Fact-checked by Grok 2 weeks ago

Stellar dynamics

Stellar dynamics is the branch of that studies the collective motions and gravitational interactions of stars within bound systems, such as star clusters, galaxies, and galaxy clusters, treating these systems as collisionless ensembles governed primarily by Newtonian gravity. It focuses on how stars orbit within the mean created by the distribution of mass in these systems, rather than individual two-body encounters, which are rare due to the vast distances. The field employs statistical methods to describe the evolution of stellar systems over time, using tools from including the collisionless , which governs the phase-space distribution function of stars, and the Jeans equations, derived as moments of this equation to relate density, velocity dispersions, and . A key concept is the relaxation time, the timescale over which cumulative weak gravitational encounters randomize stellar orbits; for galaxies with ~10^{11} stars, this is on the order of hundreds of billions of years, far exceeding the age of the , allowing collisionless approximations to hold effectively. The provides a fundamental relation for steady-state systems, stating that twice the total equals the negative of the energy, enabling estimates of total mass from observed velocities. Stellar dynamics has broad applications in understanding galactic structure and evolution, including the inference of dark matter distributions from discrepancies between visible mass and dynamical tracers like rotation curves, as well as the dynamics of globular clusters and the effects of mergers in hierarchical galaxy formation. It draws analogies to plasma physics for handling long-range forces and to statistical mechanics for large-N systems, bridging celestial mechanics with modern cosmology. Advances in computational simulations have further refined models of relaxation processes and asymmetric drift, where older stellar populations lag behind circular rotation speeds due to velocity dispersion support.

Fundamentals

Definition and Scope

Stellar dynamics is the branch of that studies the collective motions of in self-gravitating systems, such as galaxies and star clusters, by modeling their gravitational interactions and treating as collisionless point-mass particles. This approach focuses on the statistical behavior of large ensembles of rather than individual trajectories, enabling the analysis of how gravitational forces shape the overall structure and evolution of these systems. The scope of stellar dynamics spans a broad range of scales, from compact star clusters with $10^4 to $10^6 members to vast galaxies containing up to $10^{12} , and even extends to assemblies like globular clusters and galactic nuclei. It distinctly differs from , which examines the internal physical processes within individual , by emphasizing the orbital dynamics and of across these systems. Central to this field is the approximation of the for large N, where direct computation of all pairwise interactions is infeasible, leading to statistical methods for describing system-wide properties. Key assumptions in stellar dynamics include treating stars as test particles that respond to the smooth mean produced by the entire system, rather than discrete encounters with individual stars. Most applications operate in the collisionless regime, where the relaxation time due to two-body encounters vastly exceeds the dynamical timescale, allowing stars to follow nearly collision-free orbits; this contrasts with collisional regimes in denser environments where cumulative small-angle deflections significantly alter motions over time. Fundamental terminology includes , a six-dimensional continuum of position and coordinates that fully specifies the state of stars in the system. The , often denoted f(\mathbf{r}, \mathbf{v}), quantifies the phase-space density by giving the number of stars per unit volume in position \mathbf{r} and \mathbf{v}. dispersion, a measure of the random velocity spread (e.g., \sigma_v), characterizes the kinetic support against and is integral to concepts like the .

Historical Development

The foundations of stellar dynamics were laid by Isaac Newton's formulation of the law of universal gravitation in 1687, which provided the mathematical framework for analyzing gravitational interactions among celestial bodies, including stars. further advanced these ideas through his comprehensive work on in the early 19th century, treating planetary and stellar systems as perturbed multi-body problems under Newtonian gravity. contributed significantly in the early 20th century by introducing the ellipsoidal hypothesis in 1907, proposing that stellar velocity distributions could be modeled as anisotropic ellipsoids to explain observed proper motions and deviations from isotropic random distributions. In 1919, James Jeans published his seminal work on and stellar dynamics, applying kinetic theory to assess the of gaseous spheres and derive criteria for the of clouds into or galaxies, marking a shift toward statistical treatments of large stellar ensembles. During the 1920s and 1930s, Jan Oort's studies on galactic rotation, particularly his 1927 analysis confirming in the based on stellar radial velocities, integrated observational data with dynamical models to reveal the Galaxy's rotational structure. These efforts established stellar dynamics as a field bridging theoretical and astronomical observations. In the mid-20th century, applied rigorously to stellar systems in his 1942 monograph, developing the collisionless and relaxation timescales for star clusters, which formalized the probabilistic description of stellar encounters. advanced the field in the 1960s through theoretical investigations into cluster evolution and interstellar dynamics, culminating in his comprehensive 1987 textbook that synthesized relaxation processes and evolutionary pathways for globular clusters. Influential figures like Chandrasekhar and Spitzer emphasized the role of long-range gravitational forces in shaping stellar distributions over cosmic timescales. Key milestones in the 1970s included the development of algorithms by Sverre Aarseth, whose direct summation methods and neighbor schemes enabled numerical modeling of few-body interactions in evolving clusters, paving the way for computational stellar dynamics. Donald Lynden-Bell's 1967 theory of violent relaxation provided a basis for understanding the rapid equilibration of collisionless systems post-perturbation, influencing models of formation. From the 1980s onward, stellar dynamics integrated advanced numerical techniques, such as particle-mesh codes for simulating large-scale gravitational potentials, with high-precision observations; the mission, launched in 2013, has supplied astrometric data for billions of stars, enabling dynamical mapping of the Milky Way's velocity field and refining models of orbital structures.

Connections to Other Disciplines

Fluid Dynamics Analogy

In stellar dynamics, the analogy treats as particles within a self-gravitating "fluid," where the random motions of , characterized by their , mimic the in a gas. This conceptual framework allows for the description of large-scale in stellar systems, such as galaxies or star clusters, by approximating the stellar as a continuous medium. The collisionless , equivalent to the in this context, functions as the foundational kinetic equation, analogous to the Navier-Stokes equations for but without collisional terms, enabling the modeling of pressure-like effects from . This analogy has historical roots in the work of , who in 1929 described galaxies as gaseous spheres composed of stars swarming like molecules in a gas, providing an early hydrodynamic of galactic and . Jeans' swarming hypothesis emphasized the equilibrium between gravitational attraction and the dispersive forces from stellar velocities, treating the galactic system as a self-gravitating fluid sphere to explain its form and dynamics. For dense stellar systems, hydrodynamic approximations are derived by taking successive moments of the in . The zeroth moment yields the governing mass density evolution, the first moment produces the Euler equation for momentum transport with velocity dispersion acting as , and the Poisson equation relates the to the density; higher moments introduce terms akin to and , though closure at low orders is often assumed for tractability. These moment equations provide a fluid-like description valid when the system is sufficiently relaxed, allowing conceptual parallels to compressible fluid flows under self-gravity. A key difference from traditional lies in the collisionless nature of stellar systems, where the between gravitational encounters vastly exceeds the system size, rendering viscous dissipation negligible compared to the short-range collisions in molecular fluids. Consequently, the omits transport coefficients like , leading to inviscid-like behavior with persistent anisotropies in the velocity dispersion tensor. This distinction is particularly relevant in numerical simulations, where the analogy informs (SPH) methods; here, discrete particles represent smoothed density and velocity fields, bridging collisionless stellar dynamics with gaseous components in hybrid models of galaxy formation. Stellar dynamics builds upon the classical , which describes the gravitational interaction between two bodies as the cornerstone for approximating stellar motions in more complex environments. In this two-body framework, the relative orbit follows an elliptical path with the primary mass at one , characterized by the semi-major axis a and e < 1 for bound systems. The radial distance is given by r = \frac{a(1 - e^2)}{1 + e \cos \phi}, where \phi is the true anomaly, derived from the inverse radial coordinate substitution u = 1/r in the orbital equation d^2u/d\phi^2 + u = GM/h^2, with h as the specific angular momentum. This solution ensures conservation of total energy E = -GM/(2a) < 0 and angular momentum, underpinning : orbits close as ellipses, equal areas swept in equal times, and periods scaling as T^2 \propto a^3. In stellar contexts, such orbits approximate binary star systems or restricted motions around a dominant mass, like the S2 star orbiting with a period of about 16 years. The three-body problem extends this to interactions among three gravitationally bound bodies, revealing inherent instabilities and chaotic dynamics that preclude analytical solutions in the general case. Henri Poincaré's seminal work in the 1890s demonstrated the non-integrability of the planar restricted circular three-body problem, showing exponential divergence of trajectories due to sensitivity to initial conditions, with no additional conserved quantities beyond energy, linear momentum, and angular momentum. This chaos manifests in close encounters within dense stellar clusters, where three-body interactions drive orbital perturbations, leading to ejections of one body at high velocities while the other two form a bound pair. In globular clusters, such encounters reverse core collapse by expelling stars, with error growth in simulations scaling as t_{cr}/\ln N for N particles, highlighting the role of three-body processes in cluster evolution. Perturbation theory bridges the three-body problem to the general N-body case in stellar dynamics, enabling analysis of secular evolution—long-term changes in orbital elements like eccentricity and inclination—while averaging over short-period oscillations. In hierarchical configurations, where an inner binary is perturbed by a distant third body, quadrupole-order approximations yield the Kozai-Lidov mechanism, oscillating the inner orbit's eccentricity and inclination; higher octupole terms introduce chaos, such as prograde-to-retrograde flips, especially prominent in stellar systems with comparable masses compared to planetary hierarchies. Unlike planetary contexts dominated by test-particle approximations and circular outer orbits, stellar N-body dynamics emphasizes resonant scattering and secular resonances, as simulated in barred galaxy halos or black hole environs, where perturbations drive slow evolution over gigayears. These principles find direct application in hierarchical triple systems, common among field stars where an inner binary is orbited by a tertiary, influencing binary evolution through that enhance tidal friction and potential mergers. In massive stellar triples, three-body dynamics coupled with stellar evolution can lead to compact object formation, with octupole effects destabilizing orbits and producing gravitational-wave sources. Similarly, three-body interactions in globular clusters eject stars via binary formation, imparting velocities of hundreds to over 1,000 km/s, particularly catalyzed by black holes in non-core-collapsed systems, accounting for up to 4% of escape rates over a Hubble time and explaining observed .

Gravitational Framework

Potential Field Concept

In stellar dynamics, the gravitational potential \Phi serves as the fundamental field that governs the motions of stars within self-gravitating systems. It is defined such that the acceleration \mathbf{g} of a test particle, such as a star, is given by \mathbf{g} = -\nabla \Phi, where the gradient points in the direction of the force. This potential arises from the collective mass distribution of the stellar system and is sourced by the mass density \rho, linking the gravitational field directly to the spatial arrangement of matter. A key simplification in modeling these systems is the mean field approximation, which posits that individual stars orbit within the smooth, time-averaged potential generated by the entire ensemble of other stars, while disregarding short-lived fluctuations from discrete two-body encounters. This approach is valid for large-N systems where the number of stars is sufficiently high that statistical averaging dominates over stochastic perturbations, enabling the treatment of stellar motions as predictable paths in a continuous field rather than chaotic interactions. Equipotential surfaces, defined by contours of constant \Phi, play a crucial role in delineating regions of gravitational influence. In binary star systems, for instance, the Roche lobe represents the equipotential surface that passes through the inner Lagrangian point L1, bounding the volume around each star where material is primarily attracted to that component; overflow beyond this surface can lead to mass transfer between companions. Potentials in stellar systems can exhibit different symmetries, such as axisymmetric forms common in disk galaxies, where the field is invariant under rotation around a principal axis, or triaxial potentials in elliptical galaxies, which lack such rotational symmetry and support more complex orbital families like box orbits. Among the essential properties of the potential field are energy conservation for stars in steady-state configurations and the foundational role it plays in the , which equates twice the total kinetic energy to the virial of the potential energy, providing a global constraint on system equilibrium (as explored further in subsequent sections). In steady states, the depth of the potential well determines the , beyond which stars can leave the system unbound.

Poisson Equation Basics

The Poisson equation forms the cornerstone of the gravitational framework in stellar dynamics, relating the gravitational potential \Phi to the mass density \rho of a stellar system through the differential equation \nabla^2 \Phi = 4\pi G \rho, where G is the gravitational constant. This equation governs the self-gravitational field in collisionless systems such as star clusters and galaxies, enabling the computation of potentials that dictate stellar orbits and system evolution. The derivation of the Poisson equation follows from Gauss's law for gravity, which states that the surface integral of the gravitational field \mathbf{g} over any closed surface equals -4\pi G times the enclosed mass M_\text{enc}: \oint \mathbf{g} \cdot d\mathbf{A} = -4\pi G M_\text{enc}. Defining \mathbf{g} = -\nabla \Phi, the gravitational field as the negative gradient of the potential, and applying the divergence theorem to convert the surface integral to a volume integral yields \nabla \cdot \mathbf{g} = -4\pi G \rho. Substituting the expression for \mathbf{g} then gives \nabla^2 \Phi = 4\pi G \rho. This form is particularly useful in stellar dynamics for modeling the mean gravitational field experienced by stars in a statistically averaged sense. For isolated stellar systems, such as or , appropriate boundary conditions are essential for solving the . The potential must satisfy \Phi \to 0 as r \to \infty, ensuring the system is unbound from external influences and the total mass is finite. At large distances from the system, the potential can be approximated using a : \Phi(\mathbf{r}) = -\frac{GM}{r} + \sum_{l=1}^\infty \frac{1}{r^{l+1}} \int \rho(\mathbf{r}') (r')^l P_l(\cos \alpha) \, d^3\mathbf{r}', where M is the total mass, and higher-order terms (dipole, quadrupole, etc.) describe deviations from spherical symmetry; for systems with vanishing center-of-mass motion, the dipole term is zero. This expansion facilitates efficient calculations of tidal interactions or flyby perturbations in multi-system dynamics. A key physical quantity derived from the potential is the escape speed v_\text{esc}, defined as the minimum velocity required for a star to reach infinity with zero kinetic energy, given by v_\text{esc} = \sqrt{-2\Phi}. This follows from conservation of mechanical energy: \frac{1}{2} v^2 + \Phi = 0 at infinity, where \Phi < 0 inside the system. In stellar dynamics, v_\text{esc} sets the boundary for bound versus unbound orbits, influencing evaporation rates in clusters. An illustrative analytical solution to the Poisson equation arises for a uniform-density sphere of radius R and constant density \rho, modeling idealized star clusters or planetary cores. Inside the sphere (r < R), the potential is quadratic: \Phi(r) = -\frac{2\pi G \rho}{3} (3R^2 - r^2). This satisfies \nabla^2 \Phi = 4\pi G \rho and matches the exterior point-mass potential at r = R, with \Phi(0) = -2\pi G \rho R^2 at the center. Such solutions provide benchmarks for understanding density-potential relations in spherically symmetric systems. For realistic stellar systems with non-uniform densities, analytical solutions are often infeasible, necessitating numerical methods to solve the Poisson equation self-consistently with the stellar distribution function. King models, introduced for globular clusters, achieve this by assuming a lowered isothermal distribution truncated at a finite energy, yielding solutions via iterative integration of the Poisson equation in dimensionless form. These models reproduce observed cluster profiles with a concentration parameter W_0, capturing tidal truncation effects essential for bounded systems.

Key Dynamical Processes

Gravitational Encounters

In stellar dynamics, gravitational encounters refer to the mutual perturbations between stars arising from their two-body gravitational interactions, which introduce stochastic variations in stellar velocities and orbits over time. These encounters are fundamental to the evolution of stellar systems, as they gradually redistribute energy and angular momentum among stars without direct physical collisions, given the vast interstellar distances. Encounters are classified as strong or weak based on the impact parameter b, which is the perpendicular distance between the asymptotic paths of the two interacting stars. Strong encounters occur when b is small, leading to close hyperbolic deflections with large changes in direction, often exceeding 90 degrees. In contrast, weak encounters involve larger b, resulting in distant passages with minimal individual deflections, typically much less than 1 radian, but their cumulative effect becomes significant over many interactions. The boundary between strong and weak is often defined by the impact parameter b_{90} \approx GM/v^2, where G is the gravitational constant, M is the mass of the perturbing star, and v is the relative velocity; encounters with b \lesssim b_{90} are strong, while those with b \gtrsim b_{90} are weak. The deflection angle \theta in a two-body hyperbolic encounter can be approximated for small angles as \theta \approx 2GM/(b v^2), where this relation holds in the limit of weak scattering and provides a 90° deflection when b \approx GM/v^2. For strong encounters, the cross-section \sigma is roughly \pi b_{90}^2, and the mean free path \lambda—the average distance a star travels before such an encounter—is given by \lambda = 1/(n \sigma), with n the stellar number density. Although strong encounters remain rare even in dense systems like globular clusters, weak encounters dominate the long-term dynamics due to their frequency. The characteristic timescale for velocity changes from encounters is the relaxation time t_{\rm rel} \propto v^3/(G^2 m^2 n \ln \Lambda), where m is the typical stellar mass and \ln \Lambda is the Coulomb logarithm accounting for the range of impact parameters, typically \ln \Lambda = \ln(b_{\rm max}/b_{\rm min}) with b_{\rm max} on the order of the system's size and b_{\rm min} \approx b_{90}. This logarithm, often \ln \Lambda \approx 5–$15 in realistic systems, arises because very distant (b \gg b_{90}) and very close (b \ll b_{90}) encounters contribute logarithmically to the total deflection. Through repeated weak encounters, stars experience diffusion in velocity space, where random small deflections accumulate to produce systematic changes in speed and direction, akin to Brownian motion but governed by gravity. This process, first rigorously described by Chandrasekhar, underpins the collisional nature of stellar systems despite their apparent collisionlessness on short timescales.

Dynamical Friction

Dynamical friction refers to the gravitational drag experienced by a massive object, known as the test particle or perturber, as it moves through a background medium of lighter stars or particles in a stellar system. This drag arises from repeated gravitational encounters that systematically decelerate the perturber, transferring its momentum to the surrounding field stars. The effect was first theoretically derived by Subrahmanyan Chandrasekhar in 1943, who showed that the friction stems from the statistical imbalance in scattering events, where field stars with velocities lower than the perturber are more likely to interact and be deflected forward, effectively slowing the perturber. The underlying mechanism involves gravitational focusing: as the massive object moves, it attracts nearby field stars, but due to the finite relative velocities, the deflections create an asymmetric density enhancement, or "wake," trailing behind the object. This wake generates a net gravitational pull opposite to the direction of motion, resulting in deceleration. In stellar systems, this process is diffusive and cumulative, relying on the collective effect of many weak, distant encounters rather than strong collisions. The standard quantitative description is given by Chandrasekhar's formula for the rate of change of the velocity \mathbf{v} of a perturber of mass M moving at speed v through a uniform background of density \rho and isotropic velocity dispersion \sigma: \frac{d\mathbf{v}}{dt} = -\frac{4\pi G^2 M \rho \ln\Lambda}{v^2} \left[ \erf(X) - \frac{2X}{\sqrt{\pi}} \exp(-X^2) \right] \hat{\mathbf{v}}, where G is the , \ln\Lambda is the approximating the ratio of the maximum to minimum impact parameters (typically \ln\Lambda \approx 5-10 in stellar systems), X = v / (\sqrt{2} \sigma), \erf(X) is the , and \hat{\mathbf{v}} is the unit vector in the direction of \mathbf{v}. This expression accounts for contributions only from field stars slower than the perturber, with the term in brackets approaching 1 for X \gg 1 (supersonic motion) and scaling as X^2 for X \ll 1 (subsonic motion). For equal-mass stars, the formula generalizes by replacing M with M (1 + m/M) where m is the field star mass, but it assumes M \gg m. The associated timescale for significant deceleration is approximately t_\mathrm{fric} \approx \frac{v^3}{4\pi G^2 M \rho \ln\Lambda}, derived by integrating the formula under the assumption of constant \rho and high-speed motion where the bracketed term is near unity. In virialized stellar systems, where the density \rho \sim \sigma^2 / (G R^2) from the virial theorem and the crossing time is t_\mathrm{cross} = R / v \sim R / \sigma, the ratio t_\mathrm{fric} / t_\mathrm{cross} \sim M_\mathrm{sys} / M (with M_\mathrm{sys} the system mass). Thus, t_\mathrm{fric} \gg t_\mathrm{cross} for light perturbers (M \ll M_\mathrm{sys}), making friction negligible on dynamical timescales, while heavy objects (e.g., black holes or globular clusters) sink rapidly. This timescale exceeds the two-body relaxation time by a factor of order M_\mathrm{sys} / m, highlighting friction's role in segregating massive objects. A more rigorous treatment frames dynamical friction within the Fokker-Planck equation, which describes the evolution of the distribution function f(\mathbf{v}) of stars due to small-angle gravitational deflections. The friction arises as the first-order moment of the velocity change diffusion coefficients: \left< \Delta v_i \right> = -\frac{4\pi G^2 m^2 \ln\Lambda}{v^3} \rho \int_0^v 4\pi v'^2 dv' f(v') \left( \erf(X) - \frac{2X}{\sqrt{\pi}} e^{-X^2} \right) v_i, where the angular brackets denote the average per unit time, and the diffusion tensor captures both frictional drag and velocity diffusion (). This formulation, originally developed by Chandrasekhar, allows for non-uniform backgrounds and anisotropic distributions, enabling numerical simulations of relaxation processes in clusters and galaxies. The coefficients ensure that the Maxwellian velocity distribution is preserved in equilibrium, with friction balancing diffusion for the perturber's .

Relaxation and Equilibrium

Relaxation Mechanisms

In stellar systems, two-body relaxation arises from the cumulative effect of numerous weak gravitational encounters between stars, which perturb their velocities in a manner analogous to a in space. This process leads to of stellar energies and velocities, gradually driving the system toward by redistributing among the stars. The stochastic nature of these encounters ensures that individual deflections are small, but over many interactions, they result in significant changes to the velocity distribution function. A key consequence of two-body relaxation is the establishment of the , whereby stars achieve equal average kinetic energies regardless of , implying that heavier stars attain lower velocities than lighter ones. This velocity- dependence promotes mass segregation, with more massive stars preferentially sinking toward the system's while lighter stars occupy the outer regions. In systems with a significant range, such as globular clusters, this segregation can lead to the formation of compact nuclei dominated by heavy stars. The timescale for local relaxation, t_{\rm rel}, characterizes the duration required for these processes to significantly alter stellar velocities and is approximated by t_{\rm rel} \approx \frac{N}{8 \ln N} t_{\rm cross}, where N is the number of stars and t_{\rm cross} is the crossing time across the system. Relaxation proceeds more rapidly in denser regions, such as cores, where encounter rates are higher due to elevated stellar densities. Dynamical friction represents the ordered, systematic component of relaxation, manifesting as a drag force on a test star due to the asymmetric wake of perturbed field stars, while the diffusive spreading constitutes the disordered fluctuations. This dichotomy draws an analogy to the , where random fluctuations enable dissipative evolution toward equilibrium in stellar systems. As a simplified conceptual overview, relaxation mechanisms can be understood through the in six-dimensional : in the collisionless limit, the governs incompressible flow of the without diffusion, whereas the collisional incorporates a relaxation term that drives the toward via two-body .

Loss Cone Dynamics

In stellar dynamics, the loss cone refers to the region in occupied by stars on orbits with sufficiently low such that their pericenters bring them within a critical radius r_{lc} of a central massive object, leading to capture, tidal disruption, or other loss processes. This critical angular momentum is given by L_{lc}^2 \approx G M_\bullet r_{lc}, where M_\bullet is the mass of the central object (e.g., a supermassive black hole) and r_{lc} is the loss radius, such as the tidal disruption radius for stars or the capture radius for compact objects. In velocity space at a given position \mathbf{r}, the loss cone manifests as a double with opening angle \theta_{lc} \approx \sqrt{r_{lc}/r}, encompassing orbits that intersect the loss region. The dynamics of the loss cone are governed by gravitational scattering events that populate this low-angular-momentum region from the surrounding , balanced by the depletion due to losses at the center. In , the of stars into the loss cone at E is F(E) \approx \bar{N}(E) D(E) / \ln(1/R_{lc}), where \bar{N}(E) is the average phase-space , D(E) is the coefficient from two-body relaxation, and R_{lc} = (r_{lc}/r_p)^2 with r_p the pericenter influence radius. This can be approximated as proportional to the of the f(\mathbf{v}) over the loss cone volume in velocity space, F \propto \int_{cone} f(\mathbf{v}) d^3v, reflecting the rate at which stars are scattered into losing orbits. The process relies on relaxation mechanisms, where and gradually alter stellar orbits, with the critical radius for loss cone entry at position r given by r_c = L_{crit}^2 / (G M r). Refilling of the loss cone occurs through diffusive scattering, with the characteristic diffusion time t_{diff} \approx t_{rel} \theta_{lc}^2, where t_{rel} is the two-body relaxation time, approximately t_{rel} \approx 0.34 \sigma^3 / (G^2 m \rho \ln \Lambda) for velocity dispersion \sigma, stellar mass m, density \rho, and Coulomb logarithm \ln \Lambda. Two regimes describe the loss cone state: the empty loss cone, where diffusion is slow compared to depletion (q \ll 1, with q = P(E) / [D(E) R_{lc}(E)] and orbital period P), leading to sparse population and flux F_{elc} \approx q \ln(1/R_{lc}) \bar{N} R_{lc} / P; and the full loss cone, where refilling matches losses (q \gg 1), resulting in a densely populated cone and flux F_{flc} \approx \bar{N} R_{lc}^2 / P. These regimes determine the overall consumption rate, with empty cones typical in dense galactic nuclei around supermassive black holes. Loss cone dynamics connect to gas accretion physics through analogous processes, where stellar inflows mimic Bondi-Hoyle accretion but for collisionless systems, driving the feeding of central s via captured or disrupted . The steady-state rates from loss cone flux provide the primary source of material for low-level accretion in quiescent galactic nuclei, scaling with black hole mass and stellar density profiles established by relaxation.

Statistical and Kinetic Approaches

Collisionless Boltzmann Equation

The collisionless Boltzmann equation (CBE) governs the evolution of the phase-space distribution function f(\mathbf{r}, \mathbf{v}, t) for a system of stars treated as collisionless test particles moving in a mean gravitational potential \Phi(\mathbf{r}). This equation arises from the conservation of phase-space density along stellar trajectories, as established in the foundational treatment of stellar dynamics. The CBE is expressed as \frac{df}{dt} = \frac{\partial f}{\partial t} + \mathbf{v} \cdot \nabla_{\mathbf{r}} f - \nabla \Phi \cdot \nabla_{\mathbf{v}} f = 0, where the total time reflects the convective derivatives in and space due to streaming and , respectively. This form embodies , which states that the remains constant along the unperturbed orbits of stars in , ensuring incompressibility of the flow for collisionless systems. In steady-state conditions, where \partial f / \partial t = 0, the CBE simplifies to \mathbf{v} \cdot \nabla_{\mathbf{r}} f = \nabla \Phi \cdot \nabla_{\mathbf{v}} f, balancing the of through configuration with changes induced by the force field in velocity space. Solutions to this equation are constrained by Jeans' theorem, which asserts that for a time-independent potential, the depends only on the isolating integrals of motion, such as the E = \frac{1}{2} v^2 + \Phi(\mathbf{r}) and the \mathbf{L}. In axisymmetric or spherical systems, typical forms include f(E, L_z) or f(E, L), enabling the construction of self-consistent models where the potential is generated by the stellar density via . For isotropic systems, thermal distributions often take a Maxwellian form in velocity space, f \propto \exp(-E / \sigma^2), where \sigma^2 is the velocity acting as an effective , analogous to kinetic in gases but adapted to self-gravitating stellar ensembles. Higher moments of the distribution, such as the velocity dispersion tensor \langle v^i v^j \rangle, quantify the pressure support and , with \langle \mathbf{v} \rangle representing the streaming . Taking the zeroth moment of the CBE in steady state yields the for the density \rho(\mathbf{r}) = \int f \, d^3\mathbf{v}, leading to : \nabla \cdot (\rho \langle \mathbf{v} \rangle) = -\rho \nabla \Phi. This relation underscores how the CBE provides a phase-space foundation for macroscopic fluid-like descriptions in stellar systems, distinct from moment-based approximations like the Jeans equations.

Jeans Equations and Virial Theorem

The equations arise from taking the first moments of the collisionless , providing a set of hydrodynamic-like equations that describe the and dispersion in a stellar system. In Cartesian coordinates, the general form of the first-moment equation is \frac{\partial (\rho \langle v_i \rangle)}{\partial t} + \frac{\partial (\rho \langle v_i v_j \rangle)}{\partial x_j} = -\rho \frac{\partial \Phi}{\partial x_i}, where \rho is the stellar , \langle v_i \rangle is the mean in the i-direction, \langle v_i v_j \rangle is the second-moment tensor (related to the velocity dispersion), and \Phi is the . This equation balances the rate of change of density with the of the momentum and the gravitational density. For steady-state systems with no bulk streaming (\langle \mathbf{v} \rangle = 0), it simplifies to \frac{\partial (\rho \sigma_{ij}^2)}{\partial x_j} = -\rho \frac{\partial \Phi}{\partial x_i}, where \sigma_{ij}^2 = \langle v_i v_j \rangle represents the components of the anisotropic tensor, allowing for in the . These equations, originally developed by James in the context of stellar motions, enable the modeling of density and profiles without solving the full . The second-moment Jeans equation, obtained by multiplying the collisionless by coordinates and integrating, relates the dispersion tensor to the and higher-order moments. In and spherical symmetry with , it reduces to a form where the dispersion \sigma_r^2 satisfies \frac{d (\rho \sigma_r^2)}{dr} + \rho \frac{2 \beta \sigma_r^2}{r} = -\rho \frac{d\Phi}{dr}, with \beta = 1 - (\sigma_\theta^2 + \sigma_\phi^2)/(2 \sigma_r^2) measuring anisotropy (\beta = 0 for isotropy). Here, \sigma^2 traces the depth of the potential well, as deeper potentials require higher dispersions to maintain equilibrium against gravity. This connection allows inference of the mass distribution from observed dispersions, assuming a form for the anisotropy. The virial theorem provides a global constraint on the energies of a self-gravitating stellar system in steady state, stating that twice the total kinetic energy K plus the total potential energy W vanishes: $2K + W = 0. For a system of N stars, K = \frac{1}{2} \sum m_i \langle \mathbf{v}_i^2 \rangle \approx \frac{3}{2} M \langle \sigma^2 \rangle in the isotropic case with total mass M, while W = -\sum_{i<j} G m_i m_j / r_{ij}. In the time-dependent form, applicable to evolving systems, it becomes $2K + W = \int \mathbf{F} \cdot \mathbf{r} \, dt, accounting for external forces or tidal effects. This theorem, extensively applied by Chandrasekhar to stellar dynamics, links observable velocity dispersions to the system's mass and size, facilitating mass estimates for clusters and galaxies. A worked example illustrates the virial theorem for a uniform-density sphere of radius R and density \rho, where the potential energy is W = -\frac{3}{5} \frac{G M^2}{R} with M = \frac{4\pi}{3} \rho R^3. Applying $2K + W = 0 yields the average one-dimensional velocity dispersion \langle \sigma^2 \rangle = \frac{4\pi G \rho R^2}{15}, or equivalently \langle \sigma^2 \rangle = \frac{G M}{5 R}. This result assumes isotropy and highlights how phase-space density is conserved along stellar orbits (via ), ensuring the distribution function remains constant in the steady state despite varying local dispersions. Such global relations underpin applications like thick disk modeling, where virial equilibrium constrains vertical dispersions.

Relativistic and Advanced Topics

Relativistic Approximations

In stellar dynamics near compact objects, special relativistic effects become significant when velocities approach the speed of light, necessitating a reformulation of the phase space to ensure Lorentz invariance. The traditional Newtonian phase space volume is not preserved under Lorentz transformations, but in special relativity, the invariant phase space element is given by d^3\mathbf{x} \, d^3\mathbf{p} / E, where E is the particle energy, maintaining conservation across inertial frames. This invariance underpins the relativistic distribution functions used to describe stellar ensembles, such as the , which generalizes the to relativistic speeds: f(\mathbf{p}) \propto \exp\left( -\frac{E}{kT} \right), where E = \gamma m c^2 and \gamma = (1 - v^2/c^2)^{-1/2}. In dense stellar systems like , these distributions account for high-velocity tails that influence relaxation processes and dynamical friction, altering encounter rates compared to non-relativistic approximations. General relativistic corrections are essential for stellar dynamics in strong gravitational fields, such as around black holes, where post-Newtonian expansions provide a perturbative bridge from Newtonian gravity. The post-Newtonian formalism expands the metric and equations of motion in powers of (v/c)^2 and (GM/rc^2), incorporating terms up to 3.5PN for accurate N-body simulations of star clusters harboring intermediate-mass black holes. These expansions modify orbital energies and pericenter advances, enhancing merger rates in relativistic regimes by factors of up to 10 in dense environments. For non-rotating black holes, the Schwarzschild metric governs stellar orbits, describing geodesics in the vacuum spacetime outside a spherically symmetric mass M: ds^2 = -\left(1 - \frac{2GM}{c^2 r}\right) c^2 dt^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} dr^2 + r^2 d\Omega^2, where the Schwarzschild radius R_s = 2GM/c^2 defines the event horizon. Stellar trajectories in this metric exhibit precession and plunging orbits, fundamentally altering relaxation timescales near the black hole. Key general relativistic effects include the innermost stable circular orbit (ISCO), beyond which stars cannot maintain stable circular motion, located at r_{\rm ISCO} = 3 R_s for prograde orbits in the . This radius sets a boundary for accretion disks and influences the dynamics of stars scattered into close orbits, leading to tidal disruptions or captures. , or the Lense-Thirring effect, arises in rotating (Kerr) black holes, where the spacetime vorticity induces nodal precession of stellar orbits at rates \Omega_{\rm LT} \approx (G J / c^2 r^3), with J the black hole angular momentum; this effect warps the loss cone and enhances binary formation in galactic nuclei. The Eddington limit further constrains relativistic stellar dynamics by balancing gravitational infall with radiation pressure, yielding a maximum luminosity L_{\rm Edd} = \frac{4\pi G M c}{\kappa}, where \kappa is the opacity (typically Thomson scattering \kappa_{\rm es} = 0.2(1+X) cm² g⁻¹ for hydrogen mass fraction X). In super-Eddington accretion regimes, exceeding L_{\rm Edd} drives outflows and photon bubbling, regulating black hole growth and dispersing stellar clusters; simulations show growth bursts up to 10 times \dot{M}_{\rm Edd} but limited by feedback to mild excesses of 2-3 times in quasar seeds. These dynamics imply shorter accretion phases and overmassive black holes relative to host galaxies in early universe models.

Tidal Disruption and Sphere of Influence

In stellar dynamics, the tidal disruption radius defines the critical distance at which a massive perturber begins to significantly deform a star due to differential gravitational forces. For a star of mass m and radius R encountering a perturber of mass M \gg m, this radius is approximated by the Roche lobe size, beyond which the star remains intact and within which tidal forces exceed the star's self-gravity, leading to disruption if the periastron is smaller than this value. The formula is r_t = R \left( \frac{M}{m} \right)^{1/3}, derived under the assumption of a fluid body filling its Roche lobe; for main-sequence stars, disruption occurs when the stellar envelope overflows this limit. In the context of supermassive black holes, where M \sim 10^6 - 10^8 \, M_\odot, full disruptions are common for solar-type stars approaching within r_t, but partial disruptions arise if the periastron lies between the stellar radius R and r_t, stripping only the outer layers while the core survives. The sphere of influence delineates the region around a massive perturber where its gravity dominates the motion of nearby objects, often approximated using the Hill radius in the restricted three-body problem. For a binary system of total mass m and semimajor axis a interacting with a perturber of mass M \gg m, the characteristic radius of the bound orbit following partial or full disruption is r_h \approx a (M/m)^{2/3}, marking the scale over which the perturber controls the dynamics of the debris or surviving components. This radius governs the dominance of perturbations, with objects inside experiencing strong tidal influences while those outside follow orbits primarily dictated by the larger system's potential. Post-disruption dynamics involve the stretching of the star into a thin stream due to the radial range of orbital energies, with debris spreading along highly eccentric orbits around the perturber. The bound material falls back on a timescale t \sim (r_t^3 / G M)^{1/2}, forming an accretion disk that produces characteristic light curve flares peaking near the and decaying as t^{-5/3} from viscous spreading. These streams can remain coherent over multiple orbits before shocking and circularizing, influencing subsequent accretion and jet formation in black hole environments. Applications of these concepts include the Hills radius for globular clusters, where r_h sets the boundary for bound stars, with velocities exceeding the local escape speed leading to stripping and cluster evaporation in the galactic tidal field. In dense environments like galactic nuclei, binary disruptions within the perturber's sphere of influence eject hypervelocity stars while injecting tightly bound companions, enhancing loss cone populations and contributing to the dynamical evolution of nuclear star clusters.

Applications and Examples

Galactic and Cluster Dynamics

Stellar dynamics in galaxies manifests prominently through rotation curves, which plot the orbital velocities of stars and gas as a function of radial distance from the galactic center. Observations of spiral galaxies, including the , reveal that these curves remain approximately flat at large radii, indicating that orbital speeds do not decline as expected under Keplerian motion dominated by visible matter. This flatness suggests the presence of extended dark matter halos providing additional gravitational potential to sustain the observed velocities, with typical halo masses estimated at around 10^{12} solar masses for the as of the early 2020s. The seminal work by and colleagues in the 1970s and 1980s on external galaxies confirmed this pattern across multiple systems, establishing dark matter as a key component of galactic structure. Differential rotation within the galactic disk is quantified by the , A and B, which describe the local shear and vorticity from velocity gradients. The constant A measures the differential rotation rate, typically around 15 km/s/kpc in the solar neighborhood, while B reflects the rotation direction, with values near -12 km/s/kpc indicating clockwise rotation when viewed from the north galactic pole. These parameters, derived from proper motions and radial velocities of stars, provide constraints on the local mass distribution and the enclosed mass within the solar radius, approximately 10^{11} solar masses. In star clusters, such as globular clusters, dynamics are governed by two-body relaxation processes that lead to the formation of a dense core surrounded by a diffuse halo, known as the core-halo structure. Relaxation drives energy equipartition, causing massive stars to segregate toward the core while lighter stars populate the halo, with core radii often spanning 0.1-1 parsecs in observed systems. Over time, this relaxation also induces evaporation, where low-energy stars escape the cluster's tidal boundary due to cumulative perturbations, leading to gradual cluster expansion and mass loss at rates of about 10^{-5} to 10^{-4} of the total mass per relaxation time. Comprehensive N-body simulations have shown that evaporation is limited by the relaxation timescale, which for typical globular clusters is 10^8 to 10^9 years, resulting in half-mass relaxation times that dictate long-term evolution. Observational constraints on both galactic and cluster dynamics have advanced significantly with astrometric and spectroscopic data. The Gaia mission's proper motions enable precise mapping of 3D velocities for millions of stars, yielding velocity dispersions such as in the Hyades open cluster at around 0.5 km/s (1D) and facilitating dynamical mass estimates via the virial theorem, which relates the total mass M to the velocity dispersion σ and radius R as M ≈ 3σ²R/G for isotropic systems. Spectroscopic surveys, such as those from SDSS or LAMOST, complement this by measuring line-of-sight velocities, allowing decomposition of dispersions into radial and tangential components for improved mass modeling in dwarf galaxies and clusters. The role of dark matter in these systems is further illuminated by discrepancies between predicted and observed density profiles. The Navarro-Frenk-White (NFW) profile, derived from cold dark matter simulations, posits a cuspy central density ρ ∝ r^{-1}, yet observations of dwarf galaxies and clusters often reveal cored profiles with flatter inner slopes around ρ ∝ r^{0} to r^{-0.5}, challenging pure collisionless models. This cusp-core problem may arise from baryonic feedback or alternative dark matter candidates, as evidenced by kinematic data from resolved dwarfs like . Additionally, dynamical friction from dark matter halos decelerates infalling satellites, with Chandrasekhar's formula predicting orbital decay times proportional to v^3 / (ρ ln Λ), where v is velocity and ρ is local density, explaining the observed truncation of satellite streams in the .

Thick Disk Models

Self-consistent models for the thick disk of galaxies often employ the isothermal sheet approximation, which assumes a vertically isothermal population with constant velocity dispersion σ_z. In this framework, the vertical structure is described by the potential \Phi(z) = 2 \sigma_z^2 \ln \left[ \cosh \left( \frac{z}{z_0} \right) \right] + C, where z_0 = \sigma_z^2 / (\pi G \Sigma) is the scale length, Σ is the surface density, G is the , and C is a constant. This potential captures the flattened geometry of the thick disk and facilitates the calculation of vertical frequencies, with the vertical oscillation frequency given by κ_z ≈ √(4π G ρ_0) near the midplane, where ρ_0 is the midplane density. The corresponding density profile is obtained from Poisson's equation, yielding \rho(z) = \frac{\Sigma}{2 z_0} \sech^2 \left( \frac{z}{z_0} \right), which ensures self-consistency between the potential and mass distribution. The surface density Σ integrates the vertical profile to the total column, and for a radially exponential disk with scale length R_d, the total disk mass is M = 2π Σ R_d^2, providing a measure of the overall mass scale in the model. This sech² form is particularly suitable for thick disks, as it naturally arises from the balance of self-gravity and pressure support in a collisionless system, distinguishing it from thinner components with smaller scale heights. The scale height h_z ≈ z_0 / 2. Applying the vertical Jeans equation, d(ν σ_z²)/dz = -ν dΦ/dz, where ν(z) is the stellar number density proportional to ρ(z), confirms the isothermal assumption with constant σ_z, as the equation integrates to ν ∝ exp(-Φ / σ_z²), reproducing the sech² profile self-consistently. In the epicycle approximation, radial and vertical motions are largely decoupled for axisymmetric potentials, but coupling arises through non-separable terms in Φ(R, z), allowing stars to execute coupled oscillations that maintain the disk's thickness against perturbations. The vertical frequency κ_z relates to the epicycle frequency κ_R via the disk's rotation curve, with κ_z ≈ √(4π G ρ) providing a local estimate of stability. As a worked example, consider the gravity vector in the disk: g = -∇Φ, so in the vertical direction g_z = -∂Φ/∂z = - (2 σ_z^2 / z_0) tanh(z / z_0) = -2π G Σ tanh(z / z_0) (for z > 0), which for small z approximates the restoring force. For steady-state conditions, the ∇ · (ρ v) = 0 holds, ensuring that the vertical flux balances under the assumed equilibrium distribution function, consistent with the collisionless solutions for the isothermal sheet. This framework, while idealized, underpins numerical models of evolution and kinematic studies in galaxies like the .

Uniform Sphere Examples

In a uniform density sphere of constant density ρ and radius R, the gravitational potential inside the sphere (r < R) takes the form Φ(r) ∝ r², leading to a linear restoring force and harmonic oscillator motion for bound particles. Specifically, the potential is given by \Phi(r) = -\frac{2\pi G \rho}{3} (3R^2 - r^2), resulting in orbits that are ellipses centered on the origin with angular frequency ω = √(4π G ρ / 3). This harmonic behavior arises from the equivalence of the interior field to that of a point mass at the center for the enclosed mass, as per the shell theorem, and simplifies the analysis of particle trajectories in idealized spherical systems. Applying the Jeans theorem to the collisionless Boltzmann equation (CBE) in this spherical potential allows construction of equilibrium distribution functions (DFs) that depend solely on the binding energy E for isotropic systems. For an isotropic DF of the form f(E) = constant within the sphere, the phase-space density is uniform, satisfying the steady-state CBE since all orbits are ergodic and fill the available volume uniformly. The second moment of the DF yields the velocity dispersion, with the total velocity dispersion <v²> integrated over velocity space giving <v²> = 3 σ² for isotropic conditions, where σ is the one-dimensional dispersion. The provides a for the uniform sphere, stating 2K + W = 0, where K is the total and W is the energy. For a uniform sphere, W = - (3/5) G M² / R, and assuming isotropic velocities, K = (3/2) M σ², leading to σ² = (1/5) G M / R as the average one-dimensional velocity dispersion. The radial Jeans equation further elucidates local structure: \frac{d(\rho \sigma_r^2)}{dr} + \frac{2 \beta \rho \sigma_r^2}{r} = -\rho \frac{d\Phi}{dr}, where σ_r is the dispersion, β = 1 - σ_θ² / σ_r² is the parameter (β = 0 for ), and dΦ/dr = (4π ρ r)/3 inside the sphere. For isotropic (β = 0) and uniform ρ, this implies a linearly increasing σ_r² with r to balance , consistent with the potential. Moments of the DF offer insights into average properties, such as the probability-weighted mean = ∫ f(E) E d³v / ∫ f(E) d³v, which for an isotropic ergodic DF in the uniform sphere evaluates to a negative value reflecting the , with the ensuring mass conservation. In a uniform cluster model, the maximum phase-space density f_max equals the constant DF value f, as particles uniformly occupy the accessible due to the orbits, providing a benchmark for collisionless systems. Relativistic particles, such as , in a galactic potential can mimic aspects of dynamics when modeled via the CBE, particularly in clusters where their distribution follows the potential wells despite high velocities. For instance, in or A119 clusters, self-consistent solutions for a relativistic yield profiles that contribute to the total , with velocities approaching the speed allowing partial clustering akin to collisionless components. This illustrates how relativistic effects modify Jeans-like equations without altering the spherical symmetry assumptions.

References

  1. [1]
    [PDF] Introduction - Princeton University
    Stellar dynamics is directly related to at least three other areas of theoretical physics. Superficially, it is closest to celestial mechanics, the theory of ...
  2. [2]
    [PDF] Class #4: Elements of Stellar Dynamics
    The basic idea of stellar dynamics is to determine how systems of stars (a.k.a. galaxies) evolve under the effects of their own self-gravity (and, perhaps ...
  3. [3]
  4. [4]
    [PDF] Stellar Dynamics - UMD Astronomy
    Stellar dynamics describes systems of many point mass particles whose mutual gravitational interactions determine their orbits. These particles are usually.
  5. [5]
    Problems of cosmogony and stellar dynamics - Internet Archive
    Mar 20, 2007 · Problems of cosmogony and stellar dynamics. by: Jeans, James Hopwood, Sir, 1877-1946. Publication date: 1919. Topics: Cosmogony, Nebular ...
  6. [6]
    Observational evidence confirming Lindblad's hypothesis of a ...
    Observational evidence confirming Lindblad's hypothesis of a rotation of the galactic system. Oort, J. H.. Abstract. Publication: Bulletin of the Astronomical ...
  7. [7]
    Principles of Stellar Dynamics - Subrahmanyan Chandrasekhar
    Bibliographic information ; Title, Principles of Stellar Dynamics ; Author, Subrahmanyan Chandrasekhar ; Publisher, Dover, 1942 ; Length, 313 pages.
  8. [8]
    (PDF) Continuum Equations for Stellar Dynamics - ResearchGate
    PDF | The description of a stellar system as a continuous fluid represents a convenient first approximation to stellar dynamics, and its derivation from.
  9. [9]
    [PDF] Stellar Dynamics and Structure of Galaxies - Orbits in a given potential
    Unlike two-body, three-body problem is not possible to solve analytically. However, approximate descriptions can be derived for particular situations. For ...
  10. [10]
    [PDF] The simplest complexity: The story of the three-body problem - arXiv
    Oct 21, 2025 · The second phase of the story started with Poincaré and the first signs of chaos. During it, the three-body problem motivated Poincaré in ...
  11. [11]
    [PDF] chaos in the n-body problem of stellar dynamics
    When the density in the core has become high enough, three-body interactions become sufficiently common to eject substantial numbers of stars from the cluster.
  12. [12]
    [1107.2414] Secular Dynamics in Hierarchical Three-Body Systems
    Jul 12, 2011 · Here we revisit the secular dynamics of hierarchical triple systems. We derive the secular evolution equations to octupole order in Hamiltonian perturbation ...
  13. [13]
    The bar–halo interaction – II. Secular evolution and the religion of N ...
    This paper explores resonance-driven secular evolution between a bar and dark matter halo using N-body simulations. We make direct comparisons to our analytic ...
  14. [14]
    SECULAR STELLAR DYNAMICS NEAR A MASSIVE BLACK HOLE
    We use the ARMA model, calibrated with extensive N-body simulations, to analyze the long-term evolution of stellar systems around MBHs with Monte Carlo ...
  15. [15]
    [1612.06172] The evolution of hierarchical triple star-systems - arXiv
    Dec 19, 2016 · Abstract:Field stars are frequently formed in pairs, and many of these binaries are part of triples or even higher-order systems.
  16. [16]
    Stellar Escape from Globular Clusters. I. Escape Mechanisms and ...
    Apr 5, 2023 · When a three-body interaction involving a hard binary features a high mass ratio, ejection speeds of several hundred kilometers per second ...<|control11|><|separator|>
  17. [17]
    Ejection of pulsars and binaries to the outskirts of globular clusters
    THREE-BODY interactions can eject stars, singly or in binaries, from the core of a globular cluster to its outskirts, whither dynamical friction may take more ...
  18. [18]
    [PDF] Chapter 2 Stellar Dynamics in Galaxies
    Instead stellar dynamics is mostly governed by the interaction of individual stars with the mean gravitational field of all the other stars combined. This has ...
  19. [19]
  20. [20]
    Equipotential Surfaces and Lagrangian Points in Nonsynchronous ...
    We investigate the existence and properties of equipotential surfaces and Lagrangian points in nonsynchronous, eccentric binary star and planetary systems
  21. [21]
    [PDF] Part A : Gravity
    Introducing the gravitational potential Φ such that g = ∇Φ, Gauss's Law can be written as. ∇2Φ=4πGρ,. (8) a form known as Poisson's equation. In class, we ...
  22. [22]
    [PDF] Dynamics in a Gravitational Field
    Implies that the gravitational potential Φ(x) inside the sphere is constant, since F(x) = −∇Φ(x). The easiest place to evaluate it is at the centre, which is ...
  23. [23]
    [PDF] Lecture 2: Potential Theory I
    This is Gauss's theorem: the integral of the normal compo- nent of ∇Φ over any closed surface equals 4πG times the mass contained within that surface. 9. Page ...
  24. [24]
    [PDF] Lecture XII: External field of an isolated system - Caltech (Tapir)
    Nov 16, 2011 · MULTIPOLE EXPANSION. Far from a source, it is common to do a multipole expansion: essentially a power-series expansion of Eq. (20) in powers ...
  25. [25]
    [PDF] Derivation of potential from density distribution - People at the Institute
    To determine the force due to a given density distribution ρ(r0) we split it into many point masses of size dm0 = ρ(r0)d3r0 at r0.
  26. [26]
    3.2. Escape velocity - Dynamics and Astrophysics of Galaxies
    In most potentials, bodies that move fast enough can “escape” the gravitational field of a mass distribution. This means that such bodies do not orbit the ...
  27. [27]
    Cluster models and the escape of stars
    The transformation introduces U and r as parameters, so that Poisson's equation can then be written d2U zdU + - = 4rrGp(U, r). (5) When this differential ...
  28. [28]
    [PDF] Stellar Dynamics and Structure of Galaxies - Star Clusters
    Globular clusters are approximately spherical, compact agglomerations of stars swarming around galaxies. They are as old as the galaxies themselves and provide.Missing: scope | Show results with:scope
  29. [29]
    None
    Summary of each segment:
  30. [30]
  31. [31]
    Dynamical friction and galaxy merging time-scales - Oxford Academic
    In this paper, we study the merging time-scales of extended dark matter haloes using N-body simulations. We compare these results to standard estimates.INTRODUCTION · SATELLITE DECAY FROM... · COMPUTING MERGER TIME...
  32. [32]
    The Time of Relaxation of Stellar Systems. - Astrophysics Data System
    SAO/NASA ADS Astronomy Abstract Service. THE TIME OF RELAXATION OF STELLAR SYSTEMS. I S. CHANDRASEKHAR ABSTRACT In this paper the problem of the time of ...
  33. [33]
    Equipartition and the Formation of Compact Nuclei in Spherical ...
    Printed in U.S.A. EQUIPARTITION AND THE FORMATION OF COMPACT NUCLEI IN SPHERICAL STELLAR SYSTEMS LYMAN SPITZER, JR. Princeton University Observatory ...Missing: instability paper
  34. [34]
    [PDF] 1943ApJ 97 . . 255C DYNAMICAL FRICTION I. GENERAL ...
    equation. dW= 0 V2 w dt * u. ' (2). 1 S. Chandrasekhar, Ap. /., 94, 511, 1941. 2 For a general discussion of this and related theories see a forthcoming ...
  35. [35]
    Linear response, dynamical friction and the fluctuation dissipation ...
    Abstract. We apply linear response theory to a general, inhomogeneous, stationary stellar system, with particular emphasis on dissipative processes analogo.Missing: ordered | Show results with:ordered
  36. [36]
    [1307.3268] Loss Cone Dynamics - arXiv
    Jul 11, 2013 · This article reviews the dynamical processes by which stars or stellar remnants are placed onto loss-cone orbits and the implications for feeding rates.Missing: seminal | Show results with:seminal
  37. [37]
    None
    Error: Could not load webpage.<|control11|><|separator|>
  38. [38]
    Problems of Cosmology and Stellar Dynamics
    Problems of Cosmogony and Stellar Dynamics is a theoretical prelude to Jeans's later and more mature work on the subject, Astronomy and Cosmogony.
  39. [39]
    Some Elementary Applications of the Virial Theorem to Stellar ...
    The dynamical evolution of spherical and spheroidal systems of mass points is examined with the aid of the scalar and the tensor forms of the virial theorem.
  40. [40]
    Relativistic dynamical friction in stellar systems
    In a system defined by the phase-space distribution function F = nf(uF) ... (18), is a Lorentz-invariant quantity3, where. V ≡ V 1 − uT · uF/c2 = V 1 ...
  41. [41]
    Relativistic Stellar Dynamics - Astrophysics Data System
    The collision time is a measure of the mean time between collisions for a given star traveling through the system, whereas the time of relaxation is a measure ...
  42. [42]
    Stellar Dynamics around Black Holes - IOP Science
    Collisional Stellar Dynamics around Massive Black Holes in Active Galactic Nuclei ... (Lynden-Bell 1969; cf. Hoyle et al. 1964). Hubble Space Telescope ...
  43. [43]
    Post-Newtonian N-body simulations - Oxford Academic
    Abstract. We report on the first fully consistent conventional cluster simulation which includes terms up to the third-order post-Newtonian approximation.Post-Newtonian N-Body... · 2 Wheel-Spoke Implementation · 5 Post-Newtonian Aspects<|separator|>
  44. [44]
    Influence of the Lense–Thirring effect on stellar orbits around a ...
    Rotating supermassive black holes drive Lense–Thirring (LT) effects, which induce orbit nodal and eccentricity vector precession in their vicinity.
  45. [45]
    How the super-Eddington regime regulates black hole growth in ...
    The super-Eddington accretion events remain relatively mild with typical rates of about 2−3 times the Eddington limit, because of the efficient regulation by ...Missing: dynamical implications
  46. [46]
    effects of super-Eddington accretion and feedback on the growth of ...
    We find that allowing super-Eddington accretion leads to drastic differences, with the BH going through an intense but short super-Eddington growth burst.Missing: implications | Show results with:implications
  47. [47]
    Possible power source of Seyfert galaxies and QSOs - Nature
    Mar 27, 1975 · The possible presence of massive black holes in the nuclei of galaxies has been suggested many times. In addition, there is considerable observational evidence ...
  48. [48]
    Tidal disruption of stars by black holes of 106–108 solar masses in ...
    Jun 9, 1988 · Tidal disruption of stars by black holes of 106–108 solar masses in nearby galaxies. Martin J. Rees. Nature volume 333, pages 523–528 (1988)Cite ...
  49. [49]
    Hyper-velocity and tidal stars from binaries disrupted by a ... - Nature
    Feb 25, 1988 · Hyper-velocity and tidal stars from binaries disrupted by a massive Galactic black hole. J. G. Hills. Nature volume 331, pages 687–689 (1988) ...
  50. [50]
    [PDF] arXiv:1712.03242v1 [astro-ph.GA] 8 Dec 2017
    Dec 8, 2017 · STEEPER STELLAR CUSPS IN GALACTIC CENTERS FROM BINARY DISRUPTION. Giacomo Fragione1 and Re'em Sari1. 1Racah Institute for Physics, The Hebrew ...
  51. [51]
    Rotation Curve of the Milky Way and the Dark Matter Density - MDPI
    We review the current status of the study of rotation curve (RC) of the Milky Way, and present a unified RC from the Galactic Center to the galacto-centric ...
  52. [52]
  53. [53]
    [PDF] Galactic Rotation- Oort Constants
    – flat rotation curves. – surface brightness profiles close to ... – dark matter halo ρ(r)=ρ(0)/[1+(r/a)2]. • See B&T sec 2.7 for more complex ...
  54. [54]
    [PDF] Star cluster dynamics - arXiv
    Nov 4, 2009 · In this paper, we will review the main physical ingredients driving the early and long-term evolution of clusters and describe how cluster ...
  55. [55]
    Relaxation-limited evaporation of globular clusters | Request PDF
    Evaporative evolution of stellar clusters is shown to be relaxation limited when the number of stars satisfies N > > N c N>>N_c , where N c ≃ 1600 ...
  56. [56]
    Proper motions and dynamics of the Milky Way globular cluster ...
    We use Gaia Data Release 2 to determine the mean proper motions for 150 Milky Way globular clusters (almost the entire known population)
  57. [57]
    [PDF] Accurate mass estimates from the proper motions of dispersion ...
    Mar 14, 2020 · We derive a new mass estimator that relies on internal proper motion measurements of dispersion-supported stellar systems, ...Missing: spectroscopy | Show results with:spectroscopy
  58. [58]
    Review of Solutions to the Cusp-Core Problem of the ΛCDM Model
    This review aims at proposing to the field an overview of the Cusp-core problem, including a discussion of its advocated solutions.
  59. [59]
  60. [60]
    [PDF] GRAVITATIONAL POTENTIAL
    ❖Homogeneous (uniform) sphere const for r < a ρ(r) = 0 for r > a outside ... harmonic oscillator. Because F r. = v2/r. Page 12. ❖Logarithmic potential. We ...
  61. [61]
    [PDF] Jeans Theorem & Applications
    Jeans theorem is useful in constructing equilibrium models of stellar systems. 12.2 Isotropic Models. The simplest use of the Jeans Theorem is the construction ...Missing: uniform dynamics
  62. [62]
    a comparison of mass modelling methods for spherical systems
    Nov 25, 2020 · ABSTRACT. We apply four different mass modelling methods to a suite of publicly available mock data for spherical stellar systems.
  63. [63]
    Neutrino dark matter in clusters of galaxies - IOP Science
    Abstract. We present a model calculation for the radial matter density and mass distribution in two clusters of galaxies (Coma and A119) including cold dark ...
  64. [64]
    Neutrinos as cluster dark matter - Oxford Academic
    It would seem possible that individual galaxies could possess extensive neutrino dark haloes, even if the neutrino mass is as low as 2 eV. It is true that no ...Missing: mimic | Show results with:mimic