PROTON-CAPTURE CROSS-SECTION MEASUREMENTS FOR THE ASTROPHYSICAL GAMMA PROCESS: FROM STABLE TO RADIOACTIVE ION BEAMS By Artemis Tsantiri A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Physics—Doctor of Philosophy 2025 ABSTRACT One of the key questions in nuclear astrophysics is understanding how elements heavier than iron are forged in the stars. Heavy element nucleosynthesis is primarily governed by the slow and rapid neutron capture processes. However, a relatively small group of naturally occurring, neutron- deficient isotopes, known as p nuclei, cannot be formed by either of those processes. These ∌ 30 nuclei are believed to be synthesized in the 𝛟 process, where preexisting r- and s-process seeds are “burned" through a sequence of photodisintegration reactions. The astrophysical sites where such conditions occur have been a subject of controversy for more than 60 years, and is currently believed that the 𝛟 process can take place in the O/Ne layers of core collapse supernovae, and in thermonuclear supernovae. Reproducing solar p-nuclei abundances through nuclear reaction networks requires input on a large number of mostly radioactive isotopes. However, as experimental cross sections of 𝛟- process reactions are very limited, and almost entirely unknown for radioactive nuclei, the related reaction rates are based on Hauser-Feshbach (HF) theoretical calculations and therefore carry large uncertainties. Therefore, it is crucial to develop techniques to accurately measure these reactions within the astrophysically relevant Gamow window with radioactive beams. The SuN group at the Facility for Rare Isotope Beams (FRIB) has been developing such a program for the past decade. This thesis focuses on implementing a technique to measure reaction cross sections in inverse kinematics with a radioactive beam. Specifically, this work presents data analysis from the proof- of-principle stable beam experiment for the 82Kr(p,𝛟)83Rb reaction, along with the measurement of the 73As(p,𝛟)74Se reaction in our first radioactive beam experiment. The latter reaction is particularly significant for the final abundance of the lightest p nucleus, 74Se, since the inverse reaction 74Se(𝛟,p)73As is one of the primary destruction mechanisms of 74Se. The experiments were conducted at FRIB at Michigan State University using the ReA facility. The 82Kr and 73As beams were directed onto a hydrogen gas cell located in the center of the Summing NaI(Tl) (SuN) detector and the obtained spectra were analyzed using the 𝛟-summing technique. In addition to the total cross section measurements, this thesis also presents the development of an analysis technique to extract statistical properties of the compound nucleus (nuclear level density and 𝛟-ray strength function) through a series of simulations. This approach enables the extraction of an experimentally constrained cross section across the entire Gamow window of the 𝛟 process. Finally, the experimentally constrained reaction rate for the 73As(𝑝, 𝛟)74Se reaction is used in Monte Carlo one-zone network simulations of the 𝛟 process to explore its impact on the production of the 74Se. Copyright by ARTEMIS TSANTIRI 2025 Στη ΌάΜα Όου, που Όου έΎωσε φτερά Μα πετάΟω. v ACKNOWLEDGEMENTS First and foremost, I would like to express my deepest gratitude to my thesis supervisor, Artemis Spyrou, for accepting me as a graduate student and for being an incredible mentor over the past five years. Artemis generously provided the means for an excellent education with great emphasis on professional development, encouraging me to attend multiple summer schools, conferences and workshops. Through her guidance, I learned how to work within a research group and a larger collaboration, and I am especially grateful for her support in helping me build my own professional network by introducing me to her collaborators and providing opportunities to establish connections that will be invaluable throughout my career. Most importantly, I deeply appreciate the trust she placed in me by allowing me to chose my thesis topic, and her hands-off approach to supervision that gave me the freedom to explore beyond my thesis, develop my own ideas, and grow both scientifically and personally. Looking back, I am certain that my decision to move halfway across the world to work with her has been the best academic decision I have ever made. I would like to thank Hendrik Schatz and Sean Liddick, for all their help and guidance during my graduate studies and my thesis project. I also want to express my graditude to the rest of my supervisory committee, Scott Bogner, Kendall Mahn and Brian O’Shea for their advice, questions, and offering their valuable time to serve on my guidance committee. I would like to thank all past and present members of the SuN group for their support and camaraderie. I am especially grateful to the former and current postdocs — Andrea, Erin, Mallory, and Sivi — for their guidance and friendship, with special thanks to Erin for her support and help during the execution of my experiment, and for teaching me to keep moving forward with a smile, calmness, and lightheartedness when things don’t go as planned. I also want to acknowledge former graduate students Alicia, Steven, and Alex, whose work laid the foundation for this research. To my fellow current graduate students — Hannah, Jordan, Caley, Kostas, Amal, and Sydney —thank you for making this journey both productive and enjoyable, especially to Hannah who has been my academic sister, and best office and travel mate. I wish all of them the best in their next stages of their career. I also extend my gratitude to the members of the 𝛜 group and everyone I have vi interacted with through the SuN detector. This group has truly felt like a family, and I thank them for all the great memories and adventures. I would also like to thank all the staff physicists and faculty at FRIB who contributed to the execution of my thesis experiment and this project. I am especially grateful to Jorge Pereira and Remco Zegers for their support with the hydrogen gas handling system, Aaron Chester for teaching me the ins and outs of the FRIB DAQ and helping me set up the experiment’s software, and Ana Henriques, Sam Nash, and Chandana Sumithrarachchi for assisting with the setup in ReA3 and providing such a beautiful—yet tricky—arsenic beam. I am also very grateful to Pablo Giuliani for his generous help on the mathematical description of the scores for the astrophysical calculations and for allowing me to bother him so often. Additionally, I want to thank the members of the NuGrid Collaboration for welcoming me into their community and enabling the astrophysical calculation portion of my dissertation. In particular, I am grateful to Pavel Denissenkov and Falk Herwig for hosting me during my internship at the University of Victoria, introducing me to the NuGrid codes, and engaging in many interesting discussions during my visit. I also appreciate the support of Thanassis Psaltis and Lorenzo Roberti for their help in setting up this impact study, sharing their codes, and contributing valuable ideas and discussions. Finally I am thankful for Umberto Battino and Claudia Travaglio for providing the trajectories for the Type Ia impact study. I am also grateful for the support of IReNA throughout my graduate studies, including the generous funding that allowed me to attend numerous conferences and summer schools, as well as the Visiting Fellowship that made my research stay at UVic possible. I am deeply grateful for all my close friends, both near and far. To Pely, Nikos, Yannis, and Emily — thank you for the memories, adventures, and for making Lansing feel like home. To my friends from afar—Christos, Christina, Anna, Orestis, and Eva—your unwavering support has meant the world to me, no matter the distance. You have always been with me wherever I go. A very special thank you to my partner, best friend, roommate, and colleague, Jordan. Meeting you has been the best part of this journey, and I can’t wait to see where our next adventures take us. vii Finally, I would like to express my deepest gratitude to my family—my parents, Despoina and Nikos, and my sister, Christina. I cannot overstate my appreciation for their love, support, and the countless sacrifices they have made. They have always encouraged me to pursue my goals, no matter how far they took me. Especially to my mother, to whom I dedicate this thesis—I doubt many people can say their mother followed them to a scientific conference because Austria was the closest they had been to home in a while. And, of course, to my cat, who has been a constant source of comfort, chaos, and companionship through it all. viii CHAPTER 1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 TABLE OF CONTENTS CHAPTER 2 NUCLEAR PHYSICS FOR ASTROPHYSICS . . . . . . . . . . . . . . 2.1 Nuclear Masses and Binding Energies . . . . . . . . . . . . . . . . . . . . . . 2.2 Energetics of Nuclear Reactions . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Reaction Cross Section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Nuclear Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Nuclear Reactions in Stars 2 3 5 6 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 . CHAPTER 3 . . ASTROPHYSICS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 3.1 Abundances . . . . 3.2 Stellar Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 3.3 Stellar Nucleosynthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 3.4 Production of the p Nuclei 3.5 Nuclear Networks and Uncertainties . . . . . . . . . . . . . . . . . . . . . . . 45 3.6 The Lightest p Nucleus, 74Se . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 CHAPTER 4 EXPERIMENTAL SETUP &TECHNIQUES . . . . . . . . . . . . . . . 51 4.1 Beam Delivery in ReA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 4.2 Hydrogen Gas Target . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 . 4.3 The SuN Detector . 4.4 The SuNSCREEN Detector . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 CHAPTER 5 ANALYSIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 . 5.1 Effective Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 5.2 Beam Particle Number 5.3 Experimental Yield . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 5.4 Target Particle Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 5.5 Detection Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 5.6 Theoretical Investigation with Rainier and Talys . . . . . . . . . . . . . . . . 79 5.7 Uncertainty Quantification . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 CHAPTER 6 RESULTS &DISCUSSION . . . . . . . . . . . . . . . . . . . . . . . . 88 6.1 The 82Kr(𝑝, 𝛟)83Rb cross section . . . . . . . . . . . . . . . . . . . . . . . . . 88 6.2 The 73As(𝑝, 𝛟)74Se cross section . . . . . . . . . . . . . . . . . . . . . . . . . 93 CHAPTER 7 ASTROPHYSICAL IMPACT . . . . . . . . . . . . . . . . . . . . . . . 98 . . . . . . . . . . . . . . . . . . . . . . . . . 98 7.1 Core-Collapse Supernova - SNII 7.2 Type Ia Supernova - SNIa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 CHAPTER 8 SUMMARY & CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . 106 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 ix APPENDIX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 x CHAPTER 1 INTRODUCTION Carl Sagan said, “we are made of star stuff". But how is this “star stuff" made? The carbon in our cells, the calcium in our bones, and the oxygen in our blood are all forged during the life of a star. But the silver and gold in our jewelry, the platinum in our cars, and the tungsten in our LED lights were all made during a star’s death. The origins of nuclear astrophysics trace back over a century to Eddington’s 1920 manuscript, which, based on Aston’s demonstration that the mass of helium is less than four times that of the proton [1], proposed that an unknown process in the Sun’s core converts hydrogen into helium, releasing energy [2]. Nearly 40 years later, the groundbreaking review by Burbidge, Burbidge, Fowler, and Hoyle (B2FH) presented a detailed framework of stellar nucleosynthesis, describing how the elements are synthesized in stars [3]. These discoveries highlighted the need for a field that combines astronomy, astrophysics, and nuclear physics to explore how the elements in the universe came to be, the interdisciplinary field of nuclear astrophysics. The following thesis aims to offer but a small contribution to this ongoing effort, by focusing on the nucleosynthesis of a particular isotope, namely 74Se. Given the interdisciplinary nature of nuclear astrophysics, the necessary theoretical foundations are introduced in separate chapters covering nuclear physics and astrophysics. Then, the experimental study of the destruction mecha- nism of 74Se is presented, followed by analysis of the experimental data. Finally, the impact of the measurement is investigated through astrophysical network calculations, aiming to constrain the final production of 74Se in stars. 1 CHAPTER 2 NUCLEAR PHYSICS FOR ASTROPHYSICS As per Rutherford’s description, the majority of the atom’s volume is comprised of empty space and electrons that surround a tiny, dense, positively charged central core called the nucleus. Composed of protons and neutrons, the nucleus is characterized by the mass number, 𝐎, which represents their total number. The number of protons is called the atomic number 𝑍, and the neutron number 𝑁 is defined as 𝑁 = 𝐎 − 𝑍. Different elements are distinguished by the different atomic number 𝑍, whereas nuclei of the same element with different neutron number 𝑁 are called isotopes. The 𝑍 𝑋𝑁 , or simply 𝐎 𝑋, where 𝑋 represents the chemical symbol of notation for different isotopes is 𝐎 the element. Similarly to chemicals elements in the periodic table, isotopes are represented in the chart of the nuclides. This two-dimensional chart, shown in Fig. 2.1 displays the number of protons, 𝑍, on the Figure 2.1 The chart of nuclides. The black squares correspond to stable nuclei, while the gray to radioactive. The white lines indicate proton and neutron magic numbers. Data from IAEA [4]. 2 y-axis, and the number of neutrons 𝑁 on the x-axis. Each square represents a different isotope with different properties. The black squares correspond to stable isotopes that can be found in nature, while the gray are radioactive, meaning that after some time they will lose energy by emitting radiation. The white lines correspond to proton and neutron magic numbers, which according to the nuclear shell model developed Mayer in 1948 [5], refer to specific numbers of protons or neutrons within a nucleus that result in significantly increased stability. The magic numbers for nuclei are 2, 8, 20, 28, 50, 82, and 126. 2.1 Nuclear Masses and Binding Energies One might initially assume that the nuclear mass, 𝑚nuc, equals 𝑍 𝑚 𝑝 + 𝑁 𝑚𝑛, where 𝑚 𝑝 and 𝑚𝑛 are the masses of protons and neutrons that comprise the nucleus, respectively. However, Aston’s experiments in the 1910s revealed that the mass of helium (4He) is less than 4 times the mass of hydrogen (1H) [1]. This mass difference is called the mass excess or mass defect, Δ𝑚, and is a direct result of the binding energy that holds the nucleus together. The binding energy, 𝐵(𝑍, 𝑁), represents the energy required to break a nucleus into its constituent 𝑍 protons and 𝑁 neutrons, and can be expressed as 𝐵(𝑍, 𝑁) = Δ𝑚 · 𝑐2 = (𝑍 𝑚 𝑝 + 𝑁 𝑚𝑛 − 𝑚nuc) 𝑐2 (2.1) [6], where 𝑐2 often includes a unit conversion factor so that 𝑐2 = 931.50 MeV/u, and thus the binding energy is expressed in atomic mass units. A useful nuclear property derived from the binding energy is the particle separation energy, which corresponds to the energy required to remove a proton or a neutron from a nucleus. Hence the proton separation energy is equal to the difference between the binding energy of 𝐎 𝑍 𝑋𝑁 and 𝐎−1 𝑍−1 𝑋𝑁 : 𝑆 𝑝 = 𝐵(𝑍, 𝑁) − 𝐵(𝑍 − 1, 𝑁) = [𝑚( 𝐎 𝑍 𝑋𝑁 ) − 𝑚( 𝐎−1 𝑍−1 𝑋𝑁 ) + 𝑚(1H)] 𝑐2 and similarly the neutron separation energy is: 𝑆𝑛 = 𝐵(𝑍, 𝑁) − 𝐵(𝑍, 𝑁 − 1) = [𝑚( 𝐎 𝑍 𝑋𝑁 ) − 𝑚( 𝐎−1 𝑍 𝑋𝑁−1) + 𝑚𝑛] 𝑐2 3 (2.2) (2.3) The separation and binding energies carry important information about the stability and structure of the nuclides. To allow for a systematic study of the nuclear binding energy, it is common to display the average binding energy per nucleon, 𝐵/𝐎. Figure 2.2 shows 𝐵/𝐎 as a function of mass number 𝐎 [7]. A few notable features shown in Fig. 2.2 include that, aside from the light nuclei, the average Figure 2.2 The binding energy per nucleon, 𝐵(𝑍, 𝑁)/𝐎, as a function of the mass number, 𝐎. Figure from [8]. binding energy is around 8 MeV/u. The most bound nuclei, those with the maximum 𝐵/𝐎, are found in the mass range of 𝐎 = 50 − 65. This is the so-called iron peak, with the most tightly bound nuclides being 62Ni, 58Fe, and 56Fe [9]. It follows that there are two ways to release energy through nuclear processes: for nuclei lighter than the iron peak energy is released by fusion, the assembly of light nuclei into heavier species, while for nuclei heavier than iron, energy is released by fission, the breaking of heavy nuclei into lighter ones [7]. As will be described later in Ch. 3, these are the two main mechanisms for energy generation in a stellar environment. 4 2.2 Energetics of Nuclear Reactions A binary nuclear interaction is written as 𝑎 + 𝑋 → 𝑌 + 𝑏 or 𝑋 (𝑎, 𝑏)𝑌 (2.4) where 𝑎 and 𝑋 are the two colliding nuclei (entrance channel), and 𝑏 and 𝑌 are the reaction products (exit channel). Typically, 𝑎 is an accelerated projectile of lighter mass and 𝑋 is a stationary heavy target, while 𝑏 is the light ejectile that can be directly measured and 𝑌 is the heavy recoil nucleus that stays in the target and is not observed. The various classifications of nuclear reactions are discussed in Sec. 2.4. As with every other interaction in nature, nuclear reactions are governed by fundamental conservation laws, which provide a basis for deriving various characteristic quantities to describe the system. For example the conservation of total energy and linear momentum allows us to deduce the energy of the undetected recoil nucleus from the known energies of the reactants and the measured energy of the ejectile. Other conserved quantities include the angular momentum, proton and neutron number (or baryon number), and parity [7]. The conservation of total relativistic energy for a reaction of the form shown in Eq. 2.4 yields 𝑚 𝑋 𝑐2 + 𝐞 𝑋 + 𝑚𝑎 𝑐2 + 𝐞𝑎 = 𝑚𝑌 𝑐2 + 𝐞𝑌 + 𝑚𝑏 𝑐2 + 𝐞𝑏 (2.5) where 𝐞𝑖 are the kinetic energies in the center-of-mass system and 𝑚𝑖 the rest masses. The energy available in the system for this reaction is defined as the Q value, and represents the difference in mass energy of the system before and after the reaction. 𝑄 = (𝑚initial − 𝑚final) 𝑐2 = (𝑚 𝑋 + 𝑚𝑎 − 𝑚𝑌 − 𝑚𝑏) 𝑐2 or in terms of the excess of kinetic energy 𝑄 = 𝐞final − 𝐞initial = 𝐞𝑌 + 𝐞𝑏 − 𝐞 𝑋 − 𝐞𝑎 5 (2.6) (2.7) If the 𝑄 is positive, then the reaction releases energy, and is called exoergic or exothermic. If 𝑄 is negative, then energy is consumed for the reaction to occur, and it is called endoergic or endothermic [7]. The particle separation energy introduced in Sec. 2.1 corresponds to the 𝑄 value for particle emission. The proton separation energy, 𝑆 𝑝, corresponds to the 𝑄 value for proton emission, while the neutron separation energy, 𝑆𝑛, corresponds to that for neutron emission. This will become important later in Ch. 3, as these quantities define the limits of the nuclear landscape. 2.3 Reaction Cross Section One of the most important quantities characterizing a nuclear reaction is the cross section, 𝜎, which can be broadly understood as the probability for an interaction to occur. Consider the geometry illustrated in Fig. 2.3. An incident beam of 𝐌𝑎 particles per unit time impinges on a Figure 2.3 Illustration of a typical nuclear physics cross section measurement, showing an incident beam, target and detector. Figure recreated based on [6, 7]. target of 𝑁𝑡 particles per unit area. A detector is positioned at an angle (𝜃, 𝜙) with respect to the beam axis, and its surface covers a solid angle 𝑑Ω. If the rate of outgoing ejectile particles is 𝑅𝑏, 6 then the reaction cross section is defined by 𝜎 = = (interactions per unit time) (incident particles per unit time) (target nuclei per unit area) 𝑅𝑏 𝐌𝑎 𝑁𝑡 (2.8) By this definition, the cross section has dimensions of area, but is proportional to the reaction probability and is typically measured in barns, where 1 b = 10−24 cm2. The detector of Fig. 2.3 covers but a small solid angle 𝑑Ω, and therefore could not have possibly detected all outgoing particles. The ejectiles are emitted in a non uniform manner, and if we assume an angular distribution 𝑟 (𝜃, 𝜙) for the emitted ejectiles, then the fraction of ejectiles detected would be 𝑑𝑅𝑏 = 𝑟 (𝜃, 𝜙) 𝑑Ω/4𝜋. The illustrated geometry would allow for the measurement of the differential cross section, 𝑑𝜎/𝑑Ω = 𝑟 (𝜃, 𝜙)/(4𝜋 𝐌𝑎 𝑁𝑡). The total reaction cross section can be calculated by integrating 𝑑𝜎/𝑑Ω over all angles, where 𝑑Ω = sin 𝜃 𝑑𝜃 𝑑𝜙. 𝜎 = ∫ 𝑑𝜎 𝑑Ω 𝑑Ω = ∫ 𝜋 0 sin 𝜃 𝑑𝜃 ∫ 2𝜋 0 𝑑𝜙 𝑑𝜎 𝑑Ω (2.9) In this work, as described later on in Ch. 4, the detector geometry has a solid angle of 4𝜋. Therefore this work regards to a total cross section measurement, and the analysis follows Eq. 2.8, with the addition of a detection efficiency term, 𝜖. The efficiency accounts for the fraction of the outgoing ejectiles that enter the active volume of the detector, but are not recorded, so that 𝑅𝑏 = 𝑌 /𝜖, where 𝑌 is the experimental yield, meaning the total number of particles detected. The analysis presented in Ch. 5 follows: 𝜎 = 𝑌 𝐌𝑎 𝑁𝑡 𝜖 (2.10) 2.4 Nuclear Reactions Nuclear interactions of the form 𝑋 (𝑎, 𝑏)𝑌 are categorized based on the nature of the species involved, as well as the mechanisms governing the process. If the reactants 𝑎 and 𝑋 are identical to the reaction products, the interaction is referred to as scattering. Scattering is classified as elastic if the products remain in their ground state and as inelastic if they are in an excited state. Otherwise, the reactants and products are distinct species and a nuclear reaction occurs. If particle 𝑏 is a 𝛟 ray, 7 the reaction is called radiative capture, whereas if particle 𝑎 is a 𝛟 ray, it is a photodisintegration. In cases where particles 𝑎 and 𝑏 are identical but an additional ejectile is present (resulting in three final products), the reaction is referred to as a knockout reaction. If one or two nucleons are exchanged between the projectile and target, this is classified as a transfer reaction. Transfer reactions can be further categorized as pick-up reactions, where the projectile acquires nucleons from the target, or stripping reactions, where the target removes nucleons from the projectile [7, 10]. Lastly, if the projectile exchanges a proton for a neutron or vice-versa, the process is known as a 𝑐ℎ𝑎𝑟𝑔𝑒 − 𝑒𝑥𝑐ℎ𝑎𝑛𝑔𝑒 reaction. Another important way to classify nuclear reactions is based on the governing mechanism, which determines the timescale of the interaction and the extent to which the target nucleus is affected. Imagine you’re running through a forest. If you run slowly, you have the time to observe each tree, interact with them, and maybe even touch their leaves. The entire forest feels your presence as you pass through it. But if you’re running really fast, you barely notice the trees. All the forest notices is a blur, like a bullet, and you might only interact with a single tree if you hit it directly. Most of the forest remains unaffected by your passage. Nuclear reactions work in a similar way.1 At low energies, the incoming particle has a large de Broglie wavelength, comparable to the size of the whole nucleus. This allows it to interact with the entire nucleus, forming a compound nucleus. For example, a 1 MeV proton has a de Broglie wavelength of about 4 fm, which is equal to the average radius of the Fe nucleus. At higher energies, the particle’s de Broglie wavelength becomes much smaller, and it’s more likely to interact with individual nucleons. A 50 MeV proton, for instance, with a de Broglie wavelength of about 0.6 fm, is more likely to perform a direct reaction. In between these two mechanisms are the pre-equilibrium reactions, in which the system of the reactants breaks up before it reaches statistical equilibrium. 2.4.1 Direct Reactions Direct reactions, also called peripheral, involve the interaction of one or very few particles from the target with the projectile. These reactions occur on a timescale of approximately 10−22 seconds 1This metaphor was first introduced to me by my undergraduate supervisor, Mike Kokkoris. 8 and primarily affect the surface of the target, leaving the remaining nucleons largely unaffected as spectators. An example of such reactions are transfer reactions, which are commonly used to study the structure of nuclei. They may insert or remove a particle from a specific state within the nucleus, leaving the rest of the system unperturbed. Such experiments allow to probe particle states of specific angular momentum, spin and parity, by detecting ejectile particles at different angles, and most often regard to the measurement of differential cross sections [7, 11]. 2.4.2 Compound Nucleus Reactions The bound nuclear states studied in direct reactions are stable against particle emission. There- fore, their lifetimes, 𝜏, are very long, and have a narrow width, Γ, corresponding to a small uncertainty in their energy, based on Heisenberg’s uncertainty principle Γ = Δ𝐞 = ℏ/𝜏 [12]. On the opposite side is the compound nucleus mechanism, in which the incoming particle 𝑎 and target 𝑋 merge, populating an excited state of a compound nucleus 𝐶∗. The captured particle remains in the compound system for an extended period, typically on the order of 10−16 to 10−18 seconds. Unlike direct reactions, this timescale allows the incoming particle to interact randomly with a very large number of nucleons, sharing its energy across the entire system. The resulting compound nucleus has undergone so many small interactions that loses any memory of its formation mechanism. As a result, the entrance and exit channels of the system can be treated independently, an idea described by Bohr’s independence hypothesis [13, 14]. Compound reactions can populate either the resonance region or the statistical region (also known as continuum), depending on the excitation energy of the compound nucleus and the number of available states. The resonance region corresponds to discrete nuclear states, while the statistical region consists of numerous, closely spaced states that overlap. 2.4.2.1 Resonance Reactions In resonance reactions, the incoming particle becomes “quasibound" to a nuclear state with a very high probability of formation, resulting in a very large cross section. These states of the compound nucleus often have small widths and low excitation energies, and will decay either by emitting 𝛟 rays or by re-emitting the incident particle, as in scattering. 9 The cross section 𝜎(𝐞) for a resonance with energy 𝐞𝑅 and width Γ is described through the Breit-Wigner formula as 𝜎(𝐞) = 𝜎0 Γ2/4 (𝐞 − 𝐞𝑅)2 + 1 4 Γ2 (2.11) where 𝜎0 is the cross section value at the maximum of the resonance peak [10]. An illustration of a resonance reaction is shown in Fig. 2.4, where the incident particle 𝑎 is captured by the target 𝑋 and populating the 𝐞𝑅 state of the 𝑌 compound nucleus. The right-hand-side of the illustration shows the cross section for this capture. The resonant state then decays by either emitting 𝛟 rays, or re-emitting particle 𝑎. Figure 2.4 Illustration of a resonance reaction. Figure recreated based on [6, 15]. Resonant reactions are particularly important in nuclear astrophysics, as the existence of such resonances enables reactions that would otherwise hinder nucleosynthesis, due to very low cross sections, as described in Sec. 2.5. 2.4.2.2 Statistical Model of Compound Nucleus Reactions As the excitation energy, 𝐞 𝑋, of the compound nucleus populated by the reaction increases, the number of available nuclear states grows almost exponentially. At higher energies, the states become so numerous that their spacing is much smaller than their width, leading to significant overlap and resembling a structureless continuum. Under these conditions, the resonance reaction mechanism becomes inadequate, and the reaction is instead described by the statistical model of 10 compound reactions, initially developed by N. Bohr in 1936 [13]. The most widely used implementation of Bohr’s independence hypothesis in the statistical model is the Hauser-Feshbach (HF) theory [16], that results from averaging over a large number of Breit–Wigner resonances. The central quantities in the HF formalism are the averaged transmission coefficients 𝑇, that reflect the probability for a particle’s wavefunction to cross an obstacle. In this context, the transmission coefficients, instead of a resonance behavior, they describe the formation of the system as an absorption of the incident particle’s wavefunction in the nuclear potential. The cross section 𝜎 of the reaction 𝑎 + 𝑋 → 𝐶∗ → 𝑌 + 𝑏 (proceeding via compound nucleus 𝐶), is expressed as a summation over all possible spin and parity states of the compound system as 𝜎𝑎𝑋→𝑌 𝑏 (𝐞𝑎𝑋) = 𝜋ℏ2/(2𝜇𝑎𝑋 𝐞𝑎𝑋) (2𝐜𝑋 + 1)(2𝐜𝑎 + 1) ∑ 𝐜,𝜋 (2𝐜 + 1) 𝑎 (𝐞, 𝐜, 𝜋)𝑇𝑌 𝑇 𝑋 𝑏 (𝐞, 𝐜, 𝜋) 𝑇tot(𝐞, 𝐜, 𝜋) 𝑊 𝑎𝑋→𝑌 𝑏 (2.12) where 𝐞𝑎𝑋 the center mass energy, 𝜇𝑎𝑋 the reduced mass, 𝐜 and 𝜋 the spin and parity, and 𝑇 𝑋 𝑎 and 𝑇𝑌 𝑏 the transmission coefficients for the entrance and exit channels respectively. The summation includes all individual transitions to all 𝐜𝜋 states accessible by particle or photon emission from the same compound nucleus C, accounting for the quantum mechanical spin/parity selection rules. The final term 𝑊 𝑎𝑋→𝑌 𝑏 is the width fluctuation correction (WFC) and describes non-statistical correlations between the widths of the entrance and exit channels and is close to unity [14, 17]. The transmission coefficient for the entrance channel 𝑇 𝑋 𝑎 is typically calculated numerically by solving the Schrödinger equation with an optical nucleon-nucleus potential, which represents the average nuclear potential. The development of optical model potentials that accurately describe the complexity of the potential caused by the strong nuclear force has been an active field of study for decades. As a detailed discussion of these potentials exceeds the scope of this thesis, further information can be found in Refs. [18, 19, 20]. The transmission coefficient for the exit channel can be described by assuming all possible bound and unbound states 𝜈 in all energetically accessible exit channels. 𝑇𝑌 𝑏 (𝐞, 𝐜, 𝜋) = 𝜈𝑚∑ 𝜈=0 𝑇 𝜈 𝑏 (𝐞, 𝐜, 𝜋) + ∫ 𝐞𝑖 ∑ 𝐞 𝜈𝑚 𝐜,𝜋 𝑇𝑏 (𝐞, 𝐜, 𝜋, 𝐞𝑖, 𝐜𝑖, 𝜋𝑖) × 𝜌(𝐞𝑖, 𝐜𝑖, 𝜋𝑖) 𝑑𝐞𝑖 (2.13) 11 The first term on the right-hand side represents a summation over all experimentally known discreet states, 𝜈𝑚. The second term is an integration that accounts for the transmission coefficient of all excited states above the highest experimentally known state, 𝜈𝑚, weighted by the nuclear level density 𝜌(𝐞𝑖, 𝐜𝑖, 𝜋𝑖). This density corresponds to the number of available spin-parity states within an energy region 𝑑𝐞𝑖. If the ejectile 𝑏 is a particle, 𝑇𝑏 is calculated in a similar manner as 𝑇 𝑋 𝑎 , which requires knowledge of the optical potential that the ejectile must overcome in order to escape the compound system. In the case of radiative capture, however, particle 𝑏 is a 𝛟 ray, and the exit channel will be described by the 𝛟-ray transmission coefficient 𝑇𝛟, which is directly proportional to the 𝛟-ray strength function, and represents the escape probability of a 𝛟 ray that is stuck inside the volume of a nucleus. The nuclear level density and 𝛟-ray strength function will be further discussed in the following paragraphs. 2.4.2.3 Nuclear Level Density The nuclear level density (NLD) corresponds to the available quantum levels Δ𝑁 at a specific excitation energy 𝐞𝑥, spin 𝐜, and parity 𝜋, and is defined as: 𝜌(𝐞𝑥, 𝐜, 𝜋) = Δ𝑁 (𝐞𝑥, 𝐜, 𝜋) Δ𝐞𝑥 (2.14) where Δ𝐞𝑥 is the energy interval considered, typically 1 MeV. Summing over all possible spin and parity values gives the total level density 𝜌(𝐞𝑥) [21]. For the statistical-model formalism to apply, the total level density needs to be sufficiently high. While this criterion is somewhat relative, an accuracy of 20% in the description of the level densities with numerical calculations is achievable when 𝜌 ≳ 10 MeV−1 (non-overlapping, narrow resonances) [22]. For nuclei with mass 𝐎 > 60, excitation energies above approximately 4 MeV have sufficiently high level density for the statistical model to be applicable. The first theoretical description of level density was proposed by Bethe in 1936 [23], treating the nucleus as a gas of non-interacting fermions (protons and neutrons). While simplistic, this approach captured all the essential information, apart from the influence of the pairing between the nucleons, that was realized and described almost twenty years later by Bardeen, Cooper and Schrieffer [24]. This pairing was then introduced in the description of the level density as a simple 12 constant energy shift, that was later on found to be too large of a correction. This lead to the back-shifted Fermi Gas formula (BSFG) proposed by Gilbert and Cameron in 1965 [25]: 𝜌(𝑈) = √ √ 𝜋 12 1 2𝜋𝜎2 √ exp(2 𝑎𝑈) 𝑎1/4𝑈5/4 (2.15) where 𝑈 = 𝐞𝑥 − Δ is the shifted excitation energy. The energy shift Δ is an empirical parameter closely associated with the pairing energy, accounting for odd-even effects in nuclei. The concept behind Δ is that nucleon pairs must first be separated before their individual components can be excited. In practice, Δ serves as an adjustable parameter to reproduce observables. The 𝑎 term in Eq. 2.15, referred to as the level density parameter, is, in its most simplistic form, given by 𝑎 = 𝜋 6 (𝑔𝑝 + 𝑔𝑛), where 𝑔𝑝 and 𝑔𝑛 is the spacing of the proton or neutron single-particle states near the Fermi energy. Recognizing that 𝑎 should include energy-dependent shell effects, more sophisticated expressions for 𝑎 have been developped [26, 27, 28]. The spin cut-off parameter, 𝜎2, of Eq. 2.15 represents the width of the angular momentum distribution of the level density. The description of 𝜎2 is based on the observation that the nucleus possesses collective rotational energy, and the spin cut-off parameter is related to the moment of inertia of the undeformed nucleus 𝐌0, and the thermodynamic temperature 𝑡 = √𝑈/𝑎, so that 𝜎2 = 𝐌0 𝑡. Similarly to the parameter 𝑎, energy-dependent shell effects are often included in more advanced models for 𝜎2 [22, 29, 30]. An alternative analytical description of the level density is the Constant Temperature (CT) model, introduced by Ericson in 1959, who described it as incorporating “a temperature 𝜏 which is somewhat different from the ordinary nuclear temperature 𝑇, defined by the level density", [31]. This temperature 𝜏 is related to the nuclear temperature 𝑇 by: 𝑑 𝑑𝐞 log 𝜌(𝐞) = 1 𝑇 1 𝜏 = (cid:18) 1 − (cid:19) 𝑑𝜏 𝑑𝐞 and the level density is then described as 𝜌(𝐞 𝑋) = exp[(𝐞 𝑋 − 𝐞0)]/𝜏 𝜏 (2.16) (2.17) where in practice, 𝐞0 and 𝜏 are parameters used to adjust the formula to experimental discreet levels. Since the BSFG model diverges as 𝑈 → 0, a common practice is to use the CT model at low 13 energies and the BSFG model at higher excitation energies, with parameters to ensure a smooth transition between the two models. Additional approaches include the phenomenological Generalized Superfluid Model (GSM) [32, 33], which incorporates nucleon pairing correlations according to the Bardeen-Cooper-Schrieffer theory [24], along with various microscopic models grounded in first principles and fundamental interactions. These microscopic descriptions of the NLD can capture intricate details of nuclear structure that are beyond the capabilities of analytical expressions. One microscopic approach is the shell model Monte-Carlo by Alhassid [34], as well as the approach based on mean-field theory by Demetriou and Goriely [35]. Additional examples of microscopic models that will be included in the analysis in the next chapters include the calculated NLD by Goriely from Hartree-Fock calculations [36], parity-dependent NLD based on the microscopic combinatorial model by Hilaire [37], as well as temperature-dependent Hartree-Fock-Bogoliubov calculations using the Gogny force [38]. 2.4.2.4 Radiative Decay, Transmission Coefficients and 𝛟 Strength Function Gamma rays emitted from an excited nucleus must follow selection rules to conserve the angular momentum and parity. They are classified with an electric (𝐞) or magnetic (𝑀) character, along with a multipolarity, based on the angular momentum 𝐿 they carry, and the parity change Δ𝜋 between the initial 𝑖 and final state 𝑓 : |𝐌𝑖 − 𝐌 𝑓 | ≀ 𝐿 ≀ 𝐌𝑖 + 𝐌 𝑓 (𝐿 ≠ 0) Δ𝜋 = no : even electric, odd magnetic (2.18) Δ𝜋 = yes : odd electric, even magnetic For instance, a transition from an initial state of 𝐜 𝜋 𝑓 = 0+ involves angular momentum 𝐿 = 2 without a change of parity, making it an 𝐞2 transition. When many 𝑖 = 2+ to a final state of 𝐜 𝜋 multipolarities are possible, the lower multipoles are significantly more likely to occur. For example the transition from 𝐜 𝜋 𝑓 = 5/2+ permits 𝑀1, 𝐞2, 𝑀3 and 𝐞4 transitions. Among these, 𝑀1 transition is typically a thousand times more probable than 𝐞2, 𝐞2 a thousand times more 𝑖 = 3/2+ to 𝐜 𝜋 likely than 𝑀3, and so forth [7]. 14 However, even if a photon can be emitted according to the selection rules, its probability of escaping the volume of the nucleus is much smaller than the probability of being reflected back. This escape probability is described by the 𝛟-ray transmission coefficient 𝑇𝛟, a quantity that characterizes the average electromagnetic properties of excited states, and can be described through the 𝛟-ray strength function 𝑓 (𝐞𝛟) (also called photon strength function or radiative strength function) as: 𝑇𝑋 𝐿 (𝐞𝛟) = 2𝜋 𝐞 (2𝐿+1) 𝛟 𝑓𝑋 𝐿 (𝐞𝛟) (2.19) where 𝑋 denotes the character (𝐞 or 𝑀), 𝐿 the multipolarity, and 𝐞𝛟 the energy of the 𝛟 ray. Photon strength functions are important for the description of all transitions involving 𝛟 rays, but their significance is even more apparent in (𝑛, 𝛟) and (𝛟, 𝑛) reactions, as neutrons are not affected by the Coulomb force of the nucleus, and photon strength functions directly govern the reaction cross section. They are distinguished by the upward 𝛟-strength function −−→ 𝑓XL, associated with the average photo-absorption, and the downward strength function ←−− 𝑓XL, related to the 𝛟 decay. The treatment of photon strength functions involves two key assumptions. First, the strength function is assumed to be independent of 𝐜 and 𝜋 [39], an approximation valid when the initial and final state have high excitation energies, and therefore overlap with many states of the same energy and different 𝐜𝜋 values. Second, the upward and downward strength functions are assumed to be equal, implying that the photo-absorption cross section on an excited state will have the same shape as the photo-absorption on the ground state. This assumption is known as the Brink hypothesis [40]. In calculations of the 𝛟-ray transmission coefficient for astrophysics, at least the most dominant 𝐞1 and 𝑀1 transitions have to be considered. Similar to the level density, there is a plethora of models, both analytical and microscopic, to describe the dipole (𝐞1 and 𝑀1) strength functions. The 𝐞1 transitions are calculated on the basis of the Lorentzian representation of the giant dipole resonance (GDR), that has been observed throughout the periodic table to strongly influence the strength function. Macroscopically, this strong resonance is described as a vibration of the charged (proton) matter in the nucleus against the neutral matter (neutrons). The magnetic dipole (𝑀1) strength function is also commonly described by Lorentzian reso- nance-like structures that are much smaller in magnitude compared to the GDR. Depending how 15 deformed the many-body system is, collective excitations can appear as enhancements in the 𝑀1 strength function, such as the scissors mode around 3 MeV, or the spin-flip strength around 5-9 MeV [41, 42]. Examples of phenomenological models to describe the photon strength function that are widely used in astrophysics are the Standard Lorentzian function by Brink [40] and Axel [43], and the Generalized Lorentzian model of Kopecky and Chrien [44] and Kopecky and Uhl [45]. Mi- croscopic models to describe 𝐞1 and 𝑀1 radiation include, but are not limited to, large-scale calculations based on the quasi-particle random-phase approximation (QRPA) model combined with the Hartree–Fock–Bogoliubov (HFB) method [46, 47, 48, 49], and the relativistic mean-field approach (RMF) [50, 51, 52, 53]. Additionally to the Lorentzian resonance-like structures that comprise the form of the strength function, an enhancement at low transition energies and excitation energies in the statistical region has been experimentally observed [54, 55]. This feature is called the low energy enhancement or upbend, and even though it is not clear whether it correspond to the electric or magnetic radiation [56, 57], it is believed to be of dipole character [58]. The upbend is parameterized in the form of an exponential tail as 𝑓upbend(𝐞𝛟) = 𝐶 exp(−𝜂 𝐞𝛟) (2.20) where 𝐶 and 𝜂 are adjustable parameters [55]. The existence of the upbend has shown to have significant impact on capture reaction cross sections [59, 60], and its intensity appears to be dependent on the nuclear structure [61]. 2.4.2.5 Statistical Model Calculations with Talys A software package for simulations and predictions of nuclear reactions that will be extensively used in the analysis of the following chapters is Talys [62]. A variety of nuclear reactions can be simulated using Talys including direct reactions, compound nucleus model, pre-equilibrium reactions and fission. In the context of this thesis, Talys will be used for calculating (𝑝, 𝛟) reaction cross sections based on the statistical model for compound nucleus reactions. As discussed in the previous sections, main ingredients of the HF formalism include the optical model potential (OMP), 16 the nuclear level density (NLD) and 𝛟-ray strength function (𝛟SF). In this paragraph the models used for the description of these quantities will be listed along with their respective references. Regarding the proton-OMP, the default option used in Talys is the phenomenological param- eterisation of Koning and Delaroche [63]. In addition to the default p-OMP option, a so-called “jlm-type" potential (based on the work of Jeukenne, Lejeunne, and Mahaux [64, 65, 66, 67] with later modifications by Bauge et al. [68, 69]) is utilized. The various models of the NLD anf 𝛟 SF used in Talys were discussed in Sec. 2.4.2.3 and 2.4.2.4 and are listed in Tables 2.1 and 2.2. Table 2.1 The available models for the nuclear level density in Talys [62]. Talys Keyword Model ldmodel 1 ldmodel 2 ldmodel 3 ldmodel 4 ldmodel 5 ldmodel 6 Constant Temperature & Fermi Gas Model Back-shifted Fermi Gas Model Generalized Superfluid Model Skyrme-Hartree-Fock-Bogolyubov level densities from nu- merical tables Skyrme-Hartree-Fock-Bogolyubov densities from numerical tables Temperature-dependent Gogny-Hartree-Fock-Bogolyubov combinatorial level densities from numerical tables combinatorial level Ref. [31] [25] [32, 33] [36] [37] [38] Finally, the width fluctuation correction (WFC) from Eq. 2.12 takes into account that there are correlations between the incident and outgoing wave functions. By default, Talys applies a WFC using the formalism of Moldauer (so-called “widthmode 1") [70, 71]. A much stronger WFC is obtained for the approach of Hofmann, Richert, Tepel, and WeidenmÃŒller (HRTW ap- proach, “widthmode 2") [72, 73, 74], leading to significantly lower calculated (𝑝, 𝛟) cross sections, especially at low energies. 2.5 Nuclear Reactions in Stars In the previous sections, we introduced the concept of nuclei and nuclear reactions, describing the probabilities and mechanisms through which these reactions occur. As nuclear reactions can transform nuclei while releasing energy, they play a crucial role in understanding both the production of energy and the nucleosynthesis of elements in stars. While the various stellar environments and 17 Table 2.2 The available models for the 𝛟-ray strength function in Talys [62]. Talys Keyword Transition Model strength 1 strength 2 strength 3 strength 4 strength 5 strength 6 strength 7 strength 8 strength 9 strengthM1 1 strengthM1 2 strengthM1 3 upbend y/n 𝐞1 𝐞1 𝐞1 𝐞1 𝐞1 𝐞1 𝐞1 𝐞1 𝐞1 𝑀1 𝑀1 𝑀1 𝑀1 Skyrme-Hartree- Kopecky-Uhl Generalized Lorentzian Brink-Axel Standard Lorentzian Skyrme-Hartree-Fock BCS model with QRPA Skyrme-Hartree-Fock-Bogoliubov model with QRPA Hybrid model (Lorentzian model with energy and temperature dependent width) Temperature-dependent Fock-Bogoliubov model with QRPA Temperature-dependent Relativistic Mean Field Model Gogny-Hartree-Fock-Bogoliubov model with QRPA by based on the D1M version of the Gogny force Simplified Modified Lorentzian Model Standard Lorentzian Model as parameterized in RIPL3 Library 𝑀1 normalized on 𝐞1 as 𝑓𝐞1/(0.0588𝐎0.878) Addition of spin-flip and scissors mode Flag to include upbend or not Ref. [45] [75, 76] [46] [48] [77] [48] [53] [49] [78] [79] [41, 80, 42] [55, 54, 59, 61] astrophysical processes will be explored in detail in Ch. 3, it is helpful to introduce the key concepts needed to bridge the theory of nuclear physics discussed earlier with stellar nucleosynthesis. The energy dependence of the cross section 𝜎(𝐞), can be interpreted as velocity dependence 𝜎(𝑣), where 𝑣 represents the relative velocity between the projectile and target nucleus. Instead of projectile beam and stationary target we can consider nuclear species being part of a stellar gas, where the kinetic energy available for the reaction comes from thermal movement. The reactions initiated by such motion are called thermonuclear reactions [6]. In a stellar gas that consists of 𝑁𝑎 nuclei per cubic centimeter of species 𝑎, and 𝑁𝑋 nuclei per cubic centimeter of species 𝑋 the reaction rate 𝑟 between species 𝑎 and 𝑋 is given by: 𝑟 = 𝑁𝑋 𝑁𝑎 𝑣 𝜎(𝑣) (2.21) where 𝑟 is in reactions per cubic centimeter per second. The velocities of gas particles vary over a wide range of values, described by a probability distribution 𝜙(𝑣) that is normalized to unity, 18 𝜙(𝑣)𝑑𝑣 = 1. Averaging the product 𝑣𝜎(𝑣) over this distribution gives the reaction rate per ∫ ∞ 0 particle pair: ⟚𝜎𝑣⟩ = ∫ ∞ 0 𝜙(𝑣)𝑣𝜎(𝑣)𝑑𝑢 The total reaction rate 𝑟 then becomes: 𝑟 = 𝑁𝑋 𝑁𝑎 ⟚𝜎𝑣⟩ (2.22) (2.23) Stellar matter is normally non-degenerate, and nuclei move non-relativistically. Therefore, in most cases, the velocities of nuclei can be described by the Maxwell-Boltzmann velocity distribu- tion: or in terms of energy 𝜙(𝑣) = 4𝜋𝑣2 (cid:16) 𝑚 2𝜋𝑘𝑇 (cid:17) 3/2 exp (cid:19) (cid:18) − 𝑚𝑣2 2𝑘𝑇 𝜙(𝐞) ∝ 𝐞 exp (−𝐞/𝑘𝑇) (2.24) (2.25) where 𝑇 refers to the temperature of the gas, 𝑚 the mass of the nucleus of interest, and 𝑘 the Boltzmann constant. As shown in Fig. 2.5 at low energies the function increases almost linearly Figure 2.5 The Maxwell-Boltzmann energy distribution of a gas at temperature 𝑇. with 𝑇 until it reaches its maximum value at 𝐞 = 𝑘𝑇. At higher energies, the function decreases exponentially [15]. 19 In a stellar gas, the velocities of both species 𝑎 and 𝑋 follow the Maxwell-Boltzmann distribution. By combining Eq. 2.22 and 2.24 we obtain: ⟚𝜎𝑣⟩ = (cid:19) 1/2 (cid:18) 8 𝜋𝜇 1 (𝑘𝑇)3/2 ∫ ∞ 0 𝜎(𝐞)𝐞 exp (cid:19) (cid:18) − 𝐞 𝑘𝑇 𝑑𝐞 (2.26) where 𝐞 is the center-of-mass energy and 𝜇 the reduced mass [15]. 2.5.1 Reactions at elevated temperatures In stellar plasma at elevated temperatures, nuclei are thermally excited, so a significant fraction of reacting nuclei will not be in their ground state. The fraction of nuclei in an excited state 𝜇 is given by the Boltzmann distribution: 𝑁𝜇 𝑁 = (2𝐜𝜇 + 1)𝑒−𝐞 𝜇/𝑘𝑇 (cid:205)𝜇 (2𝐜𝜇 + 1)𝑒−𝐞 𝜇/𝑘𝑇 = (2𝐜𝜇 + 1)𝑒−𝐞 𝜇/𝑘𝑇 𝐺 (2.27) where 𝐺 is the partition function that reflects the Boltzmann factor, and 𝐜𝜇 and 𝐞𝜇 the spin and excitation energy of the 𝜇 state, respectively. The ratio of the reaction rate involving thermally excited nuclei, ⟚𝜎𝑣⟩∗, to the reaction rate involving only the ground state ⟚𝜎𝑣⟩, is known as the stellar enhancement factor (SEF): SEF ≡ ⟚𝜎𝑣⟩∗ ⟚𝜎𝑣⟩ = (cid:205)𝜇 (2𝐜𝜇 + 1)𝑒−𝐞 𝜇/𝑘𝑇 (cid:205)𝜈 ⟚𝜎𝑣⟩ 𝜇→𝜈 𝐺 (cid:205)𝜈 ⟚𝜎𝑣⟩g.s.→𝜈 (2.28) Here, the summations over 𝜇 and 𝜈 include all the possible excited states of the target nucleus and all available states of the final nucleus, respectively [15]. 2.5.2 Inverse Reactions At low stellar temperatures, a nuclear reaction requires a positive 𝑄 value to proceed. However, as the temperature increases, the number of particles with energy exceeding the 𝑄 value also increases, allowing the inverse process to become energetically possible. Thus, when calculating the total reaction rate 𝑟, contributions from both reactions 𝑎 + 𝑋 ↔ 𝑏 + 𝑌 should be considered: 𝑟 = 𝑟𝑎𝑋 − 𝑟𝑏𝑌 = 𝑁𝑋 𝑁𝑎 1 + 𝛿𝑎𝑋 ⟚𝜎𝑣⟩𝑎𝑋 − 𝑁𝑌 𝑁𝑎𝑏 1 + 𝛿𝑏𝑌 ⟚𝜎𝑣⟩𝑏𝑌 (2.29) where 𝛿𝑖 𝑗 is the Kronecker delta, and the term (1 + 𝛿𝑖 𝑗 ) accounts for identical particles. 20 If the reaction rate ⟚𝜎𝑣⟩𝑎𝑋 is known, the reaction rate for the inverse reaction ⟚𝜎𝑣⟩𝑏𝑌 can be calculated using the reciprocity theorem, which relies on the invariance of the strong and electromagnetic interactions under time-reversal symmetry, meaning they are independent of the direction of time. As long as the cross sections depends on these two interactions, the ratio of the two cross sections can be written as: 𝜎𝑎𝑋 𝜎𝑏𝑌 = 𝑚𝑏𝑚𝑌 𝐞𝑏𝑌 (2𝐜𝑏 + 1) (2𝐜𝑌 + 1) (1 + 𝛿𝑎𝑋) 𝑚𝑎𝑚 𝑋 𝐞𝑎𝑋 (2𝐜𝑎 + 1) (2𝐜𝑋 + 1) (1 + 𝛿𝑏𝑌 ) Then, by using Eq. 2.26 and the relation 𝐞𝑏𝑌 = 𝐞𝑎𝑋 + 𝑄 (𝑄 > 0), we obtain: ⟚𝜎𝑏𝑌 ⟩ ⟚𝜎𝑎𝑋⟩ = (2𝐜𝑎 + 1)(2𝐜𝑋 + 1) (1 + 𝛿𝑏𝑌 ) (2𝐜𝑏 + 1)(2𝐜𝑌 + 1) (1 + 𝛿𝑎𝑋) (cid:18) 𝜇𝑎𝑋 𝜇𝑏𝑌 (cid:19) 3/2 exp (cid:19) (cid:18) − 𝑄 𝑘𝑇 which, replacing back in Eq. 2.29 leads to: (cid:34) 𝑟 = ⟚𝜎𝑎𝑋⟩ 1 + 𝛿𝑎𝑋 𝑁𝑎 𝑁𝑋 − 𝑁𝑏 𝑁𝑌 (2𝐜𝑎 + 1) (2𝐜𝑋 + 1) (2𝐜𝑏 + 1) (2𝐜𝑌 + 1) (cid:18) 𝜇𝑎𝑋 𝜇𝑏𝑌 (cid:19) 3/2 exp (cid:19)(cid:35) (cid:18) − 𝑄 𝑘𝑇 (2.30) (2.31) (2.32) It is important to note that Eq. 2.32 refers to ground state contributions. For an accurate astrophysical calculation the stellar enhancement factor described in Sec. 2.5.1 needs to be accounted for [15]. 2.5.3 Neutron-Induced Reactions Neutrons play an important role in stellar nucleosynthesis, however due to their short lifetime of about 10 minutes, they only exist in stellar environments in which they can be produced. Some important neutron producing reactions in stars are the 13C(𝛌,n)16O, 18O(𝛌,n)21Ne, and 22Ne(𝛌,n)25Mg. Neutrons produced in stars are very quickly thermalized due to elastic scattering, so their velocities are described by the Maxwell-Boltzmann distribution. Being electrically neutral, neutrons do not experience the Coulomb barrier of nuclei. Additionally, for angular momentum 𝑙 = 0 (s-wave), they do not encounter a centrifugal barrier, meaning their penetrability depends solely on their velocity. As a result, neutron capture is mainly proceeding with s-wave neutrons, and the cross section is approximated by the 1/𝑣 law: 𝜎𝑛 (𝐞𝑛) ∝ 1 𝑣𝑛 (2.33) It follows that, in the absence of resonances, the reaction rate per particle pair ⟚𝜎𝑣⟩ is approximately constant [15]. 21 Figure 2.6 Cross section for thermal s-wave neutrons follows the 1/𝑣 law. 2.5.4 Charged-Particle-Induced Reactions As will be discussed in more detail in Chapter 3, stars consist mainly of hydrogen and helium. Therefore proton and 𝛌 (4𝐻𝑒) capture reactions are some of the most common and important reactions that can happen in a star. Unlike the case of neutrons, charged particles need to overcome the repulsive Coulomb barrier of nuclei to be captured. The Coulomb barrier is of the form: 𝑉𝐶 = 𝑍1𝑍2𝑒2 𝑟 (2.34) where 𝑍𝑖 the atomic number, 𝑒 the electron charge and 𝑟 the distance between nuclei. It is apparent that the closer to the nucleus, the larger the Coulomb barrier encountered. Combined with the attractive potential caused by the strong nuclear force leads to the effective potential shown in Fig. 2.7. Classically, a charged particle reaction can occur only when the projectile energy is sufficient to overcome the Coulomb barrier. However, as will be discussed in Ch. 3, the temperatures reached in stellar environments during the majority of a star’s lifetime are not sufficient to thermally excite nuclei to such energies. Fortunately, quantum mechanics provides a solution: charged particles can penetrate the Coulomb barrier through quantum tunneling, allowing stellar nucleosynthesis to begin at lower temperatures, an idea first proposed in 1929 [81]. 22 Figure 2.7 Schematic of the combined Coulomb and nuclear potential. The incident projectile needs to penetrate the Coulomb barrier to be captured in the nucleus. Classically, the nearest allowed distance it would reach is the turning point. Figure recreated based on [15]. The probability for tunneling through the Coulomb potential can be found by solving the Schrödinger equation as 𝑃 = 𝑒−2𝜋𝜂 (2.35) where 𝜂 = 𝑍1𝑍2𝑒2 ℏ𝑣 2𝜋𝜂 = 31.29𝑍1𝑍2 is the Sommerfeld parameter. The exponent can be numerically calculated as √𝜇/𝐞. The cross section, being proportional to exp(−2𝜋𝜂), decreases sharply below the Coulomb barrier. To make experiments feasible, cross section measurements are typically performed at higher energies and then extrapolated to the lower energies relevant to astrophysical temperatures. However, the rapid decline of the cross section below the Coulomb barrier signifi- cantly limits the accuracy of this extrapolation. To address this issue, it is more useful to express the cross section in terms of the astrophysical S-factor, 𝑆(𝐞), defined as: 𝜎(𝐞) = 1 𝐞 exp(−2𝜋𝜂)𝑆(𝐞) (2.36) 23 The S-factor has much smoother variations in the non-resonant region, and contains only the nuclear information, eliminating the rapidly decreasing energy-dependent factors. Calculating the charged-particle reaction rate involves combining the probability of a particle being in a specific energy, meaning the Maxwell-Boltzmann rate from Eq. 2.26, with the probability of tunneling through the Coulomb barrier. These probabilities overlap within a narrow energy range, 𝐞0 ± Δ𝐞0/2, where the S-factor, 𝑆(𝐞0), can be considered approximately constant. Therefore, substituting Eq. 2.36 into Eq. 2.26 yields: ⟚𝜎𝑣⟩ = (cid:19) 1/2 (cid:18) 8 𝜋𝜇 1 (𝑘𝑇)3/2 𝑆(𝐞0) ∫ ∞ 0 exp (cid:18) − 𝐞 𝑘𝑇 − (cid:19) 𝑏 √ 𝐞 𝑑𝐞 (2.37) where 𝑏 = (2𝜇)1/2𝜋𝑒2𝑍1𝑍2/ℏ = 0.989𝑍1𝑍2𝜇1/2 MeV1/2. The quantity 𝑏2 is called the Gamow energy. The peak shown in Fig. 2.8 is formed by the overlap of the two probability distributions Figure 2.8 The convolution of the Maxwell-Boltzmann energy distribution and the quantum me- chanical tunneling function through the Coulomb barrier produce the Gamow peak at energy 𝐞0 ± Δ𝐞0/2. Figure from [82]. at energy 𝐞0, and is called the Gamow peak. It represents the energy range in which the reaction can happen in a star [15]. The width of the Gamow peak can be approximated by Δ = 4 √ 3 √𝐞0𝑘𝑇 = 0.2368 (cid:16) 0 𝑍 2 𝑍 2 1 𝜇𝑇 5 9 (cid:17) 1/6 (MeV) (2.38) 24 where 𝜇 is the reduced mass, 𝑍𝑖 the atomic number and 𝑇9 the stellar temperature in GK [6]. It should be noted that the resonances introduced in Sec. 2.4.2.1 play a crucial role in charged- particle reactions within stellar environments. When a resonance lies within the Gamow peak, it can contribute significantly to the total reaction rate. As will be discussed in the next chapter, the existence of such narrow resonances within the Gamow window enables the nucleosynthesis of elements vital to life, such as carbon. 25 CHAPTER 3 ASTROPHYSICS In the previous chapter, the fundamental concepts of nuclei, nuclear reactions, and the mechanisms by which these reactions occur in stellar environments were introduced. This chapter begins with a discussion of the observed element abundances, as any description of nucleosynthesis processes must ultimately account for these observations. The evolution of stars is then discussed to provide the context for the environments where nucleosynthesis takes place. The various nucleosynthesis processes are then presented, leading to the main topic of this thesis, the astrophysical 𝛟 process. 3.1 Abundances Understanding the origin of the elements in the universe involves explaining and reproducing their abundances, meaning the relative amounts of nuclides. Astronomical observations provide spectra that identify elements found in the interstellar medium and on the surface of stars. Pre-solar grains that were ejected in stellar winds, can become embedded in meteorites in our solar system. Many of those grains are carbonaceous materials such as diamond, silicon carbide (SiC), and graphite, and can offer isotopic ratios of their original environment. By analyzing the patterns of the elemental abundances, nuclear astrophysics can infer the mechanisms responsible for producing each element. Among the elemental abundances across all stars, the solar abundance pattern is the most extensively studied. The solar system formed from a uniform gaseous nebula that contained contributions from many generations of stars and explosive events. Today, 280 naturally occurring isotopes remain in 83 elements [14]. The solar system’s elemental abundances are shown in Fig. 3.1 normalized to silicon atoms. The most abundant elements, hydrogen and helium, make up approximately 98% of the solar mass. About 1.5% of the mass is carbon and oxygen, while all other elements account for the remaining 0.5%. A significant drop is observed near the very weakly bound elements lithium, beryllium and boron (𝐎 = 5 − 8), known as the mass gaps, and a peak forms in the region around iron. As discussed in Sec 2.1, this corresponds to the most tightly bound nuclides in the iron peak. The zig-zag structure reflects the differences in binding energies 26 Figure 3.1 Solar system abundances based on data from Asplund [83]. The data is normalized to 106 Si atoms. between nuclei with odd and even number of nucleons, due to pairing effects. 3.2 Stellar Evolution The life of a star begins as a collapsing molecular cloud primarily composed of hydrogen (H) and helium (He). As the cloud contracts, its temperature rises due to an increase in gravitational potential energy. This rise in temperature causes the pressure at the cloud’s center to increase, which counteracts further contraction. Smaller clouds may remain in this balanced state and never contract further, but larger ones continue collapsing until thermonuclear reactions ignite in the core, and hydrostatic equilibrium maintains stability. The different burning phases will be discussed in Sec. 3.3, but they generally follow a pattern. A primary nuclear “fuel" undergoes fusion in the core, releasing energy that temporarily halts gravitational contraction. As the fuel depletes, the burning region shifts outward, and the star transitions from core burning to shell burning. With insufficient energy production in the core to counteract gravity, contraction resumes, increasing the temperature until the next fuel ignites. The cycle repeats as long as fusion is possible [14]. The lifetime of a star varies vastly, from several billion years for smaller stars to just a few million years or less for the most massive ones. Small stars with masses below 0.08 𝑀⊙ (where 1 𝑀⊙ is the mass of the Sun) cannot ignite hydrogen and remain as brown dwarfs, supported by molecular gas pressure. Stars between 0.6–2.3 𝑀⊙, like the Sun, spend billions of years burning 27 hydrogen. Once the hydrogen in the core is depleted, these stars expand into red giants, burning hydrogen in a shell. Helium ignition leads to a core He-flash and He-shell burning during the asymptotic giant branch (AGB) phase. Unable to ignite carbon, they eventually shed their outer layers, leaving behind a white dwarf. Stars up to 8 𝑀⊙ evolve further, igniting carbon in their C-O cores. These stars are very luminous and lose most of their mass during the AGB phase through stellar winds, ending their lives as white dwarfs. Such AGB stars, as will be discussed in the next section, are responsible for synthesizing almost half of the heavy elements [6, 14]. The evolution of stars with masses larger than 8 𝑀⊙ is fundamentally different and much more spectacular than the previous cases. These massive stars have significantly shorter lifetimes, lasting only a few million years. However, within this relatively brief period, stars more massive than 12𝑀⊙ undergo all burning phases, successively fusing hydrogen, helium, carbon, neon, oxygen, and finally silicon. Each phase becomes progressively shorter, with silicon burning lasting only about a day. By the time silicon is exhausted in the core, the star has developed an “onion-like" structure, with layers of elements separated by thin nuclear burning shells, as illustrated in Fig. 3.2. Stars more massive than 25 𝑀⊙, or those that rotate rapidly, lose a significant fraction of their mass through strong stellar winds and eruptions, resulting in the loss of most of their outer envelopes. At this stage, the core primarily consists of iron-peak nuclei, which, as discussed in the previous chapter, have the highest binding energy and fusion is no longer energetically favorable. Without a nuclear source to counteract gravity, the core continues shrinking. Once the core’s mass exceeds the Chandrasekhar limit of 1.4 𝑀⊙, it collapses [6, 14]. 3.2.1 Supernovae Supernovae are among the brightest, most complex and cataclysmic events in the universe. They are mainly classified based on the type of light emitted and their detailed mechanisms remain an active area of research. The endpoint of the evolution of a red supergiant described above, is a core-collapse supernova (CCSN) known as Type II (SNII), and is characterized by strong presence of hydrogen in the emitted light spectra (light curves). As the core reaches 𝑇 > 1010 K, photons break apart 56Fe into 28 Figure 3.2 Schematic drawing of the onion shell structure of massive stars of 12 ≲ 𝑀 ≲ 25𝑀⊙ at the end of their evolution, with the dominant elements indicated. The outer envelopes of stars of 𝑀 > 25𝑀⊙ are stripped before they explode as supernovae. Note that the figure is not to scale. 𝛌 particles and neutrons, an endothermic reaction that accelerates the collapse. Within a fraction of a second, a core with size of several thousand kilometers collapses to tens of kilometers radius. Nuclear force at very small distances becomes repulsive supporting the core, along with rising density that enables electron capture, forming a degenerate neutron gas that can support very high pressure. If the collapsing core’s mass is below ∌2 𝑀⊙ (the Oppenheimer-Volkoff limit [84]), the pressure halts the collapse, forming a neutron star. If not, then it continues until a black hole forms [85]. During the collapse, several phenomena occur, leading to the ejection of the star’s outer envelopes into the interstellar medium at extreme velocities. The infalling matter encounters the very compact proto-neutron star and experiences an intense shock, that pushes matter outwards as it bounces back. The outward moving shock wave compresses and heats the outer layers for short 29 periods of time, giving grounds for explosive nucleosynthesis. The temperature in the collapsing core is so high that photons can be captured by electrons, creating neutrinos. The neutrinos either escape or get trapped due to the high density and are captured by the infalling layers further increasing the temperature. The strong neutrino and antineutrino fluxes drive a continuous flow of protons and neutrons, known as neutrino-driven wind, that enables further nucleosynthesis to take place [6]. Another important category is Type Ia (SNIa), that has no hydrogen in the light spectra, but a lot of silicon. These are created in binary systems, where a white dwarf is accreting mass from a companion star. Once the white dwarf’s mass approaches the Chandrasekhar limit, carbon ignites under degenerate conditions, increasing temperature while supporting very high pressure. Once the degeneracy is lifted, the energy generation rate is so large that an explosion occurs. In the single- degenerate (SD) scenario the companion star is a red giant, supplying hydrogen and helium-rich material onto the primary star, and after the explosion a remnant of the secondary star remains. In the double-degenerate (DD) scenario, both stars are white dwarfs that eventually merge, and the collapse leaves no remnant behind [6, 85]. Other types include SN Ib and Ic, that are most likely caused by supermassive stars that have lost almost all their envelope, called Wolf-Rayet stars, and are characterized by absence of hydrogen and silicon in their light spectra [85]. 3.3 Stellar Nucleosynthesis The previous section described the evolution of stars starting from clouds of dust. In the early stages of the universe, these molecular clouds consisted of light elements formed during the Big Bang. A few seconds after the Big Bang, the universe had cooled down sufficiently for free protons and neutrons to form. About 20 minutes later, primordial nucleosynthesis established the abundances of light elements, consisting of 75% 1H, 25% 4He, and traces of the stable deuterium (2H), 3He, and 7Li, as well as unstable tritium (3H) and 7Be that decayed shortly after [86]. Any elements heavier than A=7 were not formed until hundreds of millions of years later, when the first nuclear reactions started taking place in stellar cores. As it was pointed out in the previous 30 section, nuclear reactions allow the star to maintain hydrostatic equilibrium, preventing it from gravitational collapse. Initially, the fuel consists of nuclei with the smallest nuclear charges, as their low Coulomb barrier allows fusion to occur at lower temperatures. Once this fuel is consumed, the core contracts, increasing the temperature and enabling fusion of nuclei with progressively higher Coulomb barriers. 3.3.1 Hydrostatic Burning The first nuclear fuel is the lightest and most abundant hydrogen. H-burning occurs through the net reaction: 4𝑝 → 4𝐻𝑒 + 2𝑒+ + 2𝜈𝑒 (3.1) with a Q-value of 26.731 MeV. The two positrons are immediately annihilated with free electrons in the stellar plasma. However, the probability of four protons interacting simultaneously is too small for this reaction to occur directly. Instead, the net reaction is achieved with sequences of two-particle interactions. There are two main mechanisms for H-burning, the proton-proton (or pp) chains and the CNO-cycles. Which mechanism dominates depends on the core temperature and the availability of CNO-cycle nuclei, which act as catalysts, converting 1H into 4He. The pp chains are dominant in first-generation stars, which formed from primordial material, or in less massive stars with lower central core temperatures. In contrast, the CNO cycle becomes significant in more massive, and later-generation stars enriched with heavier elements synthesized in massive first-generation stars that have already exploded [6, 14, 15]. After the hydrogen fuel has been consumed, the stellar core consists mainly of 4He, as the creation of heavier elements is blocked by instabilities at the mass gaps 𝐎 = 5 and 8. However, 12C, the third most abundant element in the universe, is not a product of primordial nucleosynthesis and must therefore be synthesized in stars. This problem was solved by Öpik and Salpeter [87, 88, 89] who proposed that 12C is produced through a two-step process known as the triple-alpha process. In this process, two alpha particles fuse to form an unstable 8Be nucleus, which can occasionally capture another alpha particle before decaying, resulting in 12C. As the non-resonant tunneling probability for this reaction is too low to explain the observed abundance of 12C in the universe, 31 Fred Hoyle [90] proposed that the reaction proceeds via a resonance in 12C just above the 8Be+𝛌 threshold, at 𝐞 𝑋 ≃ 7.68 MeV. This resonance, now known as the Hoyle state, has been subject of extensive research since its discovery and continues to be studied today [91, 92, 93, 94, 95, 96]. After the formation of 12C, further reactions like 12C(𝛌, 𝛟)16O occur, but the subsequent 16O(𝛌, 𝛟)20Ne proceeds at an extremely low rate, blocking significant nucleosynthesis via He- burning beyond 16O. As carbon-based life forms that depend on the existence of oxygen in the atmosphere, we should appreciate that it is only through some fortunate nuclear properties of carbon and oxygen that they are produced so plentifully and survive the red giant phase of stars. With the exhaustion of helium in the core, the star transitions to helium-shell burning. As the density increases, thermal pulses, known as helium-shell flashes, can occur. Similar to the He-core flash, these thermal pulses cause mixing of the material between the He-burning and H-burning shells. These thermal pulses combined with a complicated convective mixing process in the inter-shell regions combines the p-rich material of the H-burning shell with 12C from the He-burning shell, forming 13C through the 12C(𝑝, 𝛟)13N(𝛜−)13C reaction sequence [97]. The reaction 13C(𝛌,n)16O is an important neutron source for heavy element nucleosynthesis, as will be discussed in Sec. 3.3.3. Another significant source of neutrons for heavy element nucleosynthesis is the 22Ne(𝛌,n)25Mg reaction, making the availability of 22Ne very important during He-burning. The exhaustion of He in the center of the star, leaves a core rich in C and O contracting under gravity. As discussed in Sec. 3.2, stars with mass above 8 𝑀⊙ will proceed to more advanced burning stages, while only stars above 12 𝑀⊙ undergo all burning stages. Carbon burning ignites first in the CO-rich core, followed by neon burning and then oxygen burning. While carbon and oxygen burning proceed mostly through fusion, Ne-burning mostly proceeds through photodisintegration reactions on 20Ne, making this phase particularly brief due to its lower energy output. Similarly, the final burning phase, Si-burning, is based on photodisintegrations of silicon isotopes rather than fusion, leading to a complex network of reactions. In this network, many forward and reverse reaction rates become comparable to the burning timescales, leading to the formation of certain “clusters" in the nuclear chart where particle captures and photodisintegrations are in equilibrium, 32 a state know as a quasi-statistical nuclear equilibrium (QSE). The nuclear flow is therefore largely confined within these clusters. The final abundances produced during Si-burning depend on where those clusters form, which depends on the available neutrons in the system (also known as neutron excess 𝜂, neutron to proton ratio, or electron fraction 𝑌𝑒). This dependency on 𝑌𝑒 determines the composition of the core and sets the stage for the subsequent core-collapse [6, 14, 15]. 3.3.2 Nucleosynthesis During Core-Collapse SN The collapse of a stellar core discussed in Sec. 3.2, enables two main mechanisms for nucle- osynthesis. The first is driven by the outward-moving shock wave during the explosion, and the second by the large amount of energy released from the core in the form on neutrinos. As the shock wave moves outward, it heats and compresses the star’s layers almost instanta- neously. First, it passes through the silicon-burning shell, followed by the oxygen-rich layer, and finally the region primarily composed of neon, carbon, and oxygen. Each layer undergoes a specific explosive burning process at varying peak temperatures, producing a range of nuclei. The resulting abundances depend strongly on the expansion timescale during cooling and the availability of free particles (𝛌, 𝑝, 𝑛) [6]. The launch of the shock generates a strong flux of electron neutrinos and antineutrinos, which drive protons and neutrons from the region near the proto-neutron star in what is known as the neutrino-driven wind [98]. Neutrinos interact with nuclei from infalling layers, populating excited nuclear levels that decay via particle emission (𝑝, 𝑛, 𝛌). This neutrino-driven nucleosynthesis, known as the 𝜈 process, depends on the wind’s properties, such as electron fraction 𝑌𝑒 and entropy (or photon-to-baryon ratio), determining whether the wind is proton-rich or neutron-rich. Final abundances also depend on the expansion and cooling timescales, as in hotter environments the wind will consist mainly of neutrons and protons, but in cooler environments protons and neutrons will combine to 𝛌 particles. 3.3.3 Nucleosynthesis Beyond Iron As discussed in previous sections, hydrostatic burning stages synthesize elements through fusion up to the iron region. At very high temperatures, where charged-particle capture reactions 33 are enabled, nucleosynthesis occurs in clusters where nuclear statistical equilibrium is established, favoring either iron peak nuclei or lighter elements. However, during hydrostatic burning, neutrons are produced through the 13C(𝛌, 𝑛)16O and 22Ne(𝛌, 𝑛)25Mg reactions. As neutrons are not affected by the increasingly large Coulomb barrier of heavy nuclei, neutron capture reactions are able to synthesize elements heavier than iron. In their pioneering work in 1957, Burbidge, Burbidge, Fowler, and Hoyle (henceforth B2FH) described two main neutron-capture processes for heavy-element nucleosynthesis [3]. The dis- tinction between the two lies on the vastly different timescales on which they operate. These processes are known as the slow and rapid neutron-capture process, or s- and r-process for short. These two processes are responsible for synthesizing the majority of heavy elements. However, it is now known that additional neutron-capture processes such as the intermediate (i-) [99, 100] and n-process [101], are also required to accurately reproduce the observed abundances of stars. Additional processes are also required for the production of the proton-rich elements, and those will be discussed in Sec. 3.4. One of the most compelling pieces of evidence for the two neutron-capture mechanisms is their ability to naturally explain the existence of peaks in the solar system abundance pattern of heavy elements. As shown in Fig. 3.3, double peaks appear in the regions of 𝐎 ≃ 84, 138 and 208. These patterns arise from the neutron magic numbers 𝑁 = 50, 82 and 126. The sharp peaks correspond to abundances formed by the s-process, whereas the broader peaks about 10 mass units below reflect r-process enhanced abundances. The s- and r-processes are discussed in more detail in the following sections, along with the i- and n-process. 3.3.3.1 The s Process First evidence of the slow neutron-capture process (s-process) nucleosynthesis was found in spectra of AGB stars, where radioactive Tc was observed [102]. The s-process involves a series of neutron-capture reactions followed by 𝛜− decays and is responsible for the production of almost half of the isotopes of heavy elements. The name “slow" reflects the time intervals between successive captures that are inversely proportional to the neutron capture reaction rates and the 34 Figure 3.3 Decomposition of solar s– (solid line), r– (black circles) and p–abundances (white squares) relative to silicon. Figure from M. Arnould et al., Physics Reports 450 (2007), with permission from Elsevier. neutron flux. In environments of typical neutron density ≈ 107−11 neutrons/cm3, the rate for n- captures is comparable to that of the 𝛜− decay, and therefore the synthesis path follows closely the valley of stability in the nuclear chart. The process terminates in Bi, as any heavier elements are unstable and decay by 𝛌 emission back to stability. An example of the s-process path is shown in Fig. 3.4, starting from 77Se. The path can be calculated by comparing the decay rate 𝜆 = ln 2/𝑡1/2, with the neutron-capture reaction rate. If the two are comparable, such as in 85Kr, a branching point occurs, where the path can follow both directions, leading to different abundance patterns. There are about 15-20 significant branching points along the s-process path. Branching points can provide information on the detailed conditions of the stellar environment, such as neutron density and temperature. However, achieving this requires accurate knowledge of neutron-capture reaction cross sections, decay half-lives and any temperature dependence of the rates [6, 14]. The example path of Fig. 3.4 passes through the neutron magic number 𝑁 = 50. This configu- ration is energetically more favorable than 𝑁 = 51, and therefore the neutron-capture cross section 35 Figure 3.4 An example of the s-process path in the region around 𝐎 = 85. The halflives are obtained from [4]. The gray and white squares correspond to stable and radioactive isotopes, respectively, and the circle indicates a branching point. on nuclei with 𝑁 = 50 will drop significantly, blocking the s-process path from more n-rich nuclei, and pushing the reaction flow to higher elemental chains. This results in the first peak in the solar abundance pattern from Fig. 3.3. The s process is a secondary process, as it requires the existence of iron-peak nuclei to act as seeds, unlike the hydrostatic burning phases that are primary processes, and do not depend on preexisting nuclei. As a secondary process, the produced abundances can vary significantly based on the stellar conditions. As discussed in Sec. 3.3.1, during He-burning in AGB stars, neutrons are produced by the reactions 13C(𝛌,n)16O and 22Ne(𝛌,n)25Mg. A significant amount of 13C can be produced in thermally pulsing, low mass (1.5-3 𝑀⊙) AGB stars. A complex mixing of the intershell, which is the region between the He- and H-burning shells, mixes protons with material 36 rich in 12C and 4He. This enables the sequence 12C(𝑝, 𝛟)13N(𝛜+𝜈)13C, forming a region known as the 13C pocket [97]. At temperatures near 𝑇 ≈ 0.09 GK, a neutron density in the order of ≈ 107 n/cm3 produced by the 13C(𝛌,n)16O reaction, provides fuel for s-process nucleosynthesis for a period of ≈ 20 000 years. This is known as the main s-process component and provides about 95% of the total neutron exposure, synthesizing elements up to Pb. The remaining neutron exposure is achieved with the 22Ne(𝛌,n)25Mg source. This so-called weak s-process component, is achieved in massive stars at the end of the convective He-burning core and in the C-burning shell [103]. 3.3.3.2 The r Process The presence of abundance peaks that can’t be explained by the s process, along with the existence of long-lived isotopes heavier than Bi, such as 232Th and 238U, highlights the need for an additional neutron-capture process beyond the s process. The environment for such a process can be found in extreme stellar environments, where neutron fluxes are so high (≈ 1020−22 n/cm3) that the 𝛜-decay rate of unstable nuclei is small compared the rate of neutron capture. In this case, the nucleosynthesis path may move close to the neutron dripline. Only when the neutron flux terminates, the neutron-rich nuclei decay back to stability through 𝛜− decays. This nucleosynthesis mechanism is named the rapid (r-) neutron-capture process and is responsible for the synthesis of approximately the other half of the isotopes of heavy elements [3]. Similarly to the s process, the magic neutron numbers impact the reaction flow, leading to the creation of two peaks at mass numbers 𝐎 = 130 and 195, which are about 10 mass units below the s-process peaks near 𝐎 = 138 and 208, as shown in Fig. 3.3. At neutron densities in the order of 1020−22 n/cm3 neutron captures can drive the r-process flux close to magic neutron numbers. However, the neutron separation energy 𝑆𝑛 decreases for more neutron-rich nuclei. Therefore in each isotopic chain (𝑛, 𝛟) and (𝛟, 𝑛) reactions may eventually reach a quasi-statistical equilibrium (QSE), similar to the clusters formed during Si-burning. The flow toward higher elemental chains depends on the 𝛜-decay rates of the nuclei in QSE. As these rates are slow compared to the rates in equilibrium, waiting points may be established, usually one or two per chain in QSE. These waiting points can determine the timescale of the reaction network and influence the final abundance, as 37 𝛜− decays will follow the isobaric chain (𝐎 = const). As a consequence the r-process peaks are located in mass regions below the corresponding s-process peaks. The r-process network can extend to the neutron dripline, where most nuclear properties remain experimentally unknown, especially for nuclei far from stability. Extensive efforts are underway to both theoretically and experimentally determine key nuclear properties such as masses, level schemes, halflives, 𝛜-decay rates, fission rates, and neutron-capture cross sections [21, 104, 105, 106, 107]. Improving our understanding of these quantities is essential for enhancing the predictive power of r-process models, which are critical for explaining observed abundances in the Sun and other stars. The site of the r process has been one of the most significant open questions in the field for decades, further complicated by the lack of nuclear data for exotic isotopes. Over the years, numerous potential sites have been proposed, including neutrino-driven core collapse supernovae [108, 109], electron-capture supernovae [110], magneto-rotational supernovae leading to magnetars (i.e. neutron stars with very high magnetic fields) [111, 112, 113], collapsars (massive stars that collapse into a black hole) that produce powerful relativistic jets [114, 115], as well as black hole - neutron star mergers [116]. Compact binary mergers (NS-NS mergers) have been suggested as r-process sites since the 1970s [116] with first nucleosynthesis predictions in 1999 [117]. In August 2017 LIGO and Virgo detected gravitational waves from the NS-NS merger GW170817 [118], providing the first direct evidence of an r-process event. Since then, research on r-process nucleosynthesis in NS-NS mergers has been exponentially growing [119, 120, 121]. 3.3.3.3 Other Neutron-Capture Processes The solar abundances shown in Fig. 3.3 can be sufficiently explained by a combination of the s and r processes, but this is not the case for many other stars. The abundance patterns of a group of very old, carbon-enriched stars known as carbon-enhanced metal poor (CEMP) stars can instead be explained by an alternative neutron-capture mechanism, operating at intermediate neutron densities between those of the s and r process. This intermediate (i-) neutron capture process was first proposed by Cowan in 1977 [99], and was found to match the observed abundances of CEMP 38 stars in 2016 [100]. Since then, many studies have been dedicated to the i process and its potential sites, with candidates including rapidly rotating white dwarfs accreting material from a companion red giant [122], and thermal-pulsing AGB stars [123]. The mechanism is similar to that of the s and r process, but intermediate neutron fluxes of ≈ 1012−15 n/cm3 drive neutron-capture reactions a few steps away from stability before 𝛜 decays return the nuclear flow to stable species. This proximity to stability makes the i process particularly promising for experimental studies, as many of the relevant neutron-capture reactions are accessible with current facilities [124, 125]. Another neutron-capture process, the n process, has been proposed to occur in the He shell after its composition is modified by the SN shock passage. A neutron flux of ≈ 1018 n/cm3 or higher can then be produced by the 22Ne(𝛌, 𝑛)25Mg reaction. The n process is able to reproduce anomalous Mo isotopic abundances measured in SiC meteorites, motivating further study of this mechanism [101]. 3.4 Production of the p Nuclei As discussed in the previous section, neutron-capture processes dominate heavy element nu- cleosynthesis. However, these processes cannot produce all isotopes of heavy elements. In their pioneering work, B2FH identified 35 proton-rich nuclides that are shielded by the valley of stability and cannot receive contributions from the s or r process. These isotopes were named p nuclei, and the mechanism responsible for their synthesis, the p process. The 35 classical p nuclei as identified by B2FH are listed in Table. 3.1, along with their isotopic fractions within their respective elements and their solar abundances [126]. Subsequent research has shown that many of these p isotopes also receive contributions from neutron-capture [127, 129] or neutrino-driven processes [128, 130], meaning they cannot be strictly classified as p-only isotopes. Additionally, certain unstable nuclides, such as 92Nb, 97,98Tc and 146Sm, though not part of the original list of classical p nuclei, play a significant role in studies of the p process. These isotopes, with half-lives comparable to astronomical timescales, are believed to form in the same events as stable p nuclei and can be used as cosmochronometers, providing important information on the composition of the early Solar system [131, 132]. 39 Table 3.1 The classical p nuclei, their fraction (in %) of the isotopic composition of the elements and solar abundances (relative to Si=106). Data from Lodders [126]. Isotope Element (%) Solar Abundance Comment 74Se 78Kr 84Sr 92Mo 94Mo 96Ru 98Ru 102Pd 106Cd 108Cd 113In 112Sn 114Sn 115Sn 120Te 124Xe 126Xe 130Ba 132Ba 136Ce 138Ce 138La 144Sm 152Gd 156Dy 158Dy 162Er 164Er 168Yb 174Hf 180Ta 180W 184Os 190Pt 196Hg 0.889 0.362 0.555 14.8 9.25 5.54 1.87 1.02 1.25 0.89 4.29 0.971 0.659 0.339 0.096 0.129 0.112 0.106 0.101 0.186 0.251 0.0902 3.07 0.203 0.056 0.096 0.137 1.61 0.13 0.162 0.0123 0.12 0.0198 0.0136 0.153 5.80 · 10−1 2.00 · 10−1 1.31 · 10−1 3.86 · 10−1 2.41 · 10−1 1.05 · 10−1 3.55 · 10−2 1.46 · 10−2 1.98 · 10−2 1.41 · 10−2 7.80 · 10−3 3.62 · 10−2 2.46 · 10−2 1.26 · 10−2 4.60 · 10−3 6.94 · 10−3 6.02 · 10−3 4.60 · 10−3 4.40 · 10−3 2.17 · 10−3 2.93 · 10−3 3.97 · 10−4 7.81 · 10−3 6.70 · 10−4 2.16 · 10−4 3.71 · 10−4 3.50 · 10−4 4.11 · 10−3 3.23 · 10−4 2.75 · 10−4 2.58 · 10−6 1.53 · 10−4 1.33 · 10−4 1.85 · 10−4 6.30 · 10−4 r-process contribution [127] r-process contribution [127] 𝜈-process contribution [128] s-process contribution [129] s-process contribution [129] s-process [129] and 𝜈-process contributions [130] The p process was initially described in B2FH as occurring in the hydrogen-rich layers of core collapse supernovae, through a series of (𝑝, 𝛟) and (𝛟, 𝑛) reactions on existing s- and r-process seeds during the passage of the shock wave [3]. In 1978, Woosley and Howard [133] suggested 40 that the required conditions for the process, including high densities, elevated temperatures, and extended time scales, are unlikely to exist in the hydrogen-rich regions of most stars. Alternatively, they proposed an explosive nucleosynthesis mechanism based on a series of photodisintegration reactions on s- or r-process seed nuclei, which was named the 𝛟 process. The production of p nuclei remains an active area of research. Various mechanisms have been proposed in different astrophysical environments, involving both explosive and neutrino-driven nucleosynthesis. The term p process has been retained within the astrophysics community for historical reasons and now serves as an “umbrella" term that includes these diverse processes. The following provides an overview of the primary scenarios currently under investigation. 3.4.1 The 𝛟 Process The 𝛟 process is widely regarded as the main mechanism for the synthesis of the p nuclei. It occurs in stellar environments of sufficiently high plasma temperatures through particle emission from thermally excited nuclei. Rather than the hydrogen-rich layer proposed by B2FH, it is thought to occur in a zone where hydrogen is exhausted and heavy elements are subjected to a “hot photon bath" [133]. Under these conditions, the most likely reactions are photodisintegrations, meaning (𝛟, 𝑛), (𝛟, 𝑝) and (𝛟, 𝛌), as shown schematically in Fig. 3.5. The first reactions to take place are (𝛟, 𝑛), as these dominate the photodisintegration processes for most stable nuclei [134]. As the nuclear flow progresses to more neutron-deficient nuclei, the (𝛟, 𝑛) reaction rate decreases. At the same time, the proton-richer the isotope, the less energy is needed to remove a proton or an 𝛌 particle. Consequently, (𝛟, 𝑝) and (𝛟, 𝛌) reactions take over, moving the nuclear flow to lower elemental chains and eventually the p nucleus of interest. The process is highly sensitive to temperature, as higher temperatures or prolonged exposure would completely photodissociate the seeds into iron peak nuclei. On the other hand, cooler envi- ronments would not allow thermally excited nuclei to decay by particle emission. The temperature range for 𝛟-process nucleosynthesis to occur is between 1.8 and 3.2 GK. As these photodisintegra- tion reactions are strongly influenced by the particle separation energies of the seed nuclei, lighter p nuclei require higher plasma temperatures (𝑇 ≈ 3–3.2 GK) because their seeds, being closer to 41 Figure 3.5 An schematic illustration of the 𝛟-process path. The gray and white squares correspond to stable and radioactive isotopes, respectively, and the blue square indicates the produced p nucleus. the iron peak, are more tightly bound. The heavier p nuclei are synthesized in lower temperatures (𝑇 ≈ 1.8 − 2 GK) as their seeds are less bound. Any site capable of sustaining 𝛟-process nucleosynthesis must maintain these temperatures for a short period of time, while providing an adequate supply of seed nuclei. The most viable candidates are stellar explosions that involve a rapid expansion and subsequent cooling of the material. As a result, the 𝛟-process nucleosynthesis is highly sensitive to factors such as the temperature and density profile, expansion timescales, the initial abundances of seed nuclei, and the hydrodynamic properties of the explosion. The two main explosive environments where the 𝛟 process is thought to occur are the oxygen and neon enriched layers of a core-collapse supernovae (O/Ne SNII) [133, 135, 136], and thermonuclear Type Ia supernovae [137, 138, 139]. The mechanism of SNII, as discussed in Sec. 3.2, involves the propagation of an outward moving shock wave. There are two main components of the SNII that contribute to the 𝛟 process: the explosive component during shock wave propagation [133, 135, 136, 140], and the pre-explosive component [141, 136, 142, 143]. During the explosion, the shock encounters the O/Ne burning shell, which has been enriched in s-process material. The inner layers reach higher temperatures, enabling the synthesis of the lightest p nuclei, while the heavier ones are formed in the cooler outer 42 layers [135, 136, 134]. In the final stages of stellar evolution, just before the explosion, C-rich material may be ingested in the convective O-burning shell, forming a merged convective zone. This zone, known as the C-O shell merger [136, 142], provides a sustainable environment for synthesizing p nunclei heavier than Pd [143]. The produced material is mixed throughout the extended C-O shell and as it is not fully reprocessed by the shock wave, it maintains its pre-supernova abundances in the ejecta. It is important to note that the 𝛟-process abundances depend strongly on the s-process seed distributions. Early studies assumed solar s-process distributions [133, 136], however more recent studies have shown that these seeds may be enhanced. This can occur either through the presence of additional 13C from the convective C core, which leads to enhanced s-process seed distributions [144], or through stellar rotation, which enhances the weak s-process abundances driven by the 22Ne(𝛌,n)25Mg neutron source [145]. The other main environment for 𝛟-process nucleosynthesis is thermonuclear Type Ia supernovae, which was briefly discussed in Sec. 3.2. In the single-degenerate (SD) scenario, a CO white dwarf (WD) accretes material from a main-sequence or red giant companion. The WD explodes once its growing mass approaches the Chandrasekhar limit. During the explosion, a broad range of peak temperatures are reached across different mass coordinates of the WD, including temperatures sufficient to sustain 𝛟-process nucleosynthesis in the outer layers [138]. However, it is again essential to determine the available seed distributions in the exploding WD. These seeds can be provided by the s process during the AGB or TP-AGB phase [138], by the n process from recurring H-shell flashes [146], or by the i process from recurring He-shell flashes in the WD [147]. Although there have been efforts to explore the 𝛟 process in a sub-Chandrasekhar helium detonation model, there were significant uncertainties in the seed distributions, and sufficient p-nuclei abundances could only be achieved with highly enhanced seed abundances [148]. Additionally, ongoing research is exploring the potential role of the double-degenerate scenario in Type Ia SNe. The 𝛟 process is the most established scenario for the production of the p nuclei, as it has so far been the most successful at reproducing the majority of the observed solar abundances within 43 a factor of three. However, several discrepancies arise, especially in the region near 𝐎 = 95, as the 92,94Mo and 96,98Ru are systematically underproduced by one order of magnitude compared to other 𝛟-process nuclei [131, 134, 149]. An example overproduction factor divided by the average overproduction factor of all 35 p nuclei is shown in Fig. 3.6 by Roberti et al. [143], using a 20 M⊙ SNII model by Ritter et al. [142]. Figure 3.6 p-nuclide overproduction factors divided by their average, from a 20 M⊙ SNII model [142]. The different color symbols represent nuclei explicitly produced by the by Ritter et al. 𝛟 process (blue), and nuclei that may have an additional explosive contribution (orange) or an s, r process, or neutrino-capture contribution. (Figure from Roberti et al. Astronomy & Astrophysics 677, A22 (2023), under CC BY 4.0 (https://creativecommons.org/licenses/by/4.0)) 3.4.2 Other Scenarios for the p Process Such discrepancies in the produced p-nuclei abundances lead to investigate possible contribu- tions from multiple processes other than the 𝛟 process. Alternative scenarios include the rp, 𝜈p, and 𝜈r processes, which will be briefly discussed here. An alternative mechanism for the production of the p nuclei was proposed to take place on the surface of a neutron star that accretes H- and He-rich material from a companion star. This accretion leads to a large gravitational energy release called an X-ray burst. Within this environment, some of the lighter p nuclides can be synthesized through the rp-process (short for rapid proton capture 44 process) [150]. The rp-process path proceeds through a series of proton captures on CNO nuclei toward progressively heavier nuclei, until the proton dripline is approached and 𝛜+ decay takes place. While this process can produce the lightest p nuclei, it remains uncertain whether the synthesized material can escape the strong gravitational pull of the neutron star [151]. Recent studies suggest that if the neutron star is accreting in binary common envelopes, where the expanding star envelops its companion, then the material produced by the rp-process may escape the strong gravitational field and be ejected into the interstellar medium [152]. An additional mechanism involves the neutrino-driven winds produced in CCSN. For values of the electron fraction 𝑌𝑒 > 0.5, a proton-rich neutrino-driven wind can be obtained, and the so-called 𝜈p process can occur [153, 154]. The proton-rich environment is constantly supplied by a small number of free neutrons created by antineutrino captures on free protons. The resulting nucleosynthesis flow near the heavy elements is similar to the r-process, as it is characterized by rapid proton captures in a (𝑝, 𝛟)-(𝛟, 𝑝) equilibrium, with (𝑛, 𝑝) reactions connecting the isotonic (𝑁 = const) chains. This process has been shown to produce the lightest p nuclei [155], however such calculations depend on large uncertainties in neutrino interaction cross sections, the average energies associated with different neutrino flavors, the overall neutrino luminosity, and the specific details of the stellar evolution and explosion models used in the simulations. A recent research worth highlighting proposed an new nucleosynthesis mechanism that could contribute to p-nuclei production, called the 𝜈𝑟 process [156]. This process is suggested to take place on neutron-rich ejecta, where r process occurs. In this scenario, r-process seeds experience strong neutrino irradiation, and thus the (𝑛, 𝛟) (𝛟, 𝑛) equilibrium is broken by the neutrino interactions instead of 𝛜 decays. This pushes the nuclear flow toward and beyond the valley of stability, producing p nuclei. This process is highly dependent on uncertainties on neutrino interactions, and the specific astrophysical conditions required for such strong neutrino fluxes are still uncertain. 3.5 Nuclear Networks and Uncertainties As discussed in the previous sections, stellar nucleosynthesis involves various complex pro- cesses that occur simultaneously in a stellar environment. Simulating these processes requires 45 stellar evolution codes that integrate nuclear physics with the physical mechanisms governing stars, such as gas properties, hydrodynamics for hydrostatic equilibrium, and energy transport via radia- tion or convection. An example of such a code is Modules for Experiments in Stellar Astrophysics (MESA) [157], a 1D stellar evolution code that employs modern numerical and software techniques to solve stellar structure and composition equations while incorporating nuclear physics. However, simulating an entire star is computationally intensive, particularly for environments such as the p or r process, which involve hundreds of thousands of reactions across thousands of isotopes. To simplify the problem and reduce computational costs, one approach is post-processing. Post-processing nuclear network calculations requires a prior stellar evolution simulation, such as one produced by MESA, using a reduced network of isotopes and reactions. These networks are limited to reactions critical for energy generation and those that significantly alter the star’s composition. For example, simulating hydrostatic hydrogen burning only necessitates reactions from the pp-chains and CNO cycles (see Sec. 3.3.1), as these are key to energy production. Reactions on other light nuclei, while important for accurately reproducing final abundances, can be omitted during the evolution phase without affecting the star’s structure, density, or temperature. Once the stellar evolution simulation is complete, the temperature and density profiles as functions of time (trajectories) are extracted. Post-processing then uses these trajectories along with the initial abundances of all isotopes and nuclear physics inputs, such as masses, half-lives, and reaction rates to calculate nucleosynthesis. At this stage, the problem is reduced to solving a system of ordinary differential equations that describe the production and destruction of each nuclear species. This decoupling of nucleosynthesis calculations from stellar structure and evolution allows the inclusion of extensive networks of isotopes and reactions without making the computation prohibitively expensive. An example of such a post-processing framework is provided by the nucleosynthesis grid (Nu- Grid) collaboration to perform both single-zone (PPN) and multizone parallel (MPPNP) simulations for given thermodynamic conditions [158, 159]. The difference between a single-zone and a multi- zone model lies in the mass coordinates that the model can simulate. A single-zone model would 46 follow the evolution of a single mass coordinate of the star for one trajectory, while a multi-zone model can follow multiple zones with different initial abundances and trajectories. In such complex calculations, uncertainties in input quantities naturally have a significant impact on the calculated abundances. For the 𝛟-process network, in addition to the astrophysical uncertainties described in Sec. 3.4.1 regarding the astrophysical site and the distribution of seed nuclei, numerous nuclear uncertainties affect the network calculations, as there are nearly 20 000 nuclear reactions on almost 2000 nuclei that must be considered [131]. As the nuclides involved in the 𝛟 process are predominantly stable or moderately unstable proton-rich nuclei, their masses and corresponding reaction Q-values are generally well-known. Similarly, half-lives are in principle known, aside from the dependence of electron captures and 𝛜+-decay rates on ionization and thermal excitation in the stellar plasma that require theoretical corrections [134]. The main uncertainty on the nuclear physics lies in the photodisintegration reaction rates, which must be determined with high accuracy for many possible reactions. As experimental data are scarce, uncertainties in the predicted reaction rates increase substantially for nuclei farther from stability [160]. Given that the 𝛟-process network involves thousands of possible reactions, it is crucial to focus on those with the most significant impact to address the problem effectively. To this end, several sensitivity studies have been conducted over the years [160, 161, 162] to identify reactions whose uncertainties notably influence the production of specific p nuclei, helping in the planning of nuclear physics experiments. During the recent decades, significant experimental efforts have been made to measure cross sections relevant to the 𝛟 process on stable nuclei [163, 164, 165, 166, 167, 168, 169, 170, 171, 172]. However, only one experiment involving a radioactive beam has been conducted to date [173]. For the thousands of reactions yet to be measured experimentally, reaction rates rely primarily on Hauser-Feshbach (HF) theoretical calculations. The HF model calculations most commonly adapted for 𝛟-process networks are obtained from the Non-Smoker code [174]. As discussed in Sec. 2.4.2.2, HF cross-section calculations depend on the nuclear optical model potential (OMP), nuclear level density (NLD), and 𝛟-ray strength functions (𝛟SF). For heavier p 47 nuclei, where lower temperatures are required and the relevant energies lie near the lower end of the Gamow window, uncertainties in the OMP dominate. In contrast, for lighter p nuclei, which require higher temperatures and relevant energies correspond to the higher end of the Gamow window, uncertainties in the NLD and 𝛟SF have a greater impact on the calculated reaction rates. Developing experimental techniques to directly measure 𝛟-process reactions involving unsta- ble isotopes is therefore critically important. This thesis focuses on the application of such an experimental technique, to study the destruction of the lightest p nucleus, 74Se. 3.6 The Lightest p Nucleus, 74Se 74Se is the lightest of the p nuclei, as its production is shielded from the s-process path and r-process decay chains as shown in Fig. 3.7. In sensitivity studies of the 𝛟 process during SNII Figure 3.7 The lightest p nucleus, 74Se, is shielded by the valley of stability from the s-process path and the r-process decay chain. The gray and white squares correspond to stable and radioactive isotopes, respectively, and the blue square indicates the 74Se nucleus of interest. The halflives are obtained from [4]. scenario the 74Se(𝛟, 𝑝)73As reaction has been identified as key reaction rate to impact the final abundances of 74Se [160, 161]. While stellar models for SNII [136, 176, 142, 177] show some variation, 74Se is often found to be overproduced compared to solar abundances, as shown in Fig. 3.6. 48 Figure 3.8 The 𝛟-process fluxes producing and destroying 74Se during a SNII. The sum of all production and destruction fluxes is normalized 100%. Fluxes smaller than 1% are not shown. Fluxes obtained from [175], using trajectories from [160]. The possible production and destruction mechanisms of 74Se in a SNII, shown in Fig. 3.8, have been a topic of experimental studies for many decades. Of those reactions the 74Se(𝑝, 𝛟)75Br [178, 179, 180, 181], 70Ge(𝛌, 𝛟)74Se [182], and 74Se(𝑛, 𝛟)75Se [183] have been measured directly, and the 75Se(𝛟,n)74Se can be inferred from the latter through the reciprocity theorem (see Sec. 2.5.2). The only reaction channels that immediately affect the final 74Se abundance, for which no experimental data exist are the 74Se(𝛟, 𝑝)73As and the 74Se(𝛟, 𝑛)73Se. This work focuses on the measurement of the inverse 73As( 𝑝, 𝛟)74Se reaction, that can be used to calculate the ground-state contribution of the 74Se(𝛟, 𝑝)73As reaction through the reciprocity theorem. In simulations of SNII [142], the maximum production of 74Se is found in layers with peak temperature 𝑇 ≈ 3 GK. For such temperature, the Gamow window for the 73As(𝑝, 𝛟)74Se reaction is located at center-mass-energy ranging from 1.7 to 3.6 MeV. The predicted cross section of the 73As(𝑝, 𝛟)74Se reaction in this energy range is shown in Fig. 3.9. The solid black line corresponds to standard statistical model calculations using the Non-Smoker code [174], and the blue band is calculated through Talys [62], using the various possible models for the OMP, NLD, and 𝛟SF discussed in Sec. 2.4.2.5. It can be seen that the statistical properties of the 74Se nucleus result in a cross section uncertainty of a factor of 6 at center-mass-energy near 3 MeV. This thesis aims 49 Figure 3.9 The cross section of the 73As(𝑝, 𝛟)74Se reaction using standard statistical model calculations from the Non-Smoker code [174] (black line) and Talys [62] (blue band). The Talys calculations include all possible NLD and 𝛟SF options listed in Tables 2.1 and 2.2, as well as the default and JLM OMP discussed in Sec. 2.4.2.5. The energy range covers the Gamow window for the 𝛟 process at 𝑇 = 3 GK. to investigate whether this uncertainty in the reaction rate is responsible for the overproduction of 74Se in SNII models and whether an experimental measurement can help resolve this discrepancy. Regarding the SNIa scenario, while sensitivity studies suggest that nuclear uncertainties in 74Se reactions do not significantly affect 74Se production [162], this work aims to provide a quantitative assessment of their potential role. The following chapter discuses the experimental setup for the measurement of the 73As( 𝑝, 𝛟)74Se cross section using a radioactive 73As beam. Chapter 5 presents the analysis from the proof-of- principle stable beam experiment on the 82Kr(p, 𝛟)83Rb reaction, the measured 73As( 𝑝, 𝛟)74Se reaction, as well as the development of an analysis method to constrain the NLD and 𝛟SF used in the statistical model calculations. Chapter 6 presents the results of the two cross section mea- surements, and in chapter 7 the impact of the measured 73As( 𝑝, 𝛟)74Se reaction cross section is investigated in a SNII and SNIa scenario. 50 CHAPTER 4 EXPERIMENTAL SETUP & TECHNIQUES This work regards to two experiments in inverse kinematics using the same experimental setup. The first experiment, conducted in 2017, served as the proof-of-principle stable beam experiment for the measurement of the 82Kr(𝑝, 𝛟)83Rb and 84Kr(𝑝, 𝛟)85Rb reaction cross sections. The 84Kr(𝑝, 𝛟)85Rb reaction has been analyzed and published by a former student of the group [184], therefore this work focuses only on the 82Kr(𝑝, 𝛟)83Rb measurement [185]. The second experiment is the radioactive beam experiment for the measurement 73As(𝑝, 𝛟)74Se reaction cross section that took place in 2023. The experiments took place in the ReA post-accelerator of the Facility for Rare Isotope Beams (FRIB) (formerly known as National Superconducting Cyclotron Laboratory) at Michigan State University. More details on the facility and the delivered beams are discussed in Sec. 4.1. An illustration of the experimental setup is shown in Fig. 4.1. The 82Kr and 73As beams impinged onto a hydrogen gas-cell target described in Sec. 4.2. The 𝛟 rays produced by the Figure 4.1 The experimental setup with the SuN and SuNSCREEN detectors in ReA3. The beam enters from the right, as shown by the white arrow. The inset shows a side-section view of the SuN detector, indicating the location of the hydrogen gas cell. 51 (𝑝, 𝛟) reaction were detected using the Summing NaI (SuN) detector and analyzed using the 𝛟- summing technique, as discussed in Sec. 4.3. To minimize cosmic-ray background contributions, the Scintillating Cosmic Ray Eliminating ENsemble (SuNSCREEN) was used, as explained in Sec. 4.4. 4.1 Beam Delivery in ReA The Facility for Rare Isotope Beams (FRIB) is a national scientific user facility that can provide access to thousands of nuclei far from stability [186]. It succeeded the Coupled Cyclotron Facility at the National Superconducting Cyclotron Laboratory (NSCL) that pursued a very successful science program with rare isotopes produced by projectile fragmentation until early 2021 [187]. Along with fast and stopped beams, the facility can provide reaccelerated beams in energies ranging from 300 keV/nucleon to 6 MeV/nucleon, providing unique opportunities for nuclear astrophysics experiments. Ion beams can alternatively be generated in the so-called “offline mode”, using radioactive or stable source samples [188]. Such was the case for the stable 82Kr, and the radioactive 73As beam. Firstly the source samples were evaporated in an ion source and directed toward the ReAcceler- ator charge breeder. For the 73As source, the Batch-Mode Ion Source (BMIS) was used (Fig. 4.2), which is an oven ion source combination commissioned in 2021 [188]. The 73As sample was inserted inside a cylindrical tantalum oven, and heated up to about 1000 0C. With the oven-ion source biased to a few tens of kV, the evaporated material exited through a transfer tube and was directed to the ReAccelerator charge breeder. Details on the preparation of the radioactive 73As source sample can be found in Ref. [189]. The ReA charge breeder is the Electron Beam Ion Trap (EBIT) [190]. EBIT is a superconducting magnet in which the evaporated ions were injected and charge bred to high charge states (73As+23 and 82Kr27+). After the EBIT, the beam was mass selected in a charge-over-mass (𝑄/𝐎) separator and accelerated through the Radio Frequency Quadrupole (RFQ) [191] and superconducting LINAC (short for linear accelerator) [192]. The ReA LINAC has three accelerating cryomodules that include superconducting quarter wave resonators (QWR) and superconducting solenoids (SS) that 52 Figure 4.2 The batch mode ion source: (left) internal components of the oven-ion source and (right) the fully assembled front end. accelerated the 82Kr and 73As beams. The beams were delivered to the experimental end station in the ReA3 area, shown in Fig. 4.1, with energies 3.1, 3.4 and 3.7 MeV/nucleon for the 82Kr beam, and 3.1 and 3.7 MeV/nucleon for the 73As beam. The measurement of the beam current for the 82Kr experiment was achieved by electrically isolating the beam pipe with an isolating flange upstream of SuN and using the entire beamline as a Faraday cup. This way, any electrons produced by ionization in the cell or beam pipe were still recorded and the total current did not get affected. Due to issues with grounding the beamline was not properly isolated during the 73As experiment and the beam current was measured in regular intervals through a Faraday cup upstream. More details on the current measurements are provided in Sec. 5.2. However, it is worth noting that while the proof-of-principle experiment was performed with a stable 82Kr beam of intensity on the order of 107 particles/sec, the radioactive 73As beam was around 105 particles/sec, almost two orders of magnitude lower. This highlights the inherent challenges of working with radioactive beams, particularly with a highly toxic element like arsenic. Nevertheless, as will be shown in the following chapters, the measurement was successfully achieved. 53 4.2 Hydrogen Gas Target The beam impinged on a hydrogen gas target positioned at the center of the SuN detector. The target system involved two main components: the gas cell, which held the hydrogen gas, and the gas handling system. The gas cell, designed at Hope College, consisted of two plastic halves. A schematic side section of the cell is shown in Fig.4.3, with construction images in Fig.4.4. A 2-𝜇m thick molybdenum entrance and 5-𝜇m thick exit window were glued onto the tantalum rings, as shown in the figures. Figure 4.3 Side-section illustration of the gas cell showing the tantalum rings supporting the entrance and exit molybdenum windows, along with openings for the gas supply and grounding cable. The choice of molybdenum for the target window was based on several criteria. High atomic number (𝑍) materials are required to ensure that the Coulomb barrier is high enough, keeping the threshold for fusion-evaporation reactions well above the maximum beam energy. Additionally, the target window needs to be as thin as possible to minimize beam straggling and energy loss. At the same time, the material must have high strain tolerance, allowing the very thin foil to withstand the atmospheric pressure of the hydrogen gas against the vacuum in the beamline without rupturing. Lastly, it must be available from the manufacturer as leak-tight, so the hydrogen gas remains secured in the cell. Ensuring that the components of the gas-cell interacting with the beam do not create significant 54 Figure 4.4 (left) tantalum rings with molybdenum entrance and exit windows during construction (right) assembled gas cell beam-induced background is critical. This background can become significant, especially if the beam interacts with anything plastic, as synthetic polymers consist mostly of carbon and hydrogen atoms, which can become target for the (𝑝, 𝛟) reaction of interest, or get scattered, affecting the quality of the data. For these reasons, the inner volume and front face of the cell were lined with tantalum foil to shield the plastic components from the beam. The tantalum also allowed charge collection from the entire inner volume of the cell. The back half of the cell contained two small openings: one for the grounding cable that was connected to the inner lining, which prevents charging and discharging inside the cell, and another for attaching the gas supply pipe. Both openings were sealed with epoxy glue after assembly to ensure leak-tight operation. The gas handling system ensured safe handling and proper disposal of the flammable hydrogen gas. The system, shown in Fig. 4.5, consisted of a 10 L hydrogen reservoir, flowmeters, and regulators to control the slow filling and emptying of the cell, preventing rupture of the thin window foils. It also included a dry nitrogen supply line to purge the hydrogen before venting, preventing hazardous mixtures with atmospheric oxygen. A series of valves regulated the flow, while two manometers monitored the pressure of the reservoir and gas cell. The reservoir remained overpressurized at ≈ 850-900 Torr, to mitigate the risk of atmospheric oxygen entering the small volume of the container. The cell was filled with 600 Torr of hydrogen during operation, with its pressure continuously monitored and recorded throughout the experiment. The system featured 55 detailed procedures for pumping down, filling the reservoir and gas cell with hydrogen, purging hydrogen, and venting. Multiple interlocks were incorporated to ensure that, in the event of target failure—such as a rupture of the target window—any potential hazards were effectively mitigated. Figure 4.5 The setup of the 73As(𝑝, 𝛟)74Se experiment in the ReA3 experimental area. The picture shows the SuN and SuNSCREEN detectors and the hydrogen gas supply system. 4.3 The SuN Detector Surrounding the hydrogen gas target was the Summing NaI (SuN) 𝛟-ray detector, shown in Fig. 4.1 and 4.5. SuN is a calorimeter with the shape of a 16 × 16 inch barrel with a 1.8 inch diameter borehole along its axis. The barrel is segmented into 8 NaI(Tl) crystals, each connected to three photomultiplier tubes (PMTs) [193, 194]. The large volume of the detector allows for high 𝛟-ray detection efficiency and nearly 4𝜋 solid angle coverage for a source placed in its center. Sodium iodide (NaI) crystals doped with traces thallium (Tl) are the most common inorganic scintillation material, widely used for the detection and spectroscopy of 𝛟-ray radiation. Introduced in 1948 [195], NaI(Tl) crystals are valued for their availability in large volumes at relatively low 56 costs, which often outweighs the advantages of newer inorganic scintillators offering higher light output, better energy resolution or faster timing capabilities [196]. However, NaI(Tl) is highly hygroscopic and deteriorates upon contact with atmospheric water. To prevent this, the top and bottom SuN crystals are encapsulated in aluminum casing. Additionally, the crystals are surrounded by a reflective polytetrafluoroethylene layer, therefore they are optically isolated from each other and can function as individual detectors. When ionizing radiation enters the active volume of the detector, it excites the crystal atoms, leading to emission of visible light with wavelength of approximately 415 nm during de-excitation. This light is collected by the PMTs, where it is converted to photoelectrons. These photoelectrons are accelerated and multiplied through a series of dynodes, creating an electrical signal whose magnitude is proportional to the incident radiation energy [196]. The signals from the PMTs are amplified by a pre-amplifier and processed by XIA Pixie-16 digitizers. In the Pixie-16 modules the analog electronic signal from the PMTs is converted to a digital representation through an analog to digital converter (ADC). The digitizers are configured, read out and analyzed using the Digital Data Acquisition System (DDAS) [197], a lab-supported framework built around the Pixie-16 digitizers. The FRIBDAQ software suite manages the data flow and sets up the analysis pipeline, transforming the raw data from its hexadecimal format into physically meaningful parameters, such as energy and time spectra [194]. 4.3.1 Summing Technique The large angular coverage and high detection efficiency of the detector allows for the application of the 𝛟 -summing technique [198, 193]. In this technique, the spectra obtained by the individual crystals (segments) provide sensitivity to the individual 𝛟 -ray transitions, whereas the full energy deposited in SuN provides sensitivity to the populated excitation energies. More specifically, there are three main types of spectra used in this analysis: the Sum of Segments (SoS), the Total Absorption Spectrum (TAS), and multiplicity spectrum. The SoS corresponds to the sum of the energy spectra recorded by each one of the optically isolated segments. This contains all the individual 𝛟-ray transitions that occurred within the compound nucleus and got detected by SuN. TAS includes the 57 energy deposited in all segments added up on an event-by-event basis. This represents to the full energy deposited inside the detector and reflects the populated excitation energy in the compound nucleus. Finally, the multiplicity spectrum indicates how many segments of SuN recorded energy in each event, which is indicative of the 𝛟-ray multiplicity in a 𝛟 cascade. An example of the summing technique using the spectra of a 60Co calibration source is shown in Fig. 4.6. The 60Co nucleus populates excited states in the 60Ni compound nucleus by 𝛜− decay. (a) (c) (b) (d) Figure 4.6 SuN spectra from a 60Co nucleus: (a) decay scheme of 60Co, (b) SoS spectra showing the two characteristic 𝛟 rays, (c) TAS with a sum peak at 2.5MeV and (d) multiplicity spectrum showing the majority of cascades to have multiplicity two. 58 In most cases (99.88%), the 𝛜− decay populates a state of 𝐞 𝑋 = 2.505 MeV, that will subsequently decay by the emission of two 𝛟 rays, of energies 1.173 MeV and 1.332 MeV. A small fraction (0.12%) of the decays populate the 1.332 MeV state and therefore only one 𝛟 ray is emitted. The SoS spectra (Fig. 4.6b) contain both the 1.173 MeV and 1.332 MeV, with a slightly higher intensity on the 1.332 MeV 𝛟 ray, as well as a small peak at 2.505 MeV, in case both 𝛟 rays are recorded by the same segment. The dominant feature of the TAS spectrum (Fig. 4.6c) is a sum peak at 2.505 MeV corresponding to the excitation energy of the 60Ni compound nucleus, and small peaks at 1.173 MeV and 1.332 MeV, corresponding to the 0.12% chance of populating the 1.332 MeV, as well as the few instances of incomplete summation. The multiplicity spectrum (Fig. 4.6d) shows the majority of events with multiplicity 2, and the average multiplicity is 2.06. Events with multiplicity higher than 2 correspond to scattered 𝛟 rays deposited their energy to more than one segment. Building on the simplistic example of the 60Co decay, Fig. 4.7 shows the application of the summing technique in a more complex scenario, such as a capture reaction experiment. The sum Figure 4.7 Illustration of the energetics of a 𝐎 𝑋 ( 𝑝, 𝛟) 𝐎𝑌 reaction with the summing technique. The compound nucleus 𝐎𝑌 is populated at an excitation energy 𝐞 𝑋 = 𝐞CMS + 𝑄 and a sum peak forms at the TAS spectrum at energy 𝐞 𝑋. 59 peak forms in the TAS spectrum at energy 𝐞 𝑋 = 𝐞CMS + 𝑄, where 𝐞CMS is the center-of-mass energy and 𝑄 the reaction 𝑄 value. As will be discussed in Sec. 5.5, the efficiency of the sum peak depends on the multiplicity of the cascade [193], and therefore it is important to take all three types of spectra into consideration when applying the summing technique. 4.4 The SuNSCREEN Detector In the energy region of interest for these measurements, the main source of background comes from cosmic rays, which pose a significant challenge when trying to measure very small cross sections on the order of millibarns. To address this issue and increase the sensitivity of the SuN detector, the Scintillating Cosmic Ray Eliminating ENsemble (SuNSCREEN) [199] was positioned above SuN, as shown in Fig. 4.1 and 4.5. SuNSCREEN is a plastic scintillator detector array comprised of nine bars, each with two PMTs, forming a roof-like arrangement above the SuN detector. To reduce the cosmic-ray induced background, SuNSCREEN was used as a veto detector. For this reason, all events that recorded signals in both PMTs of a SuNSCREEN bar, and at least one segment of SuN were rejected from the SuN spectra. During the SuNSCREEN’s commissioning this method was shown to reduce the cosmic-ray background contributions by up to a factor of four in the SoS spectra and a factor of two in TAS [199]. Spectra of the background radiation with and without the SuNSCREEN veto gate applied are shown in Fig. 4.8, where in the region around 10 MeV, which is the relevant region for the present work, a background reduction of a factor of three is observed in both SoS spectra and TAS. The spectra also show characteristic background peaks at 1461 keV and 2614 keV from natural radiation, as well as a small peak at 6.8 MeV in the TAS from neutron capture on the NaI crystal via the 127I(𝑛, 𝛟)128I reaction with Q-value of 6.8 MeV. Additional background reduction is achieved by taking advantage of the time structure of the beam as discussed in Sec. 5.3.4. 60 (a) (b) Figure 4.8 Typical room background (a) sum of segments and (b) total absorption spectra shown in red dashed line. The same spectra are shown with black solid line after SuNSCREEN veto rejection. The lower panels show the ratio before and after rejection. 61 CHAPTER 5 ANALYSIS As discussed in Ch. 2, the cross section formula used in this analysis is given by Eq. 2.10 𝜎 = 𝑌 𝐌𝑎 𝑁𝑡 𝜖 (5.1) where 𝑌 the experimental yield, meaning the number of reactions recorded by the detector, 𝐌𝑎 the total number of beam particles, 𝑁𝑡 the number of target nuclei and 𝜖 the detection efficiency. This chapter focuses on calculating the cross sections for the 82Kr(𝑝, 𝛟)83Rb and 73As(𝑝, 𝛟)74Se reactions. Details on the calculation of each parameter, along with the associated uncertainties are discussed, and the resulting cross section values are presented in Ch. 6. 5.1 Effective Energy Before calculating the individual parameters of Eq. 5.1, it is useful to determine the center- of-mass energy at which the cross section is measured. In thin target experiments, where energy loss through the target is minimal, it is common to assume that the reaction energy corresponds to the one in the middle of the target. However, for the 4-cm-long gas cell used in this experiment, the effective center-of-mass energy 𝐞eff needed to be calculated [15]. The effective energy 𝐞eff, represents the beam energy within the target at which half of the total yield is produced. For non-resonant reactions the astrophysical 𝑆 factor (Sec. 2.5.4) can be considered nearly constant over a relatively small energy interval of the target thickness Δ. This can be utilized to calculate the 𝐞eff from the integral of the cross section from Eq. 2.36 over the target thickness Δ as ∫ 𝐞0 𝐞0−Δ𝐞 1 𝐞 exp(−2𝜋𝜂)𝑑𝐞 = 2 ∫ 𝐞0 𝐞eff 1 𝐞 exp(−2𝜋𝜂)𝑑𝐞 (5.2) where 𝐞0 is the incident projectile energy and Δ𝐞 is the energy loss within the target. Assuming the cross section decreases linearly between 𝜎1 at 𝐞0 and 𝜎2 at 𝐞0 − Δ𝐞, the effective energy 𝐞eff can be calculated from Eq. 5.2 as 𝐞𝑒 𝑓 𝑓 = 𝐞0 − Δ𝐞 + Δ𝐞 − 𝜎2 𝜎1 − 𝜎2 + (cid:34) 𝜎2 1 + 𝜎2 2 2(𝜎1 − 𝜎2)2    (cid:35) 1/2   62 (5.3) which is a good approximation for 𝜎1/𝜎2 < 10 [15]. The values of 𝜎1 and 𝜎2 are obtained from Non-Smoker [174], and their ratios are 1.6, 1.7, and 2.0 for initial 82Kr beam energy of 3.7, 3.4, and 3.1 MeV/u, and 1.4 and 2.0 for the 73As beam at 3.7 and 3.1 MeV/u. 5.2 Beam Particle Number The design of the experimental setup includes an isolating flange upstream of SuN, and a plastic feedthrough in the end of the beamline for the gas supply. This way the beamline is electrically isolated from the rest of the setup and is used as a Faraday cup. The beam current in the 82Kr experiment was calculated from the ammeter measurement of the beamline. Unfortunately, during the 73As experiment the isolation was not successful and as shown in Fig. 5.1 a clear charge and discharge of the beamline can be seen in the measured current in the order of tens of fA. Figure 5.1 The 73As beam current measured from the ammeter connected to the improperly isolated beamline as a function of time. The blue highlighted areas correspond to data acquisition times. The small frequent drops correspond to upstream Faraday cup measurement where no beam is present. As the baseline varies by tens of fA, whereas the beam current was only a few fA, this measurement cannot be used for analysis. As the deposited beam current was in the order of a couple fA, the measured current from the dowstream ammeter cannot be used for this analysis. Instead, throughout the experiment, Faraday cup measurements were taken every 20 minutes that interrupted the beam momentarily, as can be seen from the frequent dips in Fig. 5.1 that reflect the position of the baseline during that time. 63 The beam current calculations are shown in Fig.5.2. Every 20 minutes a Faraday cup interrupted the beam upstream of the SuN detector, and the current was measured off of that cup. The top Figure 5.2 Beam current calculation for the 73As experiment. The top panel displays the upstream Faraday cup measurements taken every 20 minutes, with the inset zooming in on one minute of data and identifying the plateau region used for the calculation. The bottom panel highlights the integrated regions, with shaded areas indicating data acquisition periods. panel shows the current recorded by that collimator, were each peak corresponds to a current measurement. The beam intensity was considered at the plateau of the peak, shown in the inset. The baseline was fitted with a linear fit, reflected by the dashed line. The blue solid line in the bottom panel corresponds to the subtracted baseline from the peaks. The total deposited beam 64 charge, 𝐶tot, was calculated using square integrals between consecutive peaks, shown by the shaded region in the bottom panel. The total number of beam particles, 𝐌𝛌, was calculated as 𝐌𝛌 = 𝐶tot 𝑄beam · 𝑞𝑒 (5.4) were 𝐶tot is the integral of the beam current, 𝑄beam is the charge state of the beam (23 for 73As and 27 for 82Kr), and 𝑞𝑒 the electron charge. 5.3 Experimental Yield 5.3.1 SuN Gainmatching & Energy Calibration The energy spectra from the SuN detector, acquired using the 12-bit digitizers discussed in Sec. 4.3, are expressed in 212 = 4096 ADC channels. Each channel corresponds to a specific voltage height of a PMT output signal. The first step in analyzing these spectra is to ensure that all 24 PMTs respond consistently to 𝛟 rays of the same energy. This procedure is known as gainmatching. To eliminate position dependence of the PMTs within each crystal, gainmatching is typically performed with 𝛟 rays from natural background radiation, such as 40K. Approximately 10% of the 𝛜− decays of 40K populate an excited state of 40Ar, resulting in the emission of a characteristic 1461 keV 𝛟 ray. Gainmatching for the SuN PMTs is conducted in two stages. The first stage, hardware gain- matching, involves adjusting the voltages applied to each PMT before the experiment so that the 1461 keV 𝛟 ray from 40K appears in approximately the same ADC channel in the recorded spectrum. The second stage, software gainmatching, fine tunes the PMT gains by normalizing the obtained spectra so the 40K peak appears in the exact same channel for each PMT. Fig. 5.3 shows the spectra from all PMTs in the 1461 keV region before software gainmatching during the 73As experimemt. Gaussian fits on a linear background were applied to locate the peak centroids, and the resulting gainmatching factors, listed in Table 5.1, were calculated to align the peaks to the same ADC channel, as shown in Fig. 5.4. 65 Figure 5.3 Background spectra of all 24 SuN PMTs before gainmatching in the region near the 1461 keV 40K peak during the 73As experiment. Each panel represents one detector segment, showing spectra from its three associated PMTs, with black, red, and blue lines corresponding to PMTs 1, 2, and 3, respectively. Figure 5.4 Same as Fig. 5.3 after applying the gainmatching factors from Table 5.1. After gainmatching, all PMTs produce consistent spectra, however, the spectra remain in arbi- trary ADC units. To find the correlation between ADC units and energy, an energy calibration is performed using sources that emit characteristic 𝛟 rays of known energy. The 𝛟 rays used were the 59.5 keV 241Am peak, the 661.7 keV 137Cs peak, and the 1173.2 keV and 1332.5 keV peaks from 60Co. Additionally, 𝛟 rays from a 228Th source were used, specifically the 238.6 keV peak from the 212Pb daughter nucleus and the 583.2 keV and 2614.5 keV peaks from 208Tl. The resulting 66 calibrations are shown in Fig. 5.5 and the fit parameters are listed in Table 5.1. Table 5.1 The gainmatching and calibration factors for SuN for the 73As experiment. Gainmatching Factors PMT Factor 1.0075 B11 0.9997 B12 0.9929 B13 1.0005 B21 1.0000 B22 1.0022 B23 1.0116 B31 0.9903 B32 0.9812 B33 1.0018 B41 0.9892 B42 0.9825 B43 PMT Factor 0.9927 T11 1.0141 T12 1.0135 T13 0.9948 T21 0.9873 T22 1.0141 T23 0.9863 T31 1.0003 T32 0.9962 T33 1.0166 T41 1.0059 T42 1.0136 T43 Calibration Factors Segment B1 B2 B3 B4 T1 T2 T3 T4 Scale 0.1815 0.1839 0.1801 0.1817 0.1812 0.1837 0.1801 0.1824 Intercept -28.0864 -29.7079 -27.8599 -29.5897 -27.6117 -26.6570 -26.8405 -27.9716 Figure 5.5 SuN calibration fits. Each panel corresponds to one segment of SuN. 67 5.3.2 The Sum Peak As discussed in Sec. 4.3, in the summing technique the experimental yield 𝑌 is calculated from the integral of the sum peak, that forms in the TAS spectrum at energy 𝐞 𝑋 = 𝐞CMS + 𝑄, where 𝐞CMS is the center-of-mass energy and 𝑄 the reaction 𝑄 value. The summing technique has been successfully applied to a plethora of (𝑝, 𝛟) and (𝛌, 𝛟) reaction measurements on stable nuclei [198, 200, 201, 202, 203]. These experiments are typically conducted in regular kinematics, where a 𝑝 or 𝛌 beam impinges on a heavy, solid, stable target. For radioactive isotopes, however, it becomes necessary to transition to inverse kinematics, as constructing targets from exotic isotopes with short half-lives is highly challenging. Inverse kinematics is a widely used approach at many facilities for 𝛟-process measurements using mass separators [204] and storage rings [166], but there has so far been only one experiment with a radioactive beam [173]. The summing technique has been successfully applied in inverse kinematics using a solid target [205]. However, as the beam intensity in radioactive beams decreases by orders of magnitude compared to stable beams, the introduction of a gas target was necessary to increase the purity of the target, and allow for more efficient measurements. Transitioning to inverse kinematics and the introduction of a gas target increases the complexity of this method. Firstly, due to the experiment being conducted in inverse kinematics, the recoil nucleus has significant momentum and continues its path along with the unreacted beam, rather than remaining stationary in the target. Therefore the 𝛟 rays are emitted from a moving source, and Doppler corrections need to be applied for the detected 𝛟-ray energy [206]. Additionally, the beam’s passage through the target entrance window introduces energy straggling, resulting in a range of incident beam energies. This, in turn, populates the compound nucleus at a range of excitation energies, causing a significant widening of the resulting sum peak. 5.3.3 Doppler-Shift Corrections The Doppler effect is the change in frequency of a wave emitted by a moving source relative to a stationary observer, compared to the frequency that would be measured if the source were at rest. For a relativistic moving particle, such as light, the observed energy will be shifted according to 68 the velocity of the moving source. Therefore the energy of the 𝛟 rays emitted by the moving recoil nucleus are shifted by 𝐞0 = 1 − 𝛜 cos 𝜃 √1 − 𝛜2 𝐞 (5.5) where 𝐞0 the 𝛟-ray energy emitted by the source, 𝐞 is the detected 𝛟-ray energy by the stationary observer, 𝛜 = 𝑣/𝑐 the recoil relative velocity and 𝜃 the relative angle between the recoil and the detector. Assuming the decay of the moving recoil in the center of the detector, each SuN segment has a different angle corresponding to the center of the segment, as shown in Table 5.2. Table 5.3 shows the different beam energy for the 73As and 82Kr beam in the laboratory frame, the center-of- mass energy at the center of the target accounting for the energy loss through the entrance window and the recoil relative velocity 𝛜. Table 5.2 SuN segment angles from Ref. [194]. Segment Angle (deg) 1 2 3 4 2.550 2.024 1.118 0.592 Table 5.3 82Kr and 73As beam energies and relativistic ve- locities. Beam 82Kr 73As Energy Lab CoM energy in middle (MeV/u) 3.7 3.4 3.1 3.7 3.1 of target (MeV/u) 2.98 2.67 2.37 2.95 2.31 𝛜 = 𝑣/𝑐 0.079 0.075 0.071 0.079 0.070 Once all acquired spectra were corrected on a segment-by-segment basis, they were summed to form a Doppler-shift corrected sum peak. 5.3.4 Background Subtraction There were two main types of background contributions that could interfere with the sum peak in the energy region of interest: cosmic-ray background and beam-induced background. Minimizing those background contributions in the region where the sum peak was expected was important for the accurate determination of the experimental yield. The cosmic-ray background, as discussed in Sec. 4.4, was significantly reduced using the SuNSCREEN veto detector. Any remainder cosmic-ray background contributions, were removed by utilizing the pulsed structure of the beam. The beam was delivered in 80 𝜇s pulses every 200 69 ms. While data was recorded continuously, two distinct time gates were applied during processing. The first gate corresponded to the 80 𝜇s beam-on intervals, triggered by a signal from the EBIT charge breeder. The second gate, applied 100 ms later, captured 800 𝜇s of background data between pulses. An illustration of this structure is shown in Fig. 5.6. The background data were scaled Figure 5.6 Illustration of the pulsed-beam structure. Blue boxes correspond to 80 𝜇s beam pulses occurring every 200 ms, while grey boxes reflect the 100 ms time-shifted gate used to record 800 𝜇s of background between beam pulses. The time axis is not to scale. down by a factor of ten to account for the shorter recorded time and subtracted from the beam-on data to remove room-background contributions. The last form of background originates from the beam and regards to any beam-induced reactions other than the reaction of interest. This includes beam scattering on any beamline components, or any interaction with the gas cell and its entrance and exit windows. As discussed in Sec. 4.2 high 𝑍 materials are used to cover all parts of the cell the beam may interact with. Unfortunately, during the 73As experiment, of the two gas cell targets used, one had significant scattering background caused by epoxy glue residue that had seeped onto the entrance window. Therefore, the data acquired with that gas target could not be analyzed, as the sum peak was not visible above the background. Removing beam-induced background contributions requires isolating them from the data cor- responding to the reaction of interest. For this purpose data are acquired while the gas cell is full of hydrogen gas, as well as while the cell is empty. The empty cell data are scaled based on the 70 ratio of beam current during full cell and empty cell runs and subtracted from the full cell data. Fig.5.7 shows the Doppler corrected TAS spectrum for the 3.7 MeV/u 82Kr background subtracted sum peak, after subtracting cosmic-ray and room background. The black line corresponds to the full cell data and the red line is the empty cell data scaled based on the total deposited beam current of the full and empty cell runs. Subtracting the two gives the sum peak shown in the blue line. The blue band corresponds to the statistical uncertainty from the background subtraction. Figure 5.7 Doppler-corrected TAS spectra showing the background subtraction for the sum peak for the 82Kr(𝑝, 𝛟)83Rb reaction at the initial beam energy 3.7 MeV/nucleon. The black histogram corresponds to the gas cell filled with hydrogen gas, the red histogram corresponds to the empty gas cell scaled to the beam current, and the blue histogram is the fully subtracted sum peak that was used for the remaining analysis. The blue band corresponds to the statistical uncertainty of the background subtraction. (Figure adapted with permission from Tsantiri et al., Phys. Rev. C 107, 035808 2023. Copyright 2023 by the American Physical Society) The low energy region of the TAS spectrum shows a number of peaks from scattering of the beam on the molybdenum window. This serves as a useful tool to identify if the position of the incoming beam remains the same between full and empty cell runs, as the two spectra should overlap after scaling on the beam current. Similarly in the energy region higher than the sum peak the spectra should overlap as well. In the case of the 73As experiment, scaling on the beam current did not result in overlapping scattering peaks or high energies, indicating that the position of the 71 beam had changed during data acquisition, as shown in Fig. 5.8. For this reason, instead of Figure 5.8 Doppler-corrected TAS spectra for the 73As(𝑝, 𝛟)74Se reaction at the initial beam energy 3.7 MeV/nucleon. The black histogram corresponds to the gas cell filled with hydrogen gas, the magenta histogram corresponds to the empty gas cell, and the blue to the empty gas cell scaled to the beam current. The disagreement of the full cell and scaled empty cell spectra indicates the beam position on the cell has shifted. scaling on the beam current, the empty cell data were scaled based on the high energy region of the spectrum. Specifically, for energies between 13 and 18 MeV, where no contribution of the sum peak is expected, more than 1000 integrals over different energy ranges and regions were sampled. This created a distribution of scaling factors as shown in Fig. 5.9. The background subtraction was performed using the mean scaling factor, and the deviation of the distribution was introduced as an additional source of uncertainty for the remainder of this analysis. Figure 5.10(a) shows the background subtraction for the 3.7 MeV/u 73As beam. The challenge of running a radioactive beam experiment is apparent from the significantly lower statistics and large statistical uncertainties compared to the stable beam experiment. For the 3.1 MeV/u sum peak there was an additional challenge faced. Unfortunately, a few minutes after the beginning of full cell data acquisition in the 3.1 MeV/u beam energy the 2-𝜇m thin molybdenum entrance window ruptured, allowing only for 30-minutes-worth of data acquisition. Regardless of this unfortunate 72 (a) (b) Figure 5.9 Sampling of various high energy regions for empty cell data scaling. Figure (a) shows the full cell data spectrum with black, the various empty cell scaled data with green and the resulting background subtracted sum peaks with blue, where each line corresponds to a different scaling factor. Figure (b) reflects the distribution of scaling factors. More details in text. event, the data were deemed worth analyzing, as this reaction has never been measured before, and any new information that may be provided is important. The significantly low statistics accumulated result in very large statistical uncertainty, as shown by the blue error band in the resulting sum peak of Fig. 5.10(b). The uncertainty quantification is discussed in detail in Sec. 5.7. (a) (b) Figure 5.10 Same as Fig. 5.7 for the 73As(𝑝, 𝛟)74Se reaction at the initial beam energy of (a) 3.7 MeV/nucleon and (b) 3.1 MeV/nucleon. The blue bands correspond to statistical uncertainty. The uncertainty introduced by the empty cell scaling methodology is not included here. 73 5.4 Target Particle Density The next parameter in Eq. 5.1, the target particle density 𝑁𝑡, is calculated based on the average target pressure. The pressure was recorded throughout the experiment and it can be seen for the 73As experiment in Fig. 5.11. Figure 5.11 The gas cell pressure recorded during the 73As experiment. The highlighted regions correspond to the full cell and empty cell runs of the 3.7 and 3.1 MeV/u beam energies. From the target density, the effective target thickness can be calculated using LISE++ [207], an FRIB software used primarily for beam production and transmission calculations. It features many useful calculators, including a physics calculator used in this analysis that calculates effective target thickness of gas target based on the target material, pressure and width. The target particle density 𝑁𝑡, can be calculated through: where 𝑡 is the target thickness in grams/cm3 from LISE++, 𝑁 𝐎 is the Avogadro’s number, and 𝑚𝐻 𝑁𝑡 = 𝑁 𝐎 · 𝑡 𝑚𝐻 (5.6) the hydrogen mass. 5.5 Detection Efficiency The final term to be calculated in Eq.5.1 is the detector efficiency, 𝜖, which represents the fraction of particles recorded relative to the total number emitted. For the SuN detector, a 4𝜋 74 calorimeter, the geometric efficiency is nearly 100% from its design. As a result, the detection efficiency is primarily governed by the intrinsic efficiency, defined as the ratio of detected particles to incident particles. Since uncharged radiations such as 𝛟 rays can travel large distances before interacting, or may interact with materials other than the scintillator, such as the casing, scintillating detectors are typically less than 100% efficient [196]. The intrinsic efficiency varies with energy, 𝜖 (𝐞), and for a total absorption spectrometer like SuN, it also depends on the multiplicity of the cascade [193]. For example, the efficiency of detecting a single 𝛟 ray of 𝐞𝛟 = 10 MeV is higher than detecting two 𝛟 rays of 5 MeV each, emitted from a state of 𝐞𝑥 = 10 MeV. Beyond this multiplicity dependence, regular kinematics experiments are characterized by narrow sum peaks, which allow the efficiency to be approximated for a single excitation energy. However, in this work, the sum peak is significantly broader, as discussed in Sec.5.3.2. The Doppler shift corrections were applied assuming the angle at the center of each segment, which is an approximation given the large angular coverage of each segment. Beyond any incomplete Doppler broadening corrections, substantial energy straggling also occurs as the beam passes through the target window foil. This results in sum peak widths around 2 MeV, as shown in Fig.5.7 and 5.10. Consequently, the efficiency must be calculated as a function of all energies contributing to the sum peak, while also accounting for the multiplicity of the cascades. For this reason, the detection efficiency is determined through a series of simulations, as outlined in the following paragraphs. 5.5.1 Rainier Simulations The first step in calculating the detection efficiency is to simulate the 𝛟-ray deexcitation of the compound nucleus for all possible excitation energies populated during the reactions. The 𝑄 values for the reactions of interest are 8.55 MeV for 73As(𝑝, 𝛟)74Se and 5.77 MeV for 82Kr(𝑝, 𝛟)83Rb. These values are sufficiently high for the excitation energy 𝐞𝑥 of the compound nucleus to be in the nuclear continuum region, where statistical model calculations are applicable (see Sec. 2.4.2.2). To simulate the 𝛟-ray deexcitation of the compound nucleus, the Rainier code [208] was used. Rainier is a Monte Carlo code that simulates the deexcitation of a compound nucleus using 75 statistical nuclear properties. For these simulations, the nuclear level structure of the compound nucleus is provided as input. The low-energy level schemes for 74Se and 83Rb were taken from Ref. [79], up to approximately 3.3 MeV and 1.8 MeV, respectively, where the level schemes are considered complete. The upper portion of the level scheme can be constructed using the analytical nuclear level density (NLD) models described in Sec. 2.4.2.3, namely the constant temperature (CT) model [31] and the back-shifted Fermi gas (BSFG) model [25]. Rainier also requires a description of the 𝐞𝑥 and 𝐜𝜋 of the entry state. The values of 𝐞𝑥 were considered throughout the range that the experimental sum peak extends. For example in the 3.7 MeV/u beam energy, this corresponds to energies between 9.8 and 11.8 MeV for 74Se, and between 7.0 and 8.8 MeV for 83Rb. For the 𝐜𝜋 of the state, 𝑠-wave proton capture (1/2+) on the ground state of 82Kr (0+) was considered, while for the 74Se compound nucleus, higher order corrections were needed. The 𝐜𝜋 population was obtained from Talys (Sec. 2.4.2.5) by enabling the outdecay and outpopulation options, that output detailed information of the population and statistical decay of the compound nucleus to all possible states. The 𝐜𝜋 distribution used is shown in Fig. 5.12. Figure 5.12 The 𝐜𝜋 distribution of populated excitation energies in the 74Se compound nucleus, calculated using Talys . Once the level scheme is built and the entry state defined, the deexcitation of the nucleus is governed by the 𝛟-ray strength function (𝛟SF ), where a generalized Lorentzian of the form of 76 Kopecky and Uhl [45] (Sec. 2.4.2.4) was adopted. As Rainier is a Monte Carlo code, for each excitation energy component, 10 or 20 realizations of 1000 cascades were calculated for the 74Se and 83Rb compound nucleus, respectively. 5.5.2 Geant4 Simulations The 𝛟 rays obtained by the deexcitation of each contributing 𝐞𝑥 of the compound nucleus through Rainier were then input in Geant4 simulations [209] to account for the detector’s response function. An example of TAS, SoS and multiplicity spectra obtained from Geant4 for the decay a 83Rb compound nucleus at an 𝐞𝑥 = 8 MeV is shown in Fig. 5.13. (a) (b) Figure 5.13 Simulated (a) TAS, (b) SoS and (c) multiplicity spectra of a 𝐞𝑥 = 8 MeV 83Rb state decay using Geant4. (c) 77 5.5.3 Chi-Square Minimization The last step is to determine the contribution of each possible excitation energy in the exper- imental spectra. For this purpose, a 𝜒2 minimization algorithm was implemented [210]. The 𝜒2 code uses the simulated TAS, SoS, and multiplicity spectra, along with the experimental spectra gated on the sum peak, to minimize the following global 𝜒2 value 𝜒2 global = ∑ ∑ 𝑖 𝑗 (cid:169) (cid:173) (cid:173) (cid:171) exp − (cid:205)𝑘 𝑓𝑘𝐶𝑖 𝑗 𝐶𝑖 𝑗 sim 𝐶𝑖 𝑗 exp √ 2 (cid:170) (cid:174) (cid:174) (cid:172) (5.7) where the summations are over 𝑖 types of histograms (TAS, SoS and multiplicity) and 𝑗 number of bins in the 𝑖-th histogram. 𝐶𝑖 𝑗 exp and 𝐶𝑖 𝑗 sim are the counts of the 𝑗-th bin in the 𝑖-th experimental or simulated histogram respectively [210]. The simulated histograms are summed over the 𝑘 different components contributing in the spectra, with 𝑓𝑘 their respective scaling factor. The code utilizes the Minuit algorithm [211] from the Root data analysis toolkit [212], to assign values to the 𝑓𝑘 factors until the minimum global 𝜒2 is achieved. Eq. 5.7 assumes that the error in 𝐶𝑖 𝑗 exp is equal to its square root, which is a valid assumption for spectra that do not carry uncertainty from subtractions. However, in the case of 73As, the error was dominated by statistical uncertainty from the background subtractions (Sec. 5.3.4). Therefore, for the 73As analysis, the denominator in Eq. 5.7 was replaced with the statistical uncertainty shown in Fig. 5.10 for each bin. The detection efficiency is inherently accounted for in the Geant4 simulations, as the detector’s material and geometry are included in the model. Therefore, from the 𝜒2 minimization output, the efficiency can be extracted in the form of a ratio 𝑌 /𝜖, where 𝑌 is the experimental yield. In more detail, based on Eq. 5.7, the 𝜒2 code calculates scaling factors to match the integral of the sum peak with a weighted linear combination of the simulated spectra. For example, consider a sum peak with 100 counts, where two energies contribute to the peak with weights of 60% and 40%. For 10 000 simulated events at each energy, the scaling factors would be 0.006 and 0.004 respectively, so that 0.006 · 10 000 + 0.004 · 10 000 = 100 counts. However, if the detector had a 50% intrinsic efficiency, only 5 000 of the 10 000 simulated events would contribute to the sum 78 peak. In that case, the factors would be 0.012 (= 0.006/0.5) and 0.008 (= 0.004/0.5), so that (0.012 · 10 000 + 0.008 · 10 000) × 0.5 = 100 counts. Based on the example, the efficiency-corrected yield 𝑌 /𝜖, can be expressed as the product of the sum of all scaling factors and the number of simulated events as 𝑌 𝜖 ∑ = 𝑘 𝑓𝑘 · 𝑁sim (5.8) where 𝑓𝑘 represents the scaling factor of the 𝑘-th energy contributing to the peak as calculated through Eq. 5.7, and 𝑁sim the number of simulated events for each energy. T The number of 𝐞𝑥 components simulated for each case reflects the resolution of the sum peak. For 83Rb where the sum peak has significant statistics, ∌20 energy components were simulated, one every 100 keV. For the case of 74Se, where the statistics are small and large energy binning was required (Fig. 5.10), only 5 energy components were simulated, one for every 400 keV. Simulating more than 5 components, accounting for the statistical uncertainty of the spectra, resulted in the minimizer not properly converging. The segmentation of SuN allows to extract valuable information on the statistical properties of the compound nucleus. Specifically, the choice of NLD and 𝛟SF model parameters that are input in Rainier, significantly affects the 𝛟 rays that can be emitted through the deexcitation of a nuclear level in the continuum. This dependence is mostly apparent in the shape of the SoS spectra, and as discussed in detail in the next section, it allowed for the development of an analysis method to constrain the products of NLD and 𝛟SF used in statistical model calculations, and provide predictions for the cross section in a much larger energy range than the experimentally measured. 5.6 Theoretical Investigation with Rainier and Talys The shape of the SoS spectrum depends on the product of the NLD and 𝛟SF . The theoretical models for the NLD (Sec. 2.4.2.3) and 𝛟SF (Se. 2.4.2.4) include multiple adjustable parameters. The default parameters for the constant temperature (CT) and back-shifted Fermi gas (BSFG) models for the NLD, as well as the Generalized Lorentzian for the 𝛟SF found in literature [79, 213, 62], do not appear to reproduce the shape of the SoS spectra, as shown in Fig.5.14. In particular, the simulations overestimate high-energy 𝛟 rays while, in the case of 83Rb, underestimating low-energy 79 (a) (b) Figure 5.14 The 𝜒2 minimization SoS fits for the (a) 82Kr(𝑝, 𝛟)83Rb and (b) 73As(𝑝, 𝛟)74Se reactions at initial beam energy 3.7 MeV/u. The black lines are the experimental spectra and the red and blue lines correspond to the default initial parameters of the BSFG and CT model NLD from Ref. [213] and 𝛟 SF from Ref. [45, 62]. (Figure (a) adapted with permission from Tsantiri et al., Phys. Rev. C 107, 035808 2023. Copyright 2023 by the American Physical Society) ones. These discrepancies could stem from several factors. One possibility is that, for the same 𝛟SF , a higher NLD is needed to reproduce the spectra. In that case, instead of emitting fewer high-energy 𝛟 rays, the compound nucleus would emit multiple lower-energy ones, de-exciting in smaller steps. Alternatively, for the same NLD, the 𝛟SF may assign a lower probability to emitting low-energy 𝛟 rays. An enhancement in the lower-energy region of the 𝛟SF could then resolve the discrepancy, especially in the case of 83Rb. Since this method is sensitive only to the product of the NLD and 𝛟SF , rather than their individual values, this analysis focuses on identifying suitable combinations of these two quantities that reproduce the experimental spectra, and not absolute values for the two quantities independently. 5.6.1 Constraining the Statistical Properties of 83Rb For the 82Kr analysis, five NLD and 𝛟SF combinations were identified as listed in Table 5.4. The first two rows show the literature values for default parameters shown in Fig. 5.14(a). In the first three combinations, the NLD is modified using different parameterizations of the CT and BSFG models (Eqs. 2.17 and 2.15), while the 𝛟SF follows the Generalized Lorentzian model of Kopecky and Uhl [45]. Specifically combination number 2 includes CT parameters obtained from 80 Ref. [214]. In the last two combinations the NLD is considered to follow the default BSFG model parameters, while an upbend in the M1 strength function was implemented, following Eq. 2.20. In combination number 4 the upbend follows data from Ref. [215]. Simulations using those five different combinations create the bands shown in Fig. 5.15 for the TAS, SoS, and multiplicity spectra. Table 5.4 Parameters for modeling the NLD and 𝛟 SF of the 83Rb nucleus. The default parameters in the first two rows are shown in Fig. 5.14(a) with a red and blue lines. The rest of the parameters were chosen for the calculation of the ratio 𝑌 /𝜖, and form the band shown in Fig. 5.15. See text for details on parameters. (Table with permission from Tsantiri et al., Phys. Rev. C 107, 035808 2023. Copyright 2023 by the American Physical Society) NLD Model CT default BSFG default 1. 2. 3. 4. 5. CT CT BSFG BSFG BSFG NLD Model Details Upbend in 𝛟 SF [213] [213] [214] T = 0.824 𝐞0 = -1.16 𝛌 = 10.17 Δ = - 0.54 T = 0.824 𝐞0 = -2.2 T = 0.861 𝐞0 = -3.34 𝛌 = 10.17 Δ = -1.6 𝛌 = 10.17 Δ = -0.54 𝛌 = 10.17 Δ = -0.54 No No No No No [215] a = 1.5 c = 8.7 × 10−8 a = 1.0 c = 1.0 × 10−7 5.6.2 Constraining the Statistical Properties of 74Se There are infinite possible combinations of NLD and 𝛟SF that produce the same transmission coefficient (Eq.2.13). In the previous section, only five suitable combinations were identified. Given the large number of free parameters in this problem, the 74Se analysis aimed not just to find a set of possible solutions, but to systematically characterize the different model combinations available in Talys . As seen in Fig. 3.9, the variations in NLD, 𝛟SF and pOMP result in a cross-section uncertainty of a factor of six at a center-of-mass energy of approximately 3 MeV, which is the 81 (a) (b) (c) Figure 5.15 The 𝜒2 minimization fits for the SoS (a), TAS (b) and multiplicity (c) for the 82Kr(p,𝛟 )83Rb reaction at an initial beam energy 3.7 MeV/nucleon. The black lines are the experimental spectra and the light blue bands indicate the simulated spectra for the combinations of the NLD and 𝛟 SF models listed in Table 5.4. In (a) the red and blue lines are the same as in Fig. 5.14(a). (Figure adapted with permission from Tsantiri et al., Phys. Rev. C 107, 035808 2023. Copyright 2023 by the American Physical Society) energy region relevant for the production of 74Se in the 𝛟 process. Thus, aside from extracting the detection efficiency, the goal of this analysis was to identify suitable NLD and 𝛟SF combinations within the available Talys models, in an attempt to the constrain the cross section of the reaction even further. As shown in Tables 2.1 and 2.2, there are six models available to describe the NLD, nine for the 𝐞1 SF, three for the 𝑀1, and an option to include an 𝑀1 upbend or not. This produces 324 82 possible combinations. Additionally, there two main models for the 𝑝OMP (Sec. 2.4.2.5), and while fitting the experimental spectra is only sensitive to the NLD and 𝛟SF , these will be included in reproducing the experimental cross section in the next chapter, resulting in 648 total possible combinations. Talys is able to output the NLD and 𝛟SF used in each calculation, and therefore those tables were directly input in Rainier. The ability of each Talys model to reproduce the experimental data can be expressed by adapting the concept of a likelihood function from Bayesian analysis [216] as (cid:34) 𝑃(𝑌 |𝑚) ∝ exp − 𝑁 ∑ 𝑗 (𝑊 𝑗 − 𝑓𝑚 (𝑥 𝑗 ))2 2𝜎2 𝑗 (cid:35) (5.9) where 𝑃(𝑌 /𝑚) represents the likelihood of model 𝑚 to reproduce the data 𝑌 , the summation goes over the 𝑁 data-points 𝑊 𝑗 of the dataset 𝑌 , 𝜎𝑗 is the uncertainty of 𝑊 𝑗 , and 𝑓𝑚 (𝑥 𝑗 ) is the prediction of model 𝑚 at value 𝑥 𝑗 . This is often expressed in terms of the log-likelihood log 𝑃(𝑌 |𝑚). The log-likelihood was applied in this analysis as log 𝑃(𝑌 |𝑚) ∝ − 𝑁 𝑏𝑖𝑛𝑠 ∑ 𝑗 (𝐶 𝑗 exp − (cid:205)𝑘 𝑓𝑘𝐶 𝑗 sim(𝑚))2 2𝜎2 𝑗 = − 𝜒2 2 (5.10) where for each Talys model 𝑚, the three histograms 𝑌 (TAS, SoS, and multiplicity) were compared with the simulated ones on a bin-by-bin basis. The values 𝐶 𝑗 exp were the experimental counts per bin 𝑗 and 𝜎𝑗 was the corresponding statistical uncertainty from the background subtraction. The simulated histograms, as described in Sec. 5.5.3, were a linear combination of all energy components weighted by 𝑓𝑘 as calculated in Eq. 5.7. As each of the three types of histograms is treated individually the first summation of Eq. 5.7 is omitted. A few things should be considered in this analysis. Firstly, treating the TAS, SoS and multiplicity spectra independently is an assumption. A more accurate approach would require including correlations between them, but this is beyond the scope of this thesis. Additionally, the log- likelihood obtained for each of the three types of spectra was normalized to the number of data points (bins) in the respective spectrum, 𝑁bins. This normalization ensures that each type of spectrum contributes equally to the overall likelihood, independently of how many bins it contains 83 [217]. The “score" derived by the modified log-likelihood for each type of spectrum is Score = exp (cid:19) (cid:18) log 𝑃 𝑁bins = exp (cid:19) (cid:18) − 𝜒2 2𝑁bins (5.11) The simulated spectra obtained from the various Talys model combinations for initial beam energy of 3.7 MeV/u, are shown in Fig. 5.16, color-coded based on their score, exp(log 𝑃/𝑁bins). For the (a) (b) (c) (d) Figure 5.16 The 𝜒2 minimization fits for the TAS (a), multiplicity (b), and SoS (c) for the 73As(p,𝛟 )74Se reaction at an initial beam energy 3.7 MeV/nucleon. The black lines represent the experimental spectra and the various green lines correspond to the simulated spectra for the combinations of the NLD and 𝛟 SF models from Talys . The varying shades of green in each line reflect different scores from Eq. 5.11. Darker tones represent higher scores, as shown by the color bar (d). 3.1 MeV/u datapoint, due to the low statistics, the 𝑌 /𝜖 was extracted using the NLD and 𝛟SF model 84 Figure 5.17 Beam energy distributions calculated using Srim for 73As at an initial beam energy of 3.7 MeV/u. The blue distribution represents the beam energy after the Mo entrance window, and the red to energy at the end of the cell. The mean of each distribution was used to calculate the effective energy, while the maximum and mean-3𝜎 values determined the upper and lower uncertainty, respectively. combination that produces the lowest 𝜒2 for the 3.7 MeV/u data. No further investigation was performed for that energy. 5.7 Uncertainty Quantification The following describes the uncertainties that contribute to the cross-section calculation for all the quantities discussed in the previous sections. The uncertainty in the effective energy 𝐞eff is mainly attributed to energy straggling as the beam passes through the Mo foil and hydrogen gas. The beam’s energy distribution was calculated using Srim[218]. Fig 5.17 shows an example distribution for 73As at an initial beam energy of 3.7 MeV/u. The blue distribution represents the energy straggling after the 2-𝜇m-thick Mo entrance window, while the red corresponds to the energy at the end of the 4-cm-long gas cell filled with 600 torr of hydrogen. The mean of each distribution was used as 𝐞0 and 𝐞0 − Δ𝐞 in Eq. 5.3. The upper and 85 lower uncertainties in 𝐞eff were determined using the maximum and mean-3𝜎 of each distribution, respectively. Due to the asymmetrical energy straggling distribution, which has a long low-energy tail, the errors in the effective energy are similarly asymmetric. An additional 1% uncertainty was included to reflect the uncertainty in the delivered beam by the facility. The uncertainty in the total number of beam particles 𝐌𝛌 (Sec. 5.2) in the 82Kr measurement was considered to be 5% from the beam-charge accumulation. For the 73As measurement in addition to the 5% uncertainty for the ammeter measurement, there were two methods used for the integration of the current shown in Fig. 5.2, introducing an additional 2% uncertainty, as well as a 6% uncertainty from the baseline fits. Overall the uncertainty in the beam current for the 73As measurement was 8.3%. The target particle density 𝑁𝑡 was considered to carry an uncertainty of 5%. This includes systematic uncertainty from the calibration of the manometer to the atmospheric pressure and the zero offset, as well as the random error from the instrument resolution. The uncertainty in the 𝑌 /𝜖 for the 82Kr measurement includes statistical uncertainty varying between 1% and 4%, along with uncertainty from the various parameters chosen for the NLD and 𝛟SF models varying between 26% and 17%, with the latter value corresponding to the smaller energy. For the 73As measurement the uncertainty is dominated by the statistical uncertainty from the background subtraction varying between 18% and 72%, for the larger and smaller beam energies, respectively. An additional 6% uncertainty was assumed for the 2𝜎 deviation of the scaling factors distribution for the empty cell measurement (Fig. 5.9(b)). Finally the various Talys combinations produced a range of 𝑌 /𝜖 values. In Fig. 5.18, for the 3.7 MeV/u measurement, the various 𝑌 /𝜖 values produced by all the different fits are shown compared to their total fitting score, defined as the product of all three scores from Eq. 5.11. The higher scores that reproduce the best fits, converge near 𝑌 /𝜖 ≈ 6000. By weighing each 𝑌 /𝜖 with the score of the corresponding model, the distribution shown in Fig. 5.19 was obtained. An uncertainty of 13.5% was obtained from the 3𝜎 deviation of that distribution. The overall uncertainty for 𝑌 /𝜖 of the 73As measurement was 24% and 74% for the 3.7 and 3.1 MeV/u beam energies. 86 Figure 5.18 The 𝑌 /𝜖 values produced using the various available Talys combinations of NLD ang 𝛟SF compared to the fitting score, defined as the product of all three exp(log 𝑃/𝑁bins) for the three types of histograms (TAS, SoS and multiplicity). The fits with the highest score converge around 𝑌 /𝜖 ≈ 6000. Figure 5.19 The distribution of 𝑌 /𝜖 values shown in Fig. 5.18 weighted by their respective score. The value of 𝑌 /𝜖 used in the analysis was obtained from the best fit (dashed line), and the 3𝜎 uncertainty of the distribution was assumed as model uncertainty for this measurement (dot-dashed line). 87 CHAPTER 6 RESULTS & DISCUSSION 6.1 The 82Kr(𝑝, 𝛟)83Rb cross section The cross section for the 82Kr(𝑝, 𝛟)83Rb reaction as calculated using Eq. 5.1 is presented in Table 6.1 along with all calculated quantities described in Ch. 5. In Fig. 6.1 the cross section is shown along with statistical model calculations using the Non-Smoker code [174] and Talys , for all available NLD and 𝛟SF models. Figure 6.1 The measured cross section of the 82Kr(𝑝, 𝛟)83Rb reaction (black dots) compared with standard Non-Smoker theoretical calculations [174] (blue solid line) and default Talys 1.96 calculations [62] (orange band). (Figure with permission from Tsantiri et al, Phys. Rev. C 107, 035808 2023. Copyright 2023 by the American Physical Society) 88 Table 6.1 Measured cross section of the 82Kr(𝑝, 𝛟)83Rb reaction. The first column represents the initial beam energy in the lab system, and the second column the center-of-mass effective energy of the reaction. The third and fourth columns represent the total number of incident and target particles, and the fifth the total number of reactions that occurred 𝑌 /𝜖. The last two columns represent the efficiency in detecting 𝛟 -rays from the de-excitation of the 83Rb compound nucleus, and the measured cross section. (Table with permission from Tsantiri et al., Phys. Rev. C 107, 035808 2023. Copyright 2023 by the American Physical Society) Initial Beam Energy (MeV/u) 3.7 3.4 3.1 Effective Energy 𝐞𝑒 𝑓 𝑓 (MeV) 2.99+0.03 −0.06 2.68+0.03 −0.06 2.38+0.02 −0.09 Total Incident Particles 𝐌𝛌 (2.15 ± 0.11)×1011 (2.05 ± 0.10)×1011 (2.07 ± 0.10)×1011 Total Target Particles 𝑁𝑡 (7.44 ± 0.39)×1019 (7.43 ± 0.39)×1019 (7.39 ± 0.39)×1019 Number of Reactions 𝑌 /𝜖 26079 ± 6707 11011 ± 2460 3578 ± 595 Efficiency 𝜖 (%) 51.6 ± 5.6 51.3 ± 5.7 52.6 ± 6.0 Cross Section 𝜎 (mb) 1.63 ± 0.40 0.72 ± 0.16 0.23 ± 0.04 89 The comparison in Fig. 6.1 shows that standard statistical model calculations with default model parameters generally overestimate the cross section for the 82Kr(𝑝, 𝛟)83Rb reaction relative to the experimental measurements. The discrepancies between the experimental data and predictions from the Non-Smoker code range from 23% to 47%, with the largest deviation observed at the lowest beam energy. A similar trend has been reported by Lotay et al. [173] for the (𝑝, 𝛟) reaction on the neighboring 83Rb nucleus, as well as by GyÃŒrky et al. [219] on various proton-rich Sr isotopes. In both instances, the experimental cross sections were systematically lower than those predicted by the Hauser-Feshbach (HF) theory, indicating the existence of a trend in this mass region. This substantial overestimation by theoretical models motivated further investigation on the model parameters used in Talys . As discussed in Sec. 5.6.1, specific parameters were identified to model the NLD and 𝛟 SF of the 83Rb nucleus (Table 5.4). These parameters were selected to reproduce the experimental spectra, and therefore, should provide a more accurate representation of the calculated cross section. Interestingly, the various combinations of NLD and 𝛟SF primarily impact the upper range of the calculated 82Kr(𝑝, 𝛟)83Rb cross sections, while the lower energy range near the 3.1 MeV/u measurement appears to not be as sensitive. Consequently, a better description could be obtained by modifying other input quantities in Talys . As discussed in Sec. 2.4.2.2, the cross section in the statistical model includes the transmission 𝑎 and 𝑇𝑌 coefficients 𝑇 𝑋 𝑏 between the 𝑋 + 𝑎 entrance and 𝑌 + 𝑏 exit channel, and the width fluctuation correction 𝑊 𝑎𝑋→𝑌 𝑏 (WFC). At the low energies examined in this experiment, the only available exit channels are the proton and 𝛟 emission. The neutron channel opens just above 5 MeV, and the 𝛌 channel remains negligible throughout the energy range under study, due to the higher Coulomb barrier. Therefore in this case, Eq. 2.12 simplifies to 𝜎( 𝑝, 𝛟) ∌ 𝑇𝑝𝑇𝛟 𝑇𝑝 + 𝑇𝛟 𝑊 𝑝𝛟 (6.1) where the 𝑇𝑝 corresponds to the proton capture and 𝑇𝛟 to the 𝛟 decay. It is significant to examine the competition between the available open channels. Fig. 6.2 shows a decomposition of the total reaction cross section into the different exit channels, expressed in terms of astrophysical 𝑆 factor 90 (Sec. 2.5), for better readability. The dashed lines in Fig. 6.2 correspond to a Talys calculation Figure 6.2 Decomposition of the astrophysical 𝑆 factor for proton capture on 82Kr. Dashed lines correspond to a standard talys calculation, while solid lines use an optimized set of parameters. See text for more details on the optimized parameters. The contribution of the 82Kr(p,𝛌)79Br reaction is below the scale of the figure. (Figure adapted with permission from Tsantiri et al., Phys. Rev. C 107, 035808 2023. Copyright 2023 by the American Physical Society) using the NLD and 𝛟SF parameter set No.5 from Table 5.4. Since these models are fixed, the cross section remains dependent only to the 𝑇𝑝, meaning the proton optical model potential (pOMP), as well as the WFC. As discussed in Sec. 2.4.2.5, another option provided in Talys for the description of the pOMP is the “jlm-type" potential, based on the work of Jeukenne, Lejeunne, and Mahaux [64, 65, 66, 67] with later modifications by Bauge et al. [68, 69]). This potential predicts a slightly lower cross section at lower energies, providing a better agreement with the experimental data. The WFC accounts for correlations between the incident and outgoing wave functions, which typically enhance the compound-elastic channel. The WFC becomes most significant when only a few channels are open, which is the case for lower energies. At higher energies, where many channels are accessible, the relevance of the WFC becomes negligible. The default WFC used in Talys is based on Moldauer’s formalism [70, 71] (widthmode 1). An alternative approach is 91 that of Hofmann, Richert, Tepel, and WeidenmÃŒller (HRTW, widthmode 2) [72, 73, 74]. This approach provides a stronger correction, enhancing the compound-elastic channel and reducing the calculated (𝑝, 𝛟) cross section. Combining a HWRT-based WFC with a “jlm-type" potential yields the solid lines in Fig. 6.2, showing an improved agreement with the experimental data, particularly at lower energies. This combination along with the different NLD and 𝛟SF models of Table 5.4 produce the teal band in Fig. 6.3. Figure 6.3 Same as Fig. 6.1, with the addition of the optimized Talys band, resulting from the NLD and 𝛟SF combinations of Table 6.1 along with a HWRT-based WFC and a “jlm-type" potential. (Figure with permission from Tsantiri et al., Phys. Rev. C 107, 035808 2023. Copyright 2023 by the American Physical Society) This result demonstrates the successful proof-of-principle implementation of the summing tech- nique in inverse kinematics using this experimental setup and a gas cell target for the measurement of (𝑝, 𝛟) reactions. Additionally, this study shows that a consistent description of the cross section, SoS, TAS, and multiplicity spectra can be achieved through a careful choice of parameters in the statistical model. The ability to simultaneously describe multiple observables provides stronger 92 constraints on the model parameters than a comparison with cross-section data alone, enabling a more constrained cross section over a broader energy range than the one directly measured. 6.2 The 73As(𝑝, 𝛟)74Se cross section Similarly to the previous section, the cross section for the 73As(𝑝, 𝛟)74Se reaction as calculated using Eq. 5.1 is presented in Table 6.2 along with all calculated quantities. Due to the low statistics of the low-energy data point, the measurement efficiency was assumed to be the value calculated for the 3.7 MeV/u data point for both energies. The asymmetry in the errors of 𝑌 /𝜖 results from avoiding negative counts in specific bins. In Fig. 6.4 the cross section is shown along with statistical model calculations using the Non-Smoker code [174] and Talys , for all available NLD and 𝛟SF models, and the two pOMP options (default and JLM). Figure 6.4 The measured cross section of the 73As(𝑝, 𝛟)74Se reaction (black dots) compared with standard Non-Smoker theoretical calculations [174] shown in a solid line and default Talys 1.96 calculations [62] creating a band. 93 Table 6.2 Same as Table 6.1 for the 73As(𝑝, 𝛟)74Se reaction measurement. Initial Beam Energy (MeV/u) 3.7 3.1 Effective Energy 𝐞𝑒 𝑓 𝑓 (MeV) 2.92+0.04 −0.05 2.32+0.03 −0.05 Total Incident Particles 𝐌𝛌 (2.51 ± 0.21)×1010 (1.67 ± 0.14)×109 Total Target Particles 𝑁𝑡 (7.94 ± 0.40)×1019 (7.94 ± 0.40)×1019 Number of Reactions 𝑌 /𝜖 6193 ± 1480 152+112 −95 Efficiency 𝜖 (%) 48.2 ± 6.5 Cross Section 𝜎 (mb) 3.11 ± 0.80 1.15+0.86 −0.73 94 As discussed in the previous chapters, the challenge of running a radioactive beam experiment is apparent from the large statistical uncertainties compared to the stable beam experiment, especially in the case of the low-energy datapoint, as the measurement was limited to just 30 minutes of data acquisition (see Sec. 5.3.4). Nevertheless, this is the first experimental data provided on the 73As(𝑝, 𝛟)74Se reaction cross section, and as the measurement is performed within the Gamow window window for the astrophysical 𝛟 process, it is particularly relevant for calculating the production of 74Se. The calculated cross section is in agreement with the theoretical prediction by Non-Smoker within the experimental uncertainty, however the higher energy measurement is ∌18% higher than the Non-Smoker calculation. As with the 82Kr analysis, the 73As(𝑝, 𝛟)74Se cross section measured can be combined with the description of the TAS, SoS and multiplicity spectra obtained in Sec. 5.6.2, to provide a constrained Talys band that can be used in network calculations. For this analysis, only the higher-energy data point will be utilized, as the uncertainty of the low-energy point exceeds that of Talys . As shown in Sec. 5.6.2 for each of the various Talys models, the three different types of spectra (TAS, SoS and multiplicity) have been evaluated by a score described by Eq. 5.11. The calculated cross section from Fig. 6.4 is used to evaluate one last factor that contributes to the score. However, the likelihood as defined in Eq. 5.9, includes only uncertainty on the y-axis, whereas the cross section also includes uncertainty in the effective energy. The uncertainty in the effective energy reflects the energy straggling of the beam in the gas cell and hydrogen gas, and that distribution of energies has been calculated using Srim (Fig. 5.17). This distribution can be used to weight the energies on the x-axis as shown in Fig. 6.5. The error energy range is divided into 20 bins, and the likelihood is calculated for each one of the bins weighted by its normalized probability distribution as (cid:34) 𝑃(𝜎exp|𝑚) = exp − 20 ∑ 𝑗 𝑀 𝑗 (𝜎exp − 𝑓 (𝐞 𝑗 , 𝑚))2 2 · 𝛿𝜎2 exp (cid:35) (6.2) where the summation goes over the 20 energy bins with weight 𝑀 𝑗 shown in Fig. 6.5, 𝜎exp and 𝛿𝜎exp is the experimental cross section with its uncertainty from Table 6.2, and 𝑓 (𝐞 𝑗 , 𝑚) is the predicted cross section from Talys model 𝑚 at the central energy 𝐞 𝑗 of each bin. Fig. 6.6(top) 95 Figure 6.5 The effective energy uncertainty of the 73As(𝑝, 𝛟)74Se reaction as a function of the energy straggling distributions from Fig. 5.17. shows the calculated cross section with the Talys models shown in Fig 6.4, color-coded based on the likelihood calculated through Eq. 6.2. Combining this likelihood with the total score for the TAS, SoS and multiplicity spectra from Sec. 5.6.2, produces the combined score shown in Fig. 6.6(bottom). One can observe that this combination leads to a significantly narrower favored cross-section band. A complete table of the scores for all Talys models is shown in the Appendix. This distribution is used to constrain the production of 74Se in network calculations of the 𝛟 process in the following chapter. It can be seen that the favored distribution is located slightly higher than the Non-Smoker calculation shown in red, which is the reaction adopted in astrophysical calculations, even though this deviation is within the experimental uncertainty. This is consistent with the experimental measurement being ∌18% higher than the one calculated by Non-Smoker. 96 Figure 6.6 The measured cross section of the 73As(𝑝, 𝛟)74Se reaction in the 3.7 MeV/u initial beam energy compared to Talys calculations color-coded based on (top) the likelihood calculated through Eq. 6.2 and (bottom) the combined score to reproduce all types of data (cross section, TAS, SoS and multiplicity). 97 CHAPTER 7 ASTROPHYSICAL IMPACT As discussed in Sec. 3.4.1, the two main explosive environments where the 𝛟 process is thought to occur are the oxygen and neon enriched layers of a core-collapse supernovae (SNII) [133, 135, 136], and thermonuclear Type Ia supernovae (SNIa) [137, 138, 139]. The impact of the measured cross section will be examined in both cases using Monte Carlo simulations of both scenarios. 7.1 Core-Collapse Supernova - SNII For the SNII environment, NuGrid stellar data set II models [142] were used. The most massive progenitors with solar metallicity available in these models were stars with masses of 15, 20, and 25 𝑀⊙, calculated using MESA [157]. The initial solar composition was from Grevesse and Noels [220] with isotopic ratios from Lodders [126], while the nucleosynthesis during stellar evolution and explosion was computed using the Multi-zone Post-Processing Network – Parallel (MPPNP) code [159]. The nuclear network beyond iron consisted of 74 313 reactions, for which experimental rates from the KADoNIS compilation [221] were incorporated wherever available, and any missing reaction rates were obtained from the JINA REACLIB library [222]. Among the three available mass models, the 25 𝑀⊙ model ended in a failed supernova due to significant fallback, trapping all the material produced during the explosion in the remnant star. The 15 𝑀⊙ model experienced a C-O shell merger event, significantly enhancing the pre-supernova production of p nuclei. Since 74Se is expected to be primarily produced by the explosive component, even in the presence of a C-O shell merger [143], the 20 𝑀⊙ model was chosen as a representative case, where the explosive component dominates p-nuclei production, similarly to other SNII models used in 𝛟-process studies [136, 176, 177]. The impact of the measured cross section was expected to be similar for any SNII model reaching comparable peak temperatures during the shock wave propagation. The produced mass fraction of 74Se as the shock wave propagates through the ONe layer of the star is shown in Fig. 7.1. The maximum production of 74Se occurs in the inner ONe layer, at mass coordinate 𝑀 = 2.93𝑀⊙. At this location, the temperature and density profiles are presented in 98 Figure 7.1 The final mass fractions of the 20 𝑀⊙ mass model by Ritter et al. [142] as a function of mass coordinate. The 16O, 20Ne, 12C, 4He and 1H lines indicate the different layers of the star. 74Se is produced in the inner ONe layer, where the higher peak temperatures in the Gamow window for the 𝛟 process are achieved. Fig. 7.2, where the temperature trajectory exhibits a plateau at 𝑇 = 3.08 GK. To investigate the impact of the measured cross section under such conditions, the cross section predictions of the measured 73As(𝑝, 𝛟)74Se reaction from Fig. 6.6 were converted to reaction rates In Fig. 7.3 the reaction rate of the 73As(𝑝, 𝛟)74Se reaction is shown as a function of using Talys . temperature, with models color-coded by the total score calculated in Sec. 6.2. The solid red line is the Jina-Reaclib [222] reaction rate, while the dotted line marks the temperature of interest, 𝑇 = 3.08 GK. A projection of the total scores along the 𝑇 = 3.08 GK temperature is shown in Fig. 7.4, as a function of each model’s total score. The various points represent the different Talys models, and the blue bars are the normalized distribution of those rates grouped in 50 bins, between the minimum and maximum Talys reaction rate. For Monte Carlo simulations, 10 000 rates were randomly sampled from this distribution, forming the set of rates shown by the magenta line. The Monte Carlo one-zone nucleosynthesis simulations were performed using the PPN code 99 (a) (b) Figure 7.2 The temperature and density trajectories at mass coordinate 𝑀 = 2.93𝑀⊙ in the 20 𝑀⊙ mass model by Ritter et al. [142]. The arrows in (b) indicate the direction of time. Figure 7.3 The reaction rate of the 73As(𝑝, 𝛟)74Se reaction as a function of temperature calculated using Talys . The Talys models are color coded based on their total score, similarly to Fig. 6.2. The red line indicates the reaction rate of the Jina-Reaclib [222] library, typically used in network calculations. The dotted line represents the peak temperature for the SNII model from Fig. 7.2a. from the NuGrid framework [159]. The temperature and density profiles (Fig. 7.2), as well as the initial abundances were obtained from the 20 𝑀⊙ mass model shown in Fig. 7.1 for the mass 100 Figure 7.4 A projection of the reaction rates calculated by Talys along the 𝑇 = 3.08 GK temperature of interest, as a function of each model’s total score. The blue bars correspond to the Talys models grouped in 50 bins between the minimum and maximum rate calculated by Talys . The magenta line shows the sampled rates used in Monte Carlo simulations of the SNII scenario. coordinate 𝑀 = 2.93𝑀⊙. The only variable varied in each of the 10 000 realizations of the code was the multiplication factor for the 73As(𝑝, 𝛟)74Se reaction rate (and its inverse 74Se(𝛟, 𝑝)73As). This multiplication factor was calculated as the ratio of the sampled reaction rate over the Jina-Reaclib rate that is used as a default value. The produced 74Se mass fraction as a function of time is shown in Fig. 7.5. The various blue lines correspond to the 10 000 runs with varied 73As(𝑝, 𝛟)74Se reaction rate, with the darker colors reflecting more lines that overlap in that region. The orange line corresponds to the mass fraction produced using the Jina-Reaclib rate, and the red dashed line are the mass fractions obtained with the minimum and maximum rate obtained from Talys . Fig. 7.6 shows final mass fraction of 74Se relative to the default Jina-Reaclib model. The experimentally measured 73As(𝑝, 𝛟)74Se cross section from Sec. 6.2 is ∌18% higher than the Non-Smoker rate adopted in the Jina-Reaclib library but still consistent within the 101 Figure 7.5 74Se mass fraction as a function of time calculated using PPN one-zone simulations. The various blue lines correspond to different 73As(𝑝, 𝛟)74Se reaction rates sampled from the distribution of Fig. 7.4, and the darker color correspond to more overlapping trajectories. The orange line corresponds to simulation using the default Jina-Reaclib reaction rate and the red dashed lines to the minimum and maximum reaction rates by Talys . experimental uncertainty (Fig. 6.4). Consequently, the final 74Se abundance in SNII models remains in good agreement with calculations using the Jina-Reaclib rate, though a slightly lower mean 74Se production is suggested. However, this deviation is too small to indicate that the observed overproduction of 74Se is driven by uncertainties in this reaction rate. Comparing the full width at half maximum of the distribution of Fig. 7.8 to the Talys uncertainty shows that the measurement significantly reduces theoretical uncertainties in reaction rates by approximately a factor of two, providing a more constrained and reliable input for future sensitivity studies. 102 Figure 7.6 Comparison of the final 74Se mass fraction distribution obtained in the Monte Carlo simulations varying the 73As(𝑝, 𝛟)74Se reaction rate relative to the default model using the Jina- Reaclib reaction rate. The dashed lines correspond to the maximum and minimum reaction rate from Talys . 7.2 Type Ia Supernova - SNIa For the Type Ia scenario the models by Travaglio et al. [138] were adopted. In their work, they explored two-dimensional SNIa models of a white dwarf (WD) accreting mass from a companion star in the single degenerate (SD) scenario. The s-process seed distribution is assumed to be produced during the AGB phase leading to the formation of the WD, through thermal pulses that mix material from the H-rich envelope to the C-rich layers, resulting in a 13C pocket (see Sec. 3.3.3.1). From the different explosion mechanisms studied by Travaglio et al., the model adopted in this work is the so-called DTT-a. DTT-a is a delayed detonation model assuming the deflagration-to-detonation criterion of Kasen et al. [223] and a CO-WD structure by Domínguez et al. [224] with solar metallicity and a progenitor mass of M = 1.5 𝑀⊙. The nucleosynthesis in such multidimensional simulations was calculated using the tracer parti- cle method [225, 226] by placing 51 200 tracer particles, uniformly distributed in mass coordinate. 103 For this work, the trajectories of three tracer particles were provided from the external layers of the model, where the peak temperatures reached allowed for 𝛟-process nucleosynthesis. The isotopic abundances of elements heavier than Ge were also provided, while for the lighter elements solar abundances by Asplund [83] were used. The temperature and density profiles for one of the three tracers (number 2876) is shown as an example in Fig. 7.7. The other two tracer trajectories had minimal differences, an thus the tracer shown in Fig. 7.7 was chosen as a representative case. (a) (b) Figure 7.7 The temperature (a) and density (b) profiles of tracer 2876 of the DTT-a SNIa model by Travaglio et al. [138]. The tracer is located on the external layers of the SNIa model, where the peak temperatures reached allow for 𝛟-process nucleosynthesis. The peak temperature obtained is 𝑇 = 3.27 GK. Similar to the SNII scenario one-zone nucle- osynthesis simulations were performed using the PPN code for the Jina-Reaclib rate, the minimum and maximum rate predicted by Talys models at the peak temperature of interest, and a sample of the reaction rate distribution weighted by the total score. The produced 74Se mass fraction as a function of time, shown in Fig. 7.8, confirms that variations in the 73As(𝑝, 𝛟)74Se reaction rate within the nuclear uncertainty predicted by Talys do not significantly impact the final 74Se abundance. This is consistent with Ref. [162], indicating that 74Se production it SNIa is affected by astrophysical conditions rather than nuclear uncertainties in this reaction. As the 73As(𝑝, 𝛟)74Se 104 Figure 7.8 Same as Fig. 7.5 for the SNIa trajectories shown in Fig. 7.7. reaction is a destruction mechanism for 74Se, it is possible that the material is not exposed to the peak temperature for a long enough time, for this destruction mechanism to become relevant. As seen in Fig. 7.7 compared to Fig. 7.2, the material is exposed to peak temperatures for fractions of a second, while in the SNII scenario the temperature exhibits a plateau that lasts for multiple seconds, allowing the 73As(𝑝, 𝛟)74Se reaction to destroy the produced 74Se and impact the final abundances. 105 CHAPTER 8 SUMMARY & CONCLUSIONS To summarize, the astrophysical 𝛟 process, the most established scenario for the production of the p nuclei, involves a vast network of primarily radioactive isotopes. Developing experimental techniques to directly measure photodisintegration reactions using unstable beams is essential, as reaction rates in network calculations are largely estimated through statistical model calculations, that carry significant uncertainties away from stability. This work focused on implementing such a technique to measure reaction cross sections in inverse kinematics using a radioactive beam, the summing technique and a hydrogen gas target. The proof-of-principle application of this method was successfully demonstrated with the first measurement of the 82Kr(𝑝, 𝛟)83Rb reaction cross section. The experimental results suggest a smaller cross section than theoretically predicted, but a more consistent description of the statistical model parameters was obtained through simulations to reproduce the experiment spectra. The ability to simultaneously describe multiple observables provides stronger constraints on the model parameters than a comparison with cross-section data alone, enabling a more constrained cross section over a broader energy range than the one directly measured. The same methodology was applied using a radioactive 73As beam, leading to the first mea- surement of the 73As(𝑝, 𝛟)74Se reaction cross section. The measured cross section showed good agreement with the theoretical prediction from Non-Smoker. The cross-section data, along with experimental spectra, were used to characterize various nuclear level density and 𝛟-ray strength function models from Talys , allowing for the extraction of an experimentally constrained cross section across the entire Gamow window of the 𝛟 process. This characterization enabled the deter- mination of an experimentally constrained reaction rate for the 73As(𝑝, 𝛟)74Se reaction, which was then used in Monte Carlo one-zone network simulations of the 𝛟 process. Although the production of 74Se in the Type Ia supernova scenario showed no sensitivity to the constrained reaction rate, a significant impact on the final 74Se abundance in Type II supernovae was observed. The uncertainty in the 74Se production was reduced by a factor of two, providing a 106 more constrained and reliable input for future sensitivity studies. The successful implementation of this technique opens the possibility of extending similar mea- surements to reactions involving heavier radioactive beams. Such studies would further constrain reaction rates relevant for the 𝛟 process, improving the accuracy of nucleosynthesis models and reducing uncertainties in astrophysical simulations. 107 BIBLIOGRAPHY F. W. Aston, “Nobel Lecture,” 1922. Accessed from https://www.nobelprize.org/prizes/ chemistry/1922/aston/lecture/. A. S. Eddington, “The Internal Constitution of the Stars,” The Scientific Monthly, vol. 11, no. 4, pp. 297–303, 1920. E. M. Burbidge, G. R. Burbidge, W. A. Fowler, and F. Hoyle, “Synthesis of the Elements in Stars,” Rev. Mod. Phys., vol. 29, pp. 547–650, 1957. IAEA, nds.iaea.org/relnsd/vcharthtml/VChartHTML.html. Nuclides.” Chart “Live of Accessed from https://www- [1] [2] [3] [4] [5] M. G. Mayer, “On Closed Shells in Nuclei,” Phys. Rev., vol. 74, pp. 235–239, 1948. [6] [7] C. Iliadis, Nuclear Physics of Stars. John Wiley & Sons, Ltd, 2007. K. S. Krane, Introductory Nuclear Physics. John Wiley & Sons, Inc, 1987. [8] Wikimedia, “Binding Energy Curve,” 2022. This work has been released into the public domain by urlhttps://en.m.wikipedia.org/wiki/File:Binding_energy_curve_- _common_isotopes.svg . [9] M. Wang, G. Audi, A. Wapstra, F. Kondev, M. MacCormick, X. Xu, and B. Pfeiffer, “The Ame2012 atomic mass evaluation,” Chinese Physics C, vol. 36, no. 12, p. 1603, 2012. [10] ΑσηΌακόπουλος, Π. Α., ΠυρηΜική Ίυσική, ch. 9, pp. 259–311. ΙωάΜΜιΜα: Επτάλοφος, 1984. [11] C. A. Bertulani and A. Bonaccorso, "Direct Nuclear Reactions", pp. 1415–1449. Singapore: Springer Nature Singapore, 2023. [12] W. Heisenberg, “Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik,” Zeitschrift fÃŒr Physik, vol. 43, pp. 172–198, 1927. [13] N. Bohr, “Neutron Capture and Nuclear Constitution,” Nature, vol. 137, pp. 344–348, 1936. [14] T. Rauscher, Essentials of Nucleosynthesis and Theoretical Nuclear Astrophysics. 2514- 3433, IOP Publishing, 2020. [15] C. Rolfs and W. Rodney, Cauldrons in the Cosmos: Nuclear Astrophysics. Theoretical Astrophysics, University of Chicago Press, 1988. [16] W. Hauser and H. Feshbach, “The Inelastic Scattering of Neutrons,” Phys. Rev., vol. 87, pp. 366–373, 1952. [17] T. Rauscher and F.-K. Thielemann, “Astrophysical Reaction Rates From Statistical Model Calculations,” Atomic Data and Nuclear Data Tables, vol. 75, no. 1, pp. 1–351, 2000. 108 [18] C. Mahaux and H. A. Weidenmuller, “Recent Developments in Compound-Nucleus Theory,” Annual Review of Nuclear and Particle Science, vol. 29, no. Volume 29, 1979, pp. 1–31, 1979. [19] P. E. Hodgson, “The nuclear optical model,” Reports on Progress in Physics, vol. 34, no. 2, p. 765, 1971. [20] W. Dickhoff and R. Charity, “Recent developments for the optical model of nuclei,” Progress in Particle and Nuclear Physics, vol. 105, pp. 252–299, 2019. [21] A. Larsen, A. Spyrou, S. Liddick, and M. Guttormsen, “Novel techniques for constraining neutron-capture rates relevant for r-process heavy-element nucleosynthesis,” Progress in Particle and Nuclear Physics, vol. 107, pp. 69–108, 2019. [22] T. Rauscher, F.-K. Thielemann, and K.-L. Kratz, “Nuclear level density and the determination of thermonuclear rates for astrophysics,” Phys. Rev. C, vol. 56, pp. 1613–1625, 1997. [23] H. A. Bethe, “An Attempt to Calculate the Number of Energy Levels of a Heavy Nucleus,” Phys. Rev., vol. 50, pp. 332–341, 1936. [24] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, “Theory of Superconductivity,” Phys. Rev., vol. 108, pp. 1175–1204, 1957. [25] A. Gilbert and A. G. W. Cameron, “A composite nuclear-level density formula with shell corrections,” Canadian Journal of Physics, vol. 43, p. 1446, 1965. [26] 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, vol. 217, no. 2, pp. 269–298, 1973. [27] H. Baba, “A shell-model nuclear level density,” Nuclear Physics A, vol. 159, no. 2, pp. 625– 641, 1970. [28] A. V. Ignatyuk, G. N. Smirenkin, and A. S. Tishin, “Phenomenological description of energy dependence of the level density parameter,” Yad. Fiz., v. 21, no. 3, pp. 485-490, 1975. [29] T. von Egidy and D. Bucurescu, “Statistical nuclear properties (level densities, spin distri- butions),” Journal of Physics: Conference Series, vol. 338, no. 1, p. 012028, 2012. [30] S. Goriely, “A new nuclear level density formula including shell and pairing correction in the light of a microscopic model calculation,” Nuclear Physics A, vol. 605, no. 1, pp. 28–60, 1996. [31] T. Ericson, “A statistical analysis of excited nuclear states,” Nuclear Physics, vol. 11, pp. 481– 491, 1959. [32] A. V. Ignatyuk, K. K. Istekov, and G. N. Smirenkin, “Collective effects in level density, and the probability of fission,” Sov. J. Nucl. Phys. (Engl. Transl.); (United States), vol. 30:5, 1979. 109 [33] A. V. Ignatyuk, J. L. Weil, S. Raman, and S. Kahane, “Density of discrete levels in 116Sn,” Phys. Rev. C, vol. 47, pp. 1504–1513, 1993. [34] Y. Alhassid, M. Bonett-Matiz, S. Liu, and H. Nakada, “Direct microscopic calculation of nuclear level densities in the shell model Monte Carlo approach,” Phys. Rev. C, vol. 92, p. 024307, 2015. [35] P. Demetriou and S. Goriely, “Microscopic nuclear level densities for practical applications,” Nuclear Physics A, vol. 695, no. 1, pp. 95–108, 2001. [36] S. Goriely, F. Tondeur, and J. Pearson, “A Hartree–Fock Nuclear Mass Table,” Atomic Data and Nuclear Data Tables, vol. 77, no. 2, pp. 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, vol. 78, p. 064307, 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, vol. 86, p. 064317, 2012. [39] G. A. Bartholomew, E. D. Earle, A. J. Ferguson, J. W. Knowles, and M. A. Lone, Gamma-Ray Strength Functions, pp. 229–324. Boston, MA: Springer US, 1973. [40] D. Brink, Some Aspects of the Interaction of Light with Matter. PhD thesis, Oxford University, 1955. [41] N. L. Iudice and F. Palumbo, “New Isovector Collective Modes in Deformed Nuclei,” Phys. Rev. Lett., vol. 41, pp. 1532–1534, 1978. [42] K. Heyde, P. von Neumann-Cosel, and A. Richter, “Magnetic dipole excitations in nuclei: Elementary modes of nucleonic motion,” Rev. Mod. Phys., vol. 82, pp. 2365–2419, 2010. [43] P. Axel, “Electric Dipole Ground-State Transition Width Strength Function and 7-MeV Photon Interactions,” Phys. Rev., vol. 126, pp. 671–683, 1962. [44] [45] J. Kopecky and R. Chrien, “Observation of the M1 giant resonance by resonance averaging in 106Pd,” Nuclear Physics A, vol. 468, no. 2, pp. 285–300, 1987. J. Kopecky and M. Uhl, “Test of gamma-ray strength functions in nuclear reaction model calculations,” Phys. Rev. C, vol. 41, pp. 1941–1955, 1990. [46] S. Goriely and E. Khan, “Large-scale QRPA calculation of E1-strength and its impact on the neutron capture cross section,” Nuclear Physics A, vol. 706, no. 1, pp. 217–232, 2002. [47] M. Martini, S. Péru, S. Hilaire, S. Goriely, and F. Lechaftois, “Large-scale deformed quasi- particle random-phase approximation calculations of the 𝛟-ray strength function using the Gogny force,” Phys. Rev. C, vol. 94, p. 014304, 2016. [48] S. Goriely, E. Khan, and M. Samyn, “Microscopic HFB + QRPA predictions of dipole strength for astrophysics applications,” Nuclear Physics A, vol. 739, no. 3, pp. 331–352, 2004. 110 [49] S. Goriely, S. Hilaire, S. Péru, and K. Sieja, “Gogny-HFB+QRPA dipole strength func- tion and its application to radiative nucleon capture cross section,” Phys. Rev. C, vol. 98, p. 014327, 2018. [50] D. Vretenar, N. Paar, P. Ring, and G. A. Lalazissis, “Pygmy dipole resonances in the relativistic random phase approximation,” Phys. Rev. C, vol. 63, p. 047301, 2001. [51] E. Litvinova and N. Belov, “Low-energy limit of the radiative dipole strength in nuclei,” Phys. Rev. C, vol. 88, p. 031302, 2013. [52] X. Roca-Maza, G. Pozzi, M. Brenna, K. Mizuyama, and G. Colò, “Low-lying dipole response: Isospin character and collectivity in 68Ni, 132Sn, and 208Pb,” Phys. Rev. C, vol. 85, p. 024601, 2012. [53] I. Daoutidis and S. Goriely, “Large-scale continuum random-phase approximation predic- tions of dipole strength for astrophysical applications,” Phys. Rev. C, vol. 86, p. 034328, 2012. [54] 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., vol. 93, p. 142504, 2004. [55] A. Simon, M. Guttormsen, A. C. Larsen, C. W. Beausang, P. Humby, J. T. Harke, R. J. Casperson, R. O. Hughes, T. J. Ross, J. M. Allmond, and et al., “First observation of low-energy 𝛟-ray enhancement in the rare-earth region,” Phys. Rev. C, vol. 93, p. 034303, 2016. [56] S. W. Jones, H. Möller, C. L. Fryer, C. J. Fontes, R. Trappitsch, W. P. Even, A. Couture, M. R. Mumpower, and S. Safi-Harb, “60Fe in core-collapse supernovae and prospects for X-ray and gamma-ray detection in supernova remnants,” Monthly Notices of the Royal Astronomical Society, vol. 485, no. 3, pp. 4287–4310, 2019. [57] R. Schwengner, S. Frauendorf, and B. A. Brown, “Low-Energy Magnetic Dipole Radiation in Open-Shell Nuclei,” Phys. Rev. Lett., vol. 118, p. 092502, 2017. [58] A. C. Larsen, N. Blasi, A. Bracco, F. Camera, T. K. Eriksen, A. Görgen, M. Guttormsen, T. W. Hagen, S. Leoni, B. Million, and et al., “Evidence for the Dipole Nature of the Low-Energy 𝛟 Enhancement in 56Fe,” Phys. Rev. Lett., vol. 111, p. 242504, 2013. [59] A. C. Larsen and S. Goriely, “Impact of a low-energy enhancement in the 𝛟-ray strength function on the neutron-capture cross section,” Phys. Rev. C, vol. 82, p. 014318, 2010. [60] A. Spyrou, D. Richman, A. Couture, C. E. Fields, S. N. Liddick, K. Childers, B. P. Crider, P. A. DeYoung, A. C. Dombos, P. Gastis, and et al., “Enhanced production of 60Fe in massive stars,” Nature Communications, vol. 15, 2024. [61] M. Guttormsen, K. O. Ay, M. Ozgur, E. Algin, A. C. Larsen, F. L. Bello Garrote, H. C. Berg, L. Crespo Campo, T. Dahl-Jacobsen, F. W. Furmyr, and et al., “Evolution of the 𝛟-ray strength function in neodymium isotopes,” Phys. Rev. C, vol. 106, p. 034314, 2022. 111 [62] A. Koning, S. Hilaire, and S. Goriely, “TALYS: modeling of nuclear reactions,” The Euro- pean Physical Journal A, vol. 59, p. 131, 2023. [63] A. Koning and J. Delaroche, “Local and global nucleon optical models from 1 keV to 200 MeV,” Nuclear Physics A, vol. 713, no. 3, pp. 231–310, 2003. [64] [65] [66] J. P. Jeukenne, A. Lejeune, and C. Mahaux, “Optical-model potential in nuclear matter from Reid’s hard core interaction,” Phys. Rev. C, vol. 10, pp. 1391–1401, 1974. J. Jeukenne, A. Lejeune, and C. Mahaux, “Many-body theory of nuclear matter,” Physics Reports, vol. 25, no. 2, pp. 83–174, 1976. J. P. Jeukenne, A. Lejeune, and C. Mahaux, “Microscopic calculation of the symmetry and Coulomb components of the complex optical-model potential,” Phys. Rev. C, vol. 15, pp. 10–29, 1977. [67] J.-P. Jeukenne, A. Lejeune, and C. Mahaux, “Optical-model potential in finite nuclei from Reid’s hard core interaction,” Phys. Rev. C, vol. 16, pp. 80–96, 1977. [68] E. Bauge, J. P. Delaroche, and M. Girod, “Semimicroscopic nucleon-nucleus spherical optical model for nuclei with 𝐎 > 40 at energies up to 200 MeV,” Phys. Rev. C, vol. 58, pp. 1118–1145, 1998. [69] E. Bauge, J. P. Delaroche, and M. Girod, “Lane-consistent, semimicroscopic nucleon-nucleus optical model,” Phys. Rev. C, vol. 63, p. 024607, 2001. [70] P. A. Moldauer, “Evaluation of the fluctuation enhancement factor,” Phys. Rev. C, vol. 14, pp. 764–766, 1976. [71] P. A. Moldauer, “Statistics and the average cross section,” Nuclear Physics A, vol. 344, no. 2, pp. 185–195, 1980. [72] J. W. Tepel, H. M. Hofmann, and H. A. WeidenmÃŒller, “Hauser-Feshbach formulas for medium and strong absorption,” Physics Letters B, vol. 49, no. 1, pp. 1–4, 1974. [73] H. M. Hofmann, J. Richert, J. W. Tepel, and H. A. WeidenmÃŒller, “Direct reactions and Hauser-Feshbach theory,” Annals of Physics, vol. 90, no. 2, pp. 403–437, 1975. [74] H. M. Hofmann, T. Mertelmeier, M. Herman, and J. W. Tepel, “Hauser-Feshbach calcu- lations in the presence of weakly absorbing channels with special reference to the elastic enhancement factor and the factorization assumption,” Zeitschrift fÃŒr Physik A Atoms and Nuclei, vol. 297, no. 2, pp. 153–160, 1980. [75] D. Brink, “Individual particle and collective aspects of the nuclear photoeffect,” Nuclear Physics, vol. 4, pp. 215–220, 1957. [76] P. Axel, “Electric Dipole Ground-State Transition Width Strength Function and 7-MeV Photon Interactions,” Phys. Rev., vol. 126, pp. 671–683, 1962. 112 [77] S. Goriely, “Radiative neutron captures by neutron-rich nuclei and the r-process nucleosyn- thesis,” Physics Letters B, vol. 436, no. 1, pp. 10–18, 1998. [78] V. Plujko, O. Gorbachenko, R. Capote, and P. Dimitriou, “Giant dipole resonance parameters of ground-state photoabsorption: Experimental values with uncertainties,” Atomic Data and Nuclear Data Tables, vol. 123-124, pp. 1–85, 2018. [79] R. Capote, M. Herman, P. ObloÅŸinskÜ, P. Young, S. Goriely, T. Belgya, A. Ignatyuk, A. Koning, S. Hilaire, V. Plujko, and et al., “RIPL – Reference Input Parameter Library for Calculation of Nuclear Reactions and Nuclear Data Evaluations,” Nuclear Data Sheets, vol. 110, no. 12, pp. 3107–3214, 2009. Special Issue on Nuclear Reaction Data. [80] D. Bohle, A. Richter, W. Steffen, A. Dieperink, N. Lo Iudice, F. Palumbo, and O. Scholten, “New magnetic dipole excitation mode studied in the heavy deformed nucleus 156Gd by inelastic electron scattering,” Physics Letters B, vol. 137, no. 1, pp. 27–31, 1984. [81] R. E. Atkinson and F. G. Houtermans, “Zur Frage der Aufbaumöglichkeit der Elemente in Sternen,” Zeits. f. Physik 54, 656, 1929. [82] Wikimedia, “Gamow Peak.” This work is licensed under the Creative Commons Attribution 4.0 International License. To view a copy of this license, visit http://creativecommons. org/licenses/by/4.0/. [83] M. Asplund, N. Grevesse, A. J. Sauval, and P. Scott, “The Chemical Composition of the Sun,” Annual Review of Astronomy & Astrophysics, vol. 47, no. 1, pp. 481–522, 2009. [84] J. R. Oppenheimer and G. M. Volkoff, “On Massive Neutron Cores,” Phys. Rev., vol. 55, pp. 374–381, 1939. [85] H. J. Lamers and E. M. Levesque, Understanding Stellar Evolution. 2514-3433, IOP Pub- lishing, 2017. [86] R. A. Alpher, H. Bethe, and G. Gamow, “The Origin of Chemical Elements,” Phys. Rev., vol. 73, pp. 803–804, 1948. [87] E. J. Öpik, “Collision probabilities with the planets and the distribution of interplanetary matter,” Pattern Recognition and Image Analysis, vol. 54, pp. 165–199, 1951. [88] E. E. Salpeter, “Nuclear Reactions in the Stars. I. Proton-Proton Chain,” Phys. Rev., vol. 88, pp. 547–553, 1952. [89] E. E. Salpeter, “Nuclear Reactions in Stars. Buildup from Helium,” Phys. Rev., vol. 107, pp. 516–525, 1957. [90] F. Hoyle, D. N. F. Dunbar, R. E. Pixley, W. A. Wenzel, and W. Whaling, “The 7.68-MeV State in C12,” Phys. Rev., vol. 92, pp. 649–650, 1953. [91] D. E. Alburger, “Gamma-Ray Decay of the 7.66-Mev Level of C12,” Phys. Rev., vol. 124, pp. 193–198, 1961. 113 [92] J. A. Nolen and S. M. Austin, “Measurement of the excitation energy of the 7.654 MeV state of 12C and the rate of the 3𝛌 reaction,” Phys. Rev. C, vol. 13, pp. 1773–1779, 1976. [93] P. Descouvemont and I. Baraffe, “Microscopic cluster study of the 12B and 12N systems and application to a hot pp chain in zero-metal stars,” Nuclear Physics A, vol. 514, no. 1, pp. 66–86, 1990. [94] T. A. LÀhde, U.-G. Meißner, and E. Epelbaum, “An update on fine-tunings in the triple-alpha process,” The European Physical Journal A, vol. 56, 2020. [95] S. Jin, S. M. Roberts, Luke F. Austin, and H. Schatz, “Enhanced triple-𝛌 reaction reduces proton-rich nucleosynthesis in supernovae,” Nature, vol. 588, 2020. [96] J. Bishop, C. E. Parker, G. V. Rogachev, S. Ahn, E. Koshchiy, K. Brandenburg, C. R. Brune, R. J. Charity, J. Derkin, N. Dronchi, and et al., “Neutron-upscattering enhancement of the triple-alpha process,” Nature Communications, vol. 13, 2022. [97] F. Herwig, Evolution of Solar and Intermediate-Mass Stars, pp. 397–445. Dordrecht: Springer Netherlands, 2013. [98] R. C. Duncan, S. L. Shapiro, and I. Wasserman, “Neutrino-driven Winds from Young, Hot [99] Neutron Stars,” Astrophysical Journal, vol. 309, p. 141, 1986. J. J. Cowan and W. K. Rose, “Production of 14C and neutrons in red giants.,” The Astrophys- ical Journal, vol. 212, pp. 149–158, 1977. [100] M. Hampel, R. J. Stancliffe, M. Lugaro, and B. S. Meyer, “The Intermediate Neutron-capture Process and Carbon-enhanced Metal-poor Stars,” The Astrophysical Journal, vol. 831, no. 2, p. 171, 2016. [101] M. Pignatari, P. Hoppe, R. Trappitsch, C. Fryer, F. Timmes, F. Herwig, and R. Hirschi, “The neutron capture process in the He shell in core-collapse supernovae: Presolar silicon carbide grains as a diagnostic tool for nuclear astrophysics,” Geochimica et Cosmochimica Acta, vol. 221, pp. 37–46, 2018. Astrophysical Implications of Extraterrestrial Materials: A Special issue for Ernst K. Zinner. [102] P. W. Merrill, “Spectroscopic Observations of Stars of Class S,” The Astrophysical Journal, vol. 116, p. 21, 1952. [103] M. Pignatari, R. Gallino, M. Heil, M. Wiescher, F. KÀppeler, F. Herwig, and S. Bisterzo, “The Weak s-Process in Massive Stars and its Dependence on the Neuton-Capture Cross Sections,” The Astrophysical Journal, vol. 710, p. 1557, feb 2010. [104] M. R. Mumpower, R. Surman, D.-L. Fang, M. Beard, P. Möller, T. Kawano, and A. Apra- hamian, “Impact of individual nuclear masses on 𝑟-process abundances,” Phys. Rev. C, vol. 92, p. 035807, 2015. [105] T. Marketin, L. Huther, and G. Martínez-Pinedo, “Large-scale evaluation of 𝛜-decay rates of 𝑟-process nuclei with the inclusion of first-forbidden transitions,” Phys. Rev. C, vol. 93, p. 025805, 2016. 114 [106] J. J. Cowan, C. Sneden, J. E. Lawler, A. Aprahamian, M. Wiescher, K. Langanke, G. Martínez-Pinedo, and F.-K. Thielemann, “Origin of the heaviest elements: The rapid neutron-capture process,” Rev. Mod. Phys., vol. 93, p. 015002, 2021. [107] K. A. Lund, J. Engel, G. C. McLaughlin, M. R. Mumpower, E. M. Ney, and R. Surman, “The Influence of 𝛜-decay Rates on r-process Observables,” The Astrophysical Journal, vol. 944, no. 2, p. 144, 2023. [108] S. E. Woosley, J. R. Wilson, G. J. Mathews, R. D. Hoffman, and B. S. Meyer, “The r-Process and Neutrino-heated Supernova Ejecta,” The Astrophysical Journal, vol. 433, p. 229, 1994. [109] K. Takahashi, J. Witti, and H. T. Janka, “Nucleosynthesis in neutrino-driven winds from protoneutron stars II. The r-process,” Astronomy and Astrophysics, vol. 286, pp. 857–869, 1994. [110] S. Wanajo, H.-T. Janka, and B. MÃŒller, “Electron-capture Supernovae as The Origin of Elements Beyond Iron,” The Astrophysical Journal Letters, vol. 726, no. 2, p. L15, 2011. [111] C. Winteler, R. KÀppeli, A. Perego, A. Arcones, N. Vasset, N. Nishimura, M. Liebendörfer, and F.-K. Thielemann, “Magnetorotationally Driven Supernovae as the Origin of Eearly Galaxy r-Process Elements?,” The Astrophysical Journal Letters, vol. 750, no. 1, p. L22, 2012. [112] P. Mösta, C. D. Ott, D. Radice, L. F. Roberts, E. Schnetter, and R. Haas, “A large-scale dy- namo and magnetoturbulence in rapidly rotating core-collapse supernovae,” Nature, vol. 528, 2015. [113] N. Nishimura, H. Sawai, T. Takiwaki, S. Yamada, and F.-K. Thielemann, “The Intermediate r-process in Core-collapse Supernovae Driven by the Magneto-rotational Instability,” The Astrophysical Journal Letters, vol. 836, no. 2, p. L21, 2017. [114] A. I. MacFadyen and S. E. Woosley, “Collapsars: Gamma-Ray Bursts and Explosions in “Failed Supernovae”,” The Astrophysical Journal, vol. 524, no. 1, pp. 262–289, 1999. [115] J. C. McKinney, A. Tchekhovskoy, and R. D. Blandford, “Alignment of Magnetized Accretion Disks and Relativistic Jets with Spinning Black Holes,” Science, vol. 339, no. 6115, pp. 49– 52, 2013. [116] J. M. Lattimer and D. N. Schramm, “Black-Hole-Neutron-Star Collisions,” The Astrophysical Journal, vol. 192, p. L145, 1974. [117] C. Freiburghaus, S. Rosswog, and F.-K. Thielemann, “r-Process in Neutron Star Mergers,” The Astrophysical Journal, vol. 525, no. 2, p. L121, 1999. [118] B. P. Abbot et al. (LIGO Scientific Collaboration and Virgo Collaboration), “GW170817: Observation of Gravitational Waves from a Binary Neutron Star Inspiral,” Phys. Rev. Lett., vol. 119, p. 161101, 2017. 115 [119] F. K. Thielemann, M. Eichler, I. V. Panov, and B. Wehmeyer, “Neutron Star Mergers and Nucleosynthesis of Heavy Elements,” Annual Review of Nuclear and Particle Science, vol. 67, pp. 253–274, 2017. [120] L. Baiotti and L. Rezzolla, “Binary neutron star mergers: a review of Einstein’s richest laboratory,” Reports on Progress in Physics, vol. 80, no. 9, p. 096901, 2017. [121] Rosswog, S., Sollerman, J., Feindt, U., Goobar, A., Korobkin, O., Wollaeger, R., Fremling, C., and Kasliwal, M. M., “The first direct double neutron star merger detection: Implications for cosmic nucleosynthesis,” Astronomy & Astrophysics, vol. 615, p. A132, 2018. [122] B. CÃŽté, P. Denissenkov, F. Herwig, A. J. Ruiter, C. Ritter, M. Pignatari, and K. Belczynski, “i-process Contribution of Rapidly Accreting White Dwarfs to the Solar Composition of First-peak Neutron-capture Elements,” The Astrophysical Journal, vol. 854, no. 2, p. 105, 2018. [123] Choplin, A., Siess, L., and Goriely, S., “The intermediate neutron capture process - III. The i-process in AGB stars of different masses and metallicities without overshoot,” Astronomy & Astrophysics, vol. 667, p. A155, 2022. [124] A. Spyrou, D. MÃŒcher, P. A. Denissenkov, F. Herwig, E. C. Good, and et al., “First Study of the 139Ba(𝑛, 𝛟)140Ba Reaction to Constrain the Conditions for the Astrophysical 𝑖 Process,” Phys. Rev. Lett., vol. 132, p. 202701, 2024. [125] V. W. Ingeberg, S. Siem, M. Wiedeking, and et al (The ISOLDE Collaboration), “Nuclear level density and 𝛟-ray strength function of 67Ni and the impact on the 𝑖 process,” Phys. Rev. C, vol. 111, p. 015803, 2025. [126] K. Lodders, “Solar System Abundances and Condensation Temperatures of the Elements,” The Astrophysical Journal, vol. 51, 2003. [127] I. Dillmann, T. Rauscher, M. Heil, F. KÀppeler, W. Rapp, and F.-K. Thielemann, “p-Process simulations with a modified reaction library,” Journal of Physics G: Nuclear and Particle Physics, vol. 35, no. 1, p. 014029, 2007. [128] Goriely, S., Arnould, M., Borzov, I., and Rayet, M., “The puzzle of the synthesis of the rare nuclide 138La,” Astronomy & Astrophysics, vol. 375, no. 2, pp. L35–L38, 2001. [129] S. Bisterzo, C. Travaglio, R. Gallino, M. Wiescher, and F. KÀppeler, “Galactic Chemical Evolution and Solar s-Process Abundances: Dependence on the 13C-Pocket Structure,” The Astrophysical Journal, vol. 787, no. 1, p. 10, 2014. [130] A. Sieverding, K. Langanke, G. Martínez-Pinedo, R. Bollig, H.-T. Janka, and A. Heger, “The 𝜈-process with Fully Time-dependent Supernova Neutrino Emission Spectra,” The Astrophysical Journal, vol. 876, no. 2, p. 151, 2019. [131] M. Arnould and S. Goriely, “The p-process of stellar nucleosynthesis: astrophysics and nuclear physics status,” Physics Reports, vol. 384, no. 1, pp. 1–84, 2003. 116 [132] M. Lugaro, U. Ott, and ÂŽA. Kereszturi, “Radioactive nuclei from cosmochronology to habit- ability,” Progress in Particle and Nuclear Physics, vol. 102, pp. 1–47, 2018. [133] S. E. Woosley and W. M. Howard, “The p-processes in supernovae.,” The Astrophysical Journal Supplement Series, vol. 36, pp. 285–304, 1978. [134] T. Rauscher, N. Dauphas, I. Dillmann, C. Fröhlich, Z. FÃŒlöp, and G. GyÃŒrky, “Constraining the astrophysical origin of the p-nuclei through nuclear physics and meteoritic data,” Reports on Progress in Physics, vol. 76, no. 6, p. 066201, 2013. [135] N. Prantzos, M. Hashimoto, M. Rayet, and M. Arnould, “The p-process in SN 1987A.,” Astronomy and Astrophysics, vol. 238, pp. 455–461, 1990. [136] T. Rauscher, A. Heger, R. D. Hoffman, and S. E. Woosley, “Nucleosynthesis in Massive Stars with Improved Nuclear and Stellar Physics,” The Astrophysical Journal, vol. 576, no. 1, pp. 323–348, 2002. [137] W. M. Howard, B. S. Meyer, and S. E. Woosley, “A New Site for the Astrophysical Gamma- Process,” Astrophysical Journal Letters, vol. 373, p. L5, 1991. [138] C. Travaglio, F. K. Röpke, R. Gallino, and W. Hillebrandt, “Type Ia Supernovae as Sites of the p-process: Two-dimensional Models Coupled to Nucleosynthesis,” The Astrophysical Journal, vol. 739, no. 2, p. 93, 2011. [139] N. Nishimura, T. Rauscher, R. Hirschi, A. S. J. Murphy, G. Cescutti, and C. Travaglio, “Uncertainties in the production of p nuclides in thermonuclear supernovae determined by Monte Carlo variations,” Monthly Notices of the Royal Astronomical Society, vol. 474, no. 3, pp. 3133–3139, 2017. [140] Roberti, L., Pignatari, M., Fryer, C., and Lugaro, M., “The 𝑔𝑎𝑚𝑚𝑎-process nucleosynthesis in core-collapse supernovae - II. Effect of the explosive recipe,” Astronomy & Astrophysics, vol. 686, p. L8, 2024. [141] M. Arnould, “Possibility of synthesis of proton-rich nuclei in highly evolved stars. II.,” Astronomy and Astrophysics, vol. 46, no. 1, pp. 117–125, 1976. [142] C. Ritter, F. Herwig, S. Jones, M. Pignatari, C. Fryer, and R. Hirschi, “NuGrid stellar data set – II. Stellar yields from H to Bi for stellar models with MZAMS = 1 − 25M⊙ and Z = 0.0001–0.02,” Monthly Notices of the Royal Astronomical Society, vol. 480, no. 1, pp. 538–571, 2018. [143] Roberti, L., Pignatari, M., Psaltis, A., Sieverding, A., Mohr, P., FÃŒlöp, Zs., and Lugaro, M., “The 𝛟-process nucleosynthesis in core-collapse supernovae - I. A novel analysis of 𝛟-process yields in massive stars,” Astronomy & Astrophysics, vol. 677, p. A22, 2023. [144] M. Pignatari, R. Hirschi, M. Wiescher, R. Gallino, M. Bennett, M. Beard, C. Fryer, F. Herwig, G. Rockefeller, and F. X. Timmes, “The 12c + 12c reaction and the impract on nucleosynthesis in massive stars,” The Astrophysical Journal, vol. 762, no. 1, p. 31, 2012. 117 [145] Choplin, A., Goriely, S., Hirschi, R., Tominaga, N., and Meynet, G., “The p-process in exploding rotating massive stars,” Astronomy & Astrophysics, vol. 661, p. A86, 2022. [146] U. Battino, M. Pignatari, C. Travaglio, C. Lederer-Woods, P. Denissenkov, F. Herwig, F. Thielemann, and T. Rauscher, “Heavy elements nucleosynthesis on accreting white dwarfs: building seeds for the p-process,” Monthly Notices of the Royal Astronomical Society, vol. 497, no. 4, pp. 4981–4998, 2020. [147] P. A. Denissenkov, F. Herwig, P. Woodward, R. Andrassy, M. Pignatari, and S. Jones, “The i-process yields of rapidly accreting white dwarfs from multicycle He-shell flash stellar evolution models with mixing parametrizations from 3D hydrodynamics simulations,” Monthly Notices of the Royal Astronomical Society, vol. 488, no. 3, pp. 4258–4270, 2019. [148] Goriely, S., José, J., Hernanz, M., Rayet, M., and Arnould, M., “He-detonation in sub- Chandrasekhar CO white dwarfs: A new insight into energetics and 𝑝-process nucleosyn- thesis,” Astronomy & Astrophysics, vol. 383, no. 3, pp. L27–L30, 2002. [149] M. Pignatari, K. Göbel, R. Reifarth, and C. Travaglio, “The production of proton-rich isotopes beyond iron: The 𝛟-process in stars,” International Journal of Modern Physics E, vol. 25, no. 04, p. 1630003, 2016. [150] H. Schatz, L. Bildsten, A. Cumming, and M. Wiescher, “The Rapid Proton Process Ashes from Stable Nuclear Burning on an Accreting Neutron Star,” The Astrophysical Journal, vol. 524, no. 2, pp. 1014–1029, 1999. [151] J. José, F. Moreno, A. Parikh, and C. Iliadis, “Hydrodynamic Models of Type I X-Ray Bursts: Metallicity Effects,” The Astrophysical Journal Supplement Series, vol. 189, no. 1, p. 204, 2010. [152] J. Keegans, C. L. Fryer, S. W. Jones, B. CÃŽté, K. Belczynski, F. Herwig, M. Pignatari, A. M. Laird, and C. A. Diget, “Nucleosynthetic yields from neutron stars accreting in binary common envelopes,” Monthly Notices of the Royal Astronomical Society, vol. 485, no. 1, pp. 620–639, 2019. [153] C. Fröhlich, G. Martínez-Pinedo, M. Liebendörfer, F.-K. Thielemann, E. Bravo, W. R. Hix, K. Langanke, and N. T. Zinner, “Neutrino-Induced Nucleosynthesis of 𝐎 > 64 Nuclei: The 𝜈 𝑝 Process,” Phys. Rev. Lett., vol. 96, p. 142502, 2006. [154] S. Wanajo, “The rp-Process in Neutrino-driven Winds,” The Astrophysical Journal, vol. 647, no. 2, pp. 1323–1340, 2006. [155] A. Psaltis, M. Jacobi, F. Montes, A. Arcones, C. J. Hansen, and H. Schatz, “Neutrino- driven Outflows and the Elemental Abundance Patterns of Very Metal-poor Stars,” The Astrophysical Journal, vol. 966, no. 1, p. 11, 2024. [156] Z. Xiong, G. Martínez-Pinedo, O. Just, and A. Sieverding, “Production of 𝑝 Nuclei from 𝑟-Process Seeds: The 𝜈𝑟 Process,” Phys. Rev. Lett., vol. 132, p. 192701, 2024. 118 [157] B. Paxton, L. Bildsten, A. Dotter, F. Herwig, P. Lesaffre, and F. Timmes, “Modules for Experiments in Stellar Astrophysics (MESA),” The Astrophysical Journal Supplement Series, vol. 192, no. 1, p. 3, 2010. [158] F. Herwig, S. Diehl, C. L. Fryer, R. Hirschi, A. Hungerford, G. Magkotsios, M. Pignatari, G. Rockefeller, F. X. Timmes, P. Young, and M. E. Bennet, “Nucleosynthesis simulations for a wide range of nuclear production sites from NuGrid,” in Nuclei in the Cosmos (NIC X), p. E23, 2008. [159] M. Pignatari and F. Herwig, “The NuGrid Research Platform: A Comprehensive Simulation Approach for Nuclear Astrophysics,” Nuclear Physics News, vol. 22, no. 4, pp. 18–23, 2012. [160] W. Rapp, J. Goerres, M. Wiescher, H. Schatz, and F. Kaeppeler, “Sensitivity of p-Process Nucleosynthesis to Nuclear Reaction Rates in a 25 M⊙ Supernova Model,” The Astrophysical Journal, vol. 653, pp. 474 – 489, 2006. [161] T. Rauscher, N. Nishimura, R. Hirschi, G. Cescutti, A. S. J. Murphy, and A. Heger, “Uncer- tainties in the production of p nuclei in massive stars obtained from Monte Carlo variations,” Monthly Notices of the Royal Astronomical Society, vol. 463, no. 4, pp. 4153–4166, 2016. [162] N. Nishimura, T. Rauscher, R. Hirschi, A. S. J. Murphy, G. Cescutti, and C. Travaglio, “Uncertainties in the production of p nuclides in thermonuclear supernovae determined by Monte Carlo variations,” Monthly Notices of the Royal Astronomical Society, vol. 474, no. 3, pp. 3133–3139, 2017. [163] G. GyÃŒrky, G. G. Kiss, Z. Elekes, Z. FÃŒlöp, E. Somorjai, and T. Rauscher, “Proton capture cross-section of 106,108Cd for the astrophysical p-process,” Journal of Physics G: Nuclear and Particle Physics, vol. 34, no. 5, p. 817, 2007. [164] G. GyÃŒrky, M. Vakulenko, Z. FÃŒlöp, Z. Halász, G. Kiss, E. Somorjai, and T. SzÃŒcs, “Cross section and reaction rate of 92Mo( 𝑝, 𝛟)93Tc determined from thick target yield measure- ments,” Nuclear Physics A, vol. 922, pp. 112–125, 2014. [165] S. Harissopulos, A. Spyrou, V. Foteinou, M. Axiotis, G. Provatas, and P. Demetriou, “Sys- tematic study of proton capture reactions in medium-mass nuclei relevant to the 𝑝 process: The case of 103Rh and 113,115In,” Phys. Rev. C, vol. 93, p. 025804, 2016. [166] J. Glorius, C. Langer, Z. Slavkovská, L. Bott, C. Brandau, B. BrÃŒckner, K. Blaum, X. Chen, S. Dababneh, T. Davinson, and et. al., “Approaching the Gamow Window with Stored Ions: Direct Measurement of 124Xe( 𝑝, 𝛟) in the ESR Storage Ring,” Phys. Rev. Lett., vol. 122, p. 092701, 2019. [167] A. Psaltis, A. Khaliel, E.-M. Assimakopoulou, A. Kanellakopoulos, V. Lagaki, M. Lykiar- dopoulou, E. Malami, P. Tsavalas, A. Zyriliou, and T. J. Mertzimekis, “Cross-section mea- surements of radiative proton-capture reactions in 112Cd at energies of astrophysical interest,” Phys. Rev. C, vol. 99, p. 065807, 2019. 119 [168] A. Simon, R. Kelmar, O. Olivas-Gomez, E. Churchman, P. Millican, C. S. Reingold, T. An- derson, A. M. Clark, N. Cooper, A. C. Dombos, and et al., “First measurements of cap- ture reactions for the 𝛟-process using HECTOR,” Journal of Physics: Conference Series, vol. 1308, no. 1, p. 012020, 2019. [169] A. Banu, E. G. Meekins, J. A. Silano, H. J. Karwowski, and S. Goriely, “Photoneutron reaction cross section measurements on 94Mo and 90Zr relevant to the 𝑝-process nucleosyn- thesis,” Phys. Rev. C, vol. 99, p. 025802, 2019. [170] F. Heim, J. Mayer, M. MÃŒller, P. Scholz, M. Weinert, and A. Zilges, “Experimental techniques to study the 𝛟 process for nuclear astrophysics at the Cologne accelerator laboratory,” Nu- clear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 966, p. 163854, 2020. [171] H. Cheng, B.-H. Sun, L.-H. Zhu, M. Kusakabe, Y. Zheng, L.-C. He, T. Kajino, Z.-M. Niu, T.-X. Li, C.-B. Li, and et al., “Measurements of 160Dy( 𝑝, 𝛟) at Energies Relevant for the Astrophysical 𝛟 Process,” The Astrophysical Journal, vol. 915, no. 2, p. 78, 2021. [172] P.-A. Söderström, A. Kuşoğlu, and D. Testov, “Prospect for measurements of (𝛟, n) reaction cross-sections of p-nuclei at ELI-NP,” Frontiers in Astronomy and Space Sciences, vol. 10, 2023. [173] G. Lotay, S. A. Gillespie, M. Williams, T. Rauscher, M. Alcorta, A. M. Amthor, C. A. Andreoiu, D. Baal, G. C. Ball, S. S. Bhattacharjee, and et al., “First Direct Measurement of an Astrophysical 𝑝-Process Reaction Cross Section Using a Radioactive Ion Beam,” Phys. Rev. Lett., vol. 127, p. 112701, 2021. [174] T. Rauscher and F.-K. Thielemann, “Tables of Nuclear Cross Sections and Reaction Rates: an Addendum to the Papaer “Astrophysical Reaction Rates from Statistical Model Calcula- tions”,” Atomic Data and Nuclear Data Tables, vol. 79, no. 1, pp. 47–64, 2001. [175] Webpage hosted by the Experimental Astrophysics Group of Goethe University Frankfurt am Main at https://exp-astro.physik.uni-frankfurt.de/fluxes/. p-process fluxes tool. Accessed 01/19/2025. [176] M. Pignatari, F. Herwig, R. Hirschi, M. Bennett, G. Rockefeller, C. Fryer, F. X. Timmes, C. Ritter, A. Heger, S. Jones, U. Battino, A. Dotter, R. Trappitsch, S. Diehl, U. Frischknecht, A. Hungerford, G. Magkotsios, C. Travaglio, and P. Young, “NuGrid Stellar Data Set. I. Stellar Yields from H to Bi for Stars with Metallicities Z = 0.02 and Z = 0.01,” The Astrophysical Journal Supplement Series, vol. 225, no. 2, p. 24, 2016. [177] T. V. Lawson, M. Pignatari, R. J. Stancliffe, J. den Hartogh, S. Jones, C. L. Fryer, B. K. Gibson, and M. Lugaro, “Radioactive nuclei in the early Solar system: analysis of the 15 isotopes produced by core-collapse supernovae,” Monthly Notices of the Royal Astronomical Society, vol. 511, no. 1, pp. 886–902, 2021. [178] G. A. Krivonosov, O. L. Ekhichev, B. A. Nemashkalo, V. E. Storizhko, and V. K. Chirt, “Radiative-Capture Cross Sections of Nuclides of Intermediate Atomic Weight for Low- Energy Protons,” Bull.Russian Academy of Sciences - Physics, vol. 41, p. 175, 1977. 120 [179] G. GyÃŒrky, Z. FÃŒlöp, E. Somorjai, Z. Elekes, M. Kokkoris, S. Galanopulos, P. Demetriou, S. Harissopulos, S. Goriely, and T. Rauscher, “Se( 𝑝, 𝛟) cross section measurements for p-process studies,” Nuclear Physics A, vol. 718, pp. 599–601, 2003. [180] Y. Skakun, S. Utenkov, V. Mishchenko, J. Farkas, Z. FÃŒlöp, G. GyÃŒrky, G. G. Kiss, E. Somor- jai, and T. Rauscher, “Cross sections of low energy ( 𝑝, 𝛟) and ( 𝑝, 𝑛)-reactions on selenium isotopes for the astrophysical 𝛟-process,” 3rd Int. Conf. Cur. Prob. in Nucl. Phys. Atom. Ene., p. 207, 2010. [181] V. Foteinou, S. Harissopulos, M. Axiotis, A. Lagoyannis, G. Provatas, A. Spyrou, G. Perdikakis, C. Zarkadas, and P. Demetriou, “Cross section measurements of proton capture reactions on Se isotopes relevant to the astrophysical 𝑝 process,” Phys. Rev. C, vol. 97, p. 035806, 2018. [182] Z. FÃŒlöp, A. Kiss, E. Somorjai, C. Rolfs, H. Trautvetter, T. Rauscher, and H. Oberhummer, “70Ge(𝛌, 𝛟)74Se cross section measurements at energies of astrophysical interest,” Zeitschrift fÃŒr Physik A Hadrons and nuclei, vol. 355, pp. 203–207, 1996. [183] I. Dillmann, M. Heil, F. KÀppeler, T. Rauscher, and F.-K. Thielemann, “Experimental (𝑛, 𝛟) cross sections of the p-process nuclei 74Se and 84Sr,” Phys. Rev. C, vol. 73, p. 015803, 2006. [184] A. Palmisano-Kyle, A. Spyrou, P. A. DeYoung, A. Dombos, P. Gastis, O. Olivas-Gomez, C. Harris, S. Liddick, S. M. Lyons, J. Pereira, and et al., “Constraining the astrophysical 𝑝 process: Cross section measurement of the 84Kr( 𝑝, 𝛟)85Rb reaction in inverse kinematics,” Phys. Rev. C, vol. 105, p. 065804, 2022. [185] A. Tsantiri, A. Palmisano-Kyle, A. Spyrou, P. Mohr, H. C. Berg, P. A. DeYoung, A. C. Dombos, P. Gastis, E. C. Good, C. M. Harris, and et al., “Cross-section measurement of the 82Kr( 𝑝, 𝛟)83Rb reaction in inverse kinematics,” Phys. Rev. C, vol. 107, p. 035808, 2023. [186] G. B. Thomas Glasmacher, Alexandra Gade and J. Wei, “The Facility for Rare Isotope Beams: Providing New Opportunities for Science,” Nuclear Physics News, vol. 34, no. 3, pp. 5–12, 2024. [187] A. Gade and B. M. Sherrill, “NSCL and FRIB at Michigan State University: Nuclear science at the limits of stability,” Physica Scripta, vol. 91, no. 5, p. 053003, 2016. [188] C. Sumithrarachchi, Y. Liu, S. Rogers, S. Schwarz, G. Bollen, N. Gamage, A. Henriques, A. Lapierre, R. Ringle, I. Yandow, and et al., “The new Batch Mode Ion Source for stand- alone operation at the Facility for Rare Isotope Beams (FRIB),” Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms, vol. 541, pp. 301–304, 2023. [189] A. C. Candia, G. Bollen, S. Campbell, A. Henriques, A. Lapierre, S. Nash, S. Schwarz, C. Sumithrarachchi, A. C. Villari, and K. A. Domnanich, “Preparation of a 73As source sample for application in an offline ion source,” Applied Radiation and Isotopes, vol. 217, p. 111642, 2025. 121 [190] A. Lapierre, G. Bollen, D. Crisp, S. W. Krause, L. E. Linhardt, K. Lund, S. Nash, R. Rencsok, R. Ringle, S. Schwarz, and et al., “First two operational years of the electron-beam ion trap charge breeder at the National Superconducting Cyclotron Laboratory,” Phys. Rev. Accel. Beams, vol. 21, p. 053401, 2018. [191] S. Nash and et al., “Commissioning of the FRIB/NSCL New ReA3 4-Rod Radio Frequency Quadrupole Accelerator,” in Proc. NAPAC’19, pp. 817–819, JACoW Publishing, Geneva, Switzerland, 2019. [192] A. Villari, G. Bollen, A. Henriques, A. Lapierre, S. Nash, R. Ringle, S. Schwarz, and C. Sumithrarachchi, “Gas stopping and reacceleration techniques at the Facility for Rare Isotope Beams (FRIB),” Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms, vol. 541, pp. 350–354, 2023. [193] A. Simon, S. Quinn, A. Spyrou, and et. al., “SuN: Summing NaI(Tl) gamma-ray detector for capture reaction measurements,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 703, pp. 16–21, 2013. [194] S. Quinn, Capture Cross Sections for the Astrophysical p Process. PhD thesis, Michigan State University, 2015. [195] R. Hofstadter, “Alkali Halide Scintillation Counters,” Phys. Rev., vol. 74, pp. 100–101, 1948. [196] G. Knoll, Radiation Detection and Measurement (4th ed.). Hoboken, NJ: John Wiley, 2010. [197] C. Prokop, S. Liddick, B. Abromeit, A. Chemey, N. Larson, S. Suchyta, and J. Tompkins, “Digital data acquisition system implementation at the National Superconducting Cyclotron Laboratory,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 741, pp. 163–168, 2014. [198] A. Spyrou, H.-W. Becker, A. Lagoyannis, S. Harissopulos, and C. Rolfs, “Cross-section measurements of capture reactions relevant to the p process using a 4𝜋 𝛟-summing method,” Phys. Rev. C, vol. 76, p. 015802, Jul 2007. [199] E. Klopfer, J. Brett, P. A. DeYoung, A. C. Dombos, S. J. Quinn, A. Simon, and A. Spyrou, “SuNSCREEN: A cosmic-ray veto detector for capture-reaction measurements,” Nuclear Instruments and Methods in Physics Research A, vol. 788, pp. 5–8, 2015. [200] A. Spyrou, S. J. Quinn, A. Simon, T. Rauscher, and et al., “Measurement of the 90,92Zr(𝑝,𝛟)91,93Nb reactions for the nucleosynthesis of elements near 𝐎 = 90,” Phys. Rev. C, vol. 88, p. 045802, 2013. [201] F. Naqvi, S. J. Quinn, A. Spyrou, A. Battaglia, M. Couder, P. A. DeYoung, A. C. Dom- bos, X. Fang, J. Görres, A. Kontos, and et al., “Proton capture cross section of 72Ge and astrophysical implications,” Phys. Rev. C, vol. 92, p. 025804, 2015. [202] A. Simon, M. Beard, A. Spyrou, S. J. Quinn, B. Bucher, M. Couder, P. A. DeYoung, A. C. Dombos, J. Görres, A. Kontos, and et al., “Systematic study of (𝛌, 𝛟) reactions for stable nickel isotopes,” Phys. Rev. C, vol. 92, p. 025806, 2015. 122 [203] R. Kelmar, A. Simon, O. Olivas-Gomez, P. Millican, C. S. Reingold, E. Churchman, A. M. Clark, S. L. Henderson, S. E. Kelly, D. Robertson, E. Stech, and W. P. Tan, “Searching for (𝛟, 𝛌)/(𝛟, 𝑛) branching points in the 𝛟-process path near 𝐎 = 100,” Phys. Rev. C, vol. 101, p. 015801, 2020. [204] J. Fallis, C. Akers, A. Laird, A. Simon, A. Spyrou, G. Christian, D. Connolly, U. Hager, D. Hutcheon, A. Lennarz, P. O’Malley, S. Quinn, J. Riley, A. Rojas, C. Ruiz, and M. Williams, “First measurement in the Gamow window of a reaction for the 𝛟-process in inverse kine- matics: 76Se(𝛌, 𝛟)80Kr,” Physics Letters B, vol. 807, p. 135575, 2020. [205] S. Quinn, A. Spyrou, A. Simon, A. Battaglia, M. Bowers, B. Bucher, C. Casarella, M. Couder, P. DeYoung, A. Dombos, and et al., “First application of the 𝛟-summing technique in inverse kinematics,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 757, pp. 62–66, 2014. [206] S. Quinn, A. Spyrou, A. Simon, and et. al., “First application of the gamma-summing tech- nique in inverse kinematics,” Nuclear Instruments and Methods in Physics Research Section A Accelerators Spectrometers Detectors and Associated Equipment, vol. 757, p. 62–66, 2014. [207] O. Tarasov, D. Bazin, M. Hausmann, M. Kuchera, P. Ostroumov, M. Portillo, B. Sherrill, K. Tarasova, and T. Zhang, “LISE cute++, the latest generation of the LISE ++ package, to simulate rare isotope production with fragment-separators,” Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms, vol. 541, pp. 4–7, 2023. [208] L. Kirsch and L. Bernstein, “RAINIER: A simulation tool for distributions of excited nuclear states and cascade fluctuations,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 892, pp. 30–40, 2018. [209] S. Agostinelli, J. Allison, K. Amako, J. Apostolakis, H. Araujo, P. Arce, M. Asai, D. Axen, S. Banerjee, and G. B. et.al., “Geant4—a simulation toolkit,” Nuclear Instruments and Meth- ods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 506, no. 3, pp. 250–303, 2003. [210] A. C. Dombos, 𝛜-Decay Total Absorption Spectroscopy Around A = 100-110 Relevant to Nuclear Structure and the Astrophysical r Process. PhD thesis, Michigan State University, 2015. [211] F. James, “MINUIT Function Minimization and Error Analysis: Reference Manual Version 94.1,” tech. rep., CERN, 1994. [212] R. Brun and F. Rademakers, “ROOT — An object oriented data analysis framework,” Nu- clear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 389, no. 1, pp. 81–86, 1997. New Computing Techniques in Physics Research V. [213] T. von Egidy and D. Bucurescu, “Experimental energy-dependent nuclear spin distributions,” Phys. Rev. C, vol. 80, p. 054310, 2009. 123 [214] R. Hoffman, F. Dietrich, R. Bauer, K. Kelley, and M. Mustafa, “Neutron and Charged-Particle Induced Cross Sections for Radiochemistry in the Region of Bromine and Krypton,” tech. rep., Lawrence Livermore National Lab. (LLNL), Livermore, CA (United States, 2004. [215] M. Guttormsen, R. Chankova, U. Agvaanluvsan, and et. al., “Radiative strength functions in 93−98Mo,” Phys. Rev. C, vol. 71, p. 044307, 2005. [216] A. Gelman, J. B. Carlin, H. S. Stern, D. B. Dunson, A. Vehtari, and D. B. Rubin, Bayesian Data Analysis. Chapman and Hall/CRC, 3rd ed., 2013. [217] C. D. Pruitt, J. E. Escher, and R. Rahman, “Uncertainty-quantified phenomenological optical potentials for single-nucleon scattering,” Phys. Rev. C, vol. 107, p. 014602, Jan 2023. [218] J. F. Ziegler, M. Ziegler, and J. Biersack, “SRIM – The stopping and range of ions in matter (2010),” Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms, vol. 268, no. 11, pp. 1818–1823, 2010. 19th International Conference on Ion Beam Analysis. [219] G. GyÃŒrky, E. Somorjai, Z. FÃŒlöp, S. Harissopulos, P. Demetriou, and T. Rauscher, “Proton capture cross section of Sr isotopes and their importance for nucleosynthesis of proton-rich nuclides,” Phys. Rev. C, vol. 64, p. 065803, 2001. [220] N. Grevesse and A. Noels, “Atomic data and the spectrum of the solar photosphere,” in In: Origin and Evolution of the elements, 1993. [221] I. Dillmann, M. Heil, F. KÀppeler, R. Plag, T. Rauscher, and F. K. Thielemann, “KADoNiS- The Karlsruhe Astrophysical Database of Nucleosynthesis in Stars,” in Capture Gamma-Ray Spectroscopy and Related Topics (A. Woehr and A. Aprahamian, eds.), vol. 819 of American Institute of Physics Conference Series, pp. 123–127, AIP, 2006. [222] R. H. Cyburt, A. M. Amthor, R. Ferguson, Z. Meisel, K. Smith, S. Warren, A. Heger, R. D. Hoffman, T. Rauscher, A. Sakharuk, H. Schatz, F. K. Thielemann, and M. Wiescher, “The JINA REACLIB Database: Its Recent Updates and Impact on Type-I X-Ray Bursts,” The Astrophysical Journal Supplement Series, vol. 189, no. 1, p. 240, 2010. [223] D. Kasen, F. K. Röpke, and S. E. Woosley, “The diversity of type Ia supernovae from broken symmetries,” Nature, vol. 460, no. 7257, pp. 869–872, 2009. [224] I. Domínguez, E. Bravo, L. Piersanti, O. Straniero, and A. Tornambé, “Supernovae and Dark Energy,” in Cosmology Across Cultures (J. Rubiño-Martín, J. Belmonte, F. Prada, and A. Alberdi, eds.), vol. 409 of Astronomical Society of the Pacific Conference Series, p. 42, 2009. [225] Travaglio, C., Hillebrandt, W., Reinecke, M., and Thielemann, F.-K., “Nucleosynthesis in multi-dimensional sn ia explosions*,” Astronomy & Astrophysics, vol. 425, no. 3, pp. 1029– 1040, 2004. [226] K. Maeda, F. Röpke, M. Fink, W. Hillebrandt, C. Travaglio, and F.-K. Thielemann, “Nucle- osynthesis in Two-Dimensional Delayed Detonation Models of Type Ia Supernova Explo- sions,” The Astrophysical Journal, vol. 712, no. 1, p. 624, 2010. 124 APPENDIX TALYS SCORES FOR THE 73AS(𝑝, 𝛟)74SE REACTION Table .1 The Talys model parameters and scores Model # NLD 𝛟SF 𝛟SF 𝛟SF E1 M1 upbend OMP TAS Score SoS Score Mul Score Cross Section Score (Eq. 6.2)) Total Score 0 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 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n JLM 0.4341 default 0.4341 JLM 0.4341 default 0.4341 JLM 0.3805 default 0.3805 JLM 0.3805 default 0.3805 JLM 0.6034 default 0.6034 JLM 0.5111 default 0.5111 JLM 0.6068 default 0.6068 JLM 0.6068 default 0.6068 JLM 0.5904 default 0.5904 JLM 0.5904 default 0.5904 JLM 0.6513 default 0.6513 JLM 0.6650 default 0.6650 JLM 0.1251 default 0.1251 JLM 0.1251 default 0.1251 JLM 0.1196 default 0.1196 JLM 0.1196 125 0.1215 0.1215 0.1215 0.1215 0.1132 0.1132 0.1132 0.1132 0.2326 0.2326 0.1921 0.1921 0.2244 0.2244 0.2244 0.2244 0.2119 0.2119 0.2119 0.2119 0.2892 0.2892 0.2708 0.2708 0.0336 0.0336 0.0336 0.0336 0.0306 0.0306 0.0306 0.1064 0.1064 0.1064 0.1064 0.0982 0.0982 0.0982 0.0982 0.3122 0.3122 0.1852 0.1852 0.2058 0.2058 0.2058 0.2058 0.1643 0.1643 0.1643 0.1643 0.3077 0.3077 0.2538 0.2538 0.0058 0.0058 0.0058 0.0058 0.0049 0.0049 0.0049 0.6233 0.4702 0.6233 0.4702 0.1828 0.1284 0.1828 0.1284 0.6610 0.4984 0.5090 0.3717 0.1150 0.3221 0.1150 0.3221 0.0739 0.2417 0.0739 0.2417 0.1092 0.3127 0.1344 0.3566 0.9894 0.9743 0.9894 0.9743 0.9991 0.9335 0.9991 3.50e-03 2.64e-03 3.50e-03 2.64e-03 7.73e-04 5.43e-04 7.73e-04 5.43e-04 2.90e-02 2.18e-02 9.26e-03 6.76e-03 3.22e-03 9.03e-03 3.22e-03 9.03e-03 1.52e-03 4.97e-03 1.52e-03 4.97e-03 6.33e-03 1.81e-02 6.14e-03 1.63e-02 2.39e-05 2.36e-05 2.39e-05 2.36e-05 1.79e-05 1.68e-05 1.79e-05 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 3 3 3 3 3 4 4 4 4 4 4 4 4 4 4 4 4 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 Table .1 (cont’d) n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n default 0.1196 JLM 0.1893 default 0.1893 JLM 0.1570 default 0.1570 JLM 0.2687 default 0.2687 JLM 0.2687 default 0.2687 JLM 0.2277 default 0.2277 JLM 0.2277 default 0.2277 JLM 0.3996 default 0.3996 JLM 0.3407 default 0.3407 JLM 0.5923 default 0.5923 JLM 0.5923 default 0.5923 JLM 0.5877 default 0.5877 JLM 0.5877 default 0.5877 JLM 0.6936 default 0.6936 JLM 0.6530 default 0.6530 JLM 0.2837 default 0.2837 JLM 0.2837 default 0.2837 JLM 0.2792 default 0.2792 JLM 0.2792 default 0.2792 126 0.0306 0.0519 0.0519 0.0410 0.0410 0.0755 0.0755 0.0755 0.0755 0.0663 0.0663 0.0663 0.0663 0.1296 0.1296 0.1024 0.1024 0.1996 0.1996 0.1996 0.1996 0.2218 0.2218 0.2218 0.2218 0.3048 0.3048 0.2735 0.2735 0.0897 0.0897 0.0897 0.0897 0.0873 0.0873 0.0873 0.0873 0.0049 0.0171 0.0171 0.0084 0.0084 0.0215 0.0215 0.0215 0.0215 0.0183 0.0183 0.0183 0.0183 0.0739 0.0739 0.0397 0.0397 0.1673 0.1673 0.1673 0.1673 0.2642 0.2642 0.2642 0.2642 0.4075 0.4075 0.3104 0.3104 0.0315 0.0315 0.0315 0.0315 0.0374 0.0374 0.0374 0.0374 0.9335 0.9818 0.9817 0.9993 0.9380 0.9545 0.8232 0.9545 0.8232 0.7658 0.5993 0.7658 0.5993 0.9685 0.8437 0.8989 0.7431 0.9989 0.9471 0.9989 0.9471 0.9084 0.7550 0.9084 0.7550 0.9966 0.9577 0.9926 0.8992 0.9846 0.8793 0.9846 0.8793 0.8439 0.6795 0.8439 0.6795 1.68e-05 1.65e-04 1.65e-04 5.38e-05 5.05e-05 4.17e-04 3.60e-04 4.17e-04 3.60e-04 2.12e-04 1.66e-04 2.12e-04 1.66e-04 3.71e-03 3.23e-03 1.25e-03 1.03e-03 1.98e-02 1.87e-02 1.98e-02 1.87e-02 3.13e-02 2.60e-02 3.13e-02 2.60e-02 8.58e-02 8.25e-02 5.50e-02 4.98e-02 7.90e-04 7.05e-04 7.90e-04 7.05e-04 7.69e-04 6.19e-04 7.69e-04 6.19e-04 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 6 6 6 6 7 7 7 7 7 7 7 7 7 7 7 7 8 8 8 8 8 8 8 8 8 8 8 8 9 9 9 9 9 9 9 9 9 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 Table .1 (cont’d) y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y JLM 0.4293 default 0.4293 JLM 0.3489 default 0.3489 JLM 0.2374 default 0.2374 JLM 0.2374 default 0.2374 JLM 0.1838 default 0.1838 JLM 0.1838 default 0.1838 JLM 0.3343 default 0.3343 JLM 0.2825 default 0.2825 JLM 0.2741 default 0.2741 JLM 0.2377 default 0.2377 JLM 0.2891 default 0.2891 JLM 0.2507 default 0.2507 JLM 0.3842 default 0.3842 JLM 0.3022 default 0.3022 JLM 0.3274 default 0.3274 JLM 0.3274 default 0.3274 JLM 0.3239 default 0.3239 JLM 0.3239 default 0.3239 JLM 0.4561 127 0.1508 0.1508 0.1266 0.1266 0.0712 0.0712 0.0712 0.0712 0.0661 0.0661 0.0661 0.0661 0.1170 0.1170 0.0976 0.0976 0.0703 0.0703 0.0656 0.0656 0.0690 0.0690 0.0662 0.0662 0.1075 0.1075 0.0826 0.0826 0.0837 0.0837 0.0837 0.0837 0.0785 0.0785 0.0785 0.0785 0.1359 0.1118 0.1118 0.0697 0.0697 0.0260 0.0260 0.0260 0.0260 0.0234 0.0234 0.0234 0.0234 0.0782 0.0782 0.0486 0.0486 0.0227 0.0227 0.0190 0.0190 0.0234 0.0234 0.0206 0.0206 0.0689 0.0689 0.0311 0.0311 0.0357 0.0357 0.0357 0.0357 0.0291 0.0291 0.0291 0.0291 0.0898 0.9919 0.8962 0.9483 0.8105 0.9990 0.9310 0.9990 0.9310 0.8592 0.6940 0.8592 0.6940 0.9990 0.9435 0.9848 0.8764 0.8199 0.9766 0.8515 0.9869 0.8010 0.9693 0.8335 0.9812 0.7990 0.9696 0.9057 0.9983 0.9823 0.8748 0.9823 0.8748 0.8855 0.7287 0.8855 0.7287 0.9903 7.18e-03 6.48e-03 2.92e-03 2.50e-03 4.39e-04 4.09e-04 4.39e-04 4.09e-04 2.44e-04 1.97e-04 2.44e-04 1.97e-04 3.06e-03 2.89e-03 1.32e-03 1.17e-03 3.59e-04 4.27e-04 2.52e-04 2.92e-04 3.75e-04 4.53e-04 2.85e-04 3.36e-04 2.27e-03 2.76e-03 7.03e-04 7.74e-04 9.62e-04 8.56e-04 9.62e-04 8.56e-04 6.55e-04 5.39e-04 6.55e-04 5.39e-04 5.51e-03 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 9 9 9 1 1 1 1 1 1 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 Table .1 (cont’d) y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y default 0.4561 JLM 0.3966 default 0.3966 JLM 0.7550 default 0.7550 JLM 0.7550 default 0.7550 JLM 0.7579 default 0.7579 JLM 0.7579 default 0.7579 JLM 0.8004 default 0.8004 JLM 0.8017 default 0.8017 JLM 0.7800 default 0.7800 JLM 0.7800 default 0.7800 JLM 0.7604 default 0.7604 JLM 0.7604 default 0.7604 JLM 0.8162 default 0.8162 JLM 0.8090 default 0.8090 JLM 0.3597 default 0.3597 JLM 0.3597 default 0.3597 JLM 0.3119 default 0.3119 JLM 0.3119 default 0.3119 JLM 0.4162 default 0.4162 128 0.1359 0.1131 0.1131 0.3980 0.3980 0.3980 0.3980 0.4332 0.4332 0.4332 0.4332 0.5198 0.5198 0.4793 0.4793 0.4271 0.4271 0.4271 0.4271 0.4153 0.4153 0.4153 0.4153 0.4802 0.4802 0.4547 0.4547 0.1108 0.1108 0.1108 0.1108 0.1065 0.1065 0.1065 0.1065 0.1681 0.1681 0.0898 0.0513 0.0513 0.4213 0.4213 0.4213 0.4213 0.6112 0.6112 0.6112 0.6112 0.7146 0.7146 0.6168 0.6168 0.3763 0.3763 0.3763 0.3763 0.3203 0.3203 0.3203 0.3203 0.4998 0.4998 0.4203 0.4203 0.0350 0.0350 0.0350 0.0350 0.0336 0.0336 0.0336 0.0336 0.1134 0.1134 0.8920 0.9443 0.8052 0.6055 0.8607 0.6055 0.8607 0.9987 0.9278 0.9987 0.9278 0.6417 0.8870 0.7648 0.9549 0.0002 0.0052 0.0002 0.0052 0.0001 0.0028 0.0001 0.0028 0.0003 0.0055 0.0004 0.0068 0.1684 0.4097 0.1684 0.4097 0.2385 0.5084 0.2385 0.5084 0.1836 0.4344 4.96e-03 2.17e-03 1.85e-03 7.66e-02 1.09e-01 7.66e-02 1.09e-01 2.00e-01 1.86e-01 2.00e-01 1.86e-01 1.91e-01 2.64e-01 1.81e-01 2.26e-01 3.10e-05 6.47e-04 3.10e-05 6.47e-04 1.04e-05 2.85e-04 1.04e-05 2.85e-04 5.33e-05 1.08e-03 5.67e-05 1.05e-03 2.35e-04 5.72e-04 2.35e-04 5.72e-04 2.66e-04 5.67e-04 2.66e-04 5.67e-04 1.46e-03 3.45e-03 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 3 3 4 4 4 4 4 4 4 4 4 4 4 4 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 6 6 6 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 Table .1 (cont’d) n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n JLM 0.3803 default 0.3803 JLM 0.5390 default 0.5390 JLM 0.5390 default 0.5390 JLM 0.4906 default 0.4906 JLM 0.4906 default 0.4906 JLM 0.6107 default 0.6107 JLM 0.5955 default 0.5955 JLM 0.7854 default 0.7854 JLM 0.7854 default 0.7854 JLM 0.7968 default 0.7968 JLM 0.7968 default 0.7968 JLM 0.8128 default 0.8128 JLM 0.8179 default 0.8179 JLM 0.6189 default 0.6189 JLM 0.6189 default 0.6189 JLM 0.5958 default 0.5958 JLM 0.5958 default 0.5958 JLM 0.6786 default 0.6786 JLM 0.6584 129 0.1359 0.1359 0.2134 0.2134 0.2134 0.2134 0.2055 0.2055 0.2055 0.2055 0.3340 0.3340 0.2653 0.2653 0.4463 0.4463 0.4463 0.4463 0.5018 0.5018 0.5018 0.5018 0.5518 0.5518 0.5174 0.5174 0.2926 0.2926 0.2926 0.2926 0.2853 0.2853 0.2853 0.2853 0.3997 0.3997 0.3403 0.0534 0.0534 0.0879 0.0879 0.0879 0.0879 0.0891 0.0891 0.0891 0.0891 0.3055 0.3055 0.1452 0.1452 0.4802 0.4802 0.4802 0.4802 0.6083 0.6083 0.6083 0.6083 0.6866 0.6866 0.6195 0.6195 0.1990 0.1990 0.1990 0.1990 0.2310 0.2310 0.2310 0.2310 0.4442 0.4442 0.2727 0.2441 0.5169 0.3580 0.6466 0.3580 0.6466 0.6688 0.9020 0.6688 0.9020 0.3855 0.6773 0.4893 0.7722 0.0965 0.2880 0.0965 0.2880 0.2555 0.5316 0.2555 0.5316 0.1058 0.3068 0.1423 0.3702 0.2232 0.4881 0.2232 0.4881 0.4601 0.7467 0.4601 0.7467 0.2425 0.5156 0.3172 6.74e-04 1.43e-03 3.62e-03 6.54e-03 3.62e-03 6.54e-03 6.01e-03 8.10e-03 6.01e-03 8.10e-03 2.40e-02 4.22e-02 1.12e-02 1.77e-02 1.62e-02 4.85e-02 1.62e-02 4.85e-02 6.22e-02 1.29e-01 6.22e-02 1.29e-01 3.26e-02 9.45e-02 3.73e-02 9.71e-02 8.04e-03 1.76e-02 8.04e-03 1.76e-02 1.81e-02 2.93e-02 1.81e-02 2.93e-02 2.92e-02 6.21e-02 1.94e-02 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 6 7 7 7 7 7 7 7 7 7 7 7 7 8 8 8 8 8 8 8 8 8 8 8 8 9 9 9 9 9 9 9 9 9 9 9 9 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 Table .1 (cont’d) n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n default 0.6584 JLM 0.5634 default 0.5634 JLM 0.5634 default 0.5634 JLM 0.5225 default 0.5225 JLM 0.5225 default 0.5225 JLM 0.6292 default 0.6292 JLM 0.6231 default 0.6231 JLM 0.5668 default 0.5668 JLM 0.5088 default 0.5088 JLM 0.5212 default 0.5212 JLM 0.5401 default 0.5401 JLM 0.5994 default 0.5994 JLM 0.5556 default 0.5556 JLM 0.6245 default 0.6245 JLM 0.6245 default 0.6245 JLM 0.6088 default 0.6088 JLM 0.6088 default 0.6088 JLM 0.6654 default 0.6654 JLM 0.6953 default 0.6953 130 0.3403 0.2482 0.2482 0.2482 0.2482 0.2405 0.2405 0.2405 0.2405 0.3374 0.3374 0.2853 0.2853 0.2143 0.2143 0.1901 0.1901 0.2127 0.2127 0.1860 0.1860 0.2829 0.2829 0.2240 0.2240 0.2380 0.2380 0.2380 0.2380 0.2378 0.2378 0.2378 0.2378 0.3563 0.3563 0.2917 0.2917 0.2727 0.1750 0.1750 0.1750 0.1750 0.1976 0.1976 0.1976 0.1976 0.4192 0.4192 0.2759 0.2759 0.1036 0.1036 0.0675 0.0675 0.1029 0.1029 0.0694 0.0694 0.2509 0.2509 0.1045 0.1045 0.1091 0.1091 0.1091 0.1091 0.1166 0.1166 0.1166 0.1166 0.3242 0.3242 0.1659 0.1659 0.6048 0.2049 0.4641 0.2049 0.4641 0.5237 0.8019 0.5237 0.8019 0.2229 0.4909 0.2929 0.5782 0.0409 0.1641 0.0467 0.1790 0.0368 0.1530 0.0421 0.1670 0.0451 0.1757 0.0703 0.2348 0.2357 0.5039 0.2357 0.5039 0.4106 0.6999 0.4106 0.6999 0.2558 0.5318 0.3336 0.6221 3.69e-02 5.02e-03 1.14e-02 5.02e-03 1.14e-02 1.30e-02 1.99e-02 1.30e-02 1.99e-02 1.98e-02 4.37e-02 1.44e-02 2.84e-02 5.15e-04 2.07e-03 3.05e-04 1.17e-03 4.20e-04 1.75e-03 2.93e-04 1.16e-03 1.92e-03 7.48e-03 9.15e-04 3.05e-03 3.82e-03 8.17e-03 3.82e-03 8.17e-03 6.93e-03 1.18e-02 6.93e-03 1.18e-02 1.97e-02 4.09e-02 1.12e-02 2.09e-02 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 1 1 1 1 1 1 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 4 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 Table .1 (cont’d) y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y JLM 0.6586 default 0.6586 JLM 0.6586 default 0.6586 JLM 0.6616 default 0.6616 JLM 0.6616 default 0.6616 JLM 0.7736 default 0.7736 JLM 0.7523 default 0.7523 JLM 0.7595 default 0.7595 JLM 0.7595 default 0.7595 JLM 0.7561 default 0.7561 JLM 0.7561 default 0.7561 JLM 0.7646 default 0.7646 JLM 0.7734 default 0.7734 JLM 0.2005 default 0.2005 JLM 0.2005 default 0.2005 JLM 0.1736 default 0.1736 JLM 0.1736 default 0.1736 JLM 0.2664 default 0.2664 JLM 0.2530 default 0.2530 JLM 0.3782 131 0.2592 0.2592 0.2592 0.2592 0.2761 0.2761 0.2761 0.2761 0.4157 0.4157 0.3517 0.3517 0.3203 0.3203 0.3203 0.3203 0.3050 0.3050 0.3050 0.3050 0.3777 0.3777 0.3516 0.3516 0.0523 0.0523 0.0523 0.0523 0.0496 0.0496 0.0496 0.0496 0.0819 0.0819 0.0680 0.0680 0.1190 0.2887 0.2887 0.2887 0.2887 0.3993 0.3993 0.3993 0.3993 0.6112 0.6112 0.4410 0.4410 0.2998 0.2998 0.2998 0.2998 0.2614 0.2614 0.2614 0.2614 0.4196 0.4196 0.3426 0.3426 0.0119 0.0119 0.0119 0.0119 0.0111 0.0111 0.0111 0.0111 0.0372 0.0372 0.0195 0.0195 0.0425 0.9991 0.9337 0.9991 0.9337 0.6630 0.4979 0.6630 0.4979 0.9980 0.9497 0.9817 0.8678 0.0124 0.0747 0.0124 0.0747 0.0066 0.0488 0.0066 0.0488 0.0115 0.0715 0.0152 0.0859 0.7192 0.9326 0.7192 0.9326 0.8170 0.9769 0.8170 0.9769 0.6899 0.9182 0.7980 0.9704 0.9250 4.92e-02 4.60e-02 4.92e-02 4.60e-02 4.84e-02 3.63e-02 4.84e-02 3.63e-02 1.96e-01 1.87e-01 1.15e-01 1.01e-01 9.02e-04 5.45e-03 9.02e-04 5.45e-03 3.98e-04 2.94e-03 3.98e-04 2.94e-03 1.40e-03 8.66e-03 1.41e-03 8.01e-03 9.00e-05 1.17e-04 9.00e-05 1.17e-04 7.82e-05 9.35e-05 7.82e-05 9.35e-05 5.59e-04 7.45e-04 2.68e-04 3.26e-04 1.77e-03 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 4 4 4 4 4 4 4 4 4 4 4 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 6 6 6 6 7 7 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 Table .1 (cont’d) y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y default 0.3782 JLM 0.3782 default 0.3782 JLM 0.3534 default 0.3534 JLM 0.3534 default 0.3534 JLM 0.5235 default 0.5235 JLM 0.4456 default 0.4456 JLM 0.7473 default 0.7473 JLM 0.7473 default 0.7473 JLM 0.7784 default 0.7784 JLM 0.7784 default 0.7784 JLM 0.8001 default 0.8001 JLM 0.8018 default 0.8018 JLM 0.4608 default 0.4608 JLM 0.4608 default 0.4608 JLM 0.4846 default 0.4846 JLM 0.4846 default 0.4846 JLM 0.5915 default 0.5915 JLM 0.5581 default 0.5581 JLM 0.3487 default 0.3487 132 0.1190 0.1190 0.1190 0.1071 0.1071 0.1071 0.1071 0.2008 0.2008 0.1513 0.1513 0.3399 0.3399 0.3399 0.3399 0.3781 0.3781 0.3781 0.3781 0.4469 0.4469 0.4171 0.4171 0.1674 0.1674 0.1674 0.1674 0.1600 0.1600 0.1600 0.1600 0.2519 0.2519 0.2189 0.2189 0.1293 0.1293 0.0425 0.0425 0.0425 0.0387 0.0387 0.0387 0.0387 0.1438 0.1438 0.0669 0.0669 0.3579 0.3579 0.3579 0.3579 0.4683 0.4683 0.4683 0.4683 0.6076 0.6076 0.5065 0.5065 0.1082 0.1082 0.1082 0.1082 0.1011 0.1011 0.1011 0.1011 0.2618 0.2618 0.1760 0.1760 0.0772 0.0772 0.9993 0.9250 0.9993 0.9960 0.9108 0.9960 0.9108 0.9036 0.9984 0.9718 0.9886 0.5619 0.8319 0.5619 0.8319 0.8279 0.9811 0.8279 0.9811 0.5356 0.8138 0.6387 0.8866 0.8020 0.9710 0.8020 0.9710 0.9734 0.9876 0.9734 0.9876 0.7745 0.9608 0.8735 0.9937 0.7386 0.9433 1.91e-03 1.77e-03 1.91e-03 1.46e-03 1.33e-03 1.46e-03 1.33e-03 1.37e-02 1.51e-02 4.39e-03 4.46e-03 5.11e-02 7.56e-02 5.11e-02 7.56e-02 1.14e-01 1.35e-01 1.14e-01 1.35e-01 1.16e-01 1.77e-01 1.08e-01 1.50e-01 6.69e-03 8.10e-03 6.69e-03 8.10e-03 7.63e-03 7.74e-03 7.63e-03 7.74e-03 3.02e-02 3.75e-02 1.88e-02 2.14e-02 2.57e-03 3.28e-03 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 4 4 4 7 7 7 7 7 7 7 7 7 7 8 8 8 8 8 8 8 8 8 8 8 8 9 9 9 9 9 9 9 9 9 9 9 9 1 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 Table .1 (cont’d) n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n JLM 0.3487 default 0.3487 JLM 0.3188 default 0.3188 JLM 0.3188 default 0.3188 JLM 0.4854 default 0.4854 JLM 0.4194 default 0.4194 JLM 0.4105 default 0.4105 JLM 0.3996 default 0.3996 JLM 0.4020 default 0.4020 JLM 0.3958 default 0.3958 JLM 0.4946 default 0.4946 JLM 0.4639 default 0.4639 JLM 0.5015 default 0.5015 JLM 0.5015 default 0.5015 JLM 0.4754 default 0.4754 JLM 0.4754 default 0.4754 JLM 0.6402 default 0.6402 JLM 0.5221 default 0.5221 JLM 0.7562 default 0.7562 JLM 0.7562 133 0.1293 0.1293 0.1213 0.1213 0.1213 0.1213 0.1975 0.1975 0.1675 0.1675 0.1086 0.1086 0.1008 0.1008 0.1122 0.1122 0.1015 0.1015 0.1639 0.1639 0.1251 0.1251 0.1410 0.1410 0.1410 0.1410 0.1318 0.1318 0.1318 0.1318 0.2289 0.2289 0.1771 0.1771 0.3451 0.3451 0.3451 0.0772 0.0772 0.0760 0.0760 0.0760 0.0760 0.2313 0.2313 0.1235 0.1235 0.0484 0.0484 0.0375 0.0375 0.0512 0.0512 0.0372 0.0372 0.1214 0.1214 0.0604 0.0604 0.0678 0.0678 0.0678 0.0678 0.0686 0.0686 0.0686 0.0686 0.1993 0.1993 0.1053 0.1053 0.5904 0.5904 0.5904 0.7386 0.9433 0.9764 0.9852 0.9764 0.9852 0.7102 0.9301 0.8156 0.9774 0.3808 0.6734 0.4155 0.7073 0.3606 0.6526 0.3942 0.6865 0.3595 0.6538 0.4830 0.7696 0.8351 0.9826 0.8351 0.9826 0.9639 0.9927 0.9639 0.9927 0.8085 0.9743 0.9022 0.9981 0.8579 0.9892 0.8579 2.57e-03 3.28e-03 2.87e-03 2.89e-03 2.87e-03 2.89e-03 1.57e-02 2.06e-02 7.08e-03 8.48e-03 8.22e-04 1.45e-03 6.27e-04 1.07e-03 8.32e-04 1.51e-03 5.88e-04 1.02e-03 3.54e-03 6.43e-03 1.69e-03 2.70e-03 4.01e-03 4.72e-03 4.01e-03 4.72e-03 4.14e-03 4.27e-03 4.14e-03 4.27e-03 2.36e-02 2.85e-02 8.79e-03 9.72e-03 1.32e-01 1.52e-01 1.32e-01 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 1 1 1 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 4 4 4 4 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 Table .1 (cont’d) n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n default 0.7562 JLM 0.7756 default 0.7756 JLM 0.7756 default 0.7756 JLM 0.8119 default 0.8119 JLM 0.7948 default 0.7948 JLM 0.7943 default 0.7943 JLM 0.7943 default 0.7943 JLM 0.7898 default 0.7898 JLM 0.7898 default 0.7898 JLM 0.8233 default 0.8233 JLM 0.8214 default 0.8214 JLM 0.2891 default 0.2891 JLM 0.2891 default 0.2891 JLM 0.2758 default 0.2758 JLM 0.2758 default 0.2758 JLM 0.5295 default 0.5295 JLM 0.3626 default 0.3626 JLM 0.5632 default 0.5632 JLM 0.5632 default 0.5632 134 0.3451 0.3670 0.3670 0.3670 0.3670 0.5291 0.5291 0.4647 0.4647 0.4419 0.4419 0.4419 0.4419 0.4155 0.4155 0.4155 0.4155 0.5187 0.5187 0.4829 0.4829 0.0773 0.0773 0.0773 0.0773 0.0740 0.0740 0.0740 0.0740 0.1570 0.1570 0.1036 0.1036 0.1828 0.1828 0.1828 0.1828 0.5904 0.6593 0.6593 0.6593 0.6593 0.6888 0.6888 0.7332 0.7332 0.5530 0.5530 0.5530 0.5530 0.5038 0.5038 0.5038 0.5038 0.7032 0.7032 0.6197 0.6197 0.0401 0.0401 0.0401 0.0401 0.0341 0.0341 0.0341 0.0341 0.2276 0.2276 0.0697 0.0697 0.1343 0.1343 0.1343 0.1343 0.9892 0.8012 0.6305 0.8012 0.6305 0.8553 0.9895 0.9579 0.9945 0.0002 0.0040 0.0002 0.0040 0.0001 0.0022 0.0001 0.0022 0.0002 0.0040 0.0002 0.0050 0.1750 0.4218 0.1750 0.4218 0.2478 0.5225 0.2478 0.5225 0.1742 0.4226 0.2402 0.5142 0.3434 0.6335 0.3434 0.6335 1.52e-01 1.50e-01 1.18e-01 1.50e-01 1.18e-01 2.53e-01 2.93e-01 2.59e-01 2.69e-01 3.28e-05 7.79e-04 3.28e-05 7.79e-04 1.17e-05 3.64e-04 1.17e-05 3.64e-04 5.09e-05 1.21e-03 5.79e-05 1.24e-03 1.57e-04 3.78e-04 1.57e-04 3.78e-04 1.73e-04 3.64e-04 1.73e-04 3.64e-04 3.29e-03 7.99e-03 6.29e-04 1.35e-03 4.75e-03 8.76e-03 4.75e-03 8.76e-03 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 6 6 6 6 7 7 7 7 7 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 Table .1 (cont’d) y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y JLM 0.5290 default 0.5290 JLM 0.5290 default 0.5290 JLM 0.7825 default 0.7825 JLM 0.6102 default 0.6102 JLM 0.7941 default 0.7941 JLM 0.7941 default 0.7941 JLM 0.8046 default 0.8046 JLM 0.8046 default 0.8046 JLM 0.8181 default 0.8181 JLM 0.8145 default 0.8145 JLM 0.6735 default 0.6735 JLM 0.6735 default 0.6735 JLM 0.6440 default 0.6440 JLM 0.6440 default 0.6440 JLM 0.8053 default 0.8053 JLM 0.7540 default 0.7540 JLM 0.4978 default 0.4978 JLM 0.4978 default 0.4978 JLM 0.5072 135 0.1685 0.1685 0.1685 0.1685 0.3546 0.3546 0.2481 0.2481 0.4369 0.4369 0.4369 0.4369 0.4883 0.4883 0.4883 0.4883 0.5647 0.5647 0.5357 0.5357 0.2304 0.2304 0.2304 0.2304 0.2495 0.2495 0.2495 0.2495 0.4060 0.4060 0.3112 0.3112 0.2143 0.2143 0.2143 0.2143 0.2039 0.1327 0.1327 0.1327 0.1327 0.5466 0.5466 0.2417 0.2417 0.6438 0.6438 0.6438 0.6438 0.7257 0.7257 0.7257 0.7257 0.7118 0.7118 0.7180 0.7180 0.2754 0.2754 0.2754 0.2754 0.3173 0.3173 0.3173 0.3173 0.6619 0.6619 0.4025 0.4025 0.2585 0.2585 0.2585 0.2585 0.2409 0.6498 0.8918 0.6498 0.8918 0.3418 0.6345 0.4509 0.7410 0.1850 0.4370 0.1850 0.4370 0.4489 0.7395 0.4489 0.7395 0.1845 0.4382 0.2521 0.5299 0.2794 0.5614 0.2794 0.5614 0.5540 0.8257 0.5540 0.8257 0.2783 0.5625 0.3726 0.6664 0.2180 0.4843 0.2180 0.4843 0.5508 7.68e-03 1.05e-02 7.68e-03 1.05e-02 5.19e-02 9.63e-02 1.65e-02 2.71e-02 4.13e-02 9.76e-02 4.13e-02 9.76e-02 1.28e-01 2.11e-01 1.28e-01 2.11e-01 6.07e-02 1.44e-01 7.90e-02 1.66e-01 1.19e-02 2.40e-02 1.19e-02 2.40e-02 2.82e-02 4.21e-02 2.82e-02 4.21e-02 6.02e-02 1.22e-01 3.52e-02 6.29e-02 6.01e-03 1.34e-02 6.01e-03 1.34e-02 1.37e-02 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 5 5 5 5 5 5 7 7 7 7 7 7 7 8 8 8 8 8 8 8 8 8 8 8 8 9 9 9 9 9 9 9 9 9 9 9 9 1 1 1 1 1 1 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 Table .1 (cont’d) y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y default 0.5072 JLM 0.5072 default 0.5072 JLM 0.7489 default 0.7489 JLM 0.5514 default 0.5514 JLM 0.5204 default 0.5204 JLM 0.4857 default 0.4857 JLM 0.5472 default 0.5472 JLM 0.4750 default 0.4750 JLM 0.7238 default 0.7238 JLM 0.5199 default 0.5199 JLM 0.5734 default 0.5734 JLM 0.5734 default 0.5734 JLM 0.5770 default 0.5770 JLM 0.5770 default 0.5770 JLM 0.7777 default 0.7777 JLM 0.6611 default 0.6611 JLM 0.3036 default 0.3036 JLM 0.3036 default 0.3036 JLM 0.1869 default 0.1869 136 0.2039 0.2039 0.2039 0.3448 0.3448 0.2550 0.2550 0.1786 0.1786 0.1550 0.1550 0.1742 0.1742 0.1525 0.1525 0.2818 0.2818 0.1905 0.1905 0.2016 0.2016 0.2016 0.2016 0.1886 0.1886 0.1886 0.1886 0.3632 0.3632 0.2617 0.2617 0.0750 0.0750 0.0750 0.0750 0.0707 0.0707 0.2409 0.2409 0.2409 0.5994 0.5994 0.3248 0.3248 0.1618 0.1618 0.1033 0.1033 0.1479 0.1479 0.0996 0.0996 0.4360 0.4360 0.1598 0.1598 0.1693 0.1693 0.1693 0.1693 0.1675 0.1675 0.1675 0.1675 0.5763 0.5763 0.2557 0.2557 0.0472 0.0472 0.0472 0.0472 0.0454 0.0454 0.8250 0.5508 0.8250 0.2172 0.4854 0.2952 0.5832 0.0397 0.1618 0.0460 0.1783 0.0357 0.1508 0.0414 0.1664 0.0395 0.1623 0.0653 0.2251 0.2693 0.5484 0.2693 0.5484 0.4619 0.7492 0.4619 0.7492 0.2681 0.5495 0.3599 0.6524 0.4802 0.3451 0.4802 0.3451 0.1260 0.0874 2.06e-02 1.37e-02 2.06e-02 3.36e-02 7.51e-02 1.35e-02 2.66e-02 5.96e-04 2.43e-03 3.58e-04 1.39e-03 5.03e-04 2.13e-03 2.99e-04 1.20e-03 3.52e-03 1.44e-02 1.03e-03 3.56e-03 5.27e-03 1.07e-02 5.27e-03 1.07e-02 8.42e-03 1.37e-02 8.42e-03 1.37e-02 4.37e-02 8.95e-02 1.59e-02 2.89e-02 5.16e-04 3.71e-04 5.16e-04 3.71e-04 7.56e-05 5.25e-05 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 4 4 4 4 4 4 4 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 Table .1 (cont’d) n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n JLM 0.1869 default 0.1869 JLM 0.5168 default 0.5168 JLM 0.4158 default 0.4158 JLM 0.5129 default 0.5129 JLM 0.5129 default 0.5129 JLM 0.5509 default 0.5509 JLM 0.5509 default 0.5509 JLM 0.5929 default 0.5929 JLM 0.5875 default 0.5875 JLM 0.0633 default 0.0633 JLM 0.0633 default 0.0633 JLM 0.0572 default 0.0572 JLM 0.0572 default 0.0572 JLM 0.0789 default 0.0789 JLM 0.0726 default 0.0726 JLM 0.1498 default 0.1498 JLM 0.1498 default 0.1498 JLM 0.1403 default 0.1403 JLM 0.1403 137 0.0707 0.0707 0.1586 0.1586 0.1247 0.1247 0.1558 0.1558 0.1558 0.1558 0.1469 0.1469 0.1469 0.1469 0.2068 0.2068 0.1936 0.1936 0.0205 0.0205 0.0205 0.0205 0.0193 0.0193 0.0193 0.0193 0.0344 0.0344 0.0262 0.0262 0.0464 0.0464 0.0464 0.0464 0.0422 0.0422 0.0422 0.0454 0.0454 0.1913 0.1913 0.1039 0.1039 0.1336 0.1336 0.1336 0.1336 0.1113 0.1113 0.1113 0.1113 0.2309 0.2309 0.1912 0.1912 0.0026 0.0026 0.0026 0.0026 0.0024 0.0024 0.0024 0.0024 0.0086 0.0086 0.0044 0.0044 0.0124 0.0124 0.0124 0.0124 0.0117 0.0117 0.0117 0.1260 0.0874 0.5020 0.3589 0.3803 0.2665 0.2266 0.4995 0.2266 0.4995 0.1591 0.4013 0.1591 0.4013 0.2193 0.4910 0.2553 0.5379 0.9907 0.8891 0.9907 0.8891 0.9603 0.8246 0.9603 0.8246 0.9942 0.8983 0.9651 0.8311 0.8608 0.6922 0.8608 0.6922 0.6378 0.4743 0.6378 7.56e-05 5.25e-05 7.87e-03 5.63e-03 2.05e-03 1.44e-03 2.42e-03 5.33e-03 2.42e-03 5.33e-03 1.43e-03 3.62e-03 1.43e-03 3.62e-03 6.21e-03 1.39e-02 5.55e-03 1.17e-02 3.37e-06 3.03e-06 3.37e-06 3.03e-06 2.52e-06 2.17e-06 2.52e-06 2.17e-06 2.31e-05 2.08e-05 7.98e-06 6.88e-06 7.39e-05 5.95e-05 7.39e-05 5.95e-05 4.41e-05 3.28e-05 4.41e-05 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 4 4 4 4 4 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 6 6 6 6 7 7 7 7 7 7 7 7 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 Table .1 (cont’d) n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n default 0.1403 JLM 0.2640 default 0.2640 JLM 0.2071 default 0.2071 JLM 0.4847 default 0.4847 JLM 0.4847 default 0.4847 JLM 0.5033 default 0.5033 JLM 0.5033 default 0.5033 JLM 0.5957 default 0.5957 JLM 0.6019 default 0.6019 JLM 0.2278 default 0.2278 JLM 0.2278 default 0.2278 JLM 0.1900 default 0.1900 JLM 0.1900 default 0.1900 JLM 0.3249 default 0.3249 JLM 0.2584 default 0.2584 JLM 0.1145 default 0.1145 JLM 0.1145 default 0.1145 JLM 0.1045 default 0.1045 JLM 0.1045 default 0.1045 138 0.0422 0.0871 0.0871 0.0709 0.0709 0.1357 0.1357 0.1357 0.1357 0.1501 0.1501 0.1501 0.1501 0.2289 0.2289 0.1969 0.1969 0.0558 0.0558 0.0558 0.0558 0.0566 0.0566 0.0566 0.0566 0.1036 0.1036 0.0821 0.0821 0.0435 0.0435 0.0435 0.0435 0.0416 0.0416 0.0416 0.0416 0.0117 0.0447 0.0447 0.0261 0.0261 0.1175 0.1175 0.1175 0.1175 0.1587 0.1587 0.1587 0.1587 0.3039 0.3039 0.2111 0.2111 0.0215 0.0215 0.0215 0.0215 0.0219 0.0219 0.0219 0.0219 0.0701 0.0701 0.0363 0.0363 0.0119 0.0119 0.0119 0.0119 0.0117 0.0117 0.0117 0.0117 0.4743 0.8758 0.7064 0.7857 0.6094 0.9629 0.8298 0.9629 0.8298 0.7754 0.6008 0.7754 0.6008 0.9704 0.8411 0.9206 0.7637 0.9097 0.7513 0.9097 0.7513 0.7140 0.5422 0.7140 0.5422 0.9218 0.7644 0.8468 0.6738 0.9590 0.8217 0.9590 0.8217 0.7359 0.5606 0.7359 0.5606 3.28e-05 9.01e-04 7.27e-04 3.01e-04 2.34e-04 7.44e-03 6.41e-03 7.44e-03 6.41e-03 9.30e-03 7.20e-03 9.30e-03 7.20e-03 4.02e-02 3.48e-02 2.30e-02 1.91e-02 2.49e-04 2.06e-04 2.49e-04 2.06e-04 1.68e-04 1.28e-04 1.68e-04 1.28e-04 2.17e-03 1.80e-03 6.52e-04 5.19e-04 5.68e-05 4.87e-05 5.68e-05 4.87e-05 3.74e-05 2.85e-05 3.74e-05 2.85e-05 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 6 7 7 7 7 8 8 8 8 8 8 8 8 8 8 8 8 9 9 9 9 9 9 9 9 9 9 9 9 1 1 1 1 1 1 1 1 1 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 Table .1 (cont’d) y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y JLM 0.1738 default 0.1738 JLM 0.1389 default 0.1389 JLM 0.1443 default 0.1443 JLM 0.1746 default 0.1746 JLM 0.1804 default 0.1804 JLM 0.1628 default 0.1628 JLM 0.2219 default 0.2219 JLM 0.2017 default 0.2017 JLM 0.2159 default 0.2159 JLM 0.2159 default 0.2159 JLM 0.2056 default 0.2056 JLM 0.2056 default 0.2056 JLM 0.3073 default 0.3073 JLM 0.2890 default 0.2890 JLM 0.7673 default 0.7673 JLM 0.7673 default 0.7673 JLM 0.7323 default 0.7323 JLM 0.7323 default 0.7323 JLM 0.8081 139 0.0770 0.0770 0.0593 0.0593 0.0449 0.0449 0.0410 0.0410 0.0441 0.0441 0.0403 0.0403 0.0691 0.0691 0.0517 0.0517 0.0514 0.0514 0.0514 0.0514 0.0489 0.0489 0.0489 0.0489 0.0948 0.0948 0.0722 0.0722 0.3867 0.3867 0.3867 0.3867 0.3952 0.3952 0.3952 0.3952 0.5319 0.0464 0.0464 0.0242 0.0242 0.0145 0.0145 0.0128 0.0128 0.0173 0.0173 0.0138 0.0138 0.0439 0.0439 0.0202 0.0202 0.0208 0.0208 0.0208 0.0208 0.0178 0.0178 0.0178 0.0178 0.0729 0.0729 0.0355 0.0355 0.4275 0.4275 0.4275 0.4275 0.5677 0.5677 0.5677 0.5677 0.7389 0.9670 0.8331 0.9137 0.7526 0.9455 0.9972 0.9616 0.9928 0.9343 0.9987 0.9518 0.9958 0.9368 0.9983 0.9861 0.9753 0.9060 0.7473 0.9060 0.7473 0.7648 0.5915 0.7648 0.5915 0.9184 0.7606 0.8425 0.6698 0.3713 0.6677 0.3713 0.6677 0.9942 0.9607 0.9942 0.9607 0.3685 6.00e-04 5.17e-04 1.82e-04 1.50e-04 8.86e-05 9.34e-05 8.80e-05 9.08e-05 1.29e-04 1.38e-04 8.63e-05 9.03e-05 6.31e-04 6.72e-04 2.07e-04 2.05e-04 2.09e-04 1.72e-04 2.09e-04 1.72e-04 1.37e-04 1.06e-04 1.37e-04 1.06e-04 1.95e-03 1.61e-03 6.25e-04 4.97e-04 4.71e-02 8.47e-02 4.71e-02 8.47e-02 1.63e-01 1.58e-01 1.63e-01 1.58e-01 1.17e-01 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 4 4 4 4 4 4 4 4 4 4 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 Table .1 (cont’d) y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y default 0.8081 JLM 0.8008 default 0.8008 JLM 0.8341 default 0.8341 JLM 0.8341 default 0.8341 JLM 0.8320 default 0.8320 JLM 0.8320 default 0.8320 JLM 0.8414 default 0.8414 JLM 0.8407 default 0.8407 JLM 0.4530 default 0.4530 JLM 0.4530 default 0.4530 JLM 0.4214 default 0.4214 JLM 0.4214 default 0.4214 JLM 0.5519 default 0.5519 JLM 0.4760 default 0.4760 JLM 0.6271 default 0.6271 JLM 0.6271 default 0.6271 JLM 0.6350 default 0.6350 JLM 0.6350 default 0.6350 JLM 0.7381 default 0.7381 140 0.5319 0.4783 0.4783 0.4911 0.4911 0.4911 0.4911 0.4784 0.4784 0.4784 0.4784 0.5451 0.5451 0.5200 0.5200 0.1323 0.1323 0.1323 0.1323 0.1312 0.1312 0.1312 0.1312 0.2078 0.2078 0.1645 0.1645 0.2633 0.2633 0.2633 0.2633 0.2416 0.2416 0.2416 0.2416 0.3871 0.3871 0.7389 0.6240 0.6240 0.4870 0.4870 0.4870 0.4870 0.4572 0.4572 0.4572 0.4572 0.6200 0.6200 0.5588 0.5588 0.0508 0.0508 0.0508 0.0508 0.0487 0.0487 0.0487 0.0487 0.1850 0.1850 0.0824 0.0824 0.1343 0.1343 0.1343 0.1343 0.1349 0.1349 0.1349 0.1349 0.4340 0.4340 0.6738 0.5039 0.7975 0.0000 0.0006 0.0000 0.0006 0.0000 0.0003 0.0000 0.0003 0.0000 0.0005 0.0000 0.0007 0.0119 0.0751 0.0119 0.0751 0.0203 0.1078 0.0203 0.1078 0.0117 0.0758 0.0180 0.1010 0.0662 0.2270 0.0662 0.2270 0.2051 0.4668 0.2051 0.4668 0.0646 0.2272 2.14e-01 1.20e-01 1.91e-01 1.98e-06 1.11e-04 1.98e-06 1.11e-04 6.53e-07 5.07e-05 6.53e-07 5.07e-05 2.70e-06 1.56e-04 3.26e-06 1.68e-04 3.62e-05 2.29e-04 3.62e-05 2.29e-04 5.48e-05 2.90e-04 5.48e-05 2.90e-04 2.48e-04 1.61e-03 1.16e-04 6.52e-04 1.47e-03 5.03e-03 1.47e-03 5.03e-03 4.25e-03 9.66e-03 4.25e-03 9.66e-03 8.01e-03 2.82e-02 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 4 4 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 6 6 6 6 7 7 7 7 7 7 7 7 7 7 7 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 Table .1 (cont’d) n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n n y y n JLM 0.6954 default 0.6954 JLM 0.8231 default 0.8231 JLM 0.8231 default 0.8231 JLM 0.8208 default 0.8208 JLM 0.8208 default 0.8208 JLM 0.8222 default 0.8222 JLM 0.8279 default 0.8279 JLM 0.6819 default 0.6819 JLM 0.6819 default 0.6819 JLM 0.6880 default 0.6880 JLM 0.6880 default 0.6880 JLM 0.7840 default 0.7840 JLM 0.7402 default 0.7402 JLM 0.6568 default 0.6568 JLM 0.6568 default 0.6568 JLM 0.6185 default 0.6185 JLM 0.6185 default 0.6185 JLM 0.7514 default 0.7514 JLM 0.7254 141 0.3223 0.3223 0.4822 0.4822 0.4822 0.4822 0.5158 0.5158 0.5158 0.5158 0.5742 0.5742 0.5517 0.5517 0.3159 0.3159 0.3159 0.3159 0.3103 0.3103 0.3103 0.3103 0.4414 0.4414 0.3730 0.3730 0.2637 0.2637 0.2637 0.2637 0.2533 0.2533 0.2533 0.2533 0.3744 0.3744 0.3146 0.2294 0.2294 0.5633 0.5633 0.5633 0.5633 0.6566 0.6566 0.6566 0.6566 0.7216 0.7216 0.6661 0.6661 0.2500 0.2500 0.2500 0.2500 0.2857 0.2857 0.2857 0.2857 0.5535 0.5535 0.3597 0.3597 0.2284 0.2284 0.2284 0.2284 0.2579 0.2579 0.2579 0.2579 0.5157 0.5157 0.3567 0.0967 0.2945 0.0379 0.1559 0.0379 0.1559 0.1507 0.3818 0.1507 0.3818 0.0366 0.1549 0.0551 0.2022 0.0513 0.1920 0.0513 0.1920 0.1620 0.4028 0.1620 0.4028 0.0499 0.1919 0.0752 0.2500 0.0560 0.2004 0.0560 0.2004 0.2423 0.5099 0.2423 0.5099 0.0541 0.1992 0.0811 4.97e-03 1.51e-02 8.47e-03 3.49e-02 8.47e-03 3.49e-02 4.19e-02 1.06e-01 4.19e-02 1.06e-01 1.25e-02 5.27e-02 1.68e-02 6.15e-02 2.76e-03 1.03e-02 2.76e-03 1.03e-02 9.88e-03 2.46e-02 9.88e-03 2.46e-02 9.57e-03 3.68e-02 7.46e-03 2.48e-02 2.21e-03 7.93e-03 2.21e-03 7.93e-03 9.79e-03 2.06e-02 9.79e-03 2.06e-02 7.85e-03 2.89e-02 6.60e-03 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 7 8 8 8 8 8 8 8 8 8 8 8 8 9 9 9 9 9 9 9 9 9 9 9 9 3 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 Table .1 (cont’d) n y y n n y y n n y y n n y y n n y y n n y y n n default 0.7254 JLM 0.6343 default 0.6343 JLM 0.6427 default 0.6427 JLM 0.6543 default 0.6543 JLM 0.6266 default 0.6266 JLM 0.7286 default 0.7286 JLM 0.6855 default 0.6855 JLM 0.6802 default 0.6802 JLM 0.6802 default 0.6802 JLM 0.6559 default 0.6559 JLM 0.6559 default 0.6559 JLM 0.7808 default 0.7808 JLM 0.7282 default 0.7282 0.3146 0.2487 0.2487 0.2310 0.2310 0.2480 0.2480 0.2282 0.2282 0.3404 0.3404 0.2694 0.2694 0.2770 0.2770 0.2770 0.2770 0.2671 0.2671 0.2671 0.2671 0.4074 0.4074 0.3324 0.3324 0.3567 0.1493 0.1493 0.1254 0.1254 0.1563 0.1563 0.1266 0.1266 0.3552 0.3552 0.1744 0.1744 0.1628 0.1628 0.1628 0.1628 0.1618 0.1618 0.1618 0.1618 0.4439 0.4439 0.2572 0.2572 0.2586 0.0019 0.0212 0.0022 0.0237 0.0016 0.0193 0.0019 0.0215 0.0018 0.0211 0.0033 0.0313 0.0381 0.1589 0.0381 0.1589 0.0935 0.2856 0.0935 0.2856 0.0372 0.1590 0.0562 0.2083 2.11e-02 4.46e-05 5.00e-04 4.14e-05 4.41e-04 4.18e-05 4.89e-04 3.50e-05 3.90e-04 1.62e-04 1.86e-03 1.05e-04 1.01e-03 1.17e-03 4.87e-03 1.17e-03 4.87e-03 2.65e-03 8.09e-03 2.65e-03 8.09e-03 5.25e-03 2.25e-02 3.50e-03 1.30e-02 142