Fact-checked by Grok 2 weeks ago

Radiative zone

In stellar interiors, the radiative zone is a layer where energy is transported outward primarily by radiation rather than convection. In the Sun, it is a major layer extending from roughly 25% to 70% of the solar radius outward from the center, where energy produced by nuclear fusion in the core is transported toward the surface primarily through radiative diffusion. This zone, which comprises about 32% of the Sun's volume (approximately 48% of its mass), is characterized by its high density and opacity, causing photons to undergo frequent scattering interactions with ionized particles, resulting in an energy transfer process that takes approximately one million years for light to cross the region despite traveling at the speed of light between collisions. Temperatures in the radiative zone decrease gradually from about 7 million Kelvin at its inner boundary to around 2 million Kelvin at the outer edge, while density drops from 20 g/cm³ to 0.2 g/cm³ over this span. Composed mainly of ionized hydrogen (about 73%) and helium (about 25%) plasma with trace heavier elements, the zone's stable stratification prevents convective motions, ensuring efficient but slow radiative heat flow. The radiative zone plays a crucial role in the Sun's energy dynamics, acting as a thermal buffer that diffuses the immense heat from without significant mixing, which helps maintain the of the overlying convective zone. Observations from helioseismology have confirmed its , revealing subtle variations in and nearly uniform at a rate intermediate to the convection zone's differential , slightly faster than the innermost but overall slower than the equatorial convection zone and faster than its polar regions. This layer's boundaries are marked by the tachocline at its outer edge, a thin transition region where shear forces influence solar dynamo activity and magnetic field generation. Overall, the radiative exemplifies the radiative transfer processes dominant in the interiors of many main-sequence stars like the Sun, where opacity from abundant ions governs propagation.

Fundamentals

Definition and Role

The radiative zone is a distinct layer within the interior of a where generated in is transported outward primarily through , involving the of photons that are repeatedly absorbed and re-emitted by ions and electrons, rather than by or conduction. This process dominates in regions where the and gradients allow for stable without triggering convective instability. In solar-type stars, the radiative zone occupies a structural position between the central energy-generating core and the overlying convective zone, extending from roughly 25% to 70% of the solar radius. Its inner boundary aligns with the outer edge of the core, while the outer boundary, often marked by the tachocline, occurs where the radiative flux equals the star's total luminosity divided by the surface area at that radius, transitioning to convective transport beyond. This positioning ensures a controlled outward flow of energy, preventing rapid heat loss from the core. The radiative zone plays a crucial evolutionary role as a thermal insulator, significantly slowing the escape of from —photons may take up to a million years to traverse it—thereby supporting the star's long-term and facilitating a gradual decline in and outward. This insulation is vital for sustaining rates in over billions of years. The concept of the radiative zone was first conceptualized in early 20th-century stellar models, notably by in his 1926 work The Internal Constitution of the Stars, which incorporated to explain observed stellar luminosities and mass-luminosity relations. These foundational models assumed a polytropic with radiative transport dominating the interiors of main-sequence stars.

Energy Transport Mechanism

In the radiative zone of a , energy is transported outward primarily through the of , which are repeatedly absorbed and re-emitted by ions and electrons in the . This process results in a , where each travels a short —typically on the order of millimeters to centimeters—before interacting again, causing the net energy flow to propagate slowly outward due to the . The radiative flux F_{\rm rad} is described by the diffusion approximation of , given by F_{\rm rad} = -\frac{4acT^3}{3\kappa \rho} \frac{dT}{dr}, where a is the radiation constant, c is the , T is the , \kappa is the opacity, \rho is the , and \frac{dT}{dr} is the radial . This formula arises from the balance between the gradient and absorption, assuming local and isotropic scattering. The time scale for photon diffusion across the star is estimated as t_{\rm diff} \approx \frac{R^2}{c \lambda}, where R is the stellar radius and \lambda is the mean free path of the photon. For the Sun, this yields approximately 170,000 years for energy generated in the core to reach the base of the convection zone via radiative diffusion. Conduction is negligible as an energy transport mechanism in stellar plasmas because the speeds of charged particles are much less than the and their mean free paths are much shorter than those of photons, making thermal conductivity insignificant compared to . In non-degenerate matter, radiative transport also dominates over neutrino emission, which primarily affects the energy budget in the dense but contributes little in the radiative zone.

Physical Characteristics

Temperature and Density Gradients

In the radiative zone of a star, the temperature gradient is determined by the requirement to transport energy outward via radiation while maintaining hydrostatic equilibrium. The radial temperature gradient follows the radiative transfer relation, given by \frac{dT}{dr} = -\frac{3 \kappa \rho L(r)}{16 \pi a c T^3 r^2}, where T is the temperature, \kappa is the opacity, \rho is the density, L(r) is the luminosity at radius r, a is the radiation constant, and c is the speed of light. This gradient is steeper near the core because L(r) increases with accumulated energy production from nuclear fusion, leading to a more rapid temperature decrease per unit radius in inner regions compared to outer parts of the zone. Unlike convective regions, this radiative gradient remains sub-adiabatic, ensuring stability against buoyancy-driven motions. The density profile in the radiative zone decreases monotonically outward, governed by the equation of hydrostatic equilibrium, \frac{dP}{dr} = -\frac{G M(r) \rho}{r^2}, where P is the pressure, G is the gravitational constant, and M(r) is the mass interior to radius r. For ionized gas in this region, the pressure is primarily supported by thermal motions, following the ideal gas law P = \frac{\rho k T}{\mu m_H}, with k as Boltzmann's constant, \mu the mean molecular weight, and m_H the hydrogen mass. As temperature declines outward, density must drop more sharply to balance the decreasing pressure support, resulting in a profile that is tightly coupled to the temperature structure. Observational constraints from helioseismology and standard solar models provide specific profiles for , where the radiative zone spans approximately 0.25 to 0.7 radii. At the inner boundary near edge, temperatures reach about 7 million K, dropping to roughly 2 million K at the outer boundary with the convective zone; densities similarly fall from around 20 g/cm³ to 0.2 g/cm³ over this interval. These gradual gradients, inferred from inverted sound-speed profiles, confirm the zone's role in slowly diffusing energy generated in . The smooth temperature and gradients in radiative zones support overall by preventing convective overturn, as the suffices to carry the energy load without exceeding the adiabatic limit. Composition variations, such as helium accumulation toward the core due to gravitational settling in the radiative interior, further influence these gradients by altering the mean molecular weight \mu and opacity, leading to subtle enhancements in near .

Radiative Opacity and Transfer

In the radiative zone of stars, opacity arises primarily from and processes in ionized plasmas, which determine the of photons and thus the efficiency of radiative energy transport. , known as in the non-relativistic limit applicable to stellar interiors, provides a frequency-independent opacity given by the mass extinction coefficient \kappa_{es} = 0.2(1 + X) cm²/g, where X is the mass fraction; this accounts for electrons from fully ionized and , with the Thomson cross-section per \sigma_T \approx 6.65 \times 10^{-25} cm². In regions of high , such as the inner radiative zones, bound-free ( of bound electrons) and free-free (, where photons are absorbed during electron-ion encounters) contribute significantly to the continuum opacity, particularly at temperatures below a few million where states allow these transitions. To compute the effective opacity for energy transport in stellar models, the Rosseland mean opacity \kappa_R is employed, which is a weighted by the temperature derivative of the blackbody intensity, suitable for the regime: \frac{1}{\kappa_R} = \frac{\int_0^\infty \frac{1}{\kappa_\nu} \frac{\partial B_\nu}{\partial T} \, d\nu}{\int_0^\infty \frac{\partial B_\nu}{\partial T} \, d\nu}, where \kappa_\nu is the frequency-dependent opacity and B_\nu(T) is the Planck function; this averaging emphasizes contributions near the peak of the blackbody spectrum, capturing the dominant photon interactions in optically thick environments. The underlying physics is governed by the radiative transfer equation along a path s: \frac{dI_\nu}{ds} = -\kappa_\nu \rho I_\nu + \kappa_\nu \rho S_\nu, where I_\nu is the specific intensity, \rho is the , and S_\nu is the source function (approximating the local Planck function B_\nu in local ); in the radiative zone's optically thick conditions (\tau \gg 1), this simplifies to the diffusion approximation, where the F \propto -\frac{1}{3\kappa_R \rho} \nabla (a T^4) with a the radiation constant, allowing photons to random-walk outward over mean free paths of order centimeters to meters. Opacity varies with composition and temperature, being higher in metal-rich stars due to increased bound-free and bound-bound transitions from heavier elements, which can enhance \kappa_R by factors of 2–10 compared to hydrogen-helium mixtures; opacity generally decreases with rising temperature as ionization completes and absorption cross-sections drop (following Kramers' opacity \kappa \propto \rho T^{-3.5} for free-free processes), thereby influencing the radiative zone's thickness by permitting steeper temperature gradients at higher temperatures. Modern computations rely on detailed opacity tables such as , which incorporate updated atomic data and show peaks around T \sim 2 \times 10^5 K from iron-group ions in metal-enriched plasmas, with revisions post-1996 improving accuracy for solar models.

Theoretical Models

Eddington Stellar Model

The Eddington stellar model represents a foundational analytical framework for understanding , particularly in regions dominated by radiative energy transport, such as the radiative zones of stars. Developed by in the 1920s, the model assumes throughout the star, where energy is transported solely by radiation under the influence of a constant mean opacity K. To achieve convective neutrality—ensuring the radiative matches the adiabatic gradient—the model adopts a polytropic with index n=3, equivalent to an adiabatic index \gamma = 4/3. This leads to a pressure-density relation P = K' \rho^{4/3}, where K' is a constant determined by the . The density profile is then derived by solving the Lane-Emden equation for n=3, \frac{1}{\xi^2} \frac{d}{d\xi} \left( \xi^2 \frac{d\theta}{d\xi} \right) = -\theta^3, with boundary conditions \theta(0) = 1 and \theta'(0) = 0, providing a dimensionless structure that scales to physical dimensions via the star's total mass and the polytropic constant. Central to the model is the parameter \beta, defined as the ratio of gas pressure to total pressure (P_g / P), assumed constant across the star to simplify the inclusion of both gas and radiation pressure contributions. The radiation pressure fraction $1 - \beta quantifies the role of radiation in supporting the star against gravity. By integrating the equations of hydrostatic equilibrium, radiative transfer (with the diffusion approximation for flux F = -\frac{4ac T^3}{3\kappa \rho} \frac{dT}{dr}), and the polytropic relation, Eddington derived the relation \frac{1 - \beta}{\beta^4} = \frac{3 K L \mu^4}{16 \pi a c G M^3}, which forms the basis of the quartic equation for \beta, where \mu is the mean molecular weight, L is the luminosity, M is the mass, a is the radiation constant, c is the speed of light, and G is the gravitational constant. This equation emerges from matching the integrated radiative luminosity constraint to the structural support provided by the n=3 polytrope, highlighting how higher luminosity or lower mass increases the reliance on radiation pressure. Solving the quartic for \beta yields the equilibrium configuration, with solutions existing only for certain ranges of M and \mu. Despite its simplicity, the model predicts a mass-luminosity L \propto M^3 for main-sequence stars under constant opacity, capturing the scaling observed in intermediate-mass stars and providing early insight into why massive stars are -pressure dominated. However, it remains idealized, neglecting composition gradients, variable opacity, and convective regions, which limits its applicability to detailed radiative zone modeling; contemporary numerical simulations, incorporating full opacity tables and networks, have largely superseded it for precision. Eddington's work, published in , laid the groundwork for modern stellar theory by demonstrating the interplay of and in stellar interiors.

Radiative Transfer Equations

The mathematical framework for in stellar interiors is derived from the equation of (RTE), which describes the propagation of specific intensity I_\nu along rays in a medium with and . To make the RTE tractable for stellar models, it is often expanded into a of equations by integrating over angles. The zeroth corresponds to the radiation u_\nu = \frac{4\pi}{c} J_\nu, where J_\nu = \frac{1}{2} \int_{-1}^{1} I_\nu(\mu) \, d\mu is the mean intensity and \mu = \cos\theta. The first relates to the \mathbf{F}_\nu = 4\pi H_\nu \hat{n}, with H_\nu = \frac{1}{2} \int_{-1}^{1} I_\nu(\mu) \mu \, d\mu. These satisfy coupled partial differential equations that account for , , and geometric dilution, but the is infinite and requires closure. The Eddington approximation provides this by assuming local isotropy in the radiation field at large optical depths, typical of stellar radiative zones, setting the second moment K_\nu = \frac{1}{3} J_\nu, where K_\nu = \frac{1}{2} \int_{-1}^{1} I_\nu(\mu) \mu^2 \, d\mu. This simplifies the system to two equations: the zeroth-moment equation for and the first-moment equation for , yielding a relation between and the of , \mathbf{F}_\nu = -\frac{c}{3 \kappa_\nu \rho} \nabla u_\nu. Such moment-based methods with Eddington form the basis for in hydrostatic stellar interiors, as applied analytically in models like the Eddington . In the diffusion limit, valid deep in radiative zones where the optical depth \tau \gg 1 and photon mean free paths are short compared to the scale height, the transfer equations reduce further. Steady-state energy balance implies \nabla \cdot \mathbf{F} = 0 in regions without nuclear sources or sinks, leading to the diffusive flux expression: \mathbf{F} = -\frac{c}{3 \kappa \rho} \nabla u, where u = a T^4 is the integrated radiation energy density, \kappa is the Rosseland mean opacity, \rho is density, and c is the speed of light. Equivalently, in terms of temperature, \mathbf{F} = -\frac{4 a c T^3}{3 \kappa \rho} \nabla T. This approximation captures the random-walk nature of photon transport, with luminosity L(r) = 4\pi r^2 F increasing outward to match energy generation. Numerical implementations of these equations are integral to modern stellar evolution codes, which solve the moment equations implicitly within a time-dependent framework to evolve stellar structure. For instance, the Modules for Experiments in Stellar Astrophysics (MESA) code employs a fully implicit, adaptive mesh solver for the energy and hydrostatic equations, incorporating the diffusion approximation with Rosseland opacities tabulated as functions of temperature, density, and metallicity. Frequency-dependent opacities are handled via multi-group methods or precomputed tables (e.g., from OP or OPAL projects), allowing efficient computation of radiative accelerations and gradients while conserving total energy. These methods enable modeling of complex interiors over gigayear timescales. Post-2010 advancements have extended these frameworks to include relativistic effects and , addressing limitations in standard treatments for high-mass or evolved stars. Relativistic incorporates Doppler boosting and aberration in plane-parallel flows, modifying the moment equations with Lorentz-invariant forms to better model near Eddington limits in massive star interiors. Similarly, introduce anisotropic opacity and Lorentz forces, perturbing the flux via magnetized ; recent simulations couple these to the RTE using variable Eddington factors in models, revealing enhanced transport in radiative zones. These developments, often implemented in multidimensional codes, improve predictions for stellar winds and evolution.

Stability Criteria

Conditions for Radiative Stability

The radiative zone in a maintains stability through specific conditions that ensure transport occurs primarily via without triggering instabilities. A fundamental requirement is that the local luminosity L(r) at any radius r must remain below the local L_{\rm Edd}(r) = \frac{4\pi G M(r) c}{\kappa}, where M(r) is the mass enclosed within r, G is the , c is the , and \kappa is the opacity. This condition prevents from overcoming gravitational forces, thereby upholding and avoiding dynamical instability. Additionally, the radiative \nabla_{\rm rad} = \frac{d \ln T}{d \ln P} must be less than the adiabatic gradient \nabla_{\rm ad}, ensuring that the actual temperature gradient required for radiative transport does not exceed the stable adiabatic profile, thus suppressing convective motions. Compositional uniformity plays a key role in these stability conditions, with models typically assuming a constant mean molecular weight \mu throughout the zone to simplify the equation of and maintain thermodynamic consistency. This assumption holds in regions without significant composition gradients from processing or mixing. Furthermore, is bolstered in high-temperature regimes where radiative efficiency is high due to lower opacities, allowing photons to diffuse freely and transport without steepening the excessively. The boundaries of the radiative zone are defined by energy input and gradient transitions. At the inner edge, adjacent to the core, the luminosity L(r) is supplied by fusion reactions, providing a steady heat source that initiates radiative transport outward. The outer boundary occurs where \nabla_{\rm rad} begins to approach or exceed \nabla_{\rm ad}, marking the transition to regions where alternative transport mechanisms dominate, though the zone itself remains stable internally. Observational evidence from helioseismology supports these stability conditions in the Sun's radiative zone, where acoustic wave inversions reveal a rigidly rotating with minimal and no significant , consistent with profiles satisfying \nabla_{\rm rad} < \nabla_{\rm ad} and sub-Eddington luminosities.

Distinction from Convective Instability

The distinction between radiative zones and convective instability hinges on specific criteria that determine whether a stellar layer will undergo buoyancy-driven overturning or remain stable to small perturbations. The foundational Schwarzschild criterion assesses convective by comparing the actual temperature gradient \nabla = \frac{d \ln T}{d \ln P} to the adiabatic gradient \nabla_{\rm ad}. A layer is stable against convection—and thus qualifies as radiative—if the radiative gradient satisfies \nabla_{\rm rad} < \nabla_{\rm ad}, where for an ideal monatomic gas with adiabatic index \gamma = 5/3, \nabla_{\rm ad} = \frac{\gamma - 1}{\gamma} = 0.4. This condition ensures that a displaced fluid element expands adiabatically and becomes denser than its surroundings, sinking back to its original position without triggering convection. In contrast, convective regions occur where \nabla_{\rm rad} > \nabla_{\rm ad}, leading to as the element becomes buoyant and rises, promoting mixing. The Ledoux criterion provides a more comprehensive framework by incorporating gradients in mean molecular weight \mu, which are crucial in evolving stars where composition varies due to nuclear processing. It states that stability holds if \nabla_{\rm rad} < \nabla_{\rm ad} + \frac{\phi}{\delta} \nabla_\mu, where \nabla_\mu = \frac{d \ln \mu}{d \ln P} is the molecular weight gradient, and \phi = \left( \frac{\partial \ln \rho}{\partial \ln \mu} \right)_{P,T}, \delta = -\left( \frac{\partial \ln \rho}{\partial \ln T} \right)_{P,\mu} are thermodynamic quantities typically of order unity for stellar plasmas. Unlike the Schwarzschild criterion, which assumes uniform composition (\nabla_\mu = 0), the Ledoux term \frac{\phi}{\delta} \nabla_\mu introduces additional stabilization when \nabla_\mu > 0 (a \mu-gradient with mean molecular weight increasing inward, as heavier elements settle inward), preventing full convection in regions that might otherwise be Schwarzschild-unstable. This difference is particularly relevant in models, where ignoring \mu-gradients can overestimate convective extents; the Ledoux approach better captures phenomena like semiconvection, where weak, layered instabilities arise instead of vigorous mixing. Instability thresholds are crossed when these criteria are violated: under the Ledoux , convection ensues if \nabla_{\rm rad} > \nabla_{\rm ad} + \frac{\phi}{\delta} \nabla_\mu, initiating buoyancy-driven motions that efficiently transport energy outward. In the Sun's radiative zone, extending from about 0.25 to 0.7 solar radii, stability is maintained because \nabla_{\rm rad} remains well below \nabla_{\rm ad} throughout, with nearly uniform \mu ensuring the Ledoux term provides no additional destabilization. This low \nabla_{\rm rad}, driven by efficient radiative opacity in the ionized , contrasts sharply with the overlying convective zone, where superadiabatic gradients dominate.

Occurrence in Stars

Main Sequence Stars

In main sequence stars, the presence and extent of the radiative zone exhibit a strong dependence on stellar mass, primarily due to differences in core temperatures, nuclear reaction rates, and energy transport mechanisms. Low-mass stars with masses below approximately 1.5 solar masses (M⊙) feature thick radiative cores where photon diffusion dominates energy transfer, extending outward to a significant fraction of the stellar radius before transitioning to a convective envelope. This structure arises because the lower central temperatures support the proton-proton chain reaction, producing insufficient energy flux to drive convection in the core. In contrast, high-mass stars exceeding 10 M⊙ possess large convective cores owing to the hotter conditions favoring the CNO cycle, which generates high luminosities that destabilize radiative transport; as a result, the radiative zone becomes thin or effectively absent, confined to a narrow envelope layer above the extended convective interior. During the hydrogen-burning of the , the radiative zone remains a prominent feature in stars where it exists, facilitating the outward transport of from to . As the star evolves, the growing helium core from ongoing contracts under gravitational influence, gradually shrinking the radiative zone's extent in low- and intermediate-mass stars by compressing the surrounding radiative layers. This evolutionary contraction reflects the increasing central density, altering the thermal profile without fundamentally disrupting the zone's radiative character until the post-main-sequence . Representative examples illustrate these variations, such as in F-type main sequence stars (masses around 1.0–1.6 M⊙), where the radiative zone typically spans about 70% of the stellar radius, bridging the developing convective core in higher-mass subtypes and the fully radiative interior in lower-mass ones. The zone's precise boundaries are further modulated by stellar metallicity, as higher metal abundances increase radiative opacity through enhanced electron scattering and bound-free absorption, steepening the temperature gradient and potentially reducing the radiative zone's size by promoting convective instability at its edges per the Schwarzschild criterion. Lower metallicity, conversely, lowers opacity, allowing a more extended radiative region. Observational confirmation of these radiative interiors in main sequence dwarfs comes from asteroseismology, particularly Kepler mission data, which detected solar-like oscillations in hundreds of such stars, enabling inferences about internal sound speed profiles and zone boundaries through mode frequency analysis. For instance, the Kepler Asteroseismic sample of main-sequence dwarfs revealed consistent radiative core signatures in low-mass targets via p-mode trapping. These measurements validate theoretical models and highlight the radiative zone's role in constraining stellar ages and compositions.

The Sun's Radiative Zone

The Sun's radiative zone is the thick intermediate layer of its interior, extending from approximately 0.25 solar radii (R⊙) outward from to about 0.7 R⊙ at the base of the convective zone, where the tachocline marks the outer boundary as a thin transition region of shear and compositional changes. This zone encompasses roughly 70% of the Sun's total mass despite occupying about half its radius, due to the high near the core. All energy generated by in the core—primarily through the proton-proton chain—must traverse this region via radiative diffusion, where photons undergo repeated absorption and re-emission by ionized plasma, carrying 100% of the core's output luminosity outward without significant convective mixing. Helioseismology, employing acoustic oscillations observed by instruments like the Michelson Doppler Imager (MDI) on the (), has provided detailed sound speed profiles that confirm the expected composition gradients in the radiative zone, revealing a gradual increase in mean molecular weight due to settling of heavier elements. These profiles align closely with standard solar models, showing relative sound speed variations of less than 1% in the zone's mid-depths, thus validating the radiative energy transport mechanism. Complementary solar neutrino flux measurements, such as those from the Borexino detector capturing pp-chain and CNO-cycle neutrinos, further corroborate the core's energy production rates and their efficient propagation through the radiative zone, with observed fluxes like 6.57 × 10^{10} cm^{-2} s^{-1} for pp neutrinos matching predictions within 5%. A key structural anomaly in the radiative zone arises from gravitational settling of helium produced in the core, which creates an outward-increasing mean molecular weight (μ) gradient; this compositional stratification provides a stabilizing buoyancy force that suppresses convection throughout the zone, maintaining its radiative character. Standard models incorporating gravitational, thermal, and radiative diffusion predict that this helium settling establishes the μ-gradient over a characteristic timescale of approximately 10^5 years, comparable to the photon's random-walk diffusion time across the zone (∼1.7 × 10^5 years), allowing the gradient to persist over the Sun's 4.6 billion-year age. Recent post-2020 refinements in opacity, inferred directly from helioseismic inversions rather than laboratory measurements, indicate that radiative opacities in the zone are about 10% higher than previous theoretical estimates near 2 × 10^6 , addressing discrepancies in sound speed profiles and improving model agreement with observations; these updates, bridging gaps in pre-2010s solar models, enhance predictions of energy transport without invoking altered compositions.

References

  1. [1]
    NASA: The Solar Interior
    The radiative zone is characterized by the method of energy transport - radiation. The energy generated in the core is carried by light (photons) that bounces ...
  2. [2]
    [PDF] Solar Interior
    The radiative zone is bounded by the point where temperature and density are low enough to permit atoms to hold some of their electrons. • Contains ~32% of the ...Missing: function | Show results with:function
  3. [3]
    The Structure and Composition of the Sun – MCC AST
    The light generated in the core is transported through the radiative zone very slowly, since the high density of matter in this region means a photon cannot ...
  4. [4]
    Layers of the Sun - NASA Science
    Sep 26, 2023 · The Radiative Zone – extends outward from the outer edge of the core to base of the convection zone, characterized by the method of energy ...Missing: function | Show results with:function
  5. [5]
    7.5: The Sun - Geosciences LibreTexts
    Oct 17, 2024 · The radiative zone, just outside the core, has a temperature of about 7 million degrees Celsius. The energy released in the core travels ...Missing: function | Show results with:function<|control11|><|separator|>
  6. [6]
    Heliophysics Big Idea 3.2 - - NASA Science
    Jan 29, 2025 · The Radiative Zone extends outward from the outer edge of the core to the base of the convection zone, characterized by the method of energy ...
  7. [7]
    The Sun
    Just outside the Inner Core of the sun at a distance approximately 0.25 to 0.7 solar radii lies the Radiative Zone. This zone radiates energy through the ...
  8. [8]
    Radiation Zone - Solar Physics at MSU
    A Slow Means of Energy Transport. Once energy is produced in the core of the Sun, it needs a way to travel from the solar center to the outer regions.
  9. [9]
    [PDF] EDDINGTON MODEL
    The model was proposed by Eddington in the 1920s, when very little was known about physical properties of matter in stellar interiors. The model assumes ...
  10. [10]
    [PDF] Chapter 5 - Energy transport in stellar interiors - Astrophysics
    The energy that a star radiates from its surface is generally replenished from sources or reservoirs located in its hot central region.
  11. [11]
    [PDF] Radiative energy transport in stellar interior - IUCAA
    There are three main processes responsible for energy transport: radiative diffusion, conduction and convection, all driven by the radial temperature gradient.
  12. [12]
    Lecture 3: Inside the Sun
    Jan 8, 2025 · It typically takes about 170,000 years for energy generated by fusion in the Sun's core to stagger its way to the Sun's surface.Missing: photon | Show results with:photon
  13. [13]
    [PDF] Equations of Stellar Structure
    Equations of stellar structure include: 1/ρ ∂P/∂r + GMr/r² + d²r/dt² = 0, ∂Mr/∂r = 4πr²ρ, ∂T/∂r = − 3κρLr/16πacT³r², ∂Lr/∂r = 4πr²ρ (ǫ − T dS dt ).
  14. [14]
    [astro-ph/0511008] Helioseismology, Neutrinos and Radiative Zones
    Oct 31, 2005 · In this review, we recall the clues for solar gravitational settling definitively established by the seismic determination of the photospheric ...
  15. [15]
    [PDF] 14-opacity.pdf - PHYS 633: Introduction to Stellar Astrophysics
    Such coherent electron scattering is known as Thomson scattering, and has a cross-section per electron of. σT = 8π. 3 a2. 0 ≈ 6.652 × 10−25 cm2. (9). 2. Page 3 ...
  16. [16]
    [PDF] Lecture 7 – Part 1 Sources of Stellar Opacity
    The Rosseland mean then behaves as ! = constant. Kramers Law. At temperatures lower than a few times 107 K, the important thermal processes of emission and ...
  17. [17]
    [PDF] RADIATIVE TRANSFER IN STELLAR ATMOSPHERES
    The monochromatic radiative flux at frequency ν gives the net rate of energy flow through a surface element. dE ν. ~ I ν cosθ dω → integrate over the whole ...
  18. [18]
    [PDF] Lecture 6 Radiation Transport
    The radiation field in the optically thick stellar interior is thus to high accuracy, given by the Planck function. Spherical symmetry and isotropy can be ...
  19. [19]
    Metallicity dependence of envelope inflation in massive stars
    The OPAL opacity tables from Iglesias et al. (1992) introduced an opacity peak at a temperature of T ~ 2 × 105K caused by bound-bound and bound-free ...
  20. [20]
    Updated Opal Opacities - NASA ADS
    1.-Ratio of updated OPAL and IRW Rosseland mean opacities as a function of temperature at constant log R values indicated in the Figure. · 2.-Ratio of Rosseland ...Missing: dependence | Show results with:dependence
  21. [21]
    (PDF) Updated Opal Opacities - ResearchGate
    Aug 9, 2025 · Although a specific metal composition is emphasized, opacity tables for different metal distributions can be made readily available. The updated ...Missing: variations | Show results with:variations
  22. [22]
    Eddington Solar Model - Richard Fitzpatrick
    The following ingenious model, due to Eddington (Eddington 1926), is appropriate to a star whose internal energy transport is dominated by radiation.
  23. [23]
    The internal constitution of the stars : Eddington, Arthur Stanley, Sir ...
    Jan 28, 2019 · The internal constitution of the stars. xvi, 407 p. : "An unabridged and unaltered republication of the last edition [1930]"
  24. [24]
    [PDF] ASTR3007/4007/6007, Class 7: Simple Stellar Models
    The Eddington model begins by considering the temperature structure of stars, and its implication for the importance of radiation pressure. The central insight ...
  25. [25]
    [PDF] PHYS-633: Introduction to Stellar Astrophysics
    5.2 Radiative Transfer Equation for a Planar Stellar Atmosphere . ... radiative diffusion equation. Sometimes it is instructive to write this as. F(r) ...Missing: zones | Show results with:zones
  26. [26]
    [PDF] Stellar Astrophysics Chapter 3: Heat Transfer
    Radiative Transfer. The Diffusion Equation. Radiative Opacity Sources. Radiative Opacity and Emissivity. Radiation Transfer in Stellar Atmosphere. Introduction.
  27. [27]
    MODULES FOR EXPERIMENTS IN STELLAR ASTROPHYSICS ...
    The numerical and computational methods employed allow MESAstar to consistently evolve stellar models through challenging phases, e.g., the He core flash in ...
  28. [28]
    Overview of kap module — MESA main documentation
    Radiative Opacities#. The radiative opacity is the Rosseland mean opacity. The opacity depends on the temperature, density, and composition. A set of opacity ...
  29. [29]
    Relativistic Radiative Transfer in Relativistic Plane–Parallel Flows
    Various methods and techniques of radiative transfer have been developed to treat and understand all luminous gaseous phenomenae in the universe. In the ...
  30. [30]
    Magnetic fields in stellar radiative interiors - ADS
    Internal magnetic fields, mostly ignored in stellar evolution models, are one of the most promising candidates to explain this observed strong angular momentum ...
  31. [31]
    5 5 Magnetic Fields In Stellar Interiors - Oxford Academic
    Abstract. Over the bulk of a stellar radiative zone, the magnetic force is expected to be a very small perturbation to the balance of pressure versus gravity ...Missing: transfer relativistic
  32. [32]
    [PDF] Stability of Hydrostatic Equilibrium The Schwarzschild Criterion
    This is known as the Schwarzschild 7criterion for convective stability. 7 After Martin Schwarzschild, one of the pioneers of modern stellar astrophysics ...
  33. [33]
    [PDF] Convection 2 - NMSU Astronomy
    Convection zones will be deeper for He-rich stars. • For metal-poor stars, the opacity is reduced and therefore the radiative temperature gradient gets.
  34. [34]
    [PDF] 05-stability.pdf - PHYS 633: Introduction to Stellar Astrophysics
    Equation (29) is the Ledoux stability criterion (LSC). If ∇µ = 0, then the ... so the star violates the SSC, but remains stable against convection because the.
  35. [35]
    [PDF] The Ledoux Criterion for Convection in a Star - Stony Brook Astronomy
    Aug 2, 2012 · Stability criteria will define whether small perturbation will grow or keep small.
  36. [36]
    SOHO-Gallery: Best Of SOHO - Solar and Heliospheric Observatory
    Next, there is a relatively stable radiative zone that conducts heat smoothly from the core to within 125,000 miles of the Sun's surface.<|control11|><|separator|>
  37. [37]
    [PDF] Lecture 7 Evolution of Massive Stars on the Main Sequence and ...
    So for the higher metallicity there is less hydrogen to burn. But there is also an opposing effect, namely mass loss. For higher metallicity the mass loss is ...
  38. [38]
    [PDF] Lecture 7 Evolution of Massive Stars on the Main Sequence and ...
    Despite their convective cores, the overall main sequence structure can be crudely represented as an n = 3 polytrope. This is especially true of the outer ...
  39. [39]
    The influence of radiative core growth on coronal X-ray emission ...
    These are late A to late F-type variable stars that undergo non-radial ... type main-sequence stars, as their radiative zones are beginning to grow.
  40. [40]
    the Kepler Asteroseismic LEGACY Sample. II. Radii, Masses, and ...
    Jan 30, 2017 · We have presented an asteroseismic analysis for the Kepler dwarfs LEGACY sample, comprising main-sequence stars observed for more than a ...
  41. [41]
    Asteroseismology of main-sequence F stars with Kepler
    Feb 13, 2019 · Main-sequence F-type stars exhibit short mode lifetimes relative to their oscillation frequency, resulting in overlapping radial and quadrupole modes.Missing: zone | Show results with:zone
  42. [42]
  43. [43]
    Helioseismic inference of the solar radiative opacity - Nature
    Jan 27, 2025 · We find that our seismic opacity is about 10% higher than theoretical values used in current solar models around 2 million degrees, but lower by 35%.Missing: accumulation | Show results with:accumulation
  44. [44]
    First Evidence of Solar Neutrinos by Direct Detection in Borexino
    Feb 2, 2012 · Assuming the p ⁢ e ⁢ p neutrino flux predicted by the standard solar model, we obtained a constraint on the CNO solar neutrino interaction rate ...