Fact-checked by Grok 2 weeks ago

Geopotential height

is a vertical coordinate in that represents the height above mean sea level of a constant-pressure surface, adjusted for variations in Earth's with and altitude. It is calculated as the —defined as the work done per unit mass to lift air against from to that height—divided by a standard value of gravity, g₀ = 9.80665 m s⁻², yielding units of geopotential meters (gpm). This adjustment ensures consistent comparisons across the globe, as geopotential height is approximately 0.16% shorter than geometric height (or about 16 meters lower) at 10 km altitude due to gravity's decrease with elevation. In , geopotential height is essential for mapping upper-air circulation patterns on isobaric surfaces, such as the widely used 500 hPa level, where it delineates ridges (high heights indicating warmer air aloft) and troughs (low heights signaling cooler air and potential storm development). These contours provide insights into the steering of surface systems, including cyclones and anticyclones, by revealing large-scale atmospheric waves and positions. The relates differences in geopotential height between pressure levels to the mean of the air column, enabling forecasters to infer thermal structures and predict phenomena like blocking highs or cold outbreaks. Geopotential height fields are routinely derived from radiosonde balloon measurements, aircraft reports, radiances, and assimilated into numerical models for prediction. Common pressure levels for analysis include 850 (low-level winds and moisture), 500 (mid-tropospheric flow), and 300 (upper-level jets), with typical mid-latitude values ranging from 5,000 to 6,000 gpm at 500 . This metric underpins global reanalysis datasets and studies, where anomalies in geopotential height signal teleconnection patterns like the .

Definition and Basics

Definition

The geopotential, denoted as Φ, represents the energy per unit mass required to raise a unit mass from to a given z in the Earth's atmosphere. It is defined as the of the local g along the vertical path from the surface, expressed as \Phi(z) = \int_0^z g(z') \, dz', where Φ(0) = 0 at . This quantity, with units of m² s⁻² (equivalent to J kg⁻¹), quantifies the work done against without dependence on the specific path taken, assuming a conservative . Geopotential height, denoted Z, is the height coordinate used in isobaric coordinates within , defined as the divided by the standard gravity g₀ = 9.80665 m s⁻², yielding Z = \frac{\Phi(z)}{g_0} = \frac{1}{g_0} \int_0^z g(z') \, dz'. This formulation scales the to units of (meters), providing a measure of the work done to elevate a against to that level. For regions where variations in g are small compared to its mean value, Z approximates the geometric z. Physically, height enables consistent spatial comparisons of atmospheric levels across locations with differing gravitational strengths, as the division by the constant g₀ normalizes the to a uniform reference frame. In isobaric coordinates, where serves as the vertical coordinate, defines the of constant- surfaces, facilitating of atmospheric and . It is routinely employed in meteorological applications, such as contouring levels on weather maps like the 500 chart.

Units and Conventions

In operational , geopotential height is commonly reported in decameters () on constant-pressure charts, where 1 equals 10 meters; for instance, a 500 surface height of 540 indicates 5400 meters above mean . This convention facilitates concise visualization of large-scale atmospheric features, with contours typically drawn at intervals of 4 to 8 . The base unit is the geopotential meter (gpm), defined through the relation to Φ as Z = Φ / g₀, where g₀ = 9.80665 m/s² serves as the fixed reference for global uniformity, supplanting local gravitational variations. Under standard conditions near , 1 gpm approximates 1 geometric meter, though the distinction grows slightly with altitude due to gravity's decrease. This , outlined in guidelines, ensures consistent computations in upper-air observations and numerical models. Historically, the 1960s marked a shift from imperial units like feet to metric meters in meteorological practice, aligning with international metrication; for example, the UK Meteorological Office adopted metric heights effective January 1, 1967. In aviation-related reporting, geopotential heights tied to pressure levels are often denoted in feet and rounded to the nearest 10 or 60 feet to reflect observational precision and code requirements.

Theoretical Foundations

Relation to Geopotential

The Φ at height z represents the per unit required to raise a parcel of air from to that height, defined mathematically as \Phi(z) = \int_0^z g(z') \, dz', where g(z') is the at height z', and Φ(0) = 0 at . The Z is derived from the by dividing by a standard reference g₀ (typically 9.80665 m s⁻², the globally averaged value at ), yielding Z = Φ / g₀. This ensures that Z has units of (meters) and simplifies the form of the hydrostatic equation in atmospheric dynamics, as gradients in Z correspond directly to considerations without needing to account for variations in g. For computational approximations on Earth, where the radius R ≈ 6371 km, the inverse-square law for gravity gives g(z) ≈ g₀ [R / (R + z)]², leading to the exact relation Z = [R z / (R + z)]. For altitudes z ≪ R (typical in the troposphere and stratosphere), this expands to Z ≈ z [1 - (z / R) + (z / R)² - ⋯], meaning geopotential height is slightly less than geometric height; precise calculations often incorporate centrifugal effects from Earth's rotation, which reduce effective gravity by up to 0.3% at the equator and are averaged into g₀ for standard definitions. This relation ties directly into the hydrostatic balance through the differential form dΦ = g dz = -α dp, where α is the specific volume (1/ρ, with ρ density) and dp is the pressure change; integrating this links geopotential differences to pressure levels, facilitating use in pressure-coordinate systems for atmospheric modeling.

Difference from Geometric Height

Geometric height, denoted as z, is the direct physical distance measured radially from the Earth's surface to a point in the atmosphere, commonly determined using global positioning system (GPS) receivers or radar altimetry. The primary distinction between geopotential height Z and geometric height z arises from the variation in g, which decreases with increasing altitude due to the greater distance from Earth's . This effect means that Z < z, with the relative difference growing to approximately 0.5% at 30 km altitude. For instance, at the equator, a geometric height of 10 km corresponds to a geopotential height of about 9.98 km. Gravitational acceleration also exhibits latitude dependence, achieving a maximum value of 9.832 m/s² at the poles and a minimum of 9.780 m/s² at the equator due to Earth's oblateness and rotational centrifugal force. Consequently, for a fixed geometric height, geopotential height varies by roughly 0.5% between equatorial and polar regions, reflecting the integrated effect of local gravity in the geopotential computation. Employing geopotential height in place of geometric height minimizes errors in computations involving gravitational potential energy, such as those in atmospheric hydrostatic balance, by incorporating variable gravity; this yields an accuracy improvement of about 0.5% over approximations assuming constant gravity with geometric height.

Physical Role

In Atmospheric Dynamics

In atmospheric dynamics, geopotential height plays a central role in describing balanced flows on synoptic scales, where the geostrophic wind approximation dominates. The geostrophic wind \mathbf{v}_g is given by \mathbf{v}_g = \frac{g_0}{f} \mathbf{k} \times \nabla_p Z, where g_0 = 9.80665 m s⁻² is the standard gravity, f is the , \mathbf{k} is the unit vector in the vertical direction, and \nabla_p Z is the horizontal gradient of geopotential height Z on a constant pressure surface. This relation indicates that the geostrophic wind flows parallel to contours of constant Z, approximating streamlines, with wind speed inversely proportional to the spacing between contours—tighter spacing yields stronger winds. On pressure surfaces, this balance eliminates density variations, making Z gradients a direct measure of the pressure gradient force driving the flow. The vertical variation of the geostrophic wind is linked to horizontal temperature gradients through the thermal wind equation, \frac{\partial \mathbf{v}_g}{\partial \ln p} = \frac{R}{f} \mathbf{k} \times \nabla_p T, where R is the gas constant for dry air and \nabla_p T is the temperature gradient on a pressure surface. This equation connects gradients in Z between pressure levels to thermal structure, as the geopotential thickness Z_2 - Z_1 between two levels is proportional to the mean layer temperature via the hypsometric relation. Consequently, a northward increase in temperature (warm advection) results in veering winds with height, strengthening the westerly jet in midlatitudes. In synoptic-scale flows, patterns in Z reveal dynamic processes such as ridges and troughs associated with . High Z ridges at mid-tropospheric levels like 500 hPa signify warm air subsidence and thicker atmospheric layers, promoting stable, anticyclonic conditions, while low Z troughs indicate cold air ascent and thinner layers, fostering cyclonic development. Anomalies in 500 hPa Z gradients, for instance, trace the propagation and decay of synoptic-scale , where eastward-moving troughs and ridges can lead to blocking patterns if gradients weaken sufficiently (e.g., meridional gradients below -10 m per degree latitude). These features modulate large-scale circulation, with positive Z anomalies often amplifying wave amplitudes in the Northern Hemisphere winter. Under hydrostatic balance, the horizontal gradient of Z on a pressure surface directly drives geostrophic flow, as the pressure gradient force in isobaric coordinates is represented by -g_0 \nabla_p Z, integrating density and gravity variations to ensure balanced motion without vertical accelerations on large scales.

In Planetary Fluids

In planetary fluid dynamics, the concept of geopotential height generalizes beyond Earth's atmosphere to any gravitational field, defined as Z = \frac{1}{g_{\text{ref}}} \int g \, ds along equipotential surfaces, where g_{\text{ref}} is a reference gravitational acceleration and ds is the path element. This formulation accounts for variations in gravity due to planetary shape, rotation, or distance from the center of mass, ensuring Z represents potential energy per unit mass normalized to a consistent scale. In adiabatic, frictionless flows, geopotential height is conserved following fluid parcels, as dictated by , which equates the sum of kinetic energy, pressure work, and gravitational potential along streamlines. In oceanic applications, geopotential height manifests as dynamic height in isopycnal coordinates, where surfaces of constant density serve as natural coordinates for analyzing geostrophic flows. Dynamic height D at pressure p relative to a reference pressure p_0 is given by D = \frac{1}{g} \int_{p_0}^{p} \frac{dp}{\rho}, integrating the hydrostatic balance to capture density-driven deviations from a level surface. A key example is steric height, which quantifies sea surface height anomalies due to thermal and saline expansions; it is computed as h = \frac{1}{\rho g} \int \Delta \rho \, dz, or equivalently as the pressure integral of specific volume anomalies divided by a reference gravity (9.7963 m/s²), providing insights into circulation patterns like gyres without direct altimetry. For planetary atmospheres, geopotential height adjusts for spatially varying gravity, such as on oblate bodies like , where the planet's equatorial bulge causes g to decrease poleward, necessitating integration along equipotentials to map atmospheric layers accurately. On , this adjustment reveals deep zonal winds penetrating below the visible clouds, with geopotential surfaces aligning flows to the planet's non-spherical gravity field. Similarly, in Mars' thin atmosphere, geopotential height is referenced to the MOLA-derived areoid—an equipotential surface—to model density profiles and track phenomena like dust storms, where altimetry data from the Mars Orbiter Laser Altimeter informs vertical structure amid variable topography and dust loading. A conserved quantity leveraging geopotential height in these fluids is , defined as q = \frac{ (\nabla \times \mathbf{v} + \mathbf{f}) \cdot \nabla \theta }{\rho}, where \mathbf{v} is velocity, \mathbf{f} is the Coriolis parameter vector, \theta is potential temperature, and \rho is density. This scalar is materially conserved (Dq/Dt = 0) in adiabatic, frictionless flows, particularly on geopotential height surfaces that approximate isentropic levels in barotropic conditions, enabling diagnosis of wave propagation and stability in planetary oceans and atmospheres.

Practical Applications

In Weather Analysis

Geopotential height fields are routinely visualized on constant-pressure weather charts at key levels such as 1000 hPa (near the surface), 500 hPa (mid-troposphere), and 300 hPa (upper troposphere), with contours typically drawn at 60-meter intervals to depict the large-scale flow of the atmosphere. These charts reveal the undulating patterns of the and , allowing meteorologists to assess the positioning of high- and low-pressure systems aloft. For example, at 500 hPa, contours below 540 dam often indicate cooler air masses conducive to storm development, while those above 570 dam suggest warmer, subsiding conditions. Thickness maps, calculated as the difference between geopotential heights at two pressure levels (e.g., Z_{500} - Z_{1000}), provide insights into air mass characteristics and stability, as thickness is directly proportional to the mean virtual temperature between those levels. Contours on 1000–500 hPa thickness charts, often spaced at 60 gpm intervals, delineate warm versus cold air masses; values exceeding 5520 gpm typically signify warm, stable air prone to clear skies and heat, whereas lower values below 5220 gpm indicate colder, potentially unstable air associated with precipitation. These maps are essential for forecasting precipitation types, such as rain versus snow, based on the 5400 gpm isotherm. In diagnostic weather analysis, geopotential height patterns at 500 hPa are interpreted to predict surface weather features, where troughs (negative curvature in contours) signal cyclogenesis and low-pressure development, and ridges (positive curvature) indicate anticyclones and blocking. A notable example is the 2003 European heatwave, driven by a persistent 500 hPa ridge with geopotential heights anomalously exceeding 150 meters above climatology, leading to prolonged subsidence and extreme surface temperatures. Anomaly fields, computed as departures from long-term climatologies, further aid diagnosis; positive anomalies of +100 meters or greater at 500 hPa often denote blocking highs that stall weather systems, fostering droughts or heatwaves. Teleconnection patterns, such as the , are quantified using these height anomalies, with the positive NAO phase featuring a deepened Icelandic low and strengthened Azores high in 500 hPa contours, influencing transatlantic weather variability. Modern weather analysis integrates geopotential heights derived from satellite-based GPS radio occultation (RO), which measures atmospheric bending of GNSS signals to retrieve high-vertical-resolution profiles of refractivity, from which heights of constant-pressure surfaces are computed. These RO-derived heights offer global, all-weather coverage with accuracy better than 10 meters in the upper troposphere, enabling real-time assimilation into analysis charts for improved depiction of troughs, ridges, and anomalies in data-sparse regions like the oceans.

In Numerical Modeling

In numerical weather prediction and climate models, hybrid sigma-pressure coordinate systems are widely employed, with geopotential height (Z) playing a central role in vertical interpolation across model levels. These coordinates blend pressure-based levels in the upper atmosphere with terrain-following sigma levels near the surface, allowing Z to facilitate accurate representation of atmospheric stratification and mass distribution. For instance, in the ECMWF Integrated Forecasting System (IFS), Z is computed through hydrostatic integration from the model topography upward, using temperature and specific humidity profiles on model levels to derive geopotential values that ensure consistency with the governing equations. This approach enables efficient handling of complex orography while maintaining computational stability in global simulations. Data assimilation in these models, particularly through four-dimensional variational (4D-Var) methods, incorporates observations of geopotential height to refine initial conditions and improve forecast accuracy. In 4D-Var frameworks, such as those in the ECMWF IFS and NOAA's Gridpoint Statistical Interpolation system, Z fields are incrementally adjusted to align with direct measurements from radiosondes, which provide vertical profiles of pressure and temperature from which Z is derived. Error covariance matrices in these systems explicitly include gradients of Z to impose geostrophic and hydrostatic balance constraints, reducing analysis errors in mid-tropospheric flows and enhancing the dynamical consistency of the assimilated state. This assimilation of Z observations has been shown to significantly lower forecast errors in 500 hPa geopotential height, with operational implementations demonstrating improvements of up to 10-20% in medium-range predictions relative to three-dimensional variational methods. Model outputs frequently feature ensemble forecasts of 500 hPa for probabilistic predictions of synoptic-scale patterns, such as mid-latitude cyclones and blocking highs. In the NOAA Global Forecast System (GFS), for example, these Z fields are generated at a horizontal resolution of approximately 13 km (~0.125° cubed-sphere grid) for short-range forecasts, following the 2019 upgrade to the that enhanced spatial detail for global forecasting up to 16 days ahead (with coarser resolution of ~34 km for days 10-16). Ensemble members perturb initial Z conditions and model physics to quantify uncertainty, enabling derived products like probability maps of height anomalies that inform operational decision-making in weather services. Despite these advances, challenges persist in incorporating geopotential height into numerical models, particularly regarding numerical stability and long-term biases. Z-coordinate models, which treat Z as the primary vertical coordinate, can encounter instability over steep terrain due to errors in pressure gradient computations, prompting a preference for hybrid terrain-following coordinates that mitigate truncation errors but introduce their own computational complexities. In climate projections, such as those from the Coupled Model Intercomparison Project Phase 6 (CMIP6), systematic biases in simulated Z fields—often manifesting as overestimations in extratropical heights—necessitate post-processing corrections to align model outputs with reanalysis datasets like ERA5, ensuring reliable assessments of future atmospheric circulation changes.

Historical Development

Origin of the Concept

The concept of geopotential height traces its origins to foundational ideas in 19th-century potential theory, which provided the mathematical framework for understanding gravitational influences in geophysical contexts. Pierre-Simon Laplace laid early groundwork in his 1805 derivation of the hypsometric formula within celestial mechanics, linking pressure variations to gravitational potential differences in fluid layers, as part of his broader exploration of planetary atmospheres in Mécanique Céleste. Carl Friedrich Gauss advanced this in the 1830s by formalizing the scalar potential function for gravitational fields, emphasizing its role in describing equipotential surfaces where the net force is perpendicular to the surface, a principle essential for later atmospheric applications. Alfred Clebsch extended these ideas in the 1850s through variational principles in fluid dynamics, introducing methods to minimize energy functionals for incompressible fluids under gravity, which influenced the treatment of atmospheric layers as potential fields. In meteorology, the concept gained practical traction in the early 20th century as dynamic analysis shifted toward three-dimensional atmospheric structure. Felix Maria Exner contributed to dynamic meteorology in his 1925 treatise Dynamische Meteorologie, where geopotential surfaces were used to map pressure contours in a gravity-adjusted coordinate system to better represent atmospheric flow. This approach addressed limitations in geometric height measurements by accounting for varying gravity, providing a more accurate basis for analyzing upper-air dynamics. In the 1930s, Carl-Gustaf Rossby built on this by integrating with isobaric coordinates, promoting constant-pressure surfaces as a standard framework for charting large-scale circulations, such as planetary waves, in works like his 1939 analysis of mid-latitude motions. Early computations of geopotential height relied on manual hypsometric techniques using barometric pressure readings, predating automated observations in the 1940s. Meteorologists applied the hypsometric equation—derived from hydrostatic balance—to estimate the vertical distance between pressure levels from temperature and pressure data collected via mercury barometers during balloon ascents or station networks, yielding approximate geopotential thicknesses for weather map construction. These labor-intensive methods were central to pre-war upper-air analysis, despite challenges from instrument errors and sparse data. The formal standardization of geopotential height notation (Z) occurred in the mid-1950s under the World Meteorological Organization, through Technical Regulations adopted around 1956, driven by post-World War II demands for reliable aviation altimetry across international routes, where inconsistent height references had posed safety risks during global military operations. The WMO's Technical Regulations adopted geopotential meters as the unit, ensuring uniformity in radiosonde reports and forecast charts to support high-altitude flight planning.

Evolution in Meteorology

In the post-World War II era, the development and global deployment of radiosondes during the 1940s and 1950s marked a pivotal advancement in measuring upper-air atmospheric profiles, allowing meteorologists to construct detailed charts essential for synoptic analysis. These balloon-borne instruments provided pressure, temperature, and humidity data up to the stratosphere, enabling the calculation of geopotential heights on standard pressure levels such as 500 hPa, which became standard for depicting mid-tropospheric circulation patterns. By the mid-1950s, international networks had expanded, with over 4,000 stations contributing data that supported the transition from surface-based to three-dimensional weather forecasting. The advent of computing in the 1970s integrated geopotential height calculations into numerical weather prediction (NWP) models, automating the hydrostatic integration of temperature profiles to derive height fields. NASA's Goddard Laboratory for Atmospheric Sciences (GLAS) model, operational in the late 1970s, exemplified this shift by producing global 500 hPa geopotential height forecasts from assimilated radiosonde data, improving prediction skill for large-scale features like jet streams. By the 1990s, the incorporation of Global Positioning System (GPS) technology into radiosondes enhanced height accuracy, reducing geopotential errors in the troposphere to approximately 10-20 meters through precise positioning and pressure derivations, a significant improvement over earlier barometric methods. Satellite advancements from the 2000s onward further expanded geopotential height derivation by inverting infrared radiance measurements to retrieve vertical profiles. Geostationary satellites like GOES and Meteosat Second Generation (MSG), launched in 2002, used water vapor and infrared channels to assimilate radiances into models, yielding geopotential height estimates with resolutions down to 30-50 km, enhancing nowcasting of upper-level dynamics. In the 2020s, artificial intelligence integrations at the European Centre for Medium-Range Weather Forecasts (ECMWF) have improved anomaly detection in geopotential height fields; for instance, the AI-based Forecasting System (AIFS) and GraphCast models achieve competitive anomaly correlation coefficients, often matching or exceeding traditional IFS scores up to medium-range leads (e.g., >0.9 at 5 days), with improvements in capturing patterns. Recent developments have extended geopotential height concepts beyond , incorporating them into circulation models for atmospheres analyzed via (JWST) data since 2022, where scale heights and gradients inform radiative-convective equilibria in habitable-zone worlds. In modeling, the Intergovernmental Panel on Climate Change's Sixth Assessment Report (AR6, 2021) refined projections of geopotential height changes under warming scenarios, using Coupled Model Intercomparison Project Phase 6 (CMIP6) simulations to quantify effects on 500 hPa heights, with projected decreases of 100-200 meters in the by 2100 under high-emission pathways.

References

  1. [1]
    NOAA's National Weather Service - Glossary
    ### Summary of Geopotential Height
  2. [2]
    Geopotential height - Parameter Database
    Geopotential height measures a point's height in the atmosphere relative to its potential energy, calculated by dividing geopotential by Earth's gravity. It's ...
  3. [3]
    [PDF] The Hydrostatic Equation
    We define the geopotential height Z as. Z = Φ(z) g0. = 1 g0. Z z. 0 g dz where ... distribution of geopotential height of any other pressure sur- face. 15 ...
  4. [4]
    TEMPERATURE GRADIENTS...HEIGHTS AND THICKNESSES
    Geopotential height is valuable for locating troughs and ridges which are the upper level counterparts of surface cyclones and anticyclones.
  5. [5]
    ERA5 monthly averaged data on pressure levels from 1940 to present
    Apr 18, 2019 · The geopotential height can be calculated by dividing the geopotential by the Earth's gravitational acceleration, g (=9.80665 m s-2). The ...
  6. [6]
    [PDF] ESCI 341 - Meteorology
    THICKNESS AND THE HYPSOMETRIC EQUATION. ○ In terms of geopotential height the hydrostatic equation is d dp. dT d. R. T dz dz dz ρ ρ. ⌈. ⌉. = +. │. │. ⌊. ⌋ dp g.
  7. [7]
    [PDF] Pressure as Vertical Coordinate - UCI ESS
    Jan 24, 2017 · Thus in the isobaric coordinate system the horizontal pressure gradient force is measured by the gradient of geopotential at constant pressure.<|separator|>
  8. [8]
    500 hPa geopotential height and 850 hPa temperature
    These show approximately how far one has to go up in the atmosphere before the pressure falls to 500 hPa. On average this level is around 5.5 km above sea level ...
  9. [9]
    500-hPa Heights - Climate Prediction Center - Outlooks - NOAA
    Northern Hemisphere mean and anomalous 500-hPa geopotential height (CDAS/Reanalysis). Mean heights are denoted by solid contours drawn at an interval of 8 dam.
  10. [10]
    [PDF] Guide to Meteorological Instruments and Methods of Observation
    One of the purposes of the World Meteorological. Organization (WMO) is to coordinate the activities of its 188 Members in the generation of data and.
  11. [11]
    Weather - Metrication UK
    From 1 January 1967, the Met Office switched to the metric system for all references to heights and distances. This included the measurement of horizontal ...
  12. [12]
    [PDF] International Codes, Volume I.1, Annex II to the WMO Technical ...
    Aug 15, 2020 · Definitions and explanations relating to these specifications are printed in light-face roman. WMO-No. 306. © World Meteorological Organization, ...
  13. [13]
    [PDF] Introduction to Atmospheric Dynamics Chapter 1
    The notion of geopotential height is used when height-based vertical coordinates are not. Observe: Geopotential and geopotential height vary along p surfaces. ...Missing: isobaric | Show results with:isobaric
  14. [14]
    [PDF] 2 The equations governing atmospheric flow. - Staff
    The geopotential height (Z) is related to the geopotential via Z = Φ/g and so. Eq. 45 gives an expression for the geopotential height of a particular pressure ...
  15. [15]
    [PDF] Geopotential & Geometric Altitude The Perfect Gas Law
    Sep 27, 2020 · H stands for geopotential altitude and Z stands for geometric altitude, g is the acceleration of gravity and G is the value of g at sea level.<|separator|>
  16. [16]
    geopotential_to_height — MetPy 1.7
    Calculates the height above mean sea level from geopotential using the following formula, which is derived from the definition of geopotential as given in
  17. [17]
    [PDF] NASA TR R-459 DEFINING CONSTANTS, EQUATIONS, AND ...
    In the 1975 Standard, geopotential altitude is used explicitly only at heights below 86 geometric kilometers. The transformation from Z to H (equation. 16) in ...
  18. [18]
    1.7: Atmospheric Structure - Geosciences LibreTexts
    Dec 14, 2024 · Geopotential height H is defined as the work per unit mass to lift an object against the pull of gravity, divided by the gravitational ...
  19. [19]
    Acceleration Due to Gravity - Roseke Engineering
    Dec 24, 2013 · North and South Poles: g = 9.832 m/s2 = 32.26 ft/s2. Equator: g = 9.780 m/s2 = 32.09 ft/s2. Standard Gravity. The term standard gravity ...
  20. [20]
    [PDF] The Geostrophic Wind
    As seen in the figure below, the geostrophic wind blows parallel to the geopotential height contours on a constant pressure surface (most weather maps) with ...
  21. [21]
    [PDF] Lecture 3: Applications of Basic Equations - UCI ESS
    The geopotential height is approximately equal to the actual height. However, for dynamic calculations involving the wind the geopotential height must be used ...Missing: derivation | Show results with:derivation
  22. [22]
    [PDF] Geopotential, Thickness, and Thermal Wind
    Geopotential is related to gravity variation. Thickness is the difference in geopotential height. Thermal wind is the difference in geostrophic wind between ...Missing: derivation | Show results with:derivation
  23. [23]
    Representation of Synoptic‐Scale Rossby Wave Packets and ...
    Jan 9, 2019 · Geopotential height gradients at 500 hPa are evaluated for all grid ... scale flow pattern and synoptic-dynamic characteristics.
  24. [24]
    Introduction to 500 mb height maps - ATMO336 - Spring 2016
    The height above sea level where the air pressure falls 500 mb is measured at many locations around the globe by sending instrumented weather balloons upward.
  25. [25]
    Constant Pressure Charts: Thickness - NOAA
    Apr 5, 2023 · While typically plotted on surface charts, the thickness contours are derived from constant pressure charts. ... The contours are drawn in 60- ...Missing: geopotential feet
  26. [26]
    How exceptional was the early August 2003 heatwave in France?
    May 18, 2005 · The composites of the 850 hPa temperature and 500 hPa geopotential height for the 1–15 August period show the usual pattern of summer ...
  27. [27]
    The Role Played by Blocking Systems over Europe in Abnormal ...
    Apr 21, 2013 · There exist positive anomalies more than +100 m in geopotential height over Europe through that season. This block episode oscillates over ...
  28. [28]
    Measurement of geopotential heights by GPS radio occultation - Leroy
    Mar 1, 1997 · Geopotential heights of constant pressure surfaces are retrieved from global positioning system (GPS) radio occultation data.Missing: derived | Show results with:derived
  29. [29]
    Deriving dynamics from GPS radio occultation: Three‐dimensional ...
    Sep 18, 2014 · Here we use RO data for the first time to derive climatological wind fields from sampling error-corrected geopotential height fields on isobaric surfaces.<|control11|><|separator|>
  30. [30]
    Hybrid σ–p Coordinate Choices for a Global Model in - AMS Journals
    A methodology for choosing a hybrid σ–p (sigma–pressure) vertical coordinate of the Simmons–Strüfing form for a global model is presented.
  31. [31]
    ERA5: compute pressure and geopotential on model levels, geopotential height and geometric height - Copernicus Knowledge Base - ECMWF Confluence Wiki
    ### Summary: Computing Geopotential Height in ECMWF Models with Hydrostatic Integration from Topography
  32. [32]
    [PDF] PART III: DYNAMICS AND NUMERICAL PROCEDURES - ECMWF
    Jun 27, 2023 · The vertical integral in the hydrostatic equation (i.e. from the surface upwards) can be constructed by taking the difference of the integral ...
  33. [33]
    [PDF] ECMWF's 4D-Var data assimilation system – the genesis and ten ...
    Figure 1 shows the impact on the 500 hPa geopotential height of going from 3D-Var to 4D-Var with everything else ... • Operational assimilation of GPS radio.Missing: radiosondes | Show results with:radiosondes
  34. [34]
    Comparison of Hybrid Ensemble/4DVar and 4DVar within the ...
    The globally averaged analysis and forecast verification versus radiosonde observations are presented in Figs. 7 and 8 for geopotential height (left-hand column) ...4. Results · A. Rms Vector Wind Error For... · C. Geopotential Height...<|separator|>
  35. [35]
    Documentation - GFS - Virtual Lab - NOAA VLab
    GFS started to distribute higher resolution 0.25 degree data to users with its 2015 upgrade. GFS began distributing hourly 0.25 degree data up to F120 in May ...
  36. [36]
    Extending the Numerical Stability Limit of Terrain-Following ...
    Nov 1, 2012 · The truly horizontal pressure-gradient discretization improves numerical stability over steep slopes for a wide range of horizontal mesh sizes.
  37. [37]
    Bias-corrected CMIP6 global dataset for dynamical downscaling of ...
    Nov 4, 2021 · This dataset will be useful for dynamical downscaling projections of the Earth's future climate, atmospheric environment, hydrology, agriculture, wind power, ...
  38. [38]
    Performance evaluation of CMIP6 models for application to ...
    Oct 1, 2024 · The capacity of the climate models to simulate geopotential height (ghp) and specific humidity (hus) were evaluated at 200, 500, and 850 hPa ...
  39. [39]
    METHODS AND RESULTS OF DEFINITIVE AIR ... - AMS Journals
    The final step in bhe development was tlie hypsometric, formula derived by Laplace in 1805. Lap- lace proceeded from the basic equation: in which p ...
  40. [40]
    Mountain Torques and the Equatorial Components of Global ...
    Exner, F., 1925: Dynamische Meteorologie. Springer, 421 pp. Frederiksen, J., 1982: Eastward and westward flows over the topography in nonlinear and linear ...
  41. [41]
    The Comprehensive Historical Upper-Air Network in - AMS Journals
    Current global upper-level datasets only reach back to the 1940s or 1950s and do not cover some important periods in the first half of the twentieth century.
  42. [42]
    Observation errors in early historical upper‐air observations - 2013
    Oct 18, 2013 · The data set covers air temperature, geopotential height, and wind observations from 4182 stations in operation from close to the beginning of ...Missing: charts | Show results with:charts
  43. [43]
    [PDF] notice - NASA Technical Reports Server
    Space-Time Spectra of the GLAS Model 500 mb Geopotential Height ••. D. M. ... Numerical weather prediction (NWP) is an initial-value problem for a ...
  44. [44]
    (PDF) Global Criteria for Tracing the Improvements of Radiosondes ...
    The GPS technology allowed an improvement of an order of magnitude in the quality of radiosonde geopotential heights at 30 km altitude. Large improvements ...
  45. [45]
    [PDF] Assimilation of Geostationary WV Radiances from GOES ... - ECMWF
    Abstract. This report describes the current status of geostationary water vapour radiance assimilation in the ECMWF. 4DVAR system.
  46. [46]
    AIFS: a new ECMWF forecasting system
    We show the northern hemisphere Anomaly Correlation Coefficient (ACC) for geopotential height at 500 hPa of IFS forecasts (red, dashed) and AIFS forecasts (blue) ...Missing: detection 2020s
  47. [47]
  48. [48]
    Venus as an Exoplanet: I. An Initial Exploration of the 3‐D Energy ...
    Oct 25, 2025 · Atmospheric Inflation: Dayside geopotential heights bulge to 760 km (4× NIR) versus nightside (∼250 km), a 3:1 asymmetry driven by radiative ...
  49. [49]
    Chapter 4 | Climate Change 2021: The Physical Science Basis
    This chapter assesses simulations of future global climate change, spanning time horizons from the near term (2021–2040), mid-term (2041–2060), and long term ( ...Missing: geopotential | Show results with:geopotential