Fact-checked by Grok 2 weeks ago

Thermal wind

The thermal wind is a fundamental concept in representing the vertical shear of the caused by horizontal gradients in the atmosphere. It is defined as the difference between the at an upper level and that at a lower level, such that \mathbf{V}_T = \mathbf{V}_g(\text{upper}) - \mathbf{V}_g(\text{lower}), and it arises in baroclinic atmospheres where variations with height lead to changes in and . This shear is not an actual wind but a diagnostic tool that links thermal structure to , with the wind pointing parallel to isotherms and having cold air to its left in the . The thermal wind relation is derived from the geostrophic balance equations, v_g = \frac{1}{f} \frac{\partial \Phi}{\partial x} and u_g = -\frac{1}{f} \frac{\partial \Phi}{\partial y}, combined with the hydrostatic equation \frac{\partial \Phi}{\partial p} = -\frac{R T}{p} and the , yielding \mathbf{V}_T = \frac{R}{f} \hat{k} \times \nabla_p T \log\left(\frac{p_1}{p_2}\right), where f is the Coriolis parameter, R is the for dry air, and \nabla_p T is the horizontal on a pressure surface. This formulation shows that the magnitude and direction of the wind are directly proportional to the , with stronger gradients producing greater ; for instance, the x-component is u_T = -\frac{R}{f} \left\langle \frac{\partial T}{\partial y} \right\rangle \log\left(\frac{p_1}{p_2}\right) and the y-component is v_T = \frac{R}{f} \left\langle \frac{\partial T}{\partial x} \right\rangle \log\left(\frac{p_1}{p_2}\right). In midlatitudes, the thermal wind explains the predominantly westerly (eastward) flow of surface and upper-level winds, as the poleward decrease in temperature creates an equatorward-pointing that, under geostrophic balance, results in a reinforcing westerly winds with height, contributing to phenomena like the . It is crucial for , as it ensures consistency between observed wind, , and temperature fields, and helps predict temperature —warm advection causes winds to veer with height, while cold advection causes backing. The concept applies primarily to large-scale, synoptic flows where geostrophic and hydrostatic approximations hold, contrasting with barotropic conditions lacking such .

Conceptual Foundation

Definition

The thermal wind is defined as the vector difference between the geostrophic winds at two different levels in the atmosphere, resulting exclusively from horizontal gradients in a rotating such as Earth's atmosphere. This difference represents the vertical shear of the , which is parallel to the isotherms and to lines of constant thickness between the levels, with its magnitude proportional to the strength of the horizontal contrast. In essence, the thermal wind quantifies how variations across horizontal distances induce changes in and direction with height, a key feature in baroclinic conditions where isobaric and isosteric surfaces are not coincident. The itself is the hypothetical wind that arises from a balance between the acting on moving air and the horizontal , approximating large-scale atmospheric flow in the absence of . Horizontal gradients, expressed mathematically as ∂T/∂x or ∂T/∂y, reflect spatial variations in that alter air and thus at different heights, leading to the thermal wind shear. These gradients play a central role in baroclinic atmospheres, where horizontal differences prevent surfaces from being level, fostering vertical variations in geostrophic flow that the thermal wind captures. The term "thermal wind" was proposed in 1918 by British meteorologist , emerging from early 20th-century advancements in synoptic meteorology that integrated temperature observations with upper-air wind patterns. It provided a diagnostic tool to relate surface maps to upper-level , influencing the development of . The thermal wind is typically expressed in units of meters per second (m/s) for the total vector difference between levels, though it is often conceptualized as in m/s per kilometer of height; in mid-latitudes, magnitudes commonly range from 10 to 50 m/s over the tropospheric depth, reflecting strong meridional temperature contrasts.

Physical Mechanism

The thermal wind arises from the buoyancy-driven response of the atmosphere to horizontal gradients. In regions where air is warmer, such as toward the compared to the poles, the less dense warm air expands vertically, resulting in a slower decrease of with within those columns. This expansion causes pressure surfaces to rise higher over warm areas than over adjacent cold regions, where denser air leads to a more rapid pressure drop. Consequently, at upper levels, the horizontal becomes steeper over the warm side, enhancing the aloft and driving stronger geostrophic winds at higher altitudes. Earth's rotation plays a crucial role through the Coriolis effect, which deflects the response to these pressure differences. As the enhanced aloft accelerates air masses, the turns the flow perpendicular to the gradient, aligning geostrophic winds parallel to the isotherms—the lines of constant temperature. This deflection ensures that the vertical shear in the wind, known as the thermal wind, points perpendicular to the (parallel to the isotherms), with cold air typically to the left in the [Northern Hemisphere](/page/Northern Hemisphere). The result is a systematic increase in with height, particularly in the westerly direction in mid-latitudes, without altering the wind direction at the surface. This phenomenon is fundamentally tied to baroclinicity, the condition where isotherms and isobars are not aligned, allowing temperature variations on constant-pressure surfaces. In baroclinic atmospheres, prevalent in mid-latitudes, these misalignments produce the vertical characteristic of the thermal wind. By contrast, in barotropic conditions—common in the deep tropics—temperature surfaces parallel pressure surfaces, resulting in uniform density with height and no thermal wind shear, as geostrophic winds remain constant across levels. Qualitatively, this can be visualized through sketches of sloped isobaric surfaces: over a warm region, isobars tilt upward with height, creating a funnel-like that amplifies from surface to upper , while wind vectors depict a progressive strengthening and alignment parallel to the underlying isotherms. Such diagrams illustrate how the initial surface , often modest due to balanced surface temperatures, evolves into pronounced upper-level gradients solely from effects.

Theoretical Framework

Geostrophic Balance Prerequisites

The hydrostatic approximation forms a fundamental prerequisite for deriving the thermal wind relation, positing that vertical accelerations in the atmosphere are negligible compared to , thereby simplifying the vertical equation. This assumption yields the hydrostatic balance equation, expressed as \frac{\partial p}{\partial z} = -\rho [g](/page/G), where p is , z is , \rho is air , and g is . The approximation holds because typical vertical velocities in large-scale atmospheric flows are on the order of centimeters per second, resulting in accelerations far smaller than g \approx 9.8 \, \mathrm{m/s^2}, allowing to decrease with height primarily due to the weight of the overlying air column. Complementing the hydrostatic approximation, geostrophic assumes that, for large-scale horizontal flows, the counteracts the , neglecting local accelerations, , and curvature effects. This leads to the geostrophic wind equation: f \mathbf{k} \times \mathbf{V}_g = -\frac{1}{\rho} \nabla p, where f = 2 \Omega \sin \phi is the Coriolis parameter (\Omega is Earth's and \phi is ), \mathbf{V}_g is the geostrophic velocity, and \nabla p is the horizontal . Under steady-state conditions, this implies that winds flow parallel to isobars, with speed inversely proportional to f and directly to the pressure gradient magnitude. These approximations are valid primarily for synoptic-scale motions in mid-latitudes, where horizontal scales exceed approximately 100 , ensuring a small (Ro = U / f L < 0.1) that justifies neglecting accelerations relative to Coriolis forces, and where f \neq 0 to avoid equatorial singularities. They break down near the equator due to vanishing f, in small-scale turbulent flows (length scales <10 ) where friction dominates, or during rapid transients like convective storms where accelerations become significant. Hydrostatic validity similarly requires horizontal scales much larger than the atmospheric depth scale (~10 ) to minimize vertical motion effects. The thermal wind relation is derived using the hydrostatic and geostrophic approximations. These approximations are central to the scale analysis that justifies the primitive equations, which govern large-scale atmospheric motion and include the hydrostatic balance, continuity, and thermodynamic equations while retaining time-dependent terms for prognostic evolution. These equations, derived from scale analysis of large-scale flows, provide the balanced state from which vertical wind shear due to horizontal temperature gradients—the essence of thermal wind—can be diagnosed without resolving smaller-scale dynamics. Seminal developments in primitive equation models, such as those used in global circulation simulations, underscore their role in capturing baroclinic instabilities central to thermal wind phenomena.

Derivation and Equations

The derivation of the thermal wind relation begins with the geostrophic balance equations in pressure coordinates, where the horizontal pressure gradient force is balanced by the Coriolis force. The geostrophic wind components are given by u_g = -\frac{1}{f} \left( \frac{\partial \Phi}{\partial y} \right)_p, \quad v_g = \frac{1}{f} \left( \frac{\partial \Phi}{\partial x} \right)_p, where \Phi is the geopotential height, f = 2 \Omega \sin \phi is the (assumed constant under the f-plane approximation), and subscripts denote evaluation on constant-pressure surfaces. To obtain the vertical shear, differentiate these equations with respect to \ln p, using the hydrostatic relation \frac{\partial \Phi}{\partial p} = -\frac{R T}{p}, where R is the gas constant for dry air and the ideal gas law p = \rho R T holds. Interchanging the order of differentiation (valid under horizontal incompressibility, assuming no significant horizontal variations in density affect the vertical derivatives) yields the scalar components of the thermal wind shear: \frac{\partial u_g}{\partial \ln p} = \frac{R}{f} \left( \frac{\partial T}{\partial y} \right)_p, \quad \frac{\partial v_g}{\partial \ln p} = -\frac{R}{f} \left( \frac{\partial T}{\partial x} \right)_p. In vector form, this is \frac{\partial \vec{V}_g}{\partial \ln p} = \frac{R}{f} \hat{k} \times \nabla_p T, where \vec{V}_g = (u_g, v_g) and \nabla_p is the horizontal gradient on pressure surfaces; note that \ln p increases downward, so the shear direction reflects changes from lower to upper levels. For the integrated form between two pressure levels p_1 (lower) and p_2 (upper, p_1 > p_2), subtract the geostrophic winds at these levels: \vec{V}_T = \vec{V}_{g2} - \vec{V}_{g1} = \frac{R \ln(p_1 / p_2)}{f} \hat{k} \times \nabla_p \bar{T}, where \bar{T} is the mean temperature in the layer, derived by integrating the differential form assuming constant gradients. In height coordinates, the equivalent integrated thermal wind uses the hypsometric relation for layer thickness \Delta z = \frac{R \bar{T}}{g} \ln(p_1 / p_2), leading to \vec{V}_T = \frac{g}{f} \hat{k} \times \nabla_h \Delta z = \frac{R \ln(p_1 / p_2)}{f} \hat{k} \times \nabla_h \bar{T}, where \nabla_h is the horizontal gradient and g is gravitational acceleration; the differential shear is \frac{\partial \vec{V}_g}{\partial z} = \frac{g}{f T} \hat{k} \times \nabla_h T under approximations of small temperature variations. The vector form indicates that the thermal wind is perpendicular to the horizontal temperature gradient \nabla T and parallel to contours of constant thickness (which approximate isotherms for mean layer temperature).

Meteorological Applications

Jet Stream Dynamics

Jet streams represent narrow bands of strong westerly winds in the upper troposphere, primarily driven by the thermal wind relation arising from meridional temperature gradients. In mid-latitudes, particularly along the polar front where cold polar air meets warmer mid-latitude air, these gradients create significant baroclinicity, leading to vertical wind shear through thermal wind effects. This shear typically ranges from 5 to 10 m/s per km in the troposphere, resulting in total increases of 20-50 m/s over altitudes of 5-10 km, which aligns with the formation of the polar jet around 50-60° latitude and the subtropical jet near 30° latitude in both hemispheres. The vertical structure of jet streams manifests as a maximum in the thermal wind profile, where geostrophic winds accelerate with height due to the equatorward temperature increase, reaching speeds of 25-100 m/s in the polar at approximately 9-11 km altitude near the . This height-dependent speedup continues until the tropopause level, where the diminishes, forming a that is roughly 5-10 km thick vertically and 1,000-2,000 km wide horizontally. As per the thermal wind equation, this reflects the integrated effect of horizontal temperature contrasts, concentrating the strongest winds aloft. Seasonal variations in intensity stem from changes in meridional contrasts, with stronger gradients in winter enhancing thermal wind and thus jet speeds, while summer warming reduces these contrasts and weakens the jets. In winter, polar jets achieve speeds of 25-100 m/s, positioned between 30-60°N/S, whereas in summer, speeds drop to 0-45 m/s, and the jets shift poleward to around 40-45° , often merging into a single band. These patterns are evident in long-term reanalysis data, underscoring the thermal wind's role in modulating jet persistence and location. Under , increasing global temperatures are projected to enhance meridional temperature gradients in the upper , leading to stronger thermal wind and faster upper-level jet streams, particularly in the extratropics, with increases of about 2.1% per of warming as of studies through 2023. Thermal wind also influences propagation within jet streams, where the vertical modulates wave amplitudes by altering the mean flow and baroclinicity that support planetary-scale meanders. These meanders, or , form troughs and ridges in the jet, with thermal wind-driven wind variations affecting wave growth and eastward propagation speeds, typically guiding mid-latitude weather systems.

Frontogenesis Processes

Frontogenesis involves the intensification of horizontal s, which, through the thermal wind relation, directly enhances the vertical of the and drives ageostrophic circulations that further sharpen frontal boundaries. Horizontal contraction of these gradients, often via advective processes, increases the magnitude of the vector (∇T), thereby amplifying the thermal wind and promoting baroclinic in the atmosphere. This results in a feedback loop where the strengthened induces convergent motions, concentrating isentropes and accelerating front development. Kinematic frontogenesis specifically arises from deformation fields in the horizontal wind, characterized by in one direction and in the perpendicular direction, which deform and amplify ∇T. As the temperature contrast sharpens, the thermal wind equation dictates a corresponding increase in the vertical derivative of the horizontal wind (∂V/∂z), boosting the baroclinicity and along the front. This process is particularly effective in regions of geostrophic deformation, where the non-divergent flow initially tightens gradients, setting the stage for subsequent ageostrophic enhancements. A prominent example occurs during the sharpening of cold fronts in extratropical cyclones, where the thermal wind aligns parallel to the frontal zone, generating substantial vertical shear that aligns with the temperature contrast. In such systems, this shear contributes to intense and dynamic lifting along the front. The alignment enhances the front's slope and narrows its width, exemplifying how thermal wind imbalances sustain the cyclone's vigor. The Sawyer-Eliassen equation conceptually links these dynamics by modeling the cross-frontal ageostrophic circulation as a response to geostrophic forcing, including thermal wind imbalances that drive thermally direct circulations. This circulation converges air toward the frontal surface on the warm side and diverges on the cold side, further intensifying ∇T and the associated without requiring explicit derivation of the thermal wind itself. Such circulations are essential for maintaining frontogenetic tendencies in balanced flows.

Thermal Advection Effects

Horizontal temperature advection alters thermal wind profiles by changing the horizontal over time, which directly influences the direction and magnitude of the vertical . In regions of warm , such as southwesterly geostrophic flow transporting warmer air equatorward of a front, the thermal wind vector veers clockwise with height, reflecting a rotation in the shear direction due to the modified . Conversely, cold , often associated with northeasterly flow, causes the thermal wind to back anticlockwise with height. This directional change arises because perturbs the geostrophic balance, prompting ageostrophic circulations to adjust the temperature field and restore equilibrium. The process involves advective modifications to the that rotate the thermal wind vector, typically resulting in 45-90° turns over the depth of the during active synoptic conditions. These turns are conceptually linked to Q-vector , where the pattern of Q-vectors—representing the of thermal wind by geostrophic flow—drives vertical motions that further influence the shear profile. In warm scenarios, Q-vector in the lower enhances ascent, which can amplify the veering by reinforcing the temperature perturbations. A prominent synoptic example occurs in mid-latitude s, where the advects moist, warm air northward ahead of the , strengthening the thermal wind while turning it clockwise with height to maintain baroclinicity. This advection-driven veering contributes to the 's intensification and guides its track by aligning the upper-level jet with the surface low. Such dynamics highlight how thermal not only alters local profiles but also shapes larger-scale evolution, occasionally initiating frontogenesis through gradient sharpening.

Observational and Diagnostic Uses

Measurement Techniques

observations offer a direct method for measuring vertical profiles of and as functions of height or , enabling the computation of thermal wind as the vertical in between standard levels, such as 850 and 200 . These balloon-borne instruments ascend through the atmosphere at approximately 5-6 m/s, collecting data at resolutions of 6-10 seconds for and , and deriving from GPS positioning or tracking. The resulting profiles allow application of the thermal wind to quantify horizontal gradients driving upper-level changes, with validations showing good agreement between observed and geostrophically computed shears in mid-latitude soundings. Satellite-based techniques estimate thermal wind by retrieving profiles from infrared sounders, such as the TIROS Operational Vertical Sounder (TOVS), which measure brightness temperatures to infer horizontal temperature gradients via hydrostatic and thermal wind balance. These retrievals incorporate variational methods to enforce mass conservation, combining temperature data with parameterizations for wind estimation across the . Validations against ground stations indicate mean absolute errors of approximately 1 m/s in polar regions, demonstrating utility for large-scale thermal wind mapping where direct observations are sparse. Reanalysis datasets like ERA5 from the European Centre for Medium-Range Weather Forecasts integrate global observations and model outputs to provide consistent fields of , , and on levels, facilitating thermal wind computation through differences in geostrophic winds between levels such as 1000 and 500 . Tools like Metview process ERA5 geopotential data to derive thickness fields (e.g., 500-1000 ), from which thermal wind vectors are calculated assuming geostrophic balance, enabling vertical integration for synoptic-scale maps. This approach supports global monitoring of thermal wind patterns, with ERA5's 31 km horizontal resolution and hourly temporal coverage enhancing diagnostic applications over decades. Modern advancements include Doppler lidars for high-resolution and profiling in the middle atmosphere, as demonstrated in post-2010 field campaigns like the Calibration/Validation efforts, where airborne systems measured profiles up to 80 km with accuracies of 0.6 m/s at mid-altitudes. These instruments use at 532 nm to derive line-of-sight winds and hydrostatic temperatures, allowing thermal wind estimation in targeted observing scenarios such as studies. Complementarily, GPS dropsondes deployed from research aircraft in campaigns like NASA's and Rapid Intensification Processes () provide high-vertical-resolution (5 m) thermodynamic and profiles from to the surface, particularly in tropical cyclones, supporting thermal wind computations for shear analysis.

Diagnostic Implications in Weather Systems

Thermal wind serves as a key diagnostic tool in forecasting atmospheric circulations by quantifying baroclinicity, where strong vertical in the indicates regions of significant horizontal temperature gradients conducive to development. In operational , this is often assessed through 500–1000 thickness maps, which reflect the mean temperature in that layer and relate directly to the thermal wind magnitude via the ; tighter spacing of thickness contours signals enhanced baroclinicity and potential for vorticity-driven intensification. For instance, decreasing thickness gradients in such maps highlight areas where thermal wind strength promotes rapid growth, aiding forecasters in predicting storm tracks and intensity changes. Deviations from expected thermal wind balance reveal ageostrophic components in weather systems, signaling processes such as frontolysis or intensification. In frontal zones, thermal wind imbalances ahead of cold fronts arise from asymmetric of and potential , leading to mesoscale ageostrophy that can weaken frontal structures during frontolytic phases or enhance low-level jets contributing to . Similarly, in extratropical cyclones, observed thermal winds significantly stronger than model forecasts (e.g., ratios exceeding 1.3) indicate underpredicted synoptic-scale thermal gradients, often resulting in errors in intensity and steering flows, as ageostrophic accelerations disrupt geostrophic balance. In a context, long-term shifts in thermal wind due to Arctic amplification have weakened mid-latitude jet streams, with post-2000 studies linking reduced meridional gradients to decreased vertical and track intensity. This weakening, evident in a 15% decline in synoptic kinetic since 1979, fosters persistent weather patterns that exacerbate mid-latitude extremes like cold outbreaks and heatwaves. Such changes, driven by faster Arctic warming (2–4 times the global rate), diminish the poleward , thereby reducing thermal wind and altering circulation dynamics over decades. A notable is the 1987 Great Storm over , where intense thermal wind from strong baroclinicity—stemming from a sharp temperature contrast between cold polar air and warmer Atlantic air—drove rapid deepening and unexpected track deviations, leading to forecast errors in operational models. This event underscored thermal wind diagnostics in identifying , as the enhanced amplified upper-level divergence and surface winds exceeding 100 km/h, causing widespread damage.

References

  1. [1]
    [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.
  2. [2]
    [PDF] Applications of the Basic Equations Chapter 3
    Warm. Definition: The thermal wind is a vector difference between the geostrophic wind at an upper level and a lower level. Page 7. Thermal Wind. The thermal ...
  3. [3]
    10.11 Why are midlatitude winds mostly westerly (i.e., eastward)?
    The thermal wind is the geostrophic wind vector at pressure level (p1) higher in the atmosphere minus the geostrophic wind vector at a pressure level (p2) lower ...
  4. [4]
    Glossary - NOAA's National Weather Service
    A relatively small-scale, rising air current produced when the Earth's surface is heated. Thermals are a common source of low level turbulence for aircraft.Missing: definition | Show results with:definition
  5. [5]
    Glossary of Climate-Related Terms - Physical Sciences Laboratory
    The thermal wind is caused by horizontal temperature gradients. Thermal Wind Equation. This says that the difference in wind velocity between two heights is ...
  6. [6]
    [PDF] Vertical Wind Profile Information in Severe Storms Forecasting
    wind shear is the thermal wind. The basic definition of the thermal wind is the vertical shear of the geostrophic wind. As anyone who has ever looked at a ...
  7. [7]
    Weather Glossary: T's - NOAA
    Apr 17, 2023 · Areas of low pressure that are shallow in vertical extent and are produced primarily by warm surface temperatures. Thermal Wind. It is a ...
  8. [8]
  9. [9]
    Wind, war and weathermen | MIT News
    Jun 7, 2011 · “Rossby wanted to bring weather forecasting from the 19th century into the 20th century.” Achieving that aim once made Rossby famous: Time ...
  10. [10]
    [PDF] Meteorology for Scientists and Engineers
    Thermal wind magnitude MTH is: MTH = √UTH2 +VTH². 0 west. Figure 11.9. MTH ... mid-latitudes as the tropopause is approached. Above the tropopause, tilt ...
  11. [11]
    [PDF] Air pressure and wind
    The stronger the horizontal temperature gradient, the stronger the thermal wind. The stronger the vertical wind shear. Winds increase more rapidly with height.
  12. [12]
    [PDF] Thermal Wind
    a) Barotropic (the density depends only on the pressure, therefore the isobar surfaces are parallel. The geostrophic winds at different levels are same ).
  13. [13]
  14. [14]
    [PDF] The Hydrostatic Equation
    ∂z = −gρ. This is the hydrostatic equation. The negative sign ensures that the pressure decreases with increasing height. That is, the pressure at height z is ...
  15. [15]
    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 ...
  16. [16]
    [PDF] Fluid Dynamics of the Atmosphere and Oceans
    Scale analysis shows that the hydrostatic approximation is valid on horizontal scales large compared with a typical depth scale for the atmosphere (about 10 km) ...
  17. [17]
    [PDF] 2 The equations governing atmospheric flow. - Staff
    2.5 Summary of the primitive equations. The above equations that we have derived for atmospheric flow are often known as the primitive equations. The ...
  18. [18]
    GENERAL CIRCULATION EXPERIMENTS WITH THE PRIMITIVE ...
    An extended period numerical integration of a baroclinic primitive equation model has been made for the simulation and the study of the dynamics of the ...
  19. [19]
    [PDF] ATM 316 - The “thermal wind” - Recap and isobaric coordinates
    The hydrostatic equation in isobaric coordinates is. dΦ dp. = −. RT p . (13) ... This is the thermal wind equation. It is NOT A WIND. It is a wind shear ...
  20. [20]
    [PDF] Geopotential, Thickness, and Thermal Wind
    (12) THE THERMAL WIND. ● The thermal wind is defined as the vector difference in the geostrophic wind. between two levels of the atmosphere.
  21. [21]
  22. [22]
    11.8: Jet Streams - Geosciences LibreTexts
    Dec 14, 2024 · The thermal-wind relationship tells us that the meridional variation of temperature between the cold poles and warm equator will drive westerly ...
  23. [23]
    [PDF] Jet Streaks
    • Strong vertical wind shear: At least 5-10 m/s per km; At least ½ to 1 order of magnitude greater than synoptic scale shear. • Jet streak: an isotach ...
  24. [24]
    [PDF] Jet Streams of the Atmosphere - Mountain Scholar
    The vertical shear of wind is of the order 5-10 m/sec per km and the lateral shear is of the order 5 m/sec per 100 km. An arbitrary lower limit of 30 m/ sec.
  25. [25]
    [PDF] Chapter 15 - The jet stream and climate change
    The increase is consistent (at least qualitatively) with the ther- mal wind response to the strengthened meridional temperature gradient in these parts of the.
  26. [26]
    [PDF] Fronts and Frontogenesis - University of Wisconsin–Madison
    The thermal wind relationship dictates that fronts, by virtue of their large thermal contrasts, be associated with strong vertical shear of the geostrophic wind ...<|control11|><|separator|>
  27. [27]
    [PDF] Atmospheric Frontogenesis Models
    Four mechanisms for changing horizontal temperature gradients: (i) horizontal deformation, (ii) horizontal shear, (iii) vertical deformation, (iv) differential ...<|control11|><|separator|>
  28. [28]
    [PDF] fronts.pdf
    First, the zone of strong temperature gradient is generally much wider than the zone over which the wind shift occurs. In other words, while the temperature ...
  29. [29]
    [PDF] Hydrostatic Equation and Thermal Wind
    The vectorial difference between the geostrophic wind at a given pressure and the geostrophic wind at a higher pressure. (lower height) is called the THERMAL.
  30. [30]
    [PDF] 1 MET 3502/5561 Synoptic Meteorology Lecture 13: Balanced Wind ...
    wind turning clockwise with height. Warm advection is associated with veering winds.
  31. [31]
    [PDF] QG Omega Equation - NOAA
    Regions of cyclonic vorticity advection by the thermal wind (term A > 0) are located downshear (recall that the thermal wind is the vertical shear vector by ...
  32. [32]
    [PDF] Synoptic Meteorology II: The Q-Vector Form of the Omega Equation
    Mar 5, 2015 · Examining these expressions, we see that both look like temperature advection by the horizontal shear of the geostrophic wind. Alternatively, Q1 ...
  33. [33]
    Conveyor Belts | METEO 3: Introductory Meteorology - Dutton Institute
    Here, the warm conveyor usually encounters high-altitude winds from the west, and, in response, turns eastward. Meanwhile, the western edge of the warm conveyor ...
  34. [34]
    [PDF] Precipitation and Cloud Structure in Midlatitude Cyclones
    Jan 15, 2007 · The cold mode, which can be interpreted as clouds atop the warm conveyor belt, is gradually displaced to colder temperatures east- ward from the ...
  35. [35]
    Utility of Radiosonde Wind Data in Representing Climatological ...
    Mori also found that the thermal wind estimated from the geostrophic winds at 850 and 1000 hPa agrees well with that estimated from the mean horizontal ...
  36. [36]
    New estimates of tropical mean temperature trend profiles from ...
    Apr 6, 2015 · The wind data can be used to estimate zonal mean temperature trends in the tropics using the thermal wind relationship.
  37. [37]
    Collecting Meteorological Data by Radiosonde or Weather Balloon
    As the radiosonde rises at about 1,000 feet/minute (300 meters/minute), sensors on the radiosonde measure profiles of pressure, temperature, and relative ...
  38. [38]
    An improved algorithm for atmospheric wind retrievals from satellite ...
    Mar 28, 2006 · In this study, an improved algorithm to retrieve atmospheric winds from satellite temperature soundings over the polar region is developed.
  39. [39]
  40. [40]
    GRIB - Thermal Wind — metview documentation - Read the Docs
    retrieve( "reanalysis-era5-single-levels", { "product_type": "reanalysis ... compute thermal wind tw = mv.geostrophic_wind(th) # define controur and ...
  41. [41]
    ECMWF Reanalysis v5 (ERA5)
    ERA5 provides hourly estimates of a large number of atmospheric, land and oceanic climate variables. The data cover the Earth on a 31km grid and resolve the ...Wind · Temperature · Precipitation · Atmosphere
  42. [42]
    [PDF] Doppler Rayleigh/Mie/Raman lidar for wind and temperature ... - AMT
    Abstract. A direct detection Doppler lidar for measur- ing wind speed in the middle atmosphere up to 80km with 2h resolution was implemented in the ALOMAR.
  43. [43]
    Airborne lidar observations of wind, water vapor, and aerosol ... - AMT
    Jun 10, 2021 · The study used airborne lidar (DAWN and HALO) to observe wind, water vapor, and aerosol profiles during the Aeolus Cal/Val test flight campaign.
  44. [44]
    The NCAR GPS Dropwindsonde and Its Impact on ... - AMS Journals
    Sep 16, 2023 · ABSTRACT: The global positioning system dropwindsonde has provided thousands of high-resolution kinematic and thermodynamic soundings in and ...
  45. [45]
    [PDF] Thickness and its uses - EUMeTrain
    Jun 13, 2018 · Although the principal thermal wind used in operational meteorology is that through the layer between 1000 hPa and 500 hPa, in fact thermal wind ...
  46. [46]
  47. [47]
    [PDF] using the thermal wind relationship to improve offshore and coastal ...
    A thermal-wind observation that is significantly stronger than the numerical model forecast value can indicate significant errors in the modeled cyclone ...
  48. [48]
    The influence of Arctic amplification on mid-latitude summer circulation
    Aug 20, 2018 · This weakening has been detected in the westerly jet (following the thermal-wind balance), the total kinetic energy of synoptic storm ...
  49. [49]
    Arctic Amplification Weakens the Variability of Daily Temperatures ...
    Arctic amplification (AA) reduces meridional temperature gradients (dT/dy) over the northern mid-high latitudes, which may weaken westerly winds.
  50. [50]
    BBC NEWS | Science/Nature | Lessons learned from Great Storm
    Oct 14, 2007 · "In 1987, there was obviously a very strong temperature gradient." These gradients are caused by the difference in temperature in equatorial ...