Fact-checked by Grok 2 weeks ago

Free-fall time

The free-fall time is the characteristic timescale over which a self-gravitating body, such as a uniform-density gas cloud, collapses to a point under its own in the absence of any supporting forces like thermal pressure, , or . This concept originates from classical Newtonian and represents the duration of free radial infall from an initial configuration to the center of mass. In , it provides a for dynamical processes driven by self-, particularly in the formation and evolution of , protostellar cores, and larger structures like molecular clouds. The standard formula for the free-fall time of a pressureless, uniform-density sphere is t_\mathrm{ff} = \sqrt{\frac{3\pi}{32 G \rho}}, where G is the gravitational constant and \rho is the mean mass density of the body. This expression is obtained by applying conservation of energy to the motion of infalling shells, integrating the radial velocity from the initial radius to zero while accounting for the time-dependent gravitational potential during homologous collapse. The timescale scales inversely with the square root of density, meaning denser configurations collapse more rapidly; for example, at the average density of the Sun (\rho \approx 1.4 g/cm³), t_\mathrm{ff} \approx 30 minutes. Variations of this formula exist for non-spherical geometries, such as filaments or sheets, which are common in interstellar medium structures. Free-fall time plays a central role in modeling gravitational instabilities, such as the Jeans criterion for cloud fragmentation, where collapse occurs if the cloud mass exceeds a critical value relative to its thermal support. In star formation, it sets the pace for gas accretion onto protostars, with observations indicating that only a small fraction—typically \epsilon_\mathrm{ff} \approx 0.01 or 1%—of a molecular cloud's mass converts to stars per free-fall time in nearby galaxies, highlighting the regulatory effects of feedback mechanisms like radiation and outflows. For typical molecular clouds with densities \rho \sim 10^{-20} g/cm³, t_\mathrm{ff} ranges from 1 to 10 million years, influencing the overall efficiency and duration of star cluster formation. Beyond stellar contexts, the free-fall time informs larger-scale phenomena, including the dynamical relaxation of galaxy clusters and the infall phases in cosmological structure formation.

Fundamentals

Definition

The free-fall time is the characteristic timescale for gravitational collapse under idealized conditions, defined as the duration required for a uniform, pressure-free sphere to contract to a point solely under its self-gravity. This concept establishes the fundamental rate at which unbound material would coalesce in the absence of supporting forces, serving as a baseline for analyzing dynamical processes in self-gravitating systems. In precise terms, the free-fall time represents the interval for a , initially at rest at a given within a , to fall radially inward to the center. It assumes no gradients, , or other resistive effects that could alter the motion, thereby isolating pure . The concept originates from early work in on and instability, such as analyses by James . Key assumptions underpinning this include purely radial infall trajectories, negligible environmental resistance akin to a , and an of rest at the starting from the center. Qualitatively, the free-fall time scales proportionally to the of the initial radius cubed divided by the times the enclosed , mirroring the dependence of orbital periods in Kepler's third law for bound systems.

Basic Formula

The free-fall time t_{\mathrm{ff}} provides a characteristic timescale for under self-gravity, assuming no supporting forces such as or . For a uniform-density of total M and initial radius R, the free-fall time is given by t_{\mathrm{ff}} = \pi \sqrt{\frac{R^3}{8 G M}}, where G is the gravitational constant.https://iopscience.iop.org/article/10.1088/0004-637X/744/2/190 Equivalently, in terms of the uniform density \rho = 3M / (4\pi R^3), this expression becomes t_{\mathrm{ff}} = \sqrt{\frac{3\pi}{32 G \rho}}. $$$$https://iopscience.iop.org/article/10.1088/0004-637X/744/2/190$$$$https://www.aanda.org/articles/aa/full_html/2018/03/aa31522-17/aa31522-17.html$$ This equivalence highlights how the timescale depends on the overall gravitational potential depth, parameterized either by mass and size or by mean density. Dimensional analysis confirms that the free-fall time emerges naturally as the only combination of the gravitational constant $ G $ (with units m³ kg⁻¹ s⁻²), density $ \rho $ (kg m⁻³), or mass $ M $ (kg) and radius $ R $ (m) that yields units of time (seconds), yielding the scaling $ t_{\mathrm{ff}} \propto 1 / \sqrt{G \rho} $ or $ t_{\mathrm{ff}} \propto \sqrt{R^3 / (G M)} $.$$https://iopscience.iop.org/article/10.1088/0004-637X/744/2/190$$ Here, $ G = 6.67430 \times 10^{-11} $ m³ kg⁻¹ s⁻² serves as the universal constant linking mass, distance, and gravitational acceleration across all scales.$$https://www.aanda.org/articles/aa/full_html/2018/03/aa31522-17/aa31522-17.html$$ These formulas provide order-of-magnitude estimates for the duration of radial infall in idealized [collapse](/page/Collapse) scenarios, valid for homologous contractions where the [structure](/page/Structure) remains self-similar during the process.$$https://iopscience.iop.org/article/10.1088/0004-637X/744/2/190$$ ## Derivations ### Point Mass Gravity In the context of Newtonian [gravity](/page/Gravity), the free-fall time toward a central point mass describes the duration for a [test particle](/page/Test_particle) to infall radially from an initial distance $R$ to the origin, assuming it starts at rest with no [angular momentum](/page/Angular_momentum). The [gravitational potential](/page/Gravitational_potential) is $\Phi(r) = -GM/r$, where $M$ is the mass of the central point [source](/page/Point_source) and $G$ is the [gravitational constant](/page/Gravitational_constant). The acceleration follows the [inverse-square law](/page/Inverse-square_law): $d^2 r / dt^2 = -GM / r^2$. To derive the free-fall time, apply conservation of mechanical energy. The total energy per unit mass at the initial position $r = R$ (where velocity $v = 0$) is $E = -GM/R$. At any radius $r < R$, the energy equation is $\frac{1}{2} v^2 - GM/r = -GM/R$, yielding the speed $v(r) = -\sqrt{2GM \left(1/r - 1/R\right)}$ (negative sign for inward motion). The time $t_\mathrm{ff}$ is then the integral $t_\mathrm{ff} = \int_R^0 \frac{dr}{v(r)} = \int_0^R \frac{dr}{\sqrt{2GM \left(1/r - 1/R\right)}}$. This integral can be evaluated using the substitution $r = R \cos^2 \theta$, where $\theta$ ranges from 0 at $r = R$ to $\pi/2$ at $r = 0$. The differential becomes $dr = -2R \cos \theta \sin \theta \, d\theta$, and $1/r - 1/R = (\sin^2 \theta)/(R \cos^2 \theta)$, so $\sqrt{1/r - 1/R} = (\sin \theta)/(\sqrt{R} \cos \theta)$. Substituting yields $t_\mathrm{ff} = \sqrt{\frac{R^3}{2GM}} \int_0^{\pi/2} 2 \cos^2 \theta \, d\theta$. The integral $\int_0^{\pi/2} \cos^2 \theta \, d\theta = \pi/4$, giving the exact result $t_\mathrm{ff} = \frac{\pi}{2} \sqrt{\frac{R^3}{2GM}}$. An alternative parametric approach leverages the analogy to Keplerian orbits. The radial infall corresponds to half the [orbital period](/page/Orbital_period) of a degenerate [ellipse](/page/Ellipse) (eccentricity $e = 1$) with semi-major axis $a = R/2$, determined by matching the [specific energy](/page/Specific_energy) $E = -GM/(2a) = -GM/R$. Kepler's third law provides the period $T = 2\pi \sqrt{a^3 / GM}$, so $T = \pi \sqrt{R^3 / (2GM)}$. The infall time is then $t_\mathrm{ff} = T/2 = \frac{\pi}{2} \sqrt{\frac{R^3}{2GM}}$, confirming the direct [integration](/page/Integration) result. This derivation assumes a purely radial [trajectory](/page/Trajectory) with zero initial [velocity](/page/Velocity) and no [angular momentum](/page/Angular_momentum), idealizing the motion as one-dimensional. The solution breaks down near the [singularity](/page/Singularity) at $r = 0$, where the point-mass approximation leads to infinite [velocity](/page/Velocity) and [density](/page/Density), rendering the model unphysical for real systems without additional physics such as [pressure](/page/Pressure) or finite-size effects. ### Uniform Density [Sphere](/page/Sphere) The free-fall time for a uniform [density](/page/Density) [sphere](/page/Sphere) describes the characteristic timescale for the [gravitational collapse](/page/Gravitational_collapse) of a self-gravitating, homogeneous, pressureless [cloud](/page/Cloud) under its own [gravity](/page/Gravity). Consider a [sphere](/page/Sphere) with uniform [density](/page/Density) $\rho$, initial [radius](/page/Radius) $R$, and total mass $M = \frac{4}{3} \pi R^3 \rho$. This setup assumes spherical symmetry and neglects [pressure](/page/Pressure), [rotation](/page/Rotation), and [magnetic fields](/page/The_Magnetic_Fields), focusing on purely [gravitational dynamics](/page/Dynamics). The [gravitational field](/page/Gravitational_field) inside the sphere is derived from [Poisson's equation](/page/Poisson's_equation) $\nabla^2 \Phi = 4\pi [G](/page/G) \rho$, which for uniform density and spherical symmetry yields a linear [acceleration](/page/Acceleration) $g(r) = -\frac{4\pi [G](/page/G) \rho}{3} r$ directed toward the center, where $r$ is the radial distance from the center. For a particle or [shell](/page/Shell) at position $r$, the equation of motion is $\frac{d^2 r}{dt^2} = g(r)$. In a homologous [collapse](/page/Collapse), where all radial positions [scale](/page/Scale) uniformly with the sphere's radius $R(t)$, each [shell](/page/Shell) maintains its fractional position $x = r / R(t)$ constant. This implies the density remains uniform during the [collapse](/page/Collapse), increasing as $\rho(t) = \rho / a^3(t)$, where $a(t) = R(t)/R$ is the [scale](/page/Scale) factor.[](https://arxiv.org/pdf/1207.3078) The dynamics reduce to an equation for the scale factor: $\ddot{a} = -\frac{4\pi G \rho}{3 a^2}$, obtained by substituting the time-dependent density and the homologous form into the acceleration equation. Assuming the collapse starts from rest ($\dot{a}(0) = 0$ at $a(1) = 1$), conservation of energy gives $\dot{a}^2 = \frac{8\pi G \rho}{3} \left( \frac{1}{a} - 1 \right)$. Integrating $t = \int_1^a \frac{da'}{\dot{a}(a')}$ from initial rest to the singularity at $a = 0$ yields the free-fall time $t_\mathrm{ff} = \sqrt{\frac{3\pi}{32 G \rho}}$. This result is independent of the initial radius $R$, depending only on the initial density $\rho$, as the linear field ensures homologous motion scales uniformly across shells. Under the homology assumption, all shells collapse similarly, preserving the uniform density profile until a central singularity forms at $t_\mathrm{ff}$, at which point the model breaks down. This pressureless, homologous collapse serves as a foundational precursor to more sophisticated self-similar models of [gravitational collapse](/page/Gravitational_collapse), such as the Larson-Penston solution for isothermal spheres, which incorporate pressure gradients while building on the uniform density dynamics. ## Applications ### [Star Formation](/page/Star_formation) In the process of [star formation](/page/Star_formation), the free-fall time serves as a fundamental bottleneck for the [gravitational collapse](/page/Gravitational_collapse) of dense regions within turbulent molecular [clouds](/page/Cloud), dictating the pace at which gas accumulates to form protostars. Molecular [clouds](/page/Cloud), supported against immediate [collapse](/page/Collapse) by supersonic [turbulence](/page/Turbulence), exhibit virialized structures where only subregions exceeding critical densities undergo fragmentation and infall on free-fall timescales, limiting the overall star formation efficiency to roughly 1-10% per free-fall time. This inefficiency arises because [turbulence](/page/Turbulence) dissipates energy and regulates the rate at which mass crosses the [Jeans](/page/Jeans) threshold, preventing wholesale cloud [collapse](/page/Collapse).[](https://arxiv.org/abs/1312.5365) For typical densities in star-forming cloud cores of approximately $10^{-20}$ g cm$^{-3}$, corresponding to number densities around $10^{3}$ cm$^{-3}$, the free-fall time ranges from $10^{5}$ to $10^{6}$ years, providing a rapid phase for core contraction once instability sets in. However, observed lifetimes of molecular clouds and prestellar cores often extend to 10-30 free-fall times, as turbulence and external pressures prolong the pre-collapse phase, contrasting with the shorter dynamical timescales expected in isolation. The [Jeans instability](/page/Jeans_instability) triggers this collapse in overdense regions where the thermal Jeans timescale, $\sqrt{\pi c_s^2 / (G \rho)}$, becomes comparable to or shorter than the free-fall time, allowing self-gravity to overcome support and initiate infall; subsequent core fragmentation during [collapse](/page/Collapse) can produce binary or multiple stellar systems, depending on the initial mass reservoir and [angular momentum](/page/Angular_momentum).[](https://arxiv.org/abs/1312.5365) Telescopic observations, including [ALMA](/page/Alma)'s high-resolution mapping of molecular line profiles and Hubble's imaging of silhouetted structures, confirm [collapse](/page/Collapse) [dynamics](/page/Dynamics) in [Bok globule](/page/Bok_globule)s—compact, isolated cloudlets thought to represent early star-forming sites—aligning with free-fall timescales. For example, in the [Bok globule](/page/Bok_globule) B335, interferometric data reveal infalling gas motions with velocities consistent with singular isothermal sphere [collapse](/page/Collapse) models, occurring over approximately $10^{5}$ years at core densities, indicative of active [protostar](/page/Protostar) formation. [ALMA](/page/Alma) surveys of nearby clouds, such as those in the FREJA project targeting [Taurus](/page/Taurus), further show that prestellar cores evolve toward protostellar phases on timescales approaching the local free-fall time, with minimal fragmentation in low-mass examples like [Bok globule](/page/Bok_globule)s. Modern theoretical models incorporate [magnetic fields](/page/The_Magnetic_Fields), which thread molecular clouds and resist collapse, effectively extending the timescale beyond the pure free-fall estimate via [ambipolar diffusion](/page/Ambipolar_diffusion)—the relative drift of neutrals past ionized [plasma](/page/Plasma) frozen to field lines. In magnetized environments, the [ambipolar diffusion](/page/Ambipolar_diffusion) timescale, often $t_{ad} \sim t_{ff} \times (\rho / \rho_i)$, where $\rho_i$ is ion [density](/page/Density), can exceed the free-fall time by factors of 10-100, delaying [core](/page/Core) formation until sufficient [flux](/page/Flux) loss occurs; this mechanism explains the observed longevity of quiescent cores in observations of regions like [Taurus](/page/Taurus). ### Planetary Accretion In protoplanetary disks, the free-fall time represents the characteristic timescale over which dust grains and planetesimals can radially drift toward the central star or aggregate under their mutual self-gravity, influencing the early stages of [solid body](/page/Solid_body) formation. This process occurs primarily in the dense midplane layer where particles settle, with the free-fall time dictating the efficiency of gravitational instabilities that concentrate material into larger structures. Unlike gaseous [collapse](/page/Collapse) in molecular clouds, here the dynamics are modulated by the disk's Keplerian [rotation](/page/Rotation) and differential drift, limiting direct radial infall unless triggered by local overdensities. However, in rotating disks, support from Keplerian motion extends the effective [collapse](/page/Collapse) timescale to the local [orbital period](/page/Orbital_period), typically 1–1000 years. For typical midplane densities in protoplanetary disks of ρ ≈ 10^{-9} g/cm³, the free-fall time is estimated at approximately [a few](/page/The_Few) years, a duration that aligns with the onset of [runaway](/page/Runaway) accretion phases where [a few](/page/The_Few) dominant planetesimals rapidly capture surrounding material, outpacing the growth of smaller bodies. These timescales are derived from the standard [free-fall formula](/page/Formula) applied to local disk conditions and play a pivotal role in determining whether accretion proceeds efficiently before the disk dissipates. In self-gravitating disks, such short free-fall times in enhanced density regions, like spiral arms, can boost collision probabilities by factors of up to 100, accelerating planetesimal growth to kilometer scales. The foundational Safronov-Goldreich-Ward theory integrates the free-fall time into models of [planetesimal](/page/Planetesimal) accretion by linking it to collision rates, where low relative velocities—damped by gas drag—enable gravitational focusing within the Hill radius, ensuring dynamical stability during pairwise mergers. In this framework, the free-fall time sets the pace for encounters in the particle "gas," with growth rates scaling inversely with it in dense swarms, as originally formulated for the solar nebula but applicable to exoplanet-forming disks. Hill radius considerations further constrain accretion, requiring free-fall collapse to occur before [shear](/page/Shear) disperses the material. Telescopic observations of [exoplanet](/page/Exoplanet) formation timelines, such as those from TESS and JWST targeting young systems like [PDS 70](/page/PDS_70), reveal disk lifetimes of 1–10 million years, wherein the free-fall time imposes fundamental limits on accretion efficiency by bounding the number of coalescence events possible before gas photoevaporation halts solid buildup. These data suggest that efficient [planetesimal](/page/Planetesimal) accretion requires free-fall times short enough relative to disk evolution to form cores of 5–10 [Earth](/page/Earth) masses, consistent with core accretion models for super-Earths and ice giants.[](https://pubs.geoscienceworld.org/msa/rimg/article/90/1/55/645036/Planet-Formation-Observational-Constraints) In denser disk regions, aerodynamic [drag](/page/drag) from the nebular gas alters the effective free-fall time, decelerating small [dust](/page/Dust) grains (sizes ≲1 cm) through Epstein [drag](/page/drag) while channeling intermediate planetesimals (10–100 cm) toward high-density zones like spirals, thereby shortening coalescence times by enhancing local concentrations. This modification is particularly pronounced in turbulent disks, where drag-induced [settling](/page/Settling) reduces radial velocities, allowing self-gravity to dominate over [diffusion](/page/Diffusion) on timescales comparable to the free-fall estimate.[](https://academic.oup.com/mnras/article/355/2/543/970670) ## Comparisons ### Dynamical Timescale The dynamical timescale, $ t_{\rm dyn} $, represents the characteristic time for significant gravitational changes in a self-gravitating [system](/page/System) and is defined as $ t_{\rm dyn} = R / \sigma $, where $ R $ is the [system](/page/System)'s radius and $ \sigma $ is the one-dimensional velocity dispersion of its constituents.[](https://arxiv.org/pdf/astro-ph/0603278) This timescale approximates the time required for a particle to cross the [system](/page/System) at the typical speed set by internal motions, providing a measure of how quickly the [system](/page/System) responds to perturbations. In virialized self-gravitating systems, the dynamical timescale is equivalent to the free-fall time $ t_{\rm ff} $, both scaling as $ \sqrt{R^3 / GM} $, where $ M $ is the total mass; this equivalence arises from the virial theorem, which balances kinetic and potential energies in steady-state configurations. To sketch the derivation, consider the scalar virial theorem for a self-gravitating system: $ 2K + W = 0 $, where $ K $ is the total kinetic energy and $ W $ is the gravitational potential energy. For a system with $ N $ particles, $ K \approx (3/2) M \sigma^2 $ (assuming isotropic dispersion in three dimensions), and $ |W| \approx GM^2 / R $ for a typical configuration. Balancing these yields $ \sigma^2 \approx GM / R $, so $ t_{\rm dyn} = R / \sigma \approx \sqrt{R^3 / GM} $, matching the free-fall scaling derived from collapse dynamics. In dense regions, numerical factors may differ slightly (e.g., $ t_{\rm ff} \approx 0.5 t_{\rm dyn} $ for virial parameter near unity), but the timescales remain of the same order.[](https://arxiv.org/pdf/astro-ph/0603278) This equivalence is evident in galactic dynamics, where the free-fall time estimates relaxation or crossing times; for instance, in the Milky Way's core, with densities around $ 10^4 $--$ 10^5 $ cm$^{-3} $, $ t_{\rm ff} \approx 10^5 $ years, aligning with dynamical processes in the central molecular zone.[](https://www.aanda.org/articles/aa/full_html/2014/08/aa23943-14/aa23943-14.html) While the dynamical timescale is often used to characterize orbital motions and [stability](/page/Stability) in bound systems, the free-fall time specifically emphasizes [collapse](/page/Collapse) phases, such as in cloud fragmentation, though both inform the pace of gravitational evolution. ### Freefall vs Orbital Time The Keplerian [orbital period](/page/Orbital_period) characterizes the timescale for angular motion in a [stable](/page/Stable) [circular orbit](/page/Circular_orbit) around a central point mass $ M $ at radius $ R $, given by t_\text{orb} = 2\pi \sqrt{\frac{R^3}{GM}}. This formula arises from equating gravitational attraction to the [centripetal force](/page/Centripetal_force) required for [circular motion](/page/Circular_motion) and follows from Kepler's third law generalized to Newtonian [gravity](/page/Gravity).[](https://science.nasa.gov/solar-system/orbits-and-keplers-laws/) In contrast, the free-fall time $ t_\text{ff} $ governs purely radial motion, where a [test particle](/page/Test_particle) starts from rest at $ R $ and falls inward under [gravity](/page/Gravity) alone. For a point mass, t_\text{ff} = \frac{\pi}{2\sqrt{2}} \sqrt{\frac{R^3}{GM}} \approx 0.18 , t_\text{orb}. The shorter free-fall time reflects the absence of angular momentum support in radial trajectories, allowing faster collapse compared to orbital circulation. This ratio is derived from integrating the radial equation of motion using conservation of energy, where the velocity is $ v(r) = \sqrt{2GM \left( \frac{1}{r} - \frac{1}{R} \right)} $, yielding the time as the integral $ t_\text{ff} = \int_0^R \frac{dr}{v(r)} $.[](https://arxiv.org/pdf/1005.5279) The distinction between these timescales has key implications for gravitational dynamics, particularly in systems where radial infall competes with orbital stability. Since $ t_{\rm ff} < t_{\rm orb} $, material can [collapse](/page/Collapse) inward before completing even a fraction of an [orbit](/page/Orbit), leading to rapid accretion and potential instabilities in flows assuming near-Keplerian [rotation](/page/Rotation). In astrophysical accretion disks, this drives phenomena like the viscous instability or clumping, as infalling matter disrupts azimuthal balance on dynamical timescales comparable to but shorter than full orbital periods.[](https://arxiv.org/pdf/1005.5279) Examples illustrate this radial-angular contrast in specific contexts. During the final stages of binary mergers, such as [neutron star](/page/Neutron_star) or [black hole](/page/Black_hole) systems, the inspiral follows orbital dynamics governed by $ t_\text{orb} $, but the ultimate coalescence occurs on free-fall timescales, enabling prompt merger and [gravitational wave](/page/Gravitational_wave) emission as [angular momentum](/page/Angular_momentum) is shed.[](https://arxiv.org/pdf/2201.12218) Similarly, in warped accretion disks—where misalignment between inner and outer regions induces [precession](/page/Precession)—regions with $ t_\text{ff} < t_\text{orb} $ promote infall over sustained orbiting, contributing to disk breaking or material drainage toward the central object.[](https://arxiv.org/pdf/2303.14210) In [general relativity](/page/General_relativity), for compact objects like black holes, GR corrections modify the free-fall time, reducing it relative to the Newtonian value by factors incorporating the compactness $ GM/(c^2 R) $. In the weak-field limit, GR corrections to the free-fall [proper time](/page/Proper_time) approach the Newtonian value. Near the event horizon, relativistic effects significantly shorten the proper time relative to Newtonian expectations, while the [coordinate time](/page/Coordinate_time) for distant observers diverges logarithmically. This is particularly pronounced for marginally bound geodesics from [infinity](/page/Infinity), where proper time to the horizon remains finite.[](https://arxiv.org/pdf/1707.05187)

References

  1. [1]
    Freefall Time and Star Formation Time - HyperPhysics
    Freefall Time and Star Formation Time ; r0 = x10^ m ; r0 = parsecs ; r0 = lightyears ; the initial density is ρ0 = kg/m ; the calculated freefall time is tff = x s.
  2. [2]
    [PDF] Gravitational Collapse: Jeans Criterion and Free Fall Time
    In this course we apply the equations of hydrodynamics to various phases of stellar life. We start with the onset of star formation, asking under which ...
  3. [3]
    Star Formation Efficiency per Free-fall Time in nearby Galaxies
    1. Introduction. Star formation is “inefficient”, meaning that the star formation rate (SFR) is low compared to what would be expected if cold gas collapsed ...
  4. [4]
    understanding the formation of galaxy clusters
    ... free-fall time of a uniform sphere tff = [3 pi / (32G rho ta]1/2, which means that the mean density of perturbation at turnaround is rho ta = 3 pi / (32Gtta2) ...
  5. [5]
    (PDF) The Free-fall Time of Finite Sheets and Filaments
    Aug 9, 2025 · Molecular clouds often exhibit filamentary or sheet-like shapes. We compute the free-fall time (\tff) for finite, uniform, self-gravitating ...<|control11|><|separator|>
  6. [6]
    The Molecular Cloud Lifecycle - PMC - NIH
    ... free-fall time of molecular clouds is evaluated to be ∼ 10 Myr . If all the gas is converted to stars within a free-fall time, the free-fall rate of star ...
  7. [7]
    [PDF] Introduction - Yale Astronomy
    This is related to the free-fall time, defined as the time required for a uniform, pressure-free sphere to collapse to a point, as tff = tdyn/√2. • Cooling time ...
  8. [8]
    Jeans' gravitational instability and nonextensive kinetic theory
    In a seminal paper, James Jeans discussed the conditions under which a fluid becomes gravitationally unstable under the action of its own gravity (Jeans 1929).
  9. [9]
    [PDF] arXiv:1207.3078v2 [astro-ph.GA] 27 Jul 2012
    Jul 27, 2012 · The free-fall collapse timescale of a uniform-density sphere is a well-studied problem (e.g., Binney & Tremaine 1987) and it is known that the ...
  10. [10]
    [1312.5365] The Star Formation Rate of Molecular Clouds - arXiv
    Dec 18, 2013 · We review recent advances in the analytical and numerical modeling of the star formation rate in molecular clouds and discuss the available ...
  11. [11]
    Accelerated planetesimal growth in self-gravitating protoplanetary ...
    In this paper we consider the evolution of small planetesimals (radii ∼1–10 m) in marginally stable, self-gravitating protoplanetary discs. The drag force ...
  12. [12]
    Planet Formation—Observational Constraints, Physical Processes ...
    Jul 1, 2024 · The goal of planet formation as a field of study is not only to provide the understanding of how planets come into existence.Dynamics Of Dust And Pebbles · Solid Accretion Mechanisms · Gas Accretion
  13. [13]
    [PDF] arXiv:astro-ph/0603278v1 10 Mar 2006
    Mar 10, 2006 · The free- fall time is tff = (3π/32Gρ)1/2, and the dynamical, or crossing, time is tdyn = R/σ. ... points out, the relevant dynamical timescale ...
  14. [14]
    The dynamics and star-forming potential of the massive Galactic ...
    For a molecular hydrogen density of 104 to 105 cm-3, the free fall time is several 105 years, and thus gives an upper limit for the age of G0.253+0.016, in ...
  15. [15]
    Orbits and Kepler's Laws - NASA Science
    May 2, 2024 · This is written in equation form as p2=a3. Kepler's third law implies that the period for a planet to orbit the Sun increases rapidly with the ...Kepler's Laws of Planetary... · Kepler and the Mars Problem
  16. [16]
    [PDF] Accretion Disks - arXiv
    May 28, 2010 · on the order of the free fall time scale, or Kepler orbital time scale, τd = r/vK = Ω. −1. K. = (r3/GM)1/2,. (4) where vK, ΩK are the Kepler ...
  17. [17]
    [PDF] arXiv:2201.12218v1 [astro-ph.CO] 28 Jan 2022
    Jan 28, 2022 · So, for λ > 0.9, the mergers will be dominated by the free fall time. In Fig. 7 we show the resulting merger rate. For redshifts before matter ...
  18. [18]
    [PDF] Disk settling and dynamical heating - arXiv
    Mar 24, 2023 · particle in a local free-fall time. ... Again, all of these effects are at the level of ≲ 3%, where non-equilibrium effects like disk warping and ...
  19. [19]
    [PDF] arXiv:1707.05187v2 [gr-qc] 3 Oct 2019
    Oct 3, 2019 · description of the outer an inner regions of the Schwarzschild black hole. ... and absolute “free fall” time. Even if built on Euclidean space and ...