Areal velocity
Areal velocity, also known as sectorial velocity, is a pseudovector in classical mechanics whose magnitude represents the rate at which a position vector from a fixed reference point sweeps out area as a particle moves along a curve.[1] In two-dimensional motion, its magnitude is given by \frac{dA}{dt} = \frac{1}{2} r^2 \omega, where r is the radial distance and \omega is the angular velocity, or equivalently \frac{dA}{dt} = \frac{1}{2} |\mathbf{r} \times \mathbf{v}|, with \mathbf{v} as the particle's velocity.[1] This quantity is constant for particles under central forces, such as gravitational attraction, because the torque is zero, conserving angular momentum \mathbf{L} = m \mathbf{r} \times \mathbf{v}, and thus \frac{dA}{dt} = \frac{L}{2m}.[2] In the context of planetary motion, areal velocity embodies Kepler's second law, which asserts that a line segment joining a planet to the Sun sweeps out equal areas in equal intervals of time, implying uniform areal velocity regardless of the planet's orbital speed variations.[3] This constancy arises from the gravitational force being a central force, which is purely radial and produces no torque, thus conserving the angular component of momentum.[2] For non-planar trajectories, the areal velocity generalizes to a vector perpendicular to the plane of motion, with its magnitude still tied to angular momentum conservation.[2] The concept extends beyond celestial mechanics to any system with central potentials, such as charged particles in electromagnetic fields, where it quantifies the geometric progression of motion.[1]Definition and Formulation
Basic Concept
Areal velocity refers to the rate at which area is swept out per unit time by the line segment, or position vector, connecting a fixed central point to a moving particle in a plane.[4] This measure captures how quickly the particle's path encloses successive regions of space relative to the origin point.[5] The idea of areal velocity emerged in the early 17th century as part of Johannes Kepler's second law of planetary motion, which described how planets traverse equal areas in equal times around the Sun, based on detailed analysis of observational data from Tycho Brahe.[6] Kepler published this law in his 1609 work Astronomia Nova, providing the first quantitative insight into non-uniform orbital speeds while maintaining a constant sweeping rate.[6] To illustrate intuitively, consider a particle undergoing uniform circular motion around the central point: the areal velocity remains constant and equals half the product of the circle's radius and the particle's tangential speed, reflecting the steady enclosure of annular sectors.[4] This geometric simplicity highlights how areal velocity quantifies rotational dynamics without requiring detailed path computations. While areal velocity is fundamentally a planar concept, analyzed in two-dimensional polar coordinates, it can be generalized to higher dimensions as a vector normal to the instantaneous plane of motion.[7] In such cases, its magnitude relates directly to the component of angular momentum in that plane.[4]Mathematical Expression
The instantaneous areal velocity, which measures the rate at which the position vector sweeps out area in a plane, is given in polar coordinates by the expression \frac{dA}{dt} = \frac{1}{2} r^{2} \dot{\theta}, where r is the radial distance from the origin and \dot{\theta} = d\theta/dt is the angular speed.[8] This formula arises from the infinitesimal area element dA = \frac{1}{2} r^{2} d\theta in polar coordinates.[9] The average areal velocity over a finite time interval \Delta t from t_1 to t_2 is defined as \Delta A / \Delta t, where \Delta A is the total area swept by the position vector during that interval.[3] For paths approximated as polygonal segments, such as in numerical simulations of trajectories, \Delta A can be computed using the shoelace formula applied to the sequence of position coordinates.[10] In vector notation, the areal velocity is expressed as the vector \frac{1}{2} \mathbf{r} \times \mathbf{v}, where \mathbf{r} is the position vector and \mathbf{v} is the velocity vector; the magnitude of this vector is \frac{1}{2} |\mathbf{r} \times \mathbf{v}|, corresponding to the scalar areal velocity in the plane perpendicular to the vector.[9] This form highlights its direction along the axis normal to the plane of motion. In two dimensions, areal velocity is scalar, but the vector representation extends naturally to three dimensions, projecting the swept area onto the relevant plane. The areal velocity relates to angular momentum \mathbf{L} via \mathbf{L} = 2m times the areal velocity vector for a particle of mass m.[3] In the International System of Units (SI), areal velocity has dimensions of area per time, typically expressed as square meters per second (m²/s).[11]Relation to Angular Momentum
Derivation from Position and Velocity
The angular momentum \mathbf{L} of a particle of mass m is defined as \mathbf{L} = \mathbf{r} \times \mathbf{p} = m \mathbf{r} \times \mathbf{v}, where \mathbf{r} is the position vector from the origin and \mathbf{v} is the velocity vector.$$] The magnitude of the angular momentum is L = m r v \sin \phi, with \phi the angle between \mathbf{r} and \mathbf{v}, which equals m |\mathbf{r} \times \mathbf{v}|.[12] Consider the infinitesimal area dA swept by the position vector \mathbf{r} in time dt. In polar coordinates, this area is the sector of a circle given by dA = \frac{1}{2} r^2 d\theta, where d\theta is the infinitesimal angular displacement.[ The tangential velocity component is $v_\perp = r \frac{d\theta}{dt}$, so $d\theta = \frac{v_\perp dt}{r}$, and substituting yields $dA = \frac{1}{2} r v_\perp dt$.] Since v_\perp = v \sin \phi = |\mathbf{r} \times \mathbf{v}| / r, the areal velocity follows as \frac{dA}{dt} = \frac{1}{2} |\mathbf{r} \times \mathbf{v}|.[12] Substituting the expression for angular momentum gives \frac{dA}{dt} = \frac{1}{2m} |\mathbf{L}|, demonstrating that the areal velocity is directly proportional to the magnitude of the angular momentum for a particle of mass m.[ In vector form, the areal velocity vector is $\frac{d\mathbf{A}}{dt} = \frac{1}{2} \mathbf{r} \times \mathbf{v} = \frac{\mathbf{L}}{2m}$, perpendicular to the [plane](/page/Plane) of motion.] This relation holds for planar motion, where the direction of \mathbf{L} defines the plane.[12] The time derivative of angular momentum is the torque \boldsymbol{\tau} = \frac{d\mathbf{L}}{dt} = \mathbf{r} \times \mathbf{F}, where \mathbf{F} is the net force on the particle.[ For central forces, $\mathbf{F}$ is parallel to $\mathbf{r}$, so $\boldsymbol{\tau} = 0$ and $\mathbf{L}$ is conserved.] Consequently, the areal velocity \frac{d\mathbf{A}}{dt} = \frac{\mathbf{L}}{2m} remains constant in magnitude and direction under such conditions.[4] In the two-body problem, the motion reduces to an equivalent one-body problem with reduced mass \mu = \frac{m_1 m_2}{m_1 + m_2}, where the relative position \mathbf{r} = \mathbf{r}_1 - \mathbf{r}_2 and relative velocity \mathbf{v} = \dot{\mathbf{r}} are used.[ The [angular momentum](/page/Angular_momentum) becomes $\mathbf{L} = \mu \mathbf{r} \times \mathbf{v}$, and the areal velocity is $\frac{d\mathbf{A}}{dt} = \frac{\mathbf{L}}{2\mu}$, maintaining the same proportional relationship but with $\mu$ replacing $m$.] This formulation applies to systems like planetary orbits, where the central force arises from mutual gravitation.[13]Conservation in Central Force Fields
In central force fields, where the force on a particle is directed along the line connecting it to a fixed center and depends only on the distance from that center, the areal velocity remains constant over time. This conservation arises because such forces produce no torque about the center, as the force vector is parallel to the position vector, resulting in zero cross product. Consequently, the angular momentum of the particle is preserved, directly implying a constant rate at which the position vector sweeps out area.[14][15] Examples of this phenomenon are evident in elliptical orbits under gravitational attraction, such as those of planets around the Sun. Here, the particle's speed increases as it approaches the central body (periapsis) and decreases as it recedes (apoapsis), yet the areal velocity stays uniform, meaning equal areas are swept by the position vector in equal time intervals. This holds similarly for charged particles in Coulomb fields, like electrons orbiting a nucleus in the Bohr model approximation.[16][15] The principle generalizes to any radial potential, including inverse-square laws like gravity or Coulomb forces, as well as other central potentials such as Hooke's law for isotropic harmonic oscillators. Qualitatively, the trajectory confines to a plane, and the constant areal velocity manifests as symmetric "pie slices" of equal area traced out at regular time steps, regardless of the specific form of the radial dependence. This reflects the underlying rotational invariance of the system, a symmetry that forbids torques altering the angular motion.[14][16] However, areal velocity is not conserved in non-central force fields, where torques can arise. For instance, a magnetic field perpendicular to the particle's velocity induces a Lorentz force that is not radial, leading to changes in the swept area rate and precession of the orbit.[15]Applications in Classical Mechanics
Kepler's Second Law of Planetary Motion
Kepler's second law states that a line segment joining a planet to the Sun sweeps out equal areas in equal intervals of time. This principle implies that the areal velocity of the planet relative to the Sun remains constant throughout the orbit, regardless of the elliptical shape.[4] Johannes Kepler empirically discovered this law in 1609 while analyzing precise observational data collected by Tycho Brahe on the motion of Mars. Kepler formulated the law as part of his efforts to describe planetary paths more accurately than previous circular models. Isaac Newton later provided a theoretical explanation in 1687, demonstrating that the law follows from the inverse-square nature of gravitational force, leading to constant areal velocity due to angular momentum conservation.[6][4] For Earth's orbit, the constant areal velocity can be calculated using the orbital parameters: the semi-major axis a \approx 1.496 \times 10^{11} m and the sidereal period T \approx 3.156 \times 10^7 s. For a nearly circular orbit like Earth's (eccentricity e \approx 0.017), the areal velocity is approximately \pi a^2 / T \approx 2.23 \times 10^{15} m²/s, reflecting the steady rate at which area is swept. This constancy arises from the conservation of angular momentum in the central gravitational field of the Sun.[4] The law is commonly visualized in diagrams of an elliptical orbit, where two shaded sectors of equal area—one near perihelion (closest to the Sun) and one near aphelion (farthest)—represent equal time intervals, illustrating how the planet moves faster when closer to the Sun to maintain constant areal velocity.[4]General Orbits and Central Potentials
In central force problems, orbits are classified based on the total energy E of the system relative to the effective potential, with areal velocity remaining constant across all types due to the conservation of angular momentum. For bound orbits with negative total energy (E < 0), the trajectories are elliptical, where the particle oscillates between pericenter and apocenter distances while sweeping out area at a constant rate \frac{dA}{dt} = \frac{L}{2\mu}, with L being the conserved angular momentum and \mu the reduced mass.[17] Unbound orbits occur for zero or positive energy: parabolic trajectories correspond to E = 0, allowing the particle to reach infinity with zero velocity at large distances, while hyperbolic orbits for E > 0 feature asymptotic branches and escape to infinity with finite speed, yet in both cases, the areal velocity \frac{dA}{dt} stays fixed, ensuring equal areas are swept in equal times regardless of the radial motion.[17][18] The effective potential method provides insight into how constant angular momentum governs orbit shapes through the effective potential V_{\text{eff}}(r) = V(r) + \frac{L^2}{2\mu r^2}, where V(r) is the central potential (e.g., gravitational or Coulombic). The centrifugal term \frac{L^2}{2\mu r^2} arises from the conserved L, directly linking to the fixed areal velocity, as L = 2\mu \frac{dA}{dt}; higher L steepens the barrier, favoring more circular orbits, while lower L allows deeper penetration into the potential well for eccentric paths.[17] By solving the radial equation of motion \mu \ddot{r} = -\frac{d V_{\text{eff}}}{dr}, the orbit's conic section type emerges from comparing E to V_{\text{eff}}'s minimum and shape, with the constant areal velocity dictating the angular progression independent of radial variations.[19] Consider example calculations for gravitational central potentials to illustrate: in a circular orbit, where r is constant and eccentricity e = 0, the orbital speed v = \sqrt{\frac{GM}{r}} yields areal velocity \frac{dA}{dt} = \frac{1}{2} r v = \frac{1}{2} \sqrt{GM r}, constant by design.[17] For an eccentric elliptical orbit with semi-major axis a and e > 0, the speed varies from maximum at pericenter v_p = \sqrt{\frac{GM (1+e)}{a(1-e)}} to minimum at apocenter v_a = \sqrt{\frac{GM (1-e)}{a(1+e)}}, yet the areal velocity remains \frac{dA}{dt} = \frac{\pi a b}{T} = \frac{L}{2\mu}, with b = a \sqrt{1 - e^2} the semi-minor axis and T the period, demonstrating invariance despite speed fluctuations.[17] This constancy holds analogously for non-gravitational central potentials, such as inverse-cube forces, where orbit shapes differ but areal velocity persists as fixed.[18] In multi-body systems, the two-body problem under central forces reduces to an equivalent one-body problem with reduced mass \mu = \frac{m_1 m_2}{m_1 + m_2} orbiting the center of mass, preserving the areal velocity formulation \frac{dA}{dt} = \frac{L}{2\mu} where L is now the relative angular momentum.[17] This reduction applies to any central potential, ensuring that classifications (elliptical, parabolic, hyperbolic) and effective potential analyses carry over directly, with the constant areal velocity reflecting the underlying symmetry.[19] Such a framework underlies Kepler's second law as a specific gravitational instance but extends broadly to atomic and scattering problems.[18]Extensions to Other Physical Contexts
Magnetic Dipoles and Current Loops
In the context of electromagnetism, the areal velocity of a charged particle provides a direct connection to the magnetic dipole moment through its relation to orbital angular momentum. For a particle of charge q and mass m undergoing circular motion, the orbital angular momentum \mathbf{L} = m \mathbf{r} \times \mathbf{v} is linked to the magnetic dipole moment by \boldsymbol{\mu} = \frac{q}{2m} \mathbf{L}. This arises because the moving charge constitutes an effective current loop, generating a magnetic field analogous to that of a dipole. Since \mathbf{L} = 2m \frac{d\mathbf{A}}{dt}, where \frac{d\mathbf{A}}{dt} is the vector areal velocity, the magnetic moment becomes \boldsymbol{\mu} = q \frac{d\mathbf{A}}{dt}, highlighting how the rate of area swept by the position vector determines the strength and orientation of the dipole./19:Atoms/19.02:Angular_momentum_and_magnetic_moment) For a current-carrying loop, the magnetic dipole moment is given by \boldsymbol{\mu} = I \mathbf{A}, where I is the current and \mathbf{A} is the area vector perpendicular to the plane of the loop. In the case of a single charged particle orbiting in a circular path of radius r with speed v, the motion mimics a current loop with effective current I = \frac{q v}{2\pi r}, as the charge completes one revolution in period T = \frac{2\pi r}{v}. The loop area is A = \pi r^2, yielding \mu = I A = \frac{q v r}{2}. For perpendicular motion, the areal velocity magnitude is \frac{dA}{dt} = \frac{1}{2} r v, so \mu = q \frac{dA}{dt}, demonstrating proportionality between the magnetic moment and areal velocity. In vector form, this generalizes to \boldsymbol{\mu} = \frac{q}{2} \mathbf{r} \times \mathbf{v}, which mirrors the expression for the areal velocity vector \frac{d\mathbf{A}}{dt} = \frac{1}{2} \mathbf{r} \times \mathbf{v}./05:Magnetism/5.04:Magnetic_Dipole_Moment_and_Magnetic_Dipole_Media) This connection extends to atomic physics in the Bohr model of the hydrogen atom, where electrons occupy quantized circular orbits around the nucleus. The model assumes constant angular momentum L = n \hbar for principal quantum number n, implying a constant areal velocity \frac{dA}{dt} = \frac{L}{2m} = \frac{n \hbar}{2m_e}, where m_e is the electron mass. For an electron (q = -e), the resulting orbital magnetic moment is \boldsymbol{\mu} = -\frac{e}{2m_e} \mathbf{L} = -n \mu_B \hat{n}, with \mu_B = \frac{e \hbar}{2m_e} the Bohr magneton, leading to quantized magnetic moments in discrete steps. This quantization of areal velocity thus underpins the discrete magnetic properties observed in atomic spectra and Zeeman effects./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/08:Atomic_Structure/8.03:Orbital_Magnetic_Dipole_Moment_of_the_Electron)Fluid Flow and Vorticity
In fluid dynamics, areal velocity finds application in describing the rotational and sweeping characteristics of flow fields, particularly through its connection to circulation and vorticity via Stokes' theorem. This theorem relates the line integral of the velocity field around a closed curve to the surface integral of the vorticity over the enclosed area:[ \oint_C \mathbf{v} \cdot d\mathbf{l} = \iint_S (\nabla \times \mathbf{v}) \cdot d\mathbf{A}, where the circulation $\Gamma = \oint_C \mathbf{v} \cdot d\mathbf{l}$ quantifies the net rotational flow around the curve, and [vorticity](/page/Vorticity) $\boldsymbol{\omega} = \nabla \times \mathbf{v}$ measures the local [rotation](/page/Rotation) rate of fluid elements.[](https://www.ess.uci.edu/~yu/class/ess228/lecture.4.vorticity.all.pdf) For a small fluid element in solid-body [rotation](/page/Rotation) with [angular velocity](/page/Angular_velocity) $\Omega$, the [vorticity](/page/Vorticity) magnitude is $\omega = 2 \Omega$, twice the angular velocity, distinguishing it from pure [translation](/page/Translation) where $\omega = 0$.[](https://www.me.psu.edu/cimbala/me320/Lesson_Notes/Fluid_Mechanics_Lesson_04C.pdf) The analogy between vorticity and areal velocity emerges in rotational flows, where the areal velocity $dA/dt = \frac{1}{2} |\mathbf{r} \times \mathbf{v}|$ for a fluid particle relates to the local [angular](/page/Angular) motion. In solid-body [rotation](/page/Rotation), $dA/dt = \frac{1}{2} \Omega r^2$, so $\Omega = 2 (dA/dt) / r^2$ and $\omega = 4 (dA/dt) / r^2$, highlighting how [vorticity](/page/Vorticity) captures twice the effective [rotation](/page/Rotation) implied by areal velocity in particle-like descriptions. This relation underscores [vorticity](/page/Vorticity)'s role as a distributed measure of [rotation](/page/Rotation), analogous to how constant areal velocity signals conserved [angular momentum](/page/Angular_momentum) in central-force problems for point particles. For point vortices in [2D](/page/2D) inviscid flows, the velocity field is $\mathbf{v} = \frac{\Gamma}{2\pi r} \hat{e}_\theta$ (azimuthal component, irrotational except at the [singularity](/page/Singularity)), inducing [circular motion](/page/Circular_motion) on test particles.[](https://web.stanford.edu/~cantwell/AA200_Course_Material/AA200_Lectures/AA200_Ch_10_Elements_of_potential_flow_Vortex_Sticks_and_Stokes_Flow_Brian_Cantwell.pdf) Here, the areal velocity is constant at $dA/dt = \Gamma / (4\pi)$, independent of radial distance, mirroring Kepler's second law due to the inverse-linear velocity profile and constant circulation $\Gamma$.[](https://math.mit.edu/classes/18.325/PVarticle.pdf) In irrotational flows ($\omega = 0$), areal velocity varies with the [flow](/page/Flow) geometry and reference point. For instance, in uniform [flow](/page/Flow), a [test particle](/page/Test_particle) moves rectilinearly, yielding $dA/dt = \frac{1}{2} d v \sin\theta$ (where $d$ is the [perpendicular distance](/page/Perpendicular_distance) from the reference to the streamline and $\theta$ the [angle](/page/Theta)), which changes along the path unless the reference lies on the streamline. In contrast, potential vortices maintain steady areal sweep via conserved circulation, ensuring constant $dA/dt$ on circular streamlines despite zero [vorticity](/page/Vorticity) away from the core. This implies that irrotational circulatory flows can exhibit Kepler-like areal constancy without distributed rotation, relying instead on global topology. For 2D incompressible flows, areal velocity manifests through the [stream function](/page/Stream_function) $\psi(x,y)$, defined such that [velocity](/page/Velocity) components are $u = \partial \psi / \partial y$, $v = -\partial \psi / \partial x$, satisfying [continuity](/page/Continuity) automatically. In [steady state](/page/Steady_state), the difference $\Delta \psi$ between two streamlines equals the constant [volume flow rate](/page/Flow_rate) per unit depth (areal [flux](/page/Flux)) crossing any segment connecting them, conserving the "swept area" rate along streamlines.[](https://web.mit.edu/16.unified/www/FALL/fluids/Lectures/f15.pdf) This conservation holds for inviscid, barotropic flows under [Kelvin's circulation theorem](/page/Kelvin's_circulation_theorem), where $\Gamma$ remains constant for material contours, linking to steady areal velocity in non-dissipative regimes. Unlike point-particle [mechanics](/page/Mechanics), where areal velocity directly ties to the [specific angular momentum](/page/Specific_angular_momentum) $L/m = 2 dA/dt$ for a concentrated [mass](/page/Mass) under central forces, [fluid dynamics](/page/Fluid_dynamics) involves distributed [mass](/page/Mass) density and [vorticity](/page/Vorticity) fields. There is no single equivalent [angular momentum](/page/Angular_momentum) per particle; instead, [enstrophy](/page/Enstrophy) $\int \omega^2 dV$ or total circulation governs rotational dynamics, allowing diffusion, stretching, and tilting of [vorticity](/page/Vorticity) in viscous or compressible flows without direct areal velocity [conservation](/page/Conservation). This distributed nature enables phenomena like vortex sheets or stretching in [3D](/page/3D), absent in discrete particle orbits.