Angular velocity is a fundamental concept in physics that describes the rate of change of an object's angular position with respect to time, quantifying how quickly it rotates around an axis.[1] It is mathematically defined as the derivative of the angular displacement θ with respect to time t, expressed as ω = dθ/dt, where θ is typically measured in radians.[2] The standard unit of angular velocity is radians per second (rad/s), reflecting its role in rotational kinematics analogous to linear velocity in translational motion.[1]As a vector quantity, angular velocity ω has both magnitude and direction, with the latter determined by the right-hand rule: if the fingers of the right hand curl in the direction of rotation, the thumb points along the axis in the direction of the vector.[3] For a rigid body rotating about a fixed axis, every point on the body shares the same angular velocity, simplifying the analysis of rotational dynamics.[4] This vector nature enables the application of vector calculus in describing complex motions, such as in planetary orbits or mechanical systems.Angular velocity relates directly to linear (tangential) velocity v through the equation v = r ω, where r is the radial distance from the axis of rotation, linking rotational and translational descriptions of motion.[5] In practical contexts, it is essential for understanding phenomena like the rotation of wheels in vehicles, the spin of celestial bodies, and the dynamics of gyroscopes in engineering applications.[6] Variations in angular velocity lead to angular acceleration, further extending its importance in Newton's laws for rotation.[7]
Fundamentals
Definition and scalar form
Angular velocity in its scalar form describes the rate at which an object rotates about an axis, defined as the time derivative of the angular displacement \theta, expressed as \omega = \frac{d\theta}{dt}, where \theta is measured in radians.[1] This scalar quantity captures the speed of rotation without regard to direction, serving as the rotational analog to linear velocity in kinematics.[6]In the International System of Units (SI), angular velocity is measured in radians per second (rad/s); since the radian is dimensionless, the base unit is effectively the inverse second (s⁻¹), though rad/s is the conventional notation to highlight the angular nature.[8]This definition presupposes familiarity with linear kinematics, where velocity is v = \frac{ds}{dt} and s represents arc length along the path.[6] For instance, in uniform circular motion, a constant scalar angular velocity \omega yields a constant tangential speed given by v = \omega r, with r as the radius of the path.[9]
Vector form
In three-dimensional space, angular velocity is represented as a vector \vec{\omega}, which is a pseudovector whose magnitude |\vec{\omega}| equals the scalar angular speed d\theta/dt, where \theta is the angle of rotation.[8] The direction of \vec{\omega} points along the instantaneous axis of rotation, determined by the right-hand rule: the thumb aligns with \vec{\omega} while the fingers curl in the direction of the body's rotation.[6] This vector formulation extends the scalar concept by incorporating the orientation of the rotationaxis, essential for describing arbitrary rotations beyond planar motion.[3]The angular velocity vector relates to infinitesimal rotations through the equation d\vec{\theta} = \vec{\omega} \, dt, where d\vec{\theta} is the infinitesimal rotation vector, representing an incremental change in orientation over a small time interval dt.[10] This linkage allows \vec{\omega} to parameterize the instantaneous rotational motion as a directed quantity. Unlike the angular displacement vector, which accumulates over finite intervals and does not generally add commutatively due to the non-vector nature of large rotations, \vec{\omega} describes the local rate of change and supports vector addition for composing infinitesimal rotations.[10]As a pseudovector, \vec{\omega} transforms under parity inversion (spatial reflection through the origin) by changing sign, distinguishing it from polar vectors while behaving like a true vector under proper rotations.[11] It is also termed an axial vector in certain contexts, reflecting its origin from cross products of polar vectors, such as in derivations of angular momentum.[12]For small angular displacements, the change in a position vector \vec{r} due to rotation by \vec{\omega} over time dt can be approximated using the infinitesimal form of Rodrigues' rotation formula, where the rotated vector is \vec{r}' \approx \vec{r} + d\vec{\theta} \times \vec{r}.[13] However, composing finite rotations via vector addition of \vec{\omega} is only valid infinitesimally; larger composite rotations are non-commutative, meaning the order of successive rotations affects the final orientation, as rotations do not form a vector space.[10]
Orbital angular velocity for point particles
In two dimensions
In two dimensions, the orbital angular velocity of a point particle confined to the xy-plane is defined as the scalar quantity measuring the rate of rotation around the origin. This scalar \omega can be expressed in terms of the particle's position coordinates ([x, y](/page/X&Y)) and velocity components (v_x, v_y) as\omega = \frac{x v_y - y v_x}{x^2 + y^2},where the denominator is the square of the radial distance r^2 = x^2 + y^2.[14] This expression arises from the kinematic definition of angular velocity as the component perpendicular to the plane of motion. The sign of \omega indicates the direction of rotation: positive for counterclockwise and negative for clockwise motion when viewed from the positive z-axis.For a point particle undergoing uniform circular motion in the plane, with constant tangential speed v and radius r, the angular velocity simplifies to the magnitude \omega = v / r.[15] In this case, the direction is out of the plane along the positive z-axis for counterclockwise rotation, consistent with the right-hand rule. The orbital angular momentum scalar l for the particle of mass m relates kinematically to this as l = m r^2 \omega, providing a measure of the rotational tendency without delving into dynamic forces.This formulation of \omega applies more generally to any curvilinear motion of the point particle in the plane, yielding the instantaneous angular velocity at any position along the path, regardless of whether the trajectory is circular.[14] For example, consider a particle in an elliptical orbit around a central point, such as a planet around a star; here, \omega = d\theta / dt varies along the path, reaching its maximum value at periapsis (closest approach) where the radial distance is minimized and tangential speed is maximized.[16]
In three dimensions
In three dimensions, the orbital angular velocity of a point particle generalizes the two-dimensional case to allow for non-planar trajectories, using a vector formulation that captures both the rate and axis of instantaneous rotation about a chosen origin. The angular velocity vector \boldsymbol{\omega} is defined as\boldsymbol{\omega} = \frac{\boldsymbol{r} \times \boldsymbol{v}}{|\boldsymbol{r}|^2},where \boldsymbol{r} is the positionvector of the particle relative to the origin and \boldsymbol{v} is its linear velocityvector.[17] This expression arises from decomposing the velocity into radial and tangential components, where the tangential component corresponds to rotation about the origin.[17]The magnitude of \boldsymbol{\omega} is |\boldsymbol{\omega}| = \frac{v \sin \phi}{r}, where r = |\boldsymbol{r}|, v = |\boldsymbol{v}|, and \phi is the angle between \boldsymbol{r} and \boldsymbol{v}.[17] This magnitude represents the instantaneous rotational speed, equivalent to the speed of the velocity component perpendicular to \boldsymbol{r} divided by the distance from the origin. The direction of \boldsymbol{\omega} is perpendicular to the plane instantaneously spanned by \boldsymbol{r} and \boldsymbol{v}, determined by the right-hand rule applied to the cross product.[17]For arbitrary motion, \boldsymbol{\omega} specifies the instantaneous axis of rotation passing through the origin, along which the particle appears stationary at that instant, with the motion decomposing into rotation about this axis plus possible radial motion.[18] However, \boldsymbol{\omega} is generally not constant, varying with the particle's trajectory unless the motion is uniform circular orbit perpendicular to a fixed axis.[17]A representative example is a particle following a uniform helical path, such as a charged particle in a uniform magnetic field with velocity components parallel and perpendicular to the field; here, \boldsymbol{\omega} has a constant component along the helix axis (the field direction), with magnitude determined by the cyclotron frequency \omega = qB/m, while the parallel velocity contributes no angular component.This formulation is origin-dependent: changing the reference point alters \boldsymbol{r} and thus \boldsymbol{\omega}, requiring adjustment of the position and velocity vectors relative to the new origin for consistency.[18]
Spin angular velocity for rigid bodies
Definition and properties
In rigid body dynamics, the spin angular velocity \vec{\omega} is defined as the vector that describes the instantaneous rotational motion of the entire rigid body. The linear velocity \vec{v} of any point P in the body relative to a reference point O (such as the center of mass) is then given by\vec{v} = \vec{v}_O + \vec{\omega} \times \vec{r},where \vec{r} is the position vector from O to P, and \vec{v}_O is the velocity of O.[19][20] This relation holds because the rigid body constraint requires that distances between any two points remain constant during motion, ensuring no deformation. The general motion can be decomposed into translation of O plus rotation about an axis through O parallel to \vec{\omega}.[21]A key property of \vec{\omega} is that it is identical for every point in the rigid body at any given instant, independent of the choice of origin, in contrast to the orbital angular velocity of a point particle.[22] The instantaneous axis of rotation is the line parallel to \vec{\omega} along which points instantaneously have zero velocity. Its location relative to the reference point O is displaced by \vec{d} = (\vec{v}_O \times \vec{\omega}) / |\vec{\omega}|^2 from O.[23]For example, in a spinning top, \vec{\omega} aligns with the body's symmetry axis, enabling stable rotation about that axis.[24] Similarly, for a wheel undergoing pure rolling without slipping, \vec{\omega} is directed perpendicular to the plane of the wheel and its magnitude relates to the center's velocity by v_O = \omega r, but the instantaneous axis passes through the contact point.[25]In general rigid body motion, which may combine rotation and translation, the motion corresponds to a screw motion about an instantaneous screw axis, with zero pitch for pure rotation (no translation component parallel to \vec{\omega}).[23]
Determination of components
In the body-fixed frame attached to a rigid body, the spin angular velocity \vec{\omega} is expressed in terms of its components along the body basis vectors \hat{e}_1, \hat{e}_2, \hat{e}_3 as \vec{\omega} = \omega_1 \hat{e}_1 + \omega_2 \hat{e}_2 + \omega_3 \hat{e}_3. These components characterize the instantaneous rotation of the body relative to an inertial frame, with the basis vectors themselves evolving according to the relation \frac{d \hat{e}_i}{dt} = \vec{\omega} \times \hat{e}_i (evaluated in the body frame), a fundamental kinematic relation for rigid rotation that ensures the frame rotates with the body.[26] This decomposition allows for the analysis of rotational dynamics using the body's principal axes, where the components \omega_i directly enter Euler's equations of motion.One common method to compute these components analytically is through Euler angles, which parameterize the orientation of the body frame relative to the inertial frame via successive rotations. For the 3-1-3 Euler angle convention—often used for symmetric bodies like gyroscopes—the components in the body frame are given by:\begin{align*}
\omega_x &= \dot{\phi} \sin \theta \sin \psi + \dot{\theta} \cos \psi, \\
\omega_y &= \dot{\phi} \sin \theta \cos \psi - \dot{\theta} \sin \psi, \\
\omega_z &= \dot{\phi} \cos \theta + \dot{\psi},
\end{align*}where \phi is the precession angle, \theta is the nutation angle, and \psi is the spin angle, with dots denoting time derivatives.[27] These expressions arise from composing the angular velocity contributions of each Euler rotation, projecting them onto the body axes; for instance, \dot{\psi} primarily contributes to the spin along the body z-axis, while \dot{\phi} and \dot{\theta} introduce transverse components modulated by the orientation angles.In experimental or numerical contexts, the components can be determined practically by observing the linear velocities \vec{v}_i of at least three non-collinear points on the body relative to a reference point, solving the overdetermined system \vec{v}_i = \vec{v}_O + \vec{\omega} \times \vec{r}_i (where \vec{r}_i are position vectors from the reference) via least-squares minimization to isolate \vec{\omega}.[28] This approach is particularly useful in applications like robotics or motion capture, where sensor data provides the \vec{v}_i.A illustrative example is the steady precession of a gyroscope, where the body-frame components are time-varying due to the combined effects of high spin \dot{\psi} along the symmetry axis (yielding a dominant \omega_z \approx \dot{\psi}) and slower precession \dot{\phi}, which induces oscillating transverse components \omega_x and \omega_y at the precession frequency.[29] In this case, the nutation \dot{\theta} \approx 0 for steady motion simplifies the expressions, but the precession causes the angular velocity vector to trace a path in the body frame, highlighting the dynamic nature of the components during torque-induced rotation.
Mathematical representations
In coordinate frames
Angular velocity is typically expressed in specific coordinate frames to describe the rotation of a rigid body relative to a reference. The inertial frame, fixed in space and non-rotating, provides components \vec{\omega}^i that remain constant in direction for steady rotations, while the body-fixed frame, attached to the rotating body, yields components \vec{\omega}^b that account for the body's orientation. The transformation between these frames uses the direction cosine matrix (DCM) R, which rotates vectors from the body frame to the inertial frame, such that \vec{\omega}^i = R \vec{\omega}^b. This relation ensures that the magnitude of \vec{\omega} is frame-invariant, but the components differ due to the basis change.[30]The kinematics linking angular velocity to the attitude representation involve the time derivative of the DCM. Specifically, \dot{R} = [\vec{\omega}^i \times] R, where [\vec{\omega}^i \times] denotes the skew-symmetric cross-product matrix formed from \vec{\omega}^i:[\vec{\omega}^i \times] = \begin{pmatrix}
0 & -\omega^i_z & \omega^i_y \\
\omega^i_z & 0 & -\omega^i_x \\
-\omega^i_y & \omega^i_x & 0
\end{pmatrix}.Equivalently, in body-frame components, \dot{R} = R [\vec{\omega}^b \times], allowing computation of orientation evolution from measured body rates. These equations derive from the requirement that vectors fixed in the body frame rotate with the body.[31]Under a coordinate transformation corresponding to a proper rotation, the components of angular velocity transform as those of a vector: if the frame rotates by R, then \vec{\omega}' = R \vec{\omega}. However, as an axial vector (or pseudovector), \vec{\omega} reverses sign under parity inversion but behaves like a polar vector under rotations, preserving its utility in rotational dynamics. This transformation property is essential for consistency across frames in multi-body systems.[32]In non-inertial reference frames rotating with respect to an inertial frame at angular velocity \vec{\Omega}, the effective angular velocity for describing motion includes this frame rotation. For instance, the absolute angular velocity of a body in the rotating frame becomes \vec{\omega} + \vec{\Omega}, influencing fictitious forces like the centrifugal term m \vec{\Omega} \times (\vec{\Omega} \times \vec{r}) and Coriolis term -2m \vec{\Omega} \times \vec{v}. This adjustment is crucial for analyzing dynamics in Earth-fixed coordinates, where \vec{\Omega} is the planet's rotation rate.[33]A practical application occurs in satelliteattitude control, where angular velocity components are derived from gyroscope sensor data using quaternions or DCMs to represent the spacecraft's orientation relative to an inertial frame. Quaternions propagate the attitude via \dot{\mathbf{q}} = \frac{1}{2} \mathbf{q} \otimes [\vec{\omega}^b \times] \mathbf{e}_4, enabling robust control laws that maintain pointing accuracy despite disturbances.[34]To compute orientation from angular velocity measurements, numerical integration methods integrate \vec{\omega} over time, but Euler angles suffer from gimbal lock singularities at certain orientations (e.g., pitch near \pm 90^\circ). Quaternions or rotation vectors circumvent these issues by providing singularity-free representations, with integration schemes like Runge-Kutta applied to the quaternion kinematic equations for high-fidelity simulations in aerospace applications.[35]
Tensor representation
The angular velocity \boldsymbol{\omega} in three-dimensional Euclidean space can be represented by a second-rank antisymmetric tensor \boldsymbol{\Omega}, which facilitates formulations of rotational kinematics using linear algebra. The components of this tensor are defined as \Omega_{ij} = -\epsilon_{ijk} \omega_k, where \epsilon_{ijk} is the Levi-Civita symbol and summation over repeated indices is implied. This construction ensures that the tensor applied to a positionvector \mathbf{r} reproduces the cross product: (\boldsymbol{\Omega} \mathbf{r})_i = \epsilon_{ijk} \omega_j r_k = (\boldsymbol{\omega} \times \mathbf{r})_i.[36]As a skew-symmetric tensor (\boldsymbol{\Omega}^T = -\boldsymbol{\Omega}), \boldsymbol{\Omega} has a vanishing trace, \operatorname{tr}(\boldsymbol{\Omega}) = 0. Its eigenvalues are $0, +i|\boldsymbol{\omega}|, and -i|\boldsymbol{\omega}|, arising from the characteristic equation \lambda^3 + |\boldsymbol{\omega}|^2 \lambda = 0. These properties underscore \boldsymbol{\Omega}'s role as an element of the Lie algebra \mathfrak{so}(3).[37]The tensor \boldsymbol{\Omega} serves as the infinitesimal generator of rotations, where the exponential map \exp(t \boldsymbol{\Omega}) yields a rotation matrix in the special orthogonal group \mathrm{SO}(3). For small t, this approximates as \exp(t \boldsymbol{\Omega}) \approx \mathbf{I} + t \boldsymbol{\Omega}, linking local rotational increments to the global group structure.[38]In continuum mechanics, the velocity gradient tensor \mathbf{L} = \nabla \mathbf{v} decomposes additively into a symmetric rate-of-strain tensor \mathbf{D} = (\mathbf{L} + \mathbf{L}^T)/2 and an antisymmetric spin tensor \mathbf{W} = (\mathbf{L} - \mathbf{L}^T)/2. The spin tensor \mathbf{W} captures the rotational component of the deformation, coinciding with \boldsymbol{\Omega} for rigid-body motions but representing the average rotation rate in deformable continua.[39]A key application appears in fluid dynamics, where the vorticity vector \boldsymbol{\zeta} = \nabla \times \mathbf{v} relates directly to the rotation tensor: the components satisfy \zeta_k = -\epsilon_{kij} W_{ij}, such that the vorticity magnitude corresponds to twice the angular speed encoded in \mathbf{W}. This connection quantifies local fluid element rotation as half the vorticity vector.[40]
Related concepts
Angular acceleration
Angular acceleration is defined as the time derivative of the angular velocity vector, denoted as \boldsymbol{\alpha} = \frac{d\boldsymbol{\omega}}{dt}. This vectorquantity describes the rate at which the angular velocity changes in both magnitude and direction for a rotating rigid body. If the magnitude of \boldsymbol{\omega} varies while its direction remains fixed (as in rotation about a principal axis), \boldsymbol{\alpha} aligns parallel or antiparallel to \boldsymbol{\omega}. Conversely, during precession where the direction of \boldsymbol{\omega} shifts, \boldsymbol{\alpha} includes a component perpendicular to \boldsymbol{\omega}, reflecting the instantaneous change in orientation.[41][22]In an inertial reference frame, the components of angular acceleration for a rigid body are simply \alpha_i = \frac{d\omega_i}{dt} for i = x, y, z. In the body-fixed frame, which rotates with the body, the kinematic expression for the time derivative of \boldsymbol{\omega} accounts for the frame's motion itself, leading to \left( \frac{d\boldsymbol{\omega}}{dt} \right)_{\text{inertial}} = \left( \frac{d\boldsymbol{\omega}}{dt} \right)_{\text{body}} + \boldsymbol{\omega} \times \boldsymbol{\omega}, though the second term vanishes, simplifying to the body-frame rate for the acceleration vector. This formulation emphasizes the kinematic perspective, distinct from dynamic interpretations involving torques.[42]The finite change \Delta \boldsymbol{\omega} over a time interval is exactly \int \boldsymbol{\alpha} \, dt.[43] However, due to the non-commutative nature of vector rotations in three dimensions, the corresponding finite rotation cannot be simply obtained by integrating \boldsymbol{\omega} in a commutative manner; for precise finite rotations, parametrizations like the rotation vector (or Rodrigues vector) are employed to capture the compounded effects of sequential angular changes.[44]Consider an example of a rotor accelerating under a constant axial torque: the resulting angular acceleration is \boldsymbol{\alpha} = \frac{T}{I} \hat{n}, directed along the rotation axis \hat{n}, where T is the torque magnitude and I is the moment of inertia about that axis. This follows analogously from the rotational form of Newton's second law, providing a straightforward kinematic description of the speeding-up motion.[45]The angular acceleration contributes to the linear acceleration of points on the rigid body through the kinematic relation \mathbf{a} = \boldsymbol{\alpha} \times \mathbf{r} + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}), where \mathbf{r} is the position vector from the rotation reference point to the point of interest, and \mathbf{a} is the absoluteacceleration. The term \boldsymbol{\alpha} \times \mathbf{r} represents the tangential acceleration component arising directly from \boldsymbol{\alpha}, while the second term accounts for centripetal effects; for points with relative velocity \mathbf{v}_{\text{rel}} in the body (though zero in pure rigid kinematics), an additional Coriolis term $2 \boldsymbol{\omega} \times \mathbf{v}_{\text{rel}} appears.[46]
Relation to angular momentum
In rigid body dynamics, the angular momentum \vec{L} about the center of mass is linearly related to the angular velocity \vec{\omega} through the inertia tensor \mathbf{I}, expressed as \vec{L} = \mathbf{I} \vec{\omega}. The inertia tensor \mathbf{I} is a symmetric 3×3 matrix that encapsulates the mass distribution of the body relative to the chosen coordinate frame. In the principal axis frame, where \mathbf{I} is diagonal with elements I_1, I_2, I_3, this relation simplifies component-wise to L_i = I_i \omega_i for i = 1, 2, 3.[47][48]For a single point particle of mass m at position \vec{r} from the center of rotation in a rigid body, the linear momentum is \vec{p} = m \vec{v} = m (\vec{\omega} \times \vec{r}), so the angular momentum is \vec{L} = \vec{r} \times \vec{p} = m \vec{r} \times (\vec{\omega} \times \vec{r}). Using the vector triple product identity, this expands to \vec{L} = m [r^2 \vec{\omega} - (\vec{r} \cdot \vec{\omega}) \vec{r}], where r^2 = |\vec{r}|^2. The magnitude simplifies to L = m r^2 \omega_\perp when considering the component of \vec{\omega} perpendicular to \vec{r}, highlighting the perpendicular contribution to rotation.[49][50]Angular velocity \vec{\omega} is fundamentally kinematic, characterizing the instantaneous rotation without regard to mass, whereas angular momentum \vec{L} is dynamic, incorporating the body's mass distribution via \mathbf{I}. Conservation of \vec{L} in the absence of external torques implies that variations in \mathbf{I} (e.g., due to reconfiguration) alter \vec{\omega} to maintain |\vec{L}| = I |\vec{\omega}| constant, as demonstrated by an ice skater who increases spin rate by pulling in their arms, reducing I and thus boosting \omega.[51][48]The dynamical evolution of \vec{\omega} is governed by Euler's rigid body equations in the principal frame:\begin{align*}
I_1 \dot{\omega}_1 + (I_3 - I_2) \omega_2 \omega_3 &= \tau_1, \\
I_2 \dot{\omega}_2 + (I_1 - I_3) \omega_3 \omega_1 &= \tau_2, \\
I_3 \dot{\omega}_3 + (I_2 - I_1) \omega_1 \omega_2 &= \tau_3,
\end{align*}or in vector notation, \mathbf{I} \dot{\vec{\omega}} + \vec{\omega} \times (\mathbf{I} \vec{\omega}) = \vec{\tau}, where \vec{\tau} is the applied torque and \dot{\vec{\omega}} is the angular acceleration. For torque-free motion (\vec{\tau} = 0) of an asymmetric top with distinct principal moments, \vec{\omega} undergoes free precession, tracing a closed path around the fixed \vec{L} direction in the body frame, with the precession rate depending on the differences in I_i.[52][53]