Fact-checked by Grok 2 weeks ago

Geostrophic wind

The geostrophic wind is a theoretical wind in the atmosphere that results from an exact balance between the , which drives air from high to low pressure, and the , which deflects moving air due to , causing the wind to flow parallel to isobars without any net acceleration perpendicular to the flow. This balance assumes straight, frictionless, and steady flow on large spatial scales (greater than a few kilometers) and temporal scales (longer than 12 hours), typically above the where surface friction is negligible. In the , geostrophic winds blow with low pressure to the left and high pressure to the right, while in the , the orientation is reversed with low pressure to the right. The magnitude of the geostrophic wind can be estimated using the formula v_g = \frac{1}{f \rho} \frac{\partial p}{\partial n}, where f is the Coriolis parameter (f = 2 \Omega \sin \phi, with \Omega as Earth's and \phi as ), \rho is air , and \frac{\partial p}{\partial n} is the perpendicular to the isobars; closer spacing of isobars indicates a stronger geostrophic . This approximation is most valid at mid-s away from the (beyond about 2° latitude) and breaks down in regions of significant curvature, such as around high- or low-pressure centers, where centrifugal forces require modifications like the gradient wind balance. In , the geostrophic wind serves as a foundational concept for analyzing large-scale , particularly in the free atmosphere above 1–2 km altitude, and is widely used in models to derive quantities like and from height fields, such as at the 500 hPa level. Although actual winds deviate from geostrophic due to , ageostrophic components, and other forces, the geostrophic approximation provides a close estimate for synoptic-scale flows in mid-latitudes and underpins understandings of phenomena like the and relationships.

Conceptual Foundations

Definition and Physical Balance

The geostrophic wind describes a state of geostrophic balance in rotating fluids, such as the atmosphere and oceans, where the velocity of the wind or current is directed perpendicular to the local , leading to no net of the . In this , the flow proceeds parallel to isobars—lines of constant —without crossing them, as the forces maintain a steady motion. This balance is particularly relevant in large-scale fluid systems where other influences, like , are minimal. The equilibrium arises from the interplay of two dominant forces: the pressure gradient force (PGF), which accelerates fluid parcels from regions of high pressure toward low pressure, perpendicular to the isobars, and the , a due to the rotation of that deflects moving parcels to the right of their velocity vector in the (and to the left in the ). As a parcel initially accelerates under the PGF and gains speed, the Coriolis force strengthens proportionally to the velocity until it exactly opposes the PGF in magnitude and direction, halting further deflection or acceleration. This force equilibrium can be visualized through a simple diagram: the PGF points toward lower (cross-isobar direction), the lies to the isobars ( to the PGF), and the Coriolis matches the PGF but points in the opposite direction, ensuring the is zero. In the , for instance, a westerly (blowing east) experiences a southward Coriolis force that balances a northward PGF associated with higher to the north. Geostrophic winds approximate actual flows effectively in large-scale systems like mid-latitude cyclones because these phenomena occur over vast horizontal scales where the is small, minimizing relative accelerations and allowing the Coriolis and PGF to dominate over or other perturbations. This approximation underpins much of the understanding of planetary-scale circulations in both atmospheric and oceanic contexts.

Historical Origin

The concept of geostrophic wind emerged from 19th-century investigations into the role of in fluid motions, particularly through the work of American meteorologist William Ferrel. In his 1856 publication, Ferrel described how the Coriolis effect influences , proposing a mid-latitude circulation where westerly winds arise from the balance between pressure gradients and rotational forces, laying foundational ideas for later geostrophic approximations. The formal introduction of geostrophic wind as a practical tool in occurred in the early 20th century through the efforts of Norwegian physicist and his collaborators. In their seminal 1910-1911 work Dynamic Meteorology and Hydrography, Bjerknes and Johan Wilhelm Sandström outlined the balance between Coriolis and forces in large-scale atmospheric flows, enabling the approximation of wind speeds from isobaric maps for . This framework was advanced by the Bergen School, founded by Bjerknes in 1917, which applied geostrophic principles to synoptic analysis and development during the and . In , the geostrophic approximation was contrasted with surface-layer dynamics by oceanographer Vagn Walfrid Ekman in his paper, where he developed the theory of wind-driven currents in the upper ocean, showing that frictional effects dominate near the surface while geostrophy governs deeper, frictionless layers. A key milestone in integrating geostrophy into broader large-scale dynamics came in the 1930s with Carl-Gustaf Rossby, who linked the concept to planetary waves in the atmosphere, demonstrating how zonal flow variations propagate as Rossby waves under geostrophic balance. Following , geostrophic wind principles became integral to models in the 1940s and 1950s, as exemplified by Jule Charney's quasi-geostrophic framework, which filtered high-frequency noise and enabled computational forecasts on early computers like . This evolution solidified geostrophy's role in simulating large-scale atmospheric and circulations.

Mathematical Formulation

Derivation from Equations of Motion

The derivation of the geostrophic wind begins with the horizontal momentum equations for a rotating fluid, derived from the Navier-Stokes equations in a non-inertial frame rotating with the . These equations, in Cartesian coordinates with x directed eastward and y northward, are: \frac{du}{dt} - f v = -\frac{1}{\rho} \frac{\partial p}{\partial x}, \frac{dv}{dt} + f u = -\frac{1}{\rho} \frac{\partial p}{\partial y}, where u and v are the velocity components, \rho is the fluid density, p is , and f = 2 \Omega \sin \phi is the Coriolis parameter, with \Omega the angular rotation rate of the (\Omega \approx 7.29 \times 10^{-5} s^{-1}) and \phi the . To arrive at geostrophic , the following assumptions are applied: the is in a , neglecting local and advective time derivatives (du/dt = dv/dt = 0); frictional forces are absent; vertical accelerations are negligible, restricting attention to motions; and the features a small (Ro = U / (f L) \ll 1), where U is a scale and L is the length scale, ensuring the dominates over inertial accelerations. Under these conditions, the equations simplify by setting the acceleration terms to zero, resulting in a direct balance between the and the : - f v_g = -\frac{1}{\rho} \frac{\partial p}{\partial x}, f u_g = -\frac{1}{\rho} \frac{\partial p}{\partial y}. Solving for the geostrophic velocity components u_g and v_g yields: v_g = \frac{1}{f \rho} \frac{\partial p}{\partial x}, u_g = -\frac{1}{f \rho} \frac{\partial p}{\partial y}. This component form can be compactly expressed in vector notation as \vec{v_g} = \frac{1}{f \rho} \hat{k} \times \nabla p, where \hat{k} is the vertical and \nabla p = (\partial p / \partial x) \hat{i} + (\partial p / \partial y) \hat{j} is the horizontal ; the cross product ensures the geostrophic wind is perpendicular to the , paralleling isobars in the for f > 0. The simplifies the algebra here, though the balance extends to spherical coordinates with latitude-dependent f and metric terms, without altering the core pressure-Coriolis equilibrium.

Governing Equations

The geostrophic wind arises from the balance between the and the in the horizontal momentum equations, yielding scalar components for the zonal and meridional velocities in the , where the Coriolis parameter f is positive. These are given by u_g = -\frac{1}{f \rho} \frac{\partial p}{\partial y}, \quad v_g = \frac{1}{f \rho} \frac{\partial p}{\partial x}, where u_g and v_g are the geostrophic wind components, \rho is the fluid density, and p is pressure. In vector notation, the geostrophic wind \vec{v_g} is expressed as \vec{v_g} = \frac{1}{f \rho} \mathbf{k} \times \nabla p, where \mathbf{k} is the unit vector in the vertical and \nabla p is the horizontal ; this form highlights that the wind flows parallel to isobars, with direction deflected to the right of the pressure gradient in the . The Coriolis parameter is defined as f = 2 \Omega \sin \phi, where \Omega is Earth's ($7.292 \times 10^{-5} rad ^{-1}) and \phi is ; f vanishes at the (\phi = 0^\circ) and reaches maximum values near the poles (|\phi| = 90^\circ), introducing latitudinal variation in geostrophic balance strength. Vertical variations in the geostrophic wind, or geostrophic shear, are described by the thermal wind relation under the hydrostatic approximation, which links shear to horizontal temperature gradients: \frac{\partial \vec{v_g}}{\partial z} = \frac{g}{f T} \mathbf{k} \times \nabla T, where g is gravitational acceleration, T is temperature, and \nabla T is the horizontal temperature gradient; this indicates that warmer air to the south (in the Northern Hemisphere) produces westerly shear, increasing westerly winds with height. Here, \rho represents mass per unit volume, typically around 1.2 kg m^{-3} near the surface in the atmosphere but varying with height and conditions, while pressures are in pascals and gradients in Pa m^{-1}, yielding wind speeds in m s^{-1}. In quasi-geostrophic theory, the geostrophic wind can be related to a streamfunction \psi, such that \vec{v_g} = \mathbf{k} \times \nabla \psi, providing a diagnostic for incompressible, non-divergent where \psi is proportional to or pressure perturbations.

Applications in Fluid Systems

Atmospheric Winds

In large-scale atmospheric flows, the geostrophic wind is particularly applicable to upper-level circulations, where winds parallel the or contours on constant-pressure surfaces such as the 500 level. At this mid-tropospheric altitude, typical geostrophic wind speeds range from 20 to 50 m/s, especially within jet streams that form due to strong thermal contrasts and Coriolis effects. These winds exhibit minimal deviation from geostrophic balance above the , facilitating the transport of air masses across synoptic scales. The plays a central role in synoptic for the evolution of cyclones and anticyclones, where quasi- govern the large-scale patterns and associated fields. In extratropical cyclones, geostrophic winds drive the around low- centers, while in anticyclones, they support anticyclonic flow around highs, with wind shifts often signaling the passage of fronts. This balance allows meteorologists to predict system tracks and intensity changes by analyzing anomalies, as the simplifies the of baroclinic instabilities that fuel these systems. Observationally, geostrophic winds are inferred from charts derived from data or observations, where the spacing of height contours directly indicates via the geostrophic , with low pressure to the left in the . Actual winds are then compared to these estimates using wind adjustments, which account for in flow around , typically reducing speeds by 10-20% in cyclones relative to pure geostrophic values. This comparison reveals small but systematic deviations, enhancing the accuracy of real-time analyses. In modern numerical weather prediction models like the ECMWF Integrated Forecasting System and the NOAA (GFS), geostrophy serves as a foundational constraint for initializing balanced flows, ensuring that initial conditions align pressure gradients with Coriolis forces to minimize spurious gravity waves. Post-2000 advancements in ensemble forecasting, such as the ECMWF's singular vector perturbations and GFS's updates, incorporate geostrophic balance to generate probabilistic predictions of cyclone tracks and evolutions, improving for medium-range synoptic events. Near the surface, introduces deviations from , leading to winds where causes cross-isobar flow toward low pressure and reduced speeds compared to geostrophic, as described by . However, geostrophy dominates above the , typically at heights exceeding 1 km, where frictional effects diminish and flows revert to quasi-geostrophic conditions. This vertical transition is critical for distinguishing boundary-layer phenomena from free-atmospheric circulations in weather analysis.

Oceanic Currents

In , geostrophic balance plays a central role in describing large-scale interior currents, such as the , where the balances the horizontal across sloped isopycnal surfaces. These isopycnals, or surfaces of constant , tilt in response to horizontal variations, driving geostrophic flows that dominate the subtropical gyre circulations in the absence of frictional influences near boundaries. Unlike atmospheric winds, which are primarily driven by pressure gradients in a nearly , oceanic geostrophic currents exhibit a stronger dependence on , quantified by potential density (σ_t), which contributes significantly to the horizontal (∇p). This effect manifests in both barotropic modes, where pressure gradients are depth-independent due to uniform , and baroclinic modes, where variations introduce vertical structure in the flow. The thermal wind relation further elucidates the vertical structure of these oceanic geostrophic currents, stating that the vertical of the geostrophic velocity (\partial v_g / \partial z) is proportional to the horizontal gradients of and , which alter and thus the field. In baroclinic conditions, warmer surface waters to the south create southward gradients that support northward geostrophic in the , contributing to the eastward intensification of zonal flows within gyres by enhancing velocity at upper levels relative to deeper waters. This is particularly evident in western boundary currents like the , where isopycnal slopes steepen, amplifying the baroclinic component of the transport. Geostrophic currents in the are commonly measured using hydrographic sections that profile and to infer and anomalies, with velocities computed relative to a deep reference level, typically around 1000 m, where flows are assumed negligible. Instruments such as expendable bathythermographs (XBTs) provide rapid profiles along transects, enabling estimation of geostrophic shear from , while is often climatologically inferred or measured via conductivity-temperature-depth (CTD) casts for more precise calculations (σ_t). These methods yield depth-integrated transports that capture both barotropic and baroclinic components, distinguishing applications from atmospheric ones by emphasizing effects. A prominent example of a nearly purely geostrophic current is the (), which encircles without eastern boundaries to disrupt its zonal flow, maintaining balance over its full depth. The 's transport, estimated at approximately 100 Sverdrups (Sv; 1 Sv = 10^6 m³/s) above 3000 m, arises from wind-driven pressure gradients in geostrophic equilibrium, with minimal barotropic compensation due to the circumpolar geometry. This flow highlights the dominance of baroclinic modes in the , where density fronts along isopycnals sustain intense vertical shears, paralleling but exceeding the scale of atmospheric jet streams in vertical extent.

Limitations and Validity

Approximation Conditions

The geostrophic approximation is valid for flows on large horizontal scales, typically exceeding 100 km in the atmosphere and tens of kilometers in the , where the Ro = \frac{U}{f L} is small (typically Ro < 0.1 for synoptic-scale systems), indicating that the Coriolis force dominates over inertial accelerations. Here, U represents the characteristic horizontal velocity, f = 2 \Omega \sin \phi is the Coriolis parameter with \Omega as Earth's angular velocity and \phi as latitude, and L is the horizontal length scale; for synoptic-scale mid-latitude weather systems with U \approx 10 m/s and L \approx 1000 km, Ro \approx 0.1, ensuring rotational effects prevail. As higher values amplify nonlinear advection relative to the Coriolis term, small Ro supports the balance. This approximation requires sufficient latitude, away from the equator where | \phi | > 5^\circ - 10^\circ, to ensure f is significant and avoids breakdown in the ; near the , small f leads to Ro \gg 1, invalidating geostrophy, as seen in tropical cyclones with scales of 100-500 km and winds exceeding 50 m/s yielding Ro > 1. In mid-latitudes (| \phi | \approx 30^\circ - 60^\circ), f \approx 10^{-4} s^{-1} supports the balance for large-scale flows. Flows must be steady or slowly varying, with time scales longer than the inertial period $2\pi / f, approximately 12-24 hours in mid-latitudes, allowing adjustment to geostrophic balance over several such periods. The adjustment time is on the order of $1/f, ensuring transient accelerations remain negligible compared to the . Friction is negligible when the Ekman number Ek = \frac{\nu}{f L^2} (or equivalently \frac{\nu}{2 \Omega L^2}) is small, Ek \ll 1, where \nu is , minimizing viscous effects relative to in the interior . terms are minor in regimes of low Fr = \frac{U}{N H} \ll 1, with N as the buoyancy frequency and H the vertical scale, enforcing and suppressing vertical accelerations that could disrupt the horizontal force equilibrium.

Deviations and Corrections

In real atmospheric and oceanic flows, frictional effects in the boundary layer cause significant deviations from ideal geostrophic balance, primarily by introducing drag that reduces wind or current speeds and alters directions. Near the surface, friction opposes motion, leading to a spiraling velocity profile known as the Ekman spiral, where the surface flow is deflected about 45 degrees from the geostrophic direction and speeds are reduced by approximately 20-50% compared to the free-stream geostrophic value. This effect is prominent in both the atmospheric planetary boundary layer and the oceanic surface Ekman layer, where turbulent viscosity transfers momentum downward, resulting in a net transport perpendicular to the wind stress. Curvature in the flow around high- or low-pressure systems introduces additional deviations through the centrifugal force, requiring a modification to geostrophic balance known as the gradient wind approximation. In cyclonic systems, such as extratropical lows, the gradient wind is subgeostrophic, meaning actual speeds are slower than the geostrophic wind to balance the inward centrifugal force against the pressure gradient and Coriolis terms. Conversely, in anticyclonic systems like highs, the gradient wind is supergeostrophic, with faster speeds to counteract the outward centrifugal force. These adjustments are crucial for tight curvature radii, typically on the order of 500-1000 km in mid-latitude weather systems, where the centrifugal term becomes comparable to the Coriolis force. Ageostrophic components arise from non-balanced accelerations, particularly in regions of or such as frontal zones, where the Q-vector—a diagnostic tool derived from quasigeostrophic theory—quantifies the forcing for vertical motion and associated horizontal ageostrophy. In fronts, Q-vector drives ageostrophic circulations that enhance frontogenesis, leading to divergence aloft and near the surface. During Rossby adjustment processes, initial imbalances between mass and momentum fields generate transient ageostrophic accelerations, propagating as inertia-gravity waves until geostrophic equilibrium is restored over timescales of hours to days, depending on the Rossby deformation radius. To correct for these deviations in frontal zones, the Sawyer-Eliassen equation provides a for diagnosing transverse ageostrophic circulations driven by along-front variations in geostrophic wind and , enabling predictions of secondary flows that modify the primary geostrophic balance without full numerical simulation. In (NWP) models, hybrid approaches integrate —which inherently capture ageostrophic components—with techniques to refine forecasts, reducing errors from geostrophic approximations in dynamic regions like fronts and cyclones. Recent studies highlight how exacerbates deviations from geostrophic regimes in the through poleward shifts in storm tracks and , altering the effective Coriolis parameter f (which increases with ) and intensifying activity by up to 10-20% in wind speeds since the 1950s. These shifts, observed in analyses up to 2021, lead to modified geostrophic wind patterns associated with reduced and weakened meridional temperature gradients.

References

  1. [1]
    Chapter 10: Atmospheric Forces and Wind - UH Pressbooks
    When in geostrophic balance, wind in the atmosphere has a balance between the pressure gradient force and the Coriolis force. In geostrophic balance, PGF = CF.
  2. [2]
    Geostrophic Flow
    The Geostrophic Wind Equation. Geostrophic flow in the atmosphere is the most applied balanced flow among the five that are defined.
  3. [3]
    Meteorology 2/3 - PSU E-Wall
    This type of balance, called geostrophic balance by meteorologists, causes wind to move parallel to isobars. This case can be seen in Figure 4.
  4. [4]
    [PDF] The Geostrophic Wind
    These two expressions represent geostrophic (“geo” for Earth and “strophic” for turning) balance relating the horizontal pressure and wind fields. • Letting and ...
  5. [5]
    [PDF] Geostrophic Balance • Gradient Wind
    The geostrophic wind balance assumes that the wind flow is straight. • A more general form of a balanced wind can be obtained if accelerations due to ...
  6. [6]
    10.8 A Closer Look at the Four Force Balances | METEO 300
    In geostrophic flow only the pressure gradient force and the Coriolis force are important. Pressure gradient points to low pressure on the height surface or ...Missing: definition | Show results with:definition
  7. [7]
    winds balanced by the Coriolis and Pressure Gradient forces
    Geostrophic Wind. winds balanced by the Coriolis and Pressure Gradient forces. An air parcel initially at rest will move from high pressure to low pressure ...Missing: definition | Show results with:definition
  8. [8]
    [PDF] MET 200 Lecture 10 Forces and Wind - SOEST Hawaii
    30. Geostrophic Wind. The Geostrophic wind is flow in a straight line in which the pressure gradient force balances the Coriolis force. 994 mb. 996 mb.Missing: definition | Show results with:definition
  9. [9]
    GG 140 - Lecture 13 - Global Climate and the Coriolis Force
    The majority of large-scale motion in the atmosphere is in geostrophic balance, meaning the Coriolis force acting on the motion is balanced by a pressure ...
  10. [10]
    [PDF] Applications of the Basic Equations Chapter 3
    Geostrophic balance: Coriolis and pressure gradient force in exact balance (equal magnitude, opposite direction). Flow is parallel to contours: Flow follows a ...
  11. [11]
    How Do We Understand the Coriolis Force? - AMS Journals
    This discovery attracted wide scientific and popular attention and, together with reading Newton's and Laplace's works, inspired William Ferrel in 1856 to ...
  12. [12]
    [PDF] Dynamic meteorology and hydrography - Survivor Library
    DYNAMIC METEOROLOGY AND HYDROGRAPHY. Part I. STATICS. BY. V. BJERKNES and J. W. SANDSTROM. Page 14. Page 15. CHAPTER I. SYSTEM OF UNITS. 1. Meter-Ton-Second ...
  13. [13]
    [PDF] Vilhelm and Jacob Bjerknes - the NOAA Institutional Repository
    Jan 16, 2004 · Studying these pictures led Bjerknes and other members of the Bergen School to conclude that there exist air masses that differ f each other and ...
  14. [14]
    WALFRID EKMAN (1874–1954): THEORETICAL OCEANOGRAPHER
    Aug 30, 2022 · The Swedish oceanographer Vagn Walfrid Ekman (1874–1954, see Figure 1) is best known for his work on shallow ocean circulation, ...
  15. [15]
    A New Look at the Physics of Rossby Waves: A Mechanical–Coriolis ...
    Jan 1, 2013 · Introduction. Rossby waves, named after the pioneering work of Rossby (1939), play a fundamental role in large-scale atmospheric and oceanic ...
  16. [16]
    operational numerical weather prediction - AMS Journals
    Fred Shuman (left) and Otha Fuller circa 1955 at the IBM was the first computer used by the J N W P U to produce numerical weather prediction.Missing: geostrophic | Show results with:geostrophic
  17. [17]
    [PDF] Geostrophic balance | Weather and Climate Laboratory
    We describe the theory of Geostrophic Balance, derive key equations and discuss associated physical balances.1. 1 Geostrophic balance. If the flow is such ...
  18. [18]
    [PDF] Geostrophic balance Thermal wind Dynamic height
    Oct 24, 2019 · The vertical shear in geostrophic velocity is proportional to the horizontal density derivative. Formally: - f∂v/∂z = (g/ρo)∂ρ/∂x f∂u/ ...
  19. [19]
    [PDF] The Thermal Wind
    The Thermal Wind. • As the name implies, the thermal wind is an expression for the relationship between the temperature and wind fields in the atmosphere.
  20. [20]
    [PDF] EAPS 53600, Spring 2020 Lec 05: Quasi-geostrophic (QG) system 1 ...
    Because the geostrophic wind (for f = f0 constant) is non-divergent, we can write the horizontal flow field in terms of a streamfunction, ψ: ug = −. ∂ψ. ∂y.
  21. [21]
    CHART COMPARISON
    The 500 and 300 mb charts are considered the upper levels of the atmosphere. ... Jet streaks are segments of faster wind speed within the jet stream. At ...
  22. [22]
    [PDF] Forces that Affect Atmospheric Motion - UCI ESS
    The jetstream is a narrow band of strong winds that encircles the Earth in the mid- latitude. The band of strongest winds is typically 300 to 500 km wide and ...
  23. [23]
    Revisiting an Old Concept: The Gradient Wind* in - AMS Journals
    The gradient wind is defined as a horizontal wind having the same direction as the geostrophic wind but with a magnitude consistent with a balance of three ...
  24. [24]
    [PDF] synoptic_textbook.pdf - EUMeTrain
    They are also turbulent, and the geostrophic approximation does not apply in the ... Anomalies that travel with the upper waves are called cyclones and ...
  25. [25]
    Diagnosing model performance in the tropics | ECMWF
    The recently developed MODES tool provides such a decomposition into balanced (Rossby) and unbalanced (IG) modes.
  26. [26]
    A Four-Dimensional Incremental Analysis Update for the Ensemble ...
    Jul 2, 2016 · The 4DIAU helps to reduce the imbalance caused by the EnKF and retain the propagation of analysis increments compared to 3DIAU. A dry two-layer ...
  27. [27]
    30 years of ensemble forecasting at ECMWF
    Nov 24, 2022 · ECMWF was one of the first centres to introduce operational ensemble forecasts, a move which 30 years ago initiated a distinct shift in thinking.A Brief History Of Ensemble... · 2000s And Onwards... · The Value Of Ensemble...Missing: GFS geostrophic balance balanced flows
  28. [28]
    [PDF] The Relationships among Wind, Horizontal Pressure Gradient, and ...
    Using the surface friction to represent the vertical divergence of the momentum flux, Macklin et al. (1988) investigated geotriptic balance for offshore flows.
  29. [29]
    Untitled - AMS Journals
    This is a region of transition between the friction layer and the quasi-geostrophic free atmosphere and extends to heights of order 1 km. At levels well into ...
  30. [30]
    [PDF] Chapter 3 The Coriolis force, geostrophy, Rossby waves and the ...
    The term has been adopted in oceanography; a current which is recognized by sloping isopycnals is sometimes called a "thermal wind". For completeness we ...
  31. [31]
    [PDF] a Coriolis tutorial, Part 4: - Wind-driven ocean circulation
    Geostrophic balance may be understood using an f-plane approximation in which the latitudinal dependence of the Coriolis parameter (5) is, purely for ...
  32. [32]
    [PDF] Chapter 10 The wind-driven circulation
    The resulting horizontal density gradient supports a thermal wind shear. Its baroclinic instability spawns an energetic eddy field which tends to flatten ...
  33. [33]
    [PDF] Gravitational, Symmetric, and Baroclinic Instability of the Ocean ...
    then horizontal density gradients will be induced and, within a rotation period, horizontal currents in thermal-wind balance with those gradients will be ...
  34. [34]
    [PDF] GEOSTROPHY AND POTENTIAL VORTICITY IN THE GULF ...
    Is the Gulf Stream in geostrophic balance (i.e. are the horizontal pressure gradient and the Coriolis force the only significant terms in the momentum balance)? ...
  35. [35]
    Temporal variability of the South Atlantic Meridional Overturning ...
    Sep 1, 2015 · Using the synthetic T/S profiles, the geostrophic velocities were computed from the thermal wind relationship with the reference level at 1000 m ...
  36. [36]
    Estimating the Velocity and Transport of Western Boundary Current ...
    Jun 30, 2018 · (c) Geostrophic velocity at 1000-m depth along PX30 nominal transect averaged between 2004 and 2015. Absolute velocity is estimated using ...
  37. [37]
    Meridional heat transport determined with expandable ...
    The oceanographic data used for the study consist of Expendable bathythermograph (XBT) data collected along 27 sections at nominally 35°S for the period of ...
  38. [38]
    More Than 50 Years of Successful Continuous Temperature Section ...
    Jul 23, 2019 · Geostrophic velocity estimated from each AX10 section using temperature measurements from XBTs and salinity inferred from the historical T-S ...
  39. [39]
    On the meridional extent and fronts of the Antarctic Circumpolar ...
    Between the STF and the southern boundary of the ACC, the shear transport of the circumpolar current above 3000 m is at all longitudes about 100 Sv (1 Sv = 106 ...
  40. [40]
    Transport and variability of the Antarctic Circumpolar Current in ...
    The baroclinic transport of the Antarctic Circumpolar Current (ACC) above 3000 m through Drake Passage is 107.3 ± 10.4 Sv and has been steady between 1975 and ...
  41. [41]
    [PDF] SIO 210: Southern Ocean circulation
    Antarctic Circumpolar Current encircling Antarctica, starting farthest north off S. ... Transport of about 100 Sv, top to bottom, including all fronts (and ...
  42. [42]
    [PDF] Chapter 9 Geostrophy, Quasi-Geostrophy and the Potential Vorticity ...
    From the geostrophic streamfunction we obtain directly the horizontal velocities and the density field from the x, y and z derivatives. u = −ψy,v = ψx ,ρ = −ρs.
  43. [43]
    Rossby Number - an overview | ScienceDirect Topics
    The Rossby number is defined as the ratio of characteristic inertial to Coriolis accelerations in fluid dynamics, expressed as \( R_0 = \frac{U}{fL} \) ...
  44. [44]
    5.2 Geostrophic balance - Atmospheric Physics - Fiveable
    Geostrophic balance governs large-scale atmospheric motions, balancing pressure gradient and Coriolis forces. This equilibrium results in steady-state flow ...
  45. [45]
    Ekman transport and pumping in the California Current based on the ...
    A friction correction (30% speed reduction, 15° shift) is applied to these geostrophic winds before the upwelling estimate is calculated. [11] The satellite ...Missing: spiral percentage
  46. [46]
  47. [47]
    9.3: Force Balances - Geosciences LibreTexts
    Oct 20, 2025 · Compared to geostrophic winds, gradient winds feature a balance between the Coriolis force, the pressure gradient force, and the centrifugal ...
  48. [48]
    [PDF] Lecture 4: Pressure and Wind - UCI ESS
    For low pressure system →gradient wind < geostrophic wind →subgeostropic. Why didn't the strong vertical pressure gradient push the air rise?
  49. [49]
    Ageostrophic Pseudovorticity and Geostrophic C-Vector Forcing—A ...
    Abstract. By combining the two Q-vector component equations with the third quasigeostrophic (QG) diagnostic equation (the vertical ageostrophic vorticity ...
  50. [50]
    [PDF] Laboratory experiments on nonlinear Rossby adjustment in a channel
    Behind the shock was a boundary layer of approximately one deformation radius in width, in which the flow was strongly ageostrophic with a large off-shore ...
  51. [51]
    [PDF] Sawyer-Eliassen equation - Atmospheric Sciences
    ... Sawyer-Eliassen equation (hereafter, the SE equation), provides the foundation for understanding ageostrophic circulations in the vicinity of frontal zones.
  52. [52]
    [PDF] Models, Numerical Methods, and Data Assimilation
    The quasi- geostrophic theory was derived to explain the circulations of synoptic flow in the midlatitudes. Conditions for baroclinic instability were also ...
  53. [53]
    Arctic cyclones have become more intense and longer-lived over the ...
    Oct 2, 2023 · The results indicate an intensification of Arctic cyclone activity over the last seven decades. There has been a long-term shift of the maximum cyclone counts.Missing: post- | Show results with:post-
  54. [54]
    [PDF] Poleward shift of the major ocean gyres detected in a warming climate
    Under the constraint of geostrophic balance, the Coriolis ... wind pattern change implies that the extratropical atmospheric surface winds have shifted.Missing: regimes post-