Fact-checked by Grok 2 weeks ago

Coriolis frequency

The Coriolis frequency, also known as the Coriolis parameter and denoted by f, is a key quantity in that represents the vertical component of the Earth's angular rotation rate, specifically f = 2 \Omega \sin \phi, where \Omega \approx 7.292 \times 10^{-5} is the Earth's and \phi is the . This parameter vanishes at the (\phi = 0^\circ) and reaches its maximum value of $2\Omega at the poles, reflecting the latitude-dependent influence of planetary rotation on fluid motions. In the context of rotating reference frames, the Coriolis frequency governs the apparent deflection of moving fluids and particles—the —causing rightward deflection in the and leftward in the for horizontal motions. It is central to understanding large-scale atmospheric and oceanic circulations, where it balances pressure gradient forces in geostrophic flow, a state approximated in mid-latitude systems and gyres. The frequency also sets the timescale for inertial oscillations, circular motions with period $2\pi / f (approximately 24 hours at 30° , lengthening toward the ), which contribute significantly to upper- kinetic and wind-driven currents. Beyond basic balances, variations in the Coriolis frequency with latitude introduce the beta effect (\beta = \partial f / \partial y), which drives planetary-scale phenomena like Rossby waves—long-wavelength undulations essential for mid-latitude weather patterns and ocean basin-scale dynamics. In , it influences and rotations, while in , it shapes western boundary currents and equatorial dynamics under modified approximations. These effects underscore the Coriolis frequency's role in constraining vertical motions via the Taylor-Proudman theorem, promoting columnar structures in rotating fluids that resist north-south stretching.

Fundamentals

Definition

The Coriolis frequency, denoted as f, is the associated with the vertical component of the Coriolis acceleration in a on , defined as f = 2 \Omega \sin \phi, where \Omega is the of the frame and \phi is the . For , \Omega \approx 7.292 \times 10^{-5} rad/s, corresponding to its sidereal rate. This frequency arises from the fictitious forces in non-inertial frames and characterizes the rate at which rotating motion deflects trajectories, leading to oscillatory behavior in the absence of other forces. In geophysical contexts, the Coriolis frequency serves as a local , particularly for horizontal motions where the vertical component of dominates the effect. It has units of s^{-1}, emphasizing its nature as a rather than a force. While the itself is given by \mathbf{F} = -2 m \Omega \times \mathbf{v}, where m is mass and \mathbf{v} is , the f specifically quantifies the inherent oscillatory tendency induced by this deflection, with $2\pi / f.

Historical Development

The concept of the Coriolis frequency traces its origins to the foundational work of , a and , who in 1835 analyzed the equations of relative motion in rotating systems. In his paper "Sur les équations du mouvement relatif des systèmes de corps," Coriolis introduced supplementary forces arising in non-inertial frames, including what is now termed the , initially in the context of mechanical devices like waterwheels and turbines. This formulation provided the mathematical basis for understanding rotational effects on motion, though its geophysical implications were not immediately recognized. In the mid-19th century, American meteorologist William Ferrel advanced these ideas by applying them to Earth's . In his 1856 article "An Essay on the Winds and the Currents of the Ocean," Ferrel described how the planet's rotation deflects moving air masses, linking this to the formation of prevailing wind patterns and mid-latitude circulation cells. Drawing inspiration from earlier observations like and Laplace's tidal theories, Ferrel qualitatively incorporated the latitude-dependent deflection into explanations of zonal winds, marking a pivotal step toward its use in . A significant milestone occurred in 1905 with Swedish oceanographer Vagn Walfrid Ekman's development of the theory, which explicitly employed the Coriolis parameter to model wind-driven currents in the ocean surface boundary layer. Ekman's work, published as "On the Influence of the on Ocean Currents," demonstrated how frictional forces balance with Coriolis deflection to produce a spiraling profile, influencing subsequent studies in . The early 20th century saw further formalization through the efforts of Norwegian meteorologist and the Bergen School during the 1910s and 1920s. Bjerknes integrated the into comprehensive hydrothermodynamic models for weather prediction, as outlined in his 1904 paper "Das Problem der Wettervorhersage" and subsequent works, enabling graphical and numerical analysis of atmospheric circulations. This approach, refined by collaborators like his son Jacob Bjerknes and Halvor Solberg, emphasized geostrophic balance and laid the groundwork for modern synoptic meteorology. Another key advancement came in 1939 with Carl-Gustaf Rossby's identification of planetary waves, now known as , in his paper "Relation between Variations in the Intensity of the Zonal Circulation of the Atmosphere and the Displacements of the Semi-Permanent Centers of Action." Rossby's analysis revealed how variations in the Coriolis parameter drive large-scale wave propagation in the atmosphere, connecting rotational effects to global circulation patterns.

Mathematical Formulation

Derivation from Coriolis Effect

The for a particle in an inertial reference frame are given by Newton's second law: \frac{d^2 \mathbf{r}}{dt^2} = \frac{\mathbf{F}}{m}, where \mathbf{r} is the position vector, \mathbf{F} is the net physical force, and m is the mass. In a with constant angular velocity \boldsymbol{\Omega}, the observed position \mathbf{r}', velocity \mathbf{v}', and acceleration \mathbf{a}' differ from their inertial counterparts due to the frame's rotation. The for the time of a vector \mathbf{A} between the inertial (subscript i) and rotating is \left( \frac{d\mathbf{A}}{dt} \right)_i = \frac{d\mathbf{A}}{dt} + \boldsymbol{\Omega} \times \mathbf{A}. Applying this twice yields the in the rotating : \mathbf{a}' = \mathbf{a} - 2 \boldsymbol{\Omega} \times \mathbf{v}' - \boldsymbol{\Omega} \times (\boldsymbol{\Omega} \times \mathbf{r}'), where \mathbf{a} is the inertial . Thus, the equation of motion becomes m \mathbf{a}' = \mathbf{F} - 2m \boldsymbol{\Omega} \times \mathbf{v}' - m \boldsymbol{\Omega} \times (\boldsymbol{\Omega} \times \mathbf{r}'), introducing the Coriolis -2m \boldsymbol{\Omega} \times \mathbf{v}' and the centrifugal -m \boldsymbol{\Omega} \times (\boldsymbol{\Omega} \times \mathbf{r}'). The Coriolis term -2 \boldsymbol{\Omega} \times \mathbf{v}' (per unit mass) is perpendicular to both \boldsymbol{\Omega} and \mathbf{v}', causing deflection of moving objects without changing their speed. For geophysical applications on , \boldsymbol{\Omega} points along the rotation axis with magnitude \Omega = 7.292 \times 10^{-5} s^{-1}. In the f-plane approximation for horizontal motions at a fixed \phi, the vertical component of \boldsymbol{\Omega} dominates, yielding the Coriolis parameter (or frequency) f = 2 \Omega \sin \phi. The vector form simplifies to \mathbf{f} = f \hat{k}, where \hat{k} is the local vertical , so the Coriolis acceleration is - \mathbf{f} \times \mathbf{v}'. This assumes the horizontal velocity components u' (eastward) and v' (northward) are primary, with the centrifugal term often absorbed into an effective gravity. For two-dimensional geophysical flows, such as in shallow atmospheres or oceans, the Ro = \frac{U}{f L} (where U is a and L is a ) is assumed small (Ro \ll 1), indicating rotation dominates over nonlinear . The momentum equations then linearize to \frac{D u'}{Dt} - f v' = -\frac{1}{\rho} \frac{\partial p}{\partial x} and \frac{D v'}{Dt} + f u' = -\frac{1}{\rho} \frac{\partial p}{\partial y}, where \frac{D}{Dt} is the and p is , focusing the dynamics on the Coriolis frequency f.

Latitude and Vertical Variations

The Coriolis frequency, denoted as f, exhibits a strong dependence on \phi, arising from the geometry of . It is given by the f(\phi) = 2 \Omega \sin \phi, where \Omega is Earth's angular rotation rate, approximately $7.292 \times 10^{-5} rad s^{-1}. This expression reflects the projection of the planetary rotation vector onto the local vertical axis. At the , where \phi = 0^\circ, f = 0, meaning no horizontal deflection from the Coriolis effect occurs for northward or southward motions. Conversely, at the poles (\phi = \pm 90^\circ), f reaches its maximum value of \pm 2 \Omega, leading to the strongest deflection. This latitudinal variation is crucial for understanding the absence of large-scale cyclones near the and their prevalence at higher latitudes. In three-dimensional flows, a vertical component of the Coriolis frequency becomes relevant, particularly in the non-traditional approximation where vertical motions are not negligible. This vertical Coriolis parameter is f_v = 2 \Omega \cos \phi, which influences the vertical deflection of horizontal velocities. For instance, an eastward velocity u' experiences a vertical force component $2 \Omega \cos \phi \, u' directed along the local vertical. This term is zero at the poles and maximum at the equator (f_v = 2 \Omega), but it is often small compared to the horizontal component in typical geophysical contexts due to modest vertical velocities. Its inclusion is essential for modeling phenomena like inertial waves or flows near the equator where the traditional horizontal approximation breaks down. To study mid-latitude dynamics, the beta-plane approximation simplifies the latitudinal variation of f by linearizing it around a reference \phi_0. Here, f(y) \approx f_0 + \beta y, where y is the northward from the reference, f_0 = 2 \Omega \sin \phi_0, and \beta = \frac{2 \Omega \cos \phi_0}{a} with a \approx 6371 km being Earth's . This \beta captures the meridional gradient of planetary , enabling analysis of large-scale motions like Rossby waves without the full . For example, at 45° , f \approx 10^{-4} s^{-1}, providing a timescale for inertial oscillations with period $2\pi / f \approx 17 hours.

Physical Interpretation

Role in Rotating Reference Frames

In rotating reference frames, the Coriolis frequency f = 2 \Omega \sin \phi, where \Omega is the angular velocity of the frame and \phi is the effective colatitude for the component perpendicular to the motion plane (in geophysical contexts, Earth's latitude), governs the apparent deflection of particles in motion relative to the frame, resulting in curved trajectories that can exhibit cyclic patterns. This deflection arises because the rotating frame introduces kinematic terms that alter the observed acceleration of objects, transforming straight-line inertial motion into spiraling or circular paths when viewed from the non-inertial perspective. The Coriolis term manifests as a perpendicular to both the particle's and the axis of rotation, causing an apparent deflection to the right of the in frames rotating counterclockwise (analogous to the orientation) or to the left in clockwise-rotating frames (analogous to the ). Unlike genuine forces such as or , this lacks a corresponding reaction pair and does not arise from physical interactions; instead, it compensates for the frame's rotation in the , preserving the underlying Newtonian dynamics in an inertial frame. Laboratory demonstrations, such as those conducted in rotating tanks filled with , illustrate this deflection clearly: a freely moving object, like a on a rotating platform, traces out inertial circles due to the continuous Coriolis turning, with the radius and period determined by the initial speed and the frame's rotation rate governed by f. These setups, often used in experiments, highlight how f dictates the scale of cyclic motion without altering the particle's speed. The role of the Coriolis frequency is most significant in systems where rotational effects overwhelm advective inertia, characterized by a low Ro = U / (f L) \ll 1, with U as a typical and L as a scale; in such low-Ro regimes, the deflection leads to nearly balanced, rotationally constrained flows, whereas higher Ro values diminish its influence.

Inertial Oscillations

Inertial oscillations represent free motions in a rotating where the is the dominant influence, resulting in circular trajectories without external forcing. These oscillations arise in geophysical contexts, such as the atmosphere and , when initial velocities are imparted in the absence of gradients or , leading to periodic motion at the Coriolis frequency f. The phenomenon illustrates the fundamental role of in deflecting fluid parcels, producing closed loops that conserve . The governing equations for horizontal velocities u (eastward) and v (northward) in the , under the f-plane approximation, simplify to: \frac{du}{dt} = f v, \quad \frac{dv}{dt} = -f u, where f = 2 \Omega \sin \phi > 0 is the Coriolis parameter, \Omega is Earth's , and \phi is . These linear equations describe the perpendicular to the , with no change in speed. The general solution is a circular oscillation: u = U \cos(ft + \psi), \quad v = -U \sin(ft + \psi), where U is the constant speed (amplitude) and \psi is the phase angle determined by initial conditions. This yields an inertial period T = 2\pi / f, independent of U. For example, at 30° latitude where f \approx 7.3 \times 10^{-5} s^{-1}, T \approx 24 hours. The motion traces a circle of radius r = U / f. In the , the particle path is clockwise, with the velocity vector rotating to the right relative to its direction of motion; in the (f < 0), the path is anticlockwise. The amplitude U remains constant throughout, as the Coriolis force does no work and preserves kinetic energy. These characteristics hold for anticyclonic rotation aligned with the sense of Earth's rotation. This idealized description is valid for frictionless flows free of pressure gradients, typically on scales where the f-plane approximation applies (local regions much smaller than Earth's radius) and for time scales near T. Observations confirm these oscillations in the upper ocean following transient winds, though real-world damping introduces decay.

Applications in Geophysics

Atmospheric Dynamics

In atmospheric dynamics, the Coriolis frequency, denoted as f = 2 \Omega \sin \phi where \Omega is 's angular rotation rate and \phi is latitude, plays a central role in governing large-scale motions by balancing the pressure gradient force in horizontally non-divergent flows. This balance is particularly evident in the geostrophic approximation, which dominates synoptic-scale phenomena spanning thousands of kilometers, where the Rossby number (a measure of inertial to Coriolis forces) is much less than unity, rendering rotational effects paramount over local accelerations. The geostrophic balance equation is given by f \mathbf{k} \times \mathbf{V}_g = -\frac{1}{\rho} \nabla p, where \mathbf{V}_g is the geostrophic wind vector, \rho is air density, p is , and \mathbf{k} is the vertical unit vector. This arises from the steady-state momentum equations in a rotating frame, where the Coriolis force -f \mathbf{k} \times \mathbf{V} exactly opposes the pressure gradient force -\frac{1}{\rho} \nabla p, neglecting friction and acceleration terms valid for large scales. Solving for \mathbf{V}_g, the geostrophic wind flows parallel to isobars (constant pressure contours) with magnitude |\mathbf{V}_g| = \frac{1}{\rho f} |\nabla p| and direction perpendicular to the pressure gradient, such that in the Northern Hemisphere (f > 0), the wind veers with low pressure to the left. A key extension is the thermal wind relation, which connects vertical variations in the geostrophic wind to horizontal temperature gradients under hydrostatic balance. It is expressed as \frac{\partial \mathbf{V}_g}{\partial z} = \frac{g}{f T} \mathbf{k} \times \nabla T, where g is gravitational acceleration and T is temperature. This follows by differentiating the geostrophic balance vertically and substituting the hydrostatic equation \frac{\partial p}{\partial z} = -\rho g along with the ideal gas law p = \rho R T (with R the gas constant for dry air), yielding a shear that increases westerly winds with height poleward of warm anomalies, as seen in midlatitude jet streams. The sign of f dictates the curvature of balanced flows: in the , positive f supports cyclonic (counterclockwise) circulation around low-pressure centers, where the relative aligns with planetary to enhance total rotation, while anticyclonic () flow around highs opposes it, leading to weaker or reversed compared to non-rotating cases. This rotational constraint, rooted in f, ensures that geostrophic winds around cyclones exhibit tighter spacing and stronger speeds than around anticyclones for equivalent gradients. On synoptic scales of 1000–5000 km, such as extratropical cyclones, the dominates because the deformation radius N H / f (with N the and H a scale height) exceeds the size, promoting geostrophically balanced structures over inertial oscillations, whose is $2\pi / f \approx 12–24 hours at midlatitudes.

Oceanic Circulation

In oceanic circulation, the Coriolis frequency f is fundamental to the dynamics of the surface , where drives currents in a rotating frame. The balance between the and vertical turbulent friction results in a characteristic Ekman spiral, with velocities rotating clockwise with depth in the Northern Hemisphere and the surface current directed approximately 45° to the right of the wind. Deeper within the layer, currents align more closely with the direction 90° to the right of the wind, leading to net mass transport perpendicular to the . The integrated Ekman transport \mathbf{M} is given by \mathbf{M} = \frac{\boldsymbol{\tau}}{\rho f}, where \boldsymbol{\tau} is the vector, \rho is the seawater density, and the transport is directed 90° to the right of \boldsymbol{\tau} in the Northern Hemisphere; this relation highlights how f scales the magnitude and direction of wind-driven flow in the upper ocean. In the interior of ocean basins, far from lateral boundaries, the Sverdrup balance describes the large-scale meridional circulation in wind-driven gyres, incorporating the role of f. This balance arises from the vorticity equation, where the meridional advection of planetary vorticity by the mean flow is counteracted by the stretching of planetary vorticity due to vertical velocity convergence: \beta v = f \frac{\partial w}{\partial z}, with \beta = \frac{\partial f}{\partial y} representing the latitudinal variation of the Coriolis frequency and v the meridional velocity. Wind stress curl drives this interior flow, inducing upwelling in subtropical gyres and downwelling in subpolar regions, thereby establishing the broad-scale circulation patterns observed in major ocean basins like the North Atlantic and North Pacific. This framework explains how variations in f influence the strength and extent of gyre-scale transports. The latitudinal variation of the Coriolis frequency, through the \beta effect, also drives western intensification in ocean gyres, concentrating intense s on the western sides of basins. In the North Atlantic, this manifests in the , a swift western that returns poleward the equatorward from the subtropical gyre. The \beta effect causes relative to accumulate on the western side due to the conservation of , requiring a narrow, intense current to balance the planetary vorticity gradient imposed by wind forcing. This asymmetry arises because southward flow gains positive relative from stretching against the varying f, while northward flow loses it, leading to stronger western boundaries for Sverdrup transports. Observations confirm this intensification, with the exhibiting speeds exceeding 2 m/s over widths of about 100 km. Mesoscale eddies in the , with scales of 10–100 km, often exhibit near-inertial oscillations influenced by the local Coriolis frequency, forming structures akin to inertial rings. In anticyclonic eddies, where the relative reduces the effective inertial frequency, these oscillations propagate downward through a process known as the inertial chimney effect, with periods approximately equal to $2\pi / f. Such features are prominent in regions like the Gulf of Mexico's Loop Current eddies, where inertial rings contribute to enhanced vertical mixing and from the surface to deeper layers, modulating eddy lifetimes and nutrient upwelling. This interaction underscores f's role in eddy energetics and the broader dissipation of mesoscale variability.

Rossby Parameter

The Rossby parameter, denoted \beta, quantifies the meridional of the Coriolis f and is defined as \beta = \frac{\partial f}{\partial y}, where y is the northward-directed along the Earth's surface. This parameter arises in the beta-plane , which linearizes the variation of f with for analyses at mid-latitudes. The Coriolis parameter itself is given by f(\phi) = 2 \Omega \sin \phi, where \Omega is the of (\Omega \approx 7.29 \times 10^{-5} s^{-1}) and \phi is the . To derive \beta, consider small latitudinal variations where the northward y approximates a \phi (with \phi in radians and a the Earth's mean radius, a \approx 6.37 \times 10^6 m). Differentiating f with respect to \phi yields \frac{\partial f}{\partial \phi} = 2 \Omega \cos \phi, so \beta = \frac{\partial f}{\partial y} = \frac{2 \Omega \cos \phi}{a}. This expression captures the increase in the local component of planetary rotation with northward displacement. The Rossby parameter introduces the planetary vorticity gradient, which plays a crucial role in the westward propagation of large-scale waves in rotating fluids, such as those observed in Earth's atmosphere and . This gradient effect, first highlighted in analyses of zonal circulation variations, distinguishes planetary-scale dynamics from uniform rotation scenarios. At 45° latitude, \beta \approx 1.6 \times 10^{-11} m^{-1} s^{-1}, providing a characteristic scale for these processes.

Vertical Coriolis Frequency

The vertical Coriolis frequency, often denoted as f_v = 2 \Omega \cos \phi, where \Omega is Earth's and \phi is , quantifies the horizontal component of the planetary that influences vertical motions in rotating fluids. This contrasts with the horizontal Coriolis frequency f_h = 2 \Omega \sin \phi, which primarily governs deflections in horizontal flows; f_v arises in the vertical momentum equation as a term proportional to horizontal velocities, such as f_v u for eastward flow. In many geophysical models, f_v is smaller than f_h at mid-to-high latitudes and often neglected under the traditional , but it becomes comparable or dominant near the where f_h vanishes. In three-dimensional flows, particularly within stratified fluids, f_v introduces horizontal and modifies the coupling between vertical and horizontal velocities, breaking the north-south inherent in traditional treatments. A key example is its role in the Taylor-Proudman theorem, which posits that in rapidly rotating, low-Rossby-number flows, fluid parcels tend to move uniformly along columns parallel to the rotation axis, suppressing vertical variations; however, f_v alters this by inducing zonal dependencies and tilting these columns equatorward. Applications of f_v are prominent in wave dynamics and convective processes where vertical shear is significant. In acoustic-gravity waves (or inertio-gravity waves) in stratified media, f_v expands the allowable frequency range, enabling subinertial propagation (\omega < |f_h|) in weakly stratified layers by shifting the dispersion relation and facilitating energy transfer across stratification gradients. Similarly, in deep convection, such as in oceanic or atmospheric plumes, f_v generates equatorward tilts and vertical shears that organize convective structures, enhancing upscale momentum transport and vortex formation beyond traditional geostrophic balances. These effects underscore f_v's importance in equatorial and low-latitude regimes, where it compensates for the weak f_h in maintaining rotational constraints on vertical motions.

References

  1. [1]
  2. [2]
    [PDF] Coriolis effects: introduction to dynamics of fluids
    These surfaces give us the true definition of 'horizontal' on the Earth. We ... Thus the natural frequency of this system is f, the Coriolis frequency.
  3. [3]
    Transition from Geostrophic Flows to Inertia–Gravity Waves in the ...
    where ωi is the intrinsic frequency, N the buoyancy frequency, f = 2Ω the Coriolis frequency ... Coriolis frequency ωM/f are on the y axis. The first ...
  4. [4]
    Explanation: parameters of the Earth rotation
    It is defined as a prturbational ngle with respect to a uniform rotation with nominal angular velocity 7.292115146706979 · 10-5 rad/sec. Units: seconds.
  5. [5]
    [PDF] 1 A Fluid Dynamical Hierarchy - UCLA
    V is a characteristic horizontal velocity magnitude, f = 2Ω is Coriolis frequency (Ω is Earth's rotation rate), N is buoyancy frequency for the stable ...
  6. [6]
    Inertial Waves | SpringerLink
    ... Coriolis frequency (“near-inertial waves”), and (ii) when the frequency of oscillations is small compared to the Coriolis frequency 2Ω. Lastly, experimental ...
  7. [7]
    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.
  8. [8]
    Gustave Coriolis - Linda Hall Library
    May 21, 2019 · Gaspard-Gustave de Coriolis, a French engineer and mathematician, was born May 21, 1792. Coriolis was a well-respected instructor in the French engineering ...
  9. [9]
    [PDF] A brief history of the Coriolis force
    In 1856, Ferrel adopted the term s (I) and (III) from Laplaces tidal theory and introduced them in meteor ology [ 1]. In a qualitative way, the effect of term ...Missing: parameter William
  10. [10]
    Ekman Layer - an overview | ScienceDirect Topics
    Ekman layers are boundary layers in which there is a balance between the viscous force and the Coriolis acceleration. They are typically quite thin.
  11. [11]
    Weather forecasting - MacTutor History of Mathematics
    Most of Bjerknes's important results were published in On the Dynamics of the Circular Vortex with Applications to the Atmosphere and to Atmospheric Vortex Wave ...
  12. [12]
    Vilhelm Bjerknes' Vision for Scientific Weather Prediction
    Aug 5, 2025 · With the Bergen School of Meteorology he renewed weather forecasting throughout the world. Bjerknes' work inspired Louis Fry Richardson to his ...
  13. [13]
    A New Look at the Physics of Rossby Waves: A Mechanical–Coriolis ...
    Jan 1, 2013 · Rossby waves, named after the pioneering work of Rossby (1939) ... Rossby wave motions propagating to the left of the Coriolis parameter gradient.
  14. [14]
    [PDF] The Equations of Motion in a Rotating Coordinate System
    the Coriolis force acts normal to the rotation vector and normal to the velocity. is directly proportional to the magnitude of u and Ω. Note: the Coriolis force ...
  15. [15]
    [PDF] a Coriolis tutorial, Part 1: - Woods Hole Oceanographic Institution
    Mar 11, 2022 · Hence the Coriolis force does no work. Nevertheless the Coriolis force has a profound importance for the circulation of the atmosphere and.
  16. [16]
    [PDF] http://www.youtube.com/watch?v=_R7L5DNvgfU http://www.youtube ...
    - Coriolis parameter is f; f = 2 Ω sin(ϕ). Ω = 7.29 x 10. -5 s-1. ; ϕ is latitude. - It is a matter of frame of reference, there is NO Coriolis “force”. Page 17 ...
  17. [17]
    [PDF] AOSS321_L09_020509_scale_a...
    with the Coriolis parameter f = 2Ω sinϕ. Note: There is no D( )/Dt term ... f = 2 Ω sin(ϕ). Ω ≈ 7.292 × 10-5 s-1. Page 37. Rossby number. • Rossby number ...
  18. [18]
    7.2: Rotating Reference Frames - Physics LibreTexts
    Dec 30, 2020 · The Coriolis force is present whenever a particle is moving with respect to the rotating coordinates, and tends to deflect particles from a ...
  19. [19]
    [PDF] Rotating reference frames and the Coriolis force Part II - DSpace@MIT
    Sep 24, 2010 · Part I uses a very simple single parcel model to derive and illustrate the equation of motion appropriate to a steadily rotating reference ...
  20. [20]
    [PDF] Chapter 6 The equations of fluid motion - Weather in a Tank
    In the rotating frame, the puck undergoes inertial oscillations consisting of small circular orbits passing through the initial position of the unper- turbed ...
  21. [21]
    The Coriolis force and Rossby Number - UW Pressbooks
    The Coriolis force is an apparent force perpendicular to velocity in a rotating frame. The Rossby number is the ratio of inertial to Coriolis terms.Missing: frequency | Show results with:frequency<|control11|><|separator|>
  22. [22]
    [PDF] Inertial Oscillations
    The solution of these equations for u and v (check and see!) is a circular oscillation with period T = 2π/f. = f is called the inertial frequency. u= U sin(ft).
  23. [23]
    Inertial motion on the earth's spheroidal surface - AIP Publishing
    Nov 8, 2022 · The loop is clockwise in the northern hemisphere and counterclockwise in the southern hemisphere and does not pass through or around a pole.
  24. [24]
    [PDF] An Introduction to Dynamic Meteorology - webspace.science.uu.nl
    The development of the Coriolis force in Section 1.5 has been substantially improved from previous editions. The dis- cussion of the barotropic vorticity ...
  25. [25]
    [PDF] 151 02749 8728 Elanan, Vagn Waif rid On the influence of the ...
    The paper discusses the influence of the earth's rotation on ocean currents, noting that surface currents deviate 45° to the right of wind in the northern ...
  26. [26]
    [PDF] 1 GEOPHYSICAL FLUID DYNAMICS-I OC512/AS509 2015 P ...
    Mar 10, 2015 · u·ł(2ΩΩ) where now β =df/dy = 2ΩΩ/α cos φ. a is the Earth's radius. Under the geostrophic approximation, we neglect ζ compared with f on the.Missing: parameter | Show results with:parameter
  27. [27]
    [PDF] Relation between variations in the intensity of the zonal circulation of ...
    This paper attempts to interpret, from a single point of view, several at first sight independent phenomena brought into focus through the synoptic.
  28. [28]
    Balanced Models and Dynamics for the Large- and Mesoscale ...
    Dimensionally, we choose a square geometry with L = 6000 km, H0 = 10 km, f0 = 1.03 × 10−4 s−1, β = 1.62 × 10−11 (m s)−1, and g = 2.5 m s−2. The values of f0 and ...
  29. [29]