Fact-checked by Grok 2 weeks ago

Rotating reference frame

A rotating reference frame is a non-inertial that rotates at a relative to an inertial frame, where the standard form of does not apply directly due to the frame's . In such frames, the observed motion of objects appears altered, requiring the inclusion of fictitious forces to reconcile descriptions with physical reality. These frames are fundamental in physics for analyzing systems involving rotation, such as planetary motion, engineering designs, and geophysical phenomena. The transformation between an inertial frame and a rotating frame involves relating , , and vectors across the frames. The in the inertial frame is given by \mathbf{v} = \mathbf{v}' + \boldsymbol{\Omega} \times \mathbf{r}, where \mathbf{v}' is the relative to the rotating frame, \boldsymbol{\Omega} is the angular vector, and \mathbf{r} is the vector. transforms as \mathbf{a} = \mathbf{a}' + \boldsymbol{\Omega} \times (\boldsymbol{\Omega} \times \mathbf{r}) + 2 \boldsymbol{\Omega} \times \mathbf{v}', introducing terms that correspond to the centrifugal and Coriolis effects. In the rotating frame, Newton's second law becomes m \mathbf{a}' = \mathbf{F} - m \boldsymbol{\Omega} \times (\boldsymbol{\Omega} \times \mathbf{r}) - 2 m \boldsymbol{\Omega} \times \mathbf{v}', where the additional terms act as fictitious forces. The , -m \boldsymbol{\Omega} \times (\boldsymbol{\Omega} \times \mathbf{r}), points radially outward from the axis of rotation and depends on the distance from that axis, contributing to effects like the of rotating planets. The Coriolis force, -2m \boldsymbol{\Omega} \times \mathbf{v}', is perpendicular to both the and the object's velocity in the rotating frame, causing deflections such as the rotation of hurricanes (counterclockwise in the ). These forces, though not real interactions, enable accurate predictions of motion within the rotating frame and are crucial in fields like , , and .

Fundamentals of Reference Frames

Inertial versus Non-Inertial Frames

In , an inertial reference frame is defined as one in which the motion of a body not subject to external forces is rectilinear and uniform, allowing to hold without modification. This frame moves at constant velocity relative to the , providing a standard against which other motions can be measured, as the distant stars serve as an approximate inertial backdrop due to their vast separation and minimal relative acceleration. In such frames, the first law of motion states that an object at rest remains at rest, and an object in motion continues in a straight line at constant speed unless acted upon by a . Non-inertial reference frames, by contrast, undergo relative to an inertial frame, either linearly or rotationally, causing Newton's laws to appear invalid without additional corrections. Examples include a accelerating forward, where objects inside seem to press backward against the seats, or a merry-go-round rotating steadily, where riders experience an outward tendency. In these frames, observers perceive apparent or fictitious forces that account for the observed deviations from inertial motion, effectively restoring the form of Newton's second law by including these pseudo-forces as if they were real interactions. The concept of inertial frames traces back to Newton's Philosophiæ Naturalis Principia Mathematica (1687), where he introduced absolute space as an unchanging, immovable backdrop independent of external relations, serving as the foundation for true motion and the validity of his laws. Newton argued that relative motions alone could not distinguish absolute rest from uniform motion, but absolute space provided the necessary inertial structure, with the offering a practical reference for identifying such frames. This framework resolved earlier debates on motion by positing that only in absolute space do free particles follow straight paths without forces. A qualitative illustration of non-inertial effects occurs in an accelerating : a released from rest relative to the elevator appears to accelerate downward faster than alone would predict in an inertial frame, as if an additional backward force acts on it to maintain the form of Newton's laws within the elevator's perspective. Rotating reference frames represent a specific subclass of non-inertial frames, where the acceleration arises from , leading to curved paths for free particles as observed from the rotating system.

Defining a Rotating Reference Frame

A rotating reference frame is a non-inertial that rotates with a \vec{\omega} relative to an inertial reference frame. This setup typically employs Cartesian coordinates with the origin located at the axis of rotation, where the z-axis of the rotating frame aligns with the direction of \vec{\omega}. The rotation is assumed to be rigid, meaning all points in the frame maintain fixed relative positions as the entire system rotates uniformly. The angular velocity vector \vec{\omega} specifies both the magnitude \omega (in radians per second) and the direction of the rotation axis, with the direction determined by the right-hand rule: curling the fingers of the right hand in the direction of rotation points the thumb along \vec{\omega}./9:_Rotational_Kinematics_Angular_Momentum_and_Energy/9.7:_Vector_Nature_of_Rotational_Kinematics) A key assumption is that the rotation is steady, with constant \vec{\omega} in both magnitude and direction, which simplifies the kinematic description. Conceptually, consider a point fixed in the rotating frame; in the inertial frame, this point traces a circular path centered on the , with the radius equal to its from the and the period of motion given by $2\pi / \omega. This circular trajectory illustrates the relative motion between the frames. A practical example is the , which serves as an approximately rotating reference frame for local observations, with \vec{\omega} directed along its north-south and magnitude \omega \approx 7.29 \times 10^{-5} rad/s. Such frames, being non-inertial, require the introduction of fictitious forces to describe dynamics accurately.

Fictitious Forces in Rotating Frames

Centrifugal Force

In a rotating reference frame, the centrifugal force is a fictitious force acting on an object of mass m, expressed in vector form as \vec{F}_\text{cent} = -m \vec{\omega} \times (\vec{\omega} \times \vec{r}), where \vec{\omega} is the angular velocity vector of the frame and \vec{r} is the position vector of the object relative to the origin on the rotation axis. This expression simplifies using the vector triple product identity, yielding a force directed radially outward from the axis of rotation, perpendicular to \vec{\omega}. The magnitude of the force is F_\text{cent} = m \omega^2 \rho, where \rho = |\vec{r}_\perp| is the perpendicular distance from the rotation axis, highlighting its dependence solely on position and the frame's rotation rate, independent of the object's velocity in the rotating frame. The originates from the inertial tendency of objects to maintain straight-line motion in an inertial frame, which, when observed from the accelerating rotating frame, appears as an outward deflection away from the axis. This apparent force ensures consistency with the conservation of : as an object moves relative to the rotating frame, its path curves outward to preserve the it holds in the inertial frame, mimicking an expansive push. Unlike real forces from interactions, the has no physical source but emerges purely from the non-inertial nature of the frame. A practical illustration occurs on , where produces a at the of approximately $0.034 \, \text{m/s}^2, reducing the effective from $9.832 \, \text{m/s}^2 (polar value) to about $9.780 \, \text{m/s}^2, a decrease of roughly 0.3% in . This effect diminishes toward the poles, where \rho = 0 and the force vanishes. Although the is non-conservative in general rotating frames where \vec{\omega} may vary, for cases of , it derives from a V_\text{cent} = -\frac{1}{2} m \omega^2 \rho^2, allowing along paths independent of and confirming its conservative character under steady ./29%3A_Non-Inertial_Frame_and_Coriolis_Effect/29.02%3A_Uniformly_Rotating_Frame) This potential facilitates analysis of and motion in systems like rotating fluids or planetary atmospheres.

Coriolis Force

The is a velocity-dependent that arises in a rotating reference frame with . It acts on an object of m with \vec{v} relative to the rotating frame and is mathematically expressed as -2m \vec{\omega} \times \vec{v}, where \vec{\omega} is the vector of the frame. This force is always directed perpendicular to both \vec{v} and \vec{\omega}, resulting in a deflection of the object's path without altering its speed in the rotating frame./12%3A_Non-inertial_Reference_Frames/12.08%3A_Coriolis_Force) The magnitude of the force is given by $2m \omega v \sin\theta, where \theta is the angle between \vec{v} and \vec{\omega}./03%3A_The_Coriolis_Force) In physical terms, the Coriolis force causes a deflection of moving objects that appears to the right of their velocity vector in the due to and to the left in the . This deflection is a consequence of the frame's and is most pronounced for motions perpendicular to the rotation axis. Because the force is perpendicular to the velocity, it performs no work on the object, thereby conserving while redirecting and altering the trajectory. The effect vanishes at the , where \sin\theta = 0 for horizontal motions, and maximizes at the poles./03%3A_The_Coriolis_Force) A classic example is the eastward deflection of objects in free fall on , where the causes the path to deviate from vertical due to the planet's ; for a drop from rest at mid-s, this eastward shift can be on the order of centimeters for heights of tens of meters. Another demonstration is the , whose plane of oscillation at a rate \Omega = \omega \sin\phi, with \phi denoting the , visibly rotating once per day at the poles and slower elsewhere. This rate directly reflects the local vertical component of 's . The is named after , who first quantified it in 1835 while analyzing the energy transfer in rotating machinery such as waterwheels, introducing the necessary corrections to Newton's laws for such systems.

Euler Force

The is a that arises in a rotating reference frame when the angular velocity \boldsymbol{\omega} is not constant, specifically due to the angular acceleration \dot{\boldsymbol{\omega}}. It is defined as \mathbf{F}_E = -m \dot{\boldsymbol{\omega}} \times \mathbf{r}, where m is the mass of the object, \dot{\boldsymbol{\omega}} is the time derivative of the angular velocity vector, and \mathbf{r} is the position vector from the rotation axis to the object. This force accounts for the effects of torque-induced changes in the frame's rotation rate, appearing only when \dot{\boldsymbol{\omega}} \neq 0. In the inertial frame, a real torque \boldsymbol{\tau} produces \dot{\boldsymbol{\omega}} = \boldsymbol{\tau}/I for a rigid body with moment of inertia I; in the rotating frame, this manifests as the Euler force, which adjusts Newton's second law to include this term for apparent equilibrium. The magnitude of the Euler force is m |\dot{\boldsymbol{\omega}}| r \sin \theta, where \theta is the angle between \dot{\boldsymbol{\omega}} and \mathbf{r}, and its direction is perpendicular to both \dot{\boldsymbol{\omega}} and \mathbf{r}, following the right-hand rule for the cross product. A representative example occurs in a spinning top slowing due to , where \dot{\boldsymbol{\omega}} points opposite to the spin axis, producing a tangential that contributes to the top's wobbling or loss of stability as the rotation rate decreases. The is typically absent in steady-state analyses of constant-\boldsymbol{\omega} rotations but is essential for scenarios involving variable rotation, such as the of planetary axes, where the gradual change in \boldsymbol{\omega}'s direction (e.g., Earth's over 26,000 years) introduces a small but nonzero \dot{\boldsymbol{\omega}}, influencing long-term orbital dynamics. In such cases, the force's scale is often negligible compared to gravitational effects but provides critical insight into non-steady rotational motion.

Mathematical Relations Between Frames

Position and Coordinate Transformations

In a rotating reference frame undergoing pure rotation with constant angular velocity \vec{\omega} relative to an inertial frame, the origins of both frames are assumed to coincide, with no relative translation between them. The position vector \vec{r} of any point is identical as a physical entity in both frames, but its representation in coordinates differs due to the orientation of the basis vectors. The components in the inertial frame \vec{r} are related to those in the rotating frame \vec{r}' by the time-dependent rotation matrix R(t), such that \vec{r} = R(t) \vec{r}', where the rotation angle is \theta = \omega t for rotation about a fixed axis. The rotation matrix R(t) is orthogonal, satisfying R^T(t) R(t) = I, which preserves vector lengths and angles under the transformation. Consequently, the inverse relation is \vec{r}' = R^{-1}(t) \vec{r} = R^T(t) \vec{r}, allowing coordinates in the rotating frame to be obtained directly from inertial coordinates. For rotation about the z-axis with angular speed \omega, the explicit coordinate transformations are given by \begin{align*} x' &= x \cos(\omega t) - y \sin(\omega t), \\ y' &= x \sin(\omega t) + y \cos(\omega t), \\ z' &= z. \end{align*} These equations express the rotating-frame coordinates (x', y', z') in terms of the inertial-frame coordinates (x, y, z). A representative example is a point fixed at (x, y, z) = (a, 0, 0) in the inertial frame. In the rotating frame, this point appears to move in a circle of radius a around the z-axis, with coordinates x' = a \cos(\omega t), y' = a \sin(\omega t), z' = 0. This apparent circular motion illustrates how the rotation of the frame induces perceived movement for stationary points. The position mapping forms the basis for deriving velocity transformations via time differentiation.

Time Derivatives and Angular Velocity

In a rotating reference frame, the time derivative of a vector \mathbf{A} differs from that in an inertial frame due to the frame's rotation. The fundamental relation is given by \left( \frac{d\mathbf{A}}{dt} \right)_{\text{inertial}} = \left( \frac{d\mathbf{A}}{dt} \right)_{\text{rot}} + \boldsymbol{\omega} \times \mathbf{A}, where \left( \frac{d\mathbf{A}}{dt} \right)_{\text{rot}} is the derivative as measured in the rotating , and \boldsymbol{\omega} is the of the rotating frame relative to the inertial frame. This equation, known as the , applies to any vector quantity and accounts for the additional rotational contribution. The \boldsymbol{\omega} is a directed along the of , with equal to the instantaneous of d\theta/dt. In three dimensions, it has components \omega_x, \omega_y, and \omega_z along the respective axes of the inertial frame. For a frame rotating with constant angular speed about a fixed , \boldsymbol{\omega} remains constant in both and direction. To derive the transport theorem, consider the position vector \mathbf{r} of a point, which transforms between frames via a time-dependent \mathbf{R}(t), such that \mathbf{r}_{\text{inertial}} = \mathbf{R}(t) \mathbf{r}_{\text{rot}}. Differentiating with respect to time using the chain rule yields \left( \frac{d\mathbf{r}}{dt} \right)_{\text{inertial}} = \dot{\mathbf{R}} \mathbf{r}_{\text{rot}} + \mathbf{R} \left( \frac{d\mathbf{r}}{dt} \right)_{\text{rot}}. The term \dot{\mathbf{R}} \mathbf{R}^{-1} is the representation of \boldsymbol{\omega}, leading to the cross-product form \dot{\mathbf{R}} \mathbf{r}_{\text{rot}} = \boldsymbol{\omega} \times \mathbf{r}_{\text{rot}}, and thus the general relation for any vector. For scalar quantities, which lack directional dependence, the time is of the : \frac{d\phi}{dt}_{\text{inertial}} = \frac{d\phi}{dt}_{\text{rot}}. This follows directly from the absence of a cross-product term. The also applies to the unit basis vectors \hat{\mathbf{e}}' of the rotating , which evolve as \frac{d\hat{\mathbf{e}}'}{dt}_{\text{inertial}} = \boldsymbol{\omega} \times \hat{\mathbf{e}}'. In the rotating itself, these basis vectors appear fixed, so \left( \frac{d\hat{\mathbf{e}}'}{dt} \right)_{\text{rot}} = 0, highlighting the rotational contribution.

Velocity Transformations

In a rotating reference frame with angular velocity \vec{\omega} relative to an inertial frame, the velocity of a particle transforms according to the relation \vec{v} = \vec{v}' + \vec{\omega} \times \vec{r}, where \vec{v} is the velocity in the inertial frame, \vec{v}' is the velocity as measured in the rotating frame, and \vec{r} is the position vector from the common origin of the frames (assuming the origins coincide and are fixed). This transformation arises from the general rule for time derivatives in rotating frames: the inertial derivative of a vector \vec{A} is \left( \frac{d\vec{A}}{dt} \right)_{\text{inertial}} = \left( \frac{d\vec{A}}{dt} \right)_{\text{rotating}} + \vec{\omega} \times \vec{A}. Applying this to the position vector \vec{r} yields \vec{v} = \frac{d\vec{r}}{dt}_{\text{inertial}} = \frac{d\vec{r}}{dt}_{\text{rotating}} + \vec{\omega} \times \vec{r} = \vec{v}' + \vec{\omega} \times \vec{r}, since \vec{v}' = \frac{d\vec{r}}{dt}_{\text{rotating}}. The term \vec{\omega} \times \vec{r} represents the additional velocity imparted by the rotation of the frame itself, which is perpendicular to both \vec{\omega} and \vec{r} due to the cross-product operation, and has magnitude \omega r \sin\theta, where \theta is the angle between \vec{\omega} and \vec{r}. This perpendicularity ensures that the rotational contribution does not alter the radial component of velocity but adds a tangential component. For steady rotation about a fixed axis, the transformation can be expressed in components aligned with the rotating frame's coordinates, though the vector form suffices for most analyses. Consider a particle at rest in the rotating frame, so \vec{v}' = 0; in the inertial frame, its velocity is then purely \vec{\omega} \times \vec{r}, describing uniform circular motion around the rotation axis with speed \omega r \sin\theta. This example illustrates how observers in the rotating frame might perceive the particle as stationary, while inertial observers see the motion induced by the frame's rotation. The velocities \vec{v} and \vec{v}' are vectors with units of length per time, and the cross product maintains dimensional consistency.

Acceleration Transformations

The acceleration of a particle in an inertial \mathbf{a}_I relates to its in a rotating \mathbf{a}_R through the transformation \mathbf{a}_I = \mathbf{a}_R + 2 \boldsymbol{\omega} \times \mathbf{v}_R + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) + \dot{\boldsymbol{\omega}} \times \mathbf{r}, where \mathbf{v}_R is the relative to the rotating , \mathbf{r} is the from the rotation , \boldsymbol{\omega} is the of the rotating relative to the inertial , and \dot{\boldsymbol{\omega}} is the of the rotating . This equation arises in for analyzing motion across frames rotating with respect to each other. In this relation, \mathbf{a}_R represents the acceleration as measured by an observer in the rotating frame, while the additional terms account for effects induced by the frame's rotation and any change in its angular velocity. These frame-induced terms modify the apparent dynamics without altering the underlying physics in the inertial frame. To derive this transformation, begin with the velocity relation between the frames: \mathbf{v}_I = \mathbf{v}_R + \boldsymbol{\omega} \times \mathbf{r}. The time derivative of any vector \mathbf{C} in the inertial frame follows the rule \left( \frac{d\mathbf{C}}{dt} \right)_I = \left( \frac{d\mathbf{C}}{dt} \right)_R + \boldsymbol{\omega} \times \mathbf{C}, known as the transport theorem or Coriolis theorem. Apply this rule to differentiate \mathbf{v}_I: \mathbf{a}_I = \left( \frac{d\mathbf{v}_I}{dt} \right)_I = \left( \frac{d\mathbf{v}_I}{dt} \right)_R + \boldsymbol{\omega} \times \mathbf{v}_I. Substitute \mathbf{v}_I = \mathbf{v}_R + \boldsymbol{\omega} \times \mathbf{r} into the rotating-frame derivative: \left( \frac{d\mathbf{v}_I}{dt} \right)_R = \left( \frac{d}{dt} \right)_R (\mathbf{v}_R + \boldsymbol{\omega} \times \mathbf{r}) = \mathbf{a}_R + \dot{\boldsymbol{\omega}} \times \mathbf{r} + \boldsymbol{\omega} \times \mathbf{v}_R, assuming differentiation acts component-wise in the rotating frame. Now substitute back and expand \boldsymbol{\omega} \times \mathbf{v}_I = \boldsymbol{\omega} \times \mathbf{v}_R + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}), yielding the full acceleration relation after collecting terms. This second application of the derivative rule introduces the cross-product terms absent in the first derivative (velocity transformation). The key terms in the transformation are the Coriolis-like acceleration $2 \boldsymbol{\omega} \times \mathbf{v}_R, which depends on the velocity in the rotating frame; the centrifugal-like term \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}), directed radially outward from the ; and the Euler-like term \dot{\boldsymbol{\omega}} \times \mathbf{r}, arising from changes in the . These terms highlight how distorts measurements between frames. For example, consider a particle undergoing linear motion with constant velocity \mathbf{v}_R in the rotating frame, so \mathbf{a}_R = 0. In the inertial frame, the path curves due to the nonzero frame-induced accelerations $2 \boldsymbol{\omega} \times \mathbf{v}_R + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) + \dot{\boldsymbol{\omega}} \times \mathbf{r}, assuming \dot{\boldsymbol{\omega}} = 0 for steady rotation. This illustrates the kinematic coupling between frames.

Dynamics in Rotating Frames

Newton's Second Law Adaptation

In an inertial reference frame, Newton's second law states that the net real force acting on a particle of mass m is equal to the mass times the acceleration observed in that frame:
\mathbf{F}_\text{real} = m \mathbf{a}_\text{inertial}.
When transitioning to a non-inertial reference frame rotating with angular velocity \boldsymbol{\omega} relative to the inertial frame, the observed acceleration \mathbf{a}_\text{rot} in the rotating frame requires modification to the law to account for the frame's motion. The adapted form incorporates additional terms representing fictitious forces:
\mathbf{F}_\text{real} - m \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) - 2 m \boldsymbol{\omega} \times \mathbf{v}_\text{rot} - m \frac{d \boldsymbol{\omega}}{dt} \times \mathbf{r} = m \mathbf{a}_\text{rot},
where \mathbf{r} is the position vector from the rotation origin, \mathbf{v}_\text{rot} is the velocity relative to the rotating frame, and d \boldsymbol{\omega}/dt is the angular acceleration of the frame.
This equation interprets the dynamics such that the real forces, augmented by the fictitious centrifugal, Coriolis, and Euler forces, balance the times the measured in the rotating frame. The fictitious terms arise solely from the of the frame transformation and vanish in inertial frames, allowing 's second law to be applied as if the rotating frame were inertial. The adaptation assumes a classical mechanical framework with no relativistic effects, treating the rotation as rigid and the masses as point particles without quantum considerations. It holds for moderate angular velocities where special relativity is negligible. A representative example is the analysis of planetary motion in a frame rotating with the angular velocity of the planet's orbit around the Sun, centered on the Sun. For a circular orbit, the planet appears stationary in this frame (\mathbf{a}_\text{rot} = 0, \mathbf{v}_\text{rot} = 0), so the gravitational force balances the centrifugal term: \mathbf{F}_\text{grav} = m \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}), yielding GMm / r^2 = m \omega^2 r and Kepler's third law relation \omega^2 = GM / r^3.

Derivation of Fictitious Forces from Transformations

To derive the fictitious forces in a rotating reference frame, begin with Newton's second law in the inertial frame, where the physical \mathbf{F} equals m times the absolute acceleration \mathbf{a}_I: \mathbf{F} = m \mathbf{a}_I. \tag{1} This holds for any system, regardless of the observer's frame. The absolute acceleration \mathbf{a}_I relates to quantities measured in the rotating frame through the kinematic derived from differentiating and vectors while accounting for the frame's rotation. Assuming the origins of the inertial and rotating frames coincide (or any translational of the origin is separately treated), the vector \mathbf{r} is the same in both frames, the relative is \mathbf{v}_{rot} = \frac{d\mathbf{r}}{dt}\big|_{rot}, and the angular is \boldsymbol{\omega}. The is \mathbf{v}_I = \mathbf{v}_{rot} + \boldsymbol{\omega} \times \mathbf{r}. Differentiating again yields the : \mathbf{a}_I = \mathbf{a}_{rot} + \dot{\boldsymbol{\omega}} \times \mathbf{r} + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) + 2 \boldsymbol{\omega} \times \mathbf{v}_{rot}, \tag{2} where \mathbf{a}_{rot} = \frac{d\mathbf{v}_{rot}}{dt}\big|_{rot} is the relative and \dot{\boldsymbol{\omega}} = \frac{d\boldsymbol{\omega}}{dt} is the angular of the frame. This equation arises from applying the differentiation rule in rotating frames twice. Substitute Equation (2) into Equation (1): \mathbf{F} = m \left[ \mathbf{a}_{rot} + \dot{\boldsymbol{\omega}} \times \mathbf{r} + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) + 2 \boldsymbol{\omega} \times \mathbf{v}_{rot} \right]. \tag{3} Rearranging to isolate the relative acceleration \mathbf{a}_{rot} gives the adapted form of Newton's second law in the rotating frame: m \mathbf{a}_{rot} = \mathbf{F} - m \dot{\boldsymbol{\omega}} \times \mathbf{r} - m \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) - 2 m \boldsymbol{\omega} \times \mathbf{v}_{rot}. \tag{4} Here, \mathbf{F} remains the physical (real) force, while the additional terms on the right-hand side act as effective forces required to explain the observed motion in the rotating frame. These are the fictitious forces: the Euler force -m \dot{\boldsymbol{\omega}} \times \mathbf{r}, the centrifugal force -m \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}), and the Coriolis force -2 m \boldsymbol{\omega} \times \mathbf{v}_{rot}. For clarity in the centrifugal term, apply the vector triple product identity (BAC-CAB rule): \mathbf{a} \times (\mathbf{b} \times \mathbf{c}) = \mathbf{b} (\mathbf{a} \cdot \mathbf{c}) - \mathbf{c} (\mathbf{a} \cdot \mathbf{b}). Thus, \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) = \boldsymbol{\omega} (\boldsymbol{\omega} \cdot \mathbf{r}) - \mathbf{r} (\boldsymbol{\omega} \cdot \boldsymbol{\omega}), \tag{5} which highlights the term's dependence on the angular speed \omega = |\boldsymbol{\omega}| and the radial component, confirming its outward-directed nature perpendicular to \boldsymbol{\omega}. This identity simplifies analysis without altering the derivation. The generalizes to cases of \boldsymbol{\omega}(t), where the \dot{\boldsymbol{\omega}} term captures changes in rotation rate, such as during frame spin-up or spin-down; for constant \boldsymbol{\omega}, this term vanishes, reducing to the standard centrifugal and Coriolis forces. If the rotating frame's origin accelerates translationally with \mathbf{a}_0 = \frac{d^2 \mathbf{R}}{dt^2} relative to the inertial frame, an additional -m \mathbf{a}_0 appears, but the rotational terms remain unchanged. Verification confirms the result's consistency: each fictitious term has dimensions of force (m \times acceleration, with \boldsymbol{\omega} in rad/s, \dot{\boldsymbol{\omega}} in rad/s², \mathbf{r} and \mathbf{v}_{rot} in m and m/s), matching \mathbf{F} and m \mathbf{a}_{rot}. In the limiting case \boldsymbol{\omega} = 0 and \dot{\boldsymbol{\omega}} = 0, Equation (4) recovers the inertial law \mathbf{F} = m \mathbf{a}_{rot}, as expected.

Effective Potential and Conservative Forces

In a rotating reference frame with constant angular velocity \mathbf{\omega}, the equations of motion for a particle can incorporate an that accounts for the real potential V and the centrifugal contribution. The is defined as V_\text{eff}(\mathbf{r}) = V(\mathbf{r}) - \frac{1}{2} m |\mathbf{\omega} \times \mathbf{r}|^2, where the second term arises from integrating the , enabling a conservative formulation for steady rotation. The is conservative, derivable as the negative gradient of the centrifugal potential V_c = -\frac{1}{2} m |\mathbf{\omega} \times \mathbf{r}|^2, so \mathbf{F}_\text{cent} = -\nabla V_c. In contrast, the is velocity-dependent and non-conservative, preventing its inclusion in any . In cylindrical coordinates (\rho, \phi, z) aligned with the rotation axis, where \rho is the perpendicular distance from the axis, the centrifugal force takes the form \mathbf{F}_\text{cent} = m \omega^2 \rho \hat{\mathbf{e}}_\rho. This effective potential framework is applied to analyze stability in the restricted three-body problem, where the test particle moves in the effective potential generated by two massive bodies in circular orbits; equilibrium points, known as Lagrange points, occur where \nabla V_\text{eff} = 0, with stability determined by the Hessian of V_\text{eff}. The conservativity holds only for steady, axisymmetric rotation with constant \mathbf{\omega}; time-varying angular velocity introduces the , which is non-conservative and disrupts the time-independent potential structure.

Applications of Rotating Frames

In Terrestrial Mechanics and Geophysics

In terrestrial and , rotating reference frames are essential for understanding phenomena influenced by Earth's rotation, particularly through fictitious forces like the Coriolis effect. In ocean and atmospheric dynamics, the Coriolis effect deflects moving air masses, shaping global wind patterns such as the . In the Hadley circulation, warm air rises near the and flows poleward aloft, sinks in the around 30° , and returns equatorward at the surface; this surface flow is deflected to the right in the by the Coriolis effect, resulting in the northeast , while in the Southern Hemisphere, deflection to the left produces the southeast . This deflection also governs the rotation of large-scale weather systems; for hurricanes in the , inbound winds toward the low-pressure center are turned rightward, creating counterclockwise , with the effect requiring sufficient away from the where Coriolis force vanishes. Similar dynamics apply in the Southern Hemisphere, yielding clockwise rotation. Geodesy employs rotating frames to explain Earth's non-spherical shape, where centrifugal forces from counterbalance , leading to an oblate spheroid form. The arises as material at the experiences maximum centrifugal acceleration, reducing effective and allowing outward expansion until is achieved. This results in an equatorial approximately 21 km larger than the polar , with Earth's ratio about 1/298. The oblateness influences and geodetic measurements, such as variations in by roughly 0.5% from pole to . For local-scale mechanics, rotating frames account for deflections in falling objects and pendulums due to Coriolis forces. A body dropped from rest experiences an eastward deflection in the , approximately \frac{1}{3} \omega g t^3, where \omega is Earth's , g is , and t is fall time; this arises from the vertical coupling with the horizontal component of . Projectiles and Foucault pendulums similarly deviate eastward, with observable shifts in experiments confirming the effect's at mid-latitudes. In applications, rotating reference frames analyze in machinery like turbines and shafts, where gyroscopic effects from Coriolis and centrifugal forces influence dynamic balance. Gyroscopes, for instance, resist perturbations through induced by these forces, enhancing in rotating systems such as rotors or . A key quantitative measure in geophysical flows is the , Ro = \frac{U}{f L}, where U is characteristic velocity, L is length , and f = 2 \omega \sin \phi is the Coriolis parameter with latitude \phi; small Ro (e.g., < 0.1) indicates dominant rotational effects, as in large-scale currents or atmospheric vortices.

In Magnetic Resonance and Spectroscopy

In (NMR) spectroscopy, the rotating reference frame provides a powerful tool for analyzing the dynamics of nuclear spins in a . This frame rotates synchronously with the of the spins at the frequency \omega_L = \gamma B_0, where \gamma is the of the and B_0 is the applied static along the z-axis. By adopting this perspective, the rapid oscillatory motion of the spins around B_0 in the laboratory frame becomes stationary, allowing the focus to shift to slower, more manageable interactions induced by radiofrequency (RF) fields. This transformation, introduced in the foundational work on nuclear induction, greatly simplifies the mathematical description of spin evolution during resonance experiments. In the rotating frame, the effective magnetic field experienced by the spins, denoted \mathbf{B}_\mathrm{eff}, combines the RF field component \mathbf{B}_1 (typically applied along the x-axis in the rotating coordinates) with an offset term arising from any deviation between the rotation frequency and the exact Larmor frequency. Specifically, \mathbf{B}_\mathrm{eff} = (B_1, 0, \Delta \omega / \gamma), where \Delta \omega represents the frequency offset from . The spins then precess around this tilted effective field at a reduced angular frequency \omega_\mathrm{eff} = \gamma |\mathbf{B}_\mathrm{eff}|, which is much slower than the Larmor frequency when \Delta \omega and B_1 are small compared to B_0. This effective field framework is central to understanding conditions and spin tipping. The Bloch equations, which govern the time evolution of the macroscopic magnetization vector \mathbf{M}, take a particularly simple form in the rotating frame: \frac{d\mathbf{M}}{dt} = \gamma \mathbf{M} \times \mathbf{B}_\mathrm{eff} - \frac{M_x \mathbf{\hat{i}} + M_y \mathbf{\hat{j}}}{T_2} - \frac{(M_z - M_0) \mathbf{\hat{k}}}{T_1}, where T_1 and T_2 are the longitudinal and transverse relaxation times, respectively, and M_0 is the equilibrium . The torque term \gamma \mathbf{M} \times \mathbf{B}_\mathrm{eff} describes coherent , while the relaxation terms account for energy dissipation and . These equations, derived phenomenologically, enable precise modeling of signal in NMR detectors. The primary advantages of the rotating frame lie in its ability to eliminate the high-frequency Larmor oscillations, making the effects of RF pulses—such as and selective excitation—far easier to analyze and design. This simplification is essential for developing complex pulse sequences in multidimensional NMR spectroscopy and for spatial encoding in (MRI), where gradients modulate the effective field to produce images. Without the rotating frame, the rapid would obscure these subtler dynamics, complicating both theoretical predictions and experimental implementation. The approach was instrumental in the pioneering NMR experiments by at Stanford and Edward Purcell at Harvard, earning them the 1952 for their discoveries concerning magnetic resonance in solids.

In Astrophysics and Celestial Mechanics

In the restricted , a rotating reference frame is employed where the two primary masses, such as a and a , are fixed at constant separation, allowing analysis of the motion of a negligible third mass under their combined gravitational plus fictitious forces. This frame rotates with determined by Kepler's third law for the primaries' , transforming the into equilibria at the five Lagrange points (L1–L5), where the gravitational, centrifugal, and Coriolis forces balance. The collinear points L1, L2, and L3 lie along the line joining the primaries, while L4 and L5 form equilateral triangles with them. The in this frame, combining gravitational and centrifugal terms, governs the motion; its contours reveal the topology around the Lagrange points. L1 and appear as saddle points, permitting unstable orbits that require station-keeping for , whereas L4 and L5 correspond to minima, supporting librations for small third-body masses when the mass ratio of the primaries exceeds approximately 25:1. These stability properties arise from the deflecting perturbations, enabling long-term confinement near L4 and L5. In binary star systems, tidal forces in the rotating frame define the as the closed equipotential surface surrounding each star, shaped by the balance between gravitational attraction and . Overflow of this lobe leads to , with the lobe's teardrop form elongating toward the companion due to tidal distortion. This configuration is crucial for understanding accretion processes in close binaries. Galactic dynamics in the solar neighborhood utilize a rotating frame aligned with differential rotation, incorporating Coriolis-like terms from the epicycle approximation to describe stellar motions around the rotation curve. The Oort constants A and B quantify local shear and vorticity, with A ≈ 14 km/s/kpc representing half the shear rate and B ≈ -12 km/s/kpc related to angular velocity, derived from proper motions and radial velocities. These constants embed fictitious forces analogous to those in rigid-body rotation, aiding analysis of orbital perturbations. A prominent example is Jupiter's Trojan asteroids, which librate around the L4 and L5 points, maintaining longitudes approximately 60° ahead or behind the planet in its around the Sun. These objects, numbering over 10,000, remain dynamically stable over billions of years due to the favorable in the Sun-Jupiter system, illustrating the practical utility of rotating frames in .

References

  1. [1]
    Rotating reference frames
    ... rotating) reference frame. Let us observe the motion of this object in a non-inertial reference frame that rotates with constant angular velocity $\Omega ...
  2. [2]
    10.3 Effects of Earth's Rotation: Apparent Forces | METEO 300
    We have thus related velocity in the absolute reference frame to velocity in the rotating reference frame. Now we can consider acceleration. Since →U=D→rDt ...Missing: tutorial | Show results with:tutorial
  3. [3]
    Space and Time: Inertial Frames
    Mar 30, 2002 · A body that is rotating with respect to a reference frame (and dial-traveller) is rotating with respect to any other frame in uniform motion ...
  4. [4]
    5.2: Newton's First Law – University Physics Volume 1
    All frames moving uniformly with respect to this fixed-star frame are also inertial. For example, a nonrotating reference frame attached to the Sun is, for all ...
  5. [5]
    Frames of Reference and Newton's Laws - Galileo
    An example of a non-inertial frame is a rotating frame, such as a carousel. The “laws of physics” we shall consider first are those of Newtonian mechanics ...
  6. [6]
    6.4 Fictitious Forces and Non-inertial Frames: The Coriolis Force
    Rotating and accelerated frames of reference are non-inertial. Fictitious forces, such as the Coriolis force, are needed to explain motion in such frames.
  7. [7]
    38. Fictitious Forces and Non-inertial Frames: The Coriolis Force
    In contrast, a non-inertial frame of reference is one that is accelerating. Observers in such frames perceive fictitious forces to explain observed motion.Missing: mechanics | Show results with:mechanics
  8. [8]
    Newton's views on space, time, and motion
    Aug 12, 2004 · Newton defined the true motion of a body to be its motion through absolute space. Those who, before or shortly after Newton, rejected the ...
  9. [9]
    [PDF] Newton's Scholium on Time, Space, Place and Motion-I. The Text
    The scholium following the eight opening definitions of the Principia sets out. Newton's views on absolute time, space, place and motion and proceeds to.
  10. [10]
    The relativity principle - Richard Fitzpatrick
    An inertial frame is a Cartesian frame in which free particles move without acceleration, in accordance with Newton's first law of motion. There are an infinite ...Missing: mechanics | Show results with:mechanics
  11. [11]
    [PDF] Chapter 5 Force and Motion - Physics for Scientists and Engineers
    Feb 25, 2008 · ▫ In the accelerating plane, the ball accelerates backward with no horizontal force acting on it. ⇒ This is a non-inertial frame. Page 23. 25 ...
  12. [12]
    The Basics of Reference Frames Relevant to Physics
    Rotating frames are non-inertial frames, but NOT simple ones. Every small region in them over a short enough time scale is a simple non-inertial frame (i.e., a ...
  13. [13]
    Rotating Frame of Reference | Physics in a Nutshell
    Two coordinate systems, one being inertial and one rotating with constant angular velocity Ω Ω with respect to the inertial one.
  14. [14]
    [PDF] Chapter 31 Non-Inertial Rotating Reference Frames
    An object is called an isolated object if there are no physical interactions between the object and the surroundings. According to Newton's First Law an ...
  15. [15]
    Vector Properties of Rotational Quantities
    Left with two choices about direction, it is customary to use the right hand rule to specify the direction of angular quantities. Direction of other angular ...Missing: physics | Show results with:physics
  16. [16]
    Angular Speed of the Earth - The Physics Factbook - hypertextbook
    The Earth rotates at a moderate angular velocity of 7.2921159 × 10−5 radians/second. Venus has the slowest angular velocity (2.99 × 10−7rad/sec) and Jupiter has ...
  17. [17]
    Centrifugal Acceleration - Richard Fitzpatrick
    Consider an object which appears stationary in our rotating reference frame: i.e., an object which is stationary with respect to the Earth's surface.
  18. [18]
  19. [19]
    [PDF] Noninertial Reference Frames - Physics Courses
    acts only on moving particles, whereas the centrifugal force is present even when ˙r = 0. ... To conserve angular momentum, the object must speed up as it falls.
  20. [20]
    Centrifugal Force on the Equator — Collection of Solved Problems
    Aug 4, 2016 · the Earth attracts the person with the force of gravity Fg,; the ... centrifugal force comprises 0.3% of the weight on the Equator). 2 ...
  21. [21]
    [PDF] a Coriolis tutorial - Woods Hole Oceanographic Institution
    On the other hand, the same runner making a moderately sharp turn, radius R = 15 m, will undoubtedly feel the centrifugal force,. (V 2/R)M ≈ 0.15gM, and ...Missing: omega (omega
  22. [22]
    The Coriolis Effect - Currents - NOAA's National Ocean Service
    Because the Earth rotates on its axis, circulating air is deflected toward the right in the Northern Hemisphere and toward the left in the Southern Hemisphere.
  23. [23]
    Falling eastward
    Calculate the eastward deflection due to the Coriolis effect if the object is allowed to free-fall from an initial state of rest to the earth's surface.
  24. [24]
    The Foucault pendulum - the physics (and maths) involved - UNSW
    At the poles, the (apparent) precession period is 23.9 hours, whereas elsewhere it is longer by a factor of the reciprocal sine of the latitude. But why? And is ...Missing: omega phi
  25. [25]
    PC1672: 2.5 Foucault's pendulum
    The rate of precession is equal to the vertical component of the Earth's angular velocity. (From an inertial observer's point of view, the pendulum is just ...
  26. [26]
    Gaspard-Gustave de Coriolis (1792 - 1843) - Biography - MacTutor
    Gaspard-Gustave de Coriolis is best remembered for the Coriolis force. He showed that the laws of motion could be used in a rotating frame of reference.<|separator|>
  27. [27]
    [PDF] 6. Non-Inertial Frames - DAMTP
    In a rotating frame, V has the interpretation of the potential energy associated to a particle. The potential V is negative, which tells us that particles want ...
  28. [28]
    [PDF] Euler Force
    Notice that the Euler d in our text. rence frames exhibit fictitious forces. Rotating reference frames are e fictitious forces. [1].
  29. [29]
    On the “Forgotten” Euler Inertial Force | Resonance
    Jun 13, 2024 · On the “Forgotten” Euler Inertial Force ... Article PDF. Download to read the full article text. Similar content being viewed by others ...
  30. [30]
    [PDF] Mechanics (UCSD Physics 110B)
    Jan 1, 2009 · r / + m~ω × ~ω × r /. ~F / = ~F − m~ω × ~ω × r / − 2m~ω × v /. This ... There is a simple fictitious force −m ~A if the coordinate system is ...
  31. [31]
    [PDF] Some Basics on Frames and Derivatives of Vectors
    The following equation, known as the 'Transport Theorem', is a formula for the derivative of any vector r frame of reference O, dr dr. = + ωP/O × r. (11) dt.
  32. [32]
    [PDF] 11. Dynamics in Rotating Frames of Reference
    Nov 5, 2015 · −m~ω × (~ω × r) (centrifugal force). If the origin of frame R undergoes a lateral motion in addition to the rotation, then a term −m(d2R/dt2) ...
  33. [33]
    [PDF] Relative Motion using Rotating Axes - MIT OpenCourseWare
    To determine the acceleration, we again discuss the general formula for the time derivative of a vector in a rotating frame, Coriolis Theorem. We consider the ...
  34. [34]
    [PDF] Newtonian Dynamics - Richard Fitzpatrick
    planetary orbit is an ellipse. Suppose that the major and ... that fictitious forces can always be distinguished from non-fictitious forces in Newtonian.
  35. [35]
    None
    ### Summary of Fictitious Forces Derivation in Rotating Frames
  36. [36]
    [PDF] Work-Energy theorem in rotational reference frames - arXiv
    Dec 30, 2010 · the mechanical energy has an effective potential given by (kr2. 2. − 1. 2 mω2r2) where the last term is related to the conservation of angular ...
  37. [37]
    [PDF] Phys410, Classical Mechanics Notes Ted Jacobson - Physics - UMD
    Dec 20, 2012 · 3.2 Effective potential; example of spherical pendulum . . . . . . . ... Let's see how to describe motion in a rotating frame, and how that is ...
  38. [38]
    [PDF] 8.09(F14) Advanced Classical Mechanics - MIT OpenCourseWare
    Figure 4.1: Plot of the effective potential Veff along with the different qualitative orbits ... rotating frame. This argument can be repeated for other points r ...
  39. [39]
    [PDF] Classical Mechanics Charles B. Thorn1 - UF Physics
    sin(Ωt), ω2 = S sinθ. I1 cosψ = ω⊥ cos(Ωt). (253) in agreement with the ... rotation of the coordinate frame relative to an inertial frame. If the frame ...
  40. [40]
    [PDF] Effective Potential - UCSB Physics
    we have shifted into a rotating reference frame co-rotating with the particle. Thus this effective potential term is what is known as the centrifugal ...
  41. [41]
    Rotating frame - Questions and Answers ​in MRI
    The rotating frame of reference is a concept used to simplify the complex motion of precessing spins before, during, or after RF-excitation.
  42. [42]
    [PDF] The Lagrange Points - Wilkinson Microwave Anisotropy Probe
    Today we know that the full three-body problem is chaotic, and so cannot be solved in closed form. Therefore, Lagrange had good reason to make some approx-.Missing: seminal | Show results with:seminal
  43. [43]
    [PDF] The Astrophysical Journal, 268:368-369, 1983 May 1
    APPROXIMATIONS TO THE RADII OF ROCHE LOBES. Peter P. Eggleton. Institute of Astronomy, Cambridge, England. Received 1982 September 14; accepted 1982 October 20.
  44. [44]
    Roche-lobe Overflow in Eccentric Planet–Star Systems - IOPscience
    Many giant exoplanets are found near their Roche limit and in mildly eccentric orbits. In this study, we examine the fate of such planets through Roche-lobe ...