Fact-checked by Grok 2 weeks ago

Stellar structure

Stellar structure refers to the internal physical configuration and processes of , which are self-gravitating spheres of primarily composed of and , where , , , and vary radially from the core to the surface, enabling sustained and output. The study of stellar structure relies on fundamental equations that describe the mechanical and of , including balancing gravitational forces with gradients, mass continuity relating enclosed mass to local , generation from reactions, and transport via or . These equations, solved numerically for specific stellar masses and compositions, predict key observables such as (L ∝ M^{3.5} to M^{4} for main-sequence ), radius (R ∝ M^{0.8}), and , which position on the Hertzsprung-Russell diagram. In stellar interiors, the core reaches temperatures of about 10^7 K and densities exceeding 100 g/cm³, where fuses into via processes like the proton-proton chain or , generating the energy that powers the star over billions of years. Energy is transported outward primarily through radiative diffusion in stable regions, where photons diffuse against opacity from ions and electrons, or in unstable layers where superadiabatic temperature gradients drive bulk plasma motion. The equation of state, incorporating pressure, , and partial electron degeneracy in denser regions, links these variables and influences structure, with opacity (κ) playing a critical role in determining transport efficiency. Stellar models reveal distinct zones: a convective core in massive stars for efficient mixing, radiative envelopes in low-mass stars, and a thin photosphere where optical depth drops to unity, emitting the blackbody-like spectrum observed externally. Composition gradients arise from nuclear processing, with helium and heavier elements accumulating centrally, while the overall metallicity (Z ≈ 0.02 for solar-like stars) affects opacity and fusion rates. These principles, validated against observations like solar neutrinos and helioseismology, form the foundation for understanding stellar evolution and diversity across the cosmic stellar populations.

Fundamental Principles

Hydrostatic Equilibrium

Hydrostatic equilibrium describes the condition in which the inward gravitational on stellar material is precisely balanced by the outward arising from the at every point within a , thereby maintaining its structural stability against collapse or expansion. This balance is essential for a to remain in a quasi-static state over much of its lifetime, allowing it to sustain its size and density profile. The concept was formalized in early 20th-century stellar theory, building on Newtonian applied to spherical distributions. The equation of is derived by considering the forces acting on a thin of thickness dr at radius r from the star's center. The net outward force due to the difference across the must counteract the gravitational attraction from the m(r) enclosed within radius r. For a of surface area $4\pi r^2, the mass element is dm = 4\pi r^2 \rho dr, where \rho is the local . Balancing these forces yields the : \frac{dP}{dr} = -\frac{G m(r) \rho(r)}{r^2}, where P is the , G is the , and the negative sign indicates that pressure decreases outward. This equation, first systematically applied to stellar interiors in classical treatments, relates the directly to the local and must be solved alongside the mass continuity equation to determine the star's internal structure. Physically, implies that stellar stability depends on the ability of internal s to support the star's weight against self-gravity, with the central typically as P_c \sim \frac{G M^2}{R^4} for a star of M and R, reaching values exceeding $10^{16} Pa in solar-type stars. The supporting arises from multiple mechanisms, depending on local conditions of , , and . In the non-degenerate envelopes and outer layers of most stars, dominates, given by P = \frac{\rho k T}{\mu m_H}, where k is Boltzmann's , T is , \mu is the mean molecular weight, and m_H is the ; this arises from the of ions and electrons in thermal motion. In hotter, denser interiors, contributes significantly, expressed as P_\mathrm{rad} = \frac{1}{3} a T^4 with a the , providing support through photon momentum in opaque regions. For compact objects like white dwarfs, electron degeneracy becomes crucial, a quantum mechanical effect where fermions occupy discrete energy states, yielding P_\mathrm{deg} \propto \rho^{5/3} independent of , enabling support at high densities without thermal reliance. The virial theorem provides a global perspective on hydrostatic equilibrium for self-gravitating systems like stars, stating that for a stable configuration in steady state, twice the total kinetic energy (primarily thermal) equals the absolute value of the gravitational potential energy: $2K + W = 0. This relation implies that about half of the gravitational energy released during a star's contraction is radiated away, while the remainder heats the interior to maintain pressure support, underscoring the interplay between mechanical equilibrium and energy balance. Deviations from this equilibrium, such as during rapid evolution phases, can lead to dynamical instabilities if the effective polytropic index falls below $4/3. Energy transport mechanisms briefly sustain the required temperature gradients for thermal pressure, but the core balance remains mechanical.

Mass Continuity

In stellar structure, the principle of mass continuity expresses the conservation of mass under the assumption of spherical symmetry, stating that the total mass m(r) enclosed within a radius r from the center is obtained by integrating the local density \rho(r') over the volume interior to that radius. This is mathematically formulated as m(r) = \int_0^r 4\pi {r'}^2 \rho(r') \, dr', which ensures that no mass is created or lost within the star, with the mass coordinate increasing monotonically outward. The differential form, \frac{dm}{dr} = 4\pi r^2 \rho, directly links the incremental mass to the density at each shell. The implications of this equation for profiles are profound, as \rho(r) uniquely determines the cumulative distribution m(r). Stars with higher central densities \rho_c exhibit a more rapid buildup of enclosed near , resulting in a steeper g(r) = -\frac{[G](/page/G) m(r)}{r^2}, which influences the overall pressure support required throughout the star. For instance, in centrally condensed configurations typical of main-sequence stars, m(r) remains small in the outer envelopes but rises sharply inward, constraining the possible forms of \rho(r) to avoid unphysical divergences or discontinuities. The total stellar mass M_* is simply the value of m(r) evaluated at the stellar radius R_*, so M_* = m(R_*) = \int_0^{R_*} 4\pi r^2 \rho(r) \, dr, serving as an essential boundary condition that normalizes the distribution. This relation underscores how the integrated profile sets the global scale, directly impacting the star's gravitational and . Mass continuity integrates seamlessly with , where the enclosed m(r) from this equation determines the gravitational force term in the \frac{dP}{dr} = -\frac{[G](/page/G) m(r) \rho}{r^2}, thereby constraining viable structures that maintain balance against . In simplified analytical models, assuming a polytropic P = K \rho^{1 + 1/n} (with polytropic index n), the combination yields the Lane-Emden equation \frac{1}{\xi^2} \frac{d}{d\xi} \left( \xi^2 \frac{d\theta}{d\xi} \right) = -\theta^n, where \rho = \rho_c \theta^n and dimensionless variables scale the profile, providing approximate solutions for density in stars like convective cores (n \approx 1.5) or radiative envelopes (n \approx 3). These polytropic models, pioneered in early 20th-century work, illustrate how mass continuity shapes realistic stellar interiors without requiring full numerical integration.

Energy Generation and Transport

Nuclear Energy Sources

Nuclear fusion processes in the cores of provide the primary source of energy that counteracts and sustains stellar , converting into while releasing energy through the defect according to the relation E = \Delta m c^2. This energy generation occurs at temperatures exceeding 10 million , where quantum tunneling enables protons to overcome electrostatic repulsion and fuse. The specific fusion pathways depend on the star's , with low-mass stars relying predominantly on the proton-proton () chain and higher-mass stars favoring the carbon-nitrogen-oxygen (. In low-mass stars like , the pp chain converts four protons into one nucleus through a series of reactions. The initiating step is the p + p \rightarrow ^2\mathrm{H} + e^+ + \nu_e + 0.42\,\mathrm{MeV}, followed by ^2\mathrm{H} + p \rightarrow ^3\mathrm{He} + \gamma + 5.49\,\mathrm{MeV}, and subsequent branches leading to the net reaction $4p \rightarrow ^4\mathrm{He} + 2e^+ + 2\nu_e + 26.7\,\mathrm{MeV}. Of this energy, approximately 26.2 MeV is retained in the star as , while 0.5 MeV is carried away by neutrinos. The pp chain accounts for over 99% of 's energy production due to its efficiency at core temperatures around 15 million K. For stars with masses above approximately 1.5 solar masses, higher core temperatures (around 20 million K or more) favor the , a catalytic process that also fuses four protons into but uses , , and oxygen as intermediaries. The net reaction is 4\,^1\mathrm{H} \rightarrow ^4\mathrm{He} + 2e^+ + 2\nu_e + 26.7\,\mathrm{MeV}, with , , and oxygen acting as catalysts that are regenerated to sustain the . The proceeds through steps such as ^{12}\mathrm{C} + p \rightarrow ^{13}\mathrm{N} + \gamma, followed by and further proton captures, with the rate limited by the slowest reaction involving ^{14}\mathrm{N} + p \rightarrow ^{15}\mathrm{O} + \gamma. This pathway dominates in massive stars because its reaction rates increase more rapidly with temperature than those of the pp chain. The local energy generation rate \epsilon, measured in erg g^{-1} s^{-1}, is a function of density \rho, temperature T, and hydrogen mass fraction X, typically expressed as \epsilon = \epsilon(\rho, T, X). For the pp chain at solar core conditions, it approximates \epsilon \propto \rho X^2 T^\nu with temperature exponent \nu \approx 4, reflecting the chain's moderate sensitivity to thermal conditions. In contrast, the CNO cycle exhibits stronger temperature dependence with \nu \approx 18, making it negligible in low-mass stars but crucial for the rapid evolution of massive ones. These dependencies ensure that energy production is concentrated in the dense, hot core, where \rho \approx 150 g cm^{-3} and T \approx 1.5 \times 10^7 K in the Sun. As hydrogen fusion progresses, the core composition evolves, with helium accumulating and hydrogen depleting, which decreases X and thereby reduces \epsilon over time, contributing to the star's departure from the . This buildup of inert helium increases the mean molecular weight, altering support and necessitating structural adjustments. In later evolutionary stages, after core hydrogen exhaustion, helium burning ignites in stars above about 0.5 masses via the : $3 ^4\mathrm{He} \rightarrow ^{12}\mathrm{C} + \gamma + 7.3\,\mathrm{MeV}. This reaction requires temperatures above 100 million K and proceeds rapidly once initiated, producing carbon as the primary product and enabling further in more massive stars.

Radiative and Convective Transport

In stellar interiors, generated primarily through in the core is transported outward to the surface via two main mechanisms: radiative diffusion and . Radiative transport dominates in regions where the is shallow enough for photons to carry effectively, while becomes prevalent in zones of steep gradients, where unstable motions efficiently redistribute . Radiative transport occurs through the diffusion of photons, governed by the radiative flux equation: F_{\rm rad} = -\frac{4acT^3}{3\kappa\rho} \frac{dT}{dr}, where a is the radiation constant, c is the speed of light, T is temperature, \kappa is opacity, \rho is density, and r is radius. This formula arises from the diffusion approximation in radiative transfer, assuming isotropic radiation and frequent scattering or absorption events that randomize photon directions. The mean free path of photons is typically short (~1 cm in dense interiors), leading to a random walk that slows outward propagation. Opacity \kappa, which quantifies the resistance to photon flow, depends on , temperature, and composition. Key sources include by free electrons, with \kappa_{\rm es} \approx 0.2(1+X) cm²/g for ionized (X is hydrogen mass fraction) and , dominant at high temperatures (>10^7 ) where interiors are fully ionized. Free-free absorption () follows Kramers' opacity law, \kappa_{\rm ff} \propto \rho T^{-7/2} (1+X), scaling linearly with due to electron-ion interactions and decreasing with temperature as thermal velocities reduce cross-sections. Bound-free absorption, from of ions, similarly obeys \kappa_{\rm bf} \propto \rho T^{-7/2} Z (1+X), where Z is , and peaks near ionization edges but contributes significantly in partially ionized regions. These opacities determine the radiative gradient \nabla_{\rm rad} = (d \ln T / d \ln P)_{\rm rad}, steeper in high-opacity zones. Convective transport activates when radiative transfer alone cannot carry the required energy flux, leading to buoyancy-driven instabilities. It ensues if the radiative exceeds the adiabatic gradient, per the Schwarzschild criterion: \nabla_{\rm rad} > \nabla_{\rm ad}, where \nabla_{\rm ad} = (\gamma - 1)/\gamma \approx 0.4 for an monatomic gas (\gamma = 5/3). In unstable regions, rising hot parcels expand adiabatically and continue ascending if their temperature remains above surroundings, driving vigorous mixing and on timescales of weeks to months, far more efficient than radiation. The convective flux can be approximated using mixing-length theory, though exact forms vary. In massive stars (>8 M_\odot), high central energy generation rates produce steep \nabla_{\rm rad} near the core, resulting in convective cores that mix fresh fuel efficiently. Conversely, low-mass stars like the Sun (~1 M_\odot) feature radiative interiors but convective envelopes in outer layers, where ionization zones increase opacity and destabilize the gradient. The total luminosity L = 4\pi r^2 F(r) integrates the local flux F(r) = F_{\rm rad} + F_{\rm conv} from core to surface, ensuring energy balance.

Equations of Stellar Structure

The Core Equations

The core equations of stellar structure form a system of four coupled, nonlinear ordinary differential equations that describe the radial variation of P, m, T, and L as functions of r within a spherically symmetric , incorporating hydrostatic support, mass distribution, energy generation, and energy transport. These equations are derived from fundamental physical principles and must be solved simultaneously with constitutive relations for the equation of state and opacity, which depend on \rho, T, and . The first equation is the condition of , which balances the inward gravitational force against the outward : \frac{dP}{dr} = -\frac{[G](/page/Gravitational_constant) m(r) \rho(r)}{r^2}, where [G](/page/Gravitational_constant) is the . The second equation expresses mass continuity, relating the enclosed mass to the local density: \frac{dm}{dr} = 4\pi r^2 \rho(r). The third equation governs energy transport, which can occur via or ; in the radiative case, it takes the form \frac{dT}{dr} = -\frac{3 \kappa(r) \rho(r) L(r)}{16 \pi a c r^2 T^3(r)}, where \kappa is the opacity, a is the radiation constant, and c is the , while convective transport requires mixing-length theory to relate the to convective . The fourth equation describes local energy generation and its contribution to the luminosity gradient: \frac{dL}{dr} = 4\pi r^2 \rho(r) \epsilon(\rho, T, \text{composition}), with \epsilon representing the nuclear energy generation rate per unit mass. To close the system, the equation of state relates pressure to density, temperature, and composition, typically P = P(\rho, T, X_i), where X_i are the abundances of , encompassing gas, radiation, and degeneracy pressures. Similarly, opacity is a function \kappa = \kappa(\rho, T, X_i), determining the efficiency of . Boundary conditions specify the central values at r = 0: m(0) = 0, L(0) = 0, and finite P(0) and T(0); at the surface r = R_*, P(R_*) = 0, L(R_*) = L_\text{total}, and T(R_*) = T_\text{eff}, where R_* is the stellar radius, L_\text{total} the luminosity, and T_\text{eff} the . The equations are highly coupled and nonlinear, as each variable appears in multiple relations, necessitating iterative numerical solutions; common methods include the shooting technique, which integrates from the center outward while adjusting central parameters to match surface conditions, and relaxation methods, which iteratively refine an initial mesh-based guess across the entire domain until convergence. Early progress in solving these equations came from Arthur Eddington's in , which assumed constant mean molecular weight and opacity to simplify the system into an analytical form resembling a polytrope of index 3.

Solutions and Approximations

Polytropic models offer an analytical framework for approximating solutions to the stellar structure equations by adopting a polytropic , in which P relates to \rho as P \propto \rho^{1 + 1/n}, where n is the polytropic index. This assumption simplifies the and mass continuity equations, reducing them to the Lane-Emden in dimensionless form: \frac{d}{d\xi} \left( \xi^2 \frac{d\theta}{d\xi} \right) = -\xi^2 \theta^n, where \theta(\xi) is a dimensionless function describing the density profile (\rho = \rho_c \theta^n), \xi is a scaled radial coordinate proportional to the physical radius, \rho_c is the central density, and the boundary conditions are \theta(0) = 1 and d\theta/d\xi|_{ \xi=0 } = 0. The Lane-Emden equation must be solved numerically for most n, but exact analytical solutions exist for specific integer values, providing insights into stellar density distributions and stability. Solutions for particular polytropic indices correspond to idealized stellar interiors. For n=0, the model assumes constant density throughout the star, yielding a uniform \theta = 1 - \xi^2/6 and a simple spherical profile with the surface at \xi_1 = \sqrt{6}. The n=1 case, solvable with \theta = \sin(\xi)/\xi, represents an isothermal sphere with a density that falls off gradually, relevant for cores or planetary atmospheres. For n=1.5, the solution approximates fully convective stars under an adiabatic equation of state, showing moderate central concentration where most mass resides in the core; this index is often used for low-mass main-sequence stars. The n=3 polytrope, with its surface at \xi_1 \approx 6.9, models radiative stars or white dwarf envelopes, exhibiting high central concentration (over 50% of mass within 20% of the radius) and is crucial for understanding energy transport in massive stars. Across these models, central concentration increases with n, from uniform for n=0 to highly peaked for n \geq 3, while the stellar radius scales inversely with central density as R \propto \rho_c^{(1-n)/2n} for a fixed mass, highlighting how denser cores lead to more compact stars. Homology relations provide scaling laws for properties of stars with self-similar internal structures, assuming homologous and profiles that differ only by a scaling factor. These relations derive from the stellar structure equations under assumptions of opacity and energy generation rates, yielding power-law dependencies on M. For low-mass main-sequence stars dominated by radiative transport and the Kramers opacity law, scales as L \propto M^3, while for higher-mass stars with convective cores and similar opacity, the relation steepens to L \propto M^{3.5}. Radius follows R \propto M^{0.8} in the low-mass regime, reflecting the balance between gravitational contraction and pressure support. Such scalings enable quick estimates of stellar properties without full integration, though they break down for extreme masses or varying compositions. For realistic stellar models beyond these approximations, numerical methods integrate the full set of structure equations iteratively from center to surface. Stellar evolution codes employ schemes on discretized radial grids, often with adaptive refinement to resolve steep gradients in convective zones or burning regions. The Modules for Experiments in Stellar Astrophysics (MESA) is a prominent open-source code that modularizes physics inputs like networks and opacity tables, solving the equations via relaxation methods to achieve at each evolutionary timestep. MESA's flexibility allows simulations of diverse stellar types, from protostars to supernovae progenitors, by coupling structure solvers with time-dependent composition changes. While polytropic and homologous models excel in illustrating basic physics for uniform-composition stars, they are limited to idealized cases and fail to capture or gradients in mean molecular weight from nuclear processing. Convective mixing and overshooting, which alter energy transport and profiles, necessitate comprehensive numerical approaches to accurately model real stellar interiors.

Structural Variations and Evolution

Main Sequence Structure

The main sequence phase of stellar evolution is characterized by stable hydrogen fusion in the core, where stars spend the majority of their lifetimes in hydrostatic and . The zero-age main sequence (ZAMS) denotes the initial configuration upon the onset of core hydrogen burning, following the end of protostellar , at which point the central reaches approximately $10^7 K, sufficient to ignite the proton-proton chain or depending on mass. At this stage, the star's internal structure is largely homogeneous in composition, with a central concentration of and that supports sustained generation. Stellar structure on the exhibits strong dependence on mass, influencing the extent of convective and radiative zones. Low-mass stars with M < 1.5\, M_\odot possess radiative cores where transport occurs primarily via , enveloped by extensive convective zones that extend inward from near the surface; this configuration arises because their lower core temperatures favor slower rates and higher opacities in the . In contrast, stars with M \gtrsim 1.5\, M_\odot feature convective cores due to the high temperatures and fluxes that drive vigorous mixing, paired with radiative s where dominates transport; the convective core typically encompasses 10-20% of the star's mass, facilitating efficient fuel mixing. Intermediate-mass stars display transitional structures, with the boundary between convective and radiative regions shifting based on opacity and . A representative example is the Sun, a $1\, M_\odot main-sequence star whose structure is well-modeled by the (SSM). The solar core reaches a central of about $1.5 \times 10^7 K and density of roughly 150 g/cm³, enabling the proton-proton chain to produce the necessary energy output. The interior consists of a radiative zone extending from the center to approximately 0.7 R_\odot, where energy is transported by radiation, transitioning to a convective envelope in the outer layers; this boundary is marked by the tachocline, a thin shear layer of about 0.05 R_\odot thickness that accommodates differential rotation between the rigidly rotating interior and the differentially rotating convection zone. Throughout the main sequence, composition gradients develop due to nuclear processing, particularly in the core. On the ZAMS, the mean molecular weight \mu is uniformly around 0.6, reflecting the primordial hydrogen-helium mix. As hydrogen fuses into helium, \mu increases in the core to values up to approximately 0.86 for solar-mass stars at mid-sequence, enhancing pressure support and contributing to gradual structural adjustments. Stellar evolution models reveal mass-dependent scaling relations for main-sequence stars near solar composition: the radius follows R \propto M^{0.8}, while luminosity scales as L \propto M^{3.5}, reflecting the steeper rise in nuclear energy output with mass.

Post-Main Sequence Changes

After the exhaustion of core fuel, stars transition to shell burning, where occurs in a thin shell surrounding the contracting inert helium core. This process causes the core to contract under gravity, increasing its density and temperature, while the overlying envelope expands dramatically due to reduced support and altered energy transport. The stellar radius can increase by factors of 10 to several hundred, transforming low- to intermediate-mass stars into red giants with radii up to about 250 solar radii for solar-mass progenitors. In low-mass stars (typically below 2 solar masses), the contracting becomes electron-degenerate, leading to the at the tip of the . Here, the ignites explosively in the degenerate , producing a sudden release of energy as three nuclei fuse into carbon-12. However, the degeneracy pressure resists , containing the outburst and preventing a ; instead, it drives convective mixing that adjusts the structure without disrupting the envelope. For higher-mass stars (above about 2 solar masses), core helium burning occurs under non-degenerate conditions, forming a convective core where the sustains stably. This phase establishes a more extended convective region in the core compared to the . In even more massive stars (above 8 solar masses), advanced nuclear burning stages ensue, such as carbon burning, which requires central temperatures exceeding 6 × 10^8 K to ignite into heavier elements like and magnesium. Post-main-sequence structures exhibit a dichotomy between low- and high-mass stars. Low-mass stars develop thin and burning shells around an inert, degenerate , maintaining a relatively stable envelope despite expansion. In contrast, massive stars form an onion-like layering of successive burning shells, with heavier elements fusing in deeper, hotter regions surrounded by shells of lighter elements, driven by the star's high central temperatures and convective mixing. Mass loss intensifies during these phases through enhanced stellar winds, particularly in the expanded envelopes of red giants and supergiants, which strip outer layers and alter the overall structure. In low-mass stars, this culminates in the ejection of the envelope as a , exposing the hot . For massive stars, extreme winds can erode the hydrogen envelope entirely, revealing Wolf-Rayet stars characterized by helium and heavy-element surfaces and high mass-loss rates up to 10^{-5} solar masses per year. These structural changes occur on contrasting timescales: core evolution proceeds rapidly in massive stars, with helium and advanced burning phases lasting 10^5 to 10^6 years due to high temperatures accelerating , while envelope adjustments in low-mass stars evolve more slowly over 10^7 years or longer, dominated by thermal . Nuclear energy sources shift from the proton-proton chain to helium burning via the during this transition.

References

  1. [1]
    [PDF] STELLAR STRUCTURE AND EVOLUTION
    This course aims to understand the structure and evolution of stars, their physics, and observational properties, using known laws of physics. A star radiates ...
  2. [2]
    [PDF] STELLAR ASTROPHYSICS - Non-secure http index page
    To begin our study of stellar structure, let us first consider the star that we know best, our sun. The planetary orbits and the gravita- tional constant G ...
  3. [3]
    The internal constitution of the stars : Eddington, Arthur Stanley, Sir ...
    Jun 18, 2022 · The internal constitution of the stars. by: Eddington, Arthur Stanley, Sir, 1882-1944. Publication date: 1926. Topics: Stars.
  4. [4]
    An Introduction To The Study Of Stellar Structure - Internet Archive
    Jan 22, 2017 · An Introduction To The Study Of Stellar Structure. by: University Of Chicago Press. Publication date: 1939. Topics: Other. Collection ...
  5. [5]
    [PDF] POLYTROPES
    Polytropes are self-gravitating gaseous spheres that were, and still are, very useful as crude approximation to more realistic stellar models.
  6. [6]
    Star Basics - NASA Science
    Stars are giant balls of hot gas – mostly hydrogen, with some helium and small amounts of other elements. Every star has its own life cycle.
  7. [7]
    Stellar Radiation & Stellar Types - ESA Science & Technology
    Nuclear Fusion. A star forms out of a slowly condensing cloud of gas. As the pressure and density in the core increase so does the temperature.
  8. [8]
    [PDF] Fusion Cycles in Stars and Stellar Neutrinos - arXiv
    Oct 1, 2002 · Proton-proton reactions are the main source of stellar ... The proton-proton chain is seen to generate the largest luminosity fractions.<|control11|><|separator|>
  9. [9]
    [PDF] arXiv:astro-ph/9405040v1 18 May 1994
    The net reaction for the proton-proton chain and the CNO cycle is the conversion of hydrogen into helium, 4p −→ α + 2νe + 2e+ + Q, Q = 26.7MeV, where two ...
  10. [10]
    Stellar Processes and Evolution - ESA Science & Technology
    Proton-Proton Chain. The simplest reaction that occurs in stars is the ... Six protons go into the cycle, but two come back out. The overall process ...
  11. [11]
    [PDF] Stellar models and stellar stability
    Stellar models are constructed using physical processes and equations, and the chapter analyzes these equations and their stability.Missing: solar | Show results with:solar<|control11|><|separator|>
  12. [12]
    Hydrostatic Burning Phases - Book chapter - IOPscience
    7.3. Helium Burning. 7.3.1. Reactions. When the hydrogen in the core of a star has been exhausted and hydrogen burning has moved into a shell, the mass of ...
  13. [13]
    [PDF] An update on fine-tunings in the triple-alpha process - arXiv
    Jun 3, 2019 · Abstract. The triple-alpha process, whereby evolved stars create carbon and oxygen, is believed to be fine-tuned to a high degree.
  14. [14]
    [PDF] Chapter 5 - Energy transport in stellar interiors - Astrophysics
    The transport of energy in stars is the subject of this chapter, which will lead us to two additional differential equations for the stellar structure. 5.1 ...Missing: continuity | Show results with:continuity
  15. [15]
    [PDF] Radiative energy transport in stellar interior - IUCAA
    If we now define an effective opacity κ through. 1 κ. = 1 κrad. +. 1 κcd then the temperature gradient can still be written as in equations (6) and (7) by.
  16. [16]
    [PDF] Lecture 7 – Part 1 Sources of Stellar Opacity
    At temperatures higher than a few tens of million K, stellar interiors are virtually completely ionized, and electron scattering is the dominant opacity source.
  17. [17]
    [PDF] 1. Convection in Stars
    Schwarzschild's criterion for stability. If the logarithmic derivative of the temperature with respect to the pressure exceeds 0.4, convection will commence.
  18. [18]
    [PDF] Equations of Stellar Structure
    The set of equations (1.11) describes the time evolution of a spherically symmetric star with a given distribution of chemical composition with mass, Xi(Mr), ...Missing: continuity papers
  19. [19]
    Stellar Structure and Evolution - SpringerLink
    Front Matter. Pages 87-87. Download chapter PDF · The Differential Equations of Stellar Evolution. Rudolf Kippenhahn, Alfred Weigert, Achim Weiss. Pages 89-92 ...
  20. [20]
    [PDF] Solving the Equations - PHYS 633: Introduction to Stellar Astrophysics
    The first is known as 'shooting'; basically, we shoot a solution from one boundary to the other, adjusting our aim until we satisfy both sets of boundary.
  21. [21]
  22. [22]
  23. [23]
    Star formation - Astronomy 1101 - The Ohio State University
    As the collapse proceeds, the core temperature gets hotter until it reaches ~10 Million K: ... Star reaches the Zero-Age Main Sequence as a full-fledged star in ...
  24. [24]
    An Atlas of Convection in Main-sequence Stars - IOPscience
    Aug 24, 2022 · All mass ranges are approximate and metallicity dependent. To compute Ek we use Ω corresponding to a surface velocity of 150 km s−1 for O–F ...Abstract · Introduction · Convection Zones · Discussion
  25. [25]
    [PDF] Stellar evolution and the Standard Solar Model
    Regarding solar neutrino production, at the temperature and density characteristic of the solar interior (T ≈ 107 K, ρ ≈150 g/cm3), hydrogen burns with ...<|control11|><|separator|>
  26. [26]
    Dynamics of the Tachocline | Space Science Reviews
    Dec 14, 2023 · The solar tachocline is an internal region of the Sun possessing strong radial and latitudinal shears straddling the base of the convective envelope.
  27. [27]
    [PDF] Today in Astronomy 241: stellar evolution on, and slightly off, the ...
    Mar 19, 2024 · Stellar evolution on the main sequence. • As hydrogen burns in the core, the mean molecular weight slowly increases. At a given temperature ...
  28. [28]
    [PDF] Stellar Masses
    This mass-luminosity relationship is considered one of the most fundamental descriptions of stellar properties; the ability to reproduce this by stellar models ...
  29. [29]
    Why Do Stars Turn Red? I. Post-Main-Sequence Expansion ... - arXiv
    Jul 31, 2024 · We identify the key physical criteria and mechanisms that govern envelope expansion when stars evolve from the core hydrogen-burning to helium-burning stage.
  30. [30]
    The core helium flash revisited III. From Pop I to Pop III stars - arXiv
    Mar 18, 2010 · Abstract:Degenerate ignition of helium in low-mass stars at the end of the red giant branch phase leads to dynamic convection in their ...
  31. [31]
    [1512.04845] The treatment of mixing in core helium burning models
    Dec 15, 2015 · Abstract:The treatment of convective boundaries during core helium burning is a fundamental problem in stellar evolution calculations.Missing: massive | Show results with:massive