Fact-checked by Grok 2 weeks ago

Stellar rotation

Stellar rotation is the angular motion of a about its own , typically characterized by the equatorial rotational v and the inclination i of the rotation relative to the line of , with surface velocities ranging from near zero to several hundred km/s depending on the , age, and evolutionary stage. This arises from the conservation of during the from collapsing protostellar clouds and persists throughout its lifetime, influencing internal , chemical mixing, and surface activity. Measurements of stellar primarily rely on the broadening of spectral lines due to the , quantified as v \sin i, or photometric variations from starspots and other surface features in cooler stars. Rotation plays a pivotal role in stellar evolution by driving meridional circulation and shear instabilities that enhance element transport, leading to altered chemical gradients and more homogeneous compositions in massive stars. In upper main-sequence stars, rapid initial rotation—often 200–400 km/s—can elongate the star into an oblate spheroid, modify via the von Zeipel theorem, and amplify mass loss through anisotropic stellar winds, thereby removing and extending evolutionary tracks in the Hertzsprung-Russell diagram. For low-mass stars like , rotation generates dynamo-driven magnetic fields that brake the star over time, slowing rotation from periods of days in youth to months in old age, as evidenced by observations of open clusters. Notable phenomena linked to stellar rotation include the formation of Be stars through equatorial mass ejection and the spin-up of white dwarfs from interactions, while —faster at the equator than poles—complicates internal transport via magnetic torques and turbulence. Across spectral types, early-type O and B stars rotate fastest (up to 0.8–1.0 of breakup velocity), while cooler , , and dwarfs exhibit slower rates modulated by convection zones, with historical observations tracing back to solar sunspot rotations in the . Overall, incorporating rotation into stellar models is essential for accurate predictions of , progenitors, and galactic chemical enrichment.

Fundamentals and History

Definition and Basic Principles

Stellar rotation refers to the angular motion of a star about its . This motion is quantified by the angular velocity \Omega, defined as the rate of change of angular position with time, typically expressed in radians per second. Equivalently, rotation can be characterized by the equatorial velocity v_{\rm eq} = \Omega R_{\rm eq}, where R_{\rm eq} is the radius at the stellar . In an isolated , L = I \Omega—with I denoting the —is conserved due to the absence of external torques, leading to slower as the expands during . However, external torques from magnetized stellar winds or interactions in systems can alter this , resulting in loss or transfer that spins down the over time. Rotation fundamentally influences by perturbing , where the centrifugal acceleration reduces effective gravity in the equatorial plane, and by modifying energy transport mechanisms, as the warped isobaric surfaces affect radiative and convective fluxes. One common observable is the projected rotational velocity v \sin i, with i the inclination of the rotation axis relative to the . The rotational , a measure of the star's energy, is expressed as E_{\rm rot} = \frac{1}{2} I \Omega^2, which contributes significantly to the total energy budget in rapidly rotating stars. Furthermore, stellar rotation is essential for dynamo action in convective zones, where the combination of rotation, convection, and shear generates and maintains magnetic fields through the \alpha-\Omega dynamo mechanism. This process couples rotation to magnetic activity, influencing stellar winds and angular momentum loss.

Historical Development

The theoretical groundwork for measuring stellar rotation through broadening was established in 1929 by G. Shajn and O. Struve, who demonstrated how the from a star's axial rotation would distort absorption line profiles, allowing the first quantitative estimates of rotational velocities. Their analysis applied this method to , revealing its relatively low projected rotational velocity and marking the initial observational breakthrough in quantifying rotation for individual stars beyond . In the 1930s, advanced this work by formalizing the parameter v \sin i, where v is the equatorial rotational and i is the inclination of the rotation axis relative to the , enabling systematic measurements of projected velocities across diverse stellar types using high-resolution . This concept became foundational for cataloging rotational properties, with Struve's studies of B- and A-type stars highlighting rapid rotators and influencing early models of . Post-World War II, the advent of photoelectric photometry in the and facilitated the detection of periods in stars with surface spots, as brightness modulations from spot rotation across the stellar disk provided direct period estimates for active late-type stars. By the 1970s and 1980s, theoretical models began integrating rotation into frameworks, emphasizing its role in driving internal mixing and transport. A seminal contribution came from Endal and Sofia in , whose numerical simulations of rotating low-mass stars predicted surface velocity evolutions and linked rotational shear to enhanced convective mixing, thereby influencing and surface abundances. These models underscored rotation's importance in resolving discrepancies between observed and predicted stellar properties, such as lithium depletion. In recent decades, space-based photometry has revolutionized the field by enabling large-scale surveys of rotation periods. The Kepler mission (2009–2018) delivered high-precision light curves for over 150,000 stars, yielding rotation periods for more than 34,000 main-sequence targets through analysis of photometric variability, which revealed empirical relations like gyrochronology for age estimation. Building on this, the (TESS), launched in 2018, has extended such measurements to brighter, nearby stars across wider sky coverage, detecting periods for tens of thousands via full-frame images and sector light curves, thus facilitating studies of rotation in solar analogs and young clusters. The Sun's , with faster equatorial speeds than polar regions observed via sunspots since the , has long served as a prototypical example for interpreting these stellar patterns.

Measurement Methods

Spectroscopic Techniques

Spectroscopic techniques provide an indirect measure of stellar rotation by analyzing the broadening of spectral lines caused by the Doppler effect from the star's surface motions. As different parts of the rotating stellar disk approach or recede from the observer, the emitted light experiences wavelength shifts, resulting in a broadened line profile whose width is proportional to the projected equatorial rotation velocity, v \sin i, where v is the true equatorial velocity and i is the inclination of the rotation axis relative to the line of sight. The full width at half maximum (FWHM) of the rotationally broadened line directly relates to this parameter, allowing estimation of rotational speeds typically in the range of a few to several hundred km/s for main-sequence stars. The rotational broadening is described by the convolution of the intrinsic line profile with a broadening function G(v), which for a uniformly bright disk approximates G(v) = \frac{2}{\pi v \sin i} \sqrt{1 - \left( \frac{v}{v \sin i} \right)^2}, where v is the coordinate. This function produces an elliptical line profile symmetric about the line center, with the broadening extent limited to \pm v \sin i. However, the inclination i introduces an inherent ambiguity, as only \sin i \leq 1 is observable, meaning the true equatorial satisfies v_{\rm eq} \geq v \sin i, with the minimum v_{\rm eq} occurring when the star is viewed equator-on (i = 90^\circ). Accurate determination of v \sin i requires corrections for confounding effects that can mimic or alter rotational broadening. Macroturbulence, representing large-scale fields in the stellar atmosphere, broadens lines similarly to and must be disentangled through profile fitting or methods, often assuming a Gaussian . Variations in elemental abundances can distort line strengths and shapes, necessitating detailed atmospheric modeling to isolate rotational contributions. , where the stellar disk appears brighter at the center than the edges due to deeper atmospheric penetration of radiation, also modifies the broadening function, requiring inclusion of a linear or quadratic limb-darkening coefficient in the . These corrections are typically applied via least-squares fitting of synthetic spectra or of observed profiles to achieve precisions of 1-5 km/s. These techniques are particularly effective for hot O and B-type stars, where high effective temperatures (T_{\rm eff} > 10,000 K) result in intrinsically sharp spectral lines with minimal pressure or thermal broadening, making rotational effects prominent. For instance, in field O stars, v \sin i values range from slow rotators at ~10 km/s to near-breakup speeds exceeding 300 km/s, as measured from He I and metal lines in high-resolution spectra. A well-studied example is Vega (α Lyr), an A0V star with sharp lines amenable to precise analysis, yielding v \sin i \approx 21.6 km/s from Fourier transforms of multiple metallic lines.

Photometric and Direct Imaging Methods

Photometric methods measure stellar periods by detecting periodic variations in a star's brightness caused by dark starspots rotating into and out of view. These modulations in light curves arise from the contrast between cooler spots and the surrounding , allowing the P_{\rm rot} to be determined from the dominant periodicity in the flux . For solar-like main-sequence stars, the Kepler mission provided extensive data, yielding periods for over 34,000 targets spanning 0.2 to 70 days, with typical periods around 20-25 days for G-type stars similar to . These measurements reveal a broad distribution influenced by age and spectral type, enabling studies of rotational evolution without relying on line-of-sight projections. Direct imaging techniques, particularly optical long-baseline , resolve the shapes of rapidly rotating stars, providing geometric constraints on equatorial velocities. The array has been instrumental in this, achieving resolutions below 1 milliarcsecond to image stellar surfaces. A landmark observation of in 2007 using four telescopes reconstructed its near-infrared image, revealing an photosphere with an equatorial-to-polar radius ratio of approximately 1.25, corresponding to a ~25% equatorial excess, and confirming gravity darkening where the equator appears ~60-70% as bright as the poles. This approach derives true rotation velocities by combining the observed flattening with models of centrifugal distortion, offering an independent check against projected equatorial speeds v \sin i from . Gravitational microlensing events provide rare opportunities to probe in distant stars by temporarily magnifying their and allowing the to the stellar disk. During such a , the lens's motion across the star induces apparent shifts due to the Rossiter-McLaughlin effect, modulated by the star's . For K-giant lenses, where turbulent velocities often mask spectroscopically, this method can measure v \sin i to the precision of individual velocity observations, typically revealing slow rotators with periods of hundreds of days. Polarimetric techniques, particularly Zeeman-Doppler imaging (ZDI), map surface magnetic fields using the in high-resolution spectropolarimetric time series, from which and velocity fields can be inferred. ZDI inverts Stokes parameter variations across rotation phases to reconstruct vector magnetic topologies and associated surface flows, revealing latitudinal shear in young solar-type stars where equatorial regions rotate faster than poles by up to 20-30%. This tomographic approach resolves velocity fields down to ~1 km/s precision for rapidly rotating targets, complementing spot-based photometry. Recent advances include the mission, scheduled for launch in 2026, which will deliver ultra-precise photometry of up to one million bright stars to measure rotation periods with uncertainties below 1% for solar-like hosts of . By monitoring long-term light variations, PLATO will characterize surface rotation and activity in exoplanet systems, improving age estimates and obliquity constraints via cross-checks with v \sin i.

Physical Effects on Stars

Equatorial Bulge and Oblateness

Stellar rotation induces a deformation in the shape of stars, causing them to deviate from spherical as centrifugal forces oppose gravitational contraction. In , the effective gravity g_{\rm eff} at a point on the stellar surface balances the and is given by \mathbf{g}_{\rm eff} = \mathbf{g}_{\rm grav} - \Omega^2 r \sin\theta \, \hat{\phi}, where \mathbf{g}_{\rm grav} is the , \Omega is the , r is the radial distance, \theta is the , and \hat{\phi} is the azimuthal perpendicular to the axis. This modification leads to an oblate configuration, with the equatorial radius R_{\rm eq} exceeding the polar radius R_{\rm pole}, as the centrifugal term is zero at the poles but maximum at the . For slowly rotating stars, the oblateness parameter \varepsilon = (R_{\rm eq} - R_{\rm pole})/R_{\rm eq} provides a first-order approximation of this distortion, expressed as \varepsilon \approx (\Omega^2 R^3)/(3 G M), where R is the mean stellar radius, G is the , and M is the . This perturbative formula assumes uniform rotation and small rotational velocities relative to the critical breakup speed, capturing the leading quadrupole deformation without higher-order effects. Differential rotation can complicate bulge estimation by introducing latitude-dependent variations, but for uniform rotation, the approximation holds well for most main-sequence stars. The model offers a simple approximation for the shape of uniformly rotating by assuming constant and treating the as an incompressible in . In this model, the surface follows an where gravitational and centrifugal potentials balance, yielding a maximum oblateness of R_{\rm eq}/R_{\rm pole} = 1.5 at the critical rotation rate. This uniform-density assumption simplifies calculations and is particularly useful for rapidly rotating objects, though it overestimates flattening for with realistic profiles concentrated toward . Observational evidence for such deformations comes from interferometric measurements of rapidly rotating stars. For (\alpha Eri), Interferometer (VLTI) observations in 2003 revealed an shape with R_{\rm eq}/R_{\rm pole} = 1.56 \pm 0.05, corresponding to \varepsilon \approx 0.36, indicating near-critical rotation. Similarly, for A (\alpha Leo A), CHARA Array interferometry yielded R_{\rm eq} = 4.16 \, R_\odot and R_{\rm pole} = 3.14 \, R_\odot, giving R_{\rm eq}/R_{\rm pole} \approx 1.32 and \varepsilon \approx 0.24, consistent with its equatorial velocity of about 317 km/s. These deformations influence stellar photometry through asymmetries in and surface brightness. Rapid rotation causes darkening, where the equator cools relative to the poles due to reduced effective , leading to asymmetric profiles that distort light curves, particularly in transiting observations. In asteroseismology, rotational deformation perturbs oscillation modes, revealing internal structure changes.

Differential Rotation

Differential rotation refers to the variation in a star's across its surface, where the rotation rate depends on , typically faster at the and slower toward the poles. In , this manifests as an equatorial rotation period of approximately 25 days compared to about 35 days at the poles, making the equator roughly 40% faster in angular speed. This phenomenon arises primarily from internal dynamics in the star's , where turbulent convection and meridional circulation—poleward flows at the surface and equatorward at depth—redistribute , leading to latitudinal gradients in . In the underlying radiative zones, the Taylor-Proudman theorem constrains flows to be aligned with the rotation axis, promoting columnar structures that contribute to differential rotation by inhibiting latitudinal shear. Observational evidence for is well-established for through tracking migration, which reveals the latitude-dependent rates. For other , techniques like Doppler map surface features and infer profiles; for instance, the K giant ζ Andromedae exhibits solar-like , with its equator rotating faster than higher latitudes as determined from cross-correlating sequential Doppler maps. A common parameterization for solar-type stars describes the angular velocity as \Omega(\theta) = \Omega_{\rm eq} \left(1 - \alpha \cos^2 \theta \right), where \theta is the , \Omega_{\rm eq} is the equatorial rate, and \alpha \approx 0.2--$0.3$ quantifies the relative . The latitudinal shear from plays a crucial role in stellar dynamos by winding poloidal magnetic fields into strong toroidal components through the \Omega-, enabling cyclic activity like the Sun's 11-year cycle.

Evolution of Stellar Rotation

Rotation During Star Formation

Molecular cloud cores, the precursors to stars, typically exhibit very low levels of rotation, with specific angular momenta on the order of 10^{-3} to 10^{-2} pc km s^{-1}. As these cores collapse under gravity to form protostars, angular momentum conservation causes the rotation rate to increase dramatically; for a uniform density sphere, the angular velocity \Omega scales as \Omega \propto r^{-2}, where r is the radius, leading to centrifugal support preventing further central collapse without angular momentum removal. This spin-up process is crucial during the early phases of protostellar formation, transforming slowly rotating envelopes into rapidly rotating central objects. In the protostellar phase, magnetic braking plays a dominant role in regulating this excess through coupling between the forming star's and the surrounding . Strong s thread the disk, exerting torques that transfer outward, allowing material to accrete inward while slowing the protostar's ; simulations indicate this mechanism can reduce the rotation rate by a factor of up to $10^3 compared to unmagnetized scenarios. The efficiency of this disk-star coupling depends on the field strength and ionization levels, with non-ideal MHD effects like modulating the braking to permit small-scale disk formation. Outflows and bipolar jets further contribute to angular momentum ejection during this phase, primarily via magneto-centrifugal winds launched from the inner disk regions. These , driven by the twisting of lines anchored in the disk, carry away significant , with observations and models showing that jets can remove up to 90% of the accreted onto the . The Blandford-Payne mechanism underpins this process, where material is flung outward along inclined field lines, facilitating continued accretion and preventing the protostar from reaching breakup speeds prematurely. High-resolution ALMA observations of Keplerian rotation in protoplanetary disks around young stars provide key evidence for these processes, revealing well-ordered gas motions consistent with conservation in the disk after protostellar braking, with typical disk sizes of 10-100 AU. For instance, the disk around TW Hydrae, a ~10 Myr-old , shows Keplerian disk rotation, while independent photometric and spectroscopic measurements indicate an equatorial rotation speed v_\mathrm{eq} \approx 16 km s^{-1} for the central star. Similar studies in clusters like and confirm these disk properties via ALMA, while stellar rotation rates of 10-20 km/s are derived from other observations, supporting the efficiency of early angular momentum loss mechanisms. As a result, low-mass stars are born rotating at typically 5-30% of their breakup , with initial \Omega / \Omega_\mathrm{crit} \approx 0.05-0.3, where \Omega_\mathrm{crit} corresponds to the velocity at the stellar surface where balances . This range arises from the combined effects of collapse spin-up tempered by magnetic and outflow braking, setting the stage for subsequent evolutionary slowdown.

Main Sequence Braking Mechanisms

During the main sequence phase, stars lose angular momentum primarily through magnetic braking mediated by their stellar winds. This process occurs as the star's magnetic field threads through the ionized wind material, enforcing corotation out to the Alfvén radius, r_A, where magnetic stresses balance the wind's ram pressure. Beyond this radius, the wind carries away angular momentum, exerting a torque on the star given approximately by \Gamma \approx \dot{M} \Omega r_A^2, where \dot{M} is the mass-loss rate and \Omega is the stellar (with r_A depending on surface magnetic field strength B). This mechanism dominates rotational evolution for low-mass stars, leading to a progressive spin-down over billions of years. For massive stars, braking is weaker due to different wind and field properties, allowing sustained higher rotation rates. The empirical manifestation of this braking is encapsulated in the Skumanich law, which describes the rotation rate of solar-type stars as \Omega \propto t^{-1/2}, where t is the stellar age. This relation arises from integrating the torque over time, assuming a wind-driven loss that scales with the square of the rotation rate in the unsaturated regime. Validation through gyrochronology, which calibrates -age relations using open clusters, confirms the law's applicability for F-, G-, and early K-type dwarfs up to ages of several gigayears. Braking efficiency exhibits a strong mass dependence, with higher-mass F and dwarfs experiencing more rapid spin-down compared to lower-mass dwarfs due to differences in strengths, convective envelopes, and wind properties. For instance, dwarfs retain higher rotation rates at equivalent ages, reflecting weaker or less efficient magnetic coupling to their winds. Observational evidence from open clusters, such as the Hyades (age ≈ 600 ), reveals clear age- sequences where main-sequence follow the expected spin-down trend, with rotation periods increasing from a few days in younger clusters to tens of days in the Hyades. At young ages, braking saturates for rapidly rotating with \Omega \gtrsim 10 \Omega_\odot, where the becomes independent of rotation rate, leading to a plateau in spin-down before the unsaturated Skumanich regime takes over. Recent models have refined earlier formulations, such as those by Kawaler (1988), by incorporating Zeeman-Doppler imaging (ZDI) measurements of large-scale magnetic field geometries and strengths, which show that non-dipolar fields reduce the effective Alfvén radius and torque compared to simple dipole assumptions. These updates improve predictions for diverse spectral types, particularly for active stars where field complexity influences braking rates.

Changes in Advanced Evolutionary Stages

As stars evolve off the onto the (RGB), the rapid expansion of their convective envelopes significantly alters the distribution of . The envelope's growth reduces its , causing the surface rotation to slow dramatically, while the contracting conserves and spins up. Asteroseismic observations indicate that rotation rates (Ω_core) can reach up to 100 times the surface rate, with typical periods of a few weeks compared to surface periods exceeding months. This is maintained through internal processes that redistribute , preventing full coupling between and . Meridional circulation, driven by rotation-induced instabilities, transports angular momentum outward from the , while in radiative zones provide "locking" that resists further spin-up of the . These mechanisms, inferred from models incorporating hydrodynamic simulations, explain the observed persistence of rapid rotation throughout much of the RGB . Upon reaching the (HB) after helium ignition, and progressing to the (AGB), surface rotation resumes slowing due to reactivated magnetic braking from the expanded envelope. By the time stars become progenitors, their surface projected rotational velocities are typically low, with v sin i < 10 km/s, reflecting efficient angular momentum loss over the post-main-sequence lifetime. In contrast, low-mass ultracool dwarfs experience diminished braking efficiency owing to weaker dynamo activity and altered magnetic field topologies, leading to sustained rapid rotation with periods of approximately 1–10 days even in older objects. Asteroseismic analyses of red giants observed by the K2 mission provide direct evidence of this core-surface decoupling, revealing rotation contrasts through mode splittings in solar-like oscillations. For instance, in a sample of over 2000 RGB stars, core-to-envelope rotation ratios average around 20, with significant scatter highlighting the role of internal transport processes in shaping evolutionary outcomes.

Rotation in Interacting Systems

Tidal Effects in Close Binaries

In close binary systems, the gravitational interaction between components raises tidal bulges on each star, which are slightly misaligned due to the finite response time of the stellar interior to the perturbing potential. This misalignment, arising from internal viscous friction or turbulent dissipation, generates a tidal torque that transfers angular momentum from the orbit to the stellar spin (accelerating rotation) or vice versa (decelerating rotation), depending on whether the stellar rotation is sub- or super-synchronous with the orbital motion. The direction of this transfer acts to drive the system toward rotational synchronization, where the stellar rotation periods match the orbital period, and circularization of the orbit. The equilibrium tide theory provides the foundational framework for modeling these dissipative processes, originally developed by Darwin in his analysis of bodily tides and later refined for stellar contexts. In this approach, the tidal deformation is assumed to adjust quasi-statically to hydrostatic equilibrium, with dissipation occurring through a viscous lag in the bulge position. For stars with radiative envelopes, the characteristic tidal timescale for synchronization follows \tau_\text{tide} \propto (a/R)^8, where a is the semi-major axis and R the stellar radius, emphasizing the rapid efficiency of tides in compact systems where a \approx 10R. For convective envelopes, the dependence weakens to \propto (a/R)^6 due to turbulent viscosity, but the overall formalism highlights how closer separations dramatically shorten timescales, often to millions of years. Prominent examples of tidal effects are seen in RS CVn binaries, evolved systems with late-type giants or subgiants in orbits of 1–10 days, where both components typically exhibit synchronized rotation with P_\text{rot} \approx P_\text{orb} \approx 2–20 days. This synchronization maintains rapid rotation despite the stars' advanced age, countering the spin-down expected from isolated magnetic braking. In such systems, the tidal torques not only align spins but also enhance chromospheric and coronal activity through the dynamo amplification from fast rotation, manifesting as strong Ca II H and K emissions, X-ray flares, and radio bursts. In mass-transferring semidetached binaries, tidal effects interact with accretion to produce over-synchronization, where the mass-gaining component rotates faster than the orbital period. For instance, the illustrates this dynamic: the gainer experiences spin-up from accreted material with specific angular momentum exceeding the orbital value, potentially leading to super-synchronous rotation, though subsequent tidal friction and magnetic braking often reduce it to near-synchronism in observed systems like Algol itself. These processes result in observational signatures of heightened magnetic activity, including asymmetric light curves from starspots and enhanced H\alpha emission, directly linked to the rapid equatorial velocities induced by tides.

Synchronization and Spin-Orbit Coupling

In close binary systems, tidal interactions lead to rotational synchronization, where the stellar spin period aligns with the orbital period, resulting in corotation of the stars with their orbit. This equilibrium state minimizes tidal dissipation and is commonly observed in systems with short orbital periods. For instance, the majority of eclipsing binaries with orbital periods less than 10 days exhibit synchronized rotation, as determined from large photometric surveys. The timescale for achieving synchronization depends on the strength of tidal torques, which scale with the separation between the stars and the structural properties of their envelopes. A characteristic synchronization time scales as t_{\rm sync} \propto \left( \frac{a}{R} \right)^6 for convective envelopes, as derived from equilibrium tide theory, indicating that synchronization occurs rapidly—often within $10^8 years—for close binaries with convective envelopes, such as late-type stars. In binaries with eccentric orbits, full synchronization is not possible due to varying tidal forces over the orbit; instead, pseudosynchronization arises, where the stellar spin angular velocity averages to match the time-averaged orbital motion, particularly near periastron. This state, predicted by dynamical tide models, occurs when the spin rate \Omega_{\rm ps} satisfies \Omega_{\rm ps} = f(e) \Omega_{\rm orb}, with f(e) a function increasing with eccentricity e (e.g., \Omega_{\rm ps} \approx 1.5 \Omega_{\rm orb} for e \approx 0.3). Observations of eccentric "heartbeat" stars confirm that many achieve this equilibrium faster than orbital circularization. Spin-orbit coupling in binary systems governs the alignment of the stellar spin axis with the orbital angular momentum vector, characterized by the obliquity \psi. When misaligned (\psi > 0^\circ), the stellar spin precesses around the orbital axis on a timescale set by the gravitational moment, while drives evolution toward alignment or maintains misalignment in some cases. In systems, which serve as analogs for close stellar binaries, obliquity tides can sustain nonzero \psi through resonant capture, leading to enhanced . For example, the WASP-12b orbits its host star with a projected misalignment of approximately $59^\circ, as measured via the Rossiter-McLaughlin effect, indicating incomplete realignment despite its short orbital period of about 1 day. Such misalignments are observed in roughly 30-50% of around cool stars, highlighting the role of migration history in spin-orbit dynamics.

Rotation in Degenerate Remnants

White Dwarfs

White dwarfs primarily inherit their from the contracting core of their progenitor stars, with this core spin largely preserved during the envelope ejection phase that forms the remnant. Observations indicate that single white dwarfs typically rotate with periods on the order of 1 day, corresponding to equatorial velocities well below 10^4 km/s, as the initial core is diluted but not substantially altered by subsequent processes. White dwarfs in the mass range 0.51–0.73 M⊙ average rotation periods of ~35 hours, while higher-mass white dwarfs show some exceptions with faster . For isolated white dwarfs, rotational braking is minimal, primarily due to the absence of strong magnetic winds, as these remnants possess tenuous atmospheres and generally weak surface below 1 MG. In cases of rapid , gravitational wave emission from the star's oblateness provides a potential spindown , though this is inefficient for the typically slow rotators and becomes relevant only near the breakup limit. This weak braking allows the initial post-formation to persist over the white dwarf's cooling timescale of billions of years. The theoretical upper limit on white dwarf rotation is the surface Keplerian angular velocity, expressed as \Omega_K = \sqrt{\frac{GM}{R^3}}, where M and R are the and , respectively. For a canonical 0.6 M_\odot with R \approx 0.01 R_\odot, \Omega_K \approx 0.35 rad s^{-1}, equivalent to a breakup period of roughly 18 seconds; more massive white dwarfs approach shorter limits, around 3 seconds for 1.4 M_\odot models. Spectroscopic surveys such as the (SDSS) predominantly reveal slow rotators among s, with rotation periods derived from Zeeman splitting in magnetic cases or line profile variations typically exceeding 10 hours for the majority of isolated examples. Fast rotators remain rare, exemplified by the white dwarf in the cataclysmic variable AE Aquarii, which spins with a period of approximately 33 seconds, approaching a significant fraction of its Keplerian limit. In the context of Type Ia supernovae, rapid rotation enables white dwarfs to surpass the nominal 1.44 M_\odot Chandrasekhar mass through centrifugal support at the equator, potentially allowing accretion beyond the non-rotating limit and contributing to overluminous explosions observed in some events. This mechanism is particularly relevant for differentially rotating models, where redistribution sustains up to masses of 1.5 M_\odot or more.

Neutron Stars and Pulsars

Neutron stars form through the core-collapse supernovae of massive stars with initial masses exceeding about 8 solar masses, where the iron-nickel core, rotating with periods typically ranging from 400 seconds to several thousand seconds, undergoes rapid contraction. Conservation of during this compresses the core from roughly 1500 km to 10-15 km in radius, accelerating the to initial periods as short as approximately 1 , equivalent to spin frequencies up to 1000 Hz, though observed birth spins are often in the range of 10-100 ms depending on progenitor profiles. These extreme initial rotations store significant , on the order of 10^{47} erg s, powering early energetic emissions like gamma-ray bursts in some cases. Following formation, experience gradual spin-down primarily due to the electromagnetic from magnetic dipole , which extracts and aligns the magnetic axis with the axis over time. The magnitude of this braking is given by \Gamma = \frac{2}{3} \frac{B^2 R^6 \Omega^4}{c^3} \sin^2 \alpha, where B is the surface magnetic field strength (typically 10^{12}-10^{14} G for pulsars), R \approx 10 km is the radius, \Omega = 2\pi / P is the with period P, c is the , and \alpha is the obliquity angle between the magnetic and axes. This leads to a characteristic spin-down evolution where the period increases as P \propto t^{1/2} for a constant magnetic field and , with characteristic ages ranging from 10^4 to 10^7 years for young to old pulsars. Pulsars, rapidly rotating stars that emit beamed observable as pulses, display a wide range of periods reflecting their evolutionary : ordinary pulsars have periods of 0.1 to 10 seconds, slowing from their birth spins over millennia, while a subset of millisecond pulsars with periods under 10 owe their rapid to "recycling" via prolonged mass accretion from companion stars in low-mass X-ray binaries, which spins them up and weakens their magnetic fields to around 10^8-10^9 . Superposed on this secular slowing are abrupt spin-ups known as glitches, sudden increases in by fractions of a percent (typically \Delta \nu / \nu \sim 10^{-6} to $10^{-3}), attributed to the impulsive transfer of from the pinned superfluid interior to the crust when vortex lines unpin and move outward, a first proposed for explaining the restless of young pulsars like the , which exhibits glitches roughly every 3 years with \Delta \nu / \nu \approx 10^{-6}. In the 2010s, the Neutron Star Interior Composition Explorer (NICER) mission on the has advanced understanding of rotation by precisely measuring masses and radii—such as 1.44 M_\odot and 12.7 km for J0030+0451—through pulse profile modeling, providing constraints on the equation of of supranuclear matter that influence rotational stability and dynamics. These observations link the dense matter properties to rotational phenomena, confirming that softer equations of support more compact stars with potentially higher maximum spin rates before instability.

Black Holes

Black holes formed through the of massive stars inherit significant from the rotating core of their progenitor, resulting in rapidly spinning Kerr black holes rather than non-rotating Schwarzschild ones. The rotation of these black holes is quantified by the dimensionless spin parameter a = \frac{J c}{G M^2}, where J is the black hole's , M its , G the , and c the ; this parameter ranges from 0 (non-rotating) to 1 (maximal prograde spin). High spins, often a > 0.8, arise because the collapsing core conserves much of the progenitor's , with limited loss during the explosion unless asymmetric ejection occurs. Direct measurement of spin is impossible, but it can be inferred from indirect proxies involving accretion and general relativistic effects. One key method analyzes spectra from the , where the iron Kα emission line (at ~6.4 keV) is broadened and asymmetrically shifted due to Doppler boosting, , and near the (ISCO), whose radius shrinks with higher spin. For instance, in the stellar-mass , spectral fitting of relativistic reflection features from and observations yields a high spin of a \gtrsim 0.9, indicating near-extremal . The manifests in observable phenomena tied to the Kerr geometry, including the —a region exterior to the event horizon where is rigidly dragged by the black hole's rotation, forcing infalling objects to co-rotate. Within this zone, the allows extraction of up to 29% of the black hole's rest mass as : a particle entering the decays into two fragments, one with relative to distant observers that falls inward (reducing the black hole's ), while the other escapes with excess . In actively accreting systems, this powers relativistic jets and outflows through the Blandford-Znajek mechanism, where anchored magnetic fields threading the are twisted by , inducing an that accelerates along open field lines and extracts at rates up to \sim 10^{45} erg/s for supermassive black holes. Recent advances in imaging have further constrained black hole spins. The Event Horizon Telescope's 2019 observations of the M87* revealed its event-horizon-scale , a dark region ~42 μas in diameter surrounded by a bright ring of emission, consistent with predictions for a . Modeling of the 's asymmetry and , incorporating spin-dependent ray-tracing, imposes a lower limit of a > 0.5, supporting the presence of significant rotation inherited from the progenitor and necessary to launch the observed jet. As of 2025, updated EHT analyses estimate the spin parameter at approximately a \approx 0.8.

References

  1. [1]
    [PDF] STELLAR ROTATION
    1.1. Historical development. The study of stellar rotation began at the turn of the seventeenth century, when sunspots were observed for the first time ...
  2. [2]
    [PDF] the evolution of rotating stars - arXiv
    The full evolution of the rotation law has been studied with the non–stationary scheme for a 9 M⊙ star by Talon et al (1997), and for stars from 5 to 120 M⊙ in-.
  3. [3]
    [PDF] EFFECTS OF ROTATION ON INTERNAL STRUCTURE AND ...
    Effect of stellar rotation on the mean velocity profiles for rotation periods of 1 day. (solid curves) and 14 days. (dotted curves) at three.
  4. [4]
    Physics, Formation and Evolution of Rotating Stars - SpringerLink
    Book Title: Physics, Formation and Evolution of Rotating Stars · Authors: André Maeder · Series Title: Astronomy and Astrophysics Library · Publisher: Springer ...
  5. [5]
    Towards a unified model of stellar rotation - Oxford Academic
    As a consequence of increasing stellar radius and angular momentum conservation, all models for stellar rotation predict a rapid decay in the surface ...
  6. [6]
    Scaling and Evolution of Stellar Magnetic Activity
    Oct 30, 2023 · Following star formation and disk dispersal, rapid stellar rotation coupled with convection leads to strong magnetic activity, which is ...
  7. [7]
    Evolution of Solar and Stellar Dynamo Theory
    Jun 28, 2023 · We review the evolution and current state of dynamo theory and modelling, with emphasis on the solar dynamo.
  8. [8]
    On the rotation of the stars - ADS - Astrophysics Data System
    On the rotation of the stars. Shajn, G. ;; Struve, O. Abstract. Publication: Monthly Notices of the Royal Astronomical Society. Pub Date: January 1929 ...Missing: stellar Vega line broadening
  9. [9]
    ROTATION PERIODS OF 34030 KEPLER MAIN-SEQUENCE STARS
    Our automated autocorrelation-based method detected rotation periods between 0.2 and 70 days for 34,030 (25.6%) of the 133,030 main-sequence Kepler targets ( ...
  10. [10]
    Quantifying the Limits of TESS Stellar Rotation Measurements with ...
    We find that uncertainties on TESS-derived rotation periods are typically below 3% for stars with periods <10 days. Rotation periods are generally reliable out ...Introduction · TESS Rotation Pipeline · Results · Application to Stellar...
  11. [11]
    Rotational velocities of A-type stars - Astronomy & Astrophysics
    Fourier transforms of several line profiles in the range 4200–4500 Ε are used to derive v sin i from the frequency of the first zero. Statistical analysis of ...
  12. [12]
  13. [13]
    Accurate stellar rotational velocities using the Fourier transform of ...
    We propose a method for measuring the projected rotational velocity vsini with high precision even in spectra with blended lines.
  14. [14]
    Critical study of the distribution of rotational velocities of Be stars
    The statistical corrections found on the distribution of ratios V/Vc for overestimations of Vsini, due to macroturbulent motions and binarity, produce a shift ...
  15. [15]
    A SPECTROSCOPIC STUDY OF FIELD AND RUNAWAY OB STARS
    The resulting TeA, log g, and v sin i from the O star fits are listed in Table 6. Our measurement errors are typically 400 K in TeA, 0.10 in log g, and 10 km s ...
  16. [16]
    [PDF] Determination of Vega's rotational velocity based on the Fourier ...
    May 11, 2021 · The final parameters of Vega's rotation were concluded as ve sin i = 21.6(±0.3) km s−1, ve = 195(±15) km s−1, and i = 6.4(±0.5)◦. Key words: ...
  17. [17]
    Kepler main-sequence solar-like stars: surface rotation ... - Frontiers
    This review summarizes the rotation and magnetic activity properties of the single main-sequence solar-like stars within the Kepler field.<|separator|>
  18. [18]
    [PDF] Imaging the Surface of Altair - arXiv
    Jun 7, 2007 · The study used optical long-baseline interferometry to image Altair, revealing gravity darkening and an angular resolution of ~0.64 ...
  19. [19]
    Measuring the Rotation Speed of Giant Stars from Gravitational ...
    By making a series of radial velocity measurements, one can typically determine v sin i to the same accuracy as the individual radial velocity measurements.
  20. [20]
    Magnetic fields and differential rotation on the pre-main sequence
    Zeeman Doppler imaging (ZDI) (Donati et al. 2003) makes the study of stellar magnetic fields using spectropolarimetry possible and offers a window into the ...
  21. [21]
    Shape of a slowly rotating star measured by asteroseismology
    Nov 16, 2016 · Stars are not perfectly spherically symmetric. They are deformed by rotation and magnetic fields. Until now, the study of stellar shapes has ...
  22. [22]
  23. [23]
    Lesson: Differential Rotation of the Sun
    However, unlike Earth which rotates at all latitudes every 24 hours, the Sun rotates every 25 days at the equator and takes progressively longer to rotate at ...
  24. [24]
    Convective differential rotation in stars and planets – I. Theory
    Differential rotation is one of the key complications in the study of stars. It involves breaking symmetries, structure formation and both heat and momentum ...
  25. [25]
    MODELING OF DIFFERENTIAL ROTATION IN RAPIDLY ROTATING ...
    The reason why differential rotation in rapidly rotating stars is close to the Taylor–Proudman state is that meridional flow does not become fast with large ...Missing: circulation | Show results with:circulation
  26. [26]
    Measuring differential rotation of the K-giant ζ Andromedae
    Surface differential rotation (DR) of a spotted star can be measured by cross-correlation of subsequent Doppler images (Donati & Collier Cameron 1997).Missing: zeta | Show results with:zeta
  27. [27]
    (PDF) The differential rotation of stars - ResearchGate
    PDF | The difference in rotation laws between top and bottom of the solar convection zone as revealed by helioseismology may result from an instability.<|control11|><|separator|>
  28. [28]
    The Sun as a Dynamo | High Altitude Observatory
    Rotational shear also plays an essential role, stretching and amplifying fields into ribbons and sheets of toroidal flux (see figure). This is particularly ...
  29. [29]
    MAGNETIC BRAKING AND PROTOSTELLAR DISK FORMATION
    ... factor of ∼3. This indicates significant loss of angular momentum. This loss can be understood by examining the ratio of toroidal to poloidal magnetic fields.
  30. [30]
    Signatures of magnetic braking in Class 0 protostars
    The strongly magnetized model produces a profile with a steep decrease in the angular momentum, of around one order of magnitude, from the outer envelope (>1000 ...
  31. [31]
    The Role of Magnetic Fields in Protostellar Outflows and Star ...
    Through the magnetic field there is then a back reaction on the disk, forcing material in the disk to lose angular momentum and continue to spiral inwards.
  32. [32]
    Magnetocentrifugal Origin for Protostellar Jets Validated through ...
    Mar 14, 2022 · Jets can facilitate the mass accretion onto the protostars in star formation. They are believed to be launched from accretion disks around the ...
  33. [33]
    Rotation periods for stars of the TW Hydrae association
    We have conducted a photometric study of late-type members of the TW Hydrae association (TWA) and measured the rotation periods for 16 stars in 12 systems. For ...Missing: protostar | Show results with:protostar
  34. [34]
    RADIUS-DEPENDENT ANGULAR MOMENTUM EVOLUTION IN ...
    Angular momentum evolution in low-mass stars is determined by initial conditions during star formation, stellar structure evolution, and the behavior of ...
  35. [35]
    Seismic evidence for a rapidly rotating core in a lower-giant-branch ...
    Jun 14, 2012 · Seismic evidence for a rapidly rotating core in a lower-giant-branch star observed with Kepler. Authors:S. Deheuvels, R. A. Garcia, W. J. ...
  36. [36]
    Spin down of the core rotation in red giants - Astronomy & Astrophysics
    ... core rotation periods are much smaller than observed. The expansion of the convective envelope provides favorable conditions for internal gravity waves to ...
  37. [37]
    Angular momentum redistribution by mixed modes in evolved low ...
    This is performed for a benchmark model of 1.3 M⊙ at three evolutionary stages, representative of the evolved pulsating stars observed by CoRoT and Kepler. We ...Missing: locking | Show results with:locking
  38. [38]
    Rotation velocities of white dwarfs - Astronomy & Astrophysics
    In this paper, the third in a series on white dwarf rotation, we concentrate preferentially on DA white dwarfs with convective atmospheres, i.e. with Teff < 14 ...
  39. [39]
    Asteroseismic measurement of core and envelope rotation rates for ...
    May 20, 2024 · For these stars, the core-to-envelope rotation ratios are around 20 and show a large spread with evolution. Several stars show extremely mild ...Missing: K2 mission decoupling
  40. [40]
    Tidal friction in close-in satellites and exoplanets: The Darwin theory ...
    This report is a review of Darwin's classical theory of bodily tides in which we present the analytical expressions for the orbital and rotational evolution ...
  41. [41]
  42. [42]
    Rotation in close binary stars. - NASA ADS
    ... RS CVn type are to a large degree due to this effect. Synchronism operates in full only in systems in which the components are close to each other ...
  43. [43]
    [PDF] Magnetic activity in close binary systems
    Tidal effects can also work against the slowing down of rotation during stellar evolution, when the star inflates into a red giant. A notable manifestation ...
  44. [44]
    Mass loss out of close binaries - The formation of Algol-type systems ...
    During the slow phase of mass transfer, the rotation of the gainer synchronizes rapidly in the case of strong tidal interaction. For weak tidal interaction, ...
  45. [45]
    Tidal Synchronization and Differential Rotation of Kepler Eclipsing ...
    We report rotation periods for 816 EBs with starspot modulations, and find that 79% of EBs with orbital periods of less than 10 days are synchronized.
  46. [46]
  47. [47]
    Tides in asynchronous binary systems - Astronomy & Astrophysics
    Zahn (1977, 1989) found the characteristic timescale for synchronization as a function of a, the major semi-axis of the orbit; τsync ∼ a6 for stars with ...
  48. [48]
    The Pseudosynchronization of Binary Stars Undergoing Strong Tidal ...
    Highly eccentric binary systems are expected to synchronize and circularize at a rapid rate because their tidal interactions are efficient at dissipating ...
  49. [49]
  50. [50]
    Formation and stellar spin-orbit misalignment of hot Jupiters from ...
    Observed hot Jupiter (HJ) systems exhibit a wide range of stellar spin-orbit misalignment angles. This paper investigates the inward migration of giant planets ...
  51. [51]
    [PDF] Origin of the rotation rates of single white dwarfs - arXiv
    White dwarfs are observed to rotate with typical periods of a day. The main sequence progenitors of these stars are also rotating, and it is generally assumed ...
  52. [52]
    Kepler Observations of 27 Pulsating DA White Dwarfs through K2 ...
    White Dwarf Rotation as a Function of Mass and a Dichotomy of Mode Line Widths: Kepler Observations of 27 Pulsating DA White Dwarfs through K2 Campaign 8.Missing: breakup | Show results with:breakup
  53. [53]
    [gr-qc/9807036] Gravitational waves from rapidly rotating white dwarfs
    Aug 25, 1998 · Abstract page for arXiv paper gr-qc/9807036: Gravitational waves from rapidly rotating white dwarfs. ... white dwarf systems. Comments: 13 ...Missing: braking | Show results with:braking
  54. [54]
    On the evolution of rapidly rotating massive white dwarfs towards ...
    We demonstrate that pre-explosion and pre-collapse conditions of both rigidly and differentially rotating white dwarfs are well established by the present work, ...
  55. [55]
    Magnetic White Dwarfs in the SDSS 100 pc Sample - IOP Science
    Aug 22, 2025 · Abstract. We conduct a model atmosphere analysis on all magnetic white dwarfs (MWDs) in the Sloan Digital Sky Survey (SDSS) 100 pc sample.
  56. [56]
  57. [57]
    THE SPIN RATE OF PRE-COLLAPSE STELLAR CORES
    Sep 3, 2015 · The core rotation rates of massive stars have a substantial impact on the nature of core-collapse (CC) supernovae and their compact remnants ...
  58. [58]
    The Spin Periods and Rotational Profiles of Neutron Stars at Birth
    Aug 22, 2005 · Based on our calculated PNS spins, at ~ 200-300 milliseconds after bounce, and assuming angular momentum conservation, we estimate final neutron ...
  59. [59]
    Chapter 6 Pulsars
    Conservation of angular momentum during collapse increases the rotation rate by about the same factor, 1010, yielding initial rotation periods P0 in the ...
  60. [60]
    Modelling pulsar glitches with realistic pinning forces
    The model is based on the formalism for superfluid neutron stars of Andersson & Comer and on the realistic pinning forces of Grill & Pizzochero. We show that ...<|control11|><|separator|>
  61. [61]
    Pulsar glitches and restlessness as a hard superfluidity phenomenon
    ... Published: 03 July 1975. Pulsar glitches and restlessness as a hard superfluidity phenomenon. P. W. ANDERSON &; N. ITOH. Nature volume 256, pages 25–27 ...
  62. [62]
    Implications of latest NICER data for the neutron star equation of state
    Dec 8, 2024 · The recently published NICER data of PSR J0437-4751 is examined. Including the mass and radius distributions of this pulsar in our data base results in modest ...Missing: mission 2020s
  63. [63]
    Using final black hole spins and masses to infer the formation history ...
    In this paper, we propose a novel technique to constrain the progenitor binary black hole (BBH) formation history using the remnant masses and spins of merged ...2 Creating Mock Bbhs For... · 2.2 Coupling Bhs In Binaries · 2.3 Remnant Masses And Spins<|separator|>
  64. [64]
    effect of stellar rotation on black hole mass and spin - Oxford Academic
    To examine the effect of stellar rotation on the resulting black hole mass and spin, we conduct a one-dimensional general relativistic study.
  65. [65]
    THE EXTREME SPIN OF THE BLACK HOLE IN CYGNUS X-1
    We have determined that Cygnus X-1 contains a near-extreme Kerr black hole with a spin parameter a * > 0.95 (3σ).