernstein, M Krtiˇcka, D L Bleuel, J M Allmond, M S Basunia, J T Burke, P Fallon, R B Firestone, B L Goldblum, R Hatarik, P T Lake, I-Y Lee, S R Lesher, S Paschalis, M Petri, L Phair, and N D Scielzo. Low-Energy Enhancement in the Photon Strength of 95Mo. Phys. Rev. Lett., 108:162503, 2012. [46] A. C. Larsen, M. Guttormsen, R. Chankova, F. Ingebretsen, T. L¨onnroth, S. Messelt, J. Rekstad, A. Schiller, S. Siem, N. U. H. Syed, and A. Voinov. Nuclear level densities and γ-ray strength functions in Sc44,45. Physical Review C, 76(4):044303, 2007. [47] A. B¨urger, A. C. Larsen, S. Hilaire, M. Guttormsen, S. Harissopulos, M. Kmiecik, T. Konstantinopoulos, M. Krtiˇcka, A. Lagoyannis, T. L¨onnroth, K. Mazurek, M. Norrby, H. T. Nyhus, G. Perdikakis, S. Siem, A. Spyrou, and N. U.H. Syed. Nuclear level density and γ-ray strength function of 43Sc. Physical Review C - Nuclear Physics, 85(6):1–5, 2012. [48] B. V. Kheswa, M. Wiedeking, F. Giacoppo, S. Goriely, M. Guttormsen, A. C. Larsen, F. L. Bello Garrote, T. K. Eriksen, A. G¨orgen, T. W. Hagen, P. E. Koehler, M. Klinte- fjord, H. T. Nyhus, P. Papka, T. Renstrøm, S. Rose, E. Sahin, S. Siem, and T. Tornyi. Galactic production of 138La: Impact of 138,139La statistical properties. Phys. Lett. B, 744:268–272, 5 2015. [49] A Simon, M Guttormsen, A C Larsen, C W Beausang, P Humby, J T Burke, R J Casperson, R O Hughes, T J Ross, J M Allmond, R Chyzh, M Dag, J Koglin, E Mc- cleskey, M Mccleskey, S Ota, and A Saastamoinen. First observation of low-energy γ-ray enhancement in the rare-earth region. Phys. Rev. C, 93(3):034303–undefined, 2016. [50] A. C. Larsen and S. Goriely. Impact of a low-energy enhancement in the γ-ray strength function on the neutron-capture cross section. Physical Review C - Nuclear Physics, 82(1):1–7, 2010. [51] A C Larsen, N Blasi, A Bracco, F Camera, T K Eriksen, A G¨orgen, M Guttormsen, T W Hagen, S Leoni, B Million, H T Nyhus, T Renstrøm, S J Rose, I E Ruud, S Siem, T Tornyi, G M Tveten, A V Voinov, and M Wiedeking. Evidence for the Dipole Nature of the Low-Energy Enhancement in 56 Fe. Phys. Rev. Lett., 111:242504, 2013. [52] M D Jones, A O Macchiavelli, M Wiedeking, L A Bernstein, H L Crawford, C M Camp- bell, R M Clark, M Cromaz, P Fallon, I Y Lee, M Salathe, A Wiens, A D Ayangeakaa, D L Bleuel, S Bottoni, M P Carpenter, H M Davids, J Elson, A G¨orgen, M Guttormsen, R V F Janssens, J E Kinnison, L Kirsch, A C Larsen, T Lauritsen, W Reviol, D G 109 Sarantites, S Siem, A V Voinov, and S Zhu. Examination of the low-energy enhancement of the γ-ray strength function of 56Fe. Phys. Rev. C, 97:24327, 2018. [53] B. Alex Brown and A.C. Larsen. Large Low-Energy M1Strength for Fe-56,57 within the Nuclear Shell Model. Physical Review Letters, 113(25), 12 2014. [54] S Goriely and E Khan. Large-scale QRPA calculation of E1-strength and its impact on the neutron capture cross section. Nuclear Physics A, 706:217–232, 2002. [55] S. Goriely, E. Khan, and M. Samyn. Microscopic HFB + QRPA predictions of dipole strength for astrophysics applications. Nuclear Physics A, 739:331–352, 7 2004. [56] J E Escher, J T Burke, R O Hughes, N D Scielzo, R J Casperson, S Ota, H I Park, A Saastamoinen, and T J Ross. Constraining neutron capture cross sections for unstable nuclei with surrogate reaction data and theory. Phys. Rev. Lett., 121:052501, 2018. [57] H. Utsunomiya, S. Goriely, H. Akimune, H. Harada, F. Kitatani, S. Goko, H. Toyokawa, K. Yamada, T. Kondo, O. Itoh, M. Kamata, T. Yamagata, Y.-W. Lui, I. Daoutidis, D. P. Arteaga, S. Hilaire, and A. J. Koning. γ-ray strength function method and its application to 107Pd. Phys. Rev. C, 82:064610–undefined, 2010. [58] H Utsunomiya, S Goriely, T Kondo, C Iwamoto, H Akimune, T Yamagata, H Toyokawa, H Harada, F Kitatani, Y.-W Lui, A C Larsen, M Guttormsen, P E Koehler, S Hilaire, S P´eru, M Martini, and A J Koning. Photoneutron cross sections for Mo isotopes: A step toward a unified understanding of (γ,n) and (n,γ ) reactions. Phys. Rev. C, 88:15805, 2013. [59] M Guttormsen, T S Tveter, L Bergholt, F Ingebretsen, and J Rekstad. The unfolding of continuum y-ray spectra. Nuclear Instruments and Methods in Physics Research A, 374:371–376, 1996. [60] M Guttormsen, T Ramsoy, and J Rekstad. The First Generation γ-rays From Hot Nuclei. Nuclear Instruments and Methods in Physics Research, 255:518–523, 1987. [61] A. Schiller, L. Bergholt, M. Guttormsen, E. Melby, J. Rekstad, and S. Siem. Extraction of level density and γ strength function from primary γ spectra. Nuclear Instruments and Methods in Physics Research, Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 447(3):498–511, 2000. [62] A. C. Larsen, M. Guttormsen, M. Krtiˇcka, E. Bˇet´ak, A. B¨urger, A. G¨orgen, H. T. Nyhus, J. Rekstad, A. Schiller, S. Siem, H. K. Toft, G. M. Tveten, A. V. Voinov, and K. Wikan. Analysis of possible systematic errors in the Oslo method. Physical Review C - Nuclear Physics, 83(3):1–23, 2011. 110 [63] A. Spyrou, S. N. Liddick, A. C. Larsen, M. Guttormsen, K. Cooper, A. C. Dombos, D. J. Morrissey, F. Naqvi, G. Perdikakis, S. J. Quinn, T. Renstrøm, J. A. Rodriguez, A. Simon, C. S. Sumithrarachchi, and R. G. T. Zegers. Novel technique for Constraining r-Process ( n,γ) Reaction Rates. Physical Review Letters, 113(23):232502, 2014. [64] A Ratkiewicz, J A Cizewski, J E Escher, G Potel, J T Burke, R J Casperson, M Mc- cleskey, R A E Austin, S Burcher, R O Hughes, B Manning, S D Pain, W A Peters, S Rice, T J Ross, N D Scielzo, C Shand, and K Smith. Towards Neutron Capture on Exotic Nuclei: Demonstrating (d,pγ) as a Surrogate Reaction for (n,γ). Phys. Rev. Lett., 122:052502, 2019. [65] T A Laplace, F Zeiser, M Guttormsen, A C Larsen, D L Bleuel, L A Bernstein, B L Goldblum, S Siem, F L Bello Garotte, J A Brown, L Crespo Campo, T K Eriksen, F Giacoppo, A G¨orgen, K Hadynska, Hadynska-Klek Klek, R A Henderson, M Klinte- fjord, M Lebois, T Renstrøm, S J Rose, E Sahin, T G Tornyi, G M Tveten, A Voinov, M Wiedeking, J N Wilson, and W Younes. Statistical properties of 243 Pu, and 242 Pu(n,γ ) cross section calculation. Phys. Rev. C, 93:014323–undefined, 2016. [66] T. G. Tornyi, M. Guttormsen, T. K. Eriksen, A. G¨orgen, F. Giacoppo, T. W. Hagen, A. Krasznahorkay, A. C. Larsen, T. Renstrøm, S. J. Rose, S. Siem, and G. M. Tveten. Level density and γ-ray strength function in the odd-odd Np238 nucleus. Phys. Rev. C, 89:044323–undefined, 2014. [67] B V Kheswa, M Wiedeking, J A Brown, A C Larsen, S Goriely, M Guttormsen, F L Bello Garrote, L A Bernstein, D L Bleuel, T K Eriksen, F Giacoppo, A G¨orgen, B L Goldblum, T W Hagen, P E Koehler, M Klintefjord, K L Malatji, J E Midtbø, H T Nyhus, P Papka, T Renstrøm, S J Rose, E Sahin, S Siem, and T G Tornyi. 137,138,139 La(n,γ) cross sections constrained with statistical decay properties of 138,139,140 La nuclei. Phys. Rev. C, 95:045805, 2017. [68] S N Liddick, A C Larsen, M Guttormsen, A Spyrou, B P Crider, F Naqvi, J E Midtbø, F L Bello Garrote, D L Bleuel, L Crespo Campo, A Couture, A C Dombos, F Giacoppo, A G¨orgen, K Hadynska-Klek, T W Hagen, V W Ingeberg, B V Kheswa, R Lewis, S Mosby, G Perdikakis, C J Prokop, S J Quinn, T Renstrøm, S J Rose, E Sahin, S Siem, G M Tveten, M Wiedeking, and F Zeiser. Benchmarking the extraction of statistical neutron capture cross sections on short-lived nuclei for applications using the beta-Oslo method. Phys. Rev. C, 100:024624, 2019. [69] D. J. Morrissey, B. M. Sherrill, M. Steiner, A. Stolz, and I. Wiedenhoever. Commis- sioning the A1900 projectile fragment separator. Nuclear Instruments and Methods in Physics Research, Section B: Beam Interactions with Materials and Atoms, 204:90–96, 2003. [70] A. Stolz, T. Baumann, T. N. Ginter, D. J. Morrissey, M. Portillo, B. M. Sherrill, M. Steiner, and J. W. Stetson. Production of rare isotope beams with the NSCL 111 fragment separator. Nuclear Instruments and Methods in Physics Research, Section B: Beam Interactions with Materials and Atoms, 241:858–861, 12 2005. [71] A. Simon, S. J. Quinn, A. Spyrou, A. Battaglia, I. Beskin, A. Best, B. Bucher, M. Couder, P. A. Deyoung, X. Fang, J. Gorres, A. Kontos, Q. Li, S. N. Liddick, A. Long, S. Lyons, K. Padmanabhan, J. Peace, A. Roberts, D. Robertson, K. Smith, M. K. Smith, E. Stech, B. Stefanek, W. P. Tan, X. D. Tang, and M. Wiescher. SuN: Summing NaI(Tl) gamma-ray detector for capture reaction measurements. Nuclear In- struments and Methods in Physics Research, Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 703:16–21, 2013. [72] H. Bateman. The solution of a system of differential equations occurring in the theory of radioactive transformations. Proc. Cambridge Philos. Soc., 15:423–427, 1910. [73] Daniel Abriola, Melih Bostan, Sefa Erturk, Manssour Fadil, Monica Galan, Sakari Juu- tinen, Tibor Kib´edi, Filip Kondev, Aurelian Luca, Alexandru Negret, Ninel Nica, Bernd Pfeiffer, Balraj Singh, Alejandro Sonzogni, Janos Timar, Jagdish Tuli, Tsanka Venkova, and Kazimierz Zuber. Nuclear Data Sheets for A = 84. Nuclear Data Sheets, 110:2815– 2944, 2009. [74] E. A. McCutchan. Nuclear data sheets for A = 83. Nuclear Data Sheets, 125:201–394, 2015. [75] Rebecca L Lewis. Indirect Neutron-capture Cross Sections for the Weak r-process. PhD thesis, Michigan State University, 2019. [76] M. Guttormsen. Matrix Manipulation Program (MAMA), 2019. [77] M. Guttormsen, A. C. Larsen, A. Gorgen, T. Renstrom, S. Siem, T. G. Tornyi, and G. M. Tveten. Validity of the Generalized Brink-Axel Hypothesis in Np 238. Physical Review Letters, 116(1):1–5, 2016. [78] J. Allison, K. Amako, J. Apostolakis, P. Arce, M. Asai, T. Aso, E. Bagli, A. Bag- ulya, S. Banerjee, G. Barrand, B. R. Beck, A. G. Bogdanov, D. Brandt, J. M.C. Brown, H. Burkhardt, Ph Canal, D. Cano-Ott, S. Chauvie, K. Cho, G. A.P. Cirrone, G. Cooper- man, M. A. Cort´es-Giraldo, G. Cosmo, G. Cuttone, G. Depaola, L. Desorgher, X. Dong, A. Dotti, V. D. Elvira, G. Folger, Z. Francis, A. Galoyan, L. Garnier, M. Gayer, K. L. Genser, V. M. Grichine, S. Guatelli, P. Gu`eye, P. Gumplinger, A. S. Howard, I. Hˇrivn´aˇcov´a, S. Hwang, S. Incerti, A. Ivanchenko, V. N. Ivanchenko, F. W. Jones, S. Y. Jun, P. Kaitaniemi, N. Karakatsanis, M. Karamitrosi, M. Kelsey, A. Kimura, T. Koi, H. Kurashige, A. Lechner, S. B. Lee, F. Longo, M. Maire, D. Mancusi, A. Man- tero, E. Mendoza, B. Morgan, K. Murakami, T. Nikitina, L. Pandola, P. Paprocki, J. Perl, I. Petrovi´c, M. G. Pia, W. Pokorski, J. M. Quesada, M. Raine, M. A. Reis, A. Ribon, A. Risti´c Fira, F. Romano, G. Russo, G. Santin, T. Sasaki, D. Sawkey, J. I. Shin, I. I. Strakovsky, A. Taborda, S. Tanaka, B. Tom´e, T. Toshito, H. N. Tran, P. R. 112 Truscott, L. Urban, V. Uzhinsky, J. M. Verbeke, M. Verderi, B. L. Wendt, H. Wen- zel, D. H. Wright, D. M. Wright, T. Yamashita, J. Yarba, and H. Yoshida. Recent developments in GEANT4. Nucl. Instr. Meth. Phys. Res. A, 835:186–225, 11 2016. [79] M. Guttormsen, A.C. Larsen, A. Spyrou, S.N. Liddick, and F. Zeiser. Unfolding of excitation-energy spectra from total absorption γ-ray spectroscopy for the β-Oslo method. In Preparation. 2020. [80] A.V. Ignatyuk. Contribution to the Second CRP Meeting on RIPL. Technical report, Vienna, 11 1996. [81] P. Carlos, H. Beil, R. Bergere, J. Fagot, A. Lepretre, A. Veyssiere, and G. V. Solodukhov. A Study of the Photoneutron Contribution to the Giant Dipole Resonance of Nuclei in the 64 ¡ A ¡ 86 Mass Region. Nuclear Physics, 258:365–387, 1976. [82] S.F. Mughabghab. Atlas of Neutron Resonances: Resonance Parameters and Thermal Cross Sections Z = 1-100. Elsevier, 5th edition, 2006. [83] A.V. Ignatyuk, J.L. Well, S. Raman, and S. Kahane. Density of discrete levels in 116Sn. Phys. Rev. C, 47:1504–undefined, 1993. [84] R. Capote, M. Herman, P. Obloˇzinsk´y, P. G. Young, S. Goriely, T. Belgya, A. V. Ignatyuk, A. J. Koning, S. Hilaire, V. A. Plujko, M. Avrigeanu, O. Bersillon, M. B. Chadwick, T. Fukahori, Zhigang Ge, Yinlu Han, S. Kailas, J. Kopecky, V. M. Maslov, G. Reffo, M. Sin, E. Sh Soukhovitskii, and P. Talou. RIPL - Reference Input Parameter Library for Calculation of Nuclear Reactions and Nuclear Data Evaluations. Nuclear Data Sheets, 110(12):3107–3214, 2009. [85] S. Goriely. Radiative neutron captures by neutron-rich nuclei and the r-process nucle- osynthesis. Phys. Lett. B, 436(1-2):10–18, 1998. [86] M. Heil, R. Reifarth, M. M. Fowler, R. C. Haight, F. K¨appeler, R. S. Rundberg, E. H. Seabury, J. L. Ullmann, J. B. Wilhelmy, and K. Wisshak. 4π BaF2 detector for (n,γ) cross-section measurements at a spallation neutron source. Nuclear Instruments and Methods in Physics Research, Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 459:229–246, 2001. [87] C J Prokop, A Couture, S Jones, S Mosby, G Rusev, J Ullmann, and M Krtiˇcka. Measurement of the 65Cu(n, γ) cross section using the Detector for Advanced Neutron Capture Experiments at LANL. Phys. Rev. C, 99, 2019. [88] S. Mosby, F. Tovesson, A. Couture, D. L. Duke, V. Kleinrath, R. Meharchand, K. Meier- bachtol, J. M. ODonnell, B. Perdue, D. Richman, and D. Shields. A fission fragment de- tector for correlated fission output studies. Nuclear Instruments and Methods in Physics 113 Research, Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 757:75–81, 9 2014. [89] P W Lisowski, C D Bowman, G J Russell, and S A Wender. The Los Alamos National Laboratory Spallation Neutron Sources. Nuclear Science and Engineering, 106(2):208– 218, 10 1990. [90] D Habs, F S Stephens, and R M Diamond. A Proposal for a Crystal-Ball Detector System. Technical report, 1979. [91] M. Jandel, T. A. Bredeweg, A. Couture, M. M. Fowler, E. M. Bond, M. B. Chadwick, R. R.C. Clement, E. I. Esch, J. M. O’Donnell, R. Reifarth, R. S. Rundberg, J. L. Ullmann, D. J. Vieira, J. B. Wilhelmy, J. M. Wouters, R. A. Macri, C. Y. Wu, and J. A. Becker. GEANT4 simulations of the DANCE array. Nuclear Instruments and Methods in Physics Research, Section B: Beam Interactions with Materials and Atoms, 261(1-2 SPEC. ISS.):1117–1121, 8 2007. [92] A. Couture and R. Reifarth. Direct measurements of neutron capture on radioactive isotopes. Atomic Data and Nuclear Data Tables, 2007. [93] Saint-Gobain Crystals. BaF 2 Barium Fluoride Scintillation Material, 2016. [94] Jan M. Wouters, Ana Alpizar Vicente, Todd A. Bredeweg, Ernst Esch, Robert C. Haight, Robert Hatarik, John M. O’Donnell, Rene Reifarth, Robert S. Rundberg, Jon M. Schwantes, Steven A. Sheets, John L. Ullmann, David J. Vieira, and Jerry B. Wilhelmy. Acquisition - Analysis system for the DANCE (detector for advanced neutron capture experiments) BaF 2 gamma-ray calorimeter. IEEE Transactions on Nuclear Science, 53(3):880–885, 6 2006. [95] Trace Sciences International. Stable Isotopes Supplier. [96] M. B. Chadwick, M. Herman, P. Obloˇzinsk´y, M. E. Dunn, Y. Danon, A. C. Kahler, D. L. Smith, B. Pritychenko, G. Arbanas, R. Arcilla, R. Brewer, D. A. Brown, R. Capote, A. D. Carlson, Y. S. Cho, H. Derrien, K. Guber, G. M. Hale, S. Hoblit, S. Holloway, T. D. Johnson, T. Kawano, B. C. Kiedrowski, H. Kim, S. Kunieda, N. M. Larson, L. Leal, J. P. Lestone, R. C. Little, E. A. McCutchan, R. E. MacFarlane, M. MacInnes, C. M. Mattoon, R. D. McKnight, S. F. Mughabghab, G. P.A. Nobre, G. Palmiotti, A. Palumbo, M. T. Pigni, V. G. Pronyaev, R. O. Sayer, A. A. Sonzogni, N. C. Summers, P. Talou, I. J. Thompson, A. Trkov, R. L. Vogt, S. C. van der Marck, A. Wallner, M. C. White, D. Wiarda, and P. G. Young. ENDF/B-VII.1 nuclear data for science and technology: Cross sections, covariances, fission product yields and decay data. Nuclear Data Sheets, 112(12):2887–2996, 12 2011. 114 [97] F. Becvar. Simulation of gamma cascades in complex nuclei with emphasis on assessment of uncertainties of cascade-related quantities. Nuclear Instruments and Methods in Physics Research A, 417:434–449, 1998. [98] I. Dillmann, R. Plag, F Kappeler, and T. Rauscher. The third update of the ”Karl- sruhe Astrophysical Database of Nucleosynthesis in Stars”. EFNUDAT Fast Neutrons, Proceedings of the Scientific Workshop on Neutron Measurements, 2009. 115