ABSTRACT Title of Dissertation: PROTON ENERGIZATION DURING MAGNETIC RECONNECTION IN MACROSCALE SYSTEMS Zhiyu Yin Doctor of Philosophy, 2024 Dissertation Directed by: Professor James F. Drake Department of Physics Magnetic reconnection is a widespread process in plasma physics that is crucial for the rapid release of magnetic energy and is believed to be a key factor in generating non-thermal particles in space and various astrophysical systems. In this dissertation, a set of equations are developed that extend the macroscale magnetic reconnection simulation model kglobal to include particle ions. The extension from earlier versions of kglobal, which included only particle elec- trons, requires the inclusion of the inertia of particle ions in the fluid momentum equation. The new equations will facilitate the exploration of the simultaneous non-thermal energization of ions and electrons during magnetic reconnection in macroscale systems. Numerical tests of the prop- agation of Alfvén waves and the growth of firehose modes in a plasma with anisotropic electron and ion pressure are presented to benchmark the new model. The results of simulations of mag- netic reconnection accompanied by electron and proton heating and energization in a macroscale system are presented. Both species form extended powerlaw distributions that extend nearly three decades in energy. The primary drive mechanism for the production of these nonthermal parti- cles is Fermi reflection within evolving and coalescing magnetic flux ropes. While the powerlaw indices of the two species are comparable, the protons overall gain more energy than electrons and their power law extends to higher energy. The power laws roll into a hot thermal distribution at low energy with the transition energy occurring at lower energy for electrons compared with protons. A strong guide field diminishes the production of non-thermal particles by reducing the Fermi drive mechanism. In solar flares, proton power laws should extend down to 10’s of keV, far below the energies that can be directly probed via gamma-ray emission. Thus, protons should carry much more of the released magnetic energy than expected from direct observations. In Encounter 14 (E14), the Parker Solar Probe encountered a reconnection event in the he- liospheric current sheet (HCS) that revealed strong ion energization with power law distributions of protons extending to 500keV. Because the energetic particles were streaming sunward from an x-line that was anti-sunward of PSP, the reconnection source of the energetic ions was unambigu- ous. Using upstream parameters based on the data observed by PSP, we simulate the dynamics of reconnection applying kglobal and analyze the resulting spectra of energetic electrons and protons. Power law distributions extending nearly three decades in energy develop with proton energies extending to 500keV, consistent with observations. The significance of these results for particle energization in the HCS will be discussed. RECONNECTION IS ALL YOU NEED by Zhiyu Yin Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2024 Advisory Committee: Professor James F. Drake, Chair/Advisor Dr. Michael M. Swisdak, Co-Advisor Professor Thomas M. Antonsen, Jr. Professor Alexander Philippov Professor Christopher Reynolds (Dean’s Representative) © Copyright by Zhiyu Yin 2024 Preface The journey to completing this dissertation has been both challenging and rewarding. My interest in magnetic reconnection was sparked during a cosmic ray seminar in my early graduate studies, where a talk by Jim led me to explore the acceleration mechanisms of plasma particles further. This initial curiosity evolved into a deep commitment to understanding the energization process of plasma particles during macro reconnection. Throughout this research, I encountered numerous challenges, from debugging code to en- suring numerical simulation stability, each serving as an opportunity for growth and learning. The support and guidance of my advisors and colleagues have been invaluable, and their contributions have greatly enhanced the quality of this work. This dissertation is organized into several chapters. Chapter 1 provides an overview of magnetic reconnection, motivations for building the new model, brief reviews of the electron ver- sion of the new model, and a detailed outline of the thesis. Chapter 2 delves into the theoretical background and testing of the new simulation model kglobal, while Chapter 3 provides the sim- ulation results of the model. Chapter 4 gives the applications of kglobal. Chapter 5 offers the conclusions of this research and discusses the future potential applications of the model. I hope that this work will contribute to the field of plasma and heliophysics and inspire future research in these areas. ii Zhiyu Yin July 16, 2024 iii Dedication To my parents and my love, Ping Ren. iv Acknowledgments It hath been a long and arduous quest, At last, it finds its destined place of rest. My heart doth overflow with joy and zest, Emotions deep within, I do attest. With gratitude, my soul is richly blessed, For trials and triumphs that we have addressed. Too many are the words I wish to test, For this endeavor hath been our very best. First of all, I want to extend my most sincere thanks to my research supervisors, Prof. Jim Drake and Dr. Marc Swisdak. Both of you have guided me from confusion to clarity, helping me grow from immaturity to a more seasoned researcher. You provided numerous opportunities I had never dreamed of, broadening my horizons in the realm of real scientists. Your belief in me as a scientist and physicist has been instrumental. Without your support, I couldn’t have come this close to achieving my childhood dream. Thank you, Jim and Marc. This colorful journey we shared will stay with me throughout my life. I am deeply grateful to the members of my thesis and dissertation committee: Prof. Tom Antonsen, Prof. Sasha Philippov, and Prof. Chris Reynolds. Your guidance, assistance, and v advice have been invaluable, greatly enhancing my understanding of physics and supporting my future career. I also want to express my gratitude to my collaborators, Prof. Mihir Desai and Dr. Tai Phan, for providing me with the opportunity to apply my contributions to real-world scenarios. Your collaboration and support allowed me to see the practical impact of my research, bridging the gap between theoretical studies and real-world applications. I would also like to thank my group members: Dr. Meriem Alaoui, Anna Fitzmaurice, Haotian Da, Hanqing Ma, Dr. Dominic Payne, and Qianyi Ruan. Joining the group during the Covid-19 pandemic was challenging, but your communication, advice, and support made all the difference. Thank you for your academic companionship and constant encouragement. My heartfelt thanks go to the alumni of my group: Prof. Michael Shay, Prof. Paul Cas- sak, Prof. Yi-Hsin Liu, Dr. Kevin Schoeffler, Dr. Tak Chu Li, Dr. Joel Dahlin, Dr. Gareth Roberg-Clark, and Dr. Harry Arnold. Your guidance, support, and shared experiences have been invaluable. It is an honor to be part of this distinguished family. My deepest thanks also go to Prof. Peter Shawhan, my academic advisor, and Prof. Eun- Suk Seo, my former research supervisor. Both of you have been instrumental in helping me navigate the challenging early days of my graduate studies. Your support and guidance eased my transition and made me more resilient. I also want to thank the members of my former groups, Dr. Leandro de Paula, Dr. Kichun Kim, Dr. Jayoung Wu, and Dr. Hongguang Zhang, for helping me overcome those difficulties and for your unwavering encouragement. I could not have endured the challenges of this Ph.D. journey without the unwavering sup- port of my friends. My heartfelt thanks go to Dr. Mingshu Zhao for your steadfast support and the countless joyful moments we shared. Dr. Yihang Wang, your kindness and sincerity were vi a constant source of strength, helping me navigate the darkest times. Jiateng Huang and Ziqi Qin, your comforting presence and the relaxation you provided allowed me to find peace under immense pressure. Shuyang Wang, your guidance and assistance in both academic and personal matters were crucial to my progress. Wance Wang, our stimulating discussions, both academic and otherwise, fostered my intellectual growth. Dr. Ziyue Zou, your readiness to assist in times of difficulty was a great reassurance. Thanks to Haoying Dai for your daily interesting conversa- tions and for listening and reasoning through my tough times. Thanks to Shuzhe Zeng for your academic help, which has given me much to think about. I am deeply grateful to Kaixin Huang for every encouraging conversation we had; your support provided the motivation I needed dur- ing challenging times. Dr. Xuchen You and Dr. Xiaoyun Huang, your consistent support and practical assistance were instrumental in overcoming obstacles. Dr. Yalun Yu and Xiangyu Li, your care and the opportunities you shared were greatly appreciated, significantly enhancing my experience. Dr. Haonan Xiong and Dr. Tong Wei, your companionship and thoughtful conversa- tions alleviated my loneliness, bringing much-needed joy and perspective. Dr. Chongchong He, Dr. Kungang Li, Dr. Yixu Wang, and Dr. Gong Cheng, your intriguing ideas and perspectives were always a source of enjoyment and inspiration. Hua Huang and Xi Wang, thank you for your amazing and heartwarming companionship as roommates. I want to express my gratitude to Prof. Yi Wang, Dr. Hongyu Wang, Prof. Jiali Huang, Jie Zhai and Dr. Jing Ao for the insightful conversations we had and the valuable information we shared. During my graduate studies in computer science at the Georgia Institute of Technology, I had the pleasure of gaining incredible experiences working with Yun Zhang, Yisong Chen, Xiaoyue Chen, Azhar Masani, Daniel Lesser, and Jae Hyeon Lee. vii I would also like to extend my heartfelt gratitude to Prof. Stéphane Coutu, my under- graduate research supervisor. You provided exceptional opportunities to work in the lab and introduced me to the fascinating field of physics and the universe. Before your guidance, my dream of becoming a physicist was merely a lofty aspiration. Under your mentorship, I gained not only fundamental knowledge and techniques but also developed the right attitude essential for a physicist. Thank you for your unwavering support and invaluable lessons. I feel fortunate to have attended Pennsylvania State University for my undergraduate stud- ies, where I learned to build my life as an adult and consider my future seriously. During this time, I extend my heartfelt thanks to Zangnan Yu for your companionship wherever you are. I also express my gratitude to Shixin Zhao and Jiaming Zhang for our gatherings and for looking after my family. Thank you, Terry Tang, for your career advice, and Runyu Mao, for our relaxing conversations and your support. I appreciate Wei Cheng for your heartfelt company and Shuhan Tian for your initial companionship when we came to the United States. Thanks to Jiayi Wang and Hui Wang for all your help, communication, and warm invitations to spend time with your family. Thank you, Boris Huang and Mike Wu, for your support and companionship during my difficult time. 星垂平野阔,月涌大江流。作为北大附中的校友,每当回想我的高中,我还是难以 平复自己的内心。令公桃李满天下,何用堂前更种花。感谢张济达老师,为我指引物理 学的道路;感谢金钟鸣老师,既是良师也是益友;感谢王莹老师,在暗黑时期给予我唯 一的光明。这些恩情,铭刻在心,永不忘怀。孤帆远影碧空尽,唯见长江天际流。谢谢 你,给我长久陪伴的杨天骐;谢谢你,给我温柔交流的蔡宇坤,谢谢你;给我情绪抚慰 的郑峒;谢谢你,给我无尽帮助的王天雄;谢谢你,给我有趣灵魂的郝逸洲,谢谢你, 给我中肯鼓励的杨一博;谢谢你们,付震、黄无忌,谢谢你们,袁兆杰、原宇彤,谢谢 viii 你们,秦靖、李思恒,谢谢你,李家鑫,谢谢你,尹义平。 云山苍苍,江水泱泱,先生之风,山高水长。作为北京十五中分校的学生,也作为 曾经十五中初中的实习教师,无论是陶然亭畔的湖光山色,还是盆儿胡同的四季更替, 点滴细节,皆铭记于心。自蒙半夜传衣后,不羡王祥得佩刀感谢郭玉敏老师的深信不 疑,郭莎玲老师的关怀备至,吴晓光老师的雪中送炭,彭硕老师的鼎力相助,王会敏老 师的慧眼识珠,卢宇老师的仗义执言,范晓芸老师的细致入微,以及池晓波老师的教诲 如山。这些点点滴滴,皆为我成长的珍贵财富。还有高军老师,肖颖老师,于冬霞老 师,尽管不是你们的学生,但作为同事,也感谢您们的信任与照顾。还有我的挚友们, 荷笠带斜阳,青山独归远。把感谢送给肖迪,随着海风洒在星海的海滩;把感谢送给吴 慧超,随着夏夜留在丰台的车站;把感谢送给胡井暄,随着红魔呐喊在工体的国安;把 感谢送给王可,随着冬日驻在永恒的泰山;把感谢送给王梦妍,随着浓雾抚慰着天路的 草原;把感谢送给张帆,随着广西回到记忆的云南;把感谢送给李楠,随着单车畅游在 喧闹的二环;把感谢送给郜汝闯,随着哨声回响在费城的激战;把感谢送给周昊,送给 刘翔,送给董毅松,送给每一个你们,我的朋友们、同学们,还有我的学生们。 浮云游子意,落日故人情。我也十分有幸,不仅在学校中遇到了良师益友,还在这 一路上还结识了许多朋友,并得到了无数的帮助和支持。我想感谢陈晓航、谢小妍、何 雯,想感谢王丹,想感谢郭鹏,吴振媛,也想感谢揭妤。感谢不只学校中的田野,还有 你们这世界给予我的关切。 这里也想怀念我家的皮皮(狗),自从你离开后,我常常想起你,心里充满感激, 也常在梦中见到你。天堂可好?你一定也在想我吧。 最后,也要感谢我所有的家人,陈星姨妈、钱大行姨夫,陈震夏哥哥、夏蓉容姨 妈、陈健舅舅,陈越舅舅,韩超哥哥、郭碧薇姐姐、邸明明大姑、尹玫二姑、尹丹小 姑、韩晓风姑父、郭是忠姑父,还有阿姨马淑英。空床卧听南窗雨,谁复挑灯夜补衣, ix 特别要感谢我的姥爷陈昌谦,姥姥陈嬗忱,奶奶邸金俊,爷爷尹旭。因为在落日时我曾 看到看到你们抹出粉色的天空,也在睡梦里探索了你们创造的树洞。 搴帷拜母河梁去,白发愁看泪眼枯。惨惨柴门风雪夜,此时有子不如无。感谢我的 父亲尹怡青和母亲陈晓千,让我把这一切献给你们。人生三十载,来美十一年,我收到 你们的所给予不只是生命,也有去探索世界的勇气、责任和努力。你们的一句“照顾好自 己,养活好自己”,是送给我成为一个无忧无虑孩子的祝福,而我也在努力成为一个能超 出你们期待的人。 君游东山东复东,安得奋飞逐西风。愿我如星君如月,夜夜流光相皎洁。最后的最 后,把我的感谢留给任凭。曾经的信件,代表我过去的思念;如今的文献,也写下和你 的诗篇;未来的百年,都会是美丽的晴天。感谢你这一路的鼓励、鞭策和陪伴,还有见 证着我的成长,还有还有我们那未来未知却有趣的每一天、每一刻、每一点。 敬谢诸般万象,亦谢此生千事。无论吾身何人,前途何往,皆因这一切,铸吾今时 之身,亦必成未来之才。千磨万击还坚劲,任尔东西南北风。在此留下吾之姓名,尹志 宇是也。 x Table of Contents Preface iv Dedication iv Acknowledgements v Table of Contents xi List of Tables xiii List of Figures xiv List of Abbreviations xvi Chapter 1: Introduction 1 1.1 History of Magnetic Reconnection . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1.1 Historical Theoretical Development . . . . . . . . . . . . . . . . . . . . 2 1.1.2 Multi X-line Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 1.2 Particle Energy Gain in Magnetic Reconnection . . . . . . . . . . . . . . . . . . 10 1.2.1 Fermi Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 1.2.2 Parallel Electric Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 1.2.3 Gradient-B Drift and Betatron Acceleration . . . . . . . . . . . . . . . . 24 1.2.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 1.3 Limitations of Common Plasma Models . . . . . . . . . . . . . . . . . . . . . . 26 1.4 The Original kglobal Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 1.5 Outline of Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 Chapter 2: The Upgraded Computational Model for Particle Energization during Macroscale Reconnection 34 2.1 Model Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 2.1.1 kglobal with Particle Electrons . . . . . . . . . . . . . . . . . . . . . . 34 2.1.2 kglobal with particle electrons and ions . . . . . . . . . . . . . . . . . . 38 2.2 Alfvén Wave Propagation and the Growth of Firehose Modes in an Anisotropic Plasma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Chapter 3: Simultaneous Proton and Electron Energization during Macroscale Mag- netic Reconnection 49 xi 3.1 Simulation Model Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 3.2 Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 Chapter 4: Applications 73 4.1 Solar Flare . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 4.2 Heliospheric Current Sheet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 4.2.1 Simulations of Magnetic Reconnection . . . . . . . . . . . . . . . . . . 78 Chapter 5: Conclusion 81 5.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 5.2 Future Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 5.2.1 Earth’s Magnetotail . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 5.2.2 Coronal Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 5.2.3 Anomalous Cosmic Rays . . . . . . . . . . . . . . . . . . . . . . . . . . 90 5.3 Future Developments and Enhancements of the kglobal Model . . . . . . . . . . 93 Appendix A: Theoretical Derivations 96 A.1 Derivation of the Fluid Ion Momentum Equation . . . . . . . . . . . . . . . . . 96 A.2 Energy conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 Appendix B: Spectral Fitting 106 B.1 Kappa Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 Appendix C: HCS Related Topics 109 C.1 PSP Instrumentation and Current Sheet Coordinate System . . . . . . . . . . . . 109 C.2 Details of HCS Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 Bibliography 116 xii List of Tables 3.1 Parameters for Simulation Domains . . . . . . . . . . . . . . . . . . . . . . . . 52 3.2 Density Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 3.3 Relative Energy and Spectra Parameters . . . . . . . . . . . . . . . . . . . . . . 72 5.1 Upstream temperature of electrons and protons in magnetotail . . . . . . . . . . 87 xiii List of Figures 1.1 Sweet-Parker Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.2 Petschek Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 1.3 Multi-Island Magnetic Reconnection Model . . . . . . . . . . . . . . . . . . . . 11 1.4 Fermi Reflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 1.5 Contracting Magnetic Island . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 1.6 Particle Energy Gain from the Merging of Two Islands . . . . . . . . . . . . . . 23 1.7 Fermi Acceleration vs Electric Fields . . . . . . . . . . . . . . . . . . . . . . . . 25 1.8 Spectra from Original kglobal . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.1 Alfvén Wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 2.2 Firehose Instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.1 Evolution of Particle Proton Density . . . . . . . . . . . . . . . . . . . . . . . . 54 3.2 Energy and Firehose Parameter at Late Time . . . . . . . . . . . . . . . . . . . . 56 3.3 Time Evolution Spectra of Energetic Protons . . . . . . . . . . . . . . . . . . . 58 3.4 Spectra vs Different Simulation Domain Sizes . . . . . . . . . . . . . . . . . . . 60 3.5 Spectra vs Fraction of Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 3.6 Comparison of Initial and Late time Spectra between Protons and Electrons . . . 64 3.7 Spectra Fitting and Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . 66 3.8 Linear-Linear Plot of Initial and Late-Time Spectra for Protons and Electrons . . 69 4.1 PSP Crossing Data and 2D image of Simulation . . . . . . . . . . . . . . . . . . 77 4.2 Comparison of Spectra from HCS Observations vs Simulations . . . . . . . . . . 79 5.1 Comparison of Initial and Late-Time Spectra between Protons and Electrons for Magnetotail Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 5.2 Time Evolution Spectra with Pickup Ions in ACR Simulations . . . . . . . . . . 94 C.1 Time development of flux ropes in a simulation of the E14 HCS reconnection event113 C.2 The firehose parameter and the out-of-plane magnetic field Bz late in time from the simulation in Fig C.1. The color bar indicates the values of α = 1− 4π(p∥ − p⊥)/B 2 and Bz. Negative α values identify regions that are firehose unstable, where the plasma has undergone significant energization. In these areas, the en- ergetic particles influence the dynamics of flux ropes, suppressing reconnection and the merging of islands. Within the flux ropes, Bz increases significantly compared to its upstream value due to plasma compression in the reconnecting current layer. This enhancement in Bz reduces the magnetic shear during flux rope mergers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 xiv C.3 Differential flux of electrons and protons at early and late times from the simula- tion of Fig. C.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 xv List of Abbreviations ACR Anomalous Cosmic Rays DSA Diffusive Shock Acceleration E14 Encounter 14 GCR Galactic Cosmic Rays HCS Heliospheric Current Sheet KGLOBAL Kinetic Global MHD Magnetohydrodynamic MMS Magnetospheric Multiscale Mission NASA National Aeronautics and Space Administration NSF National Science Foundation PIC Particle-In-Cell PSP Parker Solar Probe RHESSI Reuven Ramaty High Energy Solar Spectroscopic Imager SCS Small-scale Current Sheet xvi Chapter 1: Introduction 1.1 History of Magnetic Reconnection Plasmas, the fourth state of matter, constitute approximately 99.999% of all observable matter, highlighting the importance of their study in understanding the fundamental nature of the universe [1, 2]. Central to plasma physics is the process of magnetic reconnection, which enables the transformation of magnetic energy into plasma energy by altering the magnetic topology [3, 4]. Magnetic reconnection allows for the rapid reconfiguration of magnetic fields, facilitating the release of magnetic energy into kinetic and thermal energy, as well as the acceleration of charged particles into non-thermal distributions [5, 6, 7, 8], particularly power-law distributions that extend over many orders of magnitude in energy [5, 7, 9, 8, 10]. This process is crucial in the dynamics of a wide range of astrophysical systems, including the solar corona, planetary magnetospheres, magnetic sectors of the solar wind, pulsar wind nebulae, and black holes [11, 6, 5, 12, 13, 7, 8, 14, 15, 16, 17]. In laboratory settings, magnetic reconnection is identified as the primary cause of the sawtooth crash in fusion devices that leads to disruptions in confinement [18, 19, 20, 21]. The significance of magnetic reconnection in both astrophysical and laboratory contexts emerged from early observations of solar flares, which are sudden bursts of brightness above the sun’s surface [22, 23, 24]. The first recorded solar flare, known as the ‘Carrington Flare,’ was 1 documented in 1859 by R. C. Carrington and R. Hodgson [25, 26, 27]. This event disrupted telegraph systems and produced intense aurorae. Later, Giovanelli noted that solar flares are associated with magnetic null points, regions where the magnetic field reverses direction [28]. Building on these observations, James Dungey proposed a model in which oppositely oriented magnetic field lines separate and reconnect due to diffusion, forming a thin current sheet, and applied reconnection to Earth’s magnetosphere [29], which laid the foundation for the modern understanding of magnetic reconnection. 1.1.1 Historical Theoretical Development 1.1.1.1 The Sweet-Parker Model The Sweet-Parker model is a foundational framework in the theory of magnetic reconnec- tion, particularly within the context of astrophysical plasmas such as those found in the solar atmosphere [30, 31]. Developed in the mid-20th century by P.A. Sweet and E.N. Parker, this model provides a steady-state description of the reconnection process, where magnetic field lines of opposite polarity are brought together, reconfigure, and release stored magnetic energy. In the Sweet-Parker model (shown in Fig. 1.1), reconnection occurs in a two-dimensional geometry where oppositely directed magnetic fields flow towards each other, forming a thin, elongated current sheet at the boundary between the opposing magnetic fields. The magnetic fields then reconnect across this current sheet. The resulting bent field lines expand downstream, converting magnetic energy into kinetic and thermal energy, as well as accelerating particles. Magnetic reconnection is self-driven since the expanding, reconnected field lines lead to a pres- sure reduction at the magnetic x-line, which pulls in the upstream plasma and magnetic field to 2 sustain the process. The Sweet-Parker model operates under several key assumptions that simplify the complex dynamics of magnetic reconnection. First, the model assumes a steady-state scenario (shown in Fig. 1.1), where the inflows and outflows of plasma (with velocities Vin and Vout, respectively) and magnetic fields (Bin) are balanced, meaning the amount of material entering the current sheet is equal to the amount leaving it. Additionally, the model presumes an elongated current sheet structure (L ≫ δ), where both the kinetic energy of the inflowing plasma and the magnetic energy of the outflow are neglected. Another important assumption is that resistivity (η) is only considered within the current sheet, ignoring any effects outside this region. The model also disregards 3D effects, focusing only on 2D dynamics. First, the mass flux into the current sheet must equal the mass flux out, which can be expressed as: LVin ∼ δVout (1.1) Similarly, the magnetic energy flux entering the current sheet is balanced by the kinetic energy flux exiting it. This conservation of energy can be expressed as: LVin ( B2 in 4π ) ∼ δVout ( ρV 2 out 2 ) (1.2) By combining these equations, it is possible to show that the outflow velocity, Vout, scales with the upstream Alfvén speed VA, defined as: 3 2L 2𝛿 𝐵𝑖𝑛 𝑉𝑖𝑛 𝑉𝑜𝑢𝑡 𝑉𝑖𝑛 𝐵𝑖𝑛 Figure 1.1: Cartoon diagram of the Sweet-Parker Model. 4 Vout ∼ VA ≡ Bin√ 4πρ (1.3) Additionally, due to the 2D and steady-state assumptions, the out-of-plane electric field is uniform inside and outside the layer, which gives: VinBin c ∼ ηJ (1.4) Using Ampère’s law, the current density J is obtained as: J ∼ c 4π Bin δ (1.5) The inflow velocity is then determined by balancing the resistive diffusion of magnetic flux with the convective inflow, yielding: Vin ∼ Dη δ (1.6) where Dη ≡ ηc2 4π is the resistive diffusivity. The Sweet-Parker model provides a scaling law for the reconnection rate, which is propor- tional to the inflow velocity and inversely proportional to the square root of the Lundquist number S. The dimensionless reconnection rate r is given by: r ∼ Vin VA ∼ 1 S1/2 (1.7) where the Lundquist number, defined as: 5 S ≡ LVA Dη (1.8) measures the ratio of the resistive diffusion time scale to the Alfvén wave crossing time scale. Typically, S is very large in astrophysical plasmas, often between 109 and 1020, leading to the prediction of slow reconnection rates in the Sweet-Parker model. This slow reconnection rate is inadequate to explain the rapid energy release observed in phenomena like solar flares. Thus, while the Sweet-Parker model successfully demonstrates that reconnection occurs faster than by diffusion alone, it requires additional mechanisms such as turbulence or 3D effects to account for the fast energy dissipation observed in explosive astrophysical events [32, 33]. 1.1.1.2 The Petschek Model and Advances in Collisionless Reconnection The Petschek model, proposed by H.E. Petschek, advanced the theory of magnetic recon- nection by suggesting that the presence of slow-mode shocks between the inflow and outflow regions (shown in Fig.1.2) can substantially increase the rate of magnetic energy dissipation by shortening the length of the current sheet (L ≫ L∗), which was taken to be macroscopic in the Sweet-Parker model [34]. These shocks facilitate the conversion of magnetic energy into kinetic and thermal energy. This model departs from the earlier Sweet-Parker framework by allowing for faster reconnection with reconnection rate r ∝ 1 lnS , which could potentially explain the rapid energy release observed in various astrophysical phenomena. 6 Despite the theoretical appeal of the Petschek model, however, numerical simulations have shown that this configuration only emerges as a solution of the MHD equations under certain conditions. Specifically, the formation of slow-mode shocks and the resulting fast reconnection rate are sustained only when there is a localized enhancement in the resistivity within the current sheet [35]. In the absence of such localized resistivity, the system tends to revert to the slower Sweet-Parker configuration, where the reconnection rate is limited by the resistive diffusion of the magnetic field [36]. On the other hand, in the high Lundquist limit, magnetic reconnection in large systems becomes highly turbulent [36, 37, 38]. Instead of a single, smooth reconnection region, the current sheet can fragment and break up into multiple localized structures called magnetic flux ropes. These flux ropes are small-scale magnetic loops or “islands” (also known as plasmoids) of magnetic field lines that are twisted around each other. This fragmentation happens due to instabilities such as the tearing mode instability, which allows the current sheet to break into smaller regions of intense current [36, 39, 40, 41, 42, 43]. The presence of multiple flux ropes suggests a more complex reconnection process, with many small regions where reconnection is happening simultaneously rather than a single large reconnection outflow. This model will be presented in the next subsection. 1.1.2 Multi X-line Model Unlike the Sweet-Parker and Petschek models, the fragmentation of current sheets and the development of plasmoids seems to be a common feature of systems undergoing reconnection [36, 37]. Plasmoids have been detected both in solar flares [44, 45] and the Earth’s magnetotail [46, 47, 48, 49, 50, 51]. Numerous simulations, ranging from fluid models [36, 52] to fully ki- 7 2𝐿 2𝐿∗ Figure 1.2: Cartoon diagram of the Petschek Model. 8 netic ones [39], have reported the occurrence of plasmoid formation. Plasmoids have become an important component in theoretical models of magnetic reconnection. For instance, they can increase the rate of magnetic energy release by reducing the effective length of the Sweet-Parker current sheet and/or activating anomalous resistivity effects associated with small-scale plasma dynamics. The scenario involving multiple plasmoids has been proposed to explain the acceler- ation of electrons during reconnection events [53]. Recently, a magnetic reconnection plasmoid model has been developed for flares associated with Sagittarius A*, the supermassive black hole at the center of our Galaxy [54, 55]. This process results in what is known as multi-island reconnection, where multiple mag- netic islands or structures emerge during reconnection. A schematic representation of multi- island magnetic reconnection is provided in Fig. 1.3. The size of these islands ranges from electron-scale to above the ion-scale. Neighboring islands with oppositely directed magnetic fields give rise to localized reconnection events. Such small-scale reconnections can be initiated by turbulence caused by the twisting and braiding of magnetic flux tubes. In large-scale recon- nection events, such as those associated with solar flares, plasmoids created by current-driven instabilities like the tearing mode instability occupy the magnetic field reversal region. In multi- island magnetic reconnection, beyond the acceleration facilitated by the reconnection-induced electric field, the contraction and merging of the magnetic islands also contribute to particle ac- celeration [39, 56, 57, 58, 59, 60], as will be discussed further in Sec. 1.2.1. 9 1.2 Particle Energy Gain in Magnetic Reconnection To investigate the various factors influencing electron energy evolution, we begin by con- sidering the guiding-center approximation as described by Northrop [62]. For simplicity, we specifically treat electrons here, although the extension to protons/ions is straightforward. The change in energy, ϵ, of an electron in the guiding-center frame can be expressed as: dϵ dt = ( µ γ ) ∂B ∂t − e(v∥b+ vc + vg) · E (1.9) Here, b = B/|B| is the unit vector along the magnetic field, µ = meγ 2v2⊥/2B represents the magnetic moment, γ is the relativistic Lorentz factor, v∥ = v · b denotes the parallel component of the velocity, and vc and vg are the curvature drift and gradient-B drift velocities, respectively, given by: vc = v2∥b Ωce × κ (1.10) vg = v2⊥b 2Ωce × ∇B B (1.11) In these equations, Ωce = eB/γmec is the electron cyclotron frequency, and the magnetic curva- ture is κ = b · ∇b. Equation (1.9) can be reformulated in a more insightful manner: dϵ dt = γmev 2 ∥(uE · κ) + qE∥v∥ + µ γ ( ∂B ∂t + uE · ∇B ) (1.12) 10 Figure 1.3: Left Panel: (a) A diagram illustrating volume-filling magnetic islands generated by tearing instability in the magnetic field reversal region. (b) A PIC simulation depicting island for- mation during magnetic reconnection. Here, jez represents the out-of-plane current density[53]. Right Panel: (c) A schematic illustration of neighboring magnetic islands with locally oppositely directed magnetic fields (shown by arrows), undergoing magnetic reconnection[59]. The entire figure is reprinted from the reference[61]. 11 This expression highlights three distinct mechanisms by which particles can gain energy: 1. Fermi Acceleration: The first term, γmev 2 ∥(uE · κ), describes Fermi acceleration, where particles gain energy by reflecting off contracting magnetic field lines. The term uE · κ indicates that this process is related to the curvature of the outflow magnetic fields. In this scenario (shown in Fig. 1.4), as a magnetic field line contracts away from an X-line with an Alfvénic outflow velocity CA, a charged particle moving along the field line experiences a head-on collision with the moving field. For instance, if the curved field line moves to the left with the Alfvén velocity and a charged particle travels rightward with a parallel velocity v∥ = −v0, the particle reflects elastically from the field line. After the reflection, the particle’s velocity increases to v∥ = v0 + 2CA, resulting in a potentially significant energy gain. 2. Acceleration by Parallel Electric Fields (E∥): The second term, qE∥v∥, represents the acceleration of particles due to electric fields that are aligned with the magnetic field. There are two types of electric fields associated with this term: the kinetic-scale electric field, a highly localized inductive parallel electric field that exists only at small scales because highly mobile electrons tend to short-circuit these fields, limiting their range [63]; and the macroscale electric field, generated by parallel electron pressure gradients, which plays a significant role in particle heating[64, 65]. 3. Betatron and Gradient-B Drift Acceleration: The third term, (µ/γ) (∂B/∂t+ uE · ∇B), encompasses both betatron acceleration and gradient-B drift acceleration, both of which are associated with the conservation of the magnetic moment µ. • Betatron Acceleration: The component (µ/γ)∂B/∂t corresponds to betatron ac- 12 celeration. This process occurs as the magnetic field strength B changes over time, with the particle’s perpendicular velocity adjusting to maintain the conservation of µ. During magnetic reconnection, the perpendicular magnetic field decreases, and as a result, the perpendicular energy ϵ⊥, becomes smaller. • Gradient-B Drift Acceleration: The term (µ/γ)uE ·∇B represents gradient-B drift acceleration. This mechanism takes place when a particle moves across magnetic field gradients, causing B to vary as a function of position. Due to the conservation of µ, the particle gains or loses energy as it drifts through regions with different magnetic field strengths via the E×B drift. 13 These mechanisms accelerate particles by affecting different components of their energy. The first and second mechanisms primarily enhance the parallel component of the energy. Par- allel electric fields tend to generate beams of particles aligned with the magnetic field, which may eventually thermalize due to various instabilities. Fermi acceleration, which is directly tied to parallel energy, increases the parallel temperature T∥. In contrast, the betatron acceleration and gradient-B drift mechanisms primarily increase the perpendicular energy in conjunction with a rising magnetic field strength. Significant energy gain via betatron acceleration as well as gradient-B drift typically requires a substantial increase in the magnetic field, which does not oc- cur during magnetic reconnection. Thus, betatron acceleration and gradient-B drift are typically energy sinks during reconnetion. 1.2.1 Fermi Acceleration What is now known as Fermi acceleration was first proposed by Enrico Fermi in 1949 to explain the origin of cosmic rays [66]. In his original formulation, Fermi proposed that charged particles could gain energy through interactions with randomly moving ‘magnetic irregularities’ in the interstellar medium. When a particle collides head-on with a moving magnetic irregularity, it gains energy, whereas a tail-on collision results in a loss of energy. Fermi argued that head-on collisions are statistically more likely than tail-on collisions, leading to a net average acceleration of the particles. The average gain in energy per collision can be expressed as: δϵ ∝ ( V c )2 ϵ (1.13) where ϵ is the energy of the particle and V is the velocity of the magnetic irregularity. Fermi 14 also considered that energetic particles have a finite lifetime, during which they can lose en- ergy through collisions with ambient particles. The evolution of the distribution f ≡ f(ϵ, t) of energetic particles can be described by the following equation: ∂f ∂t + ∂ ∂ϵ (ϵ̇f) = − f T . (1.14) Here, T represents the characteristic lifetime of the energetic particles, and ϵ̇ is the rate of energy gain which, for interactions with magnetic irregularities, is given by: ϵ̇ = ϵ ( V c )2 1 τ , (1.15) where τ is the average time between interactions. Assuming a steady state (∂/∂t = 0), the above equation simplifies to: ∂ ∂ϵ (αϵf) = − f T , (1.16) where α = (V/c)2/τ . The solution to Eq. 1.16 takes the form of a power law: f(ϵ) ∝ 1 ϵ1+(αT )−1 , (1.17) which is the functional form required to explain the observed cosmic ray spectrum. Fermi’s original concept was later adapted to describe acceleration at astrophysical shocks, leading to the development of what is now known as Diffusive Shock Acceleration (DSA) [67, 68, 69, 70]. A shock propagating through a plasma compresses and slows the upstream plasma as it crosses the shock. Particles undergo acceleration as they reflect back and forth across the shock 15 [71, 72]. This process is a specific case of Fermi acceleration, often referred to as first-order Fermi acceleration since the rate of energy gain of a particle is proportional to the shock velocity. It is the leading model for the acceleration of galactic and intergalactic cosmic rays [73]. In addition to the specific case of first-order Fermi acceleration at shocks, Fermi’s gen- eral concept has broader applications in particle acceleration processes. The second-order Fermi acceleration mechanism occurs in more turbulent environments, where particles gain energy through repeated interactions with randomly moving magnetic irregularities [74, 75]. Unlike the first-order process, where energy gain is proportional to the velocity of the shock, the second- order process is proportional to the square of the velocity (V/c)2. This makes second-order Fermi acceleration less efficient but still significant in environments dominated by turbulence. Fermi acceleration, whether first-order or second-order, fundamentally relies on the con- cept of particles gaining energy through interactions with moving magnetic structures. These in- teractions can be modeled in various astrophysical settings, such as supernova remnants [76, 77], the solar wind [78], and the interstellar medium [79, 80], where they contribute to the acceleration of particles to high energies, forming the power-law energy spectra observed in cosmic rays. 1.2.1.1 Fermi Acceleration in Reconnection During magnetic reconnection, magnetic islands or flux ropes are formed as elongated magnetic structures that contract at both ends with Alfvén speed, CA (See Fig. 1.5). Particles trapped within these islands can reflect multiple times between the ends, gaining energy with each reflection. If the particle’s velocity is significantly greater than the Alfvén speed, which is typically the case for electrons, the energy gain per reflection can be substantial. 16 During magnetic reconnection with adiabatic invariance, the conservation of parallel action can be applied to determine the particle energy gain in an elongated magnetic island. The parallel action is given by: J = v||l (1.18) where l is the circumference of the ellipse, whose expression was given by Ramanujan[81]: l = π[3(L+ ω)− √ (3L+ ω)(L+ 3ω)] (1.19) In Fig. 1.5, the circumference of the ellipse of the island for L ≫ w is: l ≃ (3− √ 3)πL ≃ 4L (1.20) And thus, with the contracting rate L̇ = CA, the derivative of l is as follows, l̇ ≃ 4L̇ = −4CA (1.21) On the other hand, the time derivative of J is: dJ dt = 0 = v̇||l + v||l̇ (1.22) which leads to the change in the parallel velocity: v̇∥ = −v|| l̇ l = −v∥ (−4CA) 4L = v|| CA L (1.23) 17 As a result in Eq.1.23, particle energy gain in a contracting island follows a first-order Fermi process. 18 𝐶! 𝑣" + 2𝐶! 𝑣" Figure 1.4: Cartoon diagram of a charged particle reflecting from an outflow magnetic field line at the Alfvén speed. The particle velocity increases by 2CA and particle’s final velocity is v0+2CA. 19 Another perspective on the energy increase can be understood through the merging of two magnetic islands. As illustrated in Fig. 1.6, two islands, each with an initial radius ri, merge to form a larger island with a final radius rf . Initially, the path length is approximately li ≃ 2(2πri), and after the merger, the final path length becomes lf = 2πrf . During this process, the total area enclosed by the islands remains conserved [82], leading to the equation: 2(πr2i ) = πr2f (1.24) from which we obtain: rf = √ 2ri (1.25) As a result, the ratio of the final to initial path lengths is given by: lf li = rf 2ri = 1√ 2 . (1.26) By conserving the parallel action, we can then deduce the velocity scaling: vf = √ 2vi (1.27) Thus, the energy (∝ v2) doubles during the merging of the two islands. 20 2𝐿 2𝜔 𝐶𝐴𝐶𝐴 Figure 1.5: Cartoon diagram of a magnetic island contracting at the Alfvén speed. 21 1.2.2 Parallel Electric Fields A seemingly straightforward mechanism for particle acceleration in reconnection regions involves parallel electric fields. Two types of parallel electric fields will be explored here, large- scale parallel fields associated with the gradient of electron pressure, and kinetic-scale parallel fields. The role of these fields in particle acceleration has been the subject of numerous studies [8, 83, 64]. In a previous study exploring the strong guide field regime [84], the kinetic-scale paral- lel electric field (E∥) drove essentially all of the electron energy gains, yet it failed to generate an energetic non-thermal component, and in particular did not produce a power-law distribu- tion. Fig.1.7 shows the cumulative energy gain of electrons resulting from different accelera- tion mechanisms—including parallel electric fields (E∥), Fermi reflection, and betatron acceler- ation—across varying guide field strengths (bg), highlighting that E∥ plays an important role in driving electron energy gains in strong guide field regimes. In the scenario with a weak guide field (as shown in Fig.1.7(a)), the dominant contributions come from Fermi reflection (positive) and betatron acceleration (negative), which partially offset each other, resulting in the bulk of the electron energy gain, while the effect of parallel electric fields is negligible. In the simula- tion with bg = 4 (as shown in Fig.1.7(c)), all terms, including parallel electric fields (E∥), have insignificant contributions to the electron energy gain. In the intermediate case with bg = 1 (as shown in Fig.1.7(b)), both Fermi reflection and E∥ play significant roles. It is important to note that the total electron energy gain is approximately 50% higher in the bg = 0.2 case compared to the bg = 4.0 case. In summary, Fig.1.7 reveals how the contribution of E∥ compares to other mechanisms, showing that the influence of E∥ only remains dominant for higher bg (under higher 22 𝑟𝑖 𝑟𝑖 𝑟𝑓 Figure 1.6: Cartoon diagram of the merger of two islands with identical radii and magnetic field strengths. 23 guide fields, the reconnection produce almost no energetic electrons [84, 85]), with implications for understanding the efficiency of electron energization during magnetic reconnection. In a later study [86] it was found that for low guide fields the electric field acts as an energy sink instead of a driver for highly energetic electrons and primarily contributes to heating the lower energy range. (Note: In this dissertation, energization refers to accelerating particles to non-thermal energies, often leading to the formation of a power-law distribution, while heating refers to accelerating particles thermally, resulting in a Maxwellian distribution.) On the other hand, large-scale parallel electric fields, extending beyond localized boundary layers, enhance the process described in Fig.1.4 by confining electrons within the reconnection exhaust, allowing them to experience multiple Fermi accelerations[87, 64, 65, 88]. Furthermore, the study [85] shows that the heating by electric fields remains nearly constant for electrons under different guide field strengths at late times. 1.2.3 Gradient-B Drift and Betatron Acceleration The betatron mechanism and gradient-B drift are two fundamental processes that play a significant role in the acceleration of charged particles in both laboratory[89, 90] and astrophysi- cal plasmas[91]. The magnetic moment of a particle µ is an adiabatic invariant. This invariance holds as long as the temporal scales involved are much longer than the particle’s cyclotron period. Consequently, changes in the magnitude of the magnetic field directly influence the evolution of the particle’s perpendicular energy ϵ⊥. Betatron acceleration occurs when the magnetic field strength B increases over time, causing the particle to adjust its perpendicular velocity to main- tain a constant magnetic moment. 24 Figure 1.7: (a)–(c) Cumulative electron energy gain from different acceleration mechanisms for various guide field strengths (bg). The red, blue, and magenta curves represent contributions from Fermi reflection, parallel electric fields (E∥), and betatron acceleration, respectively. The black dashed line shows the total energy gain. Panels (a), (b), and (c) correspond to bg = 0.2, bg = 1.0, and bg = 4.0, respectively. (d) Normalized total electron energy gain as a function of guide field strength (bg). Symbols represent different contributions: squares for total energy, diamonds for Fermi reflection, and stars for E∥. The red line indicates a fit of the form ∼ [1 + 4b2g] −1. Reprint from [84]. 25 During magnetic reconnection, the release of magnetic energy generally leads to a reduc- tion in the magnetic field strength. Given that betatron acceleration and gradient-B drift rely on an increase in B to effectively impart energy to particles, these mechanisms are not expected to play a significant role in the context of the simulations presented in this dissertation[92, 86]. Therefore, the impact of betatron acceleration and gradient-B drift is considered negligible in the modeled scenarios. 1.2.4 Conclusion Based on the discussions above, the dominant drive mechanism of particle energy gain during magnetic reconnection is Fermi reflection in growing and merging magnetic flux ropes [39, 56, 92, 93, 86, 94]. Magnetic reconnection produces bent magnetic field lines that form an Alfvénic exhaust, carry energy away from the x-line, and transfer energy from the field to the surrounding plasma [31, 95]. Particles gain energy through repeated reflections in contracting magnetic field lines, thereby leading to the observed power-law tails in particle energy distribu- tions [39, 92, 86]. The rate of energy gain from Fermi reflection is proportional to a particle’s energy and is therefore greatest for the most energetic particles. 1.3 Limitations of Common Plasma Models Both particle-in-cell (PIC) and magnetohydrodynamic (MHD) [36, 96, 97] models have limitations that compromise their utility for studying non-thermal particle energization. PIC is a widely used computational technique for simulating plasma dynamics in both lab- oratory and astrophysical contexts. This method is particularly powerful for studying systems 26 where the behavior of individual charged particles and their interactions with electromagnetic fields are crucial, such as in magnetic reconnection, shock waves, and wave-particle interactions. PIC simulations correctly model kinetic physics by resolving kinetic scales [98, 99, 100, 101, 86, 94]. “Macro-particles” that represent a collection of real particles move through the simulation domain. The electric and magnetic fields are evaluated at the gridpoints surrounding the cells containing macro-particles and in turn accelerate the particles via Lorentz forces. This model is ideal for simulations at kinetic scales, but it has limitations at macroscales due to the computa- tional expense, which prevents the simulation of systems significantly larger than kinetic scales. The spatial scales associated with solar flares can exceed 104 km, while the Debye length—a measure of the scale of kinetic effects—can be on the order of centimeters, representing a scale separation of 1010. Hence, computational costs drastically limit the ability of PIC simulations to simulate the extensive range of scales occurring in magnetic reconnection events. A consequence of this inadequate separation of scales is the demagnetization of energetic electrons and ions in PIC simulations of reconnection, which in turn may inhibit particle energy gain and prevent the formation of the extended power laws that are observed in nature [86, 94]. MHD combines the principles of fluid dynamics and electromagnetism to describe the be- havior of electrically conducting fluids such as plasmas. The MHD model is particularly useful for studying large-scale plasma dynamics in astrophysical and laboratory settings, where the fluid approximation is valid and kinetic effects are less dominant. Hence, in MHD, small-scale motions are ignored. While capable of representing macroscopic events, MHD simulations fre- quently assume the plasma is thermal (i.e., represented by a single temperature) and hence by definition cannot produce non-thermal particles. The electric and magnetic fields from MHD models have been used to explore non-thermal energization of test particles. However, these test 27 particles do not interact self-consistently with the fields so energy is not conserved, which limits the test particle model’s fidelity [102]. 1.4 The Original kglobal Model To address these challenges, a new formalism was developed that permitted the first self- consistent simulations of electron energization during magnetic reconnection in a macroscale system [103, 88, 85]. The model was named kglobal, standing for ”kinetic global” and originally did not consider ion energization. To clearly distinguish it from the version that does consider ions (and serves as the main focus of this dissertation), the version that explores electron acceler- ation exclusively is referred to as the original version. The theoretical background and details of the original version of the simulation model will be presented in Sec 2.1.1. The original kglobal model combines aspects of both the PIC and MHD descriptions to address electron acceleration in large-scale astrophysical reconnection [103, 88]. Particle electrons are distributed on the MHD grid, as in PIC models, but are evolved with MHD fields using the guiding-center approximation, where all kinetic scales, including the Larmor radii of particles, are ordered out of the equations. Notably, kglobal includes the effects of Fermi reflection within evolving magnetic flux ropes and the pressure anisotropy that feeds back on the MHD fields to reduce the tension force driving magnetic reconnection. Magnetic reconnection simulations with kglobal produced, for the first time, extended power-law spectra of non-thermal electrons that aligned with existing observa- tional data [85]. This notable result can be seen in Fig. 1.8, which presents a log-log plot of the differential number density versus the normalized energy. The results show the temporal evolu- tion of the electron distribution in a simulation with a modst guide field, Bg/B0 = 0.25, where 28 an initial Maxwellian spectrum develops an energetic tail that hardens over time, eventually sta- bilizing into a power law. The figure also demonstrates that smaller guide fields lead to harder spectra that extend to higher energies. Additionally, the late-time distributions of energetic elec- trons indicate that while the power-law indices vary weakly with the domain size, the maximum energy of the electrons increases with larger system sizes. 29 The implications of the original kglobal simulations were noteworthy. However, observa- tions show that magnetic reconnection can simultaneously produce non-thermal electrons and non-thermal protons [5, 104]. Hence, developing a self-consistent simulation model is vital to elucidate the mechanisms of ion and electron energization during magnetic reconnection and the resulting partitioning of energy between the two species. In Chapter 2, we extend the kglobal equations to include particle ions, which requires the inclusion of the inertia of the particle ions in the bulk fluid ion momentum equation. The resulting equations will facilitate the exploration of the self-consistent energization of both ions and electrons during magnetic reconnection. Based on these equations, the current version also includes particle protons and produces non-thermal spectra for both species. The proton distribution features a non-thermal tail that extends over more than three decades in energy. 1.5 Outline of Thesis In Chap. 2, we will revisit the original kglobal model in detail, show additions made in the upgraded model and present tests of the upgraded model that verify its capabilities. The model consists of an MHD backbone with macroparticle protons and electrons distributed on the MHD grid. The guiding center equations with the MHD electric and magnetic fields describe the mo- tion and energy gain of the macroparticle protons and electrons. These particles feedback onto the MHD fluids and fields and conserve energy. We show kglobal correctly simulates Alfvén waves in the presence of a pressure anisotropy, which is crucial for modeling Fermi reflection since pressure anisotropy can reduce magnetic field tension and the associated Fermi drive mech- anism. Specifically, we demonstrate that the model reproduces the correct growth rate of the 30 Figure 1.8: A log-log plot of the differential number density versus the normalized energy. Panel (a) shows the temporal evolution of the electron distribution in a simulation with Bg/B0 = 0.25. The initial Maxwellian spectrum develops an energetic tail that hardens over time and eventually stabilizes into a power law. Insert in Panel a shows the late time spectra for different guide fields. Smaller guide fields result in harder spectra that extend to higher energies. Panel b shows the late- time distributions of energetic electrons in a simulation with Bg/B0 = 0.25 for various values of Sν (effective system size) reveal that the power-law indices of the spectra vary weakly with the domain size. The notable difference is that as the size of the system increases, the maximum energy of the protons also increases. Reprinted with permission from [85] 31 linear firehose instability, which onsets when the local field-line tension goes to zero, using the same techniques as used in the original electron model [103]. In Chap. 3, we present results from 2D magnetic reconnection using the upgraded kglobal model. Reconnection begins from particle noise and produces multiple islands (or flux ropes since they consist of an axial magnetic field wrapped by the inplane magnetic field) that merge, forming larger islands and accelerating electrons and protons. The pressure anisotropy from the plasma macroparticles becomes large enough to eliminate the magnetic tension force along out- flow exhausts and within islands when particle acceleration is strong. Both species form extended powerlaw distributions that extend nearly three decades in energy. The primary drive mechanism for the production of these nonthermal particles is Fermi reflection within evolving and coalesc- ing magnetic flux ropes. While the powerlaw indices of the two species are comparable, the protons overall gain more energy than electrons and their power law extends to higher energy. The power laws roll into a hot thermal distribution at low energy with the transition energy oc- curring at lower energy for electrons compared with protons. A strong guide field diminishes the production of non-thermal particles by reducing the Fermi drive mechanism. In Chap. 4, we explore the complex processes of charged-particle acceleration during mag- netic reconnection, focusing on solar flares and the heliospheric current sheet (HCS). The chapter begins with a discussion of solar flares, highlighting the role of magnetic reconnection in trans- forming magnetic energy into particle energy and the formation of non-thermal particle popu- lations. Observational data from instruments like RHESSI are discussed to provide context for the particle acceleration mechanisms in solar flares. The chapter then shifts focus to the HCS, the largest current sheet in the solar system, analyzing data from the Parker Solar Probe (PSP) during Encounter 14. The detection of sunward-directed reconnection exhausts and the observa- 32 tion of proton energy spectra are discussed. Using the kglobal model, simulations are performed to investigate the underlying physical mechanisms driving extreme proton energization observed in the HCS. The results demonstrate that merging magnetic flux ropes during reconnection can accelerate protons to energies far exceeding the available magnetic energy per particle. This chapter combines observational insights with advanced simulations to deepen our understanding of particle acceleration in large-scale reconnection events. In Chap. 5, we present conclusions, discuss the results in the context of previous efforts to explore reconnection-driven plasma particle acceleration, and suggest future extensions and potential applications of the kglobal model. 33 Chapter 2: The Upgraded Computational Model for Particle Energization dur- ing Macroscale Reconnection The material in this chapter has been adapted from [105] with permission from the authors. 2.1 Model Equations This chapter presents the upgraded kglobal model. To ensure a comprehensive understand- ing, I will first review the governing equations of the original kglobal model before explaining the enhancements made in the updated version. The upgraded model has been tested and found to accurately represent the propagation of an Alfvén wave in the presence of ion and electron pressure anisotropy, as well as the linear firehose instability. These benchmarks are crucial be- cause bent and reconnected magnetic field lines propagate at the Alfvén speed, and reconnection is moderated by the reduction of field-line tension as a plasma nears firehose marginal stability. 2.1.1 kglobal with Particle Electrons The original model includes three distinct plasma species: the ion fluid (number density ni), the electron fluid (nef ), and particle electrons (nep) [103, 88]. No conversion between species is allowed. The ratio of the particle to electron fluid density is a free parameter. However, electron heating and energization were shown to be insensitive to this ratio, as discussed in reference [85]. 34 The particle electrons are governed by the guiding-center equations [62]. Their perpendicular velocity is determined by conservation of the magnetic moment: µep = p2ep,⊥ 2B = const. (2.1) where pep is the momentum and B is the magnitude of the local magnetic field. Their perpen- dicular motion is given by the usual MHD E×B drift while they stream parallel to the ambient magnetic field at their parallel velocity. The variation in the parallel velocity reflects contributions from Fermi reflection, magnetic mirroring, and the large-scale parallel electric field dpep,∥ dt = pep,∥vE · κ− µe γe b · ∇B − eE∥ (2.2) with vE the E × B drift velocity, b = B/B a unit vector in the direction of the magnetic field, κ = b · ∇b the field curvature, and γe the relativistic Lorentz factor. The guiding-center electrons move through the MHD grid and, crucially, do not need to have a thermal distribution. Their gyrotropic pressure tensor feeds back on the MHD momentum equation. The ion fluid forms an MHD-like backbone that satisfies the usual continuity equation ∂ni ∂t = −∇ · (nivi) (2.3) and a modified momentum equation mi ∂(nivi) ∂t = −∇ · Ti − (∇ · Tef )⊥ + J×B c − (∇ · Tep)⊥ + eniE∥b (2.4) 35 where the ion stress tensor with inertial contributions is Ti = PiI+minivivi (2.5) with Pi the ion scalar pressure and I the unit tensor. The corresponding electron fluid tensor is Tef = PefI+menefγefv 2 ef,∥bb (2.6) with Pef the electron fluid scalar pressure. Finally, Tep is the particle electron gyrotropic stress tensor, including inertial contributions, Tep = Tep∥bb+ Pep⊥(I− bb) (2.7) where Tep∥ is the component of the stress tensor along the magnetic field B, and Pep⊥ is the usual perpendicular pressure, Pep⊥ = ∫ dpe p2e⊥ 2meγe f (2.8) where in the frame drifting with vE = cE × B/B2 there are no perpendicular flows so f = f(x, pe∥, pe⊥, t). Tep∥ includes the mean parallel drifts of the particle electrons, Tep∥ = ∫ dpe p2e∥ meγe f (2.9) with pe∥ being the electron particle parallel momentum. Note that in the ion momentum equation the electron fluid and particle inertia are neglected 36 compared to the ion inertia on the left side of Eq. (2.4), while the electron fluid and particle pressures are retained on the right side of this equation. This result is based on an ordering in which the ion velocity scales as the Alfvén speed and the electron particle and fluid thermal velocities and parallel streaming velocities are of the order of the electron Alfvén speed, CAe =√ mi/meCA, where the Alfvén speed is defined as CA = √ B2/(4πmini). In this ordering the electron momentum scales as meCAe = √ me/mimiCA and so is much smaller than that of the ion fluid. In contrast, the electron pressure is meC 2 Ae = miC 2 A, which is the same as the ion pressure. Thus, the pressure terms on the right side of Eq. (2.4) must be retained even though the electron inertia in this equation is neglected. Finally, the ions follow an adiabatic equation of state d dt ( Pi nγ i ) = 0 (2.10) where the adiabatic index γ is taken to be 5/3. The number density of the electron fluid is given by nef = ni − nep (2.11) while the parallel motion follows from the requirement of vanishing parallel current nefvef,∥ = nivi,∥ − nepvep,∥ (2.12) The zero current condition follows from the constraint on the current from Ampère’s law Vd CA = J neCA ∼ Bc 4πneCAL ∼ di L ≪ 1 (2.13) 37 where Vd is the drift speed associated with the current, di is the ion inertial scale length and L is the characteristic magnetic scale length. In the kglobal model all kinetic scales are ordered out, which yields the zero current condition in Eq. (2.12). The fluid electrons also follow an adiabatic equation of state d dt ( Pef nγ ef ) = 0 (2.14) The magnetic field is advanced via Faraday’s law, ∂B ∂t = −c∇× E⊥ (2.15) with E⊥ = −1 c vi ×B (2.16) Finally, E∥ is produced by magnetic-field-aligned gradients in the electron pressure and is given by [88] E∥ = − 1 e(nep + nef ) (b · ∇ · Tep + b · ∇Tef ) (2.17) Thus, Equations (2.1)-(2.17) constitute the complete set of equations for the kglobal model with particle electrons. 2.1.2 kglobal with particle electrons and ions In the upgraded version of kglobal, particle ions are introduced as a new species and, as is the case for the particle electrons, are treated in the guiding-center limit. Similar to the original model, the populations of the four plasma species do not change after initialization. The per- 38 formance differences arising from various values of the free parameter—the ratio between the number density of particle ions (nip) and the total number density of ions (nif + nip)—will be discussed in Section 3.2. The basic ordering of the electrons and the equations describing their dynamics are basically unchanged from the original model. Minor modifications to the charge neutrality and zero current conditions are discussed below, but the details of the electron dynam- ics that were discussed earlier are not repeated here. Due to the small mass of the electrons, their inertial terms in the fluid ion momentum equation were neglected in the original kglobal equations. That simplification is not possible for the particle ions in the upgraded model, which complicates the derivation and the final form. However, the perpendicular motion of the fluid and particle ions is still governed by the same E×B drift. As a consequence, the generic form of the ion momentum equation, can be written as ∂ ∂t mi(nifvif + nipvip) = S (2.18) with vif and nif the fluid ion velocity and density, vip and nip the particle ion velocity and density, and S the usual momentum driver, including convective inertial terms. The differences in the inertia between the fluid and particle ions arises solely from the parallel motion so the total momentum equation takes the form ∂ ∂t minivif + ∂ ∂t minip(vip∥ − vif∥)b = S (2.19) where ni = nif + nip. The inertial terms parallel to B are easily evaluated, which leaves an equation for the fluid ion momentum that includes the inertia of the particle ions. The details of 39 the evaluation of the parallel inertial terms are found in the Appendix. Most of the other equations in Section 2.1.1 remain unchanged, aside from minor notation adjustments (e.g., ni → nif + nip). Specifically, Equations (2.1) and (2.2) governing the motion of the particle electrons are unchanged. Equations (2.3) and (2.10) become equations for the ion fluid density, ∂nif ∂t = −∇ · (nifvif ) (2.20) and ion fluid pressure, d dt ( Pif nγ if ) = 0 (2.21) The guiding center equations for the particle ions are essentially identical to those of the electrons in Equations (2.1) and (2.2), µip = p2ip,⊥ 2B = const. (2.22) and dpip,∥ dt = pip,∥vE · κ− µib · ∇B + eE∥ (2.23) The ions are taken to be non-relativistic. Including the new species also requires changes in the expressions for the electron fluid density nef = nif + nip − nep (2.24) and parallel electron fluid velocity nefvef,∥ = nifvif,∥ + nipvip,∥ − nepvep,∥ (2.25) 40 The adiabatic equation of state for the electron fluid, Equation (2.14), remains the same, as do equations (2.15), and (2.16) for B and E⊥. Equation (2.17) for E∥ remains the same because only the electrons contribute. In both the original and upgraded model, electron inertia can be discarded compared with that of the ions in the ion momentum equation, as discussed in Sec. 2.1.1. The distinction between the inertia of collective and individual particle electrons, along with their movements in relation to the ambient magnetic field and their specific interactions with the MHD fluid, is comprehensively discussed in [103]. However, as shown in equation (2.18), the inertia of the particle ions cannot be neglected. The derivation of the new fluid ion momentum equation is detailed in Appendix A.1, with the final result being mi ∂ (nivif ) ∂t = −∇·Tif − (∇·Tef )⊥+J×B/c−∇·Tip− (∇·Tep)⊥+eniE∥b−Ii∥ (2.26) where Ii∥ is the difference in the parallel inertia between the fluid and particle ions, given by Ii∥ = ∂ ∂t minip(vip∥ − vif∥)b = −bmivif∥ ( nip nif ∇ · nifvif∥b−∇ · nipvip∥b− nifvi,⊥ · ∇nip nif ) + nip nif bb · ∇ · Tif − bb · ∇ · Tip +minip vip∥ − vif∥ B (I− bb) · ∂B ∂t (2.27) where ni = nip+nif , and the ion fluid stress tensor, Tif , is defined in Equation (2.5). The particle ion pressure tensor is given by Tip = Pip∥bb+ Pip⊥(I− bb) +minipvipvip (2.28) with the parallel and perpendicular pressures calculated in the frame of the bulk motion of the 41 particle ions, Pip⊥ = ∫ dpi p2i⊥ 2mi f (2.29) and Pip∥ = ∫ dpi p2i∥ mi f (2.30) with pip the ion momentum. Note that in the inertia term on the left side of Equation (2.26), ni is the total ion density and not just the fluid ion density nif . In the limit nip → 0, Ii∥ → 0 and Tip → 0 and equation (2.26) reduces to equation (2.4). The ratios nip/nif and nep/nef are typically chosen at t = 0 to be identical throughout the domain, but are generally free to evolve in space and time during simulations. The electron spectra reported in previous kglobal reconnection simulations were insensitive to nep/nef [85], but the sensitivity of reconnection simulations with both particle electrons and ions to these ratios will also need to be explored. To summarize, the kglobal model with particle electrons and ions consists of Eqs. (2.1)- (2.2), (2.6)-(2.9), (2.14)-(2.17), and (2.20)-(2.30). Because of the complexity of Eqs. (2.26) and (2.27), an alternative is to advance the equation for the center-of-mass velocity in Eq. (A.26) and calculate the ion fluid velocity by subtracting the ion particle flux. 2.2 Alfvén Wave Propagation and the Growth of Firehose Modes in an Anisotropic Plasma Particle energy gain in reconnection is dominantly parallel to the ambient magnetic field, with the implication that the parallel pressure exceeds the perpendicular pressure in the heated 42 plasma as reconnection develops. As a consequence, the cores of magnetic islands approach the marginal firehose condition at late time in simulations. The tension in a bent magnetic field line, which is the fundamental driver of reconnection, goes to zero at the marginal firehose condition. Thus, the pressure anisotropy that develops as reconnection proceeds regulates reconnection and particle energy gain [53, 106, 103, 88]. Therefore, to ensure that the updated kglobal model properly reproduces the influence of pressure anisotropy on the magnetic field dynamics, we benchmark the model with a circularly polarized Alfvén wave mode propagating along a uniform magnetic field in a system with pressure anisotropy (P∥ ̸= P⊥), a problem with a well-established solution [31]. With increased pressure anisotropy, the reduction in magnetic tension of a propa- gating Alfvén wave reduces the wave phase speed. In a second benchmarking test on the dynam- ics of the pressure anisotropy of both particle ions and electrons, we explore the growth of the firehose instability. The upgraded kglobal model builds upon the original kglobal framework, the details of which can be found in the references [103, 88]. Implementation of the model uses a code written in normalized units. A reference magnetic field strength B0 and the initial total ion density ni0 = nip + nif define the Alfvén speed, CA = √ B2 0/(4πmini0). Lengths and times are normalized to an arbitrary macroscale length, L, and the Alfvén crossing time, τA = L/CA, respectively. Electric fields and temperatures are normalized to CAB0/c and miC 2 A, respectively. The kglobal model employs second-order diffusivities and fourth-order hyperviscosities (∝ ν∇4) in each fluid equation to mitigate noise at the grid scale [103]. The simulations were performed within a two-dimensional (2D) spatial domain, coupled with a three-dimensional velocity space, featuring an initially uniform magnetic field of magni- tude B0 oriented along the x-axis. The domain measures 2πL × πL with 512 × 256 cells with 43 periodic boundary conditions on all sides. Each grid cell is initialized with 200 self-consistent (i.e., not test) particles, 100 ions and 100 electrons. The particle-to-fluid density ratios are ini- tially set to nip/nif = nep/nef = 1/3 everywhere, but evolve in time at each location. The ion-to-electron mass ratio is set to 25, although the results of the simulations are insensitive to this value. The scalar temperature of both fluid species is set to 0.1. For the particle ions and electrons, the perpendicular temperature is initially set to 0.1 while the parallel temperature is adjusted to control the magnitude of the anisotropy. We introduce perturbations to the magnetic field and velocity in the form of a circularly polarized Alfvén wave, with a wavelength equal to the length of the simulation box. During a propagation interval of 3τA, we measured the wave’s velocity in order to compare it with the theoretical phase speed Cp of an Alfvén wave, which is given by Cp = CA √ 1− 4π(P∥ − P⊥) B2 ≡ αCA (2.31) where α = √ 1− 4π(P∥ − P⊥)/B2, aligns with the Chew-Goldberger-Low (CGL) framework in the linear regime where pressure perturbations are negligible [107]. Here, P∥ and P⊥ denote the combined ion and electron pressures along and perpendicular to the magnetic field, respectively. 44 The results are shown in Fig 2.1, where Cp, normalized to CA is plotted against α. There is excellent agreement between the code results and linear wave theory, highlighting the precision of the kglobal model and suggesting that the impact of ion and electron pressure anisotropy on reconnection dynamics will be properly included in the new model. 45 0.0 0.2 0.4 0.6 0.8 1.0 = 1 4 (P P )/B2 0.0 0.2 0.4 0.6 0.8 1.0 C p /C A Regression Line: y = 0.98x + 0.02 Figure 2.1: The observed phase speed of the Alfvén wave, Cp, plotted against the anisotropy parameter α. The blue dots are the results of simulations. The intersection of the vertical and horizontal dotted lines marks the isotropic Alfvén wave. 46 In the second benchmark, the rate of growth of the firehose instability was explored with a pre-set pressure anisotropy above the instability threshold (α2 = −0.2). The system was ini- tialized with small amplitude periodic disturbances with seventeen distinct wavelengths, defined by k = m/2πL, where m is the mode number. The simulations incorporated a kinematic vis- cosity of ν = 5.0 × 10−5. The growth rate, which includes the pressure anisotropy drive and viscous damping is given by γ = kCA|α| − νk4. The fourth-order hyperviscosity sets the up- per bound of the instability at small spatial scales. Figure 2.2 presents the expected theoretical growth rate (represented by a solid blue line) and the growth rates found in the simulations for a range of unstable modes. The red crosses indicate systems in which the electrons and ions contribute equally to the total anisotropy. To explore the effect of unequal pressure anisotropies, we also performed several simulations where the ions contribute 3/4 and the electrons contribute 1/4 of the total anisotropy (green squares). The growth rates measured in the unequal anisotropy simulations closely follow those from equal anisotropy simulations and the results from all the simulations are consistent with the predictions of linear theory. Note that there are no spurious unstable modes observed at short scales such as those that arise in some numerical models when the mode frequency exceeds the cyclotron frequency [108]. All waves in the kglobal model are below the cyclotron frequency. 47 2 4 6 8 10 12 14 16 m 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 A Theoretical Growth Rate Line Numerical Growth Rates with Equal Electron-Ion Pressure Anisotropy Numerical Growth Rates with Unequal Electron-Ion Pressure Anisotropy Figure 2.2: Normalized growth rate of the firehose mode, γτA, versus mode number, m = 2πkL, over a span of unstable m values. Red crosses show values from simulations with equal electron and ion pressure anisotropies; values for unequal anisotropies are shown as green squares. The blue line is the theoretical growth rate. 48 Chapter 3: Simultaneous Proton and Electron Energization during Macroscale Magnetic Reconnection The material in this chapter has been adapted from [109] with permission from the authors. 3.1 Simulation Model Setup In this chapter we present results from simulations of magentic reconnection in a macro- scopic domain (i.e., one not constrained by kinetic scales). The simulations include four distinct plasma species: fluid protons and fluid electrons (which collectively form the MHD backbone), and particle protons and electrons which are are represented by macro-particles that move through the MHD grid. The particles are treated in the guiding-center limit, thus eliminating the need to resolve their respective Larmor radii. Although the computational domain does not allow spatial variations in one direction, such a configuration is known to support reconnection and has been used in many previous simulations. Particles move across the magnetic field with their E × B drift and along the local magnetic field at their parallel velocity. As with the previous version of kglobal [85], the reconnecting component of the upstream magnetic field B0 (along the x direction) and the total proton density (the combined number den- sity of particle and fluid protons) ni0 serve as normalization parameters by defining the Alfvén speed CA0 = B0/ √ 4πmini0. Because kinetic scales are excluded in the model, lengths are nor- 49 malized to an arbitrary macroscale L0 and time scales are normalized to τA = L0/CA0. Both tem- peratures and particle energies are normalized to miC 2 A0. The perpendicular electric field in the simulations follows the MHD scaling CA0B0/c, while the parallel field scales as miC 2 A0/(eL0). Although the parallel electric field is small compared to the perpendicular component, the energy associated with the parallel potential drop over the scale L0 is of the order of miC 2 A0, making it comparable to the available magnetic energy per particle. The proton-to-electron mass ratio is set to 25, but our results are insensitive to this pa- rameter. The speed of light in our simulations is 30 times the Alfvén speed. All species (both fluid and particle) begin with an isotropic temperature of 0.0625miC 2 A0, with the particles hav- ing a Maxwellian velocity distribution. The simulations are initialized with constant densities and pressures in a force-free current sheet with periodic boundary conditions. Thus, B = B0 tanh(y/w)x̂ + √ B2 0 sech 2(y/w) +B2 g ẑ where Bg is the asymptotic out-of-plane magnetic field (the guide field) and w, the width of the current sheet, is set to 0.005L0. The initial total density of electrons (ne) and protons (ni) is normalized to unity, with particles comprising 25% of the density (nep for particle electrons and nip for particle protons) and the remaining 75% in the fluid component (nef for fluid electrons and nif for fluid protons). We demonstrate that results of the simulations are insensitive to this fraction. The simulations are conducted on grids with varying resolutions, ranging from 1024× 512 to 8192× 4096 grid points as shown in Table 3.1. The number of particles is set to 100 particles per grid. The time step δt is adjusted as the grid size is varied to ensure that particle stepping is accurate (e.g., cδt ≃ δ with δ the grid scale). The values of the time-step are given in Table 3.1. 50 In our simulations, diffusion and hyperviscosity terms are included to ensure numerical stability while reducing any high-frequency noise arising at the grid scale. The diffusion coeffi- cients, denoted as Dn for number density diffusion and Dp for pressure diffusion, are set to the values listed in Table 3.1. A hyperviscosity ν is included in the magnetic field evolution equation to facilitate reconnection while minimizing dissipation at large scales. It is applied as a fourth- order Laplacian term (∇4). The same hyperviscosity is included in the evolution equations for the fluid proton flux, fluid proton number density and fluid proton pressure. The effective Lundquist number Sν = CAL 3 0/ν associated with the hyperviscosity is varied to change the effective system size (the ratio of the macro to the dissipation scale). The hyperviscosity coefficients, denoted as νB, νnv, νn, and νp, as well as Sν , are set to the values given in Table 3.1. 3.2 Simulation Results In our simulations the hyperviscosity triggers reconnection, leading to the formation of small-scale flux ropes. These flux ropes coalesce over time and eventually form system-scale structures. Figure 3.1 shows the results of a simulation with Bg/B0 = 0.25. Shown is the density of particle protons in the x–y plane at five times, t/τA = 0, 4, 7, 10, and 13, in panels (a)–(e). 51 Table 3.1: Parameters for Simulation Domains Grid Points 1024× 512 2048× 1024 4096× 2048 8192× 4096 Time Step (in τA) 2× 10−4 1× 10−4 5× 10−5 2.5× 10−5 Proton Number Density Diffusion (Dn) 10.5× 10−5 5.25× 10−5 2.625× 10−5 1.3125× 10−5 Proton Pressure Diffusion (Dp) 10.5× 10−4 5.25× 10−4 2.625× 10−4 1.3125× 10−4 Hyperviscosity (νB, νnv, νn, and νp) 84× 10−9 10.5× 10−9 1.312× 10−9 0.1640× 10−9 Effective Lundquist Number (Sν) 1.2× 107 9.5× 107 7.6× 108 6.1× 109 52 At t = 0 the magnetic field reverses across a uniform current sheet as shown in panel (a) of Figure 3.1. As the system evolves, reconnection begins at multiple sites. The resulting reconnected magnetic field lines convect away from the x-points. The contraction of magnetic islands along the current sheet leads to the energization of particle electrons and protons trapped within them. As the simulation evolves, these islands grow and undergo coalescence, culminating in the formation of a single, large magnetic island late in time. As can be seen in Figure 3.1 the particle proton density remains relatively uniform along the field lines due to the protons’ mobility parallel to the magnetic field. 53 0.5 0 -0.5 (a) 0.5 0 -0.5 (b) 0.5 0 -0.5 (c) 0.5 0 -0.5 (d) -3 -2 -1 0 1 2 3 x/L0 0.5 0 -0.5 (e) 0.2 0.4 0.6 0.8 1.0 pa rti cl e pr ot on d en si ty y/L0 Figure 3.1: The temporal evolution of the particle proton density in the x–y plane for Bg/B0 = 0.25. Five times are displayed: t/τA = 0, 4, 7, 10, and 13 in panels (a) to (e). In each panel, magnetic field lines are shown in black. 54 Figure 3.2 shows the energy per particle (energy density divided by number density) of the particle protons and particle electrons, ⟨W ⟩, and the firehose parameter, α = 1−4π(P∥−P⊥)/B 2, at late time, t/τA = 13, from the same simulation. In the top two panels darker colors represent higher average energies per particle. Thus, the average energy of the protons is significantly greater than that of the electrons. The firehose parameter in the bottom panel of Figure 3.2 is a measure of the impact of pressure anisotropy on the local magnetic tension, given by αB · ∇B/4π. Specifically, when α approaches zero the magnetic tension, which is the driver of magnetic reconnection also goes to zero. Thus, the dominant feedback mechanism of energetic particles on reconnection dynamics is through the reduction of magnetic tension as measured by the firehose parameter. In the context of the simulation from kglobal, both P∥ and P⊥ are derived from the particle species since both the fluid proton and electrons are taken to be isotropic and do not impact the magnetic tension force. Thus, in Figure 3.2 extensive areas within the magnetic islands approach marginal stability (α = 0), punctuated by small unstable regions (red shades). Consequently, the local magnetic tension, a crucial factor in driving reconnection and particle energy gain, is strongly reduced within the large flux rope and somewhat reduced in the smaller flux ropes. This highlights the critical importance of particle feedback on the MHD fluid dynamics. Traditional MHD models that include test-particle dynamics do not include this feedback and thereby risk an uncontrolled increase of the particle energy. 55 0.5 0 -0.5 t = 13 A 0.5 0 -0.5 t = 13 A -3 -2 -1 0 1 2 3 x/L0 0.5 0 -0.5 t = 13 A 0.0 0.2 0.4 0.6 0.8 1.0 < W > /m iC 2 A0 1.0 0.5 0.0 0.5 1.0 y/L0 Figure 3.2: The energy per particle of the particle protons (top) and particle electrons (middle), ⟨W ⟩, is shown in the top two panels, respectively, for the simulation with Bg/B0 = 0.25. The darker the color, the higher the average energy of the particles so the protons contain significantly more energy than the electrons. The firehose parameter α at t = 13τA is shown in the bottom panel. The color white corresponds to marginal stability and reveals that within flux ropes where particles have been energized, the particle feedback has significant impact on the MHD environ- ment. 56 At late time, after a single large magnetic island dominates the current sheet, we calculate the proton energy spectra by aggregating particle counts throughout the entire simulation domain. Figure 3.3(a) displays the differential number densities F (W ) = dN(W )/dW plotted against normalized energy W/miC 2 A0 at various times for Bg/B0 = 0.25. F (W ) begins as a Maxwellian but develops a power-law tail, characterized by an index δ′, that first appears for low energies before eventually extending to near the maximum energy observed in the domain. The powerlaw index at late time approaches 3.4. Similar behavior was observed for electrons in the earlier electron-only version of kglobal [85]. Note that the particle distribution at low energy in Figure 3.3(a) does not change significantly in time. This is not because the protons are not heated during reconnection but because the distribution in the figure is from the entire computational domain, which includes large regions of plasma upstream of the reconnecting current layer that has not been accelerated. The heating of the plasma will be more evident when we focus on the particle distributions within the large islands that develop at late time in the simulations. In Figure 3.3(b) we show the late-time F (W ) spectra for protons for different Bg values, measured at times when approximately equal amounts of magnetic flux have undergone recon- nection. A similar figure from the electron-only version of kglobal was presented in [85]. In both cases as Bg decreases, the power-law index δ′ also decreases. This results in a harder energy spectrum and leads to the generation of a larger number of high-energy particles for small values of Bg. 57 10 2 10 1 10 0 10 1 10 2 10 3 W/miC 2 A0 10 10 10 8 10 6 10 4 10 2 10 0 10 2 F( W ) ( A rb . u ni ts ) ~W 3.4 Proton (a) t/ A = 0 t/ A = 1 t/ A = 2 t/ A = 3 t/ A = 4 t/ A = 5 t/ A = 6 t/ A = 7 t/ A = 8 t/ A = 9 t/ A = 10 t/ A = 11 t/ A = 12 t/ A = 13 10 2 10 1 10 0 10 1 10 2 10 3 W/miC 2 A0 10 10 10 8 10 6 10 4 10 2 10 0 10 2 F( W ) ( A rb . u ni ts ) Proton (b) Bg/B0 = 0.25 Bg/B0 = 0.5 Bg/B0 = 0.75 Figure 3.3: A log-log plot of the differential number density versus the normalized energy. Panel a shows the temporal evolution of the proton distribution in a simulation with Bg/B0 = 0.25. The initial Maxwellian spectrum develops an energetic tail that hardens over time and eventually stabilizes into a power law. Panel b shows the late time spectra for different guide fields. Smaller guide fields result in harder spectra that extend to higher energies. 58 In addition, we explore the late-time distributions of energetic protons for a guide field Bg/B0 = 0.25 and various values of Sν (effective system size). The results in Fig. 3.4 show that the power-law indices of the spectra are only weakly dependent on the system size for large domains. As the size of the system increases, the spectra become slightly harder and the upper limit in the particle energy gain increases. This is because in larger domains more flux ropes form in the reconnecting current sheet and it is the merger of those flux ropes that efficiently drives particle energy gain. 59 10−2 10−1 100 101 102 103 104 W/miC2 A0 10−12 10−10 10−8 10−6 10−4 10−2 100 102 F( W ) ( A rb . u ni ts ) Proton Sν=1.2×107 Sν=9.5×107 Sν=7.6×108 Sν=6.1×109 Figure 3.4: The late-time distributions of energetic protons in a simulation with Bg/B0 = 0.25 for various values of Sν (effective system size) reveal that the power-law indices of the spectra vary weakly with the domain size. The rightmost point of each curve marks the one-count level, indicating the flux when there is just one particle count per energy bin. The notable difference is that as the size of the system increases, the maximum energy of the protons also increases. 60 A free parameter in the kglobal model is the fraction of the particle electrons and protons compared with their respective fluid component [103, 88, 105]. An important question is whether reconnection simulations are sensitive to these ratios. To address this question, we investigated the spectra of protons and electrons with a guide field Bg/B0 = 0.25 and with different fractions of the particle densities. Figure 3.5 shows the late-time energy spectra of particle protons and electrons from three simulations with differing particle fractions. The power-law indices of the spectra are nearly identical, indicating that the simulation results are insensitive to this ratio. The only notable difference is that as the fraction of particles decreases, the maximum achieved energy of the particles increases modestly. This can be attributed to the fixed total energy in the simulation box. The particles gain more energy than their fluid counterparts with the parallel energy gain exceeding that of the perpendicular energy. Thus, when there are fewer particles in the simulation box, the particles can gain more energy before the firehose parameter acts to limit reconnection and energy gain. 61 10−2 10−1 100 101 102 103 W/miC2 A0 10−8 10−6 10−4 10−2 100 102 F( W ) ( A rb . u ni ts ) nip/ni = 40% nip/ni = 25% nip/ni = 10% 10−2 10−1 100 101 102 103 W/miC2 A0 10−8 10−6 10−4 10−2 100 102 F( W ) ( A rb . u ni ts ) nep/ne = 40% nep/ne = 25% nep/ne = 10% Figure 3.5: The simultaneous distributions of protons (top) and electrons (bottom) from simula- tions with a guide field Bg/B0 = 0.25 and different fractions of particle densities. The power-law indices of the spectra for the different ratios are nearly the same, indicating that the simulation is insensitive to this ratio. The only notable difference is that as the fraction of particles decreases, the maximum achieved energy of each species increases modestly. 62 We show in Figure 3.6 the initial and the late-time spectra of particle protons and electrons for the case Bg/B0 = 0.25. The initial curves for both protons and electrons are Maxwellian distributions. Late-time spectra exhibit power-law tails for both particle types. The spectral indices of the powerlaws of the two species are nearly the same. Although the length of the power-law tail for the protons is shorter than that of the electrons, it extends nearly a full decade higher in energy. A key feature is the significantly greater energy gain of the protons compared to the electrons, which is consistent with the increased energy content of protons compared to electrons shown in Figure 3.2. 63 10 2 10 1 10 0 10 1 10 2 10 3 W/miC2 A0 10 11 10 9 10 7 10 5 10 3 10 1 10 1 F( W ) ( A rb . u ni ts ) Electron Proton Figure 3.6: The simultaneous distributions of protons (blue curves) and electrons (orange curves) for a guide field Bg/B0 = 0.25. The curves on the left indicate the initial Maxwellian distribu- tions for both particle types. The hard spectra are the late time results, showing extended power law distributions for both protons and electrons with powerlaw indices that are comparable. 64 We next explore the details of particle heating and energization by examining the rela- tive numbers and energy content of the non-thermal and hot thermal particles. The spectra are modeled with the sum of a kappa [110] and a Maxwellian distribution (See Appendix B.1 for a detailed explanation of kappa distribution. The kappa distribution turns into a powerlaw at high energy which can represent the extended powerlaw tails seen in the simulation data [8]. This combination is particularly useful in modeling the distributions of particles in space and astrophysical plasmas where non-thermal behaviors are observed [111, 7, 8]. In Figure 3.7 we show F (W ) (black) for protons and electrons at late time after the dis- tribution functions have reached a steady state. The data is taken from within the large late-time island shown in Figure 3.1(e). Thus, the spectra shown include only particles that have undergone heating and energization during reconnection. Shown in Figure 3.8 are the intial and final elec- tron and proton distributions from the same large island but using a linear scale. This figure more clearly shows the changes in the distribution of low energy particles as they are heated during reconnection. The spectra in Figure 3.7 are fit (solid red) by the sum of a Maxwellian distribution and a kappa distribution [85]. This procedure is designed to capture the characteristics of both hot thermal and nonthermal particle populations. From this fitting approach, we extract several parameters: 1. The spectral index δ′ of the powerlaw of the non-thermal particles. 2. The relative densities and total energy of the hot thermal and non-thermal electrons and protons. 3. The break point energy (vertical line in Figure 3.7) , expressed in units of miC 2 A0, that marks the transition from thermal to non-thermal particle dominance in the spectra. At the 65 10 2 10 1 100 101 102 103 W/miC2 A0 10 10 10 8 10 6 10 4 10 2 100 102 F( W ) ( Ar b. u ni te s) ~W 3.4 x = 1.34 Proton Fitting Kappa Thermal Kappa Nonthermal Maxwellian 10 2 10 1 100 101 102 103 W/miC2 A0 10 10 10 8 10 6 10 4 10 2 100 102 F( W ) ( Ar b. u ni ts ) ~W 3.3 x = 0.28 Electron Fitting Kappa Thermal Kappa Nonthermal Maxwellian Figure 3.7: Fitting and decomposition of the late time spectra of electrons and protons. The solid black curves are the proton (top) and electron (bottom) distribution functions for a guide field of 0.25. The solid red curves show the fit to a sum of a Maxwellian and a kappa distribution. The green dashed curve is the thermal part of the kappa distribution, while the blue dashed curve is the non-thermal part of the kappa distribution. The yellow curves are the Maxwellian distributions of the two species. The pink vertical line indicates the energy where the nonthermal and thermal components of the kappa distribution are equal. The solid blue line is the asymptote for the distribution function in the power law region. 66 break point the number densities of the hot thermal and non-thermal components of the kappa distribution are equal (See Appendix B.1 for a detailed explanation of the procedure used to separate the hot thermal and non-thermal components of the kappa distribution). To the left of the break point, the thermal component dominates, while to the right, the thermal components decrease sharply, and the non-thermal component dominates. 67 Table 3.2: Density Distributions for Particle Acceleration at Late Time Type Guide Field Maxwellian Kappa Total Kappa Maxwellian Kappa Nonthermal (in B0) (%) (%) (%) (%) Electron 0.25 38.2 61.8 45.1 16.7 Electron 0.50 68.0 32.0 26.8 5.2 Electron 0.75 84.3 13.7 12.4 1.3 Proton 0.25 51.7 48.3 35.8 12.5 Proton 0.50 43.1 56.9 46.7 10.2 Proton 0.75 54.5 45.5 40.3 5.2 68 0.0 0.2 0.4 0.6 0.8 1.0 W/miC2 A0 0 2 4 6 8 10 F( W ) ( A rb . u ni ts ) Initial Electron Initial Proton Late-Time Electron Late-Time Proton Figure 3.8: The comparison between the initial spectra and late-time spectra of protons and electrons is presented on a linear-linear scale to more clearly reveal the changes in the particle distributions at low energy. 69 Table 3.2 documents how the strength of the guide field affects the proportion of particles that comprise the Maxwellian and kappa distributions, with the latter further subdivided into thermal and non-thermal components. As the guide field increases, the proportion of non-thermal particles, indicated by the kappa non-thermal percentage, consistently decreases. For electrons, the non-thermal component reduces significantly from a higher percentage at a lower guide field to a much smaller fraction at the higher guide field. A similar trend is observed for protons, where the non-thermal proportion diminishes as the guide field strength increases. This trend establishes that a strong guide fields suppresses the generation of non-thermal electrons and protons, which is consistent with the findings for energetic electrons in [85]. Table 3.3 shows the relative energy of thermal and non-thermal electrons and protons at late time, again as a function of guide field strength. In this data the total number density of particles is normalized to unity and the numbers in the table reflect the energy content of the various components in units of miC 2 A. As the guide field increases, the energy in the non-thermal component of the kappa distribution decreases for both electrons and protons. For electrons, the kappa non-thermal energy decreases from 0.095 at a guide field of 0.25 to 0.006 at a guide field of 0.75. For protons, the kappa non-thermal energy decreases from 0.340 at a guide field of 0.25 to 0.045 at a guide field of 0.75. The downstream thermal temperatures are also shown. For electrons the thermal temperature, represented as a weighted sum of the Maxwellian and kappa-Maxwellian distributions, remains comparable for different guide fields. The temperature increase is modest compared with the initial value of 0.0625miC 2 A0. The heating is driven by the macroscale parallel electric field as explored by the reference [85]. In contrast, protons exhibit a much larger increase in temperature as has been documented in the Earth space environment [112, 113, 114]. For the lowest guide field in Table 3.3 the ratio of the temperature increment of 70 the protons to electrons is around five, which is slightly smaller than in measurements in the Earth space environment [113]. Also noticeable is the decrease in thermal temperature of protons with increasing guide field. We emphasize that because the data in Table 3.3 is normalized to miC 2 A0, the temperature increments of both species scales with this parameter as in the observations [115, 112, 113, 114]. Also shown in Table 3.3 is the breakpoint energy, which marks the transition from t