UNIVERSIDAD NACIONAL AUTÓNOMA DE MEXICO PROGRAMA DE POSGRADO EN ASTROFÍSICA INSTITUTO DE ASTRONOMÍA ASTROFÍSICA TEÓRICA EVOLUCIÓN TÉRMICA DE ESTRELLAS DE NEUTRONES EN SISTEMAS BINARIOS DE BAJA MASA CON ALTAS TASAS DE ACRECIÓN TESIS QUE PARA OPTAR POR EL GRADO DE: DOCTOR EN CIENCIAS (ASTROFÍSICA) PRESENTA: MARTIN JAVIER NAVA CALLEJAS TUTOR DR. DANY PIERRE PAGE ROLLINET INSTITUTO DE ASTRONOMÍA - UNAM CIUDAD UNIVERSITARIA, CDMX, DICIEMBRE 2024. UNAM – Dirección General de Bibliotecas Tesis Digitales Restricciones de uso DERECHOS RESERVADOS © PROHIBIDA SU REPRODUCCIÓN TOTAL O PARCIAL Todo el material contenido en esta tesis esta protegido por la Ley Federal del Derecho de Autor (LFDA) de los Estados Unidos Mexicanos (México). El uso de imágenes, fragmentos de videos, y demás material que sea objeto de protección de los derechos de autor, será exclusivamente para fines educativos e informativos y deberá citar la fuente donde la obtuvo mencionando el autor o autores. Cualquier uso distinto como el lucro, reproducción, edición o modificación, será perseguido y sancionado por el respectivo titular de los Derechos de Autor. To my family - mom, dad and sister - for their unconditional support and love; to Prof. Dany, for his support and mentorship along the years; to Macarena, for her encouragement and friendship (ten years and beyond!); to Tais, for her support - always throughout music, lately throughout birbs, cheep; to Anna, for her kindness and for reminding me the importance of subjectivity. 1 Acknowledgements Given this is (probably) my last written thesis and thus the conclusion of my aca- demic path as a learner I wish to thank to all my key mentors along this road, both academic and personal aspects of my life. Everaldo Tapya in primary school. At CECYT 11 - “la Willy” - Prof. Alma Benítez, for introducing me to the world of research and for believing in me to reach this far. Prof. Ismael Sandoval (RIP), who taught me how wonderful physics is. At University: Prof. Mario Pacheco, my men- tor along my 8 semesters at ESFM, my alma mater ; Prof. Isaura Fuentes-Carrera, who introduced me into the marvelous world of astrophysics; Prof. Arturo Zúñiga Segundo, who motivated me to go beyond. To Profs. Ana Hidalgo-Gámez, Fernando Angulo Brown, Jaime Avendaño, José Castro Quilantán and Alfredo López Ortega, for guiding me along the different branches of physics, and to Profs. Elena Luna Elizarrarás and Luis Cisneros Ake for making me appreciate the beauty of formal mathematics. My time at the IA-CU has been wonderful. Such experience wouldn’t have been possible without the continuous support from my advisor, Prof. Dany Page, who mentored me throughout all these years of Master and PhD, believing in what I could do and always motivating me to keep exploring new ideas. My PhD project wouldn’t have been the same without the collaboration of Yuri Cavecchi, who pro- vided support and guidance as well. Thanks to Aida Kirichenko, who was always supportive and provided me valuable advices along my path as a PhD student. I wish to express my gratitude towards all the staff for welcoming me as a part of the Institution. Special thanks to Bertha Vázquez and Heike Breunig for their ample support during my time as a student, as well as to Brenda Arias. A special thank as well to the Academic Committee of the Program for allowing me to represent the PhD students’ for two years. A special grazie mille to the ECT and Residenza Arcivescoville staffs and the DTP/TALENT 2024 people I met during my stay in Trento, Italy. Forever I will treasure the memories of my first international travel, the extraordinary three weeks and the kind people I met, specially Nithish, Tiago, Milena and Rajesh, with whom I spent more time hanging around the city. Special thanks to Jorge Piekarewicz and Barbara Gazzoli for their support and attention during those days and for the organization of the event. 2 Contents 1 Equations of Stellar Evolution 14 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 1.2 Spacetime evolution of physical quantities . . . . . . . . . . . . . . . 15 1.3 Evolution in a static and spherically symmetric (SSS) spacetime - fluid in radial motion . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 1.4 On the numerical implementation of neutron star evolution . . . . . . 33 1.5 Thermonuclear Instabilities . . . . . . . . . . . . . . . . . . . . . . . 39 2 Microphysical aspects of stellar evolution 43 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 2.2 A brief survey on Thermodynamics . . . . . . . . . . . . . . . . . . . 44 2.3 The Neutron Star Equation of State . . . . . . . . . . . . . . . . . . . 50 2.3.1 The envelope, ρ ≤ 108 g cm−3 . . . . . . . . . . . . . . . . . . 54 2.3.2 Outer and inner crusts, ρ ∈ [108, 1014] g cm−3 . . . . . . . . . 62 2.4 Thermal conductivity and opacity . . . . . . . . . . . . . . . . . . . . 66 2.5 Particle reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 2.5.1 β reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 2.5.2 Chains and cycles . . . . . . . . . . . . . . . . . . . . . . . . . 78 2.6 Thermal neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 3 Nuclear networks for neutron stars 86 3.1 Overview of numerical methods . . . . . . . . . . . . . . . . . . . . . 87 3.2 On the networks for neutron stars. . . . . . . . . . . . . . . . . . . . 90 3.3 An overview of the rp-process . . . . . . . . . . . . . . . . . . . . . . 94 4 Neutron Star Envelopes 104 4.1 General features of the envelope . . . . . . . . . . . . . . . . . . . . . 105 4.2 Evolution of non-accreting envelopes . . . . . . . . . . . . . . . . . . 109 4.3 Evolution of accreting envelopes . . . . . . . . . . . . . . . . . . . . . 114 4.3.1 Stationary accreted envelopes . . . . . . . . . . . . . . . . . . 115 4.3.2 Intermezzo: Linear perturbations to stationary solutions and the origin of bursts . . . . . . . . . . . . . . . . . . . . . . . . 131 3 4.3.3 Time-dependent envelopes . . . . . . . . . . . . . . . . . . . . 135 5 Conclusions 150 6 Paper 1: “The Effect of Opacity on Neutron Star Type I X-ray Bursts Quenching” 153 6.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153 7 Paper 2 - Considered for acceptation 156 7.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156 8 Paper 3 (Third Author) 158 8.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158 A Thermodynamics 159 A.1 Propositions and theorems . . . . . . . . . . . . . . . . . . . . . . . . 159 A.2 Maxwell relations and alternative expressions for some partial deriva- tives of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 A.3 Heat capacity in superfluid systems . . . . . . . . . . . . . . . . . . . 163 B General Relativity 165 B.1 On the conservation laws . . . . . . . . . . . . . . . . . . . . . . . . . 165 B.2 SSS metric - Identities . . . . . . . . . . . . . . . . . . . . . . . . . . 167 B.3 Propositions and theorems . . . . . . . . . . . . . . . . . . . . . . . . 169 B.4 Radiative Zero Solution . . . . . . . . . . . . . . . . . . . . . . . . . 175 4 Abstract The purpose of the present work is to study the envelope of neutron stars in Low- Mass X-ray Binaries (LMXB) in the presence of high mass accretion rates, i.e. ≥ 0.3ṀEddington, in order to identify the necessary conditions to treat accreted envelopes as time-independent and thus simplify the numerical evolution of temperature and luminosity of the whole neutron star. Considering the current understanding of nuclear burning at such high accretion rates, for such study we required to construct a network of nuclear reactions, sufficiently large as to simulate in detail the rapid proton capture process (rp-process). In addition, the present study aimed to restrict the conditions under which we could regard the leakage of thermonuclear energy due to H and He burning, from the envelope to the core of the neutron star, as the source of the still-unknown shallow heating. In Chapter 1 I introduce the equations governing the structure and tempera- ture of a neutron star undergoing mass accretion. The functionals describing the microphysical aspects of the evolution, such as the Equation of State (EOS) or ther- monuclear rates, are detailed in Chapter 2. In Chapter 3 we present the numerical challenges of solving the equations of evolution for the chemical composition at the surface of the stars, as well as further details regarding the rp-process. A detailed study of the envelope, in both time-independent and -dependent scenarios is pre- sented in Chapter 4, where we show the results from the simulations computed with the stellar evolution code MESA and my own newly developed code for stationary envelopes. In Chapter 5 I summarize the results from Chapter 4 and prospects for future applications. In Chapter 6 I give further details on the research article (manuscript) presenting the newly-developed code, while in Chapter 7 I present a second research article related to modifications in opacity as a possible explanation for the burst quenching at 0.3ṀEdd. In Chapter 8, a third-author paper is intro- duced, in which I collaborated computing the evolution under electron captures of oxygen and neon bubbles. 5 Preface Neutron stars are among the most fascinating objects in the Universe. Quoting Page, D.1, “You can ask for whatever physics you want and neutron stars have it. General relativity? Solid state physics? You can have them all!” The current picture of how the interior of a neutron star looks like is illustrated in Fig. 1. The densest portion of the star, the core, is composed of neutrons, pro- tons, electrons and muons in charge and beta equilibrium, and the baryonic sector conforming a superfluid/superconductive state. While theoretical estimates point towards the existence of exotic phases at the most inner portion (marked as (?)), observational evidence is still in favor of the enlisted baryonic and leptonic compo- sition ([AAA+17, RGH+21]), and we thus stick to this picture in the present work. Between 1011 and 1014 g cm−3, in the region known as inner crust, neutrons can be either free, conforming a superfluid, or bounded with protons in nucleus of exotic shapes, going from pasta to spherical shape at low densities. At ρ ≤ 1011 we observe a transition to “standard” matter of different aggregation state: between 109 and 1011 g cm−3 nucleus are arranged in lattices imbued in a gas of free degenerate elec- trons, while between 108 and 109 g cm−3 nucleus are best described by the transition from a gas to a liquid state. While the region between 108 and 1011 g cm−3 is usually dubbed as outer crust, due to the composition and state of aggregation for some authors the portion between 108 and 1010 g cm−3 still belongs to the envelope, the region which in our Fig. 1 corresponds to the outermost and thin layer of the star. While in the present work we employ ρb to denote the density boundary between the envelope and the outer crust, to avoid confusions we always state the explicit value of this boundary for each of the employed models. Despite such versatility, placing “nuclear reactions” and “neutron stars” in the same sentence seems odd at first glance: after all, it is a well stablished fact [KKO+03] neutron stars are the last stage in the life of very massive stars, had they not collapsed into black holes. Being the last stage means nuclear reactions no longer provide sufficient heat to counterbalance the energy loss due to neutrino and photon emission. Furthermore, they do not take place anymore due to the inexistent fuel to start them. Surprisingly, however, we can place “nuclear reactions” and “neutron stars” in 1Private communication 6 npeµ (?) FR EE NE UT RO NS LIQ UID ! C RY STA L Sp he ric al Ro ds Pa st a De g. ele ct ro ns Ions in latti ce Ion s + e− + Photons ρ [g cm−3] 1014 1011 108 T sensitivity # $ τHeat difussion ≪ ≫ 99.9 % M ≪ 1 % M Figure 1: The interior of a typical neutron star the same sentence, without restricting ourselves to the subject of stellar evolution. In the vastness of the Universe, there exists systems where nuclear reactions do take place at the surface of neutron stars. These are the interacting binaries, self- gravitating system where the neutron star receives mass from a main sequence or more evolved companion star [BD23]. The transference mechanisms depends on the mass of the young star, Mcomp. If ∼ M⊙, i.e. a lower mass than the neutron star’s (Mneutron star ≈ 1.4M⊙), gravitational pull generates a disk around the neutron star and eventually, via thermal-viscous instabilities (e.g. the disk-instability model [van96, Ham20]), matter is uniformly accreted towards the surface of the compact object. On the other hand, in the presence of a massive companions the transfer of material towards the neutron star occurs via the stellar wind or equatorial disk of the companion [Cha22a]. In the present work, only the former scenario is addressed. Observationally, binary systems composed of a neutron star and a low-mass, main sequence companion are dubbed as Low-Mass X-ray Binary (LMXB) since they were discovered employing X-ray telescopes due to their prominent emission in said band due to the release of gravitational energy of the in-falling material. Such was the case of Scorpio X-I, first spotted with an X-ray luminosity of LX ∼ 104L⊙ ≈ 1037 erg s−1 [GGPR62], later identified as a LMXB [SOG+66, Shk67, Cha22b]. Following this event, many neutron stars in LMXBs have also been identified due to the observation of very energetic explosions in the X-ray band (Lx ∼ 1038 erg s−1 [Kdi+03, Kuu04]), hence named as X-ray bursts. Since then, observational capabilities in the X-ray band have significantly im- 7 proved, allowing to discover new LMXBs as well as to monitor them for either short (∼ 1 day) or long (∼ months) periods of time [BD23]. Such explosions, however, occur in the presence of accretion and represent an excellent opportunity to directly observe photons coming from the neutron star surface, and not from the output of mass accretion. This is not the only moment when the surface be- comes visible: while for some system the mass accretion occurs continuously, there also exist other systems where mass accretion ceases from time to time and, in the absence of in-falling material, the surface of the neutron star becomes visible [WMM+01, RBB+02, WDP17]. Such systems are said to exhibit transient accretion, while the period without in-falling mass towards the surface receives the name of quiesence. In practice, it is convenient to express all accretion rates in terms of the Eddington accretion rate for a 10 km neutron star, ṀEdd = 1.1 × 1018 g s−1, coming from the local rate per unit area, i.e. ṀEdd = 4πr2ṁEdd with ṁEdd = 8.8 × 104 g cm−2 s−1. In these units, for instance, we say bursts take place for the accretion rates between ∼ 0.01 and ṀEdd. Systems with Ṁ ≥ ṀEdd actually exist, such as XTE J1701-462[HvF+10] and MAXI J0556-332 [HFW+14]. But, what is the origin of the X-ray bursts? As a first step, the surface of the neutron star is enriched - via accretion - in low-Z material. For the case of main- sequence Sun-like stars, the composition is of ∼ 70% 1H, ∼ 28% 4He and the re- maining 2% of CNO metals, while in the case of a white dwarf or He-star companion matter is enriched in 4He and poor in 1H ([KiA+10]). Given the extreme gravita- tional acceleration at the neutron star surface (gs ∼ 1014 cm s−2, while at Earth gs ∼ 9.8 cm s−2), a temperature build-up is likely to take place: from Tmax ∼ 106 K to Tmax ∼ 108 K. Consequently, thermonuclear reactions eventually start operating at the surface, mainly converting 1H into 4He via pp-chains and CNO cycles, al- though synthesis of small amounts of 40Ca is still viable at moderate accretion rates (e.g. ∼ 0.1ṀEdd). However, this burning might become unstable: as long as fuel is added, the initial temperature build up due to accretion is further enhanced by the deposited heat from thermonuclear burning. If the cooling mechanisms, such as radiation and conduction transport, are not capable of distributing this heat then an extremely fast (≤ 1s) temperature build-up takes place (from T ∼ 108.25 to 109.25 K) and induces a thermonuclear explosion at ∼ 106 g cm−3, i.e. P ≈ 1022.5 erg cm−3. The energy transported at the surface increases the local temperature up to 107.25 K, e.g. L ∼ 1038 erg s−1, and during the following ∼ 100 seconds the surface relaxes and temperature decreases while the luminosity coming from the surface overcomes that from mass accretion. In the aftermath of this explosion, the thin matter layer of fuel at P ≈ 1022.5 erg cm−3 transformed, via thermonuclear reactions, into ashes enriched in high-Z elements of A = 56, A = 72, and even of A = 107, albeit in little amounts. This process of rapid captures of 1H and transformation into metals (e.g. less than 103 seconds) is known as rp-process. Under continuous accretion these explosions occur periodically, resulting in the accumulation of ashes at ρ ≥ 106 g cm−3. Despite its high-Z composition, these X-ray burst ashes between 106 and 1011 8 g cm−3 are far from being inert. Instead, heavy-Z burning proceeds via pycnonu- clear reactions, i.e. sensitive to density. As examples of these reactions we have carbon burning (12C + 12C) or electron-captures, which enrich the densest layers of the envelope (or the upper layers of the outer crust) with Z ≥ 20 material and injects up to ∼ 1 MeV per accreted baryon. Pristine composition 70 % 1H, 28 % 4He, 2 % Z ⊙ 12C, 24Mg, 40Ca 56Fe, 64Zn, . . . , 104Cd? 56Fe → 56Co → 56Ti 20Ne → 20O ρ ∼ 108 g cm−3 ρ ∼ 1011 g cm−3 ρ ∼ 106 g cm−3 · M Stable/unstable burning ashes: Fuel: Fusion/pycnonuclear Figure 2: A brief look at the composition of the accreted envelope. At this point, it would seem the picture is complete. Turns out, it is far from such state. For instance, there are two open questions regarding the end of the accretion phase, referred to as outburst as well, and the behavior of the bursts as a function of the mass accretion rate. In the first case, the necessity of invoking an additional heating source of unknown nature, between 0.1 and 10 MeV per accreted baryon at ρ ∈ [108, 1011] g cm−3 in order to explain the cooling of the neutron star after an outburst episode This constitutes the so-called shallow heating paradigm [BC09, DBW11, DWB+13, DCBP15, Mei18, CFZH20, PHN+22]. A still existing discrepancy between theory and observations is the apparent presence of a critical mass accretion rate above which bursting behavior is suppressed: while observations suggest 0.3ṀEdd [CiV+03, WM07, GMH+08, GK21], several theoretical models stop bursting at around and above ṀEdd [HCW07, FGWD06a, FTG+07, GK21]. There are, however, sources which still exhibit bursting behavior between 0.3 and 1 times ṀEdd, such as CygnusX-2 [Kv95], GX 17+2 [KHv+02], Circinus X-1 [LWA+10] and Terzan 5 [LAC+12]. Given the majority of simulations are in one spatial dimension only, such discrepancies leave open the possibility that additional effects - besides the shallow heating - such as diffusion of 4He or rotation ([PB07, IS10, CGG+20]) 9 might play an important role in order to unify these apparently different bursting regimes. Probably the biggest open question, among the existing ones, is the absence of a self-consistent long-term modeling of the neutron star in transient accretion. While the existing models rely on reasonable approximations regarding the physics of the star, some of the other open issues re-appear in these models and complicate the picture. Thus, it is important to settle why, despite the incredibly detailed picture given above, there still exist open questions. First of all, implementing nuclear reactions is numerically challenging since the system of partial differential equations (PDEs) is stiff, i.e. has many different scales involved and is thus difficult to integrate employing explicit methods (such as the workhorse 4th order Runge- Kutta) [PTVF92]. Assuming we find a method to integrate these equations (which have been extensively developed in the past 50 years and belong to the category of implicit methods), the next question to worry about is the number of species we need to incorporate into the network to obtain as realistic values as possible of generated energy and of chemical concentrations, which have a profound effect on the opacity κ (thus in the energy transport) and the equation of state (EOS) of the star. In order to have both proton- and neutron-rich isotopes, necessary to follow the whole picture described above and illustrated in Fig. 2, we require nearly 1000 species. And there lies the next big issue: we must solve ≥ 1000 PDEs (one for each species) at every spatial location in the star, i.e. ≥ 1000 cells. Assuming we can neglect the evolution of the network at ρ ≥ 1011 g cm−3, i.e. above the neutron drip point, we are still left with the task of simultaneously modeling the dense core and the surface, each with ≥ 1000 cells. The final issue concerns the timescale: it turns out the characteristic timescale at the surface is of ≤ hr, while the core evolves in ∼ months. Consequently, a full simulation of the whole neutron star undergoing accretion, even in 1D, is challenging. A quick review on the literature seems to contradict the above assertion. Be- fore and around 1978 [WZW78, FHM81] simulations of neutron stars undergoing accretion were indeed performed, although in retrospective their nuclear synthesis is inaccurate. For instance, the limited size of their networks allowed large amounts of 1H to be still present at ≥ 106 g cm−3 (deep hydrogen burning), the heaviest iso- tope to be synthesized was 56Fe and thus the possibility of pycnonuclear reactions was still out of the picture. Later works showed that the size of the network had an important impact on the chemical composition of the ashes, no longer rich in 1H [vGI+94, SBCW99]. Furthermore, they put into relevance the rp-process as the mechanism to synthesize metals far heavier than 56Fe. In addition, the enrichment in metals of the dense layers of the surface put into question the inert character of the crust, now possibly undergoing electron captures as a consequence of the com- pression and thus injecting energy. These pycnonuclear reaction gained relevance as well due to the discovery of the superbursts, which called out for the identifica- tion of fuel capable of liberating such tremendous amount of energy in a period of 10 time of days. In the modern era of networks, computers and telescopes, a review on the literature shows the simulation of neutron stars undergoing accretion is now focused on punctual aspects which restrict the time interval to be simulated, and thus alleviates the tension on the numerical resources: • For an accurate modeling of bursts, it suffices to model the envelope, i.e. ρ ≤ 107 g cm−3, by either multi-zone or one-zone simulations for a timespan sufficiently large as to obtain ≈ 20 bursts, i.e. t ≤ 1 week, in order to contrast them with observations [Mei18]. The size of the network can be as short as 300 or as large as 1500 [SBCW99, FGWD06a]. Techniques such as adaptative networks have been developed as well [HCW07]. • The end of accretion can be followed by either a detailed and large numeri- cal network, or by simple estimates on the secular burning of remaining fuel [SAB+01]. Since the typical time between the end of accretion and the relax- ation towards a “typical”, pre-accretion state, is of ∼ weeks, only the crust is usually evolved. The core, on the other hand, remains at a fixed temperature. • Existing long-term simulations of transient accretion and quiescent phases assume the heating can be modeled as ∝ Ṁ [BC09, PHN+22], and thus intro- duce this parametrization instead of solving a nuclear network. The envelope is treated as an atmospheric-like boundary condition, that is, giving an ex- plicit relation between the observed effective temperature Teff and the internal temperature Tb. This treatment regards the exiting luminosity from the core as constant in the envelope. Objective of the present work Then, what is the purpose of the present work? In principle, to find out if it is possible to accelerate the simulations of long-term evolution of accreting neutron stars by means of a Tb(Teff) scheme. This consists on replacing the whole simulation of the envelope by an atmospheric-like boundary condition. Under the assumption no heating and cooling sources exist in this region, we can know at all times the temperature at the surface as a function of the temperature at the envelope-crust boundary by simply solving dT/dP . Given the time independency of such equation, it thus suffice to solve it once, at the beginning of the simulation, extract the desired Ts(Tb) function (or data, and store them into a table) and implement them into the numerical code for the evolution of the whole star. The underlying assumption of no heating/cooling sources, however, prevents this approach to be employed in self- consistent simulations of accreting neutron stars. However, it has not been proved such scheme cannot be implemented for accreting neutron stars, naturally by taking into account the existence of the heating and cooling mechanisms. Indeed, given 11 the luminosity at all points depend on these sources, we could expect to have not only one function as boundary conditions, but two, one for Tb(Ts) and the other for Lb(Ts), i.e. the luminosity at the boundary becomes now a function of the surface temperature as well. As we shall see in the present work, the implementation is not straightforward, and the main complication arises due to the nature of the time- independent solutions to the stellar evolution equation at the neutron star envelope, which forced us to consider the proper time evolution of the system. For that part of the work, we employed an open-source code which is suitable for simulating neutron star envelopes, given its similar nature as with the standard stellar matter. For simulating the whole star, however, such publicly available code is not suited due to the absence of the high-density sector and fully general relativistic effects. In the process of studying time-dependent envelopes - and of course, of contrasting our stationary state results - additional properties of the envelope were re-examinated. For instance, given some stationary states predict Lb < 0, a quite unusual value for the standard works on the subject, we explored for their physical meaning and, naturally, whether they occur as part of time-dependent simulations or not. While the present work is devised as a compilation of research papers, it is best to provide the essential aspects of the context surrounding the results which we present in the compilation. Therefore, Chapters 1 to 4 are devoted to cover these essential aspects: in the first one, we summarize the equations of structure and thermal evolution governing the neutron star in the presence of accretion and, in particular, the surface. In Chapter 2 we analyze the microphysics of the neutron star envelopes. This is, concepts such as the Equation of State (EOS) or the nuclear reactions. Chapter 3 provides a brief summary on the nuclear networks employed in the research papers, as well as their physical justification. In Chapter 4 we wrap-up the content of the previous chapters into the study of neutron star envelopes, both in time dependent and independent scenarios. In Chapter 5 we provide a brief word summarizing the whole research, and the perspectives for the future. On the numerical codes in the present work All the contents of this work, related to the evolution of accreting neutron star envelopes, were obtained from different numerical codes, enlisted below: • TOV code. Developed by the author of this work, solves the relativistic struc- ture equations for a neutron star (TOV), as well as the stationary equation for quadrupolar deformation (see for instance [Hin08]), given as input an EOS in tabulated form. The output from this private code was employed in [NCPB23] for the GR structure. • Developed code. In order to simulate one zone models and stationary neutron star envelopes. The driver routines for adjusting the parameters, the subrou- 12 tines to implement the EOS, opacity and reaction rates were designed by the author of this work, while subroutines related to numerical methods for solving equations (as for instance stifbs) were taken from [PTVF92]. • MESA (Modules for Experiments in Stellar Astrophysics). Public code which solves the time-dependent stellar evolution equations in one spatial dimension [PBD+11, PMS+15]. Let us stress that MESA is not designed to evolve complete neutron stars. The capabilities of this code extend to up to ρ ≈ 1010 g cm−3, i.e. at the envelope-outer crust boundary, where relativistic corrections of the form e±Φ are still viable. • NSCool. Developed by Dr. Page and collaborators [Pag16]. This public code is designed to model one-dimensional thermal evolution of isolated neutron stars. In the presence of accretion, this code implements an approximate treatment of the heating sources as Qheat ∝ Ṁ , with Ṁ the mass accretion rate, and considers the envelope is described via the standard Tb−Teff relations for non-accreting envelopes [BC09, PHN+22]. As work in progress is the im- plementation of the results from the present thesis into NSCool, thus allowing it to handle thermonuclear burning as self-consistently as possible. 13 Chapter 1 Equations of Stellar Evolution “And yet [...] there are mathematicians of extraordinary genius who doubt the whole universe [...] was created only according to Euclidean geometry [...]” F. Dostoievsky Abstract. In this chapter the differential equations governing the neutron star under mass accretion - and specifically the envelope - are introduced and discussed. 1.1 Introduction The modeling of stellar evolution is possible through the fluid approximation, which consists on following the spacetime evolution of a set of physical quantities A (such as pressure, temperature and composition), characterizing the state of a fluid element, i.e. a blob of ≳ 1023 particles, on a spatial scale much larger than the intrinsic size of the constituent particles, which for a good approximation can be assumed as their Compton wavelength, λc = h/mc. An advantage of this approximation is the possibility to separately construct relations, to some extent regarded as functionals, among the physical quantities. These are further referred to as microphysics, and the actual values of them as locals since they are calculated and measured on an inertial frame. As important examples of these relationships we can mention the Equation of State (EoS), which relates the pressure, temperature and density, and the opacity, which is a measure of how much energy photons lose when passing through matter and depends on temperature and density. 14 To propagate A along spacetime, a set of equations compatible with general relativity is needed. This is the task of section X, where we introduce and discuss in detail this set of partial differential equations. A review on the indispensable concepts and mathematical framework occupies the first subsection. This chapter introduces the basic set of equations which are employed to describe thermal evolution of compact objects. In this work, we stick to the metric convention (−,+,+,+). We denote 3 and 4-vectors by V and V⃗ respectively. Similarly, we reserve ∇µ to denote the 4-divergence. As all the systems discussed in this work obey the conservation of baryons, for any physical variable of interest X we adopt the following conventions: x̃(= X/Nb) denotes X per baryon, x(= X/V ) the X-density, and x̆(= x̃/mu) the specific-X, i.e. X per unit mass. As baryons is a surrogate name for the mixture of protons and neutrons, it is appropriate to regard the atomic mass unit mu as the baryonic mass since mp ≈ mn ≈ mu. Recalling the value of Avogadro’s number NA = 6.02214076 × 1023 particles mol−1 , (1.1) the atomic mass unit mu (or amu for simplicity) is thus defined as the 12th part of the ratio of 12 g of 12C and 6.022×1023, which is the amount of atoms in a mole of this ion: mu = 1 g 6.022 × 1023 particles = 1 g mol−1 NA . (1.2) 1.2 Spacetime evolution of physical quantities So far, the best theory for describing self-gravitating objects is General Relativity (GR) [Wal84], [MTW17], which serves as a generalization of Newton’s Mechan- ics. A geometric theory by definition, GR postulates the structure of spacetime is completely determined by the components of a metric tensor gµν , related to the line-element of a coordinated system by ds2 = gµνdx µdxν , influenced by a collection of local mass-energy fields Ψ exerting mechanical work. From a modern point of view, the PDEs governing all these fields can be deduced from variational principles over the associated action functional S = 1 c ∫ d4x √−g [Lgrav(gµν) + Lmatter(gµν ,Ψ)] . (1.3) Indeed: from δSGR/δgµν = 0, we obtain the following set of non-linear PDEs, com- monly referred to as Einstein’s Field Equations Gµν = 8πG c4 T µν . (1.4) The left-hand side, the Einstein tensor, is purely geometric and thus involves gµν and their partial derivatives. By construction, it satisfies ∇µG µν = 0 and Gµν = Gνµ, 15 thus implying the existence of 10 independent PDEs in 4 dimensions. On the right hand side we have several constants allowing to recover the Newtonian equations in the non-relativistic limit: G, the Cavendish constant of gravitation, and c the speed of light in vacuum. The tensor appearing in the right hand side is defined as the energy-momentum tensor Tµν = − 2√−g δ( √−gLmatter) δgµν , (1.5) and by definition contains the contributions from the energy-matter fields Ψ. Due to the the enumerated properties of the Einstein tensor, Tµν = Tνµ and 1√−g∂µ (√−gT µν ) = 0, (1.6) an identity which yields four additional PDEs, for Ψ’s spacetime evolution. While the original formulation of GR demanded T µν to solely contain fields exerting me- chanical work [Wey22], a role heat fluxes clearly do not fulfill, more recent works on the subject have relaxed this condition in order to self-consistently incorporate the First and Second Laws of Thermodynamics into the grand scheme of GR, an extension not necessarily free of controversies [GS06]. While in principle we can sep- arate T µν as a sum T µνmechanical + T µνheat, from the classic point of view only T µνmechanical, not the sum, satisfies Eq. 1.6, a vision we shall adopt in order to later incorporate thermodynamical quantities into the fluid description. The most frequently used model to describe stellar energy-matter is the perfect fluid, an inviscid system with zero exchange of entropy, moving with unitary 4- velocity uµ (i.e. uµu µ = −1, or equivalently Uµ = cuµ). Its corresponding energy- momentum tensor is Tµν = εuµuν + PΠµν , (1.7) where Πµν = gµν + uµuν is the orthogonal projection tensor satisfying uµΠµν = 0, ε is the total energy density and P the thermodynamical pressure. These scalars are interconnected via the Equation of State, a functional relating all thermody- namical degrees of freedom with P . For instance, in the non-relativistic limit, P is typically related to the temperature T 1 and the individual number densities ni (i ∈ ¶1, . . . , Nspecies♢) for multi-species systems. On the other hand, for degenerate fermionic systems (i.e. where T/TF ≪ 1, with TF the Fermi temperature) P be- comes a functional of the total density ρ := ε/c2 only, i.e. P (ρ). Finally, for some discussions in the latter example it is useful to consider both P and ε as explicit functionals of the baryon number density nB, i.e. P (nB) and ε(nB). This latter 1T is also usually employed as the notation for the trace of T µν . While this quantity is not employed in the present work, whenever potential confusions might arise we shall explicitly state what we are denoting by T . 16 quantity plays an important role in the description of stellar matter, as in the non- relativistic limit ρ ≈ ρB, with ρB ≈ munB. Since experimental evidence support the notion of conservation for this quantity, in General Relativity we must also include this fact via a covariant identity, 1√−g∂µ [√−gnBu µ ] = 0 , (1.8) an expression which can also be expressed, via the Gauss theorem, as the conserva- tion of a “charge” on a Cauchy hypersurface Σ, the total number of baryons: NB = ∫ Σ d3x √ γ nBkµu µ, (1.9) with kµ a unitary normal 4-vector to Σ. Another important property from Eq. 1.8 we must emphasize is that high-density environments tend to have low fluid velocities. Conceptually, n (1) B uµ(1) ∼ n (2) B uµ(2), hence if n (1) B > n (2) B we have ♣uµ(2)♣ > ♣uµ(1)♣ for all µ, i.e. the fluid moves faster in low-density environments than in high-density ones. Roughly speaking, we thus have v ∝ n−1 B , with v the norm of the 3-velocity field. In spite of its simplicity, Eq. 1.7 has earned his place as the workhorse by excel- lency for studying stellar structure, a situation we can easily understand by analyzing the projections of the 4-vector expressing its conservation, Eq. 1.6. The first is along the fluid 4-velocity, uµ∇νT νµ = 0. Taking into consideration Eq. 1.8, we obtain uµ∂µε = ε+ P nB uµ∂µnB, (1.10) closely resembling the First Law of Thermodynamics in the absence of terms related to entropy exchange and variations in the individual species’ number densities, i.e. either changes in entropy are zero or T → 0, and dni → 0. On the other hand, the contraction Παβ∇µT µα = 0 leads to (ε+ P )aα = − [∂αP + uαu µ∂µP ] , (1.11) where we identified the 4-acceleration, aα = uµ∇µuα. (1.12) Eqs. 1.11 are usually referred to as momentum equations [MTW17] since they gen- eralize the non-relativistic behavior for ρu. Notice enthalpy per unit volume, ε+P , serves as the relativistic generalization of the matter density, indicating pressure plays an important role in fully-relativistic systems. Apropos of this statement, let us examine this non-relativistic limit, where ♣u0♣ ≫ ♣ui♣ and ε ≫ P . For the spatial components we have ε [ u0∂0uj + ui∂iuj + u0Γ0 0ju0 + O(u0ui) ] = − [ ∂jP + O(u0ui) ] ε [ u0∂0uj + ui∂iuj + u0Γ0 0ju0 ] ≈ −∂jP. 17 Identifying Γ0 0j = ∂j(Φnr/c 2), with Φnr the non-relativistic “gravitational poten- tiall”, we obtain ε [ 1 c2 ∂vj ∂t + 1 c2 vi∂ivj − 1 c2 ∂jΦnr ] ≈ −∂jP, ρ [ ∂t + vi∂i ] vj ≈ ρ∂jΦnr − ∂jP, which is the standard form of Euler’s equation. Here, we see the Christoffel symbols give rise to the usual contribution due to the gradient of the gravitational field, while the “kinetic” term, uν∂νuj, encloses the usual variation in time of uj but also the contribution from the so-called ram pressure, ρvi∂ivj. Given this construction, by looking back to Eq. 1.11, taking into consideration the explicit form of aα, Eq. 1.12, we see the relativistic analogous of the ram pressure is given instead by (ε+P )ui∂iuj. For the majority of applications, it is safe to neglect this term as is small in com- parison with either temporal variations of the 4-velocity or the “gravitational field” enclosed in the Christoffel symbols. As EFE were envisioned to describe the structure of self-gravitating objects the next, natural question is: how can we analyze thermal degrees of freedom within GR? The first step is to recall in the non-relativistic (i.e.Newtonian) regime, for each member of fluid’s collection of physical variables A we can associate a single or a collection of 4-currents, for simplicity denoted as J µ A , obeying ∂µJ µ A = ΓA , (1.13) where the right-hand side can be interpreted as a source or sink of A according to sgn ΓA, the particular case ΓA = 0 defining the the conservation of A. Notice the scalar character of the left-hand side demands all contributions on the other side to be scalars as well. By inspection, Eq. 1.13 is valid within Special Relativity due to its invariance with respect to the coordinated system, feature which allows its incorporation into GR as long as the partial derivatives are promoted to covariant ones, 1√−g∂µ [√−gJ µ A ] = ΓA , (1.14) as we have already seen exemplified by the conservation of baryon number density, Eq. 1.8, expressing the zero divergence of a 4-vector, but also with Eq. 1.6, a 4-vector resulting from the conservation of a tensor, T µν . Since Eq. 1.14 automatically sat- isfy the requisites of being covariant and to preserve causality, within GR 4-currents must play a critical role in describing the spacetime evolution of A. There are, how- ever, further requisites “the correct”, fully relativistic version of Thermodynamics must fulfill: expressions for the 4th Laws, causal propagation of heat fluxes and the absence thereof within the non-relativistic regime (e.g. Fourier’s law). To date, one of the most successful attempt for constructing Relativistic Thermo- dynamics was introduced by [Isr76, IS79]. At its core, this model promotes several 18 thermodynamical quantities of interest to 4-vectors and tensors and exploits the covariant character of Eq. 1.13 to propose a generalized version of the 1st Law of Thermodynamics employing the classic Gibbs’ differential for a system composed of multiple species, dSµ = ∑ i αidN µ i + βνdT νµ. (1.15) In this differential, Sµ and Nµ i are the entropy and i-th species’ number density 4-vectors, αi are the generalized chemical potentials, T νµ is the complete energy momentum tensor and βµ = T−1uµ defines the inverse temperature Killing 4-vector, where T is the temperature measured at equilibrium. Conceptually, this relativistic Gibbs relation is expected to be valid at both reversible and irreversible conditions since all tensorial quantities admit an expansion of the form Tequilibrium + Tirreversible, where Tequilibrium is typically related to the fluid’s 4-velocity uµ, and the irreversible part Tirreversible is a sum of first, second or n-th order contributions, albeit for the majority of applications is sufficient to retain up to first order. As a competing formulation for relativistic thermodynamics, we must mention the variational approach of [Car91], as well as related works on the subject concern- ing the equations for a perfect fluid [Bro93], [KB97], [Sch70] (see also [LA11] for an excellent review on this subject). The advantage of such scheme is the easiness on the incorporation of additional fluid components, as well as couplings among them as in the superfluid scenario. In spite of their apparent differences, both Is- rael’s and Carter’s approaches are in good correspondence up to second order in deviations from equilibrium [Pri91], although both formalism differ when significant dissipation takes place. This similarity also extends to the heat transport, also re- ferred to as relativistic Cattaneo equation as in the non-relativistic regime we expect τ Ḟ + F = −K∇⃗T , expression with finite-speed propagation of heat. From the pre- dictions of both theories, it becomes easy to see Fourier’s law can still be kept as “correct” within relativistic and astrophysical systems as long as the flux is steady on a thermal timescale [LA11], [LA18]. In spite of both models’ success on constructing causal thermodynamics, as well as in their similarities under thermal equilibrium, the most prominent sources of controversies are the apparent excess of parameters needed for both theories and their physical interpretation, the non-mechanical terms appearing in the energy- momentum tensor and possible instabilities due to the presence of 4-acceleration [GS06]. While the source of these observations do not seem to be impediments for regarding Israel & Stewart and Carter theories as correct, in particular as some terms admit conceptual justifications [LA11], [Bro93], there exists a strong objec- tion not only for these approaches, but for all those attempts to relativistic ther- modynamics: the “correct” transformation law for actual temperature. This is even more accentuated in the expressions found in both physical and astrophysical litera- ture where, apparently, the relationship between observed and local temperatures is T∞ = eΦTlocal, i.e. objects are colder when observed at an infinite distance, an affir- 19 mation more closer to the Einstein-Planck point of view than to Ott’s or Landsberg [DHH09], [Bec16], [FPM17]. This “transformation law”, as we explain later in the text, remains valid as long as one is interested in determining the temperature asso- ciated with the peak in frequency of the Planck function. However, the postulation of an entropy 4-current and the existence of the 4-momentum, in combination with the First Law of Thermodynamics, lead us to the conclusion that it is safer to as- sume the actual temperature as a scalar function. This point of view is supported by numerical simulations of particle interactions in the relativistic regime [CCPD+07], [FPM17] where, apparently, the definition of a frame-independent temperature is within reach. However, we advise the reader to notice that such treatment of tem- perature strongly depends on the set of initial assumptions, a neutral point of view which is valid as long as controversies over the correct formulation of relativistic thermodynamics exist [FPM17]. Let us now describe Israel & Stewart approach for relativistic thermodynamics in terms of 4-divergences, taking uµ as the 4-velocity of the fluid element. Taking T as the temperature measured in a local frame, βµ = T−1uµ defines a Killing vector (e.g. ∇µβν = −∇νβµ) taken as the inverse of temperature. From EFEs, we already know the energy-momentum obeys a conservation law, Eq. 1.6. In this regard, we also know the baryons obey a conservation law, Eqs. 1.8 and 1.9. For systems composed of Nspecies different species, it is necessary to attach a 4-current to each one of them, Nµ i := niU µ i +Qµ, together with a corresponding PDE of the form 1√−g∂µ [√−g niUµ i ] = Gi , (1.16) [Gi] = Baryon Volume−1 Time−1, where the non-equilibrium term has been absorbed into the contributions from the right-hand side to emphasize Gi expresses the net rate at which all other species contribute with creation, annihilation, scattering or diffusion for the i-th species, which in general are functionals of temperature, rest- mass density or the individual number densities. In general, Uµ i ̸= Uµ, a difference which can be accentuated in non-vacuum environments where the background den- sity determines which species fall at faster rates than others2. However, at supersonic velocities encountered as for instance in typical surfaces of neutron stars it is a valid approximation to consider Uµ i ≈ Uµ ∀i ∈ ¶1, . . . , Nspecies♢ in the left-hand side of Eqs. 1.16. Diffusive effects can thus be treated as corrective terms and included as a source in the right-hand side. An important consequence of this approximation is the possibility to write Eqs. 1.16 in terms of abundances Yi := ni/nB and, employing Eq. 1.8, obtain a scalar-like equation for each species, Uµ∂µYi = Rlocal i + Rdiffusive i , (1.17) with Rx := Gx/nB for x =local, diffusive, and where, for further convenience, the contributions in the right hand side have been split as local if they solely depend on 2Terrestrial experiments are a quick corroboration of this assertion 20 variables such as nB, T or Yi, and diffusive if they include first or second derivatives of any Yi. From an astrophysical point of view, Rlocal i represents the variations in abundances due to reaction processes, in particular, of nuclear reactions. Finally, we must note in environments where ρ ≈ ρB holds, Eqs. 1.17 are constrained by the mass-conservation condition, i.e. Nspecies ∑ i AiYi = Nspecies ∑ i Xi = 1, (1.18) with Ai the number of baryons for the i-th species and Xi := AiYi denotes the mass fraction. The next quantity of interest is the entropy 4-current, Sµ = nBs̃U µ +Qµ s , where the second term denotes all non-equilibrium contributions. While the corresponding PDE is again of the form in Eq. 1.14, the presence of nBU µ in Sµ allows to employ the conservation of baryon density, Eq. 1.8, to simplify this equation into nBU µ∂µs̃ = Gs ≥ 0 , (1.19) where the contribution due to ∇µQ µ s has been absorbed in the right-hand side term Gs, referred to as the net entropy production rate. In order to generalize the Second Law of Thermodynamics this term must be non-negative, hence the restriction in Eq. 1.19. If Gs > 0 then we say our system is irreversible, and whenever the equality holds we have a reversible one. From conceptual grounds, it is useful to explicitly categorize all contributions to Gs as heating or cooling mechanisms, i.e. Gs = Gheat - Gcool, in order to understand reversibility as the limiting scenario when the net rates of heating and cooling balance each other, Gheat = Gcool. Within the astrophysical context, as heating mechanisms we can mention the energy release from nuclear reactions, which tends to increase the temperature of the system, or the compres- sion of magnetic fields. For cooling mechanisms, on the other hand, we have the emission of photons, described by radiation and conduction fluxes, and the escape of neutrinos due to its small cross section for interactions with baryonic matter. As a consequence of Eq. 1.15, the general relativistic version of all these processes must be described either by variations in the species’ densities Nµ i or the changes in the energy-momentum tensor for the irreversible case. From these considerations, let us write Gs = Gs,N + Gs,T and separately analyze each term. Albeit neutrinos cannot be handled in the same foot as any of the other species in the mixture being out-of-thermal equilibrium with the rest of the system, con- ceptually its contribution to Gs resembles a variation in an hypothetical abundance 21 Yν . Thus Gs,N := ρB T [ ˙̆εspecies − ˙̆εν ] , ˙̆εspecies = ˙̆εnuc + ˙̆εdiff ˙̆εx = − Nspecies ∑ j=1 µ̆jRx j , x ∈ ¶local, diffusive♢ , with µ̆j the specific chemical potential per unit baryon for the j-th species. As previously stated, within the astrophysical context the x = local term can be taken as the definition for the energy generation rate due to nuclear reactions, as long as neutrino losses due to electroweak interactions are included in ˙̃εν . Since Rlocal j is typically < 0, the minus sign at the front allows a positive energy generation rate. Regarding Gs,T , we have the contributions due to heat to the energy momentum tensor. To incorporate energy diffusion due to radiative and conductive processes, we follow the approach of Tauber, Weinberg, Eckart and Misner [Eck40], [MS69], which is correct to first-order within the context of Relativistic Thermodynamics, despite some differences on notation [Isr76]. Let F µ be the flux 4-vector resulting from the addition of radiative and conductive contributions, satisfying F µFµ ≥ 0 and F µuµ = 0, i.e. heat propagates in a perpendicular direction to the motion of the fluid element. Defining a radiation energy-momentum tensor T µνrad = F µuν + F νuµ, it is immediate to write a frame-invariant rate of energy generation [MS69], Gs,T = T−1uν∇µT µν rad = −T−1 [∇µF µ + F νaν ] . (1.20) By demanding a positive-definite rate, as well as a first order expression in temper- ature gradient ∂νT and 4-acceleration aν , Eq. 1.12, it is possible to deduce F µ = −KΠµν [uα∇α(Tuν) − uα∇ν(Tu α)] (1.21) = −KΠµν [∂νT + T aν ] . (1.22) where K is the usual thermal conductivity, expected to reproduce its non-relativistic form 16σSBT 3/3κρB (but consequently neglecting the possible presence of anisotropies due to magnetic fields) and Πµν is the orthogonal projection tensor already intro- duced. To complete our discussion of radiation, we introduce the luminosity, energy per unit time carried away by photons. For some authors [Eck40], [MTW17], as the stellar structure is frequently regarded as purely radial the luminosity is frequently traded by 4πr2F , with F the absolute value of FµF µ, or by its simplified form 4πr2√grrF r. Save for r2, this term is invariant under the exchange of coordinates, but limits the angular dependence to be constant. Instead, we adopt as definition of luminosity the natural L := − ∫ Σ d3x √ γ uν∇µT µν rad, (1.23) 22 which can be shown to recover the well-known relation L = 4πr2F for spherically symmetric systems (see Proposition 7). In summary: Eq. 1.19, considering the contributions from variations of species’ densities and radiation, can be written as nBTU µ∂µs̃ = ρB [ ˙̆εspecies − ˙̆εν ] − [∇µF µ + F νaν ] . (1.24) On the temperature gradient. Due to the Equivalence Principle, since both T and P are local functionals (or scalar functions), in principle we can still adopt the non-relativistic definition for the actual temperature gradient ∇T , i.e. ∂T ∂s := ∇T T P ∂P ∂s , (1.25) with s a spatial coordinate (such as the radial direction r), and apply the same formalism of standard stellar evolution to determine whether ∇T = ∇r or ∇T = ∇conv, with ∇r and ∇conv the radiative and convective gradients respectively: ∇T =    ∇r, ∇ad > ∇r ∇conv, ∇ad < ∇r, (1.26) with the adiabatic gradient ∇ad, provided by the EoS. Notice this compels us to find an actually local expression for ∇r, and to adopt any of the existent treatments for convection to compute ∇conv whenever radiation exceeds the adiabatic contribution. 1.3 Evolution in a static and spherically symmet- ric (SSS) spacetime - fluid in radial motion Employing spherical coordinates (ct, r, θ, ϕ), the line element of this spacetime is ds2 = −e2Φ(r)c2dt2 + e2Λ(r)dr2 + r2dθ2 + r2 sin2 θdϕ2 . (1.27) Being part of the g00 component of the metric, and thus related to variations in time at fixed spatial coordinates, eΦ(r) is referred to as the redshift factor, or redshift for shortness. While eΛ does not have a specific name, in vacuum it fulfills the condi- tion eΛ(r) = e−Φ(r), usually referred to as the Schwarzschild metric as it describes the exterior spacetime of a black hole, and even more generally, of any spherically symmetric object such as a neutron star or the Sun itself. For a complete list of factors related to this metric (determinant, Christoffer symbols), see Appendix B.2. When discussing vectors in this spacetime, it is customary to distinguish between local, i.e. as measured by an orthonormal frame of reference, and at infinity quanti- ties. Since the SSS metric is diagonal, we can easily construct an orthonormal frame of reference: denoting such coordinates with hats, the relationship among basis are eα̂ = 1 √ ♣gαα♣ eα (1.28) 23 for α ∈ ¶ct, r, θ, ϕ♢ and where no summation is intended for these indices. As any 4-vector admits an expansion in terms of both basis, we have V⃗ = V µeµ = V µ̂eµ̂. Consequently, the relationship between the components is V µ̂ = √ ♣gµµ♣V µ, where once again no summation is intended for the repeated indices. Thus, we formally identify local quantities as Vlocal = V µ̂ = √ ♣gµµ♣V µ . (1.29) As we expect these variables to represent a measurable physical quantity, a non- orthonormal (or non-inertial) observer can only attain physical meaning if these variables are propagated along a timelike curve. Thus, the quantities measured at a distance, or simply measured at infinity, are related to the local ones as V ∞ = √ ♣g00♣Vlocal . (1.30) It is important to notice this “transformation law” is valid for vector components. From the followed steps, we can easily see Eq. 1.28 defines the “transformation law” for covector components. To exemplify these concepts, let us examine three important scenarios. First of all, for massive particles we know its line element can be also expressed as ds2 = −c2dτ 2, where τ is defined as the proper time of the particle. Notice τ plays the role of an orthonormal coordinate, so we can identify it as the local measured time tlocal = τ . Now, due to the invariance of the line element, for events occurring at a fixed location in space, i.e. dr = dθ = dϕ = 0, we have an explicit relation between tlocal and the time measured by the t coordinate, t∞ = e−Φtlocal, hence t is the time measured by an observer at infinity, as expected by the covector “transformation law”, Eq. 1.28. Now, let us study the energy for a single particle. Let p⃗ denote its 4-momentum, p0 having units of Energy divided by c. An observer in an inertial frame of reference seeking to measure the energy of such particle, according to the general expressions given above, will measure Elocal = eΦp0c. On the other hand, an observer at infinity will measure an energy E∞ = eΦElocal = e2Φp0c = −p0c, in agreement with the “conservation of energy” arising from the time-independency of the SSS metric (see for instance [Sch09, MTW17]). A typical application of this analysis is the case of photon’s frequencies, in particular those coming from Wien’s displacement law, the maximum effective temperature assuming a blackbody radiation: considering the standard relations E = hν and νpeak = c1Teff, with c1, h and kB constants, (the latter known as Planck and Boltzmann constants, respectively), we have ν∞ = eΦνlocal (1.31) T∞ eff = eΦTlocal,eff . (1.32) 24 The final scenario for consideration is the ratio of physical quantities, such as the the luminosity, i.e. energy per unit time. Employing the identities so far deduced, we have L∞ = E∞ t∞ = eΦElocal e−Φtlocal = e2ΦElocal tlocal = e2ΦLlocal, (1.33) i.e. the luminosity at infinity is twice redshifted with respect to the locally measured value. Before leaving this discussion, we must stress out that, despite the usefulness of the local-at infinity relations so far introduced, attempts to generalize the stellar structure and thermal evolution equations from Newton to Einstein formalism via attachment of redshift factors has a limited range of applicability. For instance, for fluids at rest and constraining the system to remain spherically symmetric, i.e. most quantities depending on ct and r. As we shall see below, however, not in all scenarios it is possible to construct such promotion. Fluid in radial motion. Let us now deduce the aspect of the normalized 4- velocity of a fluid element moving along the radial direction in the SSS metric3. Due to the diagonal character of this metric, we can easily identify orthonormal coordinates dt̂ = eΦdt, dr̂ = eΛdr. Let v be the 3-velocity measured by a local observer at a coordinate r, e.g. v = dr̂/dt̂ = eΛ−Φ dr dt . Denoting the proper time by τ , the general expression for the components of the fluid 4-velocity vector U⃗ is (U ct, U r, U θ, Uϕ) = dt dτ ( c, veΦ−Λ, 0, 0 ) . (1.34) To obtain the front factor, we consider the normalization restriction, UµU µ = −c2. This leads to dt dτ = γSe −Φ, (1.35) γS = [ 1 − v2 S ]−1/2 , (1.36) where γS denotes the traditional Lorentz factor, vS = v/c, and we added a subindex S for the SSS metric in order to avoid confusion with other γ’s and v’s. Thus, the desired expression for the normalized 4-velocity uµ = Uµ/c is (uct, ur, uθ, uϕ) = γS ( e−Φ, e−ΛvS, 0, 0 ) . (1.37) Eq. 1.37, upon the explicit replacement v → v(t, r), can be adopted as the definition of the normalized fluid 4-velocity field in spherical coordinates. Notice the only 3In order to reduce the amount of mathematical details, many of the expressions employed in this part are given without proof. For further details on their origin and intermediate steps, see Appendix B. 25 non-zero components of the orthogonal projection tensor are: Π00 = e−2Φγ2 Sv 2 S (1.38) Πrr = e−2Λγ2 S (1.39) Π0r = γ2 Se −Φ−ΛvS. (1.40) For further reference, it is useful to note the projection of the 4-gradient along uµ and Πµν are uµ∂µ = γSe −Φ c [ ∂t + veΦ−Λ∂r ] , (1.41) Dt := e−Φ∂t + ve−Λ∂r (1.42) Π0ν∂ν = γ2 Se −2ΦvS [ vS c ∂t + eΦ−Λ∂r ] , (1.43) Πrν∂ν = γ2 Se −Λ−Φ [ vS c ∂t + eΦ−Λ∂r ] , (1.44) where Dt can be identified in orthonormal coordinates t̂, r̂ as the Lagrangian deriva- tive. From its coordinates, and the Christoffel symbols for the SSS metric, it is straightforward to see the fluid 4-acceleration a⃗, whose components (see Prop. 5) are given by a0 = u0∂0u0 + ur∂ru0 = γ2 SvS { γSe Φ c uµ∂µv − eΦ−ΛdΦ dr } (1.45) = γ2 SvS { γ2 S c2 [ ∂tv + eΦ−Λv∂rv ] − eΦ−ΛdΦ dr } (1.46) ar = u0∂0ur − u0∂ru0 = −u0 ur a0 (1.47) = γ2 Se Λ−Φ { −γSe Φ c uµ∂µv + eΦ−ΛdΦ dr } (1.48) aθ = 0 (1.49) aϕ = 0. (1.50) Let us examine now the continuity equation. By direct computation, there are several forms we can express it, each one useful in its on way: ∂t(nBγS) + 1 r2eΛ ∂r(e Φr2nBγSv) = 0 (1.51) [∂t + eΦ−Λv∂r](nBγS) + nBγS r2eΛ ∂r(e Φr2v) = 0 (1.52) [∂t + eΦ−Λv∂r]nB − γ2 SnBv c2 [∂t + eΦ−Λv∂r]v + nB r2eΛ ∂r(e Φr2v) = 0. (1.53) 26 The second form of the continuity equation, Eq. 1.52, results useful for a quick analysis within Minkowski spacetime, where both eΦ and eΛ approach unity. Taking the limit γS → 1 we directly recover the traditional form of the continuity equation, i.e. ∂tnB + u · ∇⃗nB + nB∇⃗ · u = 0. A more careful approach, within a Special Relativistic context, can be extracted from Eq. 1.53 as we have explicitly stated the action of the partial derivatives over the velocity field v(r, t). Assuming time-independency of both nB and v, the first of these expressions, Eq. 1.51, implies the existence of a conserved quantity, the redshifted mass accretion rate Ṁ∞ := −4πr2eΦmunBγSv, which encloses both special and general relativistic effects via γS and eΦ respectively. Notice we have chosen a negative sign for the definition as for accreting matter we expect v < 0. To consider the impact of these terms over the surface of a typical neutron star, let us fix munB ∼ 10−3 g cm−3, the gravitational mass and radius as m∗ ∼ 1.4M⊙ and r∗ ∼ 12 km respectively, thus having eΦ ≈ √ 1 − 2Gm∗ c2r∗ ≈ 0.80 and the extreme case Ṁ∞/4πr2 ∗ = 8.8 × 104 g cm−2, i.e. the Eddington rate of accreted mass per unit time and surface. These lead to γSvS ∼ 3.67 × 10−3, i.e. vS ∼ 3.66 × 10−3. On the other hand, if Ṁ∞/4πr2 ∗e Φ = 8.8×104 g cm−2, we obtain γSvS ∼ 2.94×10−3, i.e. vS ∼ 2.94×10−3. Consequently, the redshift factor has a slightly more prominent impact over v than the SR correction due to γS, at least as long as ρ ≥ 10−3 g cm−3. Due to the smallness of vS, in this density regime it is safe to set γS ≈ 1. Furthermore, from the continuity equation we also expect v to become progressively smaller than its surface value as we move inwards since v ∝ ρ−1. From these considerations, we thus adopt the First-order approximation (FOA): Given vS ≪ 1, γS → 1 and all vS/c and v2 S terms will be considered as small perturbations and thus neglected. Within the FOA, the baryon conservation (continuity equation) we adopt is a re- duced version of Eq. 1.53, namely ∂tρB − e−Λ 4πr2 ∂rṀ ∞ = 0, (1.54) with ρB = munB and Ṁ∞ = −4πr2eΦρBv (1.55) as the redshifted mass accretion rate. In what follows, we employ the general expressions for the physical quantities of interest, deduce the general expressions for the complete 4-velocity uµ, Eq. 1.37, and then apply the FOA. Einstein Field Equations. For the analysis of these and subsequent equations, 27 it is useful to introduce the gravitational mass function m(r) := c2r 2G (1 − e−2Λ(r)). (1.56) For the SSS metric, the only non-vanishing components of the Einstein tensor are G00, Grr, Gθθ and Gϕϕ. Due to the symmetries of the problem, only the first two are necessary for the discussion, hence we shall only need T00 and Trr from the energy-momentum tensor. Given G00 = 2Ge2Φ c2r2 dm dr Grr = −2Gme2Λ c2r3 + 2 r dΦ dr T00 = γ2 Se 2Φ [ ε+ v2 SP ] Trr = γ2 Se 2Λ [ P + v2 Sε ] , from EFE we obtain two differential equation, one for the gravitational mass of the star and the other for the redshift function, dm dr = 4πr2 c2 γ2 S [ ε+ v2 SP ] (1.57) dΦ dr = Gre2Λ c4 [ mc2 r3 + 4πγ2 S ( P + v2 Sε ) ] . (1.58) For v ̸= 0, we see the 3-velocity provides additional contributions to ε in the grav- itational mass case, and to P for the redshift function. However, within the FOA these can be neglected, and we thus recover the standard set of equations for the v ≡ 0 case, dm dr = 4πr2ρ (1.59) dΦ dr = geΛ c2 G, (1.60) where, following Thorne et al, we have introduced some factors for further conve- nience: g := GmeΛ r2 (1.61) G := [ 1 + 4πPr3 mc2 ] . (1.62) 28 Euler equations. Considering the only nonzero components of the 4-acceleration are a0 and ar, as well as the normalization condition u0u0 + urur = −1, we have (ε+ P )a0 = ur [ur∂0P − u0∂rP ] (ε+ P )ar = −u0 [ur∂0P − u0∂rP ] , in agreement with the functional relation between ar and a0, Eq. 1.47, indicating only one equation is actually independent. Thus, it is necessary to solve only one equation, for simplicity the first one. Plugging the expanded expression for a0, given in Eq. 1.46, we obtain (ε+ P ) { γ2 Se Λ c2 Dtv − dΦ dr } = 1 γ2 S ∂rP + vSe Λ c DtP. (1.63) Several particular cases can be identified from Eq. 1.63: • Taking v ≡ 0, we recover the well-known hydrostatic equilibrium relation, ∂P ∂r = −Hc2dΦ dr (1.64) H = ρ+ P c2 . (1.65) • For v ̸= 0 and under the FOA we have HeΛ ¶Dtv − gG♢ = ∂rP. (1.66) In the left-hand side we have a competition between fluid’s 3-acceleration, Dtv, and the gravitational acceleration gG. At the surface of neutron stars, this second quantity has an order of magnitude ∼ 1014 cm s−1. If we consider accretion at the Eddington rate, we have seen v ∼ 10−3c ≈ 107 cm s−1 near the surface. Thus, unless very rapid changes in velocity occurs, i.e. in time scales of ∼ 10−7 s, this term is amply negligible with respect to the gravita- tional acceleration. Consequently, the contribution due to ram pressure can be ignored within the FOA. Consequently, within our first-order approximation we can retain Eq. 1.64 as the equation describing hydrostatic equilibrium. Thus, within the FOA we see the structure can still be described throughout the set of ODEs which, for the strictly v ≡ 0 case, are known as Tolman-Oppenheimer- Volkoff equations. Thermodynamical equations. The equations for all abundances are straight- forward considering Eq. 1.41 and the FOA, DtYi = Rlocal i , i ∈ ¶1, . . . , Nspecies♢ . (1.67) 29 As expected, the term in the left-hand side encloses the contributions from changes in time at a fixed spatial location, and those due to the fluid motion in space, i.e. advection. Besides the expected coupling with the spacetime evolution of ρB and T , in the general case we expect the right-hand side to enclose interactions among all Nspecies species, hence Eqs. 1.67 are collectively referred to as the network of reactions. The largest Nspecies is, the more difficult it becomes describing in terms of individual reactions the evolution of the whole system. Associated with the network is the specific energy generation rate, Eq. 1.20, which besides all Rlocal i depends on the specific chemical potentials per unit baryon, µ̆j = µj/mu with µj the usual (relativistic) chemical potential. For massive species, i.e. those with mj ≥ mu, µj ≈ mjc 2. On the other hand, for particle such as electrons where me ≪ mu, the contribution to the chemical potential coming from the internal energy might exceed the mec 2 term, thus we must employ retain the general expression µj. From these considerations, the specific energy generation rate due to local changes in species is given by ˙̆εspecies = − Nspecies ∑ j=1 µj mu Rlocal j . (1.68) In the context of purely temporal evolution, typically within the non-relativistic regime, it is frequent to measure the impact of each reaction in the network through- out the integrated flux for the A → B reaction, namely FA→B = ∫ I dt [ ∂YA ∂t ∣ ∣ ∣ ∣ A→B − ∂YB ∂t ∣ ∣ ∣ ∣ B→A ] , (1.69) where ∂tYj♣A→B denotes the (local) contribution on the right-hand side of ∂tYj explic- itly connecting nuclides A and B. By definition, FA→B < 0 if the A → B reaction is favorable, and > 0 if the reverse one, B → A, is the one favorable. Despite the scalar character of the integrand this integral is neither covariant nor can be extended via ∂t → Dt since the advection term requires integration along the spatial direction, not the temporal one. In practice, however, we can keep Eq. 1.69 for analyzing integrated fluxes in time, at fixed locations, and for studying the behavior along the radial direction it suffices to employ a simple replacement dt → dx/v′, with x and v′ a characteristic length and speed. Now let us focus on F µ, under the assumption F θ = F ϕ = 0. Since F µuµ = 0, F 0 = vSe Λ−ΦF r, (1.70) i.e. only one component is independent. In analogy with the v ≡ 0 case, we choose to compute F r. Within the FOA (see Proposition 6), we have F r = −Ke−2Λ−Φ∂r(e ΦT ), (1.71) 30 from where it immediately follows (Proposition 7) the associated cooling rate and luminosity, still within the FOA, are Gs,T = − 1 T { 1 e2Φ+Λr2 ∂r(e 2Φ+Λr2F r) } (1.72) L = 4πr2eΛF r = −4πr2Ke−Φ−Λ∂r(e ΦT ). (1.73) Radiative temperature gradient. Combining the expression for ∂rT , Eq. 1.73, dΦ/dr from Eq. 1.60 and the hydrostatic equilibrium condition, Eq. 1.64, we obtain (Proposition 9) an expression for this gradient within the FOA, ∇r = − 3κρBLPe Λ 64πσSBT 4r2 dr dP + ( 1 − ρ H ) . (1.74) In low-density environments, H/ρ ≪ 1 so the second contribution on the right-hand side is only relevant as we approach the regime where P ≈ ρc2. Let us now write down the entropy equation within the FOA. Considering Eqs. 1.41 and 1.72, ρBTDts̆ = ρB [ ˙̆εspecies − ˙̆εν ] − { 1 4πe2Φ+Λr2 ∂r(e 2ΦL) } , where we have introduced the luminosity, Eq. 1.73, and the definition of ρB. Some authors prefer to recast this equation, either for theoretical or numerical discussions, in favor of L introducing the so-called gravitational energy ˙̆εgrav = −TDts̆, (1.75) and even labeling the temporal and radial contributions in the right-hand side with different names, non-homologous for the ∼ ∂ts̆ term, and consequently homologous for the ∼ ∂rs̆ part, ˙̆εgrav, n.h. = −Te−Φ∂ts̆ (1.76) ˙̆εgrav, h. = −Tve−Λ∂rs̆. (1.77) The origin of the gravitational nickname comes from the interpretation that changes in s̆, related to ε and P via the 1st Law of Thermodynamics Tds ≈ dε−(P/n2 B)dnB, are a consequence of altering the internal energy via compressions and expansions. The minus sign at the front of the definition ensures heat is released(absorbed) during a contraction(expansion). As it is, this equation is seldom used since Dts̆, not DtT , appears explicitly, thus demanding to account for variations of T implicitly. As shown in Proposition 8, we can invert this situation employing thermodynamical identities, thus obtaining ∂(e2ΦL) ∂r = 4πe2Φ+Λr2ρB [ ˙̆εspecies − ˙̆εν + ˙̆εgrav ] (1.78) ˙̆εgrav = −c̆P [ DtT − T∇ad P DtP ] . (1.79) 31 Here, for instance, it is possible to recover well-known expression for neutron star cooling in the case v ≡ 0, ∂tP → 0 and the degenerate regime, c̆ = c̆P = c̆V , ∂(e2ΦL) ∂r = 4πe2Φ+Λr2ρB [ ˙̆εspecies − ˙̆εν − c̆e−Φ∂tT ] . (1.80) Another characteristic we can extract from Eq. 1.78 is that all sources/sinks on the right-hand side give rise to different redshifted luminosities along the radial direction, L∞ j = 4π ∫ I dr e2Φ+Λr2ρB ˙̆εj, j ∈ ¶species, ν, grav♢ , (1.81) where I is the appropriate integration domain. These integrals allow to interpret Eq. 1.78 as a sum of contributions to the total luminosity. In summary, the theoretical modeling of neutron stars in the presence of mass accretion is governed by two sets of variables, • Structure: m, P , Φ, • Thermodynamical: T , L, Y obeying, within the FOA, the following differential equations: ∂tρB − e−Λ 4πr2 ∂rṀ ∞ = 0, Ṁ∞ = −4πr2eΦρBv, (1.82) ∂m ∂r = 4πr2ρ, dΦ dr = geΛ c2 G, (1.83) ∂P ∂r = −geΛHG, ∂T ∂r = T∇T P ∂P ∂r , (1.84) g = GmeΛ r2 , G = [ 1 + 4πPr3 mc2 ] , (1.85) H = ρ+ P c2 , ∇r = − 3κρBLPe Λ 64πσSBT 4r2 ∂r ∂P + ( 1 − ρ H ) , (1.86) DtYi = Rlocal i , i ∈ ¶1, . . . , Nspecies♢ , (1.87) ∂(e2ΦL) ∂r = 4πe2Φ+Λr2ρB [ ˙̆εspecies − ˙̆εν + ˙̆εgrav ] (1.88) ˙̆εspecies = − Nspecies ∑ j=1 µj mu Rlocal j (1.89) ˙̆εgrav = −c̆P [ DtT − T∇ad P DtP ] (1.90) ∇T =    ∇r, ∇ad > ∇r ∇conv, ∇r > ∇ad. (1.91) 32 1.4 On the numerical implementation of neutron star evolution −2 0 2 4 6 8 10 log10ρ [ g cm−3 ] 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 lo g 1 0T [K ] P i de al = P r ad P i de al = P e ,re l,d eg P id ea l = P e, no n− re l,d eg Radiation dom ain Id ea l g as do m ai n Electron domain 〈m〉i = 1.320 〈m〉e = 1.177 log10P [ erg cm−3 ] 3.00 5.88 8.76 11.64 14.52 17.40 20.28 23.16 26.04 28.92 105 106 107 108 109 1010 ρ [ g cm−3 ] 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 e% [g s − g ] Figure 1.1: Left panel: contour plots of pressure, in terms of density and tempera- ture, for a typical stellar EOS. Right panel: Relative percentual difference between the surface and local gravity acceleration, gs and g as a function of density for a realistic EOS. While Eqs. 1.82-1.91 allow to describe the whole neutron star, the numerical implementation faces a first challenge in the selection of the appropriate number of cells, both in space and time directions, due to the different scales at which each functional exhibit variations. For instance, pressure typically varies from ∼ 1014 erg cm−3 to 1025 in ∼ 10 m and with a total mass of ∼ 10−9Mstar, while between 1025 and 1036 erg cm−3 the variations in mass and radius range from 0 to ∼ Mstar and ∼ Rstar, respectively. Similarly, the characteristic timescale for heat diffusion at a layer δr, τ = (δr)2ρc̆P/K. Near the surface, τ ≤ 1 day while above and around 1026 erg cm−3 it extends from days to weeks. Therefore, for modeling neutron stars it is customary to employ a combined strategy of splitting the whole star into two regions and employing a different spatial coordinate. Core-Envelope splitting. As its name suggests, the star is divided into two regions: the core, spanning almost 90% of the total mass and radius of the star, and 33 the envelope, the thinnest and lightest region near the surface. Formally speaking, the numerical core comprises up to three different regions from an actual neutron star: the physical core at ρ ≥ 1014 g cm−3; the inner crust, between 1014 and ∼ 1011 g cm−3, characterized by the presence of free neutrons, and the outer crust, between ∼ 1011 and 108 g cm−3. Whenever needed, the actual distinction between these regions will be explicitly stated. The boundary between the envelope and the numerical core is usually determined by two aspects: • The possibility to approximate m, r and Φ as almost constants, or have rel- ative errors with respect to their surface values less than 1%. As shown in Fig. 4.1, where we illustrate the metric function Φ and the gravitational mass m (related to the radial function Λ) for two different core-EOS and the same pristine crust EOS, below 1010 g cm−3 these metric functions practically be- come constants. Similarly, in the right panel of Fig. 1.1 we can see the relative difference between the surface gravity gs = g(Rstar,Mstar) and the local value g(r,m), for a typical crust-envelope EOS, does not exceed 3% at around 1010 g cm−3. • The dependence of the EOS between temperature and pressure. Indeed, if these are independent of each other, then we can decouple the structure and thermal equations and solve them sequentially, keeping the initial structure at all times during the simulation. For example, considering the typical EOS for a star (i.e. a sum of the pressure from photons, free electrons and an ideal gas of ions), as depicted in the left panel of Fig. 1.1, we see the pressure becomes less sensitive to changes in temperature at above and around 107 g cm−3 where degenerate electrons provide the largest contribution to pressure. At large density such degeneracy is unlikely to be lifted since the Fermi temperature for neutrons and protons is far above 1010 K, out of range for typical neutron stars. Following these requirements, the density boundary between the envelope and the numerical core ρb is typically set between 107 to 1010 g cm−3. In the old days of purely isolated neutron stars (e.g. [GPE83]), however, an additional requirement for setting the boundary was to have zero heating and cooling sources in the envelope, thus allowing to keep L as strictly constant. However, this condition can be relaxed as long as the characteristic thermal timescale for these sources is comparable or inferior to the local diffusion timescale, thus allowing to approximate L as constant. Different spatial coordinates. In one dimensional, non-relativistic stellar evolution it is frequent to replace the radial coordinate in the PDEs by the rest-mass MB = muNB. Physically, such replacement is suitable as intuitively it is simpler to visualize what happens in chunks of matter than in fixed locations in space, i.e. we are allowed to handle stellar physics in terms of intensive variable (see next Chapter 34 5 10 15 log10ρ [g cm−3] 0.55 0.60 0.65 0.70 0.75 0.80 ex p [Φ (ρ )] C ru st -C or e T ra n si ti on MS-A1 + PC, ρc,15 = 1.9152 APR + PC, ρc,15 = 1.0037 5 10 15 log10ρ [g cm−3] 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 m (ρ ) [M ⊙ ] C ru st -C or e T ra n si ti on MS-A1 + PC, ρc,15 = 1.9152 APR + PC, ρc,15 = 1.0037 Figure 1.2: Metric functions in terms of local density for two stars of the indicated EOS. In the left panel we show the redshift Φ and in the right the gravitational mass m. The transition between the core and the crust is indicated by vertical lines, and their corresponding central density are stated in units of 1015 g cm−3. for a thermodynamical justification). Mathematically, this is possible due to the inverse function theorem, i.e. we can always compute MB(t, r) from r(t,MB). By changing the independent spatial variable, both fields and differential opera- tors must change accordingly. Let A be any physically relevant field (temperature, pressure, mass), x1 and x2 the variables to be exchanged, having an explicit relation x1(t, x2) and x2(t, x1). The partial derivatives of A with respect to x1, t and x2 are related via ∂A(t, x2) ∂x2 = ∂A(t, x1) ∂x1 ∂x1(t, x2) ∂x2 (1.92) ∂A(t, x2) ∂t = ∂A(t, x1) ∂t + ∂x1(t, x2) ∂t ∂A(t, x1) ∂x1 . (1.93) These transformation laws are the formal representation of what in standard fluid mechanics is said to be at fixed x1 or x2 coordinate. For instance, if x1 = r and x2 = MB, the rate of change in time for field A at fixed baryonic mass is ∂A(t,MB) ∂t = ∂A(t, r) ∂t + ∂r(t,MB) ∂t ∂A(t, r) ∂r . (1.94) For the SSS metric in particular, if A = MB and due to Eqs. 1.54 1.55, we arrive to DtMB = 0, i.e. the Dt operator we have previously introduced corresponds to the so-called total time derivative, or the rate of change in time at a fixed MB. Mathematically, such name is ill posed since time does not have a primordial role over any other variable in multi-dimensional calculus. Physically, on the other hand, it is 35 useful to stress the analysis is carried at fixed chunks of matter, i.e. at fixed MB as in Eq. 1.94, and to state its relation with respect to variations at fixed location in space. However, this is a mere particular case of the application of chain’s rule. Therefore, Eqs. 1.93 should be taken as the formal route to exchange partial derivatives when moving from one set of coordinates ¶t, x1, x2, x3, x4♢ to ¶t, x′ 1, x2, x3, x4♢ in the case x1 = x1(t, x ′ 1). Although the mass coordinate is the most usual system in stellar astrophysics, it is far from being unique: for instance, in the study of atmospheres the altitude z in a cartesian coordinate system (x, y, z) (or the radius if the system is spherically symmetrical) is typically replaced by pressure P since variations in pressure span several orders of magnitude in contrast to the little variations for both mass and altitude. Another example is the modeling of accreting systems with a given analytic expression for the net rate of mass accretion, ṀTOT(t). Here, instead of MB the relative ratio q = MB/MTOT,B(t) is employed, with MTOT,B(t) the instantaneous total mass of the system. In order to employ any of these coordinates, Eqs. 1.82-1.91 must be properly adapted. Considering the transformation laws, Eqs. 1.93, notice the major shift only takes place in the Dt operator, which under change of spatial coordinates transforms into a sum of a temporal evolution plus an advection term. For completeness, let us summarize the required changes for Eqs. 1.82-1.91 under the different coordinates so far discussed. • Baryonic mass. Since DtMB = 0, the operator Dt already represents the rate of change at fixed MB. For the equations involving ∂r, it suffices to consider the first of Eqs. 1.93 with ∂MB ∂r = 4πr2ρBe Λ. (1.95) • Pressure. While the change over the ∂r is immediate via chain’s rule, for the time operator we have Dt = e−Φ ∂ ∂t ∣ ∣ ∣ ∣ P + e−Φ∂P ∂t ∂ ∂P = e−Φ { ∂ ∂t ∣ ∣ ∣ ∣ P + Ṁ∞gHG 4πr2ρB ∂ ∂P } . (1.96) • Mass ratio q = MB/MTOT,B(t). Similarly, changes in ∂r are immediate via ∂ ∂r = ∂q ∂r ∂ ∂q = ∂MB ∂r ∂q ∂MB ∂ ∂q = 1 MTOT,B ∂MB ∂r ∂ ∂q , (1.97) 36 while for Dt we have Dt = e−Φ ∂ ∂t ∣ ∣ ∣ ∣ q + e−Φ∂q ∂t ∂ ∂q = e−Φ { ∂ ∂t ∣ ∣ ∣ ∣ q − ṀTOT ṀTOT,B ∂ ∂ ln q } . (1.98) Initial and boundary conditions. Eqs. 1.82 and 1.91 are a combination of an Initial Value and a Boundary Value Problem (IVP and BVP respectively) in [0, rs], where rs is the total radius of the star. On the one hand, equations with explicit ∂t terms such as the one for density, temperature and composition require an initial profile and to fix either one or the two values at r = 0 and r = rs. On the other hand, we have quantities such as m and P requiring only the latter boundary conditions. Within the core-envelope splitting method, we have the possibility of simplifying this IVP+BVP problem. Indeed, if the EOS is temperature-independent at the core then we can decouple pressure, mass and Φ and solve their associated PDE only once - at the beginning of the numerical simulation - and then solve the rest of the equations with m(t, r) = m(r), P (t, r) = P (r). Here we must note that if one is just interested in the local and global structure of the star - as is done in many studies contrasting theory against observations on the gravitational mass and radius - it suffices to solve the structure equations, i.e. Eqs. 1.59, 1.60 and 1.64. These are collectively referred to as Tolman-Oppenheimer-Volkoff (TOV) equations of relativistic stellar structure [ST83]. As an example of this situation, in Fig. 1.3 we show the results of integrating the TOV equations with the same envelope- EOSs and four different core-EOSs: APR [APR98] and a family of MS models [PBG+20]. The procedure is as follows: we pick a pair of values (ρc, Pc), given by the EOS, integrate the TOV equations until P ≤ 0, i.e. the numerical surface, and obtain another pair of values, the gravitational mass of the star and its radius. The procedure is repeated for all the points in the EOS table and as a result every table generates a family of neutron stars’ mass and radius. In the left panel of Fig. 1.3, we can see the gravitational mass Mgrav, i.e. m(r = rs), plotted against the stellar radius, while on the right panel we see the relation between the central densities and the corresponding gravitational mass. The four selected EOSs satisfy the different observational constraints, such as predicting the existence of stars as massive as 2M⊙, and to have stars of 1.4M⊙ with radius between 11.5 and 13 km. Concerning thermal evolution studies, we must solve for the instantaneous evo- lution of the temperature, density, chemical composition and the luminosity. At the core, it is typical to suppose matter is in equilibrium, i.e. the chemical composition is fixed at all times, while at the envelope, specifically in the presence of mass accre- tion, we must consider the evolution of the nuclides due to reactions among them. Consequently, at the core we assume X(t, r) = X(0, r) for the composition, while 37 10 12 14 Radius [km] 0.0 0.5 1.0 1.5 2.0 2.5 M g ra v [M ⊙ ] 0 1 2 3 ρc [10 15 g cm−3] 0.0 0.5 1.0 1.5 2.0 2.5 MS-A1 MS-B1 MS-C1 APR Figure 1.3: Global structure properties of neutron stars, given four different core- EOSs. In these figures, we illustrate the gravitational mass of the star as a function of the total radius in the left panel, and against their central densities in the right panel. at the envelope we must provide a distribution of chemical components and specify the composition of the accreted matter, i.e. X(t, rs). Further details concerning the composition of the envelope and the accreted material are given in Chapters 3 and 4. Regarding the temperature it is valid to start with either T (0, r) = constant, or a temperature distribution with a small gradient from the center to the top. Finally, for the luminosity we only demand L(r = 0) = 0. At the very top of the star, where the total luminosity of the object is emitted, it is suitable to employ an atmospheric boundary condition for the envelope itself, i.e. a relation of temperature and pressure with the effective temperature via the optical depth τod. From Fs = σSBT 4 eff, the actual temperature at the surface is obtained from the analytic relation Ts(Teff, τ). For simplicity, we adopt Eddington relations [Gra08, KWW12] Ts = 3 4 ( τod + 2 3 )1/4 Teff , (1.99) Ps = τodgs κ(ρs, Ts,Xs) , (1.100) which constraint the temperature and the pressure at the surface. Under the as- sumption the chemical composition Xs is known as well, we can employ the equation of state to solve for ρs or Ps. This choice of atmospheric boundary condition is not unique: for instance, another common choice is the radiative zero solution, which 38 constraints the temperature and pressure to the surface flux via T 4 s = ( 3κ 16σSBgs × 4Fs ) Ps. (1.101) For additional comments on this expression, see Appendix B.4. 1.5 Thermonuclear Instabilities Around 1928, Semenov [Sem28], [ZBLM80], [Nov16] developed a criterion for char- acterizing, in the context of chemical reactions, the stability of steady-state burning and the potential onset of thermal explosions. The underlying physics can be syn- thesized as follows: the temperature evolution is governed by the imbalance in the amount of generated energy from heating mechanisms, such as chemical and nuclear reactions, and cooling mechanisms such as escaping photons and neutrinos. If the net change is zero, we say the system is in thermal equilibrium, as illustrated by the crossing points A and B of the fictitious mechanisms at Fig. 1.4 [Lib10], where we employ an adimensional temperature τ for simplicity. Due to the dependence of these mechanisms with τ , a small perturbation in temperature induces a per- turbation from thermal equilibrium. Suppose the system was originally at A: if temperature is slightly increased, the rate of cooling exceeds that of heating, so our system comes back to its equilibrium position. If temperature is slightly decreased, the amount of heating injected overcomes the cooling mechanisms, and the system once again comes back to equilibrium. Thus, A is a stable equilibrium. A different scenario takes place at B: if temperature is increased by a small amount, the cooling mechanisms are inefficient for taking away all the energy generated by the heating mechanisms, thus system’s temperature is likely to increase almost indefinitely. On the other hand, if temperature is slightly decreased the rate of cooling is likely to enhance a drastic drop in temperature, taking the system back to equilibrium point A. Therefore, B must be considered as an unstable equilibrium. Let us now express these ideas in formal terms. Let Γ(T ) and Λ(T ) be the rates of generation and losing of energy, respectively. The temperature of the system evolves in time according to cP ∂T ∂t = Γ − Λ . (1.102) We say T0 defines an equilibrium state of the system if Γ(T0) = Λ(T0). Due to the explicit functional form of these mechanisms, in general we can expect more than one equilibrium to exist. Now we pick one of them and construct a perturbed temperature profile, T (t) = T0 + θ(t), with θ ≪ T0. Upon linearization of the ODE, we arrive to cP ∣ ∣ ∣ ∣ 0 ∂θ ∂t = [ ∂Γ ∂T ∣ ∣ ∣ ∣ 0 − ∂Λ ∂T ∣ ∣ ∣ ∣ 0 ] θ , (1.103) 39 τ ǫ∗ A B ǫ∗heat ǫ∗cool Figure 1.4: Semi-realistic rates of cooling (blue line) and heating (red line) for a chemical system as functions of adimensional temperature. The intersections of these functions define steady-states. Based on [Lib10] where ♣0 indicates the correspondent functional is evaluated at T = T0. As the gen- eral solution to this ODE is θ ∝ eλt, and cP,0 is a positive quantity (see Appendix A), a stationary state T0 is stable against small perturbations if [ ∂Γ ∂T ∣ ∣ ∣ ∣ 0 − ∂Λ ∂T ∣ ∣ ∣ ∣ 0 ] < 0, (1.104) and we can refer to such T0 as a steady state. In case the opposite inequality holds true, we say such equilibrium is unstable. Furthermore, the nature of Γ defines which kind of thermal explosion we have: for instance, if heating is due to chemical reactions, we speak of thermo-chemical explosions, while if coming from nuclear reactions as in the case of astrophysical scenarios, such as neutron star in binary systems, these are labeled as thermonuclear explosions. Although Semenov’s model is quite useful to understand the physics of thermal instabilities, the absence of spacial-dependence prevents us to apply it for systems where dissipative fluxes are present, as is the case in stellar evolution with ∇ · F . A first step to remedy this issue was discussed in later works by Frank-Kamenetskyy [FK38], [ZBLM80], who studied the stability of the following PDE: cP ∂T ∂t = K∇2T + Γ(T ) , (1.105) where K is regarded as a constant. Following a similar perturbative analysis, to linear order we obtain cP ∣ ∣ ∣ ∣ 0 ∂θ ∂t = K∇2θ + [ ∂Γ ∂T ∣ ∣ ∣ ∣ 0 ] θ, (1.106) i.e. a Sturm-Liouville problem which can be solved by proposing a series of the form θ(t, r) = ∞ ∑ n=0 Cne −λntψ(r) . (1.107) 40 An instability in this context is essentially the same as in the Ineq. 1.104 case. Its identification, however, is now slightly more involved: as the eigenvalues λn for a Sturm-Liouville problem are real and ordered with n, the identification of the eigenmode λ = 0 must be complemented with the verification that it corresponds to n = 0. If this second aspect is ignored, there might exists a negative eigenmode which will render T0 as truly unstable [Kor79], [BFG83a]. In the astrophysical context, the constancy of K is objectionable by recalling that K = (16σ/3)T 3ρ−1κ−1, i.e. it is an implicit function of position via T and ρ. As pointed out in the context of chemical reactions [BFG83b], including this func- tional dependence is important as variations in the critical parameters for thermal explosions are significant. In principle, keeping the dependency of K with r for the linear perturbation analysis leads to a second-order PDE in θ, whose eigenmodes and eigenfunctions might be extracted by similar steps as in the K ≡ constant case. Surprisingly, the linear perturbation analysis for the surface of neutron stars can relatively overcomes this difficulty by imposing some locality over ∇ · F , i.e. ∇ · F ≈ KT d2 H ≡ εcool(T ) (1.108) with dH ≈ P/(ρgs), and performing the standard perturbation analysis to first order leading to Eq. 1.104, in the form ∂εheat ∂T > ∂εcool ∂T . (1.109) This criterion for instability practically became a standard in the field, simplifying the process from computing eigenvalues and eigenvectors (as attempted by some authors, e.g. ) to computing numerical or analytical partial derivatives. The ca- pacity of this approximation to predict the critical density and temperature for the beginning of thermonuclear explosion, provided stationary solutions have been calculated, is relatively good for a first comparison with fully time-dependent sim- ulations. The caveats, however, have become visible in the modern era where fast fully-time dependent simulations are available for desktop computers. A slightly outdated, yet complementary approach to the imbalance in perturba- tions to the heating and cooling mechanisms is the comparison of local timescales. The physical picture is simple: if the amount of time required to heat and cool a bulk of matter under certain conditions of temperature and density is the same, then the system will stay in thermal equilibrium. However, if the cooling time exceeds the heating one then an instability will take place. Given thermonuclear insta- bilities occur in the presence of mass accretion, for this approach three scales are contrasted: the diffusion timescale τdif = (δr)2ρc̆P/K, associated with the radiative and conductive fluxes; the heating timescale, principally due to nuclear reactions and given by τnuc = Enuc/εnuc, i.e. the ratio between the energy per gram and the specific energy generation rate, and finally the accretion timescale τacc = δM/Ṁ , 41 measuring the amount of time required to accrete a layer of δM at an accretion rate of Ṁ . Stability is guaranteed when τdif < τnuc and τnuc < τacc, a criterion which can also be deduced from Ineq. 1.109 by order-of-magnitude estimations. Such enforced locality, however, reduces the range of applicability of this criterion, although it is useful for a back-of-the-envelope estimation. 42 Chapter 2 Microphysical aspects of stellar evolution “Can’t repeat the past?...Why of course you can!” F. Scott Fitzgerald Abstract. In this chapter we present the characteristics which a nuclear network of reactions must have in order to describe the envelope of the star. 2.1 Introduction In the previous chapter we encountered several functionals, i.e. functions of func- tions, representing physical processes such as the rate of interactions among parti- cles or the opacity of a system. These functionals share the property of locality, i.e. not depending on derivatives of any kind or order (partial derivatives, 4-gradients, D’Alembertian, etc) of their functional arguments, a set of finite thermodynamical degrees of freedom such as nB, T , and ¶Yi♢Nspecies i=1 . From a strict point of view, their construction is based on a statistical approximation to the underlying processes: for instance, instead of following the instantaneous movement of each and every con- stituent particle of the system, we compute averages over a well-posed distribution of energies or velocities. As a consequence, this collection of functionals is referred to as microphysics. It is the purpose of the present chapter to provide an exposition on the enclosed physics by each of these microphysical functionals. Given the rich vastness of tech- niques behind their construction, as well as the extensive literature for each of the topics, it is our scope to provide the essential features as to allow an understanding of the following chapters. 43 In section II we discuss the fundamental thermodynamical degrees of freedom governing all the functionals discussed in the chapter. In section III we briefly discuss the neutron star EOS at different densities, focusing on the opacity at section IV, followed by the neutrino emission mechanisms at section V, and closing the chapter with the extensive subject of nuclear reactions. 2.2 A brief survey on Thermodynamics Thermodynamics is the branch of physics concerned with the study of systems Σ(T ) with ≥ 1023 particles through a collection of variables ¶ηi♢Nd.o.f i=1 , referred to as Degrees of Freedom (DoF), as well as the functionals of these variables, such as the total energy U or the EOS. The DoF can be classified according to their behavior under the addition or remotion of thermodynamical subsystems: let Σ (T ) 3 = Σ (T ) 1 ∪ Σ (T ) 2 . If η (3) j = η (1) j +η (2) j for some j ∈ ¶1, Nd.o.f.♢, we say this variable is extensive. Otherwise, we define it as intensive. Examples of extensive variables are the baryonic mass, number of particles, volume V , U or entropy S, while for intensive variables the temperature T , pressure P or chemical potential µ are good examples. Related to the concept of extensive variables is that of homogeneous function of k-degree, which in general is any differentiable f : R n → R, with k ∈ R and λ ∈ R +, satisfying f(λx) = λkf(x). In thermodynamical systems this property is crucial for our conception of functionals of state as it suggests the existence of a certain parameter for scaling all the extensive variables. Intuitively, this parameter can be the total number of particles or the volume. The mathematical foundation of Thermodynamics lies in Statistical Physics, which is the bridge between the microscopic predictions of Quantum Mechanics or Field Theory and the macroscopic observables, the DoF. An important question is, to what extent are mathematically consistent the predictions of these statistical theories? Quite frequently we talk about systems with 1023 particles or very large in volume. Recalling the old criticisms to differential calculus’ methods, two im- portant questions emerge here: Why 1023 particles?, and even more importantly, how large is a large system? The first question seems to possess a rather simple answer, as we learn in basic Chemistry that the Avogadro’s number NA, which in numerical terms is the inverse of the atomic mass unit mu ∼ 10−24 g, represents a useful rod to quantify the amount of particles in a system. As for the second question, by following our most basic mathematical intuition we might be tempted to claim that predictions are valid as long as we take limN,V→∞ for all thermody- namical variables. Turns out this is just half the answer: while intensive properties are indifferent to this limit, extensive ones actually depend on it as this classifica- tion precisely follows from their proportionality with N , hence if the limit is simply taken we might obtain infinite quantities! To precise these notions, and to make Statistical Physics a well-posed mathematical theory, the postulate of thermody- 44 namical limit is introduced [Dor21, Kuz14, Sty04], allowing to extend our formalism even to “trivial” problems such as the incorporation of Coulomb interactions [LL69]. In simple words, the thermodynamical limit postulates predictions from Statistical Physics are meaningful only if the following conditions hold: 1. The particle number density n := lim (N,V )→(∞.∞) N V (2.1) exists and is finite. 2. For extensive variables such as X(T, V ) or X(T,N), there exist associated functions of states referred to as X per particle and X-density (X per unit volume), respectively, satisfying x̃(T, n) := lim (N,V )→(∞.∞) X(T, V ) N (2.2) x(T, n) := lim (N,V )→(∞.∞) X(T,N) V . (2.3) In the astrophysical contexts, these postulates acquire uttermost importance since it is frequent to employ either specific quantities (i.e. X per gram of matter) x̆, X per particle x̃ = mux̆, or X-densities (i.e. per volume) x = nx̃, taking N ≡ NB and n ≡ nB due to the conservation of baryons and since the proton mp and neutron mn masses, close to mu, exceeds by almost 2000 times the electron’s me. As a consequence of this choice, for rewriting thermodynamical identities, and only within that context, it is valid to consider the replacements V ≡ VB = NB/nB or its close relative V ≡ Vm = MB/ρB with MB = muNB. Thermodynamical functionals. As described in standard Statistical Mechan- ics textbooks [LL76], the central functional in the study of an ensemble of particles whose dynamics is governed by the Hamilton operator Ĥ and which is allowed to exchange particles and energy with a thermal reservoir is the grand-partition func- tion Z. Assuming the system is composed of Nspecies species, with individual particle number operator N̂i satisfying [Ĥ, N̂i] = 0̂, this functional is given by Z = Tr    exp  −β  Ĥ − Nspecies ∑ i=1 µiN̂i        (2.4) with β = (kBT )−1 and µi is the chemical potential of the i-th species. If the eigenstates of Ĥ are known, the trace simplifies into a sum for the discrete case, or into an integral for the continuum one. Through partial differentiation of Z, we 45 obtain the expectation value for the number of particles, energy and entropy as1 Ni = 1 β ∂lnZ ∂µi , i ∈ ¶1, . . . , Ns♢ (2.5) U = −∂lnZ ∂β + µiNi (2.6) S = kB ∂ (T lnZ) ∂T . (2.7) If U is an homogeneous function of degree k = 1, as is the case in the majority of astrophysical applications, we can obtain the pressure from Z via P = −kBT V ln Z. (2.8) Formally speaking, the quantities on the left should be denoted as ⟨X⟩ since these are thermal expectation values for a given operator. However, we opt for the traditional approach of writing them as X, stressing on the fact that for the systems of interest in this work the system 2.5 - 2.7 can be written in the form X = ∫ dE F(E)hX(E) , (2.9) where dE denotes the appropriate measure of integration, hX is a specific function associated with the variable X and F is the distribution function, which is inherent to the physical system and thus common to Ni, U and S. By construction, Eqs. 2.5 - 2.7 satisfy the First Law of Thermodynamics, in terms of specific quantities given by dε̆ = Tds̆+ P ρ2 B dρB + Nspecies ∑ i=1 µi mu dYi, (2.10) which is a statement of how variations in s̆, ρB and the abundances Yi change of the total specific energy of the system, and the Second and Third Laws of Thermody- namics, ds̆ ≥ 0 (2.11) lim T→0 ∆s̆ = 0. (2.12) The first can be interpreted as the mathematically formal statement of the every- day situation that heat flows from hot to cold bodies. Taking into account the statistical foundation of Thermodynamics, however, it is equally plausible to con- sider the Second Law as the formal statement that systems will evolve from low- to 1For systems such as a photon gas, its appropriate description is obtained via Nspecies = 1 and µ → 0 in the expressions for N , U and S. 46 high-probability configurations. The second statement, also referred to as Nernst’s Postulate, is a requirement emerging from quantum mechanics. Eq. 2.10, the differential of the specific energy, also encloses the functional nature of ε̆, namely, ε̆(s̆, ρB,Y ), and states the intensive variables T , P and µi can be obtained via first-order partials of ε̆, T = ( ∂ε̆ ∂s̆ ) ρB,Y , P = ρ2 B ( ∂ε̆ ∂ρB ) s̆,Y , µi = mu ( ∂ε̆ ∂Yi ) s̆,ρB,Yk ̸=i . (2.13) It is important to stress that, given the functional dependence of ε̆ (as well as the Implicit/Inverse function theorems from standard multivariable calculus), we can expect the rest of thermodynamical variables here written to have analogous func- tional dependence, for instance P (s̆, ρB,Y ), or even to allow inversion in terms of a measurable quantity, such as ρB(s̆, P,Y ) or T (s̆, ρB,Y ) → s̆(T, ρB,Y ) → s̆(T, P,Y ). Similarly to the case for ε̆, the partial derivatives of functionals such as s̆ and P have well-defined roles in astrophysical discussions. For instance, to quantify how much heat Σ(T ) exchanges under a variation of temperature at constant η, we introduce specific heat at constant η as c̆η = T ( ∂s̆ ∂T ) η,Y . (2.14) By virtue of the Second Law, c̆η ≥ 0, a mathematical statement of the everyday experience that providing heating to a Σ(T ) increases its temperature, while removing it leads to a decrement in T . The most frequently used specific heats are those at constant volume and pressure, c̆V 2 and c̆P respectively. By virtue of the Second and Third Laws, these satisfy c̆P - c̆V ≥ 0 and c̆P ≈ c̆V in the vicinity of T = 0, a condition of extensive usage in degenerate systems such as the core and crust of Neutron Stars. Considering P (T, ρB,Y ), its adimensional partial derivatives functionals - or sensitivities - are [Cla68, HK94, JSB+21, AJL+22]3, χρ = ρB P ( ∂P ∂ρB ) T,Y , χT = T P ( ∂P ∂T ) ρB,Y , χYi = Yi P ( ∂P ∂Yi ) T,ρB,Yk ̸=i . (2.15) By definition, these functionals serve to give a qualitative measure of the EOS sen- sitivity with each DoF present at its differential, as under a small neighborhood around a fixed tuple (Tx, ρB,x,Yx) we have Px = P0T χT x ρ χρ B,x ∏N s=1 Y χYs s with P0 some constant. Due to the functional form of each coefficient, for improving numerical 2Properly speaking, it is at constant nB. However, we attach to the most common notation considering the constancy of NB in the relation V ≡ VB. 3Kippenhann’s ([KWW12]) notation for these derivatives, i.e. α = 1/χρ and δ = χT /χρ, represent an exception rather than a rule. 47 accuracy it is sometimes more convenient to express them employing the notation (d ln Y/d lnX)Z1,...,Zs . As shown in Appendix A, these adimensional χ can be em- ployed to express the specific heats avoiding an explicit dependence with s̆ c̆V = ( ∂ε̆ ∂T ) ρB,Y , (2.16) c̆P = ( ∂ε̆ ∂T ) P,Y + χT χρ P ρBT (2.17) c̆P − c̆V = χT χ2 ρ P ρBT . (2.18) The adiabatic gradient. A coefficient of interest mixing both P and s̆ is the adiabatic temperature gradient, expressing how much temperature change with respect to pressure in the absence of heat exchange, ∇ad = P T ( ∂T ∂P ) s̆,Y = χT χρ P Tρc̆P . (2.19) This quantity plays a central role in the study of convection, i.e. heat transport via the movement of matter. Intuitively, this is to be expected since ∇ad physically states how much can temperature be changed under a perturbation in the structure, i.e. pressure, without allowing such perturbation to absorb or provide heat. It can be shown, via stability analysis in 1D, that a sufficient condition for the onset of convection is [KWW12, HK94] ∇ad < ∇T , (2.20) with ∇T = (P/T ) ∂T ∂P the actual temperature gradient. This is the celebrated Schwarzschild’s criterion, which intuitively states that the maximum temperature gradient a blob can reach before starting its motion is that where no heating ex- change takes place. Since Schwarzschild’s criterion assumes composition is fixed, it is only a limit case of Ledoux’s criterion, which states that convective effects allowing mixing of material takes place when [HK94, BM08, GNMM14] ∇ad + χρϕ χT ∇µ,s < ∇T , (2.21) with χρϕ χT ∇µ,s the gradient of mass fraction with respect to pressure at fixed entropy. For systems such as stellar interiors, subject to gravitational acceleration and consequently to pressure gradients, convection is typically assumed to predominantly operate along the direction of maximum growth of pressure, i.e along the radial direction. This also responds to our everyday experiences at Earth (for instance, in boiling water) where mass blobs move in the vertical direction, from the heated base to the surface. While pressure gradients in the horizontal direction have a role 48 for the size of the blob and in current 3D models, employing Ineq. 2.20 and in single-dimensional simulations of stars is still a reliable approach, in combination with models computing the energy flux due to convection and an appropriate ∇T [PBD+11, GNMM14, AJL+22, JT23]. The chemical potential. As is well-known, a typical Ĥ consists on contri- butions from a kinetic K̂ and a potential energy V̂ operators. Regarding the first, in momentum space it is possible to expand it as ∑Nspecies i=1 [mic 2 + p2 i /2mi + O(p4 i )], where the first term in the series is referred to as the “rest-mass” energy. Retaining all O(p2n i ) terms in the expansion, it is possible to write ε̆ = ε̆rest + ε̆int, the first term given by ε̆rest = Nspecies ∑ i=1 mic 2Yi mu . (2.22) The second term, the “internal energy” ε̆int, encloses the contributions to energy given by the remaining terms in the series of p and the contributions from potential energy. As a consequence, by explicit calculation we see the chemical potential for the j-th species is µj = mjc 2  1 + mu mjc2 ( ∂ε̆int ∂Yj ) s̆,ρ̆B,Yk ̸=j   , (2.23) with the leading contribution coming from particle’s rest-mass, which introduces the notion that µj is the relativistic chemical potential while contribution due to changes in ε̆ represent the non-relativistic chemical potential. Due to its origin as thermal average of terms ∝ p2n, we can get further insight of its importance according to the relativistic status of the underlying particles. For instance, for particles obeying a Boltzmann distribution we have mu∂ε̆int/∂Yj = 3kBT/2, i.e. chemical potential acquires a contribution from the internal energy as long as kBT ∼ mjc 2, requiring large temperatures as long as mj ≈ mu. For these species, µj ≈ mjc 2. On the other hand, for particles obeying the Fermi-Dirac distribution, such as e−/e+, the contribution from the internal energy becomes prominent due to the smallness of me at different densities and temperatures. Consequently, both rest-mass and internal contributions must be included in the actual computation of µe. Thermodynamical potentials. Quite often, in both theoretical and experi- mental sides, we have systems where intensive variables, such as P or T , can be regarded as independent instead of ρB or s̆. Furthermore, in some theoretical and experimental cases ε̆ might not be known, a requisite in principle we must have in order to construct, via the 1st Law of Thermodynamics, quantities such as P , s̆ and c̆P . To solve this problem, we employ Legendre Transformations to construct adequate thermodynamical potentials regarding intensive variables as conjugated of the extensive ones. Once constructed, we can extract the functionals of interest such as P , s̆ or c̆P from their first and second derivatives, constituting a set of intercon- 49 nected relations referred to as the Maxwell relations. For a full list of them, we refer the reader to Appendix A. Given the set of variables s̆, ρB and Y , the Legendre transformations for ε̆ are: • Helmholtz specific free energy f̆(ρB, T,Y ): defined as f̆ = ε̆ − T s̆, its differ- ential satisfies df̆ = P ρ2 B dρB − s̆dT + Nspecies ∑ i=1 µi mu dYi. (2.24) By construction, this potential is suitable for systems where ρB and T are accessible for measurements. In the context of standard, non-relativistic stel- lar evolution, this potential is of frequent use as it is easy to compute from first principles, via the standard partition function Z := Tr { exp[−βĤ] } , and to obtain the remaining functionals such as P , s̆ and c̆P via first and sec- ond derivatives. Due to its definition, this potential is closely related to the maximum amount of work we can extract from Σ(T ). • Specific enthalpy h̆(s̆, P,Y ): defined as h̆ = ε̆+ P/ρB. Correspondingly, dh̆ = Tds̆+ 1 ρB dP + Nspecies ∑ i=1 µi mu dYi. (2.25) If the pressure of Σ(T ) is kept constant, as for instance in the case of solids, ∆h̆ = ∫ dh̆ expresses the total amount of heat injected or removed from the system. This shows the origin of its old name, “heat function”. • Gibbs specific free energy ğ(P, T,Y ): defined by ğ = ε̆ + P/ρB − T s̆, its differential is dğ = 1 ρB dP − s̆dT + Nspecies ∑ i=1 µi mu dYi, (2.26) i.e. this potential is adequate for studying processes at constant temperature or pressure. Additionally, for homogeneous U of degree k = 1, ğ can be shown equal to ∑Nspecies i=1 µiYi/mu, i.e. the Gibbs specific energy is a pondered sum of the amount of invested energy to add/remove the Nspecies species. 2.3 The Neutron Star Equation of State By simple order-of-magnitude estimations, at the core of a typical neutron star we expect a large density, ∼ 1014−15 g cm−3, contrasted against stars like our own Sun, with central densities of ∼ 102 g cm−3. From a nuclear physics point of view, neutron stars’ core is the perfect laboratory to put into test different models of 50 nuclear matter and their interactions since the core’s density is actually comparable to the nuclear saturation density, e.g. ρ0 = 2.5 × 1014 g cm−3, above and around we could simply have “ordinary” particles, such as protons, neutrons, electrons and muons, or exotic forms of matter such as de-confined quarks. This, on the other hand, brings certain uncertainty as to which model we should opt for to describe the inner portion of the neutron star. Additionally, since the surface of the star should have densities around 100 g cm−3 we must have phase transitions, from the “ordinary” stellar matter - gas like with Coulomb interactions as corrections - to the center of the star. Fortunately for the latter scenario, the large densities and the chemical composition4 allow to remove temperature from the thermodynamical degrees of freedom as it suffices to consider matter at T = 0 K, i.e. in the ground state. So far, it is difficult to speak of a “unified” equation of state, i.e. a single functional P (ρ, T,X) capable of describing the whole interior of the star. Some efforts into that direction actually exist, although it is customary as well to have the EOS in tabulated form and to implement interpolation schemes for incorporating them into stellar evolution codes. The main difficulty, we can say, is the absence of a unified model of matter behavior at all densities and temperatures. Additionally, we can cite the success of dividing the star into layers, get the EOS for each of them and then join the resulting functional/table. From a qualitative point of view, this latter path is better than the first as we can actually comprehend which are the underlying processes and the uncertainties in each of these models. Despite the different names the layers of the star receive in the literature, we must be aware there exists a simple criterion to make the distinction between one region and another: the kind of inhabiting particles and their interactions. Small digression: what kind of particles do we expect? Considering we are concerned with the study of neutron stars, it is pertinent to make a short glossary on their constituent particles in order to avoid a constant repetition of their intrinsic properties. By element we define any bounded array of Z protons, Z electrons and N neu- trons named after their Z number. For instance, a system with Z = 6 is referred to as carbon or with Z = 26 as iron, regardless of having N = 6 or N = 30. Each element, as is well known, has its on symbol: C for carbon, Fe for iron, and so on. Since clustered protons and neutrons provide the largest contribution to rest-mass and they are bounded by the nuclear force, they are collectively referred to as the nu- cleus. An isotope is an element with different number of neutrons, and are denoted as either A(Element name) or (Element name)-A, with A = Z + N . For instance, we have carbon-13 or 55Fe. By definition, an element is electrically neutral: when 4It is customary to say “chemical composition”, despite the particles under consideration have little relation with everyday chemistry. 51 107Te 106Sb Valley of stability 99Sn 100Sn 101Sn 102Sn 103Sn 104Sn 105Sn 98In 99In 100In 101In 102In 103In 104In 95Cd 96Cd 97Cd 98Cd 99Cd 100Cd 101Cd 102Cd 103Cd nuclides 94Ag 95Ag 96Ag 97Ag 98Ag 99Ag 100Ag 101Ag 102Ag 90Pd 91Pd 92Pd 93Pd 94Pd 95Pd 96Pd 97Pd 98Pd 99Pd 100Pd 101Pd 89Rh 90Rh 91Rh 92Rh 93Rh 94Rh 95Rh 96Rh 97Rh 98Rh 99Rh 100Rh nuclides 86Ru 87Ru 88Ru 89Ru 90Ru 91Ru 92Ru 93Ru 94Ru 95Ru 96Ru 97Ru 98Ru 99Ru 85Tc 86Tc 87Tc 88Tc 89Tc 90Tc 91Tc 92Tc 93Tc 94Tc 95Tc 96Tc 97Tc 98Tc 82Mo 83Mo 84Mo 85Mo 86Mo 87Mo 88Mo 89Mo 90Mo 91Mo 92Mo 93Mo 94Mo 95Mo 96Mo 97Mo 81Nb 82Nb 83Nb 84Nb 85Nb 86Nb 87Nb 88Nb 89Nb 90Nb 91Nb 92Nb 93Nb 94Nb 95Nb 96Nb 78Zr 79Zr 80Zr 81Zr 82Zr 83Zr 84Zr 85Zr 86Zr 87Zr 88Zr 89Zr 90Zr 91Zr 92Zr 93Zr 94Zr 95Zr 77Y 78Y 79Y 80Y 81Y 82Y 83Y 84Y 85Y 86Y 87Y 88Y 89Y 90Y 91Y 92Y 93Y 94Y 74Sr 75Sr 76Sr 77Sr 78Sr 79Sr 80Sr 81Sr 82Sr 83Sr 84Sr 85Sr 86Sr 87Sr 88Sr 89Sr 90Sr 91Sr 92Sr 93Sr 73Rb 74Rb 75Rb 76Rb 77Rb 78Rb 79Rb 80Rb 81Rb 82Rb 83Rb 84Rb 85Rb 86Rb 87Rb 88Rb 89Rb 90Rb 91Rb 92Rb 69Kr 70Kr 71Kr 72Kr 73Kr 74Kr 75Kr 76Kr 77Kr 78Kr 79Kr 80Kr 81Kr 82Kr 83Kr 84Kr 85Kr 86Kr 87Kr 88Kr 89Kr 90Kr 91Kr 68Br 69Br 70Br 71Br 72Br 73Br 74Br 75Br 76Br 77Br 78Br 79Br 80Br 81Br 82Br 83Br 84Br 85Br 86Br 87Br 88Br 89Br 90Br 65Se 66Se 67Se 68Se 69Se 70Se 71Se 72Se 73Se 74Se 75Se 76Se 77Se 78Se 79Se 80Se 81Se 82Se 83Se 84Se 85Se 86Se 87Se 88Se 89Se 64As 65As 66As 67As 68As 69As 70As 71As 72As 73As 74As 75As 76As 77As 78As 79As 80As 81As 82As 83As 84As 85As 86As 87As 88As 60Ge 61Ge 62Ge 63Ge 64Ge 65Ge 66Ge 67Ge 68Ge 69Ge 70Ge 71Ge 72Ge 73Ge 74Ge 75Ge 76Ge 77Ge 78Ge 79Ge 80Ge 81Ge 82Ge 83Ge 84Ge 85Ge 86Ge 87Ge 59Ga 60Ga 61Ga 62Ga 63Ga 64Ga 65Ga 66Ga 67Ga 68Ga 69Ga 70Ga 71Ga 72Ga 73Ga 74Ga 75Ga 76Ga 77Ga 78Ga 79Ga 80Ga 81Ga 82Ga 83Ga 84Ga 85Ga 86Ga 55Zn 56Zn 57Zn 58Zn 59Zn 60Zn 61Zn 62Zn 63Zn 64Zn 65Zn 66Zn 67Zn 68Zn 69Zn 70Zn 71Zn 72Zn 73Zn 74Zn 75Zn 76Zn 77Zn 78Zn 79Zn 80Zn 81Zn 82Zn 83Zn 84Zn 54Cu 55Cu 56Cu 57Cu 58Cu 59Cu 60Cu 61Cu 62Cu 63Cu 64Cu 65Cu 66Cu 67Cu 68Cu 69Cu 70Cu 71Cu 72Cu 73Cu 74Cu 75Cu 76Cu 77Cu 78Cu 79Cu 80Cu 81Cu 82Cu 50Ni 51Ni 52Ni 53Ni 54Ni 55Ni 56Ni 57Ni 58Ni 59Ni 60Ni 61Ni 62Ni 63Ni 64Ni 65Ni 66Ni 67Ni 68Ni 69Ni 70Ni 71Ni 72Ni 73Ni 74Ni 75Ni 76Ni 77Ni 78Ni 79Ni 80Ni 50Co 51Co 52Co 53Co 54Co 55Co 56Co 57Co 58Co 59Co 60Co 61Co 62Co 63Co 64Co 65Co 66Co 67Co 68Co 69Co 70Co 71Co 72Co 73Co 74Co 75Co 77Co 47Fe 48Fe 49Fe 50Fe 51Fe 52Fe 53Fe 54Fe 55Fe 56Fe 57Fe 58Fe 59Fe 60Fe 61Fe 62Fe 63Fe 64Fe 65Fe 66Fe 67Fe 68Fe 69Fe 70Fe 71Fe 72Fe 74Fe 46Mn 47Mn 48Mn 49Mn 50Mn 51Mn 52Mn 53Mn 54Mn 55Mn 56Mn 57Mn 58Mn 59Mn 60Mn 61Mn 62Mn 63Mn 64Mn 65Mn 66Mn 67Mn 68Mn 69Mn 70Mn 44Cr 45Cr 46Cr 47Cr 48Cr 49Cr 50Cr 51Cr 52Cr 53Cr 54Cr 55Cr 56Cr 57Cr 58Cr 59Cr 60Cr 61Cr 62Cr 63Cr 64Cr 65Cr 66Cr 43V 44V 45V 46V 47V 48V 49V 50V 51V 52V 53V 54V 55V 56V 57V 58V 59V 60V 61V 62V 63V 64V 40Ti 41Ti 42Ti 43Ti 44Ti 45Ti 46Ti 47Ti 48Ti 49Ti 50Ti 51Ti 52Ti 53Ti 54Ti 55Ti 56Ti 57Ti 58Ti 59Ti 60Ti 61Ti 39Sc 40Sc 41Sc 42Sc 43Sc 44Sc 45Sc 46Sc 47Sc 48Sc 49Sc 50Sc 51Sc 52Sc 53Sc 54Sc 55Sc 56Sc 57Sc 58Sc 35Ca 36Ca 37Ca 38Ca 39Ca 40Ca 41Ca 42Ca 43Ca 44Ca 45Ca 46Ca 47Ca 48Ca 49Ca 50Ca 51Ca 52Ca 53Ca 54Ca 55Ca 56Ca 35K 36K 37K 38K 39K 40K 41K 42K 43K 44K 45K 46K 47K 48K 49K 50K 51K 52K 53K 54K 31Ar 32Ar 33Ar 34Ar 35Ar 36Ar 37Ar 38Ar 39Ar 40Ar 41Ar 42Ar 43Ar 44Ar 45Ar 46Ar 47Ar 48Ar 49Ar 50Ar 30Cl 31Cl 32Cl 33Cl 34Cl 35Cl 36Cl 37Cl 38Cl 39Cl 40Cl 41Cl 42Cl 43Cl 44Cl 45Cl 46Cl 47Cl 27S 28S 29S 30S 31S 32S 33S 34S 35S 36S 37S 38S 39S 40S 41S 42S 43S 44S 45S 46S 26P 27P 28P 29P 30P 31P 32P 33P 34P 35P 36P 37P 38P 39P 40P 41P 42P 43P 44P 45P 22Si 23Si 24Si 25Si 26Si 27Si 28Si 29Si 30Si 31Si 32Si 33Si 34Si 35Si 36Si 37Si 38Si 39Si 40Si 41Si 42Si 43Si 22Al 23Al 24Al 25Al 26Al 27Al 28Al 29Al 30Al 31Al 32Al 33Al 34Al 35Al 36Al 37Al 38Al 39Al 40Al 41Al 20Mg 21Mg 22Mg 23Mg 24Mg 25Mg 26Mg 27Mg 28Mg 29Mg 30Mg 31Mg 32Mg 33Mg 34Mg 35Mg 36Mg 37Mg 38Mg 20Na 21Na 22Na 23Na 24Na 25Na 26Na 27Na 28Na 29Na 30Na 31Na 32Na 33Na 34Na 35Na 17Ne 18Ne 19Ne 20Ne 21Ne 22Ne 23Ne 24Ne 25Ne 26Ne 27Ne 28Ne 29Ne 30Ne 31Ne 32Ne 33Ne 17F 18F 19F 20F 21F 22F 23F 24F 25F 26F 27F 28F 29F 30F 13O 14O 15O 16O 17O 18O 19O 20O 21O 22O 23O 24O 25O 26O 12N 13N 14N 15N 16N 17N 18N 19N 20N 21N 22N 23N 24N 9C 10C 11C 12C 13C 14C 15C 16C 17C 18C 19C 20C 21C 22C 8B 9B 10B 11B 12B 13B 14B 15B 16B 17B 18B 19B 20B 21B 7Be 8Be 9Be 10Be 11Be 12Be 13Be 14Be 15Be 16Be 6Li 7Li 8Li 9Li 10Li 11Li 3He 4He 5He 6He 7He 8He 9He 10He 1H 2H 3H 4H 5H 6H 7H N NEUTRON RICH P R O T O N R I C H 1H 4He N Tz = − 1 α Figure 2.1: Nuclide chart one or more electrons are removed while keeping the same amount of protons and neutrons, we speak of ions and their degree of ionization. For the densities and tem- peratures of interest, we can assume all elements are fully ionized, i.e. the net charge is positive and equal to Z. Finally, by nuclide we understand any bounded array of Z protons, Z electrons and N neutrons, i.e. A = Z +N baryons, and denote them as A(Element name). For instance, 12C, 56Fe or 80Kr. In theory, the only difference with the concept of elements is the prominent role the number of neutrons acquire: for nuclides it is necessary to specify it, while for elements the number of protons is enough. A visual classification of nuclides is made by constructing nuclide charts, such as the one illustrated in Fig. 2.1. Along the vertical axis the number of pro- tons increase and along the horizontal axis the number of neutrons. The grey band at the middle is named as valley of stability since at Earth these nuclides do not undergo β+/β− decays or have estimated lifetimes exceeding thousands of years. In addition, these are the final states of nuclides at both proton- and neutron-rich sides undergoing β+/β− decays. Given an α particle is a bounded state of two protons and two neutrons (i.e. the nucleus of an 4He), nuclides having 2Z protons and 2N neutrons are classified as α nuclides. Species with (N − Z)/2 = −1 are dubbed as Tz = −1 and we remark their presence due to their involvement in the rp-process. While protons and neutrons conform the baryonic sector of the composition, we still have some words regarding the leptonic sector. Besides electrons and positrons, neutrinos and anti-neutrinos play a key role in the evolution of the neutron star. However, once the supernovae phase has finished and the proto-neutron star is born, 52 matter is transparent for them and they practically can escape from the interior, carrying away energy in the process. As such, they are not considered into the construction of the equation of state due to their practically null interaction with the rest of particles. At high densities, in order to have equilibrium the presence of muons is required, leading to the so-called npeµ matter which conforms the most accepted model for neutron star composition at high density. The presence of hy- perons, while theoretically viable, is observationally dubious at the moment given the irresolution of the hyperon puzzle, consisting on the softening of the EOS (i.e. a decreased dP/dρ) by the inclusion of hyperons, which reduces the maximum grav- itational mass a star can reach to Mmax < 2M⊙, in disagreement with existent inferences of neutron stars with masses of ∼ 2.08M⊙. From Quantum Chromodynamics (QCD) and its experimental verifications, we know protons and neutrons, contrasted against electrons, muons or neutrinos, are not fundamental particles but have structure on their own, i.e. quarks and gluons. Although their de-confinement from the neutrons and protons is a possible state of matter at ρ ≥ ρ0, from current astrophysical observations the “ordinary” matter at the core is still preferred, so in the rest of this work we shall not consider the possibility of quark matter. For theoretical purposes, we separate the neutron star layers at ρ ≤ 1011 g cm−3 into different regions, ordered according to the local density: • Atmosphere. Occupies the region ρ ≤ 10−3 g cm−3, where nuclides can be in different stages of ionization and where radiation escapes the star. As described in Chapter 1, typically it is approximated as a functional relation between the variables of interest (temperature, pressure, luminosity), or it can also be computed - based on actual observations - and stored into tabular form [ZPS96, HH09]. • Envelope. Between 10−3 and ∼ 108 g cm−3 we have this thin layer, composed primordially of fully-ionized nuclides, electrons and photons. The aggregation state of the massive particles moves from a gas-like below 104 g cm−3 towards a lattice of nuclides imbued in a gas of degenerate electrons near 108 g cm−3. Electrostatic interactions thus become important as we move towards the high density sector, and depending on the temperature it is feasible to have a liquid phase near 106 g cm−3. Precisely because of this within some contexts the envelope is divided in two parts, one belonging to the atmosphere and the other to the “ocean” at the surface of the neutron star [MC11, GBS+07, DG09]. • Outer crust. This layer occupies the region between ∼ 108 g cm−3 and ∼ 1011 g cm−3, i.e. the neutron drip point. In the outer crust we have a lattice of nuclides imbued in a gas of degenerate electrons. For numerical purposes, it is suitable to split the neutron star into two regions only, the (numerical) envelope, consisting on the actual atmosphere, envelope and a 53 portion of the outer crust below 1010 g cm−3, and the numerical core, including the rest of the star, i.e. outer and inner crust, as well as the core. 2.3.1 The envelope, ρ ≤ 108 g cm−3 In this portion of the star - the envelope - matter is composed of photons, Ntot(≥ 1) different species of fully ionized nuclides and electrons. i.e. the composition of a typical Sun-like star. From 10−3 g cm−3 (at the surface) to 108 g cm−3, at envelope- crust boundary, the aggregation state varies from a gas-like (or plasma-like) to a jelly-like environment where electrostatic interactions among ions and electrons become prominent, eventually allowing the conformation of ionic lattice structures above 108 g cm−3 and at moderate temperatures (i.e. 108 K). The prominence of electrostatic effects is quantified throughout the electrostatic-to-thermal ratio, i.e. the coupling parameter Γ = Z2e2 acellkBT , (2.27) where acell = (3ncell/4π)1/3 is a characteristic dimension associated with the local number density of the cell ncell and Z is a local average charge. According to numerical simulations, the crystallization of matter takes place at Γ ∼ 200, although the exact threshold is sensitive to composition [BP13]. The gas-like environment can be formally described via the standard non inter- active EOS, composed of the simple addition of the individual pressures of the ions, described as an ideal gas, the free electrons obeying the Fermi-Dirac distribution and the photons obeying a Planck distribution. On the other hand, the jelly-like environ- ment requires the addition of at least three terms describing Coulomb interactions among electrons (electron-electron), between electrons and ions (electron-ion) and among ions (ion-ion). However, the overall contribution to thermodynamical func- tionals is still small (corrections in the order of ∼ 10%) as to allow the retaining of the additivity property for pressure and Helmholtz free energy. Furthermore, among the three kind of corrections, ion-ion contributions has the largest impact at above 104 g cm−3, hence this term receives more attention than electron-electron and electron-ion. Mass versus number normalization. Since me ≪ mu and neutrons are typically bounded well below 108 g cm−3, it is viable to assume all mass is contained in protons and neutrons and to write the rest-mass density as ρB = munB, with nB the baryonic number density, and mp ≈ mn ≈ mu as the relative difference is of ≤ 2%. In the study of stellar plasmas, the distribution of total mass or particles among the different Ntot species leads to two different normalization schemes. The first, frequently encountered in the context of nuclear reactions, prioritizes the distribution of mass by introducing the mass density ρi for each species. A single fully ionized 54 nuclide i has Zj protons and Nj neutrons, or Aj = Nj + Zj baryons. This amounts for a total mass of Mi = Zimp + (Ai − Zi)mn ≈ Aimu. In the thermodynamical limit, ρi := lim (Ni,V )→(∞,∞) Mi Ni V = Aimuni, (2.28) related to the mass fraction Xi and the mass abundance Yi, Xi := ρi ρB = Ai ni nB , Yi := ni nB = Xi Ai , (2.29) with ni the ionic number density, obeying the mass conservation Nspecies ∑ i=1 Xi = Nspecies ∑ i=1 YiAi = 1. (2.30) Considering the Xi add up to unity, we can take the collection of Nspecies fractions as weights to compute mean values, as for instance ⟨Z⟩ = Nspecies ∑ i=1 ZjXj, ⟨A⟩ = Nspecies ∑ i=1 AjXj, (2.31) ⟨Z2⟩ = Nspecies ∑ i=1 Z2 jXj, ⟨A2⟩ = Nspecies ∑ i=1 A2 jXj. (2.32) From the individual ionic number densities nj, it is natural to define the number density of total (fully-ionized) nuclides, or number density of total ions, as nion = Nspecies ∑ i=1 ni = nB ⟨m⟩ion (2.33) 1 ⟨m⟩ion = Nspecies ∑ i=1 Yi, (2.34) where ⟨m⟩ion is referred to as the ion mean weight. For plasmas, charge neutrality holds. This imposes an additional relation between electrons and rest-mass, ne = Nspecies ∑ i=1 Zini = nB ⟨m⟩e (2.35) 1 ⟨m⟩e = Nspecies ∑ i=1 ZiYi := Ye, (2.36) with ⟨m⟩e the electron mean weight. For a mixture of 70% 1H, 28% 4He and 2% 56Fe, for instance, ⟨m⟩e ≈ 1.177, i.e. Ye ≈ 0.85, while for pure-12C or pure-4He systems we have ⟨m⟩e = 2 and Ye = 0.5. 55 The second renormalization scheme is frequent in the study of plasma interac- tions. Here, the ion number densities ni are renormalized by introducing the number fraction xj := nj/nion, resulting in an analogous to the mass conservation expression as Nspecies ∑ j=1 xj = 1. (2.37) What is the relation between Xj and xj? By direct substitution, it is easy to see xj = ⟨m⟩ionYj = Ā′Yj = Xj/Aj ∑Nspecies j=1 Xj/Aj , (2.38) i.e. the xj can be taken as a renormalized version of Xj, and the old ⟨m⟩ion now plays the role of an “average mass number”. However, in this renormalization scheme the proper average for the mass number, considering the weights are now the xi, is ⟨A⟩′ = Nspecies ∑ j=1 Ajxj. (2.39) While charge neutrality is still given by ne = ∑ j Zjnj, the xj’s induces a new definition for the electron mean weight, now having the physical meaning of being the inverse of the “average charge”, i.e. ne = Nspecies ∑ j=1 Zjnj = nion⟨Z⟩′ (2.40) ⟨Z⟩′ = Nspecies ∑ j=1 Zjxj = 1 ⟨m⟩′ e . (2.41) The coupling parameter. Old studies of plasmas consisted on modeling matter as composed of a single-species, thus receiving the name one-component-plasma, OCP. Modern studies, in contrast, focus on mixtures of species, i.e. multi-component- plasma MCP. Within the particle number renormalization, it is viable to define an appropriate sum of coupling parameters [HPY07, DG09] Γeff = Nspecies ∑ j=1 xjΓj =   Nspecies ∑ j=1 xjZ 5/3 j  Γe (2.42) Γe = e2 kB ( 4π 3mu )1/3 (ρ/⟨m⟩e)1/3 T ≈ 2.2747 × 105 K cm g−1 × (ρ/⟨m⟩e)1/3 T , (2.43) 56 with Γe the electron coupling parameter. For a pure-12C matter, for instance, at ρ = 106 and T = 108 K we have Γe ≈ 0.18 and Γeff ≈ 3.57, practically little effect of electrostatic interactions, while at T = 104 K we have Γe ≈ 1805 and Γeff = 3.5×104, i.e. a carbon crystal assuming Γcryst ≈ 200. The EOS - Helmholtz approach. Since ρB, T and the abundances Y are usually taken as independent variables in standard stellar astrophysics, the ade- quate thermodynamical potential is f̆ = f̆ideal + f̆interaction, where f̆ideal includes the ion ideal gas, free electrons and photons contributions, while f̆interaction include the Coulomb contributions. From the partial derivatives of f̆ , we recover the required thermodynamical functionals (e.g. [JSB+21]): P = ρ2 B   ∂f̆ ∂ρB   T,Y s̆ = −   ∂f̆ ∂T   ρB,Y ε̆ = f̆ − T   ∂f̆ ∂T   ρB,Y c̆V = −T   ∂2f̆ ∂T 2   ρB,Y χT = T P ( ∂P ∂T ) ρB,Y χρ = ρB P ( ∂P ∂ρB ) T,Y c̆P = c̆V + χT χ2 ρ P ρBT ∇ad = PχT TχρρBc̆P , noting that for P , s̆, c̆V and ε̆ the additive property of f̆ is preserved, while for the rest of them the interaction and ideal terms are mixed. The justification for an addi- tive f̆ or P , despite the intrinsically intensive nature of this thermodynamical func- tional is related to the experimental Dalton’s law [BB09], affirming such addition is valid as long as the individual components are non-interacting or weakly-interacting. In a substantial amount of papers, the Helmholtz free energy F and other func- tionals, such as the heat capacity, are expressed in units of NkBT , where N is taken as the total number of ions (see for instance [BST66, SDD80, IIT87, CP98, PC13]). The provided functional can be easily adapted into our notation as f̆int = kBT ⟨m⟩ionmu fpapers int . (2.44) Non-interacting sector As anticipated, in f̆ideal we include the contributions from photons, Nspecies different species of ions and electrons, i.e. f̆ideal = f̆photons + f̆ion + f̆e, subject to mass conservation, Eq. 2.30, and charge neutrality, Eq. 2.36. For photons we have f̆photons(ρB, T ) = −4σSB 3ρBc T 4, (2.45) independent of the ion abundances Y . Sometimes, the constant at the front, in- volving the Stefan-Boltzmann constant σSB, is rewritten in terms of the so-called 57 −5 −3 0 2 4 6 8 10 log10ρ [ g cm−3 ] 4 5 6 7 8 9 lo g 1 0 T [K ] P io n = P ph ot on P io n = P e, n r 〈m〉i = 1.296 〈m〉e = 1.176 Γ e ff = 20 0 log10Pideal [ erg cm−3 ] −5 −3 0 2 4 6 8 10 log10ρ [ g cm−3 ] 4 5 6 7 8 9 η = − 10 η = − 1 η = 10 η = 10 3 η = 10 5 log10|η| [Adimensional] 6.75 9.00 11.25 13.50 15.75 18.00 20.25 22.50 24.75 27.00 −2 −1 0 1 2 3 4 5 6 7 −4.32 −3.60 −2.88 −2.16 −1.44 −0.72 0.00 0.72 1.44 Figure 2.2: Left panel: contour plots for the non-interacting EOS, assuming a Solar- like composition of 70% 1H, 28% 4He and 2% 12C. Right panel: log10♣η♣ contours, considering the same composition as in left panel. a-radiation coefficient, a := 4σSB/c. Via partial differentiation, we see ε̆photons = aT 4/ρB and Pphotons = ρBε̆/3, i.e. photon pressure is independent of the local bary- onic density and strongly dependent on temperature. For ions’ free energy we have two terms: one from rest-mass and the other from the kinetic terms (see the discussion around Eq. 2.22). However, the rest-mass term solely depends on Yi, thus having no impact for other functionals such as P , c̆P and f̆ion. Therefore, ions’ free energy is given by f̆ion(ρB, T,Y ) = kBT mu Nspecies ∑ j=1 Yj [ ln ( nj nQ,j ) − 1 ] (2.46) nj = ρBYj mu , nQ,j = ( mjkBT 2πℏ2 )3/2 , (2.47) with nj the j-th ion number density, and mj its corresponding mass. By partial differentiation, we recover the standard relations [KWW12] Pion = ρBkBT ⟨m⟩ionmu , ε̆ion = 3 2 Pion ρB . (2.48) As discussed elsewhere [Hua87, JEL96], for free electrons there does not exist an exact expression for functionals such as Pe(ne, T ) since (a) the Fermi-Dirac in- tegrals take µe and T as arguments, not ne, which is actually a functional of µe and T as well, and (b) the evaluation of these integrals must be done numerically. As a consequence of this complication, for astrophysical calculations several options 58 have been introduced and employed, for instance (a) compute these integrals nu- merically and store the data as tables, implementing interpolating subroutines or fitting functionals based on them [TS00], (b) employ analytic expressions, resulting from an exact integration of ne(µe, T ) and Pe(µe, T ) under certain limiting cases [Nad74, Pac83] or (c) attempt to construct a fitting formulae based on the limiting cases [EFF73, JEL96]. Among these limiting cases, the most important is the exis- tence of a region where functionals are approximately independent of temperature, i.e. the degenerate limit, a general feature of fermionic systems. Intuitively, it can be understood in terms of the ratio T/TF, with TF the Fermi temperature of the gas. At T/TF ≪ 1, thermal effects become unimportant and functionals are of the form F (ρB, T ) ≈ F1(ρB) + c1(T/TF)s with s ∈ N, c1 constant and F1 a functional of rest-mass. Due to the smallness of the temperature ratio, changes in temperature are practically decoupled from changes in pressure. The dependency of TF with ρB makes this approximation to break down as long as T/TF ≫ 1, and calls to con- sider another parameter to quantify the degeneracy of the gas. This is done via a redefinition of the chemical potential µe in terms of the adimensional parameter η = µe −mec 2 kBT . (2.49) We say electrons are degenerate if η ≫ 1, and otherwise if η ≪ 1, region where thermal effects become important. Since µe can be negative, η as well, in good agreement with the fact electrons behave as an ideal gas, i.e. Pe ≈ nekBT , as ρ ≪ 102 g cm−3 and T > 106 K. In addition to degeneracy, for electrons we can distinguish two limits depending on whether the kinetic energy exceeds the contribution of rest- mass or not. This is decided via the kinetic-to-rest ratio x := p/mec. If x ≪ 1, rest-mass contribution exceeds the kinetic energy and we say electrons are non- relativistic. On the other hand, if x ≫ 1 the energy is predominately kinetic and thus electrons are relativistic. If electrons are highly degenerate, these relativistic limits allow to construct exact expressions for the pressure, useful for back-of-the-envelope calculations: in first place, ρB ∝ ⟨m⟩ex3 [Hua87], i.e. extremely relativistic electrons are usually associated with high-density environments and viceversa. Additionally, in the non-relativistic case, Pe ≈ 1013(ρB/⟨m⟩e)5/3 (c.g.s. units), while for the extreme relativistic case Pe ≈ 1015(ρB/⟨m⟩e)4/3 (c.g.s. units) [Pac83]. For instance, at ρB = 108 g cm−3 and for pure carbon-12 matter we have Pe ≈ 1025 erg cm−3. In the present work, for the free electrons’ EOS we adopt the fitting functionals from [JEL96], valid for a large span of densities and temperatures, i.e. ∼ [10−10, 1010] g cm−3 and ∼ [103, 1010] K respectively due to their good agreement with respect to numerical evaluation of the integrals. These fits, however, are still functionals of T (via the adimensional t = kBT/mec 2) and µe via an f parameter, a strictly positive functional of η, thus requiring the inversion of the functional ne(f, t) in order to extract, given a fixed ρB, the correct f for the remaining functionals (pressure, internal energy, etc). 59 In order to gain further insight on the non-interacting EOS, let us discuss its main functionals employing a fixed, Solar-like composition of 70% 1H, 28% 4He and 2% 12C. In Fig. 2.2 we display the pressure Pideal - resulting from the addition of Pion (Eq. 2.48), Pphotons and the fitting formulae for Pe - and the η parameter as contours in the ρB − T plane. By inspection of both panels, we see a correlation between electron degeneracy and the Pion − Pe,non-rel boundary. At the right hand side of this threshold, in left panel of Fig. 2.2, electrons provide the largest contribution to pressure. From the behavior of η in the right panel, we see electrons must have a high degree of degeneracy to dominate the pressure over the ideal gas contribution from ions. Moreover, since the envelope’s thermal profile resides in this region of the ρB −T plane, we deduce electron’s pressure controls the evolution of the largest portion of such profile. In the other extreme of the left panel, below 10−1 g cm−3 and above 106 K, photons start dominating the contribution to pressure, going as far as 1020 erg cm−3 at extremely low densities (i.e. 10−5 g cm−3) and 109 K, although in this region e+ − e− pair production becomes present as well. Now let us explore the sensitivity of the EOS to temperature and density via the χT and χρ coefficients displayed in Fig. 2.3. As expected from the behavior in pressure, χT = 1 in a diagonal band going from 10−5 g cm−3 and ∼ 104 K to 106 g cm−3 and 109 K. As electrons become degenerate and their contribution to pressure exceeds that from ions, χT slowly converges towards 10−3 near the crystallization region. On the other hand, for temperatures above this χT = 1 contour we see photon pressure rapidly taking the leading role, thus producing a drastic decrease in χρ, as shown in the right panel of Fig. 2.3. Regarding the degenerate electron region, we see the contours adopting a triangular-shaped pattern. Finally, let us comment on the heat capacity densities, Fig. 2.4. As a consequence of electron degeneracy, we see cV ≈ cP in the electron-dominated region of the ρB−T plane, in good agreement with the theoretical result on heat capacities. Since cP contains non-linear terms related to partial derivatives, the behavior with respect to cV is entirely different in the η < 0 region of the ρB − T plane, where cV exhibits a ρ-independent tendency above 106 K as a consequence of the ideal gas behavior and the progressive influence of the photon pressure. On the other hand, cP drastically increases up to 1030 erg K−1 cm−3, which we can easily attribute to the smallness of χρ. 60 −5 −3 0 2 4 6 8 10 log10ρB [ g cm−3 ] 4 5 6 7 8 9 lo g 1 0 T [K ] 3. 0 Id ea l g as , 1 .0 10 − 1 10 − 3 Γ e ff = 20 0 χT −5 −3 0 2 4 6 8 10 log10ρB [ g cm−3 ] 4 5 6 7 8 9 Id ea l g as , 1 .0 10 − 110 − 3 4/ 3 〈m〉i = 1.296 〈m〉e = 1.176 Γ e ff = 20 0 χρ 0.00 0.44 0.88 1.32 1.76 2.20 2.64 3.08 3.52 3.96 0.00 0.18 0.36 0.54 0.72 0.90 1.08 1.26 1.44 1.62 Figure 2.3: χT (left panel) and χρ (right panel) for the non-interacting EOS, assum- ing a Solar-like composition of 70% 1H, 28% 4He and 2% 12C. −5 −3 0 2 4 6 8 10 log10ρB [ g cm−3 ] 4 5 6 7 8 9 lo g 1 0 T [K ] 8.0 12.0 14.0 16.0 Γ e ff = 20 0 log10cV [ erg K−1cm−3 ] −5 −3 0 2 4 6 8 10 log10ρB [ g cm−3 ] 4 5 6 7 8 9 8.0 12.0 14.0 16.0 〈m〉i = 1.296 〈m〉e = 1.176 Γ e ff = 20 0 log10cP [ erg K−1cm−3 ] 3.09 4.62 6.15 7.68 9.21 10.74 12.27 13.80 15.33 16.86 2.88 5.94 9.00 12.06 15.12 18.18 21.24 24.30 27.36 30.42 Figure 2.4: Heat capacity densities at fixed volume (left panel) and fixed pressure (right panel) for the non-interacting EOS, assuming a Solar-like composition of 70% 1H, 28% 4He and 2% 12C. 61 2.3.2 Outer and inner crusts, ρ ∈ [108, 1014] g cm−3 Outer curst 27.5 28.0 28.5 29.0 29.5 30.0 log10P [ erg cm−3 ] 9.75 10.25 10.75 11.25 11.75 lo g 1 0 ρ [ g cm − 3 ] Neutron drip threshold HZD BPS, AME2020 BPS, HFB26 26 27 28 29 30 log10P [ erg cm−3 ] 18 21 24 27 30 33 36 Z N e u tr o n d ri p th re sh o ld HZD BPS, AME2020 BPS, HFB26 Figure 2.5: EOS for the outer crust near the neutron drip line, considering two different models for nuclei masses. Aside from present uncertainties in nuclear physics (such as binding energies), between 108 and 1011 g cm−3 the qualitative aspects of the EOS are relatively straightforward: ions conforming a lattice imbued in a relativistic gas of highly degenerate electrons, according to the behavior of the η parameter in Fig. 2.2 [BPS71, HZD89, HPY07, JSB+21]. Since nuclides conform a crystalline structure, the Wigner-Seitz approximation is employed to impose charge neutrality inside a cell of characteristic dimension ai = [3/(4πni)] 1/3, which gives rise to a lattice pressure PL = −fcZe2ni/a (c.g.s. units for e), where fc is a structure factor associated with the kind of lattice chosen5. In this picture of the outer crust, however, there still prevail several critical questions: (i) To which extent are thermal effects negligible? (ii) What kind of species are there and which are their mass fractions? (iii) What kind of crystalline structure are they arranged into? Regarding question (i): an inspection to the left panel of Fig. 2.2 shows that for T ≤ 109 K and ρ → 1010 the pressure is practically supplied by very degenerate electrons, as indicated by the high values of η in the right panel of Fig. 2.2. While claiming P = Pe is inaccurate since electrostatic interactions - neglected in the aforementioned plot - are likely to take a more prominent role in the outer crust 51.4507 for a pure spherical case, 1.4442 for bcc 62 than in the envelope due to the gas-to-liquid-to-solid phase transitions, we can expect the outer crust EOS to actually decouple pressure from temperature, i.e. structure becomes independent of thermal effects, an advantage usually exploited to simplify the simulation of thermal evolution (e.g. Chapter 1). This characteristic is shared with the inner portions of the star (inner crust and core) where neutrons and protons, being fermions as well, have large Fermi temperatures and thus require T ≥ 1010 K to require the presence of thermal effects in their description. Such decoupling allows some simplification in the process of computing the EOS: regarding the pressure, we end up with an expression of the form P (ρ), now with ρ = ε/c2 instead of ρ = ρB. On the other hand, second order quantities such as the specific heat do retain a dependency on temperature. Indeed, recalling at high degeneracy electron’s EOS behaves as a series in powers of (T/TF )2, the specific heats behaves as ∝ T/TF . However, given T/TF ≪ 1 it is customary to regard this regime as of T → 0 and take cV,e ≈ cP,e ≡ ce, i.e. we can speak of a single heat capacity (see Appendix A.3 for further details). Such T → 0 construction is customary referred to as the ground state of matter [BPS71]. To address questions (ii) and (iii), the knowledge on the composition at each den- sity or pressure is required. In principle, it can actually be a mixture of species with different abundances, as shown in the case of accreting systems [HZD89, GBS+07], or it can be a composition enriched in only one nuclear species, as for instance the “canonical” neutron star envelope of pure 56Fe matter. Single nuclear species. This scenario has been kept as the standard in the field given the simplicity of computing the EOS for single-composition layers at ρ ≤ 1011 g cm−3. Since pressure and temperature now take a decoupled role from the thermodynamical point of view, to construct the outer crust EOS and its derivatives it is useful to shift from a Helmholtz to a Gibbs free energy formalism since the independent variables are now temperature (with T → 0), pressure and composition, for the latter assuming at each pressure there exists only one kind of nuclear species [BPS71, HZD89, HPY07]. Given such single-species character of their ground-state EOS, the usage of density as independent variable is discouraged given the prominent role of discontinuities: certainly, given the continuous character of the pressure (related to the structural Φ function, see Chapter 1), as well as the predominance of electron’s pressure Pe = Kρ/⟨m⟩e)4/3, by computing its differential and taking the limit dPe → 0 we obtain a density jump with each discontinuity in composition, i.e. ∆ρ/ρ = ∆⟨m⟩e/⟨m⟩e. Concerning these single-composition approximations, it is important to notice their ground-state search pursues to describe two limiting scenarios, which nonethe- less cover the majority of neutron star cases for which thermal evolution can be addressed: (a) Composition in equilibrium [BPS71], presumed to be of isolated objects which are cooling after their birth, i.e. matter has reached an equilibrium 63 against nuclear reactions such as β-decays or captures. For more than twenty years, this model - shortly referred to as BPS given the initial letter of authors’ surnames - became known as the “canonical” outer crust for neutron stars. According to BPS, for the construction of the ground-state EOS we require a set of pairs A = { (A(j), Z(j)) }M j=1 (with A(j) ≥ 56 and for which masses are known either experimentally of theoretically), a grid of pressures P = { P (i) }K i=1 . ∀P (i) ∈ P, and to perform the following steps: (i) n(i) e and G(A(j), Z(j), n(i) e )/A(j) are calculated for each member of A. The former is obtained by solving P (i) = Pe(n (i) e ) + PL(niN) , (2.50) n(i) e = Z(j)n (i) N , (2.51) while the latter is calculated from G(A(j), Z(j), n(i) e ) = 1 n (i) N [ ρ(i)c2 + P (i) ] , (2.52) ρ(i) = 1 c2 [ n (i) NM(Z(j), n (i) N ) + εe(n (i) e ) ] (2.53) M(Z(j), n (i) N ) = WL(n (i) N ) +WN(A(j), Z(j)) , (2.54) where the nuclide mass is expressed as WN , and ρ(i) is now the associated mass-density considering relativistic terms beyond the rest-mass. For explicit expressions of εe and Pe, see [ST83]. (ii) The pair producing the lowest G(A(j), Z(j), n(i) e )/A(j) is chosen as the ground state composition for P (i), and the rest of thermodynamic vari- ables can now be deduced from such G. As ρ ̸= nbmc 2, now the baryon number density for the j-th step is n (i) b = A(j)n(i) e Z(j) . (2.55) In Fig. 2.5 we plot the resulting BPS-EOS, as generated with the algorithm described above6. In the left panel we show density against pressure, while in the right panel we observe the change in charge number as a function of pres- sure as well. For the calculations two different sets of nuclear binding energies were employed, both combining experimental data and theoretical calcula- tions: those from the recent AME2020 [WHK+21] and those from the old AME2012, dubbed as HFB26 [GCP13] given the involvement of the Hartree- Fock-Bogoliubov model in the computation of nuclear masses. In the right 6The corresponding numerical code was developed by the author of this work. 64 panel we observe some discrepancies on the Z minimizing the Gibbs free en- ergy around 1027 erg cm−3 and at 1025 erg cm−3, where the updated data suggests 56Ni is the ground state of matter instead of 56Fe. On the other hand, these discrepancies have little impact over ρ(P ), which follow an almost simi- lar track nevertheless sensitive to composition discontinuities as expected from the ∆ρ/ρ = ∆⟨m⟩e/⟨m⟩e estimation. (b) Composition out of equilibrium [HZD89, HPY07], for recurrently accret- ing neutron stars where these deep layers, composed of ashes from previous nuclear burning episodes, are compressed due to the piling up of material at the surface and thus induce pycnonuclear reactions. While this model still takes place within the Gibbs free energy formalism, the changes in composi- tion are not dictated by minimization of G/A but instead by electron captures [Sat79, HZ90a, HZ90b, HZD89]. The picture is as follows: a mass element of fixed A is pushed inwards the star due to mass accretion. On its downward path, it is possible that Z changes each time that a non-equilibrium beta cap- ture (Z → Z − 1) takes place. Being an unstable state, a second electron capture must occur at the same pressure, i.e. Z − 1 → Z − 2. This lower-Z ion keeps traveling inwards with successive electron captures dictated by this stability criterion, until it eventually reaches the neutron drip point, where its chemical potential is equal to the rest-mass energy of the neutron. The nu- merical implementation for constructing the corresponding EOS - dubbed as HZD-EOS - goes as follows: we now consider a set with fixed A and different Z, which we shall denote as Z. We still have a grid of pressures P , and at each of them charge neutrality holds as well. For P (i), we impose an initial Z(i) ∈ Z (although this proton number is not the minimum of the set), solve for n(i) e from Eq. 2.51 and study the sign of the beta-capture criterion (which is valid ∀P (j) ∈ P): Q = M(Z(i) − 1, n (i) N ) −M(Z(i), n (i) N ) − µe(n (i) e ) . (2.56) If sgn(Q) > 0, we move to the next pressure in the grid with Z(i+1) = Z(i). On the other hand, if sgn(Q) ≤ 0 Z(i) must be replaced with Z(i) − 2 and n(i) e is recalculated. This same procedure is repeated until the full grid of pressures is exhausted or the neutron drip point is reached, whichever occurs first. On Fig. 2.5 we contrast the HZ and BPS EOSs, the former constructed with the recently described algorithm7. While their ρ(P ) exhibit minor discrepancies such as the location of the jumps, and thus put into evidence the major role played by the electron’s pressure, the changes in composition result in low-Z matter for the HZD-EOS in contrast to the high-Z matter of both BPS-EOSs. 7Again, the numerical code was developed by the author of this work. 65 Multiple nuclear species at each pressure. If, however, one is following the evolution of the chemical composition then for the outer crust it is still viable to restore to EOSs such as Skye [JSB+21] or PC [CP98], based on the electrostatic interactions briefly described in the previous section, given their interacting sector take into account the transition from liquid to solids and allow multi-species sys- tems. When employing these EOSs, we must be aware of (i) limiting their range of validity to ρB ≤ 1011 g cm−3, given the neutron drip threshold, and (ii) the original composition of the functionals for which crystallization effects were computed, a limitation shared with the conductivity in the outer crust as well. Historically, this limitation can be appreciated by the custom introduction of impurity parameters in the description of the crust, emerging from the idea that single-composition layers (formed during the neutron star birth) might contain small amounts of different nuclides. It must be stressed out that, despite the disagreement of the models on the composition, there exists agreement on the interactions governing the outer crust, i.e. almost free electrons and nuclides in lattices. Equally important is to stress out that, despite such similarity on the governing interactions, quantities such as released heat and their conductivity vary from model to model as well since, in turn, they critically depend on the chemical composition of the outer crust, a source of uncertainty given the complete accretion history of a given neutron star is unknown. Inner crust Between 1011 and 1014 g cm−3 most of the existent models predict a similar behavior in the P − ρ plane, in spite of the discrepancies on the exact composition and in the arrangement and shape of ions [Lat12, LRP93, NV73]. An important prediction of these models is the gas of free neutrons, which appears since the chemical potential of these particles is µn ≈ mnc 2. The threshold density of this process, commonly referred as neutron drip point occurs at ∼ 1011 g cm−3, although the exact value is model-dependent. In this region, however, temperature effects are not negligible for the purpose of describing thermal evolution (e.g. [KWW12, JEL96]). 2.4 Thermal conductivity and opacity In the standard non-relativistic analysis of the radiative flux, thermal conductivity Krad is defined in terms of standard thermodynamical variables as Krad := 16σSBT 3 3ρBκrad(ρB, T,Y ) , (2.57) where the functional κrad is defined as the opacity. From first principles we know the mean free path lfree expresses the characteristic distance energy carriers can move 66 10−3 10−2 10−1 100 kBT/mec 2 10−1 100 l− 1 eγ σ − 1 T h om so n n − 1 e 10 7K 10 8K 10 9K η ≤ −10 η = 0 η = 5 Weaver Poutanen Paczynski Figure 2.6: Contours for the product of the mean free path multiplied by the cross section and the electron number density, for different fits for the electron opacity as a function of temperature. Based on Fig. 3 from [Pou17]. between collisions, thus we can also write opacity as κ := (ρBlfree) −1. In the presence of conduction, as in neutron star’s crust, the associated flux also contributes with an associated thermal conductivity, Kcond, which can be parametrized in a similar way as Eq. 2.57 introducing an opacity due to conduction κcond. Since fluxes are additive quantities, in the case where both radiative and conductive fluxes are present it is customary to define a single thermal conductivity and effective opacity, K = 16σSBT 3 3κeffρB (2.58) 1 κeff = 1 κrad + 1 κcond , (2.59) where the addition rule for opacities resembles that of parallel resistors. While the distinction of radiative and conductive opacities is useful for theoretical purpose, in practice both quantities face the same difficulties for their obtention as opacity is not a directly observable quantity. Indeed, as lfree is completely dependent on the exact composition of matter and all the interactions among species enclosed in their respective cross-sections, opacity is necessarily a functional of (ρB, T,Y ), thus demanding the usage of either purely theoretical models, or extrapolations from experimental constraints. In astrophysical applications this problem is even more accentuated by the apparent circularity of the situation (using for a proof the elements which are intended to be proven). While such degeneracy can be broken from independent analysis of stellar surfaces in objects like our own Sun, stellar interiors are slightly more complicated as they are not directly observable. 67 Consequently, under such scenarios analytic models become indispensable tools for stellar evolution, representing exceptional cases for the common advise of exclusively employ first-principle, self-consisntent calculations [CKM+16], [PGC17]. As the calculation of stellar opacities is an extensive subject deserving its own chapter due to the large amount of subtleties, practically an art by itself (e.g. [Car76],[PGC17]), here we limit to discuss their very essential features. • Radiative opacity. Also known as Rosseland’s, this functional takes into account the effects of bremsstrahlung (free-free electron transition) and of γ-electron scattering, referred to as Thomson’s and Compton’s in the non- relativistic and relativistic regimes respectively. Formally speaking, this opac- ity is the integral of the harmonic mean of the individual contributions, al- though for practical purposes it is possible to compute each one of them sep- arately, then add them up as (e.g. [PY01]) κrad = (κff + κes)A(κff, κes, T ). As customary, free-free opacity consists on the sum of individual contributions, one from each nuclei species and all of them factorized in terms of the Gaunt factor gff(A,Z, ρB, T ), as reported in one of the most cited expressions, e.g. [SBCW99], emerging from fits to old numerical calculations. Regarding γ- electron scattering, to date one of the best fitting formulae, correctly covering the Thomson’s and Compton’s regimes was provided by [Pou17], starting from self-consistent calculations and leading to improvements on previous fits for both sectors as discrepancies are considerable with old approaches, as illus- trated in Fig. 2.6 for a selected collection of η. Aside from these analytic fits, the results of self-consistent calculations have been reported in tabulated form by several groups of work, such as the Opac- ity Project [BBB+05] [MSB+07], the Lawrence Livermore National Labora- tory OPAL [RI92], [IR96], and at Los Alamos National Laboratory [MAC+95] [CKM+16], where the opacities correctly account for all the energy levels, ion- ization degree and local ρB−T conditions. For astrophysics, however, the most significant defect of these works is both abundances and isotopes are specified beforehand, limiting their range of applicability in spite of the Principle of Corresponding States (e.g. [Gug45] and [RI92]) as it demands the presence of the same ions in any situation to be valid. A similar problem might arise with the extension of the covered range in the ρB −T plane, which is not necessarily homogeneous for all compositions. • Conductive opacity. As previously described, this concept and its functional form is only introduced to simplify the definition of a total opacity. However, typical works on the subject focus on the calculation of the associated thermal conductivity of electrons instead [SY06], [CPP+07], [DG09], i.e. Kcond = 16σSBT 3 3ρκcond , (2.60) 68 which can be computed from first principles via Kcond = aTk2 Bne m∗ eν (2.61) where ν is the effective electron scattering/collision frequency, and a is a nu- merical factor depending on the relativistic status of the electrons. In practice, ν = νee + νei, i.e. the individual frequencies of electron-electron and electron- ion collisions are simply added, a good approximation within the context of neutron stars given conduction takes place at high densities, i.e. ρ ≥ 105 g cm−3, where electrons are highly degenerate, i.e. η ≫ 1 [HL69, Zim01] (see the right panel of Fig. 2.2). Consequently, the majority of works on this subject provide analytic fits for the associated frequencies as we can always construct κcond employing Eqs. 2.60 and 2.61. Thus, for κcond we consider the fit for νee from [SY06], and the fit from [SBCW99] for νei. 2.5 Particle reactions Let ∑N j=1 aj → ∑M j=1 bj be an arbitrary interaction process, where M particles, the products, emerge as a result of the interaction among N particles, the reactants. The associated probability per unit volume and time for this process to happen, the reaction rate, can be written in formal terms as [GW04, dDGFJ14] r∑N j=1 aj→ ∑M j=1 bj = nBR∑N j=1 aj→ ∑M j=1 bj = ∫ N ∏ i=1 dni P   N ∑ j=1 aj → M ∑ j=1 bj   , (2.62) where the measure of integration includes the incoming particles’ number densities, and the P functional contains all the information related to the nature of the in- teractions among particles, and we have explicitly stated its relation with respect to the rate R discussed in Chapter 1. Naturally, the reaction rate for the inverse process, the inverse reaction for shortness, can be written in a similar fashion consid- ering the replacement M → N in the measure of integration and the appropriate P . By construction, r can be expected to be a functional of temperature T , rest-mass density ρB, individual abundances and their charge and mass numbers (Yi, Zi, Ai) (∀i ∈ ¶1, . . . , N +M♢). Independently of the individual abundances, by employing dimensional analysis we can associate each reaction rate with an intrinsic lifetime τ−1 = ρBNA⟨σv⟩T . From a theoretical point of view, it is plausible to consider P ∝ ♣MNM ♣2, where MNM is the matrix element of the process, as well as P ∝ σ, with σ the total cross- section of the process due to σ ∝ ♣MNM ♣2 [GW04, dGFJ10]. As one might expect, as N increases, r decreases, as the available surface for N particles to interact simultaneously becomes smaller. We can get a more formal glimpse of this by 69 inspection of MNM : for two-body processes, it is equal to the sum of contributions of the form Mif (Ei−E)−1; for three-body processes, it is the sum of products of two such terms. This trend implies, in energy space, MNM ∝ E1−N , meaning that only low order terms, such as one- and two-body interactions, are important. However, since ni ∝ nb ∝ ρ in dense environments the probability of P -body interactions (P > 2) becomes significant as the proportionality ρPB can overcome the damping due to E1−P of the matrix element. Although the computation of r strongly depends on the individual distribution functions and the nature of the interactions, locally it can be expressed as r = r0 N ∏ i=1 Y si i ρ ψρ B TψT , where si is the total number of reactant particles from the i-th species. This form introduces a first classification for reaction rates: if ψT ≫ ψρ, the reaction is referred to as thermonuclear, a behavior largely seen in reactions taking place between ρB ≤ 107 g cm−3 and T ≥ 104 K. On the other hand, if ψρ ≫ ψT the reaction is labeled as pycnonuclear (from the greek πυκνoς, translated as dense or compact [Cam59]), a behavior typically taking place above 108 g cm−3 and below 107 K. From a very strict point of view, there exists neither a purely thermo- nor a purely pycno- nuclear reaction, although these categories are useful to understand which thermodynamical DoF is locally the most influential for r. Several published works have shown the feasibility, when the involved particles are nuclides and photons, of splitting the reaction rate into three terms: one purely thermonuclear rthermonuclear, regarded as the main part of the rate, one purely pyc- nonuclear, and an enhancement factor due to the electrical influence of the surround- ing cloud of electrons, i.e. screening effects. For the majority of astrophysical appli- cations, well below 108 g cm−3, pycnonuclear effects are almost negligible and such contribution is simply taken as equal to 1. We must note, however, that such split- ting might be of limited use when the number of reactants exceeds two, e.g. [FL87]. Regarding the purely thermonuclear part, an extensive compilation of them, with excellent results for numerical applications, can be found in the [JIN22, CAF+10] library, where the following parametrization is adopted: rthermonuclear = M ∑ j=1 exp[aj0 + aj1T −1 9 + aj2T −1/3 9 (2.63) + aj3T 1/3 + aj4T9 + aj5T 5/3 + aj6ln(T9)] , (2.64) where M differs from rate to rate, depending on the amount of resonances. Associated to each rate is the Q-value of the reaction, formally defined in terms of the rest-mass of the reactants and products, denoted as maj and mbj respectively, 70 as Q∑N j=1 aj→ ∑M j=1 bj =   N ∑ j=1 maj − M ∑ j=1 mbj   c2, (2.65) it serves to quantify the amount of gained (Q > 0) or lost (Q < 0) energy by the system due to the N -body interaction. The first scenario, as it provides energy for the system, is called exothermic, in contrast with the later which is referred to as endothermic as it needs to be provided with energy to take place. For nuclides, having Z protons and electrons, N neutrons (i.e. A = N + Z baryons), their rest-mass is expressed as mj(Aj, Zj) := Zjmp + Zjµe + (Aj − Zj)mn − B(Aj, Zj)/c 2, with B(A,Z) the binding energy and µe the electron’s chemical potential (see Eq. 2.23 and the discussion below for the importance of keeping terms associated with mo- mentum for electrons). Due to baryon conservation, if all reactants and products are nuclides, the Q value is simply the difference between sums of binding energies. For electroweak reactions, on the other hand, besides this difference the (mp−mn)c2 term emerges as a consequence of n ↔ p transmutations. An alternative to the bind- ing energy - and reported in tables as well - is the mass excess ∆m = mj − Ajmu, which is the difference between the actual rest-mass and the baryonic rest-mass a particle with the same number of nucleons, all with masses mu, would have [GCP13, WHK+21]. Due to the definition of the atomic mass unit, ∆m = 0 for 12C [Tay09]. An important classification scheme for nuclear reaction rates is set considering the number of reactant species: • One-body: Also referred to as particle decay. In this category we have photo- disintegration, a + γ → ∑M j=1 bj, and β-decays which are mediated by the electroweak force. In a broad sense, the correspondent reaction rate adopts the form ρBλ(ρB, T,Y ), a simplification which allows to speak in terms of lifetimes τ = λ−1. • Two-body interactions. Practically most of the reactions of interest belong to this category. From the discussion regarding ♣MNM ♣2, it can be inferred P - body (P > 2) interactions can be understood in terms of successive two-body interactions, pointing out once more the great importance such processes have. Considering two particles in the reactant side brings an important simplifica- tion to the reaction rate: either from first-principles or by reasoning in terms of fluxes and targets, it can be shown Eq. 2.62 adopts the more familiar form r a1+a2→ ∑M j=1 bj = ∫ dna1 dna2 σ(♣va1 − va2 ♣) ♣va1 − va2 ♣. (2.66) 71 Clearly, the evaluation of this integral depends on the individual velocity dis- tributions of a1 and a2. In this regard, the most important limit case, due to the large range of applicability, is that when both particles obey Maxwell- Boltzmann statistics, allowing us to work with the center-of-mass energy as integration variable and to write the reaction rate as8 [HM06] r a1+a2→ ∑M j=1 bj = Ya1 Ya2 ρ2 BN 2 A⟨σv⟩T where ⟨σv⟩T = (8/π)1/2 M ′1/2(kBT )3/2 ∫ ∞ 0 dE E σ(E) exp (−E/kBT ) , M ′ = ma1 ma2 ma1 +ma2 . Among the several possibilities of reactions between two nuclides, the most frequently encountered in literature are those between lighter ones only, e.g. p(p, e+νe) 2H or α(α, p) 7Li, as well as those involving one nucleon and a heavy (m ≫ mp) nuclide, as for instance 12C(p, γ) 13N or 55Fe(n, γ) 56Fe. This can be justified from the fact that, the heavier both nuclides’ masses are, the shorter the probability of interaction becomes under typical conditions for temperature (≤ 109 K) and density (≤ 108g cm−3). As we approach or exceed these limits, reactions such as 12C +12 C → 20Ne + α adopt a more prominent role in the evolution of the system as long as their abundances are nonzero. Nevertheless, we can associate most of the two-body reactions among nuclides with one of the following categories: 1. Proton captures (p,γ). As its name suggests, we have AZ + p → A+1(Z + 1) + γ, with Q(p,γ) = B(Z + 1, A+ 1) − B(Z,A). 2. α captures (α,γ). Here we have AZ + α → A+2(Z + 2) + γ, resulting in Q(α,γ) = B(Z + 2, A+ 2) − B(Z,A) − B(α). 8In these notes, we follow the convention of replacing mu by N−1 A as their numerical values are reciprocal. 72 3. (α,p) transmutation. In this case, A+3(Z + 1) + p → AZ + α, Q(α,p) = B(A,Z) + B(α) − B(A+ 3, Z + 1). • Three-body interactions. Properly speaking, most of the reactions in this category emerge as subsequent two-body processes, either to reduce the size of networks as for instance happens when one studies A(α, p)B(p, γ)C under the assumption DtYB = 0 [WZW78, NTM85, GZ05], or due to the shortness of the intermediate state, as in the case of the 3α reaction which is formally defined as α + α → 8Be, 8Be + α → 12C, but for simplicity is taken as 3α → 12C [FL87]. • N-body interactions, with N ≥ 4, usually provide smaller contributions to the astrophysical processes, and as such can be neglected in most calculations (for theoretical examples, see for instance [GJ21]). In the astrophysical context, it is frequent to parametrize the cross section as σ(E)E = S(E)P (E), with S(E) the astrophysical S factor, available via extrapolations from terrestrial experiments, and P (E) the penetration factor, accounting for the probability of nuclide to overcome the Coulomb barrier. For the nuclear problem, we can generally write it as [Sal54, HK94, Kam14] P (E) = exp    −2 ∫ L dr √ 2m ℏ2 [U(r) − E]    , (2.67) where L denotes the spatial interval where integration is to be performed. In the absence of screening effects (that is, the response of the background medium to the presence of charged nuclides a1 and a2), this factor can be simplified into the Gamow factor P (E) = e−2πηSomm , with the Sommerfeld parameter ηSomm = Za1 Za2 e2 ℏ ( M ′ 2E )1/2 . Enhancement due to medium: electron screening. Since nuclear reactions are an interplay between nuclear and electromagnetic (or, more properly, electric) forces, where the projectile is expected to have sufficient energy as to overcome the Coulomb barrier of the target one can ask: but can the composition of the background alter the effective Coulomb barrier? After all, in macroscopic media we learn boundaries affect the strength and geometry of the electric field. The 73 answer in this case is yes: the background responds to the electric charge/field of both interacting nuclides. And the barrier the projectile must overcome is in fact influenced by this as well, such that the total potential energy of the system is given (in c.g.s./Gaussian units) by U(r) = UCoulomb(r) − Uscreen(r), (2.68) UCoulomb(r) = Z1Z2e 2 r , (2.69) where Z1,Z2 are the charge numbers of projectile and target, r is the distance be- tween them, and Uscreen(r) is the screening potential, i.e. the response of the back- ground medium’s electric field to the presence of these charged particles. From a thermodynamical point of view, this should be expected as a functional of temper- ature, density, composition of the background, and Z1, Z2. What do we do with r then? In principle, this variable will be suppressed in practical applications for sev- eral reasons, the most important being that the actual rate is constructed by thermal averaging over the energy of the charged particles, which implies performing a set of integrals. The second reason is that we do not necessarily have to evaluate difficult integrals related to exp [−Uscreen(r)/kBT ]: as proven by [Sal54], under certain con- ditions we can safely replace Uscreen(r) by Uscreen(r = 0), or Uscreen(0) for simplicity. As a consequence, the reaction rate can be expressed as [Sal54] R = exp[Uscreen(0)/kBT ]Runscreened, i.e. this simplification allows to treat screening effects in the thermonuclear regime as a multiplicative factor over the unscreened rate, i.e we must compute/attach screening factors. Quite a nice and simple picture! So far, however, we have said anything regarding the actual calculation of Uscreen(0), sometimes denoted as H(0). That is, in fact, a quite broad subject, strongly de- pendent on microphysical inputs as well as on the classical or quantum assumption regarding the underlying distribution functions. Regardless of this, it is useful to distinguish at least three thermonuclear regimes, determined by the ratio of electro- static and thermal energy Γ [Sal54, AL22]: • Strong screening. Predominant at high densities (not in excess of 109 g cm−3), here we have Γ ≫ 1. • Weak screening. Operating at low densities and high temperatures, it is characterized by Γ ≪ 1. • Intermediate screening. In this regime, which represents a transition be- tween the first and the second of this list, we have Γ ∼ 1. 74 Probably the most famous within the astrophysical community of stellar evolution is the fudging of weak and strong screening subroutines[PBD+11, PMS+15]9, as it allows a smooth transition between the two above regimes and include the quantum corrections as introduced by [Jan77, AJ78]. This formulation is not unique: we have for instance the approach from [CDY07], which at least in MESA exhibits some issues if employed around 1010 g cm−3 and 106 K. 2.5.1 β reactions 0 2 4 6 8 10 T/T9 −2.05 −2.00 −1.95 −1.90 −1.85 −1.80 lo g 10 λ [s − 1 ] β+ dec NuDat 3.0 JINA-CEE 0 2 4 6 8 10 T/T9 −6 −4 −2 0 2 4 6 8 10 ρYe = 1011 g cm−3 ρYe = 108 g cm−3 β+ cap β− dec β− cap 17F ! 17O (Oda et al, 1994) Figure 2.7: β reactions for the 17F→ 17O process, as computed by Oda et al 1994 [OHM+94]. The different linestyles are associated with values of ρYe, as illustrated in the right-hand panel. Left panel: rate of direct reaction, β+ decay, as a function of temperature. The comparison with two databases (constant lines) is given. Right panel: remaining β processes: β− capture (direct reaction), and β+ capture and β− decay (inverse reactions, i.e. 17O→17F). This category includes ion-lepton scattering mediated by the electroweak force. As mlepton ≪ mproton, reaction rates involving leptons can be written as Pion,lepton = nionλlepton, where λlepton = ∫ dnlepton σ(vlepton)vlepton (2.70) is the natural definition of this factor, considering Eq. 2.66. For typical stellar interiors, ¶e−, νe♢ and their antiparticles are the only leptons of relevance. Furthermore, as e± in these reactions were originally called β± particles, it is frequent to label all the following as β reactions [Zub20]: 9https://cococubed.com/code_pages/codes.shtml 75 • β± decay. Consists on the emission of e± and νe/ν̄e due to the decay of a proton(neutron) inside the nucleus, A ZX → A Z−1Y + e+ + νe +Qβ+ (2.71) A ZX → A Z+1Y + e− + ν̄e +Qβ− . (2.72) • β- capture. Occurring inside neutral atoms, it consists on the capture of a bounded e−, in the lowest orbital state, by a proton, giving rise to A ZX + e− → A Z−1Y + νe +QEC . (2.73) Similar transitions can occur for electrons in higher orbital states, although transition probabilities are significantly reduced as they correspond to first- and second-forbidden transitions, according to their correspondent selection rules. • β+ capture. In this process, a bounded neutron captures a positron and decays into a proton and an antineutrino, A ZX + e+ → A Z+1Y + ν̄e +QPC . (2.74) Considering the rest-mass energy of a nuclei with Z protons and electrons plus N = A− Z neutrons is W (Z,A) = Z(mp +me)c 2 + (A− Z)mnc 2 −B(Z,A) , (2.75) with B the binding energy, the Q value for each β-channel can be written as Qβ− = W (Z,A) −W (Z + 1, A) −mec 2 (2.76) = B(Z + 1, A) −B(Z,A) + 0.782 MeV −mec 2 (2.77) Qβ+ = W (Z,A) −W (Z − 1, A) −mec 2 (2.78) = B(Z − 1, A) −B(Z,A) − 0.782 MeV −mec 2 (2.79) QEC = W (Z,A) −W (Z − 1, A) +mec 2 (2.80) = B(Z − 1, A) −B(Z,A) − 0.782 MeV +mec 2 (2.81) QPC = W (Z,A) −W (Z + 1, A) +mec 2 (2.82) = B(Z + 1, A) −B(Z,A) + 0.782 MeV +mec 2 , (2.83) where 0.782 MeV is the numerical value of [mp+me−mn]c2. By inspection, Qβ+/− = QEC/PC − 2mec 2, implying the Q value of decays must be larger than 2mec 2 for such process to dominate over electron captures. Since β+β− → 2γ is certain to occur under typical astrophysical conditions, the energy carried away by photons is accounted for in the Q-value simply by neglecting the −mec 2 term at the end of 76 Qβ± , indicating electron’s energy has become photon’s. As is customary, the net amount of specific energy production can be easily estimated as Qm−1 u λ. Although calculations for σ(vlepton) constitute a subject demanding careful ex- planations [DLC+98], [FFN82], [JLH+10], [BRSGS+06], we can grasp the essential features in actual expressions for λ through Fermi’s model of weak-Interaction aris- ing in standard perturbation theory: the total probability of transition per unit time, or essentially the rate of capture or decay of a β particle, can be expressed as [OHM+94, Kam14] λβ = Constant × ∫ dx x √ x2 − 1x2 νF (Z, x)fFD(x, T ) , (2.84) where x = Ee/mec 2, xν = Eν/mec 2. These x-terms naturally emerge from stan- dard calculations of three-body decay with the daughter nuclei remaining at rest; F (Z,Ee) is the correction factor due to Coulomb interactions, and fFD is Fermi-Dirac distribution function. The constant at the front is mainly composed of electroweak coupling constants and electron’s mass, while the limits of integration are set by the specific channel the particles are following. It is important to note that, in spite of Q factors becoming independent on neutrino’s energy, for computing the net rate of specific energy the contributions from ν/ν̄ must be subtracted. Consequently, the instantaneous rate of specific energy loss due to β± neutrinos is given by ˙̆εν,β± = m−1 u × Constant × ∫ dx x √ x2 − 1x3 νF (Z, x)fFD(x, T ) . (2.85) Although the above formalism is quite general, under certain conditions these rates can be shown to be independent of both temperature and density, although not nec- essarily of Z. This approximation is supported by experimental results [GGIH09] [GNI+10], although controversies for low-Z ions and captures still prevail [RSD+20], [LRW20], [NRB+01]. From theoretical side, it can be argued that environment can play a role via electronic screening, altering the effective Z, and even a more promi- nent one in the low-ρ and high-T regime, such as supernovae conditions, as electrons become highly non-degenerate, T/TF ≥ 1 [FFN82]. As the zero-temperature approx- imation is adequate for practical purposes, in the remaining of the text we consider λ as T -independent. We must be aware, however, that such simplification still de- mands the evaluation of at least the integral for ˙̆εν,β, as the decay rates for certain ions are experimentally known. In general, there does not exist a single formula, leading to the usage of tabulated data [GM71, OHM+94]. One of the few excep- tions for tabulated specific neutrino energy comes from the β± decays and electron captures occurring in the CNO cycle and pp chains respectively. Considering, from [FCZ75] ⟨Eν⟩FCZ β± = W± 0 2 ( 1 − 1 w2 0,± )( 1 − 1 4w0,± − 1 9w2 0,± ) (2.86) ⟨Eν⟩FCZ pp = QEC , (2.87) 77 where QEC is the Q-value of the reaction, W0,± = Qβ± ∓mec 2 and W0 = mec 2w0,±, ˙̆εFCZ ν = Yim −1 u τ−1⟨Eν⟩FCZ , (2.88) where τ denotes the lifetime of the reaction and Yi is the abundance of the parent ion. Fig. 2.7 exemplifies the typical behavior of β reaction rates, taking the particular case 17F → 17O as reported by [OHM+94]. Due to the difference in binding energies and the available phase space for integrals, we see β+ decay is the most likely of the 4 reactions at Yeρ ≤ 106 g cm−3 and T ≤ 109 K, exhibiting a very weak dependence on density and parabolic trajectory on temperature as T ≥ 4 × 109 K. As most astrophysical scenarios of interest lie below this threshold, it is usual to consider this rate as T - and ρ- independent. The chosen fixed value can be taken as either the theoretically estimated one, or as the Earth-measured inverse lifetime if available, as in the present case where the horizontal lines are taken from two database set, NuDat3 [NuD22] and JINA-CEE [JIN22, CAF+10]. On this regard we must notice the lack of absolute convergence between theoretical and experimental rates at Earth conditions, a difference which becomes more accentuated at increasing parent’s Z number. As this discrepancy might be a situation of concern, for the typical stellar environments the Earth’s rate, due to its experimentally verifiable status, might be regarded as the correct one. At progressively more extreme physical conditions the β− capture takes the lead- ing role, significantly increasing in proportion with density. As a consequence of Pauli blocking factors, we see extreme environments are certain to reduce the likeli- hood of the “inverse” processes, β+ capture and β− emission, justifying the approx- imate treatment of β reactions as non-reversible rates. 2.5.2 Chains and cycles In astrophysical scenarios, rarely just one reaction is responsible for the whole syn- thesis of elements. Instead, chains of reactions, sometimes confined to cycles, are the mechanisms controlling the overall evolution of the composition. We have seen the probability of interaction among nuclides is severely restricted by Coulomb repulsion. Considering this barrier is overcame in high-density environ- ments, of seldom appearance in standard astrophysics, it is not surprising the most discussed chains and cycles of reactions are those involving burning and exhaustion of H, He, C and even O in the context of late stellar evolution. At the envelope of accreting neutron stars, where the conditions are favorable for the nuclear burning of these elements, it is feasible to find these chains and cycles as well. Thus, let us review the main features of these reaction chains. Hydrogen burning. Being the element with lowest Z number, there are many mechanisms operating at increasing conditions of temperature and density. In pro- gressive order, these are: 78 • pp chains. This collection of reactions, primordially operating at T ≤ 107 K and ρB ≤ 103 g cm−3, have the net balance of converting four 1H atoms into a single 4He, liberating ∼ 15 MeV per process [KWW12]. We speak of them as chains since these are likely scenarios, with different probability of occurrence, emerging from the primordial chain 1H(p, e+ν) 2H (2.89) 2H(p, γ) 3He. (2.90) Of them, the first reaction is the slowest one not only between these two but among all reactions composing the pp chains, having an associated lifetime of ∼ 1010 yr. For practical purposes, this allows to approximate the specific energy generation rate as ∼ Qall chains × r1H(p,e+ν). The three channels (or pp chains) allowing to synthesize 4He after the oc- currence of Eq. 2.90, labeled according to the amount of helium produced, are: pp - I : 3He( 3He, αp) 1H (∼ 83%) ; pp - II : 3He(α, γ) 7Be(e−, ν) 7Li(p, α) 4He (∼ 17%) ; pp - III : 3He(α, γ) 7Be(p, γ) 8B(e+ν) 8Be(αα) (∼ 0.02%); pp - IV(hep) : 3He(p, e+νe) 4He. The fourth chain, usually called hep, is regarded as the highest source of neutrinos, although having the severe drawback of being very unlikely [BK98, Bah02]. • CNO cycles. These cyclic chains have, once more, the net balance of trans- forming four 1H atoms into 4He, now with C, N and O isotopes acting as catalyzers. These cycles supersede the pp-chains as 1H burning mechanisms at T ∈ [107, 108.25] K and ρ ≤ 106 g cm−3. Due to the interplay of temperature- dependent reaction rates and β-decays, we can identify cold and hot versions of the cycles, operating below and above ∼ 108.25 K respectively [WGU+10]. The chains of reactions conforming the CNO cycle are: I : 12C(p, γ) 13N(e+ν) 13C(p, γ) 14N(p, γ) 15O(e+ν) 15N(p, α) 12C II : 15N(p, γ) 16O(p, γ) 17F(e+ν) 17O(p, α) 14N(p, γ) 15O(e+ν) 15N III : 17O(p, γ) 18F(e+ν) 18O(p, α) 15N(p, γ) 16O(p, γ) 17F(e+ν) 17O IV : 18O(p, γ) 19F(p, α) 16O(p, γ) 17F(e+ν) 17O(p, γ) 18F 18O(p, α) . In the hot version of the cycle, reactions such as 13N(p, γ) 14O, 17F(p, γ) 18Ne, 18F(p, γ) 19Ne, 18F(p, α) 15O and 15O(α, γ) 19Ne, together with β+-decays of 14O, 18Ne and 19Ne, become more likely as T exceeds 108.25 K, enhancing the 79 amount of heavy-Z ions due to the short lifetimes of these hot processes. As T ∼ 108.40 K, the production of F and Ne isotopes is enhanced through breakout reactions such as 14O(α, p) 17F, 15O(α, γ) 19Ne or 12C(α, γ) 16O(α, γ) 20Ne. The left panel of Fig. 3.1 illustrate several of these temperature dependent and independent rates at ρ = 104 g cm−3, where we can infer that between 108 and 108.6 the CNO cycles, as long as some Ne and Na are present, CNO cycles can be broken and the reactions induce the enhancement of Mg and Al. On the other hand, the right panel illustrates the amount of energy generated from the CNO cycle, in the approximation introduced by [Cla68, NH03], which consists on an almost-constant section coming from the hot CNO and dominated by the approximately constant β+ decays of 14O and 15O (see the left panel). On the other hand, the T sensitive region is controlled by both β+ decays and the proton-capture reactions involving 13N - 14N. • Rapid proton capture process. Dubbed as rp-process, it consists on the synthesis of heavy elements (Z ≥ 24, A ≥ 50) as a consequence of many 1H captures, practically its exhaustion, in a timespan of t ≤ 103 s, short with respect to typical H-depletion time in standard stars, i.e. ∼ 106 yr. This rapid burning takes place at T ≥ 108.4 and ρ ≥ 105 g cm−3 for initial mass fractions of hydrogen of ∼ 70%. Although the name of this process suggests (p, γ) reactions are the sole drivers, the actual rp-process is in fact a collection of sub-processes: – CNO-breakout. This step is crucial for switching on the whole rp- process. As previously discussed, 14O(α, p)17F and 15O(α, γ)19Ne break the I-IV cycles and start the enhancement of F and Ne. However, 19Ne is more likely to decay into 19F than to produce 20Na (via p-capture) as T <∼ 108.4 K. Once this relation is inverted, the material ceases to return to CNO and synthesis of Z ≥ 10 becomes viable. – Synthesis below A = 54. In this mass range, Tz = (N − Z)/2 = −1 nuclides, namely 22Mg, 26Si, 30S, 34Ar, 38Ca, 42Ti, 46Cr, 50Fe and 54Ni be- come the main axis of hydrogen burning, which follows regular patterns. Between each of these contiguous pairs, we observe a competition between three chains: (i) a sawtooth-path of two p-captures, β+ decay, p-capture, β+ decay and a final p-capture, (ii) an α-capture like path, i.e. (α, p) followed by a p-capture, and (iii) a β-3p-β-p path, i.e. three consecutive p-captures between two β+ decays. In the left panel of Fig. 2.8 these three paths are illustrated with arrows over arbitrary Tz = −1 nuclides (of A ≤ 54). Typically, chain (i) is favored among 22Mg, 26Si and 30S, while chain (iii) is substantially prominent among the remaining Tz = −1 nuclides. – Synthesis between A = 54 and A = 64. The main axis of H burn- 80 ing in this mass range shifts from Tz = −1 to α nuclides, suppressing in the process the competition between chains (i)-(iii). The synthesis in this range proceeds by p-captures and β+ decay reactions bounded by three triangular-like structures: (i) 54Ni-54Fe-58Zn, (ii) 58Zn-58Ni-62Ge and (iii) 60Zn-60Ni-64Ge. An example of such structure, as well as the re- actions involved, is given in the right panel of Fig. 2.8. This latter isotope represents the typical endpoint at the peak of thermonuclear explosions (bursts). – Synthesis for A ≥ 64. Once the main axis of H burning has settled in α-nuclides, the synthesis of heavier species than A = 64 proceeds via p-captures and β+ decays bounded in similar triangular-like structure as in the former case. In addition, this region is characterized by cascades of β+ decays from even-even α nuclides such as 68Se or 72Kr at the peak of these triangular structures. – Endpoint. The possibility of extending the rp-process beyond A = 64 up to a maximum A depends on the local temperature: at T ∼ 108.4 K, Amax ∼ 72 while at T ∼ 108.7 K we obtain Amax ∼ 90. At the peak of thermonuclear explosions, i.e. T ∼ 109.2 K, it is feasible to synthesize up to 107Te as burning reaches a Sn-Sb-Te cycle endpoint at 105 ≤ A ≤ 108 due to the (p, γ) − (γ, p) equilibrium, (γ, α) and β+ decays. In contrast to the pp-chains and the CNO cycles, an adequate simulation of the rp-process requires the presence of many species. While a good estimation on the specific energy generation rate is given by the combination of CNO- breakout rates [WGS99], [CN06], it still must be checked if such approximation is valid for all astrophysical scenarios of interest. Helium fusion. For matter in excess of T ∼ 108 K and ρ ≥ 102 g cm−3, the reaction 4He(α, γ) 8Be∗ becomes more likely to happen. However, such excited state is unstable and has a large probability of decomposing into 2α particles or, if helium abundance is significant, has a probability of advancing to an excited state of 12C, the Hoyle state [Hoy54], via 8Be∗(α, γ) 12C∗, subsequently falling into the ground state of 12C and liberating ∼ 7.28 MeV per α particle. As the lifetime of 8Be∗ is ∼ 10−16 s [HK94], this chain of reactions is simply written as 3α →12C and referred to as triple α reaction, mechanism responsible for providing the nuclear energy for evolved stars [HHIK56, KWW12]. Although the effective reaction rate is taken as a three-body one, ⟨ααα⟩T , it can be understood as the product of consecutive two- body reactions due to their formal origin [FL87]. From the experimental point of view, on the other hand, it can be estimated from its inverse reaction 12C → 3α (e.g. [FDB+05]). As we infer from the right panel of Fig. 3.1, at very dense and hot environments with prominent presence of 4He the 3α process is the main source 81 α α Tz = − 1 Tz = − 1 α α A ≤ 54 A ≥ 60 Figure 2.8: Main flow of the rp-process, among Tz = −1 and α nuclides. of nuclear energy, significantly exceeding the amount from the CNO cycles and pp chains. 2.6 Thermal neutrinos Besides the neutrinos from β-reactions, within standard stellar matter there exist additional mechanisms - again mediated by the electroweak interaction - emitting νν̄. In Fig. 2.9 we illustrate their corresponding Feynman diagrams. In first place we have pair neutrinos, coming from the exchange of Z or W bosons by the e+e− and νeν̄e pairs [Chi66]; bremsstrahlung neutrinos [Hau75], arising as a consequence of electron scattering from a nuclide AZ; photoneutrinos, due to the absorption of one photon by an electron and the posterior emission of a neutrino/antineutrino pair; plasmon neutrinos, coming from the decay of plasma (quasi-particles) excitations [ARW63, HRW94]. An additional mechanism in this category is the recombination neutrinos, arising from the emission of a νν̄ pair upon an e− capture by a nuclide AZ. From the microphysical point of view, their overall contribution for standard astrophysical scenarios is to take energy away from the system due to their trans- parency to ordinary matter. Such losses have been computed throughout the years by different groups and eventually adjusted as functionals of temperature, density and chemical composition throughout ⟨m⟩e, ⟨A⟩ and ⟨Z⟩. For astrophysical appli- cations the compilation [IHNK96] is the standard reference, with the fits expressed as Qν in units of erg cm−3 s−1 (see, however, [KPP+99] for an updated version of the bremsstrahlung neutrinos at high densities). In the case of photo, plasma and pair neutrino processes the fits depend on T , ρYe only, recombination additionally depending on η and ⟨Z⟩ and bremsstrahlung requiring ⟨A⟩. It is straightforward to 82 Plasmon decay neutrinos Pair neutrinos Bremsstrahlung neutrinos Photo-neutrinos Figure 2.9: Feynman diagrams for the relevant neutrino processes at the surface of the neutron star. Left column: pair neutrinos (based on [Dic72]). Middle column, upper diagram: bremsstrahlung neutrinos (based on [DD10]). Middle column, lower diagram: plasmon decay (from [CCC70]). Right column, upper and lower diagrams: photoneutrinos (based on [DRP04]). 83 7.0 7.5 8.0 8.5 9.0 9.5 10.0 lo g 10 T [K ] log10 ˙̆εplasma[erg g−1s−1] Pure 12C matter log10 ˙̆εpair[erg g−1s−1] 7.0 7.5 8.0 8.5 9.0 9.5 10.0 lo g 10 T [K ] log10 ˙̆εphoto[erg g−1s−1] log10 ˙̆εrecomb[erg g−1s−1] −5.0 −2.5 0.0 2.5 5.0 7.5 10.0 log10 ρ [g cm−3] 7.0 7.5 8.0 8.5 9.0 9.5 10.0 lo g 10 T [K ] log10 ˙̆εbremss[erg g−1s−1] −5.0 −2.5 0.0 2.5 5.0 7.5 10.0 log10 ρ [g cm−3] Pair Photo P la sm on B re m ss . R e c o m b . log10 ˙̆εν,thermal[erg g−1s−1] −30 −25 −20 −15 −10 −5 0 5 10 15 −32 −24 −16 −8 0 8 16 24 32 40 −30 −20 −10 0 10 20 −32 −24 −16 −8 0 8 −20 −16 −12 −8 −4 0 4 8 12 16 −10 0 10 20 30 40 Figure 2.10: Contour plots of thermal neutrino processes, considering pure carbon- 12 matter, for densities and temperatures of a typical neutron star envelope. The name of the corresponding process is indicated at the top of each panel. For display convenience, here we took max[ ˙̆εν,process, 10−30]. 84 convert these fits into a term with the same units as the output from the nuclear generation rate, i.e. ˙̆εν = Qν/ρ, as well as to add each individual contribution, suf- ficing to compute their individual Qν value, add them and then obtain the global ˙̆εν dividing by ρ. In the grand scheme of stellar evolution, it is frequent to either add both ˙̆εthermal ν and ˙̆εβν into a single neutrino energy generation rate ˙̆εν , or reserve this latter symbol to thermal neutrinos and to remove the neutrinos due to β-reactions from ˙̆εnuc. In the present work, we adopt the first path and shall employ ˙̆εν to denote all contributions from neutrinos, both thermal and due to β-reactions. Considering the fits from [IHNK96, KPP+99], in Fig. 2.10 we display contour plots for the different neutrino processes so far discussed, as well as for their grand total ˙̆εthermal ν , considering pure 12C matter. In the upper left panel we have the con- tribution from plasmon decay, which reaches a maximum around the region where η ≈ 0 (contrast the T − ρ panel against the η contours of Fig. 2.2). In the upper right panel we observe the pair neutrinos, whose maximum evidently takes place whenever the e+/e− production becomes viable as well, i.e. low densities and high temperatures, rapidly dropping to zero outside this region. Photoneutrino contri- bution (central left panel) follow a similar tendency as well, leading to a transition of dominance with respect to pair neutrinos at around T ∼ 108.5 K at low densities. Although recombination neutrinos (central right panel) reach a maximum at high densities and temperatures - but still within the η ∼ 0 region - its contribution to the grand total only exceed the rest of thermal processes at ρ ∼ 102.5 g cm−3 and T ≤ 107.5 K, region where the K-shell of carbon-12 exists. In the lower left panel we observe the energy release due to bremsstrahlung neutrinos, which exhibits a differ- ent tendency with respect to the rest of processes by remaining above 10 erg g−1 s−1 at ρ ≥ 105 g cm−3 and the whole range of selected temperatures, thus leading to the dominion of this process over the rest of them (lower right panel). This is a consequence of the bremsstrahlung taking place within the liquid and solid phases of the envelope-crust (i.e. Γ ≥ 200, Fig. 2.2). 85 Chapter 3 Nuclear networks for neutron stars “In our time art attracts the most noble souls. Art [...] is possibly the last refuge of freedom” O. Honchar Abstract. In this chapter we present the characteristics which a nuclear network of reactions must have in order to describe the envelope of the star. Introduction At the very core of standard stellar evolution is the problem of solving a large number of non-linear PDEs. While structure, temperature and luminosity account for 5 to 6 (depending on whether we are including relativistic effects or not, see Chapter 1), for each one of the species in our system we must add one corresponding PDE. For modeling the surface of neutron stars, how many species do we need? From the nuclear physics aspects discussed in Chapter 2 concerning the rp-process, a first look suggest we require a large number of species. According to different authors this is indeed the case, requiring up to 100 species for a correct account on the nuclear generated energy and a first approximation to the distribution of ashes due to 1H and 4He burning [vGI+94, RFR+97, SBCW99, GBS+07, FST08]. Considering we must solve every PDE of stellar evolution at each spatial point inside the star - i.e. the grid of points - having a large system of equations requires significant computational power unless we develop approximation methods based on the short and long timescale evolution of the different layers inside the star, limit the spatial resolution, or reduce the number of equations to be solved. In regard to this latter aspect, we enter into the domain of the fine art of constructing approximative nuclear reaction networks. 86 Another complication arising from the inclusion of the chemical evolution to the set of structure and temperature equations is their stiffness character as a conse- quence of the different timescales of each reaction. For instance, while 14−15O mean lifetime against β+ decays is of the order of mili-seconds, above A = 56 we find isotopes with lifetimes of the order of hours to days. Thus, any standard explicit method such as the 4th order Runge-Kutta are no longer useful for modeling the evolution of a neutron star [PTVF92], and it is imperative to study which integra- tion methods we have in order to perform a fast and accurate solution of the PDEs equations. These are the implicit methods which we briefly describe in this chapter as they constitute the core part of the numerical codes employed in the following Chapter. It is then the purpose of the present chapter to provide (i) a concise view into the actual networks for studying the envelope of the neutron star, and (ii) the numerical setup employed to analyze the evolution of the chemical species, critical part to compute an accurate nuclear energy generation rate. We review in the first section the issues on the construction of a network, briefly commenting on the available numerical methods to solve stiff sets of equations. In the next section, we comment on the networks already existing in the literature and introduce our own networks, commenting on the advantages and disadvantages of each of them. The chapter concludes with a quick overview on the evolution of the rp-process employing these numerical methods, leaving the implications for the whole envelope for the next chapter. 3.1 Overview of numerical methods Even with modern computers, the simultaneous solution of Eqs. 1.16, describing the spacetime evolution of the Nspecies constituent species, and the PDEs for T (t, r) and ρ(t, r) poses a numerical challenge for several reasons. In first place, the set of PDEs DtYj = Rj is stiff : physically (or intuitively) for the solution of the set we must pay close attention to very different timescales, ranging from 10−6 to 106 depending on the species involved, their linking reactions and the local conditions for temperature and density. For instance, in the left panel of Fig. 3.1 we can observe a collection of reaction rates involved in the hydrogen burning, expressed as lifetimes, i.e. τ = R−1 since [R] = Time−1, as functions of temperature at 104 g cm−3. Even at this middle density, notice the drastic variation in values for even a single rate, from 10−3 to 106 in the small window of 108 − 109 K. Additionally, we observe at, for instance T = 108.5 K, the drastic difference between reactions such as 18Ne(α, p)21Na and 19Ne(p, γ)20Na. Mathematically, a concise definition of stiffness requires to enclose several char- acteristics [HNW93], such as the artificial oscillations appearing on attempts to solve these PDEs via explicit methods (such as forward Euler or 4th order Runge Kutta), 87 7.0 7.5 8.0 8.5 9.0 9.5 10.0 log10T [K] 10−3 10−1 101 103 105 τ − 1 [s ] 14O 15O 20Na 18Ne 19Ne 22Mg ρ = 104 g cm−3 15O (α, γ) 19Ne 14O (α, p) 17F 18Ne (α, p) 21Na 19Ne (p, γ) 20Na 20Na (p, γ) 21Mg 22Mg (p, γ) 23Al 12C (α, p) 15N 6 7 8 9 log10T [K] 0 5 10 15 20 25 lo g 1 0 ˙̆ ε n u c [e rg g− 1 s− 1 ] (X⊙, Y⊙, Z⊙) pp CNO 3α ρ = 108 g cm−3 ρ = 102 g cm−3 Figure 3.1: Left panel: Selected collection of thermonuclear rates, expressed as the inverse lifetime τ−1 at ρ = 104 g cm−3, relevant for CNO cycles and their breakout. Right panel: energy generation rates per unit gram for pp, CNO and 3α processes, as discussed in the main text. and the requirement of employing either too small or too large time-steps as a con- sequence of the different scales in the right-hand side in order to keep stability and accuracy. While for some [HM06] stiffness can be verified via the ratio of maxi- mum/minimum Jacobian matrix eigenvalues, a pragmatic definition of stiff PDEs is the failure of explicit methods to provide satisfactory solutions [CH52]. It must be noted the evolution of mass fractions is responsible for introducing the stiffness into the system: if, for instance, we beforehand know the distribution of chemical composition across our whole star, it is viable to solve the TOV equations for the structure of the star restoring to explicit methods such as 4th Order Runge Kutta. As described elsewhere [HNW93], [PTVF92], [Tim99], [THW00], the appropriate numerical methods for solving stiff PDEs are implicit, requiring the inversion of large and sparse matrices M whose dimension is determined by both the number of species and the number of cells in the spatial dimension, i.e. dim(M) ∼ Ncells × Nspecies. For astrophysical simulations, having Ncells ≥ 100 cells, the calculations become computationally expensive as Nspecies ≫ 30. For instance, if each entry in the Jacobian matrix is a double-precision number (i.e. 8 bytes), having 100 cells and 30 species implies a Jacobian matrix of 576 MB of space. In an optimistic note, regularly we do not need to include ∼ 1000 species in all simulations to consider them realistic. For instance, including Z ≥ 20 elements for simulating the evolution of mass fractions after the Big Bang (T ∼ 109 K) is not mandatory since we are dealing with the synthesis of the very first elements in the periodic table, i.e. He, Li, Be, B and at most C, N and O. On the other hand, for the evolution of Main Sequence stars p- and α- captures over Z ≥ 26 elements is quite infrequent that 88 we can neglect its presence for the major part of the simulation. Finally, unless ρ ≥ 106 g cm−3, as for instance in simulations of supernovae or the neutron star crust, the inclusion of isotopes in the neutron-rich sector of the nuclide chart is not required since β+ decays of isotopes in the valley of stability cannot proceed due to Pauli blocking, β− captures require densities near 107 g cm−3 to become viable and neutron capture processes (n, γ) are inefficient as long as these particles are confined in nucleii. Thus, the most crucial aspect for a self-consistent, yet simple, computation of stellar evolution is to identify the range of temperatures and densities of interest and to keep the indispensable amount of nuclides as to cover the processes of interest. This, however, makes it clear the most expensive part of the computation is the evolution of the network. Implicit numerical methods. The conceptual idea behind this family of meth- ods consists on solving differential equations by root-finding methods, usually involv- ing the diagonalization of some matrix - the Jacobian one, given its multi-variable context. Considering a general set of PDEs ∂tA = f(A), with A denoting the array of dependent variables and f the right-hand side functions for each equation, the idea is to approximate by finite differences the system as An+1 = An + hf(An+1) and to solve this non-linear set to extract An+1, i.e. the new configuration at a “time” tn+1 = tn + h. The advantage of these methods is the freedom to choose either large or small time-steps, in contrast to the explicit methods, where f is evaluated at An, where the time-steps must be large in order to avoid instabilities. As concrete examples of implicit methods employed for modeling neutron stars, we can cite: • Henyey method. The workhorse for many stellar evolution codes [HFG64, PBD+11, KWW12, Pag16], consisting on the construction of matrices of easy inversion - via the Newton-Raphson method - and the convergence to the updated values of the independent variables provided an initial guess for them. While its implicit nature guarantees the possibility of choosing either small or large time-steps, the initial guess for the next step must be close in order to guarantee convergence, not only in a few iterations only, but as well given the underlying Newton-Raphson method. An additional difficulty is to properly write the coefficients for handling the inversion of the large resulting matrix - which is divided into a task of inverting smaller but many sub-matrices. • High-order, semi-implicit. These methods can be considered as general- izations for stiff problems of methods for non-stiff ones. For instance, the idea behind the Rosenbrock methods is to generalize the Runge-Kutta for stiff problems, from where the subroutine stiff emerges [PTVF92]. Another generalization is the Bader-Deuflehard semi-implicit method, which relies on the conceptual idea of polynomial or rational extrapolation and the midpoint method to obtain accurate solutions employing only a few steps. As it is, this 89 method seems the fastest one while keeping a relatively low amount of required steps [Deu83], [PTVF92], [LMJ14] [FOR+00]. Consequently, we opt for con- sidering this method, as implemented in [PTVF92] via the stifbs subroutine, for constructing our one-zone model and our code for stationary envelopes. 3.2 On the networks for neutron stars. Besides keeping as few nuclides as possible, further approximations can still be made to simplify the process of solving the system of PDEs while keeping a good grasp on the physics of interest. For neutron stars, efforts are oriented to have either a de- tailed evolution of the chemical composition or a good estimation on the generated energy due to nuclear reactions. In the first category we find the so-called one-zone models (OZM), which evolve the abundances’ PDEs in a single cell, i.e. at fixed spatial coordinate (for instance, at fixed radius or pressure). This constraints the temperature and density to adopt either constant values or, at most, to be functions of time. The numerical benefit of these models is that abundances’ PDEs turn into ODEs, i.e. we go from a BVP + IVP to an IVP, allowing to reduce the dimension of the Jacobian matrices to ∼ Nspecies instead of ∼ Ncells ×Nspecies. Physically, work- ing under specified local conditions for temperature and density allows to identify which reactions are operating and, consequently, which are the most important at certain phases of the whole system evolution. The main disadvantage of OZMs is neglecting diffusion effects, which play a role in the enhancement or diminishing of mass fractions via matter movements as well as in the transport of the generated energy via nuclear reactions, and thus on the overall evolution for T . A common approximation for saving computational time while resulting in accu- rate amounts of energy generation is to assume certain species have reached a steady state, i.e. DtXi = 0. This translates into the trading of differential by algebraic equations which, after being solved, define additional relations among mass frac- tions. Although such simplification does not automatically guarantee a reduction in the computation time, due to the delay root-finding methods might bring, for some systems these approximations define whole new networks which are actually fast and are suitable for simulating large periods of time. However, whenever we restore to this approximation we must beware of physical artifacts such as low-Z survival at high depths, underestimation of α nuclides or energy from β+ decays, among others. Examples of such energy-oriented approximation are: • pp-chains and CNO cycles. A good approximation for the specific energy generation rate of these processes is ε̇eff ≈ [ ∑ i Qi ] τ−1 longest , (3.1) 90 where τlongest ∝ r−1 longest is the corresponding timescale of the slowest reaction rate from the set. For instance, the lifetime of 1H(p, e+ν)2H, denoted as τpp, is the largest in the pp-chains, thus ε̇pp ≈ τ−1 pp ×∑iQi. Below 107.5 K, 12C(p, γ)13N is the slowest reaction of the so-called cold cycle, allowing to write the specific energy generation rate as ∼ QCNOr12C(p,γ) 13N, sensitive to both temperature and density. At T ∼ 108 K, i.e. the hot CNO cycle, β+ decays from 14−15O become the slowest reactions and thus the net production of energy becomes independent of temperature and density. In both scenarios, it becomes viable to trade the detailed evolution of 12−13C, 13−15N and 14−15O by an effective rate relating 1H, 4He and 12C. Following [Cla68] [NH03], rCNO = ( 1 τ13N + 1 τ13N,β )−1 ×   ( τ−1 13N τ13N + τ14N + τ14O,β + τ15O,β ) +   τ−1 13N,β τ14N + τ13N,β + τ15O,β     , (3.2) where τX and τX,β denote the lifetime of X against a thermonuclear reaction and a β+ decay, respectively. In the right panel of Fig. 3.1 we plot these approximations to ε̇ as functions of temperature and density. In the case of the approximation to the pp-chains, we see the fit produces reasonable results well below 107 K, around the temperature at which CNO cycles replace the pp-chains as main mechanisms of 1H burning. At increasing temperatures, we observe 4He burning via the 3α process provides the largest contribution to nuclear energy instead of the hot CNO cycle, which provides up to 1015 erg g−1 s−1. For the particular case of neutron star surfaces undergoing accretion: since H and He burning are favorable below 107 g cm−3 while heavy-Z burning requires higher densities, at the envelope it suffices to consider species in the proton-rich sector of the nuclide chart and the valley of stability, where rates are prominently thermonuclear. There exist several networks of reactions which can be employed for the description of the envelope. As stressed out above, neither is free of limitations, so in principle it is advisable to understand the contents of each one of them and to implement them according to the specific requirements. • Approx21. The workhorse for MESA, this network contains 21 species, between 1H and A = 56, covering the essentials of pp-chains, CNO cycles and α-burning up to 56Fe, as originally devised by its 19-species parent network [WZW78]. Given its reduced size, simulations are actually fast even with Ncells ≥ 1000. As main drawbacks we can cite the artificial super-bursts that occasionally emerge as a consequence of 1H reaching depths well below 106 g cm−3 (i.e. not exhausting due to the absence of proton-rich nuclides), the artificial bursting 91 suppression at moderate accretion rates which, albeit observationally correct, it is numerically artificial when contrasted against other networks, and finally its limited range of proton-rich nuclides, essential in the understanding of the rp-process. While it cannot be modified in MESA, it is easy to implement in other stellar codes due to its structure. • rp153 and rp300. These were devised by [FGWD06b, FTG+07] in order to simulate X-ray bursts having as little isotopes as possible while guaranteeing a good production of nuclear energy (in both networks), or allowing to have the indispensable heavy-A species synthesized during and after the burst peak, as for instance 107Te, according to the then stablished endpoint for the rp- process [SAB+01]. Although rp153 gives a correct description of the energy output, it suffers similar issues as Approx21, allowing hydrogen to reach high depths (≤ 107 g cm−3) instead of being burnt and limiting the heaviest-Z in the network to be in the A = 56 group. While the energy production of rp300 - an extended version of rp153 - is similarly correct and more species at A > 56 are included, its main drawback is the absence of A > 56 nuclides having lifetimes against β+ decay larger than 1 day. This reduces the amount of synthesized nuclides in the valley of stability after the bursts, particularly at A ≥ 56, inducing artificial amounts of 107Te for instance. Both networks, however, result in good fittings against observations for the luminosity curve of an average bursts (see [PMS+15]). The list of species for rp153 can be seen in Table 3.1. • net381. As such, this network is essentially an extended version of rp300. The nuclear energy production is accurate, and the number of species allow for an improved distribution of ashes in terms of A and Z. For instance, the presence of 80Kr allows the whole mass fraction of 80Y (t1/2 ≈ 30 s against β+ decay) to go to the valley of stability. The advantage of having more species near this zone, in contrast to the rp300 network, is the possibility of contrasting time-dependent and -independent simulations, where the energy released by the ashes of H-burning provide the largest contribution at ρ ≥ 106 g cm−3, as described in the next chapter. Since this network neglects the presence of neutron-rich nuclides below the valley of stability, its range of validity is restricted to ρ ≤ 108 g cm−3. Similarly to rp153 and rp300, in net381 we neglect the presence of isotopes above A = 91 near the valley of stability, thus having artificial amounts of 107Te for instance. The most problematic aspect of this network is the large times required for running a complete simulation in MESA. In a time-independent scenario, however, it is feasible to run relatively fast simulations with this network and to employ its output to understand the mass fraction distributions and the generation of nuclear energy. Finally, let us note the version employed in the following chapter, net380, only differs 92 from net381 by the exclusion of neutrons in the network. From this network, we can also extract another one, net344, by removing nuclides related to the endpoint of the rp-process but which cannot β+ decay in the absence of species with A ≥ 90 in the valley of stability. The isotopes which can be removed from the net381 network to conform this sub-network are enlisted in Table 3.1. • Approx140 and Approx170. These networks were devised to (i) guarantee a reasonable production of nuclear energy, (ii) include as much β+ decays from heavy-Z species as possible, (iii) keep nuclides related to the primordial paths of the rp-process between A = 20 and A = 56, and (iv) exhaust the available hydrogen before 106 g cm−3. The complete list of species can be found in Table 3.1. Let us now give a brief summary for our motivation to introduce these networks. As a first step, pp chains were omitted since the associated energy production occurs just at the very surface of neutron stars (i.e. below 102 g cm−3), where compression is actually more energetic than nuclear reactions. Li, Be and B nuclides are thus ignored. All C, N and O isotopes for the hot CNO cycle are included, since this is an essential part for the H-burning at middle (ρ ∼ 104 g cm−3) densities and temperatures (T ≥ 107 K). Below A = 56, β+ decays take less than 1 hour to occur (a remarkable exception is 26Al). This implies once a proton-rich A isotope has been synthesized during the rp-process, it might decay towards the valley of stability in just a few minutes. Due to the local maxima at α-nuclides in the distribution of ashes due to nuclear burning (see [SBCW99]), we make the approximation of allowing the whole chain of β+ decays towards the valley of stability only to isotopes leading to α-nuclides for A ≤ 52. Considering that the integrated flow of the rp-process, either in stable or explosive burning, proceeds between Tz = −1 and α-nuclides, such neglect should not affect the evolution of ashes well below 108 g cm−3. A secondary consequence of the necessity of keeping α-nuclides in the network is to adequately simulate 4He burning. Between 22Mg and 54Ni we have “box scheme” [RFR+97], where a competition between the three chains of reactions discussed in the previous chapter and in [FGWD06a], occurring between Tz = −1 isotopes, takes place. To decide which chain is more relevant, we considered the results of [SBCW99, SAB+01] as well as our own simulations (with updated reaction rates from [CAF+10] and [JIN22]): from 22Mg to 26Si and to 30S, the sawtooth path is the most prominent one and thus the associated isotopes are included in the network. On the other hand, from 30S to 54Ni, the β-3p-β-p path is the dominant one. Between 54Ni and 62Ge (Tz = −1 nuclides), the fictitious axis of the main flow moves from Tz = −1 to α nuclides and those at two β+ decays of separation from them, as for instance 60Ni and 64Zn. At A = 64, specifically at 64Ge synthesized during the peak of the bursts, the main flow now follows a triangular-like structure: a cascade of proton captures and β+ 93 decays connect α-nuclides, while heavy isotopes such as 64Zn, 68Ge and 72Se are synthesized and do not further decay due to their long lifetimes. To reduce as much species in the network as possible, we simulate an endpoint to the rp process following two basic criteria: (i) H is fully exhausted at around 106 g cm−3, and (ii) the heaviest nuclide must have a large (above 6 days) lifetime against β+ decay. We find 80Kr, and specifically the A = 80 family, as a suitable artificial endpoint for the process. The A = 84 family is a suitable endpoint as well, although it increases the amount of nuclides. In the particular case of the present work, only the latter networks will be amply discussed. net380 allows a better distribution of species and energy production in stationary state, net344 allows to understand the essential features of the rp- process and Approx140 (or Approx170 as well) allow fast numerical simulations while fulfilling the requirements for an approximately good modeling of the rp-process. 3.3 An overview of the rp-process 0.0 0.2 0.4 0.6 0.8 1.0 Time [day] 10−14 10−12 10−10 10−8 10−6 10−4 10−2 100 M as s F ra ct io n 12C 1H stifbs eps = 10−1, N = 30 eps = 10−3, N = 29 eps = 10−4, N = 37 eps = 10−6, N = 34 0.0 0.2 0.4 0.6 0.8 1.0 Time [day] 10−14 10−12 10−10 10−8 10−6 10−4 10−2 100 Z⊙, ρ4 = 7.08, T8 = 3.86 12C 1H stiff eps = 10−1, N = 48 eps = 10−3, N = 153 eps = 10−4, N = 549 eps = 10−6, N = 5176 Figure 3.2: Comparison of stiff integrators for a OZM with different accuracies eps. The number of steps until tmax is shown as N . Left panel: stifbs. Right panel: stiff. Numerical tests. A comparison of results from two different stiff integrators is shown, for the net380 network at fixed conditions of temperature and density, in Fig. 3.2. For simplicity we adopted the solar composition at t0 = 0 s. Given both integration methods require the user-provided parameter eps, we have selected a 94 collection of values ranging from 10−1 to 10−6. In the left panel we illustrate the results from the stifbs subroutine, and in the right panel those from stiff. In both panels, we observe a correlation between the size of eps and the total number of steps taken by the subroutine, although for stifbs we see the occurrence of a local maximum of N at eps= 10−4. While stiff shows a better consistency in the mass fractions among the four models, in clear contrast with the discrepancies among different eps for stifbs, the number of steps to be taken with stiff is always larger than the other subroutine. Furthermore, the accuracy for keeping a low N seems adequate for a pure mass-fraction evolution, although it is dubious whether it would work when incorporating the rest of equations, each with their proper scale and thus with different needs for eps. Notice that stifbs, on the other hand, yields consistent results as long as one restricts the eps parameter around 10−4, while N always remain smaller than the values obtained from stiff. In the long run, this suggests stifbs provides a faster and accurate performance than stiff. Consequently, we opt for employing this subroutine for our time-independent numerical code. Evolution of the network. In order to exemplify the development of the rp-process, let us now describe in detail one simulation at fixed density and temper- ature, ρ = 105.5 g cm−3 and T = 108.7 K. For simplicity, we employed the net344, composed of the nuclides depicted in Fig. 3.4, running for t ∈ [0, 105] s. As initial conditions we chose the simplest scenario: X1H = 0.70 and X4He = 0.30, composition which allows to observe how the rest of nuclides are synthesized as a consequence of hydrogen and helium burning. In Fig. 3.3, we illustrate the evolution of the mass fraction of a selected subset of nuclides, participating at different stages of the rp- process. In Figs. 3.4 - 3.7, the integrated reaction flow at different intervals of time are illustrated, in order to identify the main reactions controlling the evolution of the species. In the upper left panel of Fig. 3.3 we appreciate the evolution of the chemical composition at t ≤ 1 s. Here we see that, as a consequence of the temperature, 12C has been synthesized and exhibits an almost constant mass fraction. By inspection of Fig. 3.4, we infer this is due to an equilibrium between the rate of creation, 3α process, and that of destruction via 12C(p, γ)13N(p, γ)14O. Such chain also explains the rapid formation of 14O. The integrated flow and the mass fractions also shows the operation of the 14O(α, p)17F and 17F(p, γ)18Ne reactions as we observe synthesis of 17F and 18Ne. On the other hand, time is still too short as to further proceed into A ≥ 18 and the carbon-12 burning results in enhancement of 15O, a sign that 18Ne β+ decay and the subsequent 18F(p, α)15O is favored over the 18F(p, γ)19Ne reaction. This building-up of 15O, as we see in the upper right panel, is fundamental for the production of 19Ne via the breakout reaction 15O(α, γ)19Ne. At t ≤ 10 s, the mass fractions of 14−15O, 18Ne, 17F and 19Ne reach a maxi- mum, followed by a rapid decreasing: the 15O(α, γ)19Ne(p, γ)20Na chain has had sufficient time as to become efficient and the synthesis of A > 20 nuclides takes 95 place, as indicating by both the upper right panel of Fig. 3.3 and the integrated flow at 10 s, Fig. 3.5. The enhancement of proton-rich nuclides occurs among contiguous nuclides, such as 24−25Si, 33−34Ar or the endpoint at this time, 37−38Ca. These species, however, are short-lived: in the mass fraction diagram, we can observe their rapid decrease just some seconds after their synthesis. Let us also note, from the integrated reaction flow, that low-Z burning plays at this stage a more promi- nent role than the synthesis of A > 20 material as for the production of heavy-Z we require 15O, which the chain of reactions 3α → 12C(p, γ)13N(p, γ)14O(α, p)17F (p, γ)18Ne (β+νe) 18F (p, α)15O keeps supplying at the cost of temporarily decrease the mass fraction of carbon-12. In the integrated flow, we can also identify some of the discussed paths for the rp-process, such as the sawtooth one between 26Si and 30S. At t ≤ 102 s (lower left panel of Fig. 3.3 and Fig. 3.6), we observe a fast synthesis of material, up to 60Zn, 63Ga and 64Ge, as well as some enhancement of 68Se. Below A = 56, we observe the presence of the sawtooth path between 30S and 34Ar, as well as the β−3p−β−p path among 34Ar-38Ca, 38Ca-42Ti, 42Ti-46Cr and 46Cr-50Fe. In the intermediate region between Fe and Ga, we observe (p, γ) and β+ decays dominating the integrated reaction flow across the whole nuclide chart. Most of the proton-rich nuclides, such as 64Ge, are even shorter-lived in contrast to those between A = 20 and A = 56, and quickly decay as the system approaches the hydrogen exhaustion point, whose occurrence can be anticipated from the integrated flow, as we see 15N(p, α)12C has started to close the CNO cycle. Around 103 s, hydrogen has been exhausted. The rapid decrease of hydrogen, in contrast to standard astrophysical scenarios, does not lead to an increasing in the amount of helium-4. Instead, 12C is synthesized, as well as other heavy elements such as 68Ge. At this point, those species at A > 56 initiate its β+ decay towards the valley of stability, as shown in both distribution of mass fractions for A = 70 in the lower right panel of Fig. 3.3 and in the integrated nuclear flow of Fig. 3.7. Given the exhaustion of hydrogen below 103 s, this chart can be useful to identify which reactions contribute with the largest portion of the rp-process. Clearly, the CNO breakout reactions play a critical role, as well as the further proton captures and the β+ decays. In Fig. 3.8, we plot the distribution of mass fractions per element (i.e. adding up the mass fractions of the individual isotopes of each element) at the beginning and at the end of the simulation. In the top panels, corresponding to our example with ρ = 105.5 g cm−3 and T = 108.7 K, we observe that in 105 seconds the system went from almost pure hydrogen to matter rich in Ge, Fe but also in C and Mg. In the lower panels, we appreciate how the synthesis changes as we reach a slightly higher temperature, typically found at the peak of the unstable thermonuclear burning. At these conditions, matter becomes deficient in C or Mg but is enriched in Ge, Zr and the elements between them. In both cases, no metals were present in the initial composition, thus it is via hydrogen and helium burning, as well as from 96 the adequate conditions of temperature, that we can synthesize, via the rp-process, heavy elements such as Ge or Zr. −10 −8 −6 −4 −2 0 lo g 10 X log10ρ = 5.5, log10T = 8.7 1H 4He 12C 68Ge 14O 15O 17F 18Ne 19Ne (XH,0, Xα,0) = (0.70, 0.30) 1H 4He 12C 68Ge 24Si + 25Si 29S + 30S 33Ar + 34Ar 37Ca + 38Ca −5 −3 −1 0 1 3 5 log10Time [s] −10 −8 −6 −4 −2 0 lo g 10 X 1H 4He 12C 68Ge 37Ca + 38Ca 58Cu + 59Cu 60Zn 63Ga 64Ge −5 −3 −1 0 1 3 5 log10Time [s] 1H 4He 12C 68Ge 70Br 70Se 70As 70Ge Figure 3.3: Mass fractions as a function of time for the rp-process example. This simulation employed the net344 network, and evolved the system at fixed tempera- ture and density. The initial conditions for the mass fractions are given in the title of the plot. 97 0 1 2 5 21 25 35 39 45 6 22 26 36 40 46 7 23 27 37 41 47 8 24 28 38 42 48 3 9 11 13 15 17 19 29 31 33 43 49 4 10 12 14 16 18 20 30 32 34 44 50 H(1) F(9) Cl(17) He(2) Ne(10) Ar(18) Li(3) C(6) Na(11) Si(14) Be(4) N(7) Mg(12) P(15) B(5) O(8) Al(13) S(16) K(19) Ca(20) Sc(21) Ti(22) V(23) Cr(24) Mn(25) Fe(26) Co(27) Ni(28) Cu(29) Zn(30) Ga(31) Ge(32) As(33) Se(34) Br(35) Kr(36) Rb(37) Sr(38) Y(39) Zr(40) Nb(41) Mo(42) Tc(43) Ru(44) Rh(45) log10 ρ = 5.5 log10 T = 8.7 ~Z = (XH,0, Xα,0) = (0.70, 0.30) ∆t = 1 s Figure 3.4: Time-integrated network flow at fixed ρ and T , corresponding to the rp-process example with net344, from t0 to 1 s. Red lines represent the reactions satisfying ♣Fij♣ ≥ 10−1♣Fmax♣, blue lines standing for 10−3♣Fmax♣ ≤ ♣Fij♣ ≤ 10−1♣Fmax♣. 98 0 1 2 5 21 25 35 39 45 6 22 26 36 40 46 7 23 27 37 41 47 8 24 28 38 42 48 3 9 11 13 15 17 19 29 31 33 43 49 4 10 12 14 16 18 20 30 32 34 44 50 H(1) F(9) Cl(17) He(2) Ne(10) Ar(18) Li(3) C(6) Na(11) Si(14) Be(4) N(7) Mg(12) P(15) B(5) O(8) Al(13) S(16) K(19) Ca(20) Sc(21) Ti(22) V(23) Cr(24) Mn(25) Fe(26) Co(27) Ni(28) Cu(29) Zn(30) Ga(31) Ge(32) As(33) Se(34) Br(35) Kr(36) Rb(37) Sr(38) Y(39) Zr(40) Nb(41) Mo(42) Tc(43) Ru(44) Rh(45) log10 ρ = 5.5 log10 T = 8.7 ~Z = (XH,0, Xα,0) = (0.70, 0.30) ∆t = 10 s Figure 3.5: Same caption as in Fig. 3.4, now for ∆t = 10 s. 99 0 1 2 5 21 25 35 39 45 6 22 26 36 40 46 7 23 27 37 41 47 8 24 28 38 42 48 3 9 11 13 15 17 19 29 31 33 43 49 4 10 12 14 16 18 20 30 32 34 44 50 H(1) F(9) Cl(17) He(2) Ne(10) Ar(18) Li(3) C(6) Na(11) Si(14) Be(4) N(7) Mg(12) P(15) B(5) O(8) Al(13) S(16) K(19) Ca(20) Sc(21) Ti(22) V(23) Cr(24) Mn(25) Fe(26) Co(27) Ni(28) Cu(29) Zn(30) Ga(31) Ge(32) As(33) Se(34) Br(35) Kr(36) Rb(37) Sr(38) Y(39) Zr(40) Nb(41) Mo(42) Tc(43) Ru(44) Rh(45) log10 ρ = 5.5 log10 T = 8.7 ~Z = (XH,0, Xα,0) = (0.70, 0.30) ∆t = 102 s Figure 3.6: Same caption as in Fig. 3.4, now for ∆t = 102 s. 100 0 1 2 5 21 25 35 39 45 6 22 26 36 40 46 7 23 27 37 41 47 8 24 28 38 42 48 3 9 11 13 15 17 19 29 31 33 43 49 4 10 12 14 16 18 20 30 32 34 44 50 H(1) F(9) Cl(17) He(2) Ne(10) Ar(18) Li(3) C(6) Na(11) Si(14) Be(4) N(7) Mg(12) P(15) B(5) O(8) Al(13) S(16) K(19) Ca(20) Sc(21) Ti(22) V(23) Cr(24) Mn(25) Fe(26) Co(27) Ni(28) Cu(29) Zn(30) Ga(31) Ge(32) As(33) Se(34) Br(35) Kr(36) Rb(37) Sr(38) Y(39) Zr(40) Nb(41) Mo(42) Tc(43) Ru(44) Rh(45) log10 ρ = 5.5 log10 T = 8.7 ~Z = (XH,0, Xα,0) = (0.70, 0.30) ∆t = 103 s Figure 3.7: Same caption as in Fig. 3.4, now for ∆t = 103 s. 101 He CO Mg Ca Fe Ge Zr 10−4 10−3 10−2 10−1 100 lo g 1 0 X t = 10−20 s log10ρ = 5.5, log10T = 8.7 He CO Mg Ca Fe Ge Zr 10−4 10−3 10−2 10−1 100 t = 105 s (XH,0, Xα,0) = (0.70, 0.30) He CO Mg Ca Fe Ge Zr 10−4 10−3 10−2 10−1 100 lo g 1 0 X t = 10−20 s log10ρ = 5.5, log10T = 9.0 He CO Mg Ca Fe Ge Zr 10−4 10−3 10−2 10−1 100 t = 105 s (XH,0, Xα,0) = (0.70, 0.30) Figure 3.8: Distribution of mass fractions per element at the beginning (left panel) and at the end (right panel) of the simulation of the rp-process with the net344 network, for the temperature and density conditions indicated in the title of the plot. Upper panels: example of the rp-process evolution, as described in the main text and in Figs. 3.3 and 3.4 - 3.7. Lower panels: evolution of the rp-process at a slightly larger temperature than for the upper panels. 102 Table 3.1: List of nuclides in the networks Approx140 and rp153. Approx140 rp153 Z A Z A Z A Z A N 1 V 44-46 N 1 Ca 36-44 H 1 Cr 45-48 H 1-3 Sc 39-45 He 4 Mn 48-50 He 3-4 Ti 40-47 C 12 Fe 49-52,54 Li 7 V 43-49 N 13-15 Co 52-56 Be 7-8 Cr 44-52 O 14-18 Ni 53-58,60 B 11 Mn 47-53 F 17-19 Cu 56-60 C 9,11,12 Fe 48-56 Ne 18-20 Zn 58-62,64 N 12-15 Co 51-56 Na 20-21 Ga 61-65 O 13-18 Ni 52-56 Mg 21,22,24 Ge 62-66,68 F 17-19 Al 23-25 As 66-69 Ne 18-21 Si 24-26,28 Se 68-70,72 Na 20-23 P 27-30 Br 70-73 Mg 21-25 S 28-32 Kr 72-74,76,80 Al 22-27 Cl 32-34 Rb 74-77,80 Si 24-30 Ar 33-36 Sr 76-78,80 P 26-31 K 36-38 Y 78-80 S 27-34 Ca 37-40 Zr 80 Cl 30-35 Sc 40-42 Ar 31-38 Ti 41-44 K 35-39 Nuclides for the rp-process endpoint (Excluded from net344) Z A Z A Z A Z A Tc 85 Pd 90-94 In 98-104 Te 107 Ru 86 Ag 94-98 Sn 99-105 Rh 89,91-93 Cd 95-99 Sb 106 103 Chapter 4 Neutron Star Envelopes “La libertad, Sancho, es uno de los más preciosos dones que a los hombres dieron los cielos[...]y, por el contrario, el cautiverio es el mayor mal que puede venir a los hombres.” M. de Cervantes Saavedra Abstract. This chapter discusses the evolution of neutron star envelopes in the presence and absence of accretion. After introducing the conditions under which the envelopes can be treated as time independent - i.e. the stationary approach - we analyze their characteristics and potential implications for the evolution of the rest of the neutron star. Introduction Modeling thermal evolution of isolated neutron stars requires the simultaneous solu- tion of the PDEs of structure and energy transport equations introduced in Chapter 1. The large variety of scales involved, both in the microphysical sector and on the spacetime coordinates, make this a challenging task as early works on the subject pointed out [NT81]. As stated in Chapter 1, for thermal evolution it is suitable to split the star into two regions, the envelope and the numerical core. In the absence of mass accretion, such splitting actually accelerates the numerical simulations: as first pointed out by [GPE83], since the envelope practically moves between steady- states of constant luminosity it is feasible to replace the evolution of the envelope by analytic or tabulated relations between the effective temperature at the surface and the temperature at the envelope-core boundary, i.e. Teff − Tb relations. 104 In the presence of mass accretion, on the other hand, we must consider the evolution of the mass fractions and, due to the onset of thermonuclear instabilities below 0.3ṀEdd (which produces drastic changes in the effective temperature within the interval of minutes), so far no attempt to produce an analogous Tb−Teff relation has been attempted. Another reason we could cite for the absence of such scheme is the usual assumption that once the envelope has ceased bursting the base luminosity becomes unique and positive, typically adopting values between 150 keV to 1 MeV per baryon. Observationally, there are systems for which we could pose into question that canonical picture: for instance, MAXI J0556-332 is presumed to have undergone accretion episodes while having a cold numerical core (e.g. Tb ∼ 106 K), and we could thus expect some heating to flow from the envelope towards the crust and core of the star. Naturally, the presence of heating sources in the right hand side of dL/dP demands to consider Lb(̸= Ls) as well. Consequently, the extension of the Tb(Teff) scheme in the presence of accretion requires it to become a Teff −Tb−Lb relation, and it is thus imperative to determine under which conditions we can actually construct and implement them into neutron star evolution codes. In the present chapter we review neutron star envelopes, both in the presence and in absence of mass accretion and within time dependent and independent scenarios, in order to determine the possibility to construct such a Teff − Tb − Lb relation at nonzero Ṁ . In Section 4.1 we review the general properties of the envelope. On Section 2 the case of envelopes without mass accretion is reviewed, and the conditions for the viability of the Tb(Teff) relations are stablished. In the next section, we analyze the envelope in the presence of mass accretion, first in time-dependent scenarios and then we move into the time-independent sector. 4.1 General features of the envelope By envelope we shall denote the section of the star between ρb = 108 g cm−3, i.e. the boundary with the numerical core, and the density at the surface ρs, set by the atmospheric boundary condition and with typical values of 10−1 to 10−3 g cm−3. Their main features can be enlisted employing the PDEs governing the structure and temperature derived in Chapter 1, as well as the thermodynamical functionals described in Chapter 2. • Mass and radius constancy. For a typical neutron star, we have gs ∼ GM/R2 ≈ 1014 cm s−2, with M and R the gravitational mass and radius of the star, respectively. Assuming hydrostatic equilibrium (Eq. 1.64) with ∆P ≈ ⟨ρ⟩gsdEnvelope, at ∆P = Pb − Ps ≈ 1024 erg cm−3 we have dEnvelope ∼ 10 m, i.e. ≈ 10−3R. Similarly, from the mass equation (Eq. 1.59), the enclosed mass in the envelope is Menvelope ≈ 4πR2dEnvelope ≈ 10−9M . Therefore, the thickness of the envelope and the enclosed mass are actually small with respect to the overall values of the star, and we can set m ≈ M and r ≈ R in the PDEs of 105 5 10 15 log10ρ [g cm−3] 0.55 0.60 0.65 0.70 0.75 0.80 ex p [Φ (ρ )] C ru st -C or e T ra n si ti on MS-A1 + PC, ρc,15 = 1.9152 APR + PC, ρc,15 = 1.0037 5 10 15 log10ρ [g cm−3] 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 m (ρ ) [M ⊙ ] C ru st -C or e T ra n si ti on MS-A1 + PC, ρc,15 = 1.9152 APR + PC, ρc,15 = 1.0037 Figure 4.1: Metric functions in terms of local density for two representative stars of the indicated core EOS. The transition between the core and the crust is indicated by vertical lines. stellar evolution at the envelope. In the presence of mass accretion, the validity of these approximations depends on the time interval of interest. For instance, at a typical accretion rate of Ṁ ≈ 0.1ṀEdd ≈ 1.1 × 1017 g s−1, the aggregated baryonic mass during one year is 3.5×1024 g ≈ 1.25×10−9M (at M = 1.4M⊙), i.e. the mass of the envelope is practically duplicated, yet remains small with respect to M . The constancy of R might be put into question considering the observational evidence for the photospheric radius expansion during the thermonuclear explosions. Typically, this changes the observed radius from ∼ 12 km to ∼ 19 km during a few seconds around the burst peak. From Chapter 1 we know low-density environments favor fluid motion, hence the outer layers of the envelope are displaced during this expansion and thus r cannot be approximated by R anymore, although for the innermost layers it is still viable to consider r ≈ R. Furthermore, since the expansion takes place during a few seconds only, for long timespans (hours to years) r ≈ R remains a good approximation for modeling envelopes. An interesting discrepancy between numerical simulations and observations is the absence of a prominent increase in radius despite the good agreement in the rest of characteristics of the bursts (peak luminosity, recurrence time, cooling phase after peak). For instance, at MESA we might go from r ∼ 11.53 km in the inter-bursting lapse to r ∼ 11.60 km at burst peak. • Constancy of the metric functions Λ(r), Φ(r). Due to the explicit relation between m(r) and Λ(r) (Eq. 1.56), the constancy of Λ is a corollary from the previous point, i.e. eΛ(r,m) ≈ eΛ(R,M). The same constancy in the envelope 106 holds for Φ(r,m) due to eΦ(R,M) = √ 1 − 2GM c2R . Moreover, since Λ(R,M) = −Φ(R,M), the combination of both results leads to the simplification that at the envelope it is viable to replace all eΛ by e−Φ or viceversa, i.e. general relativistic effects can be simply accounted by the introduction of additional redshift factors. The constancy of the metric functions near the surface of the star is nicely illustrated in Fig. 4.1, where for two different core-EOS, albeit with same crust-EOS, both eΦ and eΛ are essentially constant below 1010 g cm−3. Such result should not differ had we used another crust-EOS since between 104 and 1010 g cm−3, despite variations in the actual chemical composition, pressure and density are related via electron’s degenerate EOS. • Alternative spatial coordinates. Since variations in mass and radius are actually small throughout the envelope, as discussed above, neither of them is suitable as spatial coordinate. Reasonable alternatives are the rel- ative mass ln q1 = ln(MB/MB, TOT) or its modified version ln q2 = [(MB − MB, TOT)/MB, TOT], which have been employed in time-dependent codes for simulating stellar evolution in the presence of mass accretion with good re- sults. Another option for spatial coordinate is the pressure due to its promi- nent variations, from Pb ∼ 1025 erg cm−3 to Ps ∼ 1014 erg cm−3, as well as its continuous character. A variation of this choice, considering the constancy of M and R, is the column density coordinate y := P/gs, which allows to absorb geometrical factors in the PDEs. While variations in ρB are of 9 orders of magnitude as well, discontinuities due to changes in composition make it unsuitable as spatial coordinate. The above considerations allow for several simplifications in the discussion of stellar evolution equations at the envelope. These are useful for numerical implementations as well, albeit taking into consideration the limitations of these approximations. Thus, in order to proceed with our discussion let us first introduce the following notation: (Local mass accretion per unit surface) ṁ = Ṁ∞e−Φ 4πR2 (4.1) (Local surface gravity) gs = GMe−Φ R2 (4.2) (Proper time operator) ∂τ = e−Φ∂t (4.3) (Radial to pressure coordinate) eΦ∂r = −ρBgs∂P (4.4) (Eulerian operator for pressure) Dt = ∂τ + ṁgs∂P (4.5) (Local flux) F = −KeΦ∂rT = KρBgs∂PT (4.6) (Local luminosity) L = 4πR2F. (4.7) As a result of all these considerations, the evolution of the envelope is fundamen- tally governed by the temperature and the mass fractions for the Nspecies, according 107 to cP ∂T ∂τ = ρB ( ˙̆εnuc − ˙̆εν ) + cPT P ṁgs(∇ad − ∇T ) + ρBgs ∂F ∂P (4.8) ∂Xi ∂τ = −ṁgs ∂Xi ∂P + AiRi, i ∈ ¶1, . . . , Nspecies♢ . (4.9) By inspection of Eq. 4.8, we see temperature is governed by the interplay among three fundamental timescales: • Heat diffusion. This scale is associated with the amount of time required for radiative and conductive fluxes to transport thermal energy. Consider- ing Eq. 4.8, as well as the transformation law between P and r (hydrostatic equilibrium), this time scale can be expressed as τdiff := d2cP K . (4.10) Here, cP is the heat capacity at constant pressure per unit volume, K is the thermal conductivity, and d is the thickness of the layer of material under consideration. At ρb, for instance, considering Kb ≈ 1016 erg cm−1 s−1 K−1, cP ≈ 1014 erg cm−3 K−1 and d = dEnvelope, we get τdiff ≈ 104 s, i.e. a few hours only. As illustrated in Fig. 4.3 for different envelope models, we see this back-of-the-envelope estimation is independent of the chemical composition chosen, as well as of gs. Notice as we move from 108 g cm−3 to 1010 g cm−3, the diffusion timescale moves from days to years. • Mass accretion. Typically defined as τacc = δM/Ṁ , this scale quantifies the required amount of time to accrete a layer of mass δM at an accretion rate Ṁ . For instance, to completely replace the envelope, i.e. δM = 10−9M , at the Eddington rate ṀEdd ∼ 1018 g s−1 we must wait approximately 20 days. From Eq. 4.8, we see hydrostatic equilibrium allows to consider an alternative definition for this timescale, τacc := P ṁgs , (4.11) which serves as a local indicator of the required amount of time for accreted matter to reach a certain pressure in the envelope. For instance, at Pb ≈ 1025 erg cm−3, gs ∼ 1014 cm s−2 and ṁ ≈ 105 g cm−2 s−1, we obtain τacc ≈ 11 days. • Nuclear reactions. Assuming the total amount of energy generated by nu- clear reactions is Enuc (with units of erg g−1), it is customary to associate τnuc := Enuc ε̇nuc as the characteristic timescale for nuclear reactions at fixed den- sity and temperature (via ε̇nuc). For instance, H burning via CNO cycles yields τnuc ∼ 103 s, considering ˙̆εCNO ≈ 1015 erg g−1 s−1 and Enuc ≈ 1018 erg g−1. 108 Typically, this occurs at around ρ ∼ 106 g cm−3. On the other hand, Eq. 4.8 suggests to consider τnuc := cPT/ρB ˙̆εnuc (4.12) as an alternative definition for the nuclear timescale. As we shall illustrate below, this definition emerges in the context of thermonuclear instabilities, and we shall thus adopt it. 4.2 Evolution of non-accreting envelopes The main characteristic of these envelopes, as their name suggest, is the absence of mass increase. This state can be steady, as in the case of isolated neutron stars, or transient, as for instance during quiescent phases of binary systems. Without com- pression, the chemical composition is expected to undergo a sedimentation process, with high-Z elements displacing towards high-density layers and light elements such as H and He remaining at low densities. For neutron stars in binary systems, this sedimentation displaces the ashes from nuclear burning toward high-density layers inside the envelope. In the case of isolated stars, such sedimentation might be ex- pected during the first years after birth while at latter epochs it is reasonable to consider these layers of well-defined ⟨Z⟩ have already been settled and will remain undisturbed. In the absence of accretion, secular nuclear burning is viable as long as fuel is present and adequate conditions of temperature and density for starting ignition still exist. Keeping those conditions, however, becomes challenging and unlikely as t > 0. For instance, at T ∼ 108.25 K and ρ ≥ 104 g cm−3, 1H burning via hot CNO cycle leads to complete exhaustion of fuel in ∼ 104 s ≈ 0.11 day. Without 1H replenishment, the ashes remain inert to nuclear burning and thus heating is no longer injected at this density. Thus, any remaining fuel for thermonuclear reactions should be eventually exhausted and ε̇nuc → 0 in less than 1 day. On the other hand, electroweak reactions operating at ρ ≤ 106 g cm−3 are still likely to operate in the absence of accretion, although in the long term (i.e. timescale of years), such heating injection completely ceases. What about neutrinos? Inferred effective temperatures from isolated neutron stars, as well as during quiescent phases at binary systems, yield Teff ∼ 5 × 106 K, i.e. L ∼ 1035 erg s−1. On the other hand, thermal neutrinos at the surface contribute with Lν ≈ ε̇νMenvelope ≤ 1033 erg s−1, i.e. their contribution to the total luminosity is significantly small and their presence in the envelope can thus be neglected. We thus see the evolution of non-accreting envelopes is governed by heat diffu- sion. From Eq. 4.8, the temperature profile evolves according to a diffusion process with non-constant K, cP ∂T ∂τ ≈ ρBg 2 s ∂ ∂P ( KρB ∂T ∂P ) . (4.13) 109 Given the smallness of τdiff, i.e. < 1 day, it turns out any stationary solution to Eq. 4.13, upon setting the appropriate boundary conditions, is a steady-state so- lution. The proof relies on linear perturbation theory, where it is shown thermal perturbations decay as e−t/τdiff . In other words, non-accreting envelopes are stable against thermal perturbations. Furthermore, this stability is the mile- stone for simplifying the analysis of non-accreting envelopes, as summarized in the following statement: Steady-state approximation (SSA). In the absence of accretion and nuclear burning, thermal evolution of the envelope proceeds among stationary states of constant luminosity. 0.0 2.5 5.0 7.5 10.0 log10 ρ [g cm−3] 6.0 6.5 7.0 7.5 8.0 lo g 1 0 T [K ] ρb = 1010 g cm−3 gs,14 = 1.73 Teff = 106 K 56Fe 12C 1H 4He 0.0 2.5 5.0 7.5 10.0 log10 ρ [g cm−3] 6.0 6.5 7.0 7.5 8.0 gs,14 = 1.26 gs,14 = 3.74 56Fe 1H ρb = 1010 g cm−3 Teff = 106 K Figure 4.2: Temperature as a function of density for a selected collection of envelopes in steady state. Left panel: envelopes at fixed gs,14 = gs/1014 cm s−2 composed of a single species. Right panel: envelopes of the indicated compositions with different values of gs,14. Such result, first noticed by [GPE83], introduced a new paradigm in the modeling of neutron star cooling. Indeed: combining the constancy of L = 4πR2T 4 eff across the envelope with the approximations over mass and radius, temperature profile obeys dT dP = 3κ 16T 3 T 4 eff gs . (4.14) Since structure remains unchanged with time, for a given Teff we obtain a unique T (P ) regardless of the time at which Teff occurs. Therefore, to simplify the simu- lations at all times [GPE83] proposed to replace the envelope’s evolution, via the 110 0 2 4 6 8 10 log10 ρ [g cm −3] 0 1 2 3 4 5 6 7 lo g 1 0 τ [s ] 1 day ρb = 1010 g cm−3 gs,14 = 1.73 Teff = 106 K 1H 4He 12C 56Fe 0 2 4 6 8 10 log10 ρ [g cm −3] 0 1 2 3 4 5 6 7 ρb = 1010 g cm−3 Teff = 106 K gs,14 = 1.26 gs,14 = 3.74 56Fe 1H 1 day Figure 4.3: Heat diffusion timescale for the envelope models in Fig. 4.2. explicit solution of the corresponding PDEs of structure and temperature, by Tb−Teff relations, in a sense mimicking the procedure of employing pre-calculated tables or analytic fits as atmosphere boundary conditions. The recipe to construct Tb − Teff relations goes as follows: 1. Fix a set of values for Teff, i.e. Teff. Typically, in the absence of accretion it is usual to consider Teff = [104, 107] K. 2. For each Teff ∈ Teff, solve Eq. 4.14 in [Ps, Pb] and obtain a unique Tb. 3. Use the resulting data, in either tabular or fitting form, as surface boundary condition for the numerical core at all times. Within the Tb − Teff scheme, what is the composition of the envelope? At the time the GPE approximation was proposed [GPE83], neutron star envelopes were thought to be constituted of pure 56Fe, given its stability under β+ decays and its role as the ground state of matter, at T → 0, below 1010 g cm−3 (e.g. [BPS71]). A modern picture of the composition, taking into consideration the composition after accretion episodes and the sedimentation process, consists on separating the envelope into layers of well-defined composition, i.e. an onion-like structure (e.g. Potekhin and Yakovlev, PY [PCY97]). Quite often the amount of layers employed range from 1 to 3, so in general we can write the mass and proton numbers, as well 111 as their correspondent mass fractions as A(P ) = 4 ∑ i=1 Aiθ(P − Pi)θ(Pi+1 − P ) (4.15) Z(P ) = 4 ∑ i=1 Ziθ(P − Pi)θ(Pi+1 − P ) (4.16) X(P ) = 4 ∑ i=1 Xiθ(P − Pi)θ(Pi+1 − P ) , (4.17) where i = 4 corresponds to Pb, θ(x) is the Heaviside function and each Ui is a constant vector in R Nion . Albeit modern and reliable for modeling non-accreting envelopes, it is still an approximation which leaves several parameters unconstrained, as for instance • The transition depths Pi (i ∈ ¶1, . . .♢) at which jumps from low- to high-Z matter take place. A good estimation for the 1H/ashes boundary, for example, comes from thermonuclear burning simulations and yields Pthreshold ≈ 1022.5 erg cm−3. In contrast, a 4He/ashes boundary is considerably more challenging since the amount of 4He depends on the history of mass accretion (since at stationary conditions helium-4 survives but during thermonuclear explosions it is completely destroyed) and the maximum internal temperature reached (due to the sensitivity of the 3α reaction with temperature). • The distribution of metals in a layer. Although we can artificially set X ≈ 1 for a given species, actual simulations of explosive nuclear burning indicate the ashes are actually a mixture of metals, neither of which possess X ≈ 1. Since different mixtures of metals yield different values for the opacity and the specific heat, the composition of the layers we impose by hand should be based on the results from time-dependent simulations of accreting episodes. On Fig. 4.2 we illustrate a collection of temperature profiles for envelopes with same Teff but different composition and surface gravity. For these models, we adopted the predicted values for masses and radii considering the APR-core EOS and the TOV equations of structure. As predicted by Eq. 4.14 and exemplified on the right panel of Fig. 4.2, at fixed composition we see large values of gs translate into small temperature gradients and, consequently, into low values for Tb. Here, gs,14 = 3.74 corresponds to a 2M⊙ APR-EOS star and gs,14 = 1.26 to a M⊙ star with the same core-EOS. In this same panel, we see a low-Z envelope is considerably colder than one with high-Z, a direct consequence of having dT/dP ∝ κ since κ ∝ Z2. This behavior is also exemplified, at a gs corresponding to an 1.4M⊙ APR-EOS star, in the left panel of Fig. 4.2. Here we see an approximate progression of slopes in the temperature profiles as we move from 4He (Z = 2) to 12C (Z = 6) and finally to 56Fe 112 (Z = 26). Artificial behavior of the pure-1H envelope might be anticipated since we are placing hydrogen at unlikely depths, according to the nuclear burning and sedimentation models. On Fig. 4.3 we illustrate the behavior of τdiff as a function of density for the same steady-state envelopes in Fig. 4.2. Since dEnvelope ∝ P/ρBgs, the curves on the right panel of this figure exhibits a similar tendency as the temperature profiles of Fig. 4.2, i.e. high-gravity environments lead to both small τdiff and small T (ρ). This also implies high-gravity environments need less time to dissipate a thermal perturbation than low-gravity ones. A similar conclusion arises by changing chemical composition, as shown in the envelope models in the left panel of Fig. 4.3: a high-Z (metallic) envelope requires more time to dissipate a thermal perturbation than a low-Z one, 1H representing an exception to this rule at high densities due to the unphysical situation of pushing this element far beyond its exhaustion point by thermonuclear burning. By inspection of both panels, we are led to the universal conclusion that treating thermal evolution of neutron star envelopes as the succession of steady-states is an excellent approximation as long as heating or cooling sources are absent, or as long as such processes occur in a timescale much shorter than the local τ [BC09]. Following the steps in the recipe 4.2, we constructed several Tb - Teff relations with different chemical composition, which can be seen in Fig, 4.4. For simplicity, we adopted the surface gravity corresponding to a 2M⊙ APR-core EOS, and adopted ρb = 1010 g cm−3. The model with pure 56Fe corresponds to the GPE model, while the mixtures follow the layered description of PY (e.g. Eqs. 4.15-4.17). The impact of chemical composition on the Tb−Teff relations is quite noticeable: for a fixed value of Tb, envelopes of pure 12C have a larger effective temperature than those with pure 56Fe. Similarly, replacing carbon-12 with helium-4 leads to envelopes with even higher effective temperatures than those with pure iron-56. Physically, this means that if we have two neutron stars with same core temperature, the star with low metallicity at the surface would look brighter than the star with high metallicity at the envelope. On the other hand, at a fixed Teff we see high-Z envelopes have higher Tb than those with low-Z composition, i.e. the numerical core remains hotter as a consequence of an increased opacity at the surface (recall κ ∝ Z2). Conversely, we do not require a hot core to have a bright object, only to have a low metalicity at the envelope. Having discussed the single-species Tb − Teff relations, let us inspect the impact of having two layers, i.e. the bounded, dashed or dotted curves in Fig. 4.4. For these curves, the transition between the low and high Z species takes place at ρx = 10x g cm−3. The first noticeable change is that the lower ρx is the closest the two-layered envelopes’ Tb − Teff relation remains to the curve corresponding to the highest-Z envelopes. For instance, in the left panel of Fig. 4.4, at ρ4 the Tb − Teff relation for the two-layered envelope approaches the curve corresponding to the pure carbon-12 at around Teff ∼ 105.9 K. On the other hand, at ρ6 and ρ8 the proximity of the two-layered envelope’s relation with the pure helium-4 curve is more prominent 113 than with the pure carbon-12 curve. By comparison with the right panel, we see the higher the jump of Z in the two-layered envelope is, the closest its Tb − Teff relation is with respect to the high-Z relation. Physically, this implies two neutron stars of identical core temperature Tb ∼ 108 will look brighter if the surface is composed of pure carbon-12 than an envelope with a mixture of carbon-12 and iron-56 and, similarly, such stars would be brighter than one with a pure iron-56 envelope. 5.6 5.8 6.0 6.2 6.4 log10 Teff [K] 7.0 7.5 8.0 8.5 lo g 1 0 T b [K ] gs = 3.23×1014 cm s−2 ρb = 1010 g cm−3 12C 4He to 12C at ρ4 4He to 12C at ρ8 4He to 12C at ρ6 4He 5.6 5.8 6.0 6.2 6.4 log10 Teff [K] 7.0 7.5 8.0 8.5 9.0 56Fe 56Fe to 12C at ρ4 56Fe to 12C at ρ6 56Fe to 12C at ρ8 12C Figure 4.4: Tb − Teff relations for representative composition configurations. 4.3 Evolution of accreting envelopes In the presence of low-Z rich matter accretion, conditions become viable for the onset of thermonuclear burning via the processes discussed in the previous chap- ter. In principle, the evolution is governed by all terms in the right-hand side of Eq. 4.8 and Eq. 4.9. The physics of the system is simple: matter is accreted at a rate ṁ, temperature increases at the surface, allowing for thermonuclear burning to transform the low-Z fuel into high-Z ashes, which accumulate upon the pris- tine surface and, under constant accretion for t ∼ months, both ashes and pristine material are compressed and ignite via pycnonuclear reactions. Depending on ṁ, from observations we know thermonuclear burning takes place with or without pe- riodic explosions (short-termed increase of temperature and luminosity). Despite the qualitative simplicity of this picture, for its quantitative side we must solve the stellar evolution equations in the envelope together with the evolution of the nu- merical core. To model transient accretion episodes, such task is complicated due to the large difference in the cooling timescale: while at the surface we have τdiff ∼ seconds, at the numerical core we have τdiff ∼ yr, i.e. we must either take small time-steps to resolve the bursts at Ṁ ≤ 0.3ṀEdd or we must find a way to simulate 114 the temperature and luminosity of the envelope while correctly accounting for their evolution at the core-envelope boundary, which is no longer Teff = Ts due to the heating sources. A priori, we should restore to fully time-dependent simulations to answer the former questions. However, to fully grasp the physics involved in their output it is suitable to divide the discussion of accreting envelopes into steps. First, we look at the equilibrium points of the system, i.e. solutions of stellar equations at the envelope setting all ∂t terms to zero. Following the standard theory of ODEs and PDEs, the equilibrium points are useful to understand the dynamics of the system. In the next step, we look at the evolution of perturbations to equilibrium curves. As we shall see, they enclose the core aspects of the observed bursts. Finally, we discuss the numerical output from fully-time dependent simulations, and the implications for the long term evolution of the whole envelope and the characteristics an acceptable Tb − Teff − Lb relation in the presence of accretion must have. 4.3.1 Stationary accreted envelopes By setting all ∂t terms to zero, the envelope is described by a system of first-order ODE in the spatial variable - the pressure, for numerical convenience - and sim- plifies into a BVP. Moreover, the redshifted accretion rate Ṁ∞ becomes constant throughout the whole envelope. The resulting equations are: dr dP = − 1 geΛHG (4.18) dm dP = 4πr2ρ dr dP (4.19) dΦ dP = − 1 Hc2 (4.20) dT dP = T P ∇T (4.21) ∇T = − 3κρBLPe Λ 64πσSBT 4r2 dr dP + ( 1 − ρ H ) (4.22) dXi dP = 4πr2ρB Ṁ∞e−ΦgHGAiRi, i ∈ ¶1, . . . , Nspecies♢ (4.23) dL dP = 2L c2H + 4πr2ρBe Λ dr dP ( ˙̆εnuc + ˙̆εgrav,h − ˙̆εν ) (4.24) ˙̆εgrav,h = −Ṁ∞e−Φ−Λc̆PT 4πr2ρBP dP dr (∇ad − ∇T ) (4.25) where, following [FHM81, PMS+15], we have denoted as homogeneous (subindex h) the contribution of heating due to mass compression. 115 Consideringthat composition of accreted material for LMXB is known (i.e. Z ≈ Z⊙), we opt for specifying the boundary conditions (now turned into initial condi- tions in P ) for Eqs. 4.18 - 4.25 at the surface, which we set at the optical depth τs = 2/3. This allows us to have freedom on choosing an arbitrary - yet physically reasonable - value for Teff. In what follows, we restrict this parameter to the range [106, 107] K. Regarding the rest of surface values: we mostly employ the network with Nspecies = 380 (see Chapter 3), and unless otherwise stated, we take Z = Z⊙, X1H = 0.70 and X4He = 0.28 as the surface composition. The gravitational mass ms = M and radius rs = R are specified according to the standard APR-Core EOS M −R relation (see Chapter 1, specifically Fig. 1.3), which defines Φs as well (recall eΦs(M,R) = √ 1 − 2GM c2R ). Throughout the Eddington relation (Chapter 1), provided τs, Teff, gs and Xs we obtain Ts, as well as Ps and ρs by a combination of bisection and Newton-Raphson methods. From R and Teff, the surface luminosity is obtained via Ls = 4πR2σSBT 4 eff. Given these initial conditions, Eqs. 4.18 - 4.25 are solved in Psb := [Ps, Pb], with Pb corresponding to the pressure at which ρb = 107 g cm−3. The constant Ṁ∞ is expressed in units of ṀEdd = 1.1 × 1018 g s−1. Let us emphasize that, contrary to most - if not, all - approaches to this subject, we are neither fixing nor shooting for a single, specific value of Lb, unless otherwise stated. Given the nature of Eqs. 4.18 - 4.25 as IVP in P , doing so would imply aiming for a single, very concrete Teff yielding the desired Lb. Instead, our main goal is to analyzing the temperature, luminosity and composition profiles when we take a collection of effective temperatures Teff, i.e. their corresponding Lb’s are determined by the integration of Eqs. 4.18 - 4.25. Nuclear output at different accretion rates Let us first explore the output from full nuclear burning. Within the classical paradigm, this requires to consider a single model with Lb > 0: we opt to fol- low this approach, albeit we only demand an Lb sufficiently high as to produce an envelope profile of almost flat slope near ρb. By performing a manual shooting method in order to find such acceptable Teff to produce the desired positive base luminosity, the stationary equations were integrated for gs = 1.735, corresponding to a 1.4 M⊙ APR-Core EOS for four accretion rates: 0.01, 0.10, 1.0 and 5.0 ṀEdd. In Figs. 4.5, 4.6 and 4.7 we show three panels - functions of the local density - per accretion rate: in the upper one, we plot the temperature profile, indicating at the headers their corresponding accretion timescale (e.g. Eqs. 4.11) and column depth y. In the middle panel a collection of mass fractions for relevant species are shown, and in the lower panel the specific energy generation rate due to the whole and particular nuclear processes are plotted. Let us emphasize our choice of Lb (or its corresponding Teff) is only a requirement for an adequate comparison with existent models in the literature: we could also have started with Lb < 0, albeit such models 116 are reserved for the next subsection. At 0.01ṀEdd, left panels of Fig. 4.5, we observe hydrogen being transformed into 4He via the cold and hot CNO cycles as we move from log10 ρ = 3 to ∼ 5.6 (in g cm−3 at panel (c)). The shape of the specific energy generation rate is very distinguishable: at first, we observe a rapid increase in the energy production (between 3.0 and 4.0, in log10 ρ scale), sign of the presence of 12C(p, γ)13N, followed by a constant region of ∼ 1014 erg g−1 s−1, consequence of the β+ decays of 14−15O. Despite the tiny fractions of accreted A ≥ 40 isotopes, the temperature is still not sufficiently high (T < 108.25 K) as to allow proton captures over these nuclides, although we observe some captures between A ≥ 20 and A < 40 in panel (c), which are closely followed by the β+ decays due to the short lives of these proton-rich nuclides, as also evidenced by the rapid decrease in the Tz = −1 mass fractions in panel b. At log10 ρ = 6.25 g cm−3 helium-4 is transformed into carbon-12 via the 3α process (secondary peak in panel (c)), while α-captures occur in a minor scale, thus leading to high amounts of carbon-12 near ρb. At a ten times large accretion rate, 0.1ṀEdd, hydrogen is still transformed into 4He via CNO cycles near log10 ρ ∼ 5.5 g cm−3, although the resulting layer occupies a smaller region in the density space than its 0.01ṀEdd analogue. As shown in panels (a)/(a’) and (c)/(c’) of Fig. 4.5, this is due to the large temperature at the envelope - proportional with the accretion rate - which enhances the helium-4 burning via the 3α reaction. While its specific energy generation rate does not exceed that from the hot CNO cycle, its influence extends from log10 ρ ∼ 4.5 to 6.5 g cm−3, turning carbon-12 as one of the most prominent nuclides in the composition of the ashes. At 0.1ṀEdd we start observing the presence of proton captures over A ≥ 40 nuclides, closely followed by subsequent β+ decays. In contradistinction with the 0.01ṀEdd case, we now observe that nuclear energy production does not abruptly decay after the hydrogen exhaustion, a direct consequence of different processes, as for instance 3α (burning helium-4) or β+ decays. At slightly higher accretion rates, 0.25 and 0.35 ṀEdd in Fig. 4.6, we do not see a significant production of helium-4 right at the hydrogen exhaustion point. Instead, we see helium-4 already being burned via 3α (panels (c) and (c’)) due to T ∼ 108.5 K (panels (a) and (a’)). Moreover, we now identify an helium-4 exhaustion point near log10 ρ ∼ 6.75 g cm−3. By inspection of the accreting timescale at the top of Fig. 4.6, hydrogen is completely burnt in nearly a few hours while helium should be completely burnt in approximately 1 day. Similarly, the associated column depth differs in only an order of magnitude, i.e. y ∼ 108 at hydrogen exhaustion while y ∼ 109 at helium exhaustion. Regarding the rp-process, in the specific energy generation rate we observe a richer structure as a consequence of T ≥ 108.25 K. In first place, although the CNO cycle is still dominating the nuclear energy between log10 ρ = 3 and 5 g cm−3, proton captures over 20 < A < 40 species enrich the A = 40 sector. At these densities, however, temperature is still not high enough as to allow proton captures over A > 40 nuclides. We can also argue the occurrence of 117 the CNO cycles block the possibility to synthesize further species. However, once log10 ρ = 5 and T ∼ 108.4 K is reached, CNO cycles are no longer the primary source of energy and are overcome by proton captures over A ≥ 20. This, we infer, is a consequence of the 19Ne(p, γ)20Na(p, γ)21Mg chain, occurring right after the 15O(α, γ)19Ne CNO-breakout reaction. Once started, we see hydrogen progressively starting to decrease while proton captures rapidly occur in the A ≥ 40 sector, and in a minor scale over A ≥ 60 species. The Tz = −1 nuclides, auxiliary in these chains of proton captures and subsequent β+ decays, eventually reach a mass fraction comparable to that of the A ≥ 40 nuclides and enrich them up to the hydrogen exhaustion point, at log10 ρ ∼ 5.75 g cm−3. It is noticeable the maximum of nuclear energy due to the rp-process is an order of magnitude larger than the production due to CNO, i.e. 1016 erg g−1 s−1. At the hydrogen exhaustion point we observe a drastic decrease in the proton capture reactions and the predominance of β+ decays to the contribution of nuclear energy. Such tendency is accentuated as we move from 0.25 to 0.35 ṀEdd, where the synthesis of A ≥ 60 nuclides is followed by the subsequent decays towards the valley of stability, as suggested by panels (c) and (c’), where in the latter we appreciate the β+ decays over A ≥ 20 nuclides practically provide the majority of nuclear energy. Finally, let us analyze the mass fractions and nuclear energy output at 1 and 5 ṀEdd, depicted in the eight panels of Fig. 4.7. At the Eddington rate, practically no helium-4 is synthesized as a consequence of hydrogen burning, which leads to the small contribution to the nuclear energy from the 3α process in comparison with the β+ decays from A ≥ 20 nuclides. Similarly, the hot CNO cycle is no longer the main contributor to the energy output between log10 ρ = 3.0 and 5.0. Instead, proton captures over A ≥ 20 species and their β+ decays are the main contributors to the energy output. Notice, however, that only 12C is exhausted as a consequence of these captures, with little enhancement of 40Ca but increased amounts of 15O and 17F, an evidence in favor that CNO cycles are less energetic but they still control, between log10 ρ = 4.0 and 5.0, the production of nuclear species. It is only when temperature reaches and exceeds 108.5 K and the density is ∼ 105 g cm−3 that CNO cycles are finally broken and a cascade of proton captures - almost immediately followed by β+ decays - lead to hydrogen exhaustion near 105.75 g cm−3. Counter- intuitively, such process leads to the enhanced production of 12C (panel (b)). The explanation, however, relies on the interplay between hydrogen and helium burning: the rupture of the CNO cycles implies that whatever amounts of produced 12C - via 3α reaction - is immediately converted into 14O and reaches 19Ne, 20Na and beyond via proton captures. However, when the system starts running out of hydrogen, yet possesses sufficient amounts of 4He, 12C(p, γ)13N is no longer efficient and carbon-12 starts accumulating. As T ≥ 108.6 K, the remaining helium-4 is then transformed into carbon-12 or 16O via α captures, inducing an helium-exhaustion point around 106.5 g cm−3. As a consequence of the rp-process, nuclides in the range A ≥ 60 are produced and practically dominate the composition of the ashes. The long- 118 lived isotopes under β+ decays, above 105.75 g cm−3, start decaying and produce the largest contribution to the specific energy generation rate. The scenario at 5ṀEdd shares many aspects with that at ṀEdd, such as hydrogen and helium-4 exhaustion points, the generation of considerable amounts of 12C at the high-density sector and the largest contribution to the nuclear energy - in the aforementioned sector - coming from the β+ decays over A ≥ 20 nuclides. As differences we can cite the maximum temperature at the envelope, ∼ 108.8 K at 5ṀEdd against the 108.75 K at ṀEdd, the enhanced production of A ≥ 70 species instead of A ≥ 60, and the synthesis of 40Ca, together with the bypass of the bottleneck near 43Ti, at lower densities in comparison to the ṀEdd case. At fixed gs: different accretion rates. Having acquired a first insight to the nuclear processes in the envelope, let us now explore the impact of changing the effective temperature at the surface, an appropri- ate scenario for our final goal. Taking Teff = [106, 107] K, the stationary equations were solved for gs = 1.735, corresponding to a 1.4 M⊙ APR-Core EOS, for four accretion rates: 0.03, 0.10, 0.30 and 0.60 ṀEdd. In Figs. 4.9 and 4.10 we display four panels for each accretion rate. In the upper left panel, we show the resulting temperature profiles for each effective temperature in Teff. In the lower left panel, we display the specific energy generation rate due to nuclear reactions for four selected models, as well as the generation rate from accretion and neutrinos for a model with Lb > 0. In the upper right panel we have the luminosity at the base of the envelope (expressed in units of MeV per baryon), for each of the profiles in the upper left panel, against the surface temperature. Finally, in the lower right panel we display the phase space, i.e. T against dT/dP , for each temperature profile. The first and most notorious characteristic of the T vs ρ panel, at any of the selected accretion rate, is not all numerical solutions qualify as “complete” or phys- ical envelopes: there exists a subset T − ⊂ Teff whose corresponding temperature profiles reach a maximum value at ρ < ρb and then start decreasing inwards. To proceed with the numerical integration, all these models require T → 0. Although we could opt for setting a minimum temperature to proceed the integration, for instance Tmin = 104 K, there is not a physical justification for keeping a constant T until Tb. To prevent ambiguities, these models were simply cut at ρcut < ρb such that d log10 T/d log10 ρ = −101. Mathematically, on the other hand, the T → 0 behavior is consistent with the input: notice dL/dP (Eq. 4.24) is controlled by two heating sources (nuclear burning and mass accretion) and only one cooling mechanism, neu- trino emission. Since ˙̆εν ≪ ˙̆εnuc, whenever L becomes negative the stationary profile does not have additional mechanisms to return to L > 0 at higher densities. Due to ∇T ∝ L, we see dT/dP is and remains negative. The presence of convection in 1In what follows, we refer to those quantities at ρcut by the subindex cut, i.e. Tcut, Lcut, etc. 119 7.50 7.75 8.00 8.25 8.50 8.75 9.00 lo g 1 0 T [K ] log10y [g cm−2] 5.0 6.0 7.0 8.0 9.0 Time [y/ṁ] 1 min 10 min 1 h 10 h 1 day Teff = 5.200× 106 K −10 −8 −6 −4 −2 0 lo g 10 X 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] 8 9 10 11 12 13 14 15 16 17 lo g 1 0 ε̇ n u c [e rg s− 1 g− 1 ] Hot CNO (α, p) (α, γ) ε̇3α A ≥ 20 A ≥ 40 A ≥ 60 A ≥ 20 A ≥ 40 A ≥ 60 (p, γ) (e+νe)} } 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] 8 9 10 11 12 13 14 15 16 17 lo g 1 0 ε̇ n u c [e rg s− 1 g− 1 ] Hot CNO (α, p) (α, γ) ε̇3α A ≥ 20 A ≥ 40 A ≥ 60 A ≥ 20 A ≥ 40 A ≥ 60 (p, γ) (e+νe)} } b c b' c' 0.01 · MEdd 0.10 · MEdd CNO Tz=-1 H 4He A=20-39 A=40-59 H 4He CNO A=20-39 Tz =-1 A=40-59 A ≥60 7.50 7.75 8.00 8.25 8.50 8.75 9.00 lo g 10 T [K ] log10y [g cm−2] 5.0 6.0 7.0 8.0 9.0 Time [y/ṁ] 1 min 10 min 1 h 10 h 1 day 10 day Teff = 3.000× 106 K −10 −8 −6 −4 −2 0 lo g 10 X a'a A ≥60 Figure 4.5: Envelope models at Ṁ∞ = 0.01ṀEdd (left panels) and 0.10ṀEdd (right panels) as a function of ρ. The upper scales indicate the corresponding column depth y and time spent by the accreted matter since it started its journey from the neutron star surface. Panels (a) and (a’): temperature; panels (b) and (b’): mass fraction of selected nuclei; panels (c) and (c’): specific energy generation of dominant processes, as indicated, and with the upper thick line showing the total energy generation. The corresponding values of Qb for 0.01 and 0.10 ṀEdd are 1.09 and 0.15 MeV per baryon, respectively. 120 7.50 7.75 8.00 8.25 8.50 8.75 9.00 lo g 1 0 T [K ] log10y [g cm−2] 5.0 6.0 7.0 8.0 9.0 Time [y/ṁ] 1 min 10 min 1 h 10 h 1 day Teff = 6.500× 106 K −10 −8 −6 −4 −2 0 lo g 10 X 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] 8 9 10 11 12 13 14 15 16 17 lo g 1 0 ε̇ n u c [e rg s− 1 g− 1 ] Hot CNO (α, p) (α, γ) ε̇3α A ≥ 20 A ≥ 40 A ≥ 60 A ≥ 20 A ≥ 40 A ≥ 60 (p, γ) (e+νe)} } 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] 8 9 10 11 12 13 14 15 16 17 lo g 1 0 ε̇ n u c [e rg s− 1 g− 1 ] Hot CNO (α, p) (α, γ) ε̇3α A ≥ 20 A ≥ 40 A ≥ 60 A ≥ 20 A ≥ 40 A ≥ 60 (p, γ) (e+νe)} } b c b' c' 0.25 · MEdd 0.35 · MEdd a' CNO Tz=-1 H 4He A=20-39 A = 40 -5 9 H 4He CNO A=20-39 T z= -1 A = 40 -5 9 A ≥60 A ≥60 a 7.50 7.75 8.00 8.25 8.50 8.75 9.00 lo g 10 T [K ] log10y [g cm−2] 5.0 6.0 7.0 8.0 9.0 Time [y/ṁ] 1 min 10 min 1 h 10 h 1 day Teff = 7.050× 106 K −10 −8 −6 −4 −2 0 lo g 10 X Figure 4.6: Same as Fig. 4.5, for Ṁ∞ = 0.25ṀEdd (left panels) and 0.35ṀEdd (right panels). The corresponding values of Qb for 0.25 and 0.35 ṀEdd are 0.204 and 0.172 MeV per baryon, respectively. 121 8.25 8.50 8.75 9.00 lo g 10 T [K ] log10y [g cm−2] 6.0 7.0 8.0 9.0 Time [y/ṁ] 1 s 10 s 1 min 10 min 1 h 10 h Teff = 9.100× 106 K −4 −2 0 lo g 10 X 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 −10 −5 0 lo g 10 X 8.25 8.50 8.75 9.00 lo g 1 0 T [K ] Teff = 1.3628× 107 K log10y [g cm−2] 6.0 7.0 8.0 9.0 Time [y/ṁ] 1 s 10 s 1 min 10 min 1 h −6 −4 −2 0 lo g 1 0 X 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 −10 −5 0 lo g 10 X 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] 8 9 10 11 12 13 14 15 16 17 lo g 1 0 ε̇ n u c [e rg s− 1 g− 1 ] Hot CNO (α, p) (α, γ) ε̇3α A ≥ 20 A ≥ 40 A ≥ 60 A ≥ 20 A ≥ 40 A ≥ 60 (p, γ) (e+νe){ { 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] 8 9 10 11 12 13 14 15 16 17 lo g 1 0 ε̇ n u c [e rg s− 1 g− 1 ] Hot CNO (α, p) (α, γ) ε̇3α A ≥ 20 A ≥ 40 A ≥ 60 A ≥ 20 A ≥ 40 A ≥ 60 (p, γ) (e+νe){ { d b c d' b' c' 1 · MEdd 5 · MEdd a'a α β γ δ ε ζ α CNO Tz=-1 H 4He A=20-39 A ≥60 A=40-59 H 4He CNO A=20-39 A=40-69 A ≥70 Tz =-1 β δγ ε ζ 12C 16O 15O17F 12C 16O 15O17F 40Ca 40Ca 72Kr 64Ge 43Ti 43Ti 56Ni 56Ni Figure 4.7: Envelope models at Ṁ∞ = ṀEdd (left panels) and 5ṀEdd (right panels) as a function of ρ. The upper scales indicate the corresponding column depth y and time spent by the accreted matter since it started its journey from the neutron star surface. Panels (a) and (a’): temperature; panels (b) and (b’): mass fractions of selected light nuclei; panels (c) and (c’): mass fractions of selected heavy nuclei; panels (d) and (d’): specific energy generation of dominant processes, as indicated, and with the upper thick line showing the total energy generation. Greek labels at the top of panels (d) and (d’) indicate specific positions of events discussed in the text. Qb at the base are 32.59 and 25.54 keV per baryon for ṀEdd and 5ṀEdd, respectively. 122 0 1 2 5 21 25 35 39 45 6 22 26 36 40 46 7 23 27 37 41 47 8 24 28 38 42 48 3 9 11 13 15 17 19 29 31 33 43 49 4 10 12 14 16 18 20 30 32 34 44 50 H(1) F(9) Cl(17) He(2) Ne(10) Ar(18) Li(3) C(6) Na(11) Si(14) Be(4) N(7) Mg(12) P(15) B(5) O(8) Al(13) S(16) K(19) Ca(20) Sc(21) Ti(22) V(23) Cr(24) Mn(25) Fe(26) Co(27) Ni(28) Cu(29) Zn(30) Ga(31) Ge(32) As(33) Se(34) Br(35) Kr(36) Rb(37) Sr(38) Y(39) Zr(40) Nb(41) Mo(42) Tc(43) Ru(44) Rh(45) Pd(46) Ag(47) Cd(48) In(49) Sn(50) Sb(51) Te(52) 1ṀEdd 0 1 2 5 21 25 35 39 45 6 22 26 36 40 46 7 23 27 37 41 47 8 24 28 38 42 48 3 9 11 13 15 17 19 29 31 33 43 49 4 10 12 14 16 18 20 30 32 34 44 50 H(1) F(9) Cl(17) He(2) Ne(10) Ar(18) Li(3) C(6) Na(11) Si(14) Be(4) N(7) Mg(12) P(15) B(5) O(8) Al(13) S(16) K(19) Ca(20) Sc(21) Ti(22) V(23) Cr(24) Mn(25) Fe(26) Co(27) Ni(28) Cu(29) Zn(30) Ga(31) Ge(32) As(33) Se(34) Br(35) Kr(36) Rb(37) Sr(38) Y(39) Zr(40) Nb(41) Mo(42) Tc(43) Ru(44) Rh(45) Pd(46) Ag(47) Cd(48) In(49) Sn(50) Sb(51) Te(52) 5ṀEdd Neutrons P ro to n s Figure 4.8: Space-integrated reaction flows for Ṁ∞ = ṀEdd and ṀEdd, in units of the corresponding value of the 3α reaction. Solid lines: ≥ 0.3; dashed lines: between 0.3 and 0.01; dotted lines correspond to values between 0.01 and 0.005. Nuclides colored in lightblue illustrate the “sawtooth” path, while those in green correspond to the β − 3p − β path (see main text for further details). Some Tz = −1 nuclides are colored in blue. 123 0 2 4 6 6 7 8 9 lo g 1 0 T [K ] 6.00 6.25 6.50 6.75 7.00 log10Ts [K] −4 −2 0 2 4 Q L ,b [M eV b ar yo n − 1 ] Ṁ∞ = 0.03ṀEdd gs,14 = 1.735, Zs = Z⊙ net380 0 2 4 6 log10ρ [g cm−3] 8 10 12 14 16 lo g 1 0 ˙̆ ε n u c [e rg g − 1 s− 1 ] ˙̆ε grav ˙̆εν 6 7 8 9 log10T [K] −0.04 −0.02 0.00 0.02 0.04 (d T /d P )/ 10 − 1 3 0 2 4 6 6 7 8 9 lo g 10 T [K ] 6.00 6.25 6.50 6.75 7.00 log10Ts [K] −4 −2 0 2 4 Q L ,b [M eV b ar yo n − 1 ] Ṁ∞ = 0.10ṀEdd gs,14 = 1.735, Zs = Z⊙ net380 0 2 4 6 log10ρ [g cm−3] 8 10 12 14 16 lo g 10 ˙̆ ε n u c [e rg g − 1 s− 1 ] ˙̆ε grav ˙̆εν 6 7 8 9 log10T [K] −0.04 −0.02 0.00 0.02 0.04 (d T /d P )/ 10 − 13 Figure 4.9: Four panels, at fixed accretion rate, illustrating the envelope profiles in stationary state for different values of Teff. The upper figure corresponds to Ṁ = 0.03ṀEdd while the lower one to Ṁ = 0.10ṀEdd. Upper left panel: temperature as a function of density. Upper right panel: base or cutoff luminosity as a function of the surface temperature. Lower left panel: specific energy generation rate as a function of density, for four selected curves with the same colors as in the upper left panel. The heating/cooling contributions due to neutrinos and mass accretion are illustrated only for the red-curved envelope. Lower right panel: phase space of the temperature profile. 124 0 2 4 6 6 7 8 9 lo g 1 0 T [K ] 6.00 6.25 6.50 6.75 7.00 log10Ts [K] −4 −2 0 2 4 Q L ,b [M eV b ar yo n − 1 ] Ṁ∞ = 0.30ṀEdd gs,14 = 1.735, Zs = Z⊙ net380 0 2 4 6 log10ρ [g cm−3] 8 10 12 14 16 lo g 10 ˙̆ ε n u c [e rg g − 1 s− 1 ] ˙̆ε grav ˙̆εν 6 7 8 9 log10T [K] −0.04 −0.02 0.00 0.02 0.04 (d T /d P )/ 10 − 1 3 0 2 4 6 6 7 8 9 lo g 10 T [K ] 6.00 6.25 6.50 6.75 7.00 log10Ts [K] −4 −2 0 2 4 Q L ,b [M eV b ar yo n − 1 ] Ṁ∞ = 0.60ṀEdd gs,14 = 1.735, Zs = Z⊙ net380 0 2 4 6 log10ρ [g cm−3] 8 10 12 14 16 lo g 10 ˙̆ ε n u c [e rg g − 1 s− 1 ] ˙̆ε grav ˙̆εν 6 7 8 9 log10T [K] −0.04 −0.02 0.00 0.02 0.04 (d T /d P )/ 10 − 13 Figure 4.10: Same caption as in Fig. 4.9, now for 0.30ṀEdd and 0.60ṀEdd. 125 the model is irrelevant at the T → 0 region since, given L < 0, ∇rad < ∇ad and the system remains stable against convection. While having L < 0 at some subset P− ⊂ Psb in the envelope is a physical scenario, T → 0 should be taken with care. It might well be the case that, at most, the crust and nucleus of the star are cold (∼ 106 K), and these cut solutions correspond to transient states where matter is accreted and deposits energy at P−. Notice, however, the solutions to Eqs. 4.18 - 4.25 do not presuppose or require any information regarding Lb/Lcut or Tb/Tcut since they are obtained by integration (given its IVP status). In addition, Lb < 0 but “complete” envelopes actually exist at the four accretion rates (see the curves below the red profiles at the upper right panel of Figs. 4.9 and those lying between the red and the immediately next black curve of Figs. 4.10). Thus, we cannot discard solutions only by having Lb < 0 or Lcut < 0. Additionally, the envelopes with Teff ∈ T − should not be regarded as completely unphysical: it is important to notice the choice of ρb = 107 g cm−3 as envelope-crust boundary is an artificial construction, where we could also have chosen 105 or 106 and these solutions would have remained Lb < 0 but “complete”, or 108 g cm−3 and more Lb < 0 solutions around the red curves would have been discarded due to their T → 0 behavior. The second characteristic of relevance in the T vs ρ panel is the existence, below 0.30ṀEdd, of two branches of fully-accreted envelopes, one below and the other above the region of the Teff ∈ T − models. All envelopes in the branch with Teff < 106.5 K have QLb < 0, as shown in the upper right panel. In this same figure we observe QLb follows a concave tendency, reaching a global minimum and then moving to positive values at increasing Ts(Teff), regardless of the existence of two branches of fully- accreted envelopes. The location of the minimum in the QLb vs Ts plane is sensitive to Ṁ∞, moving from 106.5 K at 0.03 to 106.75 K at 0.6 ṀEdd. This minimum is closer to the second branch of fully-accreted solutions than to the first, low-T one, and the behavior of QLb in this second, high-T branch is almost exponential. Similarly to the existence of T − ⊂ Teff, it could be argued the emergence of two branches of solutions is an artificial construction, consequence of choosing ρb. For instance, at the very 0.3ṀEdd (Fig. 4.10) we see by choosing ρb = 108 g cm−3 the low branch is absorbed into the Teff ∈ T − subset. In clear contrast, we can suspect the upper branch is not an artifact of choosing ρb since the slope of the envelope is related to the nuclear burning: at high densities and temperatures (e.g. lower left panel of Figs. 4.9 and 4.10), ˙̆εnuc decreases from ∼ 1015 erg g−1 s−1 at its maximum to 1011 erg g−1 s−1 as a consequence of 1H and 4He - the main source of fuel - exhaustion. Here we must also consider that below 108 g cm−3 electroweak reactions do not provide sufficient energy as to significantly alter dL/dP , and higher temperatures (e.g. above ∼ 108.75 K) are required to start the fusion of 12C and 16O. Consequently, models with Lb > 0 should exist regardless of our choice of ρb, a scenario which is in agreement with the non-accreting case. What can the phase space tell us regarding the numerical solutions? From the 126 behavior of the envelopes at the four selected accretion rates, we deduce the system has a separatrix since any temperature profile either exhibits T → 0 (going to the left direction in the phase space) or T ≫ Ts (i.e. moving to the right direction of the phase space). The location of the Ts for this almost-vertical separatrix depends on the Ṁ∞: while at 0.03ṀEdd we have Ts,separatrix ∼ 108.25 K, at 0.6ṀEdd we observe Ts,separatrix ∼ 108.75 K. However, due to the density of curves at each side of the separatrix we can deduce solutions starting with low-Teff will reach a maximum and then T → 0, while those starting at high-Teff will remain with finite temperature throughout the whole profile. Without further information, we cannot regard ei- ther of these sides as stable or unstable although, intuitively, we could expect any solution close to the separatrix might become unstable under the presence of small perturbations. Furthermore, the phase space corroborates our initial assumption of the “lower” branch of solutions as a mere artifact of imposing ρb since all solutions in this branch, regardless of the accretion rate, move towards the left direction. Finally, let us discuss the specific generation rate, e.g. lower left panels of Figs. 4.9 and 4.10. Here, the red curves correspond to the model of the same color in the T vs ρ diagram. Below 103 g cm−3, the main contribution to the luminosity comes from the compression of accreted material, although the rather small mass of these outer layers produces little impact over L. In the specific generation rate due to nuclear reactions we observe a first maximum, corresponding to the pp chain processes, in particular to the 2H(p, γ)3He reaction. Its peak contribution (∼ 1014 g cm−3), however, is small in comparison with the heating from compression. A different situation takes place between 104 and 106 g cm−3, where the CNO cycles (both cold and hot versions) overtake the contribution from matter compression and, in contrast to the outermost layers, now the decreasing in luminosity can be large due to the moderate densities. This is corroborated by the blue, green and purple temperature profiles which, after reaching a global maximum, exhibit the T → 0 behavior discussed above. This feature impacts the specific energy generation rate due to nuclear reactions since, as temperature decreases, thermonuclear reactions produce progressively less energy. In clear contrast, for the red curves we see the hy- drogen burning has been completed near 106 g cm−3 and the remaining contribution to the energy comes from the electroweak sector, which is temperature-independent but is limited in extension by the composition of the ashes. Same accretion rate, different gs While we have observed the low-temperature branch of “fully accreted” envelopes is an artifact of choosing ρb, it is nevertheless important to observe how different are envelopes when we change gs. In Fig. 4.11 we fix the composition at the surface and the accretion rate at ṀEdd and change the values of stellar radius from the 1.4M⊙, APR-core EOS star, as to those corresponding to 1.1M⊙ and 2.0M⊙ for the same core EOS, thus having a smaller and larger surface gravity than the value for the 127 0 2 4 6 6 7 8 9 lo g 1 0 T [K ] 6.00 6.25 6.50 6.75 7.00 log10Ts [K] −4 −2 0 2 4 Q L ,b [M eV b ar yo n − 1 ] Ṁ∞ = 0.30ṀEdd gs,14 = 1.262, Zs = Z⊙ net380 0 2 4 6 log10ρ [g cm−3] 8 10 12 14 16 lo g 10 ˙̆ ε n u c [e rg g − 1 s− 1 ] ˙̆ε grav ˙̆εν 6 7 8 9 log10T [K] −0.04 −0.02 0.00 0.02 0.04 (d T /d P )/ 10 − 1 3 0 2 4 6 6 7 8 9 lo g 10 T [K ] 6.00 6.25 6.50 6.75 7.00 log10Ts [K] −4 −2 0 2 4 Q L ,b [M eV b ar yo n − 1 ] Ṁ∞ = 0.30ṀEdd gs,14 = 3.238, Zs = Z⊙ net380 0 2 4 6 log10ρ [g cm−3] 8 10 12 14 16 lo g 10 ˙̆ ε n u c [e rg g − 1 s− 1 ] ˙̆ε grav ˙̆εν 6 7 8 9 log10T [K] −0.04 −0.02 0.00 0.02 0.04 (d T /d P )/ 10 − 13 Figure 4.11: Same caption as in Fig. 4.9, now for a fixed accretion rate, 0.30ṀEdd, and two different values of surface gravity gs. 128 0 20 40 60 80 100 Mass Number −10 −8 −6 −4 −2 0 lo g 1 0 [A b u n d an ce ] gs,14 = 1.74 ρb,7 = 1 Tb,8 = 5.45 1.0ṀEdd XH = 0.01, 6.35×106 K XH = 0.10, 7.15×106 K XH = 0.20, 7.60×106 K XH = 0.30, 8.00×106 K XH = 0.40, 8.30×106 K XH = 0.50, 8.60×106 K XH = 0.60, 8.87×106 K XH = 0.70, 9.10×106 K 0 20 40 60 80 100 Mass Number −10 −8 −6 −4 −2 0 lo g 1 0 [A b u n d an ce ] gs,14 = 1.74 ρb,7 = 1 Tb,8 = 4.87 0.3ṀEdd XH = 0.01, 5.000×106 K XH = 0.10, 5.390×106 K XH = 0.20, 5.750×106 K XH = 0.30, 6.040×106 K XH = 0.40, 6.285×106 K XH = 0.50, 6.470×106 K XH = 0.60, 6.635×106 K XH = 0.70, 6.800×106 K Figure 4.12: Distribution of abundances for different accreted fractions of hydrogen, denoted as XH, and helium at the surface with XHe = 0.98−XH. Left panel: Ṁ∞ = ṀEdd, right panel: Ṁ∞ = 0.30ṀEdd. Here, ρb,7 = ρb/107 g cm−3 and Tb,8 = Tb/108 K. At ṀEdd, Qb ranges from 0.58 (at XH = 0.01) to 0.03 (at XH = 0.70) MeV per baryon. At 0.3ṀEdd, Qb ranges from 1.0 (at XH = 0.01) to 0.15 (at XH = 0.70) MeV per baryon. 1.4M⊙ star. In the phase-space diagram (lower right panel of Fig. 4.11) we still infer the presence of a separatrix curve, although its location in Ts is directly affected by gs, similarly to the location and the value of the minimum in the QL,b vs Ts diagram. Within the context of stable burning, we could argue a higher internal temperature is required in the presence of a strong gravitational field in order to stabilize the burning. Regarding the nuclear energetic output, we observe high surface gravity environments allow to reach the rp-process even for those incomplete envelopes of T → 0 behavior, while their counterparts at low surface gravity are starting to reach the CNO-breakout point. In the absence of accretion, at fixed chemical composition and at fixed effective temperature (Fig. 4.2), we observed an inverse correlation between Tb and gs. For the Lb > 0 envelopes in Fig. 4.11 we still observe this correlation: take the Ts = 107 K for instance. At gs,14 = 1.262, we see Tb ∼ 109 K, while at gs,14 = 3.238 we have Tb < 109 K, from where we can infer a similar correlation for all “fully accreted” envelopes below the red curves. Does the rp-process depends on the metals at the surface? Short answer: no. Figs. 4.12 illustrate this point very well: here we see the distri- bution of ashes at ρb = 107 g cm−3 for two collections of models, each panel with the indicated mass accretion rate and gs. Common to all curves is the amount and composition of metals at the surface, Z⊙ = 0.02, while specific to each model is the effective temperature and the mass fractions of hydrogen XH and helium, chosen 129 as X4He = XH+4He,⊙ − XH with XH+4He,⊙ = 0.98. In both panels we observe that H-deficient matter is less likely to change the initial distribution of metals. On the other hand, even with as small fractions of XH as 10% we start noticing the change in composition below A = 56. As the hydrogen fraction exceeds 30% the disap- pearance of 40Ca in favor of A ≥ 56 metals becomes progressively significant until reaching the distribution discussed in the first subsubsection. From these curves, we can deduce that as long as hydrogen fraction is ≈ 20% some metals can actually be synthesized, although for having a basic form of rp-process we require between 30% to 40% percent of hydrogen in the accreted matter. First wrap-up: are Tb − Lb − Teff viable? In the absence of time-dependent terms in the differential equations, we were led to Eqs. 4.18 - 4.25. Contrary to the intuitive expectation of having all models in a given Teff set with Lb > 0, we found stationary solutions with Lb < 0, as well as other curves which behaved as T → 0 at some ρcut < ρb, which we could not regard as fully-accreted envelope given their physical behavior, despite its mathematical consistency with the input functionals. In principle, at any Ṁ∞ ≥ 0.3ṀEdd (in order to be consistent with observations), we might construct Tb − Teff − Lb relations. It suffices to consider only the fully- accreted envelopes, regardless of sgnLb, and store in tables (or attempt an analytic fit) their values of Teff and the pair Tb, Lb. However, without further information regarding the time-evolution behavior of an actual envelope, the existence of an artificial category of T → 0 curves as “incomplete” envelopes prevents us from im- plementing such straightforward scheme. In particular, we could simply argue there exists certain values of Teff which the system cannot adopt, although without the temporal terms and within the given physics there does not exist a strong physical reason as to why it should be so. On the other hand, there is a physical reason to question why we cannot rule out these solutions without restoring to time-dependent simulations: the existence of the “secondary branch”, despite being artificial, points out that the equations allow envelopes of low effective temperature to inject energy to the interior (i.e. Lb < 0), which in time-dependent scenarios should lead to the heating of the core. This point is important given that, a priori, we do not know the temperature of the core and, if it starts - before accretion - well below 107 K, implementing a Tb −Teff −Lb relation only with fully-accreted envelopes lead to the existence of a gap in accessible temperatures Tb. Furthermore, directly ruling out the “secondary” branch suggests all cores should start with at least Tb ∼ 107 K and with a high effective temperature, close to 107 K as well. A secondary characteristic which requires clarification is the steep slope of Lb(Teff). Leaving aside the potential numerical difficulties of having a large difference in Lb within a small range of Teff, the behavior in the phase-space of the stationary solu- tions poses a question: are the Lb ≈ 0 curves approaching an unstable equilibrium 130 point? In that case, what is the role of the “discarded” solutions with T → 0? What is the response of the rest of the start to such unstable equilibrium? As a first patch, we could restore to instability arguments in order to impose a different cutoff to the T → 0 models. Indeed: by comparison with existent diagrams on the subject, we see these curves correspond to the “envelope just before the explosion” models described elsewhere. In these works, the main idea was to find out the location at which the classical instability criterion, e.g. ∂ϵheat/dT > ∂ϵcool/dT , predicts the stationary system is unstable against small perturbations in temperature. Coincidentally, we see the location of such instability point agrees well with the location of the maximum in temperature. However, this criterion is not exact, and the uncertainty in the location of the explosion density suggests we have to either cut at L > 0 or L < 0. Given ρcut < ρb, for the formal Tb − Teff − Lb scheme we require to join this piece of solution to another portion, not necessarily the numerical core. In the first case, Lcut > 0, joining the cut accreted envelope with a “compressed”-like solution induces a higher Tb - as in those non-accreted models of the past section. On the other hand, having Lcut < 0 is likely to lead towards the same T → 0 behavior given the absence of heating or cooling sources at ρ ∈ [ρcut, ρb]. Given the behavior of T (P ) and L(P ), without time-dependence one could also suspect the stationary solutions represent the different states the envelope takes during constant accretion, i.e. starting at low-Teff and then moving to high-Teff while injecting energy at ∼ 106 g cm−3. However, such picture would indicate the crust and core are colder than the surface at all times, and only latter they become hotter. There is, however, not sufficient evidence as to claim this is a universal behavior, as it could be the case to start with a hot core but cold exterior (as suggested by the non-accreting envelopes in the past section). In summary: while the stationary states allowed us to have a deeper grasp on the underlying physics and the overall behavior of the envelope, we still require to restore to the fully time-dependent simulations in order to understand how we can - if possible - implement adequate Tb − Teff − Lb relations since we require further knowledge on the nature of those curves of effective temperature in T −. 4.3.2 Intermezzo: Linear perturbations to stationary solu- tions and the origin of bursts Before embarking in a fully time-dependent trip, it is important to understand the underlying behavior of these simulations. In order to do so, let us now discuss per- turbations to the stationary solutions. From the general properties of the envelope (Subsection 1), the discussion of time-dependent perturbations can be carried em- ploying Eqs. 4.8 and 4.9 only - we see structure plays a minor role in contrast to temperature and the evolution of chemical composition. Given the instability analysis to the heat diffusion equation of Chapter 1, it is 131 2 3 4 5 6 log10ρ [g cm−3] −10.0 −7.5 −5.0 −2.5 0.0 2.5 5.0 7.5 10.0 lo g 10 τ [s ] Ṁ∞ = 0.30ṀEdd gs,14 = 1.735, Zs = Z⊙ net380 τdiff τacc τnuc Figure 4.13: Different timescales for the same models in Fig. 4.10. useful to classify the sources at the right hand side of Eq. 4.8 as heating or cooling mechanisms, depending on their overall sign. From the stationary solutions, we know ˙̆εnuc − ˙̆εν remains positive in practically all scenarios of interest, hence it is a heating mechanism. For simplicity, let us enclose them into a single term, ˙̆ε := ρB( ˙̆εnuc − ˙̆εν). Since ∇ad − ∇T > 0 (c.f. Chapter 2), the source term related to accretion in Eq. 4.8 acts as a heating mechanism as well. Finally, we have the flux gradient term. While its tendency in stationary state is that of a cooling mechanism, whenever dF/dP > 0 this term must be taken as a heating mechanism instead. Now, let us look at the timescales. Given heat diffusion is the only cooling mechanism, a separate comparison between τdiff with τnuc and of τacc with τdiff allows to qualitatively estimate the interplay between heating and cooling mechanisms, while comparing τnuc with τacc allows to determine which heating mechanism is more important than the other at different conditions throughout the envelope. Let us first compare τdiff and τacc, Eqs. 4.10 and 4.11. At ρb = 108 g cm−3 and ṀEdd, we have τacc ∼ 12 days, while the local diffusion timescale is approximately a 12th part of this value. Consequently, if mass accretion heats the envelope, radiation and conduction have ample time to transport the energy towards the coldest region of the star, either the surface or the deep layers. Notice in this assertion we have not made further assumptions regarding the temperature of the core just before the start of accretion: while the majority of simulations presuppose Tb ∼ 108 K, it might be the case that Tb ∼ 106 K. Under the latter scenario, some fraction of nuclear heating (generated as a consequence of mass accretion) might leak towards the interior of the star rather than to the exterior. However, the propagation of heat is still controlled by τdiff, which goes from days to weeks as we move from 107 to 132 1010 g cm−3, while at the surface this value is of ∼ hr to seconds (see also Fig. 4.3). A different scenario takes place between τnuc and τacc, Eqs. 4.12 and 4.11. At the outermost layers of the envelope (≤ 103 g cm−3), nuclear energy is predominantly generated by pp-chains, thus ˙̆εnuc ≈ 1011 g cm−3. Since the EOS is almost of an ideal gas, i.e. cP ≈ 5ρkB/2⟨m⟩imu, by definition of the nuclear timescale at T ∼ 108 K we have τnuc ≈ 105 s. On the other hand, near the conditions for 1H exhaustion (via CNO or rp-process) we know τnuc ≈ 103 s, thus nuclear timescale becomes short at increasing depths. However, τacc exhibits the opposite behavior, going from small values at the surface to high values (e.g. ∼ hrs to days) at ρB ≥ 106 g cm−3. Therefore, at the outermost layers of the envelope it takes more time to burn than to accrete fuel, while at deep layers the nuclear burning of matter is faster than the accretion. Finally, we have the competition between nuclear and diffusion timescales. From the above discussion, as well as the considerations on stability of Chapter 1, as long as τdiff < τnuc the stability against thermonuclear perturbations is guaranteed in the envelope. In the opposite limit, such thermonuclear perturbations lead to a drastic increase in the local temperature, which propagates towards the surface via radiation or convection, depending on the sign of ∇T − ∇ad. Such increase is short-lived with respect to the duration of accretion episodes, and as such is the origin of the observed X-ray bursts. The ashes of these bursts accumulate at the envelope-crust boundary, compress the pristine outer crust and eventually replace it. We can also take a look at these timescales from the actual output of our time- independent simulations. In Fig. 4.13, we observe the three timescales for the four selected envelope models from Fig. 4.10. In first place, we observe little discrepancy among the four displayed τacc, as expected since this term is practically controlled by the accretion rate and the surface gravity - equal in the four scenarios - with small influence of pressure only, as evidenced by the cut-off in the purple curve. A similar pattern is followed by τdiff, timescale which is primarily controlled by the thickness of the envelope d - which is rather small for the selected models - and secondary by the chemical composition throughout the specific heat and thermal conductivity. Given the actual temperature is higher for the red curve than for the purple one, we expect the observed inverse correlation, e.g. the timescale is larger for the coldest model, the purple one. Notice τdiff < τacc well below 105 g cm−3, indicating the envelope has ample time to relax into equilibrium. Furthermore, in this density region we observe diffusion timescale is actually the shortest among the three, allowing us to conclude the envelope there has ample time to relax into an equilibrium state. An interesting scenario takes place for the green and red curves, and consequently for all intermediate envelope models between them. As a consequence of the nuclear burning, we see an inversion of the relation between τacc and τnuc near 105 g cm−3. Having a short nuclear timescale implies more time is required into accreting than into burning it. Following this inversion trend, between 105 and 106 g cm−3 we 133 observe nuclear and diffusion timescales are equal. In this region an instability is likely to take place: if matter requires more time to diffuse than to produce heating, a temperature build up might follows and, consequently, a thermonuclear explosion. The former qualitative discussion can be put into formal terms employing the one zone model, which focuses on analyzing the stability of temperature and com- position under small perturbations. In this approximation, spatial derivatives are approximated by local values (i.e. ∂PA ≈ A/P ), and taking the flux contribution to Eq. 4.8 as a local term, i.e. ε̇cool := −ρBgs ∂F ∂P ≈ 16σSBg 2 sρBT 4 3P 2κ . (4.26) Given Nspecies ≫ 1, at the perturbative level it suffices to consider only one species as responsible for the perturbations to the system. Its mass fraction will be denoted as Xc. Under these considerations, in the one zone model approximation Eqs. 4.8 and 4.9 reduce to cP ∂T ∂τ = ε̇c − ε̇cool (4.27) ∂Xc ∂τ = −ṁgs Xc P + AcRc. (4.28) In stationary state, e.g. ∂τT = 0 and ∂τXc = 0, we obtain two expressions indicat- ing the equilibrium between heating and cooling mechanisms for the temperature equation, i.e. ε̇c = ε̇cool, and the balance between mass accretion and the rate of burning, ṁgsXc/P = AcRc. For the linear perturbation analysis, let the local rate, energy generation rate and opacity be approximated as2 Rc = R0X α1 c Tα2 (4.29) ε̇c = R0X α1 c Tα2 (4.30) κ = κ0X β1 c T β2 (4.31) By introducing T (t, P ) = T0(P ) + δT (t, P ) and Xc(t, P ) = X0(P ) + δXc(t, P ) into Eqs. 4.27 and 4.28, keeping the linear terms, employing the stationary relations among mechanisms and defining θ := δT/T and χc := δXc/Xc, we arrive to ∂θ ∂τ = 1 τnuc ¶[α2 + β2 − 4] θ + (α1 + β1)χc♢ (4.32) ∂χc ∂τ = 1 τacc ¶α2θ + (α1 − 1)χc♢ , (4.33) i.e. a coupled system of linear equations. While these can be combined into a single ODE for either θ or χc (and thus admitting solutions of the form ∝ exp(λ′τ)), 2Notice perturbations in density can easily be absorbed as perturbations in κ since, at constant pressure, ρB(T, Xc). 134 its linearity make suitable to analyze the behavior of the system by looking at its eigenvalues. By explicit calculations, it can be shown these are3 λ± = 1 2 [ α2 + β2 − 4 τnuc + α1 − 1 τacc ] ±    1 4 [ α2 + β2 − 4 τnuc − α1 − 1 τacc ]2 + α2(α1 + β1) τnucτacc    1/2 . (4.34) We thus see the system admits two kinds of solutions, oscillatory and exponentials, depending on the sign of the term inside the square root. The first case corresponds to the observed thermonuclear explosions due to the recurrence time. The second scenario corresponds to systems where nuclear burning does not induce explosions, i.e. as in sources at Ṁ ≥ 0.3ṀEdd. 4.3.3 Time-dependent envelopes In order to model the time-dependent evolution of an envelope, we employed the publicly available code MESA. Given our interest in varying parameters such as the base luminosity Lb and the boundary density ρb, we provided as initial conditions to this program the output from our time-independent envelope code in the absence or presence of accretion. In the former case, an artificially layered composition was set, prioritizing the presence of hydrogen at ρ ≤ 104 g cm−3 and of metals at higher densities. Considering the surface boundary condition is written in terms of Teff, and for non-accreting envelopes Ls = Lb, in addition to the MeV per baryon unit for the luminosity we employ as well the effective temperature. For instance, as a “cold” model we refer to the initial envelope from our stationary code having Teff = 105 K, i.e. ≈ 2.59 × 10−5L⊙. To be self-consistent with the stationary codes we employed in the past sections, we incorporated into MESA additional subroutines for the radiative opacity and called them during the execution of the simulations instead of the tabulated opacities which come by default. These fits correspond to the updated version of the electron scattering opacity from [Pou17], and the electron- ion interaction from [SBCW99]. Regarding the conductive opacity, we employed the tables provided by MESA considering they share the same formalism as that we employ in our stationary code. Simple experiments As a first step towards the understanding of the different branches and curves we found in the stationary approximation, we implemented accreted envelopes with 3Our expression for λ± differs from Heger 2007 for two reasons: (i) we consider the evolution of the mass fraction instead of the column density y = P/gs and (ii) dependency of κ is explicitly addressed. In order to recover their expressions, it suffices to consider α1 = β2 = 0 and β1 = −2. This apparent dependency of κ with y is only a consequence of choosing y instead of X at ε̇cool (e.g. Eq. 4.26): notice ∂y ε̇cool ∝ −2κ−1y−3 = β1κ−1y−3 despite κ ≡ constant, as done by Heger 2007. 135 0 2 4 lo g 1 0 L /L ⊙ 0 1 2 3 4 5 6 t [hr] 7 8 lo g 1 0 T b [K ] 〈gs,14〉 = 2.167, net381 0.30ṀEdd, Zs = Z⊙ Lb =-1.926 L⊙, ρ6.99 Lb =-56.25 L⊙, ρ5.93 Lb =-13.07 L⊙, ρ6.93 2 4 lo g 1 0 L /L ⊙ 0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00 t [hr] 8.0 8.5 lo g 1 0 T b [K ] 〈gs,14〉 = 2.167, Approx140 ṀEdd, Zs = Z⊙ Lb =-11.17 L⊙, ρ7.00 Lb =-2.413 L⊙, ρ8.00 Lb =-11.17 L⊙, ρ8.00 Lb =-48.93 L⊙, ρ7.00 Figure 4.14: Upper panels: Surface luminosity and bottom temperature for 0.3ṀEdd models with different Lb. Their respective ρb are indicated in the labels. Lower panels: same as upper ones, but for ṀEdd. different Teff as initial conditions for MESA. Considering the observed two-branched behavior of the stationary envelopes at 0.3ṀEdd, we selected three envelopes from this family of solutions: one from the “artificial”, low-Teff branch, the second from the cut-off models with T → 0 at ρcut < 107 g cm−3, and the third one from those fully-accreted envelopes (i.e. high-Teff branch) with Lb < 0. The temperatures at the bottom of the envelope as functions of time, as well as the surface luminosity, are displayed in Fig. 4.14. The models were labeled according to their base luminosity, expressed in units of L⊙, and are ordered from bottom to top according to their Teff at t = 0 s. The most remarkable feature of this plot is the rapid decrease in temperature from the T → 0 model - purple curves. Once reached Tb ∼ 103 K, MESA did not proceed with the integration due to the lack of convergence towards a satisfactory model. Although its base density is < 107 g cm−3, and we could actually expect this behavior since we are providing a negative value at the base - fixed by MESA construction - its lack of further advance in contrast to the other models (with negative luminosity as well) suggests the existence of a minimum luminosity we can inject at a certain ρ. 136 In contradistinction with the T → 0 model, we observe the remaining two proceed normally for more than one hour. In the case of the one belonging to the “low temperature branch”, we see a slow build-up of temperature, both at surface and at the bottom, until eventually a sudden increase takes place after ∼ 5 hours, i.e. a thermonuclear explosion as a consequence of having hydrogen up to 107 g cm−3. Finally, we observe the model from the “high branch” of solutions cools down for one hour before starting a thermonuclear instability. Given ρb = 107 g cm−3 for this latter model, we deduce models with Lb < 0 are indeed viable scenarios for representing the envelope of a neutron star. However, one could pose into question the viability of having a fixed (e.g. time-independent) amount of energy flowing from the surface to the interior since this scenario does not take into consideration the reaction from the numerical core, which must be heated as a consequence. Another characteristic to take into consideration is that injecting energy from the envelope to the core does not lead to a bursting suppression behavior - bursts still occur even with Lb < 0. And then another question emerges: how much energy do bursts actually deposit towards the interior of the star? How is this compatible with the classic picture of Lb > 0? In the lower panels of Fig. 4.14, we show the results of performing a similar experiment, now with stationary models corresponding to the ṀEdd mass accretion rate. Here we also allowed for different location of the negative luminosity, either ρ = 107 g cm−3 or 108 g cm−3. The cost to pay for increasing the boundary density is to have a slow numerical simulation in proportion to the amount of species in the network. Consequently, we opted for changing the net381 network by Approx140. Once again, the most notorious of the models is that where internal temperature drastically converges to zero in less than one hour, a consequence of keeping a negative luminosity at all times and, similarly as in the 0.3ṀEdd case, we can infer the existence of a minimum amount of heat that can be injected at certain densities. By keeping the same luminosity at different ρb, we observe a displacement in the rising phase of the first burst of the simulation, starting first for the ρb = 108 g cm−3 model. The variations in the recurrence time, however, are almost negligible. The second characteristic of importance is the small variation in Tb due to our selection of ρb. Having a short boundary density but keeping the luminosity as constant implies the inner temperature at such ρb varies drastically, as shown in the lowest panel of Fig. 4.14. On the other hand, the variations in Tb at the ρb = 108 g cm−3 are practically non-existing in the selected interval of time, suggesting the injection of energy from the envelope to the outer crust cannot proceed between 107 and 108 g cm−3. However, this is also a consequence of the different timescales (see Fig. 4.10), where the dissipation of a perturbation to the temperature profile requires scales in the order of days to weeks, hence the possibility of keeping Tb as practically constant for the ρb = 108 g cm−3 model. 137 Time-independent against time-dependent envelopes 8 9 lo g 10 T [K ] 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] −200 0 200 400 600 Q L [M eV b ar yo n − 1 ] Approx140, log10 ρb = 7.11 0.3ṀEdd, gs,14 = 1.73 70% 1H, 28% 4He, 2% 12C 8.0 8.5 lo g 10 T [K ] Approx140, log10 ρb = 7.11 0.3ṀEdd, gs,14 = 1.73 70% 1H, 28% 4He, 2% 12C 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] −5 0 5 Q L [M eV b ar yo n − 1 ] Figure 4.15: Temperature and luminosity (in MeV per baryon) against density for several snapshots of a time-dependent simulation with MESA (colored curves), con- trasted against the output from our time-independent envelope code (red, dashed curves). Upper panel: comparison during the rising phase of an arbitrary burst. Lower panel: comparison during the inter-bursting phase. The next logical step, given the time evolution of these “fully accreted envelopes” with negative base luminosity, is to ponder whether such states can actually be reached if we start accreting material over a pristine envelope. A related question, given the unstable nature of the mass accretion at 0.3ṀEdd, the origin of the explo- sion at around 106 g cm−3 and the similarity of the timescales - nuclear, accretion and heat diffusion - at said density, is: to what extent our accreting envelopes match with the time-evolution output from MESA? In order to answer these questions, we ran a simulation, employing the Approx140 network, of constant mass accretion at 0.3ṀEdd of a simple mixture of 70% 1H, 28% 4He and 2% 12C. The composition of the pristine envelope was chosen as pure-hydrogen below 104 g cm−3 and of pure 80Kr above this density. For simplicity, we chose ρb = 107 g cm−3 and Lb correspond- ing to an effective temperature of 105 K for a 11.57 km neutron star. As is usual, we regarded the output from the first burst as unphysical and focused on the following ones. To compare the MESA models with our own, we took their corresponding Teff and performed the usual integration until either ρ = ρb or ρ = ρcutoff, depending on 138 whether the curves exhibited a T → 0 behavior or not. In Fig. 4.15, we can see a comparison between the time-dependent and indepen- dent temperature and luminosity profiles at two different stages during this accretion episode. In the upper panels, we contrast the profiles during the rising phase of a thermonuclear instability, while in the lower panels the envelopes are contrasted dur- ing the stable mass accretion in-between bursts. The most prominent characteristic, common to both episodes, is the similarity in the temperature profile at ρ ≤ 105 g cm−3. At large densities, however, the discrepancies are notorious in both cases as well: at the explosion episode, the stationary temperature profiles with high Ts have larger inner temperature than their time-dependent counterparts, while those ex- hibiting T → 0 behavior underestimate the temperature - when contrasted against the time-dependent models - near 106 g cm−3. On the other hand, all stationary temperature profiles in-between bursts have the T → 0 behavior at 106 g cm−3 while their time-dependent counterparts follow the same trend dictated by the Lb > 0. Nevertheless, in principle we could argue in favor of setting ρb = 105 g cm−3 in order to employ these time-independent temperature profiles to construct a Tb(Teff) relation. However, not all stationary luminosity profiles match their time-dependent counterparts: for instance, during the explosion both profiles diverge from one an- other at ρ ≥ 104 g cm−3, while during the in-between bursting period the highest Ts stationary profile predicts a larger injection of energy near 105.5 g cm−3 than what is actually obtained by the time-dependent simulation. An important result to be noted from Fig. 4.15 is the amount of energy injected towards the interior during the burst episode, ≈ 104L⊙. Since MESA did not find further issues on running and completing this simulation, we might argue such results are self-consistent and do not fail to converge due to the presence of the Lb > 0 source at ρb, which prevents the temperature profile to drastically drop towards zero. However, we can also raise the question: does this amount of heat depends on our choice of ρb? Energy flowing to the interior In the previous subsection we have seen energy can flow from the envelope towards the numerical core of the star. Is there a maximum density such injection of energy can reach? Is the size of the network important for the correct account of the energy? In Figs. 4.16 4.17 we see the output of four simulations which have a common value for ρb, 109.58 g cm−34. Such value of boundary density is suitable for both observing the distribution of heat from the bursts and to relax the condition on the amount of heating from the base, as we have set it to 2.59 × 10−5L⊙ in the four models. The accretion rates we have selected are 0.3 and 1 times ṀEdd, and employed 4The original envelope, generated with our stationary code, had a boundary of 3 × 109 g cm−3. However, during the adjustment of density and mass performed by MESA this boundary density was modified. 139 two different reaction networks, Approx21 with a base of 80Kr and Approx140. Solar composition was accreted at the surface, and in the four models we ran the simulation for 8 hours (MESA time). In the four envelopes, we see the bursts deposit a considerable amount of energy at 107 g cm−3. In the case of recurring explosions, after the first one we observe an almost constant amount: between 1 and 4 MeV per baryon in the Approx140 network and between 0.2 and 0.4 in the Approx21 case. By comparing the evolution of the models from each network at same accretion rate, we appreciate the importance of having a sufficient amount of nuclides as to not underestimate the amount of heating flowing towards the interior. In addition to such punctual deposition of energy at 107 g cm−3, we observe that as t ≫ 0 energy progressively flows towards the interior, as evidenced by the QL curves (lower panels of each pair of figures) at 108 and 109 g cm−3, and the increasing in temperature (upper panels of each pair of figures). As a consequence of having more explosion episodes - and thus more heating injection - we observe that regardless of the network choice the temperature at 108 g cm−3 increased at least one order of magnitude, while temperature at 109 g cm−3 remained almost constant. While this is partially a consequence of choosing a same base - composed of 80Kr - the network also plays a role since the exact values of this temperature gradient are clearly network dependent. An additional network- dependent effect that we observe is the bursting suppression for the Approx21 net at ṀEdd, while for the Approx140 net the bursts continue to occur at a regular rate. Despite such suppression, the luminosity remains negative between 107 and 109 g cm−3: eventually, such heating will flow towards the interior and will increase the temperature. Given the Lb > 0 condition cannot be relaxed in MESA, and the evolution of the envelope with such restriction becomes less valid as the time of the simulation approaches and exceeds ∼ 1 week (typical heat diffusion scale for the outer crust, as inferred from Fig. 4.13), we cannot rule out the possibility that this stored heat at ≈ 108.5 g cm−3 will eventually flow until ρb in a simulation of the full neutron star. 5 6 7 8 lo g 10 T [K ] log10Teff 0.3ṀEdd, 〈gs,14〉 = 2.176 80Kr base, log10ρb =9.582 log10ρ = 7.0 log10ρ = 7.5 log10ρ = 8.0 log10ρ = 9.0 0 1 2 3 4 5 6 7 8 Time [hr] −0.4 −0.2 0.0 Q L [M eV b ar yo n − 1 ] Zs = Z⊙ Approx21 + 80Kr 5 6 7 8 lo g 10 T [K ] log10Teff log10ρ = 7.0 log10ρ = 7.5 log10ρ = 8.0 log10ρ = 9.0 0 1 2 3 4 5 6 7 8 Time [hr] −0.4 −0.2 0.0 Q L [M eV b ar yo n − 1 ] ṀEdd, 〈gs,14〉 = 2.176 Zs = Z⊙ 80Kr base, log10ρb =9.583 Approx21 + 80Kr Figure 4.16: Temperature and luminosity, at specific values of log10ρ, as a function of time for accretion over a pure 80Kr envelope employing the Approx21 network, for two selected accretion rates, 0.3 and 1 times ṀEdd. 140 5 6 7 8 9 lo g 10 T [K ] log10Teff log10ρ = 7.0 log10ρ = 7.5 log10ρ = 8.0 log10ρ = 9.0 0 1 2 3 4 5 6 7 8 Time [hr] −4 −2 0 Q L [M eV b ar yo n − 1 ] 0.3ṀEdd, 〈gs,14〉 = 2.176 Zs = Z⊙ 80Kr base, log10ρb =9.582 Approx140 5 6 7 8 9 lo g 10 T [K ] log10Teff log10ρ = 7.0 log10ρ = 7.5 log10ρ = 8.0 log10ρ = 9.0 0 1 2 3 4 5 6 7 8 Time [hr] −2 −1 0 Q L [M eV b ar yo n − 1 ] ṀEdd, 〈gs,14〉 = 2.176 Zs = Z⊙80Kr base, log10ρb =9.583 Approx140 Figure 4.17: Same caption as in Fig. 4.16, now for the Approx140 network. Increased opacity: a mechanism to stop bursts? Observationally, no bursts have been registered at Ṁ ≥ 0.3ṀEdd [BHM+07]. Al- though the analysis of the previous sections suggest this is feasible since temperature obeys a diffusion equation, fully time-dependent simulations have shown bursting suppression occur, theoretically, at Ṁ ≥ ṀEdd [HCW07, FTG+07]. In addition, the envelope requires very high values for Lb to suppress the oscillations. Several mechanisms have been invoked to explain either this large value of Lb or potential mechanisms for stabilization of bursts at exactly 0.3ṀEdd, such as • 2D/3D effects. So far, the majority of bursts simulation have considered a single spatial direction. There are hints that spreading of material might have an impact for their stabilization, taking into consideration gradients in the surface gravity [CWG17]. Another mechanisms might be chemical diffusion as a consequence of rotation: models in this direction have reported a reduction in the required value for Lb [KLi09]. • Different reaction rates. The 15O(α, γ)19Ne reaction has received special attention since it was found to be one, if not the most important, of the CNO- breaker reactions [FGWD06a, TFG+07]. For this reaction to take place, tem- peratures between ∼ 108.4 K (around 107 g cm−3) and ∼ 108.6 K (for 104 g cm−3) are required, becoming competitive with respect to the density and temperature independent 15O(β+νe) 15N. In the left panel of Fig. 4.18 we illus- trate several versions of this thermonuclear rate as a function of temperature, at two typical densities at the surface of neutron stars, and we can see their rather good agreement below ∼ 109.25 K. Regardless of the importance of this rate, however, the CNO-breaking must be complemented with the subsequent proton capture over 19Ne in order to prevent the return to the CNO cycle via 19Ne(β+νe) 19F(p, α)16O. As illustrated in the right panel of Fig. 4.18, in con- trast to the α-capture, this p-capture breaking path is favorable even below 108.25 K, enhanced by density around 107 g cm−3, which favors the subse- quent proton captures over progressively heavier nuclides. This in principle 141 explains why, even if CNO cycles are not broken, we can still synthesize mod- erate amounts of 40Ca at low accretion rates (e.g. ∼ 0.1ṀEdd). Going back to our main reaction of interest, if the CNO breaking occurs, we can go beyond 40Ca and 56Fe synthesis, reaching until ∼ 64Zn in the rp-process. Current experimental knowledge of this rate is limited due to the absence of strong 15O beams permitting a direct measurement. The explanation of this is due to t1/2 =122.22 s, which makes “impossible” to sustain these beams long enough as to allow significant α-captures to be detected [HdS21], although some evi- dence also exists in favor of a direct measurement [TWK+17]. Instead, some ongoing and recent collaborations have focused on 15O(7Li,t)19Ne reaction, mimicking an α-capture, from where it should be possible to pose further con- straints on the α-capture reaction [ACL+21]5. Nevertheless, available rates are usually constrained from indirect mechanisms such as the 19Ne energy lev- els [HBB+20], 19Ne(γ, α) decay [TGB+09] or the outcome from 20Mg(βp)19Ne path [GPLW+19] provide information to construct theoretical version of this rate, such as those in Fig. 4.18. Over the years, this absence of direct data has opened the possibility for theoretical speculation, probably the most no- torious in the astrophysical context is that the actual rate is few times smaller than estimated, as this would allow to reconcilie the absence of X-ray bursts above 0.3ṀEdd with theoretical predictions, pointing that bursting behavior cease at ∼ 3ṀEdd. This argument was primarily pointed out by Fisker et al [FGWD06a, FTG+07], whose PhD dissertation focused on the reaction flow during an X-ray bursts and represents a modern understanding on the rp process. Later theoretical models [CN06], recovering linear stability analysis [CN05] (which in their own words could be of “obscure” meaning), showed this argument could be used to construct an approximate network of three-to- four nuclides for simulating the rp process, and effectively found that bursting behavior should stop at ∼ 0.3ṀEdd. We must note, however, this approach overlooks the importance of the 18Ne (α, p) 21Na rate which is important for X-ray bursts as well. From stationary simulations we see this rate replaces the 15O (α, γ) as the leading CNO-breakout for accretion rates comparable with or greater than 5ṀEdd, a behavior resembling the evolution during a burst (see [SAB+01]), thus a simplification requires the presence of the 18Ne(α, p) rate. In spite of how good such argument could be, particularly in view of the observational evidence from ROSSI X-Ray Timing Explorer [GMH+08], this argument is far from being perfect. For instance, in [DCJM11], a Monte Carlo simulation simulation of the astrophysical rate was made, and this used as input for simulating X-ray bursts with a different code, and the results were 5See for instance https://indico.cern.ch/event/1013634/contributions/4313739/ or https://agenda.infn.it/event/27358/contributions/145695/contribution.pdf for examples on recent diffusion of these ongoing projects. 142 quite different with respect to [FTG+07]. Their networks were indeed different, 324 in contrast to Fisker’s 298, which could also introduces differences with respects to β+ decays due to the inclusion of isotopes in or nearby the valley of stability. But notoriously, the conclusion is overall different: no suppressing was found even for low values of the reaction rate. Finally, a critic point of view came some years later, from the same group of research who proposed this argument: in [TGB+09], taking as basis experimental measurements on excited states of 19Ne, favorable to decay into 15O + α, these authors pro- posed a revisioned rate and found their results were quite similar as those by Langanke et al. Moreover, by recalculating some models with Fisker’s code they found the unstable-stable transition in accreting systems, assuming solar composition, took place at ∼ 2 × 1018 g s−1 ≈ 1.8 ṀEdd. Considering the small deviations among versions of this rate (e.g. left panel of Fig. 4.18), we must infer this argument cannot be the “correct” answer to this puzzle, unless future experiments show otherwise. Nevertheless, as pointed out by modern simulations on X-ray bursts (e.g.[MMM19, dCC+21]), uncertainty in certain thermonuclear rates can affect the overall evolution of the accreting process. 8.0 8.5 9.0 9.5 10.0 log10T [K] 10−4 10−1 102 105 108 ρ N A 〈σ v 〉 T [s − 1 ] 15O(β+νe) 15N ρ = 104 g cm−3ρ = 107 g cm−3 15O(α, γ) 19Ne dc11 fs07 Ha96 il10 7 8 9 10 log10T [K] 10−4 10−1 102 105 108 ρ N A 〈σ v 〉 T [s − 1 ] 19Ne(β+νe) 19F ρ = 104 g cm−3ρ = 107 g cm−3 19Ne(α, γ) 20Na cf88 rath ths8 il10 Figure 4.18: Different versions of reaction rates, in units of s−1, as functions of temperature. These fits were taken from JINA-CEE Reaction Library (REACLIB) [CAF+10], preserving the assigned labels to each rate: ths8 are from T. Rauscher and part of REACLIB v1.0, cf88 from the classic [CF88], rath from [RT00], fs07 is based on [FTG+07], dc11 comes from [DCJM11], il10 is based on [ILC+10] and Ha96 is from [HGA+96]. The thickest lines correspond with the recommended version by JINA-CEE library. The horizontal lines represent the β+ decay of the parent nuclides (rates also taken from JINA-CEE). Left panel: 15O(α, γ)19Ne. Right panel: 19Ne(p, γ)20Na. We observed in the past subsection that burst suppression can be in theoretical simulations merely an artifact of the size of the network chosen. Thus, any model 143 that we propose for bursting suppression must be free of this degeneracy - i.e., the model must exhibit bursting suppression employing a large network such as Approx140 or net380. Under these considerations, in this subsection we introduce another possibility to those enlisted above, found by exploring the available options to change parameters in MESA: an increased opacity, which allows to suppress the bursting behavior near 0.3ṀEdd. As inner boundary conditions for MESA at ρb one needs to fixed and inner lumi- nosity Lb, coming from the stellar interior and we chose two extreme values, a very low one of Lb = 2.59 × 10−5L⊙ and high one of Lb = 2.49 × 10−1L⊙. The amount of cells of the spatial grid, as well as the time-step, are factors which might have some impact on the simulation results. In MESA, mesh_delta_coeff parameter controls the mesh reĄnement during a simulation: above 1.0, the number of grid cells tends to be smaller, while below 1.0 the number might reach up to 3000 cells. Unless explicitly stated, we adopt 5.0 for this coefficient. Besides the size of the mesh, MESA allows to have some control over the chosen time step. With time_delta_coeff (hereafter tdc), the user can ask for overall large steps in time, while min_timestep_factor (hereafter mtf) controls the minimum ratio between the new and previous time step. By default, their respective values are 1.0 and 0.8. By trial and error, for some simulations we have found suitable to replace these defaults by a customized conĄguration of 5.0 and 1.2, which we employ unless another combination is explicitly stated. 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 t [hr] 0 2000 4000 6000 8000 L /L ⊙ Ṁ = 5.26× 10−9 M⊙ yr−1 Ṁ = 6.26× 10−9 M⊙ yr−1 Ṁ = 7.26× 10−9 M⊙ yr−1 Ṁ = 9.26× 10−9 M⊙ yr−1 Ṁ = 1.56× 10−8 M⊙ yr−1 Figure 4.19: Luminosity (in units of L⊙) as a function of time for different values of mass accretion rate. The luminosity at the base is equal to 2.59 × 10−5L⊙, the opacity is globally increased by a factor of 10, log10ρb = 9.57, gs,14 = 2.14, approx140 network, mesh= 5.0 and 80Kr at the base. One concern could be that the bursting suppression is also a consequence of our other choices for the parameters, not only of the change in opacity. For example, setting mtf = 1.2 leads to progressively longer time steps, which may inĆuence the 144 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 t [hr] 0 2000 4000 6000 8000 L /L ⊙ Ṁ = 5.26× 10−9 M⊙ yr−1 Ṁ = 6.26× 10−9 M⊙ yr−1 Ṁ = 7.26× 10−9 M⊙ yr−1 Ṁ = 9.26× 10−9 M⊙ yr−1 Ṁ = 1.56× 10−8 M⊙ yr−1 Figure 4.20: Same as Fig. 4.19 but for Lb = 2.4910−1L⊙. evolution of the column by skipping important variations in the burning. To test whether this is the case, we ran a few simulations, keeping the accretion rate Ąxed at 7.26 × 10−9 M⊙ yr−1 (∼ 0.4ṀEdd) as well as an increased opacity of 10κ, but chaniging other parameters. In particular, we focused on the following: • Limits to the time step in the mesh. We modiĄed both time_delta_coeff and min_timestep_factor to favour longer or shorter time steps. • Number of cells in the mesh. This is partially controlled by the user via the mesh_delta_coeff parameter. In our simulations, a value around and above 5 restricts the amount of cells below 500, while with a value around 1 the cells can be as many as 2000 to 3000. • Number of species in the network. To rule out the possibility of the electroweak reactions from the omitted nuclides in the approx140 network altering the bursting behavior, we employed the same network of 380 species, adding neutrons for a fair comparison with the approx140, and thus resulting in 381 species. We refer to this larger network as net381. • Composition of the base. We chose two compositions: either the rp-ashes mixture, or a single-species with a heavy nucleus, i.e. 80Kr, common to both networks we considered. • Composition of accreted material. We explored two compositions: the Solar-like provided by MESA and a simpler mixture of 70% 1H, 28% 4He and 2% 12C, key species in the actual synthesis of heavier elements via the rp-process. 12C is necessary for a fair comparison between net381 and approx140 due to the absence of Li, Be and B isotopes directly connecting 4He with C, N and O in the latter network. 145 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 t [hr] 6.0 6.2 6.4 6.6 6.8 7.0 7.2 lo g 1 0 T [K ] log10ρb[g cm−3] = 9.571 〈gs,14〉 = 2.162 Ṁ = 7.26× 10−9 M⊙ yr−1 approx140, mesh = 5.0 80Kr base 70% 1H, 28% 4He, 2% 12C 10κ Tb Teff mtf = 1.2, tdc = 5.0 Default mtf, tdc = 1.1 Default mtf, tdc = 2.0 Default mtf, tdc = 5.0 Figure 4.21: Effective temperature as a function of time, at Ąxed opacity factor and accretion rate, for different setups of timestep. The resulting light curves can be found in Figs. 4.21 and 4.22. To test the impact of the time-step controls over the stabilization, we ran simu- lations altering both mtf and tdc, employing the approx140 network, 80Kr as the composition of the base and the simpler accretion mixture. In Fig. 4.21 we also plot the temperature at the base of our models. The models discussed in the main text have mtf = 1.2 and tdc = 5.0 (this is the red curve in the Figures), while we test combinations with MESAŠs default value of mtf = 0.8 and tdc = 1.1, 2, 5. The overall similarity of all results with respect to our Ąducial strongly suggest that our previous results are solely due the the change in opacity and not an artifact of the numerical integration. Some discrepancies can still be seen, though they donŠt change the conclusions. For instance, employing the default value for mtf in combination with a large tdc induces an additional burst and delays the stabilization for 0.5 hr. The equilibrium temperature, however, is similar to the rest of the models. The second difference is the time span of the decay phase after the burst peak, which is slightly shorter in the mtf = 1.2 case than in the rest of the simulations. The other important numerical parameter is resolution. The difference in size of the mesh does not produce appreciable deviations, neither in the Ąrst ŞnumericalŤ burst nor in the equilibrium temperature, suggesting that our Ąducial parameters for the resolution are high enough (Fig. 4.22). We now turn to more physical parameters (Fig. 4.22). Regarding the size of the network, although the decay after the peak of the Ąrst burst using net381 is faster than for the approx140 model, the overall behavior is the same: only one burst occurs, followed by a damping process Ąnally converging to a stable state. The oscillations from the net381 network around the equilibrium value are slightly more visible than those generated by the approx140 network, but this difference is 146 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 t [hr] 6.00 6.25 6.50 6.75 7.00 lo g 1 0 T eff [K ] log10ρb[g cm−3] = 9.571 〈gs,14〉 = 2.162 Ṁ = 7.26× 10−9 M⊙ yr−1 10κ approx140 , mesh = 5.0 , rpashes approx140 , mesh = 5.0 , 80Kr approx140 , mesh = 1.0 , 80Kr approx140 , mesh = 5.0 , 80Kr, HHeC net381 , mesh = 5.0 , 80Kr Figure 4.22: Effective temperature as a function of time, at Ąxed opacity factor and accretion rate, for different combinations of parameters. minor. With respect to the base composition we observe little difference in the stabi- lization properties when using either of the ashes or pure 80Kr mixtures, although some differences in the rise and decay phases of the Ąrst burst are visible. The over- all simulation reaches stabilization nonetheless and there is little difference in the equilibrium temperature with respect to the rest of the models. When considering different accreted composition, we notice that the absence of Z > 6 species in the fuel material delays the stabilization process until a second, less-energetic burst have occurred. However, no major changes are noticeable when the burning turns stable. In order to check if there is a dominant term which is responsible for the stabi- lization of the burning, we applied the factor of 10 increase only to κrad as a whole, and then to κff and κes separately. We show also these results in Figs. 4.23 and ??. Modifying only κrad simply delays the Ąrst burst, but does not avoid stabilization. However, noticeably, a global increase only to κff seems to be unable to prevent the instability and we observe multiple bursts. On the other hand, changing the electron scattering opacity by a factor of 10 produces again the observed suppression as in the global 10κrad model. To rule out this as a mere consequence of employing an old Ąt for opacity, we tested both Paczynsy and Poutanen Ąts [Pac83, Pou17] and found out that despite small differences in the rising time and cooling phase after the Ąrst burst, the stabilization still took place. In order to interpret these result, it is useful to look at the depths where each term is dominant. A typical proĄle is shown in Fig. 4.24. Incrementing only κrad has the same effect as multiply the whole κ function because the conduction term dominates only at the highest depths, after ignition and is thus not too relevant for the bursting behaviour. The distinction between κff and κes is more interesting: changing only 147 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 t [hr] 6.00 6.25 6.50 6.75 7.00 7.25 lo g 1 0 T eff [K ] Ṁ = 7.26× 10−9 M⊙ yr−1 mesh = 2.0 , 10κMESA mesh = 2.0 , 10κrad mesh = 2.0 , 10κff mesh = 2.0 , 10κes (Paczynsky) mesh = 2.0 , 10κes (Poutanen) Figure 4.23: Effective temperature as a function of time, for different opacities. Here, log10 ρb = 9.57, gs,14 = 2.14, 80Kr as composition of the base and Approx140. As base luminosity we employ 2.59 × 10−5L⊙. κff re-introduced the bursts indicating that for the opacity to affect the burst, what is really important is an increase at lower densities, above the ignition depth. We do not expect the calculations of κes to be off by a factor of 10, but our results may indicate that if a different mechanism is at work, which increases the opacity in the most superĄcial layers, this could contribute to reconcile the observations with the numerical predictions. However, is ten the minimum increase an opacity should have to suppress bursts? In Fig. 4.25 we show the luminosity curves for two families of models, with different global opacity factors, 2 and 4. Here, the base luminosity is expressed in terms of Teff, via Lb = 4πσSBR 2T 4 eff,b, taking R = 11.57 km - i.e. the radius of an 1.4M⊙ APR-core EOS. The accretion rates correspond to 0.2, 0.3 and 0.4 ṀEdd - all of them around the critical 0.3ṀEdd at which bursts observationally disappear. For simplicity, we adopted ρb = 107 g cm−3. What we observe is that doubling the opacity conĄnes the stabilization of bursts to the interval between 6.75 and 7.0 in log10Teff,b - i.e. we still require large amounts of heat coming from the base to produce bursting suppression. A four-times large opacity, however, reduces the amount of base heating required for burst suppression by 0.25 in log10Teff,b. In terms of Qb, this implies at 4κ we require between 0.3 and 3 MeV per baryon at 0.3ṀEdd to suppress bursts, while at 2κ we require between 3 and 30 MeV per baryon. 148 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 log10ρ [g cm−3] 0.00 0.05 0.10 0.15 0.20 0.25 0.30 κ [c m 2 g− 1 ] κtot κrad κcond κff κes Figure 4.24: Opacity proĄle for a typical stationary accreted envelope at ṀEdd. 10−1 103 Ṁ =3.5071×10−9 M⊙ yr−1 80Kr, approx140, opacity factor = 2.0 10−1 103 L /L ⊙ Ṁ =5.2606×10−9 M⊙ yr−1 log10Teff,b = 5.00 log10Teff,b = 6.00 log10Teff,b = 6.50 log10Teff,b = 6.75 log10Teff,b = 7.00 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 t [hr] 10−1 103 Ṁ =7.0142×10−9 M⊙ yr−1 10−1 103 Ṁ =3.5071×10−9 M⊙ yr−1 80Kr, approx140, opacity factor = 4.0 10−1 103 L /L ⊙ Ṁ =5.2606×10−9 M⊙ yr−1 log10Teff,b = 5.00 log10Teff,b = 6.00 log10Teff,b = 6.50 log10Teff,b = 6.75 log10Teff,b = 7.00 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 t [hr] 10−1 103 Ṁ =7.0142×10−9 M⊙ yr−1 Figure 4.25: Lightcurves for models with different base luminosity - parametrized in terms of Teff. The upper panels, at accretion rates 0.2, 0.3 and 0.4 ṀEdd, corre- spond to models with a doubled-opacity. The lower panels, at same accretion rates, correspond to models with a factor of 4 for the global opacity. 149 Chapter 5 Conclusions “Books are not made to be believed, but to be subjected to inquiry. When we consider a book, we mustn’t ask ourselves what it says but what it means.” U. Eco Abstract. This chapter outlines the main findings of the present thesis, as well as the work in progress for future research. Are Tb − Lb − Teff relations viable? In Chapter 4 we observed that, upon contrast between time-independent and depen- dent simulations, their region of agreement extended from the surface of the star up to ∼ 105 g cm−3 for the temperature and luminosity. Between 105 and 108 g cm−3, however, the timescales of the different processes are actually similar and it is thus necessary to restore to fully time-dependent simulations in order to obtain a correct distribution of energy, and speciĄcally on the amount Ćowing from the surface to the numerical core. Consequently, for the implementation of accreted envelopes at NSCool we can proceed as follows: • Compute, with our stationary code, the Teff − Lb − Tb relations considering ρb ∼ 105 g cm−3. • Implement these relations as boundary conditions for NSCool. • Incorporate the equations for the evolution of the mass fractions into NSCool, in order to account for the speciĄc energy generation rate due to nuclear re- actions as well. 150 These tasks are currently a work in progress. On the energy flowing from the surface to the core From the numerical experiments concerning the change in ρb, we observed that keep- ing a constant Lb had an impact on the distribution of energy Ćowing to the interior of the star. While the usage of small networks such as Approx21 provide a decent Ąrst approximation to this modeling, it is advisable to employ other approximations such as Approx140 in order to check whether bursting behavior ceases or not and, more importantly, to account for the actual energy deposited between ∼ 107 and 109 g cm−3, which can be as large as ∼ 4 MeV per baryon, in contrast to the ∼ 1 MeV per baryon reported by Approx21. Another reason to look for an improvement in the simulation is to extend the simulations beyond ∼ 1 week, i.e. the range of validity for the assumption of Lb constant, which could have an impact on the evolu- tion of the numerical coreŠs temperature given the observed increase in temperature between ∼ 107 and 109 g cm−3, even in the simplest case of the Approx21 network. Given the observed amount of injected energy and its location, ∼ 1 MeV per baryon between 107 and 109 g cm−3, we might argue this could be an additional heating source, similar as to the one required to explain the shallow heating. We must observe, however, that our reported energy emerges during accretion episodes. It might be the case, however, that such injection of energy actually takes place as the mass accretion rate decreases: if, for instance, the inner and outer crust have had ample time to cool down, they will be at lower temperature than the envelope. Thus, with the occurrence of thermonuclear explosions - elevating the temperature at 107 g cm−3 up to 109 K - some energy might Ćow towards the interior and increase its temperature in a range of ∼ 1 day, as observed in our theoretical models for constant accretion. An important take away message from these numerical simulations is that the assumption of Lb > 0, while standard in practice, is not necessarily the only one which we should consider, as states with Lb < 0 are physically viable as well and might account for the distribution of energy around 107 g cm−3, which is also the typical boundary density employed in numerous studies of accreted neutron star envelopes. Below 0.3ṀEdd, a self-consistent modeling of the envelope requires the inclusion of time-dependent terms in the PDEs in order to resolve the bursting behavior and a realistic evolution of the ashes. Above this threshold for mass accretion, however, we can expect a steady-state to be reached in order to explain the absence of bursts, considering the observational basis. However, given the future incorporation of the network into NSCool, it is still necessary to check whether the full evolution of the star actually reaches the steady states predicted by our stationary code. However, we have seen that artiĄcially increasing the opacity, albeit physically questionable due to the absence of a straightforward mechanism to explain such increase, actually 151 allows to suppress the bursting behavior at accretion rates near the critical one, 0.3ṀEdd. 152 Chapter 6 Paper 1: “The Effect of Opacity on Neutron Star Type I X-ray Bursts Quenching” 6.1 Overview The main motivation behind this paper was to test whether changes in opacity could suppress the bursting behavior observed in MESA or not. Part of the contents in this manuscript was introduced at the end of Chapter 4. Link: https://www.astroscu.unam.mx/RMxAA/accepted.html As mentioned in the Preface and in Chapter 4, there exists some tension between theory and observations regarding the Ṁcrit above which bursting suppression - also referred to as burst quenching - takes place. To resolve this issue, additional heating sources are typically invoked at the base of the envelope. Given my present work focused on examining the microphysics of the envelope, as well as in the compari- son of time dependent-independent simulations, eventually the idea of changing the opacity in MESA emerged. From one side, most of - if not all - radiative opacities provided by MESA in tabulated form assume the composition is of primary hydrogen or helium with small traces of carbon and oxygen. Similarly, the analytic Ąt for free-free opacity we employ was constructed by considering computations of hydro- gen, helium, carbon and oxygen as well. Thus, some uncertainty regarding high-Z nuclides exists on these radiative opacities, which is critical for the accreted envelope given the composition of matter after bursts is enriched in these high-Z nuclides. The simulations in this manuscript - all employing MESA - thus had the goal of testing the impact of such opacity uncertainties over the burst quenching. In partic- ular, to check whether an altered opacity could reduce the gap between Ṁcrit from theory and observations. Our results indicate that increasing the opacity by a factor ∼ 10 actually reduces the gap: bursts are no longer observed as Ṁ ≈ 0.40ṀEdd. 153 From our models we infer this behavior is independent of the base luminosity Lb, a situation which also alleviates the tension with respect to the large amounts of base heating required for the burst quenching. By running additional models, we also corroborated this suppression was not a numerical effect but physical, as long as one accepts a dramatic increase in the electron scattering opacity, not in the free-free one. Such result was surprising taken into account the uncertainties above described. Despite the robustness of our theoretical results, at present we could not came up with a physical mechanism responsible of increasing the opacity up to a factor of 10. On one hand, we can relax such stringent factor by invoking a moderate value for Lb with respect to the typical ones. On the other, we are still left without explanation for the ∼ 2 factor. After some research, we noticed our primary candidate - magnetic Ąeld - cannot be responsible as its presence decreases the opacity. The manuscript was submitted to RMxAA on September 17th 2024. On October 16th 2024 a response was received by the Reviewer, stating the manuscript was accepted with minor corrections. On November 4th 2024 the paper was officially accepted, to be published on April 2025. 154 Chapter 7 Paper 2 - Considered for acceptation 7.1 Overview This paper had the purpose of introducing the numerical code I developed for solving the time-independent envelope equations (see Chapter 4). Link: https://arxiv.org/abs/2403.13994 The manuscript was submitted to RASTI (Royal Astronomical Society - Tech- niques and Instruments) on May 28th 2024. On October 21st 2024 a response of ŞConsidered to be accepted under minor correctionsŤ was sent by the Reviewer throughout the Editors. On October 22nd 2024 a revisited version was submit- ted. By November 4th 2024, the status on RASTIŠs ScholarOne Manuscript site is still ŞAwaiting Reviewer ScoresŤ. The requested changes have been kept in red for illustrative purposes. • Introduction. Here we provide a brief review on the state-of-art regarding the modeling of thermal evolution for accreting neutron stars, and present our plans for the present paper and its relevance for future works, where we are employing the output from this time-independent code for computing the tem- perature and luminosity of the envelope in combination with a time-dependent code for the core, NSCool. • Stationary accreting neutron star envelopes. In the Ąrst part of this section we delimit the scope of our code. Since it is time-independent, it is important to state at which stages of the evolution of the whole star our approximations hold. In the second part, we introduce the ODEs governing the structure, temperature and chemical composition of the envelope. For the reasons exposed in Chapters 1 and 4, we restore to pressure as our independent variable instead of radius or mass. Further details on the derivation of these equations can be found in Chapter 1 or in the Appendix of this manuscript. 155 • Methodology. Here we enumerate the boundary conditions for the differ- ential equations and the network of reactions to be employed, net380. As mentioned in Chapter 3, this network is suitable for stationary states stud- ies since we can correctly account for both the distribution of mass fractions and the nuclear energy output. Despite the apparent computational cost (i.e. 384 ordinary differential equations), the absence of the temporal coordinate transform this IVP+BVP into a IVP, where we require to specify the condi- tions at the surface and integrate 384 equations, not 384 x Ncells as is done in time-dependent codes such as MESA. • Results. (i) Our code correctly reproduces the output from the basic refer- ence in the Ąeld, i.e. [SBCW99], although employing a smaller network with updated reactions and binding energies. In our corroboration, we expanded on the contributions to the total nuclear energy from different processes, such as proton captures or the β+ decay. (ii) We showed that variations in the amount of accreted hydrogen, below 50%, could still lead to the occurrence of the rp-process. Furthermore, in (iii) we illustrated that only the presence of hydrogen and helium was sufficient to synthesize the rest of metals, as long as Ṁ ≥ 0.1ṀEdd. While typical approximations at such low accretion rates focus on the CNO and 3α process, here we illustrate some rp-process might still occur. Finally, in (iv) we analyzed some envelopes at 1 and 5 ṀEdd accreting helium, and found out that moderate amounts of carbon-12 and heavier α nuclides were produced at these high accretion rates. • Conclusions. We closed the paper making some punctual observations on the results found. Having a stationary code capable of calculating the temperature, chemical composition and luminosity of envelopes undergoing high accretion rates will allow us to incorporate them into time-dependent codes such as NSCool. Finally, the amounts of A = 24 and A = 28 predicted by helium- 4 burning at high accretion rates provide a theoretical explanation on the possible origin of such α nuclides at high depths. However, it still requires the observational corroboration of having helium, not hydrogen, accretion over the surface of the neutron star. 156 Chapter 8 Paper 3 (Third Author) 8.1 Overview In this paper I collaborated as third author. Link: https://iopscience.iop.org/article/10.3847/1538-4357/ac72a8 My main contribution to this paper was the computation of the electron captures over the α nuclides. The methodology employed comes from [HZ90b], and is brieĆy addressed in Chapter 2 of the present work as well. In summary: under constant accretion, the original composition of the neutron star crust, enriched in metals such as 12C, 16O or 32S, undergoes compression as new material is accreted and burnt. Such compression increases the probabilities for pycnonuclear reactions, such as electron captures, to take place over this compressed matter. Besides liberating heat in the process, it changes the chemical composition by generating neutron-rich species such as 24O. According to the Monte Carlo Markov Chain simulations in the paper, such species should be the fuel for the theoretical explosion that Ąts the observational data. Hence, to corroborate or refute such assertion we restored to the model of Haensel and Zdunik, and found out that species such as 24O or 28Ne could actually exist at the densities where the explosion is likely to take place. Their parent nuclides are 24Mg and 28Si, thus another question is whether we can actually synthesize signiĄcant mass fractions of said nuclides in order to push them up to 1011 g cm−3. According to the models from Chapter 4, the synthesis of 24Mg and 28Si does take place at hydrogen and helium burning conditions, albeit the resulting mass fractions do not exceed 10−2. 157 Appendix A Thermodynamics This appendix is devoted to prove several of the thermodynamical identities and claims employed along the Chapters. As such, it is divided in three sections: in the Ąrst, several propositions and theorems which are more general are stated and proved. The second is devoted to identities which are invoked in the main discussion. A.1 Propositions and theorems The common hypothesis to all the following propositions and theorems is a thermo- dynamical system Σ(T ) with a constant number of baryons NB Proposition 1. Considering fixed abundances, c̆V = ( ∂ε̆ ∂T ) ρB,Y , (A.1) c̆P =   ∂h̆ ∂T   P,Y (A.2) = ( ∂ε̆ ∂T ) P,Y + χT χρ P ρBT . (A.3) Proof. The Ąrst identity directly follows from the First Law, working at dρB = 0 and dYi = 0 for all species, i.e. dε̆ = Tds̆. By taking its ratio with dT , and the limit dT → 0, we see c̆V = ( ∂ε̆ ∂T ) ρB,Y . Now let us consider dh̆ at dP = 0 and dYi = 0 for all species, i.e. dh̆ = Tds̆. Taking its ratio with dT and the limit dT → 0, c̆P =   ∂h̆ ∂T   P,Y . 158 Replacing h̆ = ε̆+ P/ρB and the deĄnition of χT and χρ, we obtain c̆P = ( ∂ε̆ ∂T ) P,Y + χT χρ P ρBT . ■ Proposition 2. The relation between our f̆ and those in the papers is given by f̆int = kBT ⟨m⟩imu fpapers int . (A.4) Proof. Considering the total number of ions divided by volume is given by N/V = ρB/⟨m⟩imu, we have fpapers int = Fint NkBT = 1 kBT V N Fint V = ⟨m⟩imu ρBkBT fint = ⟨m⟩imu kBT f̆int. ■ The second aspect of importance with respect to fpapers int is its implicit depen- dence with respect to both ρB and T , in concrete throughout the ŞΓ factorsŤ, the ratio between electrostatic and thermal energies, as well as the Şelectron degener- acyŤ τ = kBT/EF (following IchimaruŠs notation. Eq. (3.83) of his expresses the exchange energy due to electrons) with EF electronŠs Fermi energy. For simplicity, let us denote each of these parameters as Γi with i ∈ ¶1, . . . , Nθ♢. Such functional dependence is very useful for numerical implementations, as we can see from the following propositions. For simplicity of notation, fpapers int = f . Proposition 3. Proof. Considering f̆ = f̆ideal + f̆interaction, we see P = ρ2      ∂f̆ideal ∂ρ   T,Y +   ∂f̆interaction ∂ρ   T,Y    = Pideal + ρ2   ∂f̆interaction ∂ρ   T . From chainŠs rule, P = Pideal + ρ2kBT ⟨m⟩imu Nθ ∑ j=1 ( ∂f ∂Γj )( ∂Γj ∂ρ ) T . 159 Regarding the partial derivatives: as each ∂f/∂Γj is in itself a functional of the Nθ factors, ( ∂P ∂ρ ) T = ( ∂Pideal ∂ρ ) T + 2ρkBT ⟨m⟩imu Nθ ∑ j=1 ( ∂f ∂Γj )( ∂Γj ∂ρ ) T + ρ2kBT ⟨m⟩imu    Nθ ∑ j=1 ( ∂f ∂Γj )( ∂2Γj ∂ρ2 ) T + Nθ ∑ j=1 ( ∂Γj ∂ρ ) T Nθ ∑ k=1 ∂2f ∂Γk∂Γj ( ∂Γk ∂ρ ) T    ( ∂P ∂T ) ρ = ( ∂Pideal ∂T ) ρ + ρ2kB ⟨m⟩imu Nθ ∑ j=1 ( ∂f ∂Γj )( ∂Γj ∂ρ ) T + ρ2kBT ⟨m⟩imu    Nθ ∑ j=1 ∂f ∂Γj ∂2Γj ∂T∂ρ + Nθ ∑ j=1 ( ∂Γj ∂ρ ) T Nθ ∑ k=1 ∂2f ∂Γk∂Γj ( ∂Γk ∂T ) ρ    . Finally, for constructing the speciĄc heat at constant volume we have   ∂f̆int ∂T   ρ = kBT ⟨m⟩imu Nθ ∑ j=1 ∂f ∂Γj ( ∂Γj ∂T ) ρ + kBf ⟨m⟩imu   ∂2f̆int ∂T 2   ρ = 2kB ⟨m⟩imu Nθ ∑ j=1 ∂f ∂Γj ( ∂Γj ∂T ) ρ + kBT ⟨m⟩imu    Nθ ∑ j=1 ∂f ∂Γj ( ∂2Γj ∂T 2 ) ρ + Nθ ∑ j=1 ( ∂Γj ∂T ) ρ Nθ ∑ k=1 ∂2f ∂Γk∂Γj ( ∂Γk ∂T ) ρ    . We can notice the contributions to the partial derivatives from these interacting terms, for clarity in the discussion, can be written as ( ∂Pint ∂ρ ) T,Y = ξiρT [2T1 + ρT2] ( ∂Pint ∂T ) ρ,Y = ξiρ [ρT1 + ρTT3] c̆ (int) V = −ξiT [2T5 + TT4] , 160 where ξi = kB⟨m⟩−1 i m−1 u and T1 = 3 ∑ l=1   Nl ∑ j=1 ∂fl ∂Γj ( ∂Γj ∂ρ ) T,Y   T2 = 3 ∑ l=1    Nl ∑ j=1 ∂fl ∂Γj ( ∂2Γj ∂ρ2 ) T,Y + Nl ∑ j=1 ( ∂Γj ∂ρ ) T,Y Nl ∑ k=1 ∂2fl ∂Γk∂Γj ( ∂Γk ∂ρ ) T,Y    T3 = 3 ∑ l=1    Nl ∑ j=1 ∂fl ∂Γj ∂2Γj ∂T∂ρ + Nl ∑ j=1 ( ∂Γj ∂ρ ) T,Y Nl ∑ k=1 ∂2fl ∂Γk∂Γj ( ∂Γk ∂T ) ρ,Y    T4 = 3 ∑ l=1    Nl ∑ j=1 ∂fl ∂Γj ( ∂2Γj ∂T 2 ) ρ,Y + Nl ∑ j=1 ( ∂Γj ∂T ) ρ,Y Nl ∑ k=1 ∂2fl ∂Γk∂Γj ( ∂Γk ∂T ) ρ,Y    T5 = 3 ∑ l=1   Nl ∑ j=1 ∂fl ∂Γj ( ∂Γj ∂T ) ρ,Y   On the other hand, by direct differentiation we obtain s̆ = s̆ideal − kBT ⟨m⟩imu [ f T + T5 ] ■ A.2 Maxwell relations and alternative expressions for some partial derivatives of interest From the deĄnitions of dε̆, df̆ , dh̆, dğ and dω̆ respectively, we can show the following relations between partial derivatives hold: ( ∂T ∂ρ ) s̆,Y = 1 ρ2 ( ∂P ∂s̆ ) ρ,Y (A.5) ( ∂s̆ ∂ρ ) T,Y = − 1 ρ2 ( ∂P ∂T ) ρ,Y (A.6) ( ∂T ∂P ) s̆,Y = − 1 ρ2 ( ∂ρ ∂s̆ ) P,Y (A.7) ( ∂s̆ ∂P ) T,Y = 1 ρ2 ( ∂ρ ∂T ) P,Y (A.8) ( ∂s̆ ∂ρ ) T,µ = − 1 ρ2 ( ∂P ∂T ) ρ,µ . (A.9) 161 A.3 Heat capacity in superfluid systems In this appendix we reĄne the argument of [LL76] for the heat capacities at constant pressure and volume, namely, that their difference converges to zero as T → 0 more rapidly than them separately. For this purpose, it is necessary to deĄne what we must understand as rapid convergence, and then move to the physical aspects of the problem. Definition 1. Let f, g : (0, δ) → R + ∪ ¶0♢ be two functions, with δ ∈ R +. We say that f converges more rapidly than g to zero if the following conditions are met: • Their right limits exist: lim x→0+ f(x) = lim x→0+ g(x) = 0 ; (A.10) • ∀x ∈ (0, δ), 0 < f(x) < g(x) . (A.11) We are now ready to discuss the problem of heat capacity. As basic facts we consider that, according to [LL76]: • NernstŠs postulate (or the Third Law of Thermodynamics) holds: the entropy of the system goes to zero as T → 0. • The heat capacity at constant X, being this variable a thermodynamical one, can always be written as CX = T ( ∂S ∂T ) X (A.12) • The difference between heat capacities can be expressed as CP − CV = −T ( ∂P ∂V ) T ( ∂S ∂P )2 T (A.13) From the Ąrst point, we replace the argument in [LL76] of considering S ∝ T b, b ∈ N, by a more general expansion around T = 0: S(P, V, T ) = ∞ ∑ n=1 sn(P, V )τn (A.14) with sn(P, V ) = ( ∂S ∂T ) T=0,P,V , (A.15) 162 and τ = T/TF , with TF a positive scale parameter (which can be identiĄed with the Fermi temperature for fermion systems), [TF ] = K, that serves to work with a dimensional variables. By replacing the series expansion into the heat capacities, we can see that CP,V = ∞ ∑ n=1 sn(P, V )τn (A.16) CP − CV = ∞ ∑ n=1 σnτ n+1 (A.17) σn = − n ∑ m=1 TF ( ∂P ∂V ) T ( ∂sm ∂P ) V ( ∂sn−m ∂P ) V . (A.18) Now, if τ < 1, τn+1 < τn holds (Proof: Being τn positive, multiplying the inequality τ < 1 by this value does not alter the order of the terms. Thus, τn+1 < τn. ■). All that is left is to prove that ( ∂P ∂V ) T is negative. This can be seen from physical argumentation: if we increase the pressure (δP > 0), the occupied volume tends to decrease (δV < 0). Similarly, if we expand the region where the system exists (δV > 0), the pressure exerted diminishes (δP < 0). Thus, σn is positive ∀n ∈ N and the following inequalities hold: CP > CP − CV , (A.19) CV > CP − CV . (A.20) By deĄnition, this means that the difference between heat capacities tend to zero more rapidly than the individual capacities. In the particular case of fermion sys- tems, this result implies that the heat capacities are almost indistinguishable from one another, and it is safe to assume that CP ≈ CV . Before claiming that the proof is complete, it is necessary to examine the follow- ing subtlety: if we replace the expansions for CP,V into the difference CP − CV and compare coefficients with the expansion involving the σn-coefficients, we obtain sn(Pfixed, V ) − sn(P, Vfixed) = −T n ∑ m=1 ( ∂P ∂V ) T ( ∂sm ∂P ) V ( ∂sn−m ∂P ) V . (A.21) Apparently, the right hand side is not independent of temperature. We can be sure that it actually is: the EoS opens the possibility to write T as a function of P and V , i.e. T = f1(P, V ). In addition, the evaluation of the partial derivative of P with respect to V at a Ąxed temperature is mathematically equivalent to introduce a function f2(P, V, Tfixed) = f2(P, V ). Therefore, the right hand side is independent of temperature, as the left hand side demands, and our proof is thus consistent and complete. 163 Appendix B General Relativity This appendix introduces very speciĄc aspects of General Relativity which are in- dispensable for deducing and understanding the Ćuid-element expressions discussed in the main text. As such, they cannot be considered a review on the subject and we urge readers not familiar with the terminology to consult the excellent books on the subject. B.1 On the conservation laws When postulating the existence of physical conserved quantities, it is of universal agreement that we must start by associating to each of the variables of interest a 4-current Jµ and understand as conservation the identity 1√−g∂α (√−gJα ) = 0 , (B.1) i.e. the 4-current must have zero divergence. In the astrophysical context, for example, it can be relevant to seek the inverse approach, that is: can we deĄne a charge and claim that it is conserved? For instance, we can be interested on extending the baryon number conservation, and verify if our intuition that the total number of them can be written as Nb = ∫ Σ d3x √ γ nb , (B.2) i.e., as an integration of the baryon number density n over the proper volume element at a fixed time, i.e. on a Cauchy hyper-surface. As a starting point, let us review the conservation of the total electric charge. We can postulate (or derive, via NoetherŠs Theorem) the existence of a conserved 4-current further referred as the current density 4-vector, ∂µJ µ = 0 . (B.3) 164 Now let us consider a Cauchy hyper surface Σ, with boundary ∂Σ where the spatial part of the 4-current identically vanishes. By virtue of GaussŠ Theorem, ∫ Σ d3x ∂0J 0 = − ∫ Σ d3x ∂iJ i = − ∫ ∂Σ d2x kiJ i = 0 , (B.4) where ki is the 1-form associated with the unitary normal vector to ∂Σ. Thus, we can see that ∂Q ∂x0 = 0 , (B.5) where we have introduced the conserved electric charge as Q = ∫ Σ d3x J0 . (B.6) Now let us take an analogous route for the curvilinear coordinated case. Typically, we can expect conservation with respect to the time coordinate, so the Cauchy surface must be chosen such that ct is constant. In addition, the invariance of the integrals (with the appropriate measures) with respect to the coordinated system allows us to express the quantities involved in an orthogonal frame. This implies that the determinant of the metric tensor gαβ is related to that for the spatial-one γαβ by √−g = √ ♣ − g00♣ √ γ . (B.7) Note that √ g00 is the 0 component of the 4-dual associated with the normal unitary vector to Σ, i.e. k0 = √ ♣ − g00♣ . (B.8) Thus, our discussion can start from A = ∫ Σ d3x √ γ [ k0√−g∂0 (√−gJ0 ) ] . (B.9) The reason to choose this form is twofold: Ąrst, the factor outside ∂0 yields 1/ √ γ while the one inside simpliĄes to √ γk0J 0, and we can now introduce the spatial part of the conserved 4-current. With these considerations, it is more evident that we now have a true 3-divergence which can be cast into a 2-surface integral by virtue of GaussŠ Theorem. Denoting the induced 2-metric and orthogonal vector with a 165 super-index (2), we have A = ∫ Σ d3x √ γ [ 1√ γ ∂0 (√ γk0J 0 ) ] = ∫ Σ d3x √ γ [ 1√ γ ∂i (√ γk0J i ) ] = ∫ ∂Σ d2x √ γ(2)k0k (2) i J i . (B.10) Under the assumption that J i identically vanishes at ∂Σ, it follows that A = 0. From its deĄnition, it is evident that the conserved quantity with respect to x0 variations is Q = ∫ Σ d3x √ γk0J 0 . (B.11) Now, this expression was derived in an orthogonal frame. In order to be valid in any coordinated system, we can extend the product of 0 components to the scalar product of the unitary normal vector to Σ with the 4-current: Q = ∫ Σ d3x √ γkµJ µ . (B.12) For a more formal discussion - leading to this same expression - see [CCAW04]. B.2 SSS metric - Identities In matrix notation, the SchwarzschildŠs metric is deĄned as [gαβ] :=      − exp(2Φ(r)) 0 0 0 0 exp(2Λ(r)) 0 0 0 0 r2 0 0 0 0 r2 sin2 θ      Some of its properties are: • Inverse [gαβ] :=      − exp(−2Φ(r)) 0 0 0 0 exp(−2Λ(r)) 0 0 0 0 r−2 0 0 0 0 r−2 sin−2 θ      . • Christoffel Symbols. The entries in the following matrices respect the order of the coordinated system, i.e. (ct, r, θ, ϕ), going from left to right and from 166 top to bottom. For instance, the 1, 0 entry of [Γctαβ] is equivalent to Γctr ct. [Γtαβ] =      0 ∂rΦ 0 0 ∂rΦ 0 0 0 0 0 0 0 0 0 0 0      [Γrαβ] =      exp(2(Φ − Λ))∂rΦ 0 0 0 0 ∂rΛ 0 0 0 0 −r exp(−2Λ) 0 0 0 0 −r sin2 θ exp(−2Λ)      [Γθαβ] =      0 0 0 0 0 0 1/r 0 0 1/r 0 0 0 0 0 − sin θ cos θ      , [Γϕαβ] =      0 0 0 0 0 0 0 1/r 0 0 0 cot θ 0 1/r cot θ 0      . • Metric determinant. Being a diagonal metric, it is given by the product of all entries, g = −e2Φ+2Λr4 sin2 θ. Consequently, √−g = eΦ+Λr2 sin θ. The spatial determinant obeys √ γ = eΛr2 sin θ. Clearly, √−g = eΦ√ γ. • 4-divergence. In terms of both coordinated and non-coordinated (denoted with hats) systems, for an arbitrary 4-vector Aα we have ∇αA α = 1√−g∂α (√−gAα ) = ∂0A 0 + 1 eΦ+Λr2 ∂r ( eΦ+Λr2Ar ) + 1 sin θ ∂θ ( sin θAθ ) + ∂ϕA ϕ (B.13) = e−Φ∂0 ( A0̂ ) + e−Φ−Λ r2 ∂r ( eΦr2Ar̂ ) + 1 r sin θ [ ∂θ ( sin θAθ̂ ) + ∂ϕA ϕ̂ ] . (B.14) 167 B.3 Propositions and theorems Proposition 4. The SSS metric and the normalized 4-velocity of the fluid satisfy the following properties: (a) [ g00Γ0 r0 + grrΓr00 ] = 0 (b) u0∂rΦ = e−Φ∂r(eΦu0) − ∂ru 0 (c) ur∂rΛ = e−Λ∂r(eΛur) − ∂ru r (d) u0e −Φ = −eΦu0 (e) ure −Λ = eΛur (f) [ u0u0Γ0 0r + ururΓrrr ] = −u0∂ru 0 − ur∂ru r (g) u0∂0u0 = −ur∂0ur. Proof. For (a): employing the deĄnitions of the metric and uµ, by direct computa- tion we have [ g00Γ0 r0 + grrΓr00 ] = −e2Φ∂rΦ + e2Λe2Φ−2Λ∂rΦ = 0. For (b) and (c), we start from ∂ru 0 and ∂ru r and rearrange terms in favor of the partial derivatives of Φ and Λ: ∂ru 0 = e−Φ∂r(eΦu0) − u0∂rΦ u0∂rΦ = e−Φ∂r(eΦu0) − ∂ru 0 ∂ru r = e−Λ∂r(eΛur) − ur∂rΛ ur∂rΛ = e−Λ∂r(eΛur) − ∂ru r. It is easy to see (d) and (e) are simple consequences of the deĄnition of the 4- velocity Ąeld and the raising and lowering of components in the SSS metric, which is a diagonal one. The properties so far discussed can be now used to prove (f): [ u0u0Γ0 0r + ururΓrrr ] = u0∂rΦu0 + ur∂rΛur = −u0∂ru 0 − ur∂ru r + u0e −Φ∂r(eΦu0) + ure −Λ∂r(eΛur) = −u0∂ru 0 − ur∂ru r − 1 2 ∂r [ eΦu0 ]2 + 1 2 ∂r [ eΛur ]2 . Notice [ (eΦu0)2 − (eΛur)2 ] = γ2 S [ 1 − v2 c2 ] = 1, 168 hence [ u0u0Γ0 0r + ururΓrrr ] = −u0∂ru 0 − ur∂ru r. Finally, (g) follows from the normalization of the 4-velocity, g00∂0(u0)2 + grr∂0(ur)2 = 0, u0∂0u0 = −ur∂0ur. ■ Proposition 5. The components of the correspondent 4-acceleration to the 4-velocity field are a0 = u0∂0u0 + ur∂ru0 = γ2 Sv c { γ2 S c2 [ ∂tv + eΦ−Λv∂rv ] − eΦ−ΛdΦ dr } ar = u0∂0ur − u0∂ru0 = −u0 ur a0 aθ = 0 aϕ = 0. Proof. Let us Ąrst deduce the general expressions for the 4 components of aµ. Em- ploying the results of Proposition 4, it is easy to see a0 = u0∂0u0 + ur∂ru0 − [ u0Γσ00 + urΓσr0 ] uσ = u0∂0u0 + ur∂ru0 − uru0 [ g00Γ0 r0 + grrΓr00 ] = u0∂0u0 + ur∂ru0, ar = u0∂0ur + ur∂rur − [ u0Γσ0r + urΓσrr ] uσ = u0∂0ur + ur∂rur − [ u0u0Γ0 0r + ururΓrrr ] = u0∂0ur + ur∂rur + u0∂ru 0 + ur∂ru r = u0∂0ur + u0∂ru 0 + ∂r(urur) = u0∂0ur + u0∂ru 0 − ∂r(u0u0) = u0∂0ur − u0∂ru0 = −u0 ur [−ur∂0ur + ur∂ru0] = −u0 ur [ u0∂0u0 + ur∂ru0 ] = −u0 ur a0. For aθ and αϕ: due to uθ = uϕ = 0, the 8 partials ∂µuθ and ∂µuϕ are identically zero. On the other hand, the terms associated to the Christoffel symbols vanish as well since Γµθν and Γµϕν are identically zero for all µ, ν combinations. Thus, aθ = aϕ = 0. Finally, to show the explicit version of a0 it suffices to replace the components of the normalized 4-velocity and to consider ∂µγS = −γ3 S v c2 ∂µv 169 for all 4 coordinates µ: a0 = − [ u0∂0 + ur∂r ] (γSeΦ) = − { γS [ u0∂0 + ur∂r ] (eΦ) + eΦ [ u0∂0 + ur∂r ] (γS) } = − { γSu reΦdΦ dr − eΦγ3 S v c2 [ u0∂0 + ur∂r ] v } = γ2 Sv c { γ2 S c2 [ ∂t + veΦ−Λ∂r ] v − eΦ−ΛdΦ dr } . ■ Proposition 6. The F r component, in the first-order approximation, is given by F r ≈ −Ke−2Λ−Φ∂r(eΦT ). Proof. By straightforward calculation, we have F r = −K ¶Πrν∂νT + ΠrνTaν♢ = −K { γ2 Se −Λ−Φ [ vS c ∂t + eΦ−Λ∂r ] T + Tgrrar } = −Kγ2 Se −Λ−Φ { [ vS c ∂t + eΦ−Λ∂r ] T + T [ −γSe Φ c uµ∂µv + eΦ−ΛdΦ dr ]} . Neglecting the vs/c term (Ąrst-order approximation), setting γS ≈ 1 and rearranging terms, we obtain F r ≈ −Kγ2 Se −Λ−Φ { e−Λ∂r(eΦT ) − Tγ2 S/c 2 [ ∂t + veΦ−Λ∂r ] v } ≈ −Ke−Λ−Φ { e−Λ∂r(eΦT ) − T c2 [ ∂t + veΦ−Λ∂r ] v } . Notice the last term in the right-hand side is also of order vS/c, hence we can neglect it in our Ąrst-order approximation. Consequently, F r ≈ −Ke−2Λ−Φ∂r(eΦT ). ■ Proposition 7. Within the FOA, the contribution of the 4-flux to the entropy equa- tion and the corresponding luminosity are given by Gs,T = − 1 T { 1 e2Φ+Λr2 ∂r(e2Φ+Λr2F r) } L = −4πr2Ke−Φ−Λ∂r(eΦT ). 170 Proof. Let us separately compute each term in Gs,T . It is easy to see ∇µF µ = 1√−g [√−geΛ−Φ∂0(vSF r) + ∂r( √−gF r) ] . Regarding the projection of F µ along aµ, we have: F νaν = vSe Λ−ΦF ra0 − F ru 0a0 ur = F ra0e Λ−Φ [ vS − 1 vS ] = −F r e Λ−Φ γ2 SvS a0 = −F reΛ−Φ { γSe Φ c uµ∂µv − eΦ−ΛdΦ dr } ≈ F r dΦ dr . Notice we have employed the Ąrst-order approximation for the last line. Since 1 e2Φ+Λr2 ∂r(e2Φ+Λr2F r) = 1 eΦ+Λr2 ∂r(eΦ+Λr2F r) + F r dΦ dr , we see ∇νF ν + F νaν = eΛ−Φ c ∂t(vSF r) + 1 e2Φ+Λr2 ∂r(e2Φ+Λr2F r). Finally, the presence of vS/c in the Ąrst term allows us to write, within this Ąrst-order approximation, Gs,T = − 1 T { 1 e2Φ+Λr2 ∂r(e2Φ+Λr2F r) } . Notice our expression for Gs,T , to Ąrst order in vS, is equivalent as the one we obtain for a Ćuid at rest, uct = e−Φ, where the only non-vanishing component of the associated energy-momentum tensor is T r0rad, i.e. uν∇µT µν rad = uν√−g∂µ( √−gT µν) = u0√−g∂r( √−gT r0) = u0√−g∂r( √−gF ru0). 171 For computing the luminosity, we exploit this equivalence: L = − ∫ Σ d3x √ γ u0√−g∂r( √−gF ru0) = ∫ Σ d3x eΛr2 sin θeΦ eΦ+Λr2 sin θ ∂r(eΦ+Λr2 sin θF re−Φ) = ∫ Σ dϕdθdr sin θ∂r(eΛr2F r) = 4πr2eΛF r = −4πr2Ke−Φ−Λ∂r(eΦT ). where the angular integral was immediate due to the absence of dependence on θ and ϕ of F r, and in the last line we employed the result from Proposition 6. ■ Proposition 8. Considering s̆(P, T, ¶Yi♢Nspecies i=1 ), and under the assumption (∂s̆/∂Yi)P,T,Yk ̸=i are negligible12, TDts̆ = c̆P [ DtT − T∇ad P DtP ] . Proof. Since s̆ is a functional of several variables, by direct calculation we have TDts̆ = T ( ∂s̆ ∂P ) T,Y DtP + T ( ∂s̆ ∂T ) P,Y DtT + T Nspecies ∑ i=1 ( ∂s̆ ∂Yi ) T,P,Yk ̸=i DtYi. Since (∂s̆/∂Yi)P,T,Yk ̸=Yi are assumed negligible, it suffices to employ thermodynamical identities to simply the remaining coefficients: TDts̆ = T ( ∂s̆ ∂P ) T,Y DtP + T ( ∂s̆ ∂T ) P,Y DtT = c̆P [ DtT − T∇ad P DtP ] . ■ Proposition 9. Within the FOA, the radiative temperature gradient is ∇r = − 3κρBLPeΛ 64πσSBT 4r2 dr dP + ( 1 − ρ H ) . 1Or already part of ˙̆εspecies. 2Depending on the context, it is sometimes necessary to consider s̆(ρB , T, ¶Yi♢Nspecies i=1 ). The steps are analogous, employing the appropriate thermodynamical coefficient for DtρB . 172 Proof. ∂rT = − L 4πr2e−ΛK − T dΦ dr dT dP = − L 4πr2e−ΛK dr dP + T Hc2 ∇r = P T dT dP = − LP 4πr2e−ΛKT dr dP + P Hc2 = − LP 4πr2e−ΛKT dr dP + ( 1 − ρ H ) = − 3κρBLPeΛ 64πσSBT 4r2 dr dP + ( 1 − ρ H ) . ■ Proposition 10. For a fluid at rest, and within the SSS metric, the cooling rate can be written as Γcool = − 1 T { 1 e2ΦeΛr2 ∂r ( e2ΦeΛr2F r ) + 1 sin θ ∂θ ( sin θF θ ) + ∂ϕF ϕ } . Proof. From repetitive application of the product rule, we have F r∂rΦ = F re−ΦeΦ∂rΦ = e−Φ∂r(eΦ)F r = e−Φe−Λr−2∂r(eΦ)F reΛr2 = e−Φe−Λr−2 [ ∂r ( F reΦeΛr2 ) − eΦ∂r ( F reΛr2 )] = e−Φe−Λr−2 [ ∂r ( eΦA ) − eΦ∂rA ] , (B.15) where we have deĄned A = r2eΛF r for simplicity. Back to the cooling rate, we have Γcool = 1 T { 1 eΦeΛr2 [ −2∂r(eΦA) + eΦ∂rA ] − ∆ } , (B.16) ∆ = 1 sin θ ∂θ ( sin θF θ ) + ∂ϕF ϕ . (B.17) From ∂r(e2ΦA) = e2Φ∂rA+ 2eΦ ( ∂re Φ ) A (B.18) = eΦ∂r ( eΦA ) + ( ∂re Φ ) eΦA , (B.19) we can see that Γcool = − 1 T { 1 e2ΦeΛr2 ∂r ( e2ΦeΛr2F r ) + 1 sin θ ∂θ ( sin θF θ ) + ∂ϕF ϕ } . (B.20) This completes the proof. ■ 173 B.4 Radiative Zero Solution Let us go back to the differential equation for the temperature, dT dP = 3κ 16T 3 F σgs . (B.21) Under the assumption that the right hand side becomes temperature but not pres- sure or density dependent, we can immediately integrate this equation. If we addi- tionally assume that the lower limits of integrations are negligible in comparison to the upper ones, we arrive to T 4 = ( 3κ 16σgs × 4F ) P . (B.22) This is known as the radiative zero solution, and it expresses the relation between pressure and temperature in the case radiation transport dominates, but hydrostatic equilibrium still holds. 174 Bibliography [AAA+17] B. P. Abbott, R. Abbott, T. D. Abbott, F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso, R. X. Adhikari, V. B. Adya, C. Af- feldt, M. Afrough, B. Agarwal, M. Agathos, K. Agatsuma, N. Aggar- wal, O. D. Aguiar, L. Aiello, A. Ain, P. Ajith, B. Allen, G. Allen, A. Al- locca, P. A. Altin, A. Amato, A. Ananyeva, S. B. Anderson, W. G. Anderson, S. V. Angelova, S. Antier, S. Appert, K. Arai, M. C. Araya, J. S. Areeda, N. Arnaud, K. G. Arun, S. Ascenzi, G. Ashton, M. Ast, S. M. Aston, P. Astone, D. V. Atallah, P. Aufmuth, C. Aulbert, K. AultONeal, C. Austin, A. Avila-Alvarez, S. Babak, P. Bacon, M. K. M. Bader, S. Bae, M. Bailes, P. T. Baker, F. Baldaccini, G. Bal- lardin, S. W. Ballmer, S. Banagiri, J. C. Barayoga, S. E. Barclay, B. C. Barish, D. Barker, K. Barkett, F. Barone, B. Barr, L. Bar- sotti, M. Barsuglia, D. Barta, S. D. Barthelmy, J. Bartlett, I. Bartos, R. Bassiri, A. Basti, J. C. Batch, M. Bawaj, J. C. Bayley, M. Bazzan, B. Bécsy, C. Beer, M. Bejger, I. Belahcene, A. S. Bell, B. K. Berger, G. Bergmann, S. Bernuzzi, J. J. Bero, C. P. L. Berry, D. Bersanetti, A. Bertolini, J. Betzwieser, S. Bhagwat, R. Bhandare, I. A. Bilenko, G. Billingsley, C. R. Billman, J. Birch, R. Birney, O. Birnholtz, S. Biscans, S. Biscoveanu, A. Bisht, M. Bitossi, C. Biwer, M. A. Bi- zouard, J. K. Blackburn, J. Blackman, C. D. Blair, D. G. Blair, R. M. Blair, S. Bloemen, O. Bock, N. Bode, M. Boer, G. Bogaert, A. Bohe, F. Bondu, E. Bonilla, R. Bonnand, B. A. Boom, R. Bork, V. Boschi, S. Bose, K. Bossie, Y. Bouffanais, A. Bozzi, C. Bradaschia, P. R. Brady, M. Branchesi, J. E. Brau, T. Briant, A. Brillet, M. Brinkmann, V. Brisson, P. Brockill, J. E. Broida, A. F. Brooks, D. A. Brown, D. D. Brown, S. Brunett, C. C. Buchanan, A. Buikema, T. Bulik, H. J. Bulten, A. Buonanno, D. Buskulic, C. Buy, R. L. Byer, M. Cabero, L. Cadonati, G. Cagnoli, C. Cahillane, J. Calderón Bustillo, T. A. Callister, E. Calloni, J. B. Camp, M. Canepa, P. Canizares, K. C. Cannon, H. Cao, J. Cao, C. D. Capano, E. Capocasa, F. Carbognani, S. Caride, M. F. Carney, G. Carullo, J. Casanueva Diaz, C. Casen- tini, S. Caudill, M. Cavaglià, F. Cavalier, R. Cavalieri, G. Cella, C. B. 175 Cepeda, P. Cerdá-Durán, G. Cerretani, E. Cesarini, S. J. Chamber- lin, M. Chan, S. Chao, P. Charlton, E. Chase, E. Chassande-Mottin, D. Chatterjee, K. Chatziioannou, B. D. Cheeseboro, H. Y. Chen, X. Chen, Y. Chen, H.-P. Cheng, H. Chia, A. Chincarini, A. Chi- ummo, T. Chmiel, H. S. Cho, M. Cho, J. H. Chow, N. Christensen, Q. Chu, A. J. K. Chua, S. Chua, A. K. W. Chung, S. Chung, G. Ciani, R. CiolĄ, C. E. Cirelli, A. Cirone, F. Clara, J. A. Clark, P. Clearwater, F. Cleva, C. Cocchieri, E. Coccia, P.-F. Cohadon, D. Cohen, A. Colla, C. G. Collette, L. R. Cominsky, M. Constan- cio, L. Conti, S. J. Cooper, P. Corban, T. R. Corbitt, I. Cordero- Carrión, K. R. Corley, N. Cornish, A. Corsi, S. Cortese, C. A. Costa, M. W. Coughlin, S. B. Coughlin, J.-P. Coulon, S. T. Countryman, P. Couvares, P. B. Covas, E. E. Cowan, D. M. Coward, M. J. Cow- art, D. C. Coyne, R. Coyne, J. D. E. Creighton, T. D. Creighton, J. Cripe, S. G. Crowder, T. J. Cullen, A. Cumming, L. Cunningham, E. Cuoco, T. Dal Canton, G. Dálya, S. L. Danilishin, S. DŠAntonio, K. Danzmann, A. Dasgupta, C. F. Da Silva Costa, V. Dattilo, I. Dave, M. Davier, D. Davis, E. J. Daw, B. Day, S. De, D. DeBra, J. Degallaix, M. De Laurentis, S. Deléglise, W. Del Pozzo, N. Demos, T. Denker, T. Dent, R. De Pietri, V. Dergachev, R. De Rosa, R. T. DeRosa, C. De Rossi, R. DeSalvo, O. de Varona, J. Devenson, S. Dhurandhar, M. C. Díaz, T. Dietrich, L. Di Fiore, M. Di Giovanni, T. Di Giro- lamo, A. Di Lieto, S. Di Pace, I. Di Palma, F. Di Renzo, Z. Doc- tor, V. Dolique, F. Donovan, K. L. Dooley, S. Doravari, I. Dor- rington, R. Douglas, M. Dovale Álvarez, T. P. Downes, M. Drago, C. Dreissigacker, J. C. Driggers, Z. Du, M. Ducrot, R. Dudi, P. Dupej, S. E. Dwyer, T. B. Edo, M. C. Edwards, A. Effler, H.-B. Eggenstein, P. Ehrens, J. Eichholz, S. S. Eikenberry, R. A. Eisenstein, R. C. Es- sick, D. Estevez, Z. B. Etienne, T. Etzel, M. Evans, T. M. Evans, M. Factourovich, V. Fafone, H. Fair, S. Fairhurst, X. Fan, S. Fari- non, B. Farr, W. M. Farr, E. J. Fauchon-Jones, M. Favata, M. Fays, C. Fee, H. Fehrmann, J. Feicht, M. M. Fejer, A. Fernandez-Galiana, I. Ferrante, E. C. Ferreira, F. Ferrini, F. Fidecaro, D. Finstad, I. Fiori, D. Fiorucci, M. Fishbach, R. P. Fisher, M. Fitz-Axen, R. Flaminio, M. Fletcher, H. Fong, J. A. Font, P. W. F. Forsyth, S. S. Forsyth, J.-D. Fournier, S. Frasca, F. Frasconi, Z. Frei, A. Freise, R. Frey, V. Frey, E. M. Fries, P. Fritschel, V. V. Frolov, P. Fulda, M. Fyffe, H. Gabbard, B. U. Gadre, S. M. Gaebel, J. R. Gair, L. Gammaitoni, M. R. Ganija, S. G. Gaonkar, C. Garcia-Quiros, F. GaruĄ, B. Gateley, S. Gaudio, G. Gaur, V. Gayathri, N. Gehrels, G. Gemme, E. Genin, A. Gennai, D. George, J. George, L. Gergely, V. Germain, S. Ghonge, Abhirup Ghosh, Archisman Ghosh, S. Ghosh, J. A. Giaime, K. D. Gi- 176 ardina, A. Giazotto, K. Gill, L. Glover, E. Goetz, R. Goetz, S. Gomes, B. Goncharov, G. González, J. M. Gonzalez Castro, A. Gopakumar, M. L. Gorodetsky, S. E. Gossan, M. Gosselin, R. Gouaty, A. Grado, C. Graef, M. Granata, A. Grant, S. Gras, C. Gray, G. Greco, A. C. Green, E. M. Gretarsson, P. Groot, H. Grote, S. Grunewald, P. Grun- ing, G. M. Guidi, X. Guo, A. Gupta, M. K. Gupta, K. E. Gushwa, E. K. Gustafson, R. Gustafson, O. Halim, B. R. Hall, E. D. Hall, E. Z. Hamilton, G. Hammond, M. Haney, M. M. Hanke, J. Hanks, C. Hanna, M. D. Hannam, O. A. Hannuksela, J. Hanson, T. Hard- wick, J. Harms, G. M. Harry, I. W. Harry, M. J. Hart, C.-J. Haster, K. Haughian, J. Healy, A. Heidmann, M. C. Heintze, H. Heitmann, P. Hello, G. Hemming, M. Hendry, I. S. Heng, J. Hennig, A. W. Hep- tonstall, M. Heurs, S. Hild, T. Hinderer, W. C. G. Ho, D. Hoak, D. Hofman, K. Holt, D. E. Holz, P. Hopkins, C. Horst, J. Hough, E. A. Houston, E. J. Howell, A. Hreibi, Y. M. Hu, E. A. Huerta, D. Huet, B. Hughey, S. Husa, S. H. Huttner, T. Huynh-Dinh, N. Indik, R. Inta, G. Intini, H. N. Isa, J.-M. Isac, M. Isi, B. R. Iyer, K. Izumi, T. Jacqmin, K. Jani, P. Jaranowski, S. Jawahar, F. Jiménez-Forteza, W. W. Johnson, N. K. Johnson-McDaniel, D. I. Jones, R. Jones, R. J. G. Jonker, L. Ju, J. Junker, C. V. Kalaghatgi, V. Kalogera, B. Kamai, S. Kandhasamy, G. Kang, J. B. Kanner, S. J. Kapadia, S. Karki, K. S. Karvinen, M. Kasprzack, W. Kastaun, M. Katolik, E. Katsavounidis, W. Katzman, S. Kaufer, K. Kawabe, F. Kéfélian, D. Keitel, A. J. Kemball, R. Kennedy, C. Kent, J. S. Key, F. Y. Khalili, I. Khan, S. Khan, Z. Khan, E. A. Khazanov, N. Kijbun- choo, Chunglee Kim, J. C. Kim, K. Kim, W. Kim, W. S. Kim, Y.- M. Kim, S. J. Kimbrell, E. J. King, P. J. King, M. Kinley-Hanlon, R. Kirchhoff, J. S. Kissel, L. Kleybolte, S. Klimenko, T. D. Knowles, P. Koch, S. M. Koehlenbeck, S. Koley, V. Kondrashov, A. Kontos, M. Korobko, W. Z. Korth, I. Kowalska, D. B. Kozak, C. Krämer, V. Kringel, B. Krishnan, A. Królak, G. Kuehn, P. Kumar, R. Ku- mar, S. Kumar, L. Kuo, A. Kutynia, S. Kwang, B. D. Lackey, K. H. Lai, M. Landry, R. N. Lang, J. Lange, B. Lantz, R. K. Lanza, S. L. Larson, A. Lartaux-Vollard, P. D. Lasky, M. Laxen, A. Lazzarini, C. Lazzaro, P. Leaci, S. Leavey, C. H. Lee, H. K. Lee, H. M. Lee, H. W. Lee, K. Lee, J. Lehmann, A. Lenon, E. Leon, M. Leonardi, N. Leroy, N. Letendre, Y. Levin, T. G. F. Li, S. D. Linker, T. B. Littenberg, J. Liu, X. Liu, R. K. L. Lo, N. A. Lockerbie, L. T. Lon- don, J. E. Lord, M. Lorenzini, V. Loriette, M. Lormand, G. Losurdo, J. D. Lough, C. O. Lousto, G. Lovelace, H. Lück, D. Lumaca, A. P. Lundgren, R. Lynch, Y. Ma, R. Macas, S. Macfoy, B. Machenschalk, M. MacInnis, D. M. Macleod, I. Magaija Hernandez, F. Magaija San- 177 doval, L. Magaija Zertuche, R. M. Magee, E. Majorana, I. Maksimovic, N. Man, V. Mandic, V. Mangano, G. L. Mansell, M. Manske, M. Man- tovani, F. Marchesoni, F. Marion, S. Márka, Z. Márka, C. Markakis, A. S. Markosyan, A. Markowitz, E. Maros, A. Marquina, P. Marsh, F. Martelli, L. Martellini, I. W. Martin, R. M. Martin, D. V. Martynov, J. N. Marx, K. Mason, E. Massera, A. Masserot, T. J. Massinger, M. Masso-Reid, S. Mastrogiovanni, A. Matas, F. Matichard, L. Ma- tone, N. Mavalvala, N. Mazumder, R. McCarthy, D. E. McClelland, S. McCormick, L. McCuller, S. C. McGuire, G. McIntyre, J. McIver, D. J. McManus, L. McNeill, T. McRae, S. T. McWilliams, D. Meacher, G. D. Meadors, M. Mehmet, J. Meidam, E. Mejuto-Villa, A. Melatos, G. Mendell, R. A. Mercer, E. L. Merilh, M. Merzougui, S. Meshkov, C. Messenger, C. Messick, R. Metzdorff, P. M. Meyers, H. Miao, C. Michel, H. Middleton, E. E. Mikhailov, L. Milano, A. L. Miller, B. B. Miller, J. Miller, M. Millhouse, M. C. Milovich-Goff, O. Minaz- zoli, Y. Minenkov, J. Ming, C. Mishra, S. Mitra, V. P. Mitrofanov, G. Mitselmakher, R. Mittleman, D. Moffa, A. Moggi, K. Mogushi, M. Mohan, S. R. P. Mohapatra, I. Molina, M. Montani, C. J. Moore, D. Moraru, G. Moreno, S. Morisaki, S. R. Morriss, B. Mours, C. M. Mow-Lowry, G. Mueller, A. W. Muir, Arunava Mukherjee, D. Mukher- jee, S. Mukherjee, N. Mukund, A. Mullavey, J. Munch, E. A. Muijiz, M. Muratore, P. G. Murray, A. Nagar, K. Napier, I. Nardecchia, L. Naticchioni, R. K. Nayak, J. Neilson, G. Nelemans, T. J. N. Nel- son, M. Nery, A. Neunzert, L. Nevin, J. M. Newport, G. Newton, K. K. Y. Ng, P. Nguyen, T. T. Nguyen, D. Nichols, A. B. Nielsen, S. Nissanke, A. Nitz, A. Noack, F. Nocera, D. Nolting, C. North, L. K. Nuttall, J. Oberling, G. D. OŠDea, G. H. Ogin, J. J. Oh, S. H. Oh, F. Ohme, M. A. Okada, M. Oliver, P. Oppermann, Richard J. Oram, B. OŠReilly, R. Ormiston, L. F. Ortega, R. OŠShaughnessy, S. Os- sokine, D. J. Ottaway, H. Overmier, B. J. Owen, A. E. Pace, J. Page, M. A. Page, A. Pai, S. A. Pai, J. R. Palamos, O. Palashov, C. Palomba, A. Pal-Singh, Howard Pan, Huang-Wei Pan, B. Pang, P. T. H. Pang, C. Pankow, F. Pannarale, B. C. Pant, F. Paoletti, A. Paoli, M. A. Papa, A. Parida, W. Parker, D. Pascucci, A. Pasqualetti, R. Pas- saquieti, D. Passuello, M. Patil, B. Patricelli, B. L. Pearlstone, M. Pe- draza, R. Pedurand, L. Pekowsky, A. Pele, S. Penn, C. J. Perez, A. Perreca, L. M. Perri, H. P. Pfeiffer, M. Phelps, O. J. Piccinni, M. Pichot, F. Piergiovanni, V. Pierro, G. Pillant, L. Pinard, I. M. Pinto, M. Pirello, M. Pitkin, M. Poe, R. Poggiani, P. Popolizio, E. K. Porter, A. Post, J. Powell, J. Prasad, J. W. W. Pratt, G. Prat- ten, V. Predoi, T. Prestegard, M. Prijatelj, M. Principe, S. Privit- era, R. Prix, G. A. Prodi, L. G. Prokhorov, O. Puncken, M. Pun- 178 turo, P. Puppo, M. Pürrer, H. Qi, V. Quetschke, E. A. Quintero, R. Quitzow-James, F. J. Raab, D. S. Rabeling, H. Radkins, P. Raffai, S. Raja, C. Rajan, B. Rajbhandari, M. Rakhmanov, K. E. Ramirez, A. Ramos-Buades, P. Rapagnani, V. Raymond, M. Razzano, J. Read, T. Regimbau, L. Rei, S. Reid, D. H. Reitze, W. Ren, S. D. Reyes, F. Ricci, P. M. Ricker, S. Rieger, K. Riles, M. Rizzo, N. A. Robert- son, R. Robie, F. Robinet, A. Rocchi, L. Rolland, J. G. Rollins, V. J. Roma, J. D. Romano, R. Romano, C. L. Romel, J. H. Romie, D. Rosińska, M. P. Ross, S. Rowan, A. Rüdiger, P. Ruggi, G. Rutins, K. Ryan, S. Sachdev, T. Sadecki, L. Sadeghian, M. Sakellariadou, L. Salconi, M. Saleem, F. Salemi, A. Samajdar, L. Sammut, L. M. Sampson, E. J. Sanchez, L. E. Sanchez, N. Sanchis-Gual, V. Sand- berg, J. R. Sanders, B. Sassolas, B. S. Sathyaprakash, P. R. Saul- son, O. Sauter, R. L. Savage, A. Sawadsky, P. Schale, M. Scheel, J. Scheuer, J. Schmidt, P. Schmidt, R. Schnabel, R. M. S. SchoĄeld, A. Schönbeck, E. Schreiber, D. Schuette, B. W. Schulte, B. F. Schutz, S. G. Schwalbe, J. Scott, S. M. Scott, E. Seidel, D. Sellers, A. S. Sengupta, D. Sentenac, V. Sequino, A. Sergeev, D. A. Shaddock, T. J. Shaffer, A. A. Shah, M. S. Shahriar, M. B. Shaner, L. Shao, B. Shapiro, P. Shawhan, A. Sheperd, D. H. Shoemaker, D. M. Shoe- maker, K. Siellez, X. Siemens, M. Sieniawska, D. Sigg, A. D. Silva, L. P. Singer, A. Singh, A. Singhal, A. M. Sintes, B. J. J. Slagmolen, B. Smith, J. R. Smith, R. J. E. Smith, S. Somala, E. J. Son, J. A. Sonnenberg, B. Sorazu, F. Sorrentino, T. Souradeep, A. P. Spencer, A. K. Srivastava, K. Staats, A. Staley, M. Steinke, J. Steinlechner, S. Steinlechner, D. Steinmeyer, S. P. Stevenson, R. Stone, D. J. Stops, K. A. Strain, G. Stratta, S. E. Strigin, A. Strunk, R. Stu- rani, A. L. Stuver, T. Z. Summerscales, L. Sun, S. Sunil, J. Suresh, P. J. Sutton, B. L. Swinkels, M. J. Szczepańczyk, M. Tacca, S. C. Tait, C. Talbot, D. Talukder, D. B. Tanner, M. Tápai, A. Tarac- chini, J. D. Tasson, J. A. Taylor, R. Taylor, S. V. Tewari, T. Theeg, F. Thies, E. G. Thomas, M. Thomas, P. Thomas, K. A. Thorne, K. S. Thorne, E. Thrane, S. Tiwari, V. Tiwari, K. V. Tokmakov, K. Toland, M. Tonelli, Z. Tornasi, A. Torres-Forné, C. I. Torrie, D. Töyrä, F. Travasso, G. Traylor, J. Trinastic, M. C. Tringali, L. Trozzo, K. W. Tsang, M. Tse, R. Tso, L. Tsukada, D. Tsuna, D. Tuyenbayev, K. Ueno, D. Ugolini, C. S. Unnikrishnan, A. L. Urban, S. A. Usman, H. Vahlbruch, G. Vajente, G. Valdes, M. Vallisneri, N. van Bakel, M. van Beuzekom, J. F. J. van den Brand, C. Van Den Broeck, D. C. Vander-Hyde, L. van der Schaaf, J. V. van Heijningen, A. A. van Veg- gel, M. Vardaro, V. Varma, S. Vass, M. Vasúth, A. Vecchio, G. Ve- dovato, J. Veitch, P. J. Veitch, K. Venkateswara, G. Venugopalan, 179 D. Verkindt, F. Vetrano, A. Viceré, A. D. Viets, S. Vinciguerra, D. J. Vine, J.-Y. Vinet, S. Vitale, T. Vo, H. Vocca, C. Vorvick, S. P. Vy- atchanin, A. R. Wade, L. E. Wade, M. Wade, R. Walet, M. Walker, L. Wallace, S. Walsh, G. Wang, H. Wang, J. Z. Wang, W. H. Wang, Y. F. Wang, R. L. Ward, J. Warner, M. Was, J. Watchi, B. Weaver, L.- W. Wei, M. Weinert, A. J. Weinstein, R. Weiss, L. Wen, E. K. Wessel, P. Weßels, J. Westerweck, T. Westphal, K. Wette, J. T. Whelan, S. E. Whitcomb, B. F. Whiting, C. Whittle, D. Wilken, D. Williams, R. D. Williams, A. R. Williamson, J. L. Willis, B. Willke, M. H. Wimmer, W. Winkler, C. C. Wipf, H. Wittel, G. Woan, J. Woehler, J. Wofford, K. W. K. Wong, J. Worden, J. L. Wright, D. S. Wu, D. M. Wysocki, S. Xiao, H. Yamamoto, C. C. Yancey, L. Yang, M. J. Yap, M. Yazback, Hang Yu, Haocun Yu, M. Yvert, A. Zadro£ny, M. Zanolin, T. Zelen- ova, J.-P. Zendri, M. Zevin, L. Zhang, M. Zhang, T. Zhang, Y.-H. Zhang, C. Zhao, M. Zhou, Z. Zhou, S. J. Zhu, X. J. Zhu, A. B. Zim- merman, M. E. Zucker, and J. Zweizig. Gw170817: Observation of gravitational waves from a binary neutron star inspiral. Phys. Rev. Lett., 119:161101, Oct 2017. [ACL+21] M. Assié, E. Clément, A. Lemasson, D. Ramos, A. Raggio, I. Zanon, F. Galtarossa, C. Lenain, J. Casal, F. Flavigny, A. Matta, D. Men- goni, D. Beaumel, Y. Blumenfeld, R. Borcea, D. Brugnara, W. Cat- ford, F. de Oliveira, F. Delaunay, N. De Séréville, F. Didierjean, C.Aa. Diget, J. Dudouet, B. Fernández-Domínguez, C. Fougères, G. Frémont, V. Girard-Alcindor, A. Giret, A. Goasduff, A. Got- tardo, J. Goupil, F. Hammache, P.R. John, A. Korichi, L. Lalanne, S. Leblond, A. Lefevre, F. Legruel, L. Ménager, B. Million, C. Nicolle, F. Noury, E. Rauly, K. Rezynkina, E. Rindel, J.S. Rojo, M. Siciliano, M. Stanoiu, I. Stefan, and L. Vatrinet. The mugast-agata-vamos cam- paign: Set-up and performances. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 1014:165743, 2021. [AJ78] A. Alastuey and B. Jancovici. Nuclear reaction rate enhancement in dense stellar matter. Ap. J., 226:1034Ű1040, December 1978. [AJL+22] Evan H. Anders, Adam S. Jermyn, Daniel Lecoanet, Adrian E. Fraser, Imogen G. Cresswell, Meridith Joyce, and J. R. Fuentes. Schwarzschild and ledoux are equivalent on evolutionary timescales. The Astrophysical Journal Letters, 928(1):L10, mar 2022. [AL22] Marialuisa Aliotta and Karlheinz Langanke. Screening effects in stars and in the laboratory. Frontiers in Physics, 10, 2022. 180 [APR98] A. Akmal, V. R. Pandharipande, and D. G. Ravenhall. Equation of state of nucleon matter and neutron star structure. Phys. Rev. C, 58(3):1804Ű1828, September 1998. [ARW63] J. Barclay Adams, Malvin A. Ruderman, and Ching Hung Woo. Neu- trino pair emission by a stellar plasma. Phys. Rev., 129:1383Ű1390, Feb 1963. [Bah02] John N. Bahcall. The luminosity constraint on solar neutrino Ćuxes. Phys. Rev. C, 65:025801, Jan 2002. [BB09] S.J. Blundell and K. M. Blundell. Concepts in Thermal Physics. Ox- ford University Press, 2009. [BBB+05] N. R. Badnell, M. A. Bautista, K. Butler, F. Delahaye, C. Mendoza, P. Palmeri, C. J. Zeippen, and M. J. Seaton. Updated opacities from the Opacity Project. Monthly Not. Royal Astron. Soc., 360(2):458Ű 464, June 2005. [BC09] Edward F. Brown and Andrew Cumming. Mapping Crustal Heating with the Cooling Light Curves of Quasi-Persistent Transients. Ap. J., 698(2):1020Ű1032, June 2009. [BD23] Arash Bahramian and Nathalie Degenaar. Low-Mass X-ray Binaries. In Handbook of X-ray and Gamma-ray Astrophysics, page 120. 2023. [Bec16] F. Becattini. Thermodynamic Equilibrium in Relativity: Four- temperature, Killing Vectors and Lie Derivatives. Acta Physica Polonica B, 47(7):1819, January 2016. [BFG83a] T. Boddington, Chang-gen Feng, and Peter Gray. Thermal explosion and the theory of its initiation by steady intense light. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences, 390(1799):265Ű281, 1983. [BFG83b] Terry Boddington, Chang-Gen Feng, and Peter Gray. Thermal explo- sions, critically and transition in systems with variable thermal con- ductivity. distributed temperatures. J. Chem. Soc., Faraday Trans. 2, 79:1499Ű1513, 1983. [BHM+07] T. Belloni, J. Homan, S. Motta, E. Ratti, and M. Méndez. Rossi XTE monitoring of 4U1636-53 - I. Long-term evolution and kHz quasi- periodic oscillations. Monthly Not. Royal Astron. Soc., 379(1):247Ű 252, July 2007. 181 [BK98] John N. Bahcall and Plamen I. Krastev. Do hep neutrinos affect the solar neutrino energy spectrum? Physics Letters B, 436(3):243Ű250, 1998. [BM08] A. Sacha Brun and M. S. Miesch. Stellar convection simulations. Scholarpedia, 3(11):4278, 2008. revision #148549. [BP13] C.A. Bertulani and J. Piekarewicz. Neutron Star Crust. Space Sci- ence, Exploration and Policies Series. Nova Science Publishers, Incor- porated, 2013. [BPS71] Gordon Baym, Christopher Pethick, and Peter Sutherland. The Ground State of Matter at High Densities: Equation of State and Stellar Models. Ap. J., 170:299, December 1971. [Bro93] J D Brown. Action functionals for relativistic perfect Ćuids. 10(8):1579Ű1606, aug 1993. [BRSGS+06] A D Becerril Reyes, S Sen-Gupta, H Schatz, K L Kratz, and P Møller. Electron capture rates for neutron star crusts. PoS, NIC-IX:075, 2006. [BST66] S. G. Brush, H. L. Sahlin, and E. Teller. Monte Carlo Study of a One- Component Plasma. I. J. Chem. Phys., 45(6):2102Ű2118, September 1966. [CAF+10] Richard H. Cyburt, A. Matthew Amthor, Ryan Ferguson, Zach Meisel, Karl Smith, Scott Warren, Alexander Heger, R. D. Hoffman, Thomas Rauscher, Alexander Sakharuk, Hendrik Schatz, F. K. Thielemann, and Michael Wiescher. THE JINA REACLIB DATABASE: ITS RE- CENT UPDATES AND IMPACT ON TYPE-i x-RAY BURSTS. The Astrophysical Journal Supplement Series, 189(1):240Ű252, jun 2010. [Cam59] A. G. W. Cameron. Pycnonuclear Reations and Nova Explosions. Ap. J., 130:916, November 1959. [Car76] T. R. Carson. Stellar opacity. Ann. Rev. of Astron. and Astrophys., 14:95Ű117, January 1976. [Car91] Brandon Carter. Convective variational approach to relativistic ther- modynamics of dissipative Ćuids. Proceedings of the Royal Society of London. Series A: Mathematical and Physical Sciences, 433(1887):45Ű 62, 1991. [CCAW04] S. Carroll, S.M. Carroll, and Addison-Wesley. Spacetime and Geometry: An Introduction to General Relativity. Addison Wesley, 2004. 182 [CCC70] V. Canuto, C. Chiuderi, and C. K. Chou. Plasmon Neutri- nos Emission in a Strong Magnetic Field. I: Transverse Plasmons. Astrophys. and Space Science, 7(3):407Ű415, June 1970. [CCPD+07] David Cubero, Jesús Casado-Pascual, Jörn Dunkel, Peter Talkner, and Peter Hänggi. Thermal equilibrium and statistical thermometers in special relativity. Phys. Rev. Lett., 99:170601, Oct 2007. [CDY07] A. I. Chugunov, H. E. DeWitt, and D. G. Yakovlev. Coulomb tun- neling for fusion reactions in dense matter: Path integral monte carlo versus mean Ąeld. Phys. Rev. D, 76:025028, Jul 2007. [CF88] Georgeann R. Caughlan and William A. Fowler. Thermonuclear Reac- tion Rates V. Atomic Data and Nuclear Data Tables, 40:283, January 1988. [CFZH20] N. Chamel, A. F. Fantina, J. L. Zdunik, and P. Haensel. Experimental constraints on shallow heating in accreting neutron-star crusts. Phys. Rev. C, 102:015804, Jul 2020. [CGG+20] Y. Cavecchi, D. K. Galloway, A. J. Goodwin, Z. Johnston, and A. Heger. The efficiency of nuclear burning during ther- monuclear (Type I) bursts as a function of accretion rate. Monthly Not. Royal Astron. Soc., 499(2):2148Ű2156, September 2020. [CH52] C. F. Curtiss and J. O. Hirschfelder. Integration of stiff equations*. Proceedings of the National Academy of Sciences, 38(3):235Ű243, 1952. [Cha22a] Sylvain Chaty. High-mass x-ray binaries (hmxb). In Accreting Binaries, 2514-3433, pages 6Ű1 to 6Ű58. IOP Publishing, 2022. [Cha22b] Sylvain Chaty. Introduction. In Accreting Binaries, 2514-3433, pages 1Ű1 to 1Ű5. IOP Publishing, 2022. [Chi66] H Chiu. Neutrinos in astrophysics and cosmology. Annual Review of Nuclear and Particle Science, 16(Volume 16, 1966):591Ű618, 1966. [CiV+03] R. Cornelisse, J. J. M. inŠt Zand, F. Verbunt, E. Kuulkers, J. Heise, P. R. den Hartog, M. Cocchi, L. Natalucci, A. Bazzano, and P. Uber- tini. Six years of BeppoSAX Wide Field Cameras observations of nine galactic type I X-ray bursters. Astron. Astrophys., 405:1033Ű1042, July 2003. [CKM+16] J. Colgan, D. P. Kilcrease, N. H. Magee, M. E. Sherrill, J. Abdallah Jr., P. Hakel, C. J. Fontes, J. A. Guzik, and K. A. Mussack. A New 183 Generation of Los Alamos Opacity Tables. The Astrophysical Journal, 817(2):116, jan 2016. [Cla68] Donald D. Clayton. Principles of stellar evolution and nucleosynthesis. 1968. [CN05] Randall L. Cooper and Ramesh Narayan. Theoretical models of su- perbursts on accreting neutron stars. The Astrophysical Journal, 629(1):422, aug 2005. [CN06] Randall L. Cooper and Ramesh Narayan. On the Physics of Type I X-Ray Bursts on Accreting Neutron Stars at High Accretion Rates. Astrophys. J. Lett., 648(2):L123ŰL126, September 2006. [CP98] Gilles Chabrier and Alexander Y. Potekhin. Equation of state of fully ionized electron-ion plasmas. Phys. Rev. E, 58(4):4941Ű4949, October 1998. [CPP+07] S. Cassisi, A. Y. Potekhin, A. Pietrinferni, M. Catelan, and M. Salaris. Updated Electron-Conduction Opacities: The Impact on Low-Mass Stellar Models. Ap. J., 661(2):1094Ű1104, June 2007. [CWG17] Yuri Cavecchi, Anna L. Watts, and Duncan K. Galloway. On the dependence of the x-ray burst rate on accretion and spin rate. The Astrophysical Journal, 851(1):1, dec 2017. [DBW11] N. Degenaar, E. F. Brown, and R. Wijnands. Evidence for crust cooling in the transiently accreting 11-Hz X-ray pulsar in the glob- ular cluster Terzan 5. Monthly Not. Royal Astron. Soc., 418(1):L152Ű L156, November 2011. [DCBP15] Alex Deibel, Andrew Cumming, Edward F. Brown, and Dany Page. A strong shallow heat source in the accreting neutron star maxi j0556- 332. The Astrophysical Journal Letters, 809(2):L31, aug 2015. [dCC+21] R. J. deBoer, O. Clarkson, A. J. Couture, J. Görres, F. Herwig, I. Lom- bardo, P. Scholz, and M. Wiescher. 19F(p, γ)20Ne and 19F(p, α)16O reaction rates and their effect on calcium production in population iii stars from hot cno breakout. Phys. Rev. C, 103:055815, May 2021. [DCJM11] Barry Davids, Richard H. Cyburt, Jordi José, and Subramanian Mythili. The inĆuence of uncertainties in the 15o(α, γ)19ne reac- tion rate on models of type i x-ray bursts. The Astrophysical Journal, 735(1):40, jun 2011. 184 [DD10] J Dooling and Accelerator Systems Division. Dose calculations using mars for bremsstrahlung beam stops and collimators in aps beamline stations. 11 2010. [dDGFJ14] R. de Diego, E. Garrido, D. V. Fedorov, and A. S. Jensen. Production of 6he and 9be by radiative capture and four-body recombination. The European Physical Journal A, 50(6):93, 2014. [Deu83] Bader G. DeuĆhard, P. A semi-implicit mid-point rule for stiff systems of ordinary differential equations. Numerische Mathematik, 41:373Ű 398, 1983. [DG09] J. Daligault and S. Gupta. ELECTRONŰION SCATTERING IN DENSE MULTI-COMPONENT PLASMAS: APPLICATION TO THE OUTER CRUST OF AN ACCRETING NEUTRON STAR. The Astrophysical Journal, 703(1):994Ű1011, sep 2009. [dGFJ10] R. de Diego, E. Garrido, D. V. Fedorov, and A. S. Jensen. Few-Body Reactions in Nuclear Astrophysics: application to 6He and 9Be pro- duction. In European Physical Journal Web of Conferences, volume 3 of European Physical Journal Web of Conferences, page 04017, April 2010. [DHH09] Jörn Dunkel, Peter Hänggi, and Stefan Hilbert. Non-local observ- ables and lightcone-averaging in relativistic thermodynamics. Nature Physics, 5(10):741Ű747, October 2009. [Dic72] Duane A. Dicus. Stellar energy-loss rates in a convergent theory of weak and electromagnetic interactions. Phys. Rev. D, 6:941Ű949, Aug 1972. [DLC+98] D. J. Dean, K. Langanke, L. Chatterjee, P. B. Radha, and M. R. Strayer. Electron capture on iron group nuclei. Phys. Rev. C, 58:536Ű 544, Jul 1998. [Dor21] T.C. Dorlas. Statistical Mechanics: Fundamentals and Model Solutions. CRC Press, 2021. [DRP04] Sharada Iyer Dutta, Saša Ratković, and Madappa Prakash. Photoneu- trino process in astrophysical systems. Phys. Rev. D, 69(2):023005, January 2004. [DWB+13] N. Degenaar, R. Wijnands, E. F. Brown, D. Altamirano, E. M. Cack- ett, J. Fridriksson, J. Homan, C. O. Heinke, J. M. Miller, D. Pooley, and G. R. Sivakoff. Continued Neutron Star Crust Cooling of the 11 185 Hz X-Ray Pulsar in Terzan 5: A Challenge to Heating and Cooling Models? Ap. J., 775(1):48, September 2013. [Eck40] Carl Eckart. The Thermodynamics of Irreversible Processes. III. Rela- tivistic Theory of the Simple Fluid. Physical Review, 58(10):919Ű924, November 1940. [EFF73] P. P. Eggleton, J. Faulkner, and B. P. Flannery. An Approximate Equation of State for Stellar Material. Astron. Astrophys., 23:325, March 1973. [FCZ75] William A. Fowler, Georgeanne R. Caughlan, and Bar- bara A. Zimmerman. Thermonuclear Reaction Rates, II. Ann. Rev. of Astron. and Astrophys., 13:69, January 1975. [FDB+05] Hans Fynbo, Christian Aaen Diget, U. Bergmann, M.J.G. Borge, J. Cederkäll, P. Dendooven, L.M. Fraile, S. Franchoo, V.N. Fedosseev, B.R. Fulton, W. Huang, J. Huikari, H.B. Jeppesen, A.S. Jokinen, P. Jones, B. Jonson, U. Köster, K. Langanke, M. Meister, T. Nilsson, G. Nyman, Y. Prezado, K. Riisager, S. Rinta-Antila, O. Tengblad, M. Turrion, Y. Wang, L. Weissman, K. Wilhelmsen, and J. Äustö. Revised rates for the stellar triple-alpha process from measurement of 12c nuclear resonances. Nature, 433:136Ű139, 2005. [FFN82] G. M. Fuller, W. A. Fowler, and M. J. Newman. Stellar weak interac- tion rates for intermediate-mass nuclei. II - A = 21 to A = 60. Ap. J., 252:715Ű740, January 1982. [FGWD06a] Jacob Lund Fisker, Joachim Görres, Michael Wiescher, and Barry Davids. The importance of 15o(α,γ)19ne to x-ray bursts and super- bursts. The Astrophysical Journal, 650(1):332, oct 2006. [FGWD06b] Jacob Lund Fisker, Joachim Gorres, Michael Wiescher, and Barry Davids. The Importance of 15-O(alpha, gamma)19-Ne to x-ray bursts and superbursts. Astrophys. J., 650:332Ű337, 2006. [FHM81] M. Y. Fujimoto, T. Hanawa, and S. Miyaji. Shell Ćashes on accreting neutron stars and X-ray bursts. Ap. J., 247:267Ű278, July 1981. [FK38] David A Frank-Kamenetskii. Towards temperature distributions in a reaction vessel and the stationary theory of thermal explosion. In Doklady Acad. Nauk SSSR, volume 18, pages 411Ű412, 1938. [FL87] Ikko Fushiki and D. Q. Lamb. S-Matrix Calculation of the Triple- Alpha Reaction. Ap. J., 317:368, June 1987. 186 [FOR+00] B. Fryxell, K. Olson, P. Ricker, F. X. Timmes, M. Zingale, D. Q. Lamb, P. MacNeice, R. Rosner, J. W. Truran, and H. Tufo. FLASH: An Adaptive Mesh Hydrodynamics Code for Modeling Astrophysical Thermonuclear Flashes. Ap. J. S., 131(1):273Ű334, November 2000. [FPM17] Cristian Farías, Victor A. Pinto, and Pablo S. Moya. What is the temperature of a moving body? ScientiĄc Reports, 7:17657, December 2017. [FST08] Jacob Lund Fisker, Hendrik Schatz, and Friedrich-Karl Thielemann. Explosive Hydrogen Burning during Type I X-Ray Bursts. Ap. J. S., 174(1):261Ű276, January 2008. [FTG+07] Jacob Lund Fisker, Wanpeng Tan, Joachim Görres, Michael Wiescher, and Randall L. Cooper. The 15o(α,γ)19ne reaction rate and the stability of thermonuclear burning on accreting neutron stars. The Astrophysical Journal, 665(1):637, aug 2007. [GBS+07] Sanjib Gupta, Edward F. Brown, Hendrik Schatz, Peter Möller, and Karl-Ludwig Kratz. Heating in the Accreted Neutron Star Ocean: Implications for Superburst Ignition. Ap. J., 662(2):1188Ű1197, June 2007. [GCP13] S. Goriely, N. Chamel, and J. M. Pearson. Further explorations of skyrme-hartree-fock-bogoliubov mass formulas. xiii. the 2012 atomic mass evaluation and the symmetry coefficient. Phys. Rev. C, 88:024308, Aug 2013. [GGIH09] J. R. Goodwin, V. V. Golovko, V. E. Iacob, and J. C. Hardy. Half-life of the electron-capture decay of 97Ru: Precision measurement shows no temperature dependence. Phys. Rev. C, 80:045501, Oct 2009. [GGPR62] Riccardo Giacconi, Herbert Gursky, Frank R. Paolini, and Bruno B. Rossi. Evidence for x Rays From Sources Outside the Solar System. Phys. Rev. Lett., 9(11):439Ű443, December 1962. [GJ21] E. Garrido and A. S. Jensen. Direct and sequential four-body recom- bination rates at low temperatures. Phys. Rev. C, 103:055813, May 2021. [GK21] Duncan K. Galloway and Laurens Keek. Thermonuclear X-ray Bursts. In Tomaso M. Belloni, Mariano Méndez, and Chengmin Zhang, editors, Astrophysics and Space Science Library, volume 461 of Astrophysics and Space Science Library, pages 209Ű262, January 2021. 187 [GM71] N.B. Gove and M.J. Martin. Log-f tables for beta decay. Atomic Data and Nuclear Data Tables, 10(3):205Ű219, 1971. [GMH+08] Duncan K. Galloway, Michael P. Muno, Jacob M. Hartman, Dim- itrios Psaltis, and Deepto Chakrabarty. Thermonuclear (Type I) X- Ray Bursts Observed by the Rossi X-Ray Timing Explorer. Ap. J. S., 179(2):360Ű422, December 2008. [GNI+10] J. R. Goodwin, N. Nica, V. E. Iacob, A. Dibidad, and J. C. Hardy. Measurement of the half-life of 198Au in a nonmetal: High-precision measurement shows no host-material dependence. Phys. Rev. C, 82:044320, Oct 2010. [GNMM14] M. Gabriel, A. Noels, J. Montalbán, and A. Miglio. Proper use of Schwarzschild Ledoux criteria in stellar evolution computations. Astron. Astrophys., 569:A63, September 2014. [GPE83] E. H. Gudmundsson, C. J. Pethick, and R. I. Epstein. Structure of neutron star envelopes. Ap. J., 272:286Ű300, September 1983. [GPLW+19] B. E. Glassman, D. Pérez-Loureiro, C. Wrede, J. Allen, D. W. Bar- dayan, M. B. Bennett, K. A. Chipps, M. Febbraro, M. Friedman, C. Fry, M. R. Hall, O. Hall, S. N. Liddick, P. OŠMalley, W.-J. Ong, S. D. Pain, S. B. Schwartz, P. Shidling, H. Sims, L. J. Sun, P. Thomp- son, and H. Zhang. Doppler broadening in 20Mg(βpγ)19Ne decay. Phys. Rev. C, 99:065801, Jun 2019. [Gra08] David F. Gray. The Observation and Analysis of Stellar Photospheres. 2008. [GS06] L. S. García-Colín and Alfredo Sandoval-Villalbazo. Relativistic Non- Equilibrium Thermodynamics Revisited. Journal of Non Equilibrium Thermodynamics, 31(1):11Ű22, January 2006. [Gug45] E. A. Guggenheim. The principle of corresponding states. The Journal of Chemical Physics, 13(7):253Ű261, 1945. [GW04] M.L. Goldberger and K.M. Watson. Collision Theory. Dover books on physics. Dover Publications, 2004. [GZ05] L. V. Grigorenko and M. V. Zhukov. Three-body resonant radiative capture reactions in astrophysics. Phys. Rev. C, 72:015803, Jul 2005. [Ham20] J.M. Hameury. A review of the disc instability model for dwarf novae, soft x-ray transients and related objects. Advances in Space Research, 188 66(5):1004Ű1024, 2020. Nova Eruptions, Cataclysmic Variables and Related Systems: Challenges in the 2020 Era. [Hau75] E. Haug. Bremsstrahlung and pair production in the Ąeld of free electrons. Zeitschrift Naturforschung Teil A, 30:1099Ű1113, September 1975. [HBB+20] M. R. Hall, D. W. Bardayan, T. Baugher, A. Lepailleur, S. D. Pain, A. Ratkiewicz, S. Ahn, J. M. Allen, J. T. Anderson, A. D. Ayangeakaa, J. C. Blackmon, S. Burcher, M. P. Carpenter, S. M. Cha, K. Y. Chae, K. A. Chipps, J. A. Cizewski, M. Febbraro, O. Hall, J. Hu, C. L. Jiang, K. L. Jones, E. J. Lee, P. D. OŠMalley, S. Ota, B. C. Rasco, D. Santiago-Gonzalez, D. Seweryniak, H. Sims, K. Smith, W. P. Tan, P. Thompson, C. Thornsberry, R. L. Varner, D. Walter, G. L. Wilson, and S. Zhu. 19Ne level structure for explosive nucleosynthesis. Phys. Rev. C, 102:045802, Oct 2020. [HCW07] Alexander Heger, Andrew Cumming, and S. E. Woosley. Millihertz Quasi-periodic Oscillations from Marginally Stable Nuclear Burning on an Accreting Neutron Star. Ap. J., 665(2):1311Ű1320, August 2007. [HdS21] Faïrouz Hammache and Nicolas de Séréville. Transfer reactions as a tool in nuclear astrophysics. Frontiers in Physics, 8, 2021. [HFG64] L. G. Henyey, J. E. Forbes, and N. L. Gould. A New Method of Automatic Computation of Stellar Evolution. Ap. J., 139:306, January 1964. [HFW+14] Jeroen Homan, Joel K. Fridriksson, Rudy Wijnands, Edward M. Cack- ett, Nathalie Degenaar, Manuel Linares, Dacheng Lin, and Ronald A. Remillard. A Strongly Heated Neutron Star in the Transient Z Source MAXI J0556-332. Ap. J., 795(2):131, November 2014. [HGA+96] K. I. Hahn, A. García, E. G. Adelberger, P. V. Magnus, A. D. Bacher, N. Bateman, G. P. A. Berg, J. C. Blackmon, A. E. Champagne, B. Davis, A. J. Howard, J. Liu, B. Lund, Z. Q. Mao, D. M. Markoff, P. D. Parker, M. S. Smith, E. J. Stephenson, K. B. Swartz, S. Utku, R. B. Vogelaar, and K. Yildiz. Structure of 18Ne and the breakout from the hot cno cycle. Phys. Rev. C, 54:1999Ű2013, Oct 1996. [HH09] Wynn C. G. Ho and Craig O. Heinke. A neutron star with a car- bon atmosphere in the Cassiopeia A supernova remnant. Nature, 462(7269):71Ű73, November 2009. 189 [HHIK56] Satio Hayakawa, Chushiro Hayashi, Mitsuo Imoto, and Ken Kikuchi. Helium Capturing Reactions in Stars. Progress of Theoretical Physics, 16(5):507Ű527, 11 1956. [Hin08] Tanja Hinderer. Tidal Love Numbers of Neutron Stars. Ap. J., 677(2):1216Ű1220, April 2008. [HK94] Carl J. Hansen and Steven D. Kawaler. Stellar Interiors. Physical Principles, Structure, and Evolution. 1994. [HL69] W. B. Hubbard and Martin Lampe. Thermal Conduction by Electrons in Stellar Matter. Ap. J. S., 18:297, July 1969. [HM06] W. Raphael Hix and Bradley S. Meyer. Thermonuclear kinetics in astrophysics. Nucl. Phys. A, 777:188Ű207, October 2006. [HNW93] E. Hairer, S.P. Nørsett, and G. Wanner. Solving Ordinary Differential Equations II: Stiff and Differential-Algebraic Problems. Solving Or- dinary Differential Equations II: Stiff and Differential-algebraic Prob- lems. Springer, 1993. [Hoy54] F. Hoyle. On Nuclear Reactions Occuring in Very Hot STARS.I. the Synthesis of Elements from Carbon to Nickel. Ap. J. S., 1:121, September 1954. [HPY07] P. Haensel, A. Y. Potekhin, and D. G. Yakovlev. Neutron Stars 1 : Equation of State and Structure, volume 326. 2007. [HRW94] Martin Haft, Georg Raffelt, and Achim Weiss. Standard and Non- standard Plasma Neutrino Emission Revisited. Ap. J., 425:222, April 1994. [Hua87] Kerson Huang. Statistical Mechanics, 2nd Edition. 1987. [HvF+10] Jeroen Homan, Michiel van der Klis, Joel K. Fridriksson, Ronald A. Remillard, Rudy Wijnands, Mariano Méndez, Dacheng Lin, Diego Al- tamirano, Piergiorgio Casella, Tomaso M. Belloni, and Walter H. G. Lewin. XTE J1701-462 and Its Implications for the Nature of Sub- classes in Low-magnetic-Ąeld Neutron Star Low-mass X-ray Binaries. Ap. J., 719(1):201Ű212, August 2010. [HZ90a] P. Haensel and J. L. Zdunik. Equation of state and structure of the crust of an accreting neutron star. Astron. Astrophys., 229(1):117Ű 122, March 1990. 190 [HZ90b] P. Haensel and J. L. Zdunik. Non-equilibrium processes in the crust of an accreting neutron star. Astron. Astrophys., 227(2):431Ű436, Jan- uary 1990. [HZD89] P. Haensel, J. L. Zdunik, and J. Dobaczewski. Composition and equation of state of cold catalyzed matter below neutron drip. Astron. Astrophys., 222(1-2):353Ű357, September 1989. [IHNK96] Naoki Itoh, Hiroshi Hayashi, Akinori Nishikawa, and Yasuharu Ko- hyama. Neutrino Energy Loss in Stellar Interiors. VII. Pair, Photo- , Plasma, Bremsstrahlung, and Recombination Neutrino Processes. Ap. J. S., 102:411, February 1996. [IIT87] Setsuo Ichimaru, Hiroshi Iyetomi, and Shigenori Tanaka. Statistical physics of dense plasmas: Thermodynamics, transport coefficients and dynamic correlations. Phys. Rep., 149(2-3):91Ű205, May 1987. [ILC+10] C. Iliadis, R. Longland, A.E. Champagne, A. Coc, and R. Fitzgerald. Charged-particle thermonuclear reaction rates: Ii. tables and graphs of reaction rates and probability density functions. Nuclear Physics A, 841(1):31Ű250, 2010. The 2010 Evaluation of Monte Carlo based Thermonuclear Reaction Rates. [IR96] Carlos A. Iglesias and Forrest J. Rogers. Updated Opal Opacities. Ap. J., 464:943, June 1996. [IS79] W. Israel and J. M. Stewart. Transient relativistic thermodynamics and kinetic theory. Annals of Physics, 118(2):341Ű372, April 1979. [IS10] N. A. Inogamov and R. A. Sunyaev. Spread of matter over a neutron- star surface during disk accretion: Deceleration of rapid rotation. Astronomy Letters, 36(12):848Ű894, December 2010. [Isr76] Werner Israel. Nonstationary irreversible thermodynamics: A causal relativistic theory. Annals of Physics, 100(1):310Ű331, 1976. [Jan77] B. Jancovici. Pair correlation function in a dense plasma and pycnonu- clear reactions in stars. Journal of Statistical Physics, 17(5):357Ű370, 1977. [JEL96] S. M. Johns, P. J. Ellis, and J. M. Lattimer. Numerical Approximation to the Thermodynamic Integrals. Ap. J., 473:1020, December 1996. [JIN22] JINA Reaclib Database. JINA Reaclib Database, 2022. 191 [JLH+10] A. Juodagalvis, K. Langanke, W.R. Hix, G. Martínez-Pinedo, and J.M. Sampaio. Improved estimate of electron capture rates on nuclei during stellar core collapse. Nuclear Physics A, 848(3):454Ű478, 2010. [JSB+21] Adam S. Jermyn, Josiah Schwab, Evan Bauer, F. X. Timmes, and Alexander Y. Potekhin. Skye: A differentiable equation of state. The Astrophysical Journal, 913(1):72, may 2021. [JT23] Meridith Joyce and Jamie Tayar. A Review of the Mixing Length Theory of Convection in 1D Stellar Modeling. Galaxies, 11(3):75, June 2023. [Kam14] A. Kamal. Nuclear Physics. Graduate Texts in Physics. Springer Berlin Heidelberg, 2014. [KB97] Dallas C. Kennedy and Sidney A. Bludman. Variational Principles for Stellar Structure. Ap. J., 484(1):329Ű340, July 1997. [Kdi+03] E. Kuulkers, P. R. den Hartog, J. J. M. inŠt Zand, F. W. M. Ver- bunt, W. E. Harris, and M. Cocchi. Photospheric radius expansion X-ray bursts as standard candles. Astron. Astrophys., 399:663Ű680, February 2003. [KHv+02] E. Kuulkers, J. Homan, M. van der Klis, W. H. G. Lewin, and M. Mén- dez. X-ray bursts at extreme mass accretion rates from GX 17+2. Astron. Astrophys., 382:947Ű973, February 2002. [KiA+10] E. Kuulkers, J. J. M. inŠt Zand, J. L. Atteia, A. M. Levine, S. Brandt, D. A. Smith, M. Linares, M. Falanga, C. Sánchez-Fernández, C. B. Markwardt, T. E. Strohmayer, A. Cumming, and M. Suzuki. What ignites on the neutron star of 4U 0614+091? Astron. Astrophys., 514:A65, May 2010. [KKO+03] H. Karttunen, P. Kröger, H. Oja, M. Poutanen, and K.J. Donner. Fundamental Astronomy. Physics and Astronomy Online Library. Springer Berlin Heidelberg, 2003. [KLi09] Keek, L., Langer, N., and in Št Zand, J. J. M. The effect of rotation on the stability of nuclear burning in accreting neutron stars. A&A, 502(3):871Ű881, 2009. [Kor79] Wşodzimierz Kordylewski. Critical parameters of thermal explosion. Combustion and Flame, 34:109Ű117, 1979. 192 [KPP+99] A. D. Kaminker, C. J. Pethick, A. Y. Potekhin, V. Thorsson, and D. G. Yakovlev. Neutrino-pair bremsstrahlung by electrons in neutron star crusts. Astron. Astrophys., 343:1009Ű1024, March 1999. [Kuu04] Erik Kuulkers. The observersŠ view of (very) long x-ray bursts: they are super! Nuclear Physics B - Proceedings Supplements, 132:466Ű475, 2004. Proceedings of the 2nd BeppoSAX Conference: The Restless High-Energy Universe. [Kuz14] A. L. Kuzemsky. Thermodynamic Limit in Statistical Physics. International Journal of Modern Physics B, 28(09):1430004, 2014. [Kv95] E. Kuulkers and M. van der Klis. Detection of 26Hz quasi-periodic oscillations in the Ćaring branch of CygnusX-2. Astron. Astrophys., 303:801, November 1995. [KWW12] Rudolf Kippenhahn, Alfred Weigert, and Achim Weiss. Stellar Structure and Evolution. 2012. [LA11] C. S. Lopez-Monsalvo and N. Andersson. Thermal dynamics in gen- eral relativity. Proceedings of the Royal Society of London Series A, 467(2127):738Ű759, March 2011. [LA18] S K Lander and N Andersson. Heat conduction in rotating rela- tivistic stars. Monthly Notices of the Royal Astronomical Society, 479(3):4207Ű4215, 06 2018. [LAC+12] M. Linares, D. Altamirano, D. Chakrabarty, A. Cumming, and L. Keek. Millihertz Quasi-periodic Oscillations and Thermonuclear Bursts from Terzan 5: A Showcase of Burning Regimes. Ap. J., 748(2):82, April 2012. [Lat12] James M. Lattimer. The Nuclear Equation of State and Neutron Star Masses. Annual Review of Nuclear and Particle Science, 62(1):485Ű 515, November 2012. [Lib10] M.A. Liberman. Introduction to Physics and Chemistry of Combustion: Explosion, Flame, Detonation. Springer Berlin Heidel- berg, 2010. [LL69] J. L. Lebowitz and Elliott H. Lieb. Existence of thermodynamics for real matter with coulomb forces. Phys. Rev. Lett., 22:631Ű634, Mar 1969. [LL76] Landau, L. D. and Lifshitz, E. M. Statistical Physics, Part 1. Harcourt College Publishers, second edition, 1976. 193 [LMJ14] R. Longland, D. Martin, and J. José. Performance improvements for nuclear reaction network integration. Astron. Astrophys., 563:A67, March 2014. [LRP93] C. P. Lorenz, D. G. Ravenhall, and C. J. Pethick. Neutron star crusts. Phys. Rev. Lett., 70(4):379Ű382, January 1993. [LRW20] Elena Litvinova, Caroline Robin, and Herlik Wibowo. Temperature dependence of nuclear spin-isospin response and beta decay in hot astrophysical environments. Physics Letters B, 800:135134, 2020. [LWA+10] M. Linares, A. Watts, D. Altamirano, P. Soleri, N. Degenaar, Y. Yang, R. Wijnands, P. Casella, J. Homan, D. Chakrabarty, N. Rea, M. Armas-Padilla, Y. Cavecchi, M. Kalamkar, R. Kaur, A. Patruno, and M. van der Klis. The Return of the Bursts: Thermonuclear Flashes from Circinus X-1. Astrophys. J. Lett., 719(1):L84ŰL89, August 2010. [MAC+95] N. H. Magee, Jr. Abdallah, J., R. E. H. Clark, J. S. Cohen, L. A. Collins, G. Csanak, C. J. Fontes, A. Gauger, J. J. Keady, D. P. Kil- crease, and A. L. Merts. Atomic Structure Calculations and New LOS Alamos Astrophysical Opacities. In Saul J. Adelman and W. L. Wiese, editors, Astrophysical Applications of Powerful New Databases, vol- ume 78 of Astronomical Society of the PaciĄc Conference Series, page 51, January 1995. [MC11] Zach Medin and Andrew Cumming. Compositionally driven convec- tion in the oceans of Accreting Neutron Stars. The Astrophysical Journal, 730(2):97, mar 2011. [Mei18] Zach Meisel. Consistent modeling of gs 1826-24 x-ray bursts for mul- tiple accretion rates demonstrates the possibility of constraining rp- process reaction rates. The Astrophysical Journal, 860(2):147, jun 2018. [MMM19] Zach Meisel, Grant Merz, and Sophia Medvid. InĆuence of nuclear reaction rate uncertainties on neutron star properties extracted from x- ray burst modelŰobservation comparisons. The Astrophysical Journal, 872(1):84, feb 2019. [MS69] Charles W. Misner and David H. Sharp. Energy Transport by Radia- tive Diffusion in Relativistic Spherically Symmetric Hydrodynamics. In Quasars and high-energy astronomy, page 397, January 1969. 194 [MSB+07] C. Mendoza, M. J. Seaton, P. Buerger, A. Bellorín, M. Meléndez, J. González, L. S. Rodríguez, F. Delahaye, E. Palacios, A. K. Prad- han, and C. J. Zeippen. OPserver: interactive online computations of opacities and radiative accelerations. Monthly Notices of the Royal Astronomical Society, 378(3):1031Ű1035, 06 2007. [MTW17] Charles W. Misner, Kip S. Thorne, and John Archibald Wheeler. Gravitation. W.H. Freeman and Company, 2017. [Nad74] D. K. Nadyozhin. Asymptotic formulae for equation of state of electron-positron gas. Nauchnye Informatsii, 32:3, January 1974. [NCPB23] Martin Nava-Callejas, Dany Page, and Mikhail V. Beznogov. Probing strong Ąeld f(r) gravity and ultradense matter with the structure and thermal evolution of neutron stars. Phys. Rev. D, 107:104057, May 2023. [NH03] Ramesh Narayan and Jeremy S. Heyl. Thermonuclear Stability of Material Accreting onto a Neutron Star. Ap. J., 599(1):419Ű449, De- cember 2003. [Nov16] Vasily Novozhilov. Thermal explosion in oscillating ambient condi- tions. ScientiĄc Reports, 6(1):29730, 2016. [NRB+01] E.B Norman, G.A Rech, E Browne, R.-M Larimer, M.R Dragowsky, Y.D Chan, M.C.P Isaac, R.J McDonald, and A.R Smith. InĆuence of physical and chemical environments on the decay rates of 7be and 40k. Physics Letters B, 519(1):15Ű22, 2001. [NT81] K. Nomoto and S. Tsuruta. Cooling of young neutron stars and Ein- stein X-ray observartions. Astrophys. J. Lett., 250:L19ŰL23, Novem- ber 1981. [NTM85] K. Nomoto, F. K. Thielemann, and S. Miyaji. The triple alpha reac- tion at low temperatures in accreting white dwarfs and neutron stars. Astron. Astrophys., 149(2):239Ű245, August 1985. [NuD22] NuDat3. NuDat3, 2022. [NV73] J. W. Negele and D. Vautherin. Neutron star matter at sub-nuclear densities. Nucl. Phys. A, 207(2):298Ű320, June 1973. [OHM+94] T. Oda, M. Hino, K. Muto, M. Takahara, and K. Sato. Rate Tables for the Weak Processes of sd-Shell Nuclei in Stellar Matter. Atomic Data and Nuclear Data Tables, 56(2):231Ű403, March 1994. 195 [Pac83] B. Paczynski. Models of X-ray bursters with radius expansion. Ap. J., 267:315Ű321, April 1983. [Pag16] Dany Page. NSCool: Neutron star cooling code, September 2016. [PB07] Anthony L. Piro and Lars Bildsten. Turbulent Mixing in the Surface Layers of Accreting Neutron Stars. Ap. J., 663(2):1252Ű1268, July 2007. [PBD+11] Bill Paxton, Lars Bildsten, Aaron Dotter, Falk Herwig, Pierre Lesaffre, and Frank Timmes. Modules for Experiments in Stellar Astrophysics (MESA). Ap. J. S., 192(1):3, January 2011. [PBG+20] Dany Page, Mikhail V. Beznogov, Iván Garibay, James M. Lattimer, Madappa Prakash, and Hans-Thomas Janka. Ns 1987A in SN 1987A. Ap. J., 898(2):125, July 2020. [PC13] A. Y. Potekhin and G. Chabrier. Equation of state for magnetized Coulomb plasmas. Astron. Astrophys., 550:A43, February 2013. [PCY97] A. Y. Potekhin, G. Chabrier, and D. G. Yakovlev. Internal temper- atures and cooling of neutron stars with accreted envelopes. Astron. Astrophys., 323:415, 1997. [PGC17] Jean-Christophe Pain, Franck Gilleron, and Maxime Comet. Detailed opacity calculations for astrophysical applications. Atoms, 5(2), 2017. [PHN+22] Dany Page, Jeroen Homan, Martin Nava-Callejas, Yuri Cavecchi, Mikhail V. Beznogov, Nathalie Degenaar, Rudy Wijnands, and Aastha S. Parikh. A ŞHyperburstŤ in the MAXI J0556-332 Neutron Star: Evidence for a New Type of Thermonuclear Explosion. Ap. J., 933(2):216, July 2022. [PMS+15] Bill Paxton, Pablo Marchant, Josiah Schwab, Evan B. Bauer, Lars Bildsten, Matteo Cantiello, Luc Dessart, R. Farmer, H. Hu, N. Langer, R. H. D. Townsend, Dean M. Townsley, and F. X. Timmes. Modules for Experiments in Stellar Astrophysics (MESA): Binaries, Pulsations, and Explosions. Ap. J. S., 220(1):15, September 2015. [Pou17] Juri Poutanen. Rosseland and Flux Mean Opacities for Compton Scattering. Ap. J., 835(2):119, February 2017. [Pri91] Denis Priou. Comparison between variational and traditional ap- proaches to relativistic thermodynamics of dissipative Ćuids. Phys. Rev. D, 43:1223Ű1234, Feb 1991. 196 [PTVF92] William H. Press, Saul A. Teukolsky, William T. Vetterling, and Brian P. Flannery. Numerical Recipes in FORTRAN (2nd Ed.): The Art of ScientiĄc Computing. Cambridge University Press, USA, 1992. [PY01] A. Y. Potekhin and D. G. Yakovlev. Thermal structure and cool- ing of neutron stars with magnetized envelopes. Astron. Astrophys., 374:213Ű226, July 2001. [RBB+02] Robert E. Rutledge, Lars Bildsten, Edward F. Brown, George G. Pavlov, Vyacheslav E. Zavlin, and Greg Ushomirsky. Crustal Emis- sion and the Quiescent Spectrum of the Neutron Star in KS 1731-260. Ap. J., 580(1):413Ű422, November 2002. [RFR+97] Felix Rembges, Christian Freiburghaus, Thomas Rauscher, Friedrich- Karl Thielemann, Hendrik Schatz, and Michael Wiescher. An Approx- imation for the rp-Process. Ap. J., 484(1):412Ű423, July 1997. [RGH+21] G. Raaijmakers, S. K. Greif, K. Hebeler, T. Hinderer, S. Nissanke, A. Schwenk, T. E. Riley, A. L. Watts, J. M. Lattimer, and W. C. G. Ho. Constraints on the Dense Matter Equation of State and Neu- tron Star Properties from NICERŠs Mass-Radius Estimate of PSR J0740+6620 and Multimessenger Observations. Astrophys. J. Lett., 918(2):L29, September 2021. [RI92] Forrest J. Rogers and Carlos A. Iglesias. Rosseland Mean Opacities for Variable Compositions. Ap. J., 401:361, December 1992. [RSD+20] A. Ray, A. K. Sikdar, P. Das, S. Pathak, and J. Datta. Unexpected in- crease of 7Be decay rate under compression. Phys. Rev. C, 101:035801, Mar 2020. [RT00] Thomas Rauscher and Friedrich-Karl Thielemann. Astrophysical re- action rates from statistical model calculations. Atomic Data and Nuclear Data Tables, 75(1):1Ű351, 2000. [SAB+01] H. Schatz, A. Aprahamian, V. Barnard, L. Bildsten, A. Cumming, M. Ouellette, T. Rauscher, F. K. Thielemann, and M. Wiescher. End Point of the rp Process on Accreting Neutron Stars. Phys. Rev. Lett., 86(16):3471Ű3474, April 2001. [Sal54] E. E. Salpeter. Electrons Screening and Thermonuclear Reactions. Australian Journal of Physics, 7:373, September 1954. [Sat79] K. Sato. Nuclear Compositions in the Inner Crust of Neutron Stars. Progress of Theoretical Physics, 62(4):957Ű968, October 1979. 197 [SBCW99] Hendrik Schatz, Lars Bildsten, Andrew Cumming, and Michael Wi- escher. The Rapid Proton Process Ashes from Stable Nuclear Burning on an Accreting Neutron Star. Ap. J., 524(2):1014Ű1029, October 1999. [Sch70] Bernard F. Schutz. Perfect Fluids in General Relativity: Velocity Potentials and a Variational Principle. Phys. Rev. D, 2(12):2762Ű2773, December 1970. [Sch09] Bernard Schutz. A First Course in General Relativity. Cambridge University Press, 2 edition, 2009. [SDD80] W. L. Slattery, G. D. Doolen, and H. E. DeWitt. Improved equation of state for the classical one-component plasma. Phys. Rev. A, 21:2087Ű 2095, Jun 1980. [Sem28] NN Semenov. Theories of combustion process. Z. Phys. Chem., 48:571Ű582, 1928. [Shk67] I. S. Shklovsky. On the Nature of the Source of X-Ray Emission of Sco XR-1. Astrophys. J. Lett., 148:L1, April 1967. [SOG+66] A. Sandage, P. Osmer, R. Giacconi, P. Gorenstein, H. Gursky, J. Wa- ters, H. Bradt, G. Garmire, B. V. Sreekantan, M. Oda, K. Osawa, and J. Jugaku. On the optical identiĄcation of SCO X-1. Ap. J., 146:316, October 1966. [ST83] S. L. Shapiro and S. A. Teukolsky. Black holes, white dwarfs, and neutron stars: The physics of compact objects. Wiley-Interscience, New York, 1983. [Sty04] Daniel F. Styer. What good is the thermodynamic limit? American Journal of Physics, 72(1):25Ű29, 2004. [SY06] P. S. Shternin and D. G. Yakovlev. Electron thermal conductiv- ity owing to collisions between degenerate electrons. Phys. Rev. D, 74(4):043004, August 2006. [Tay09] Barry N Taylor. Molar mass and related quantities in the new si. Metrologia, 46(3):L16, feb 2009. [TFG+07] W. P. Tan, J. L. Fisker, J. Görres, M. Couder, and M. Wiescher. 15O(α, γ)19Ne breakout reaction and impact on x-ray bursts. Phys. Rev. Lett., 98:242503, Jun 2007. 198 [TGB+09] W. P. Tan, J. Görres, M. Beard, M. Couder, A. Couture, S. Fala- hat, J. L. Fisker, L. Lamm, P. J. LeBlanc, H. Y. Lee, S. OŠBrien, A. Palumbo, E. Stech, E. Strandberg, and M. Wiescher. Measure- ment of the decay branching ratios of the α-unbound states in 19Ne and the 15O(α, γ) reaction rate. Phys. Rev. C, 79:055805, May 2009. [THW00] F. X. Timmes, R. D. Hoffman, and S. E. Woosley. An Inexpen- sive Nuclear Energy Generation Network for Stellar Hydrodynamics. Ap. J. S., 129(1):377Ű398, July 2000. [Tim99] F. X. Timmes. Integration of Nuclear Reaction Networks for Stellar Hydrodynamics. Ap. J. S., 124(1):241Ű263, September 1999. [TS00] F. X. Timmes and F. Douglas Swesty. The accuracy, consistency, and speed of an electron-positron equation of state based on table interpolation of the helmholtz free energy. The Astrophysical Journal Supplement Series, 126(2):501, feb 2000. [TWK+17] D. Torresi, C. Wheldon, Tz. Kokalova, S. Bailey, A. Boiano, C. Boiano, M. Fisichella, M. Mazzocco, C. Parascandolo, D. Pier- routsakou, E. Strano, M. Zadro, M. Cavallaro, S. Cherubini, N. Cur- tis, A. Di Pietro, J. P. Fernández Garcia, P. Figuera, T. Glodariu, J. Grȩbosz, M. La Cognata, M. La Commara, M. Lattuada, D. Men- goni, R. G. Pizzone, C. Signorini, C. Stefanini, L. Stroe, and C. Spi- taleri. Evidence for 15O + α resonance structures in 19Ne via direct measurement. Phys. Rev. C, 96:044317, Oct 2017. [van96] J. van Paradijs. On the Accretion Instability in Soft X-Ray Transients. Astrophys. J. Lett., 464:L139, June 1996. [vGI+94] L. van Wormer, J. Görres, C. Iliadis, M. Wiescher, and F. K. Thiele- mann. Reaction Rates and Reaction Sequences in the rp-Process. Ap. J., 432:326, September 1994. [Wal84] R. M. Wald. General Relativity. The University of Chicago Press, 1984. [WDP17] Rudy Wijnands, Nathalie Degenaar, and Dany Page. Cooling of Accretion-Heated Neutron Stars. Journal of Astrophysics and Astronomy, 38(3):49, September 2017. [Wey22] Hermann Weyl. Space - Time - Matter. Dover, New York, 1922. [WGS99] M Wiescher, J Görres, and H Schatz. Break-out reactions from the cno cycles. Journal of Physics G: Nuclear and Particle Physics, 25(6):R133, jun 1999. 199 [WGU+10] M. Wiescher, J. Görres, E. Uberseder, G. Imbriani, and M. Pignatari. The Cold and Hot CNO Cycles. Annual Review of Nuclear and Particle Science, 60:381Ű404, November 2010. [WHK+21] Meng Wang, W.J. Huang, F.G. Kondev, G. Audi, and S. Naimi. The ame 2020 atomic mass evaluation (ii). tables, graphs and references*. Chinese Physics C, 45(3):030003, mar 2021. [WM07] A. L. Watts and I. Maurer. Accretion rate and the occurrence of multi-peaked X-ray bursts. Astron. Astrophys., 467(2):L33ŰL36, May 2007. [WMM+01] Rudy Wijnands, Jon M. Miller, Craig Markwardt, Walter H. G. Lewin, and Michiel van der Klis. A Chandra Observation of the Long- Duration X-Ray Transient KS 1731-260 in Quiescence: Too Cold a Neutron Star? Astrophys. J. Lett., 560(2):L159ŰL162, October 2001. [WZW78] T. A. Weaver, G. B. Zimmerman, and S. E. Woosley. Presupernova evolution of massive stars. Ap. J., 225:1021Ű1029, November 1978. [ZBLM80] I.A. B Zeldovich, G.I. Barenblatt, V.B. Librovich, and G.M. Makhvi- ladze. Mathematical Theory of Combustion and Explosions. Nauka, 1980. [Zim01] J.M. Ziman. Electrons and Phonons: The Theory of Transport Phenomena in Solids. International series of monographs on physics. OUP Oxford, 2001. [ZPS96] V. E. Zavlin, G. G. Pavlov, and Yu. A. Shibanov. Model neutron star atmospheres with low magnetic Ąelds. I. Atmospheres in radiative equilibrium. Astron. Astrophys., 315:141Ű152, November 1996. [Zub20] Kai Zuber. Neutrino Physics. Taylor & Francis, Boca Raton, 2020. 200