Fact-checked by Grok 2 weeks ago

Equivalent potential temperature

Equivalent potential temperature, denoted as θe, is a thermodynamic quantity in that represents the temperature an air parcel would reach if lifted to very low pressure to condense all its moisture and release the associated , then lowered dry-adiabatically to a standard reference pressure of 1000 hPa. This measure combines the effects of the parcel's , from , and changes, providing a conserved property for moist air parcels. It is particularly useful for analyzing the total heat content in humid environments, where dry potential alone would underestimate the energy available. The calculation of θe involves a pseudo-adiabatic process: the air parcel is first raised moist-adiabatically from its initial level (such as the surface) to very low pressure (approaching 0 hPa) until all condenses, and then descended dry-adiabatically to 1000 hPa using adjusted for the released . Approximations for this computation often neglect the heat capacities of and liquid water relative to dry air to simplify the formula. Values of θe increase with higher and content, as more is released from wetter parcels. Due to its approximate conservation during both dry adiabatic and saturated adiabatic processes, θe serves as an effective tracer for air mass identification and movement in the atmosphere. In operational , it is plotted on Skew-T log-P diagrams to evaluate (CAPE) and identify unstable layers, where steep vertical gradients in θe indicate potential for like thunderstorms. Horizontal maps of θe reveal ridges of high values associated with warm, moist air , signaling regions prone to mesoscale convective systems. Beyond , θe is applied in climate research to assess long-term trends in surface and moisture, offering a more comprehensive metric for impacts than temperature alone.

Thermodynamic Background

Stability in Incompressible Fluids

In incompressible fluids, such as those approximated in oceanographic models, is maintained when the vertical balances the gravitational force, given by the equation \frac{dP}{dz} = -\rho g, where P is , \rho is , g is , and z is the vertical coordinate increasing upward. arises from density differences between a and its surroundings; a parcel denser than the ambient sinks, while a less dense parcel rises, according to , with the buoyant force per unit volume equal to -\rho g \delta \rho / \rho, where \delta \rho is the density . Parcel theory provides a framework for assessing hydrostatic by considering the vertical of a small in a stratified , assuming no mixing or heat exchange during the motion. If displaced upward by a small \delta z from its , the parcel retains its original \rho_p = \rho(z_0), while the environmental at the new is \rho_e(z_0 + \delta z) \approx \rho(z_0) + \frac{d\rho}{dz} \delta z. The resulting acceleration is a = -g \frac{\rho_p - \rho_e}{\rho} \approx g \frac{1}{\rho} \frac{d\rho}{dz} \delta z. The parcel returns to if \frac{d\rho}{dz} < 0 (density decreasing upward, stable stratification), oscillates with neutral stability if \frac{d\rho}{dz} = 0, or accelerates away if \frac{d\rho}{dz} > 0 (unstable). The mathematical criterion for stability is encapsulated in the Brunt-Väisälä frequency N, derived from the equation of motion for the displaced parcel: \frac{d^2 \delta z}{dt^2} = g \frac{1}{\rho} \frac{d\rho}{dz} \delta z. This yields the oscillatory form \frac{d^2 \delta z}{dt^2} + N^2 \delta z = 0, where N^2 = -\frac{g}{\rho} \frac{d\rho}{dz}. Stability requires N^2 > 0, corresponding to \frac{d\rho}{dz} < 0, with the parcel oscillating at frequency N; if N^2 < 0, exponential growth indicates instability and potential convective overturning. In the ocean mixed layer, an example of incompressible fluid behavior occurs where turbulence from wind and waves homogenizes density, resulting in \frac{d\rho}{dz} \approx 0 and N^2 \approx 0, leading to neutral stability that allows vertical mixing without restoring forces. This contrasts with the underlying pycnocline, where salinity or temperature gradients produce \frac{d\rho}{dz} < 0 and positive N^2, inhibiting mixing. Such models lay the groundwork for extending stability analysis to compressible fluids like the atmosphere.

Potential Temperature in Dry Air

Potential temperature, denoted as θ, is the temperature a parcel of dry air would attain upon being adiabatically brought to a standard reference pressure of 1000 hPa, serving as a conserved thermodynamic property during dry adiabatic processes. This quantity builds on stability concepts from incompressible fluids by incorporating compressibility effects in the atmosphere, enabling consistent evaluation of buoyancy without pressure influences. The mathematical definition is given by the equation: \theta = T \left( \frac{P_0}{P} \right)^{R / c_p} where T is the parcel's current temperature in Kelvin, P is its current pressure in hPa, P_0 = 1000 hPa is the reference pressure, R = 287 J kg⁻¹ K⁻¹ is the gas constant for dry air, and c_p = 1004 J kg⁻¹ K⁻¹ is the specific heat capacity at constant pressure for dry air. This formula derives from Poisson's equation for an ideal gas undergoing adiabatic compression or expansion, where the first law of thermodynamics implies no heat exchange (dq = 0), leading to the relation T P^{-\kappa} = constant with \kappa = R / c_p \approx 0.286. Integrating this with the ideal gas law p = \rho R T yields the potential temperature as the invariant that normalizes temperature to the reference pressure. By removing pressure dependencies, potential temperature facilitates direct inter-parcel comparisons at equivalent levels, revealing buoyancy differences solely due to thermal structure. In a typical troposphere, θ increases with height, signifying stable stratification for dry processes. Atmospheric stability for dry air is assessed by comparing the environmental lapse rate \gamma (temperature decrease with height) to the dry adiabatic lapse rate \Gamma_d = g / c_p \approx 9.8 K/km, where g = 9.8 m/s² is gravitational acceleration. The atmosphere is stable if \gamma < \Gamma_d, neutral if \gamma = \Gamma_d, and unstable if \gamma > \Gamma_d, as determined by whether displaced parcels return to or diverge from their origins while conserving θ. For instance, a air parcel with initial θ = 290 lifted from the surface to 1 km altitude cools at \Gamma_d, reaching approximately 280.2 , but recompressing it adiabatically to 1000 restores it to 290 . If the environmental θ at 1 km exceeds 290 , the parcel is denser than surroundings upon displacement and sinks, confirming ; conversely, a lower environmental θ indicates .

Role of Moisture and Latent Heat

In unsaturated air, the potential temperature serves as a conserved quantity during adiabatic processes, providing a baseline for assessing stability without moisture effects. The presence of water vapor introduces significant modifications to adiabatic lapse rates through the release of latent heat during phase changes, particularly condensation. In dry air, the adiabatic lapse rate (Γ_d) is approximately 9.8 K/km, reflecting pure expansion cooling. However, when air becomes saturated and ascends, condensation occurs, releasing latent heat that offsets some of the cooling, resulting in a moist adiabatic lapse rate (Γ_m) that is substantially lower, typically around 4-6 K/km depending on temperature and pressure. This reduced lapse rate makes saturated air more stable compared to unsaturated air under the same environmental conditions, as the parcel cools less rapidly relative to its surroundings. Moisture content in the atmosphere is quantified using parameters such as the saturation mixing ratio (r_s), which represents the maximum mass of per unit mass of dry air at a given and before saturation occurs. Another key measure is the (T_w), defined as the lowest achievable by evaporating water into the air parcel at constant , which integrates both and humidity to indicate the parcel's moisture potential. These metrics highlight how water vapor loading influences the energy budget, setting the stage for effects during vertical motion. The pseudo-adiabatic process approximates real-world moist ascent by assuming that forms and is instantaneously removed from the parcel, such as through , while the released (L * dr, where L is the of and dr is the change in mixing ratio due to ) is fully added to the parcel's . This simplifies calculations by neglecting the of the liquid water, making it distinct from reversible moist adiabatic processes where remains in the parcel. The heat addition from this process further reduces the effective cooling rate, enhancing the parcel's potential in convective scenarios. Regarding atmospheric stability, unsaturated parcels follow the dry adiabatic path and may remain negatively buoyant if the environmental lapse rate is subadiabatic. In contrast, moist parcels lifted to the lifting condensation level (LCL)—the altitude where saturation is reached—experience release upon further ascent, allowing them to warm relative to the environment and become positively buoyant if the environmental exceeds Γ_m. This transition at the LCL can trigger conditional instability, where initially stable unsaturated air becomes unstable once saturation occurs, promoting and cloud development. On a , the ascent path of a dry parcel follows straight dry adiabats (typically green lines sloping at about 9.8 K/km), while a moist parcel's path shifts to curved moist adiabats (often red or magenta lines) after reaching the LCL, illustrating the reduced lapse rate in saturated conditions and the potential for the parcel to cross into positively buoyant regions above the LCL. This visual comparison underscores how moisture alters stability profiles, with the moist path diverging from the dry one to reflect latent heat's stabilizing yet conditionally destabilizing influence.

Formulation and Derivation

Core Equation for Equivalent Potential Temperature

The equivalent potential temperature, denoted as \theta_e, is defined as the temperature that a saturated air parcel would attain upon the complete removal of its moisture through condensation (in a pseudo-adiabatic process) followed by dry adiabatic descent to a standard reference pressure of 1000 hPa. This quantity extends the concept of dry potential temperature \theta by incorporating the effects of latent heat release from water vapor. A standard approximate formula for \theta_e is given by \theta_e \approx [\theta](/page/Theta) \exp\left( \frac{L r}{c_p [T_L](/page/Temperature)} \right), where \theta is the potential temperature of the air parcel (in ), L is the of vaporization (approximately $2.5 \times 10^6 J kg^{-1}), r is the total mixing ratio of (in kg kg^{-1}), c_p is the of dry air at constant pressure (1004 J kg^{-1} ^{-1}), and T_L is the temperature at the lifting condensation level (in ). This expression accounts for the additional warming due to , assuming small water vapor concentrations and constant L. An alternative form, suitable for computational applications, expresses \theta_e directly in terms of variables: \theta_e = T \left( \frac{P_0}{P} \right)^{R_d / c_{pd}} \left( 1 + \frac{0.622 r_s}{p_v} \right)^\kappa \exp\left[ \left( \frac{3.376}{T_L} - 0.00254 \right) r_s (1 + 0.81 r_s) \right], where T is the air (K), P_0 = 1000 hPa is the reference , P is the actual (hPa), R_d = 287 J kg^{-1} K^{-1} is the for dry air, c_{pd} = 1005 J kg^{-1} K^{-1} is the of dry air, r_s is the saturation mixing ratio (g kg^{-1}), p_v is the (hPa), \kappa = R_d / c_{pd} \approx 0.286, and T_L is as defined above. The exponential term incorporates corrections for the dependence of and the contribution of liquid water, while the factor involving r_s / p_v adjusts for moist air . Physically, \theta_e acts as a conserved tracer of moist for reversible processes, remaining invariant during both dry adiabatic motions and pseudo-adiabatic moist ascent or descent where occurs without mixing. It is typically expressed in , with the reference pressure standardized at P_0 = 1000 hPa to facilitate comparisons across atmospheric levels.

Derivation from Moist Adiabatic Processes

The derivation of equivalent potential temperature begins with of applied to moist air, which accounts for the of dry air, , and liquid water. states that the change in moist H satisfies dH = \delta q + V dP, where \delta q is the heat added per unit mass, V is the , and P is ; for reversible processes in moist air, this incorporates release during . Along a reversible moist adiabat, the specific entropy s of the moist air parcel is conserved, leading to the differential form ds = c_p d \ln \theta + (L / c_p) d \ln r_s = 0, where c_p is the at constant pressure (typically for dry air), \theta is the potential temperature, L is the of vaporization, and r_s is the saturation mixing ratio. This equation arises from integrating the entropy change due to temperature variations and phase changes, ensuring that the offsets the cooling from expansion without external heat addition. The conservation of s reflects the pseudo-adiabatic assumption, where condensate is removed instantaneously, maintaining near-reversibility. The equivalent potential temperature \theta_e emerges from a conceptual parcel that traces a reversible pseudo-adiabatic path. First, the unsaturated parcel is lifted dry adiabatically to the lifting condensation level (LCL), where occurs. Then, it ascends further along the moist adiabat until all has condensed out (theoretically to very low ), with released during . The resulting dry parcel is then descended dry adiabatically to the reference P_0 = 1000 hPa, where its equals \theta_e. This process conserves the total , yielding \theta_e as the potential temperature equivalent to the initial parcel's moist static . Fundamentally, \theta_e is tied to the moist , expressed as \theta_e \propto \exp(s / c_p), where s includes contributions from dry air, vapor, and the effects of ; this exponential form ensures \theta_e remains invariant along moist adiabats, serving as a conserved tracer for moist thermodynamic processes. This derivation relies on several key assumptions: reversible , where phase changes occur without irreversibility; constant L; and behavior for both dry air and components. These simplify the while capturing the essential conservation properties in the atmosphere.

Approximations for Practical Use

One widely adopted approximation for computing equivalent potential temperature (θ_e) in practical settings is the formula developed by (1980), which simplifies the iterative calculations required by the full while maintaining high accuracy across a range of conditions, including tropical environments. provides empirical formulas for the lifting level temperature and a corrected expression for the release term, typically of the form θ_e ≈ θ_L exp[ (L_v r / (c_p T_L)) ] with L_v adjusted for dependence as L_v ≈ 2.501 × 10^6 - 2.37 × 10^3 T_L (in J kg^{-1}, T_L in °C), and additional corrections for the mixing ratio effects. For saturated conditions, the LCL is at the initial level, and the formula is applied directly using the observed and saturation mixing ratio. This method avoids the need for explicit integration along pseudo-adiabatic paths, enabling efficient computation in operational settings. A simpler empirical fit, suitable for quick assessments in low-humidity scenarios, approximates θ_e as \theta_e \approx \theta + \frac{L r}{c_p} where \theta is the dry potential temperature, L is the latent heat of vaporization, r is the water vapor mixing ratio, and c_p is the specific heat capacity of dry air at constant pressure; variations of this form may incorporate dew point temperature (T_d) to refine the moisture term, such as \theta_e \approx \theta + (L r / c_p) \times (T / T_d). This linear approach stems from the first-order expansion of the exponential term in more precise formulations and is particularly useful for initial data analysis where full precision is not required. These approximations exhibit limitations in their applicability, achieving accuracy within ±1 for typical mid-latitude conditions but showing increased errors—up to 5 —in extreme or subfreezing temperatures due to assumptions about constant and specific heats. In such cases, the formulations may underestimate θ_e in highly moist air masses by neglecting higher-order contributions. In practice, these methods are integrated into meteorological software for processing observations and driving (NWP) models, including the Weather Research and Forecasting (WRF) model, where the approximation is implemented via diagnostic functions to compute θ_e from standard variables like , , and mixing ratio. This facilitates stability assessments without excessive computational overhead.

Meteorological Applications

Assessing Atmospheric Stability

Equivalent potential temperature (θ_e) serves as a key diagnostic tool for evaluating static in moist atmospheric environments, extending the principles applied to dry potential temperature in unsaturated layers. In vertical profiles, an increase in θ_e with indicates a stable atmosphere, where displaced air parcels tend to return to their original position due to forces. A constant θ_e profile suggests neutral stability, with parcels neither accelerating nor decelerating significantly upon displacement, while a decrease in θ_e with signals convective , promoting the ascent of saturated air parcels. This approach accounts for the effects of release during , providing a more comprehensive assessment than dry static stability metrics in humid conditions. A primary application of θ_e profiles is in the computation of (CAPE), which quantifies the integrated buoyant energy available for updrafts in conditionally unstable environments. CAPE is expressed as: \text{CAPE} = \int_{\text{LFC}}^{\text{EL}} \frac{g}{T} (\theta_{e,\text{parcel}} - \theta_{e,\text{env}}) \, dz where g is , T is the environmental , \theta_{e,\text{parcel}} is the equivalent potential temperature of the lifted parcel, \theta_{e,\text{env}} is the environmental equivalent potential temperature, and the occurs from the level of free convection (LFC) to the level (EL). Positive CAPE values, derived from regions where \theta_{e,\text{parcel}} > \theta_{e,\text{env}}, indicate potential for deep , with magnitudes exceeding 2000 J/kg often associated with vigorous thunderstorm development. For instance, environments featuring veering winds (clockwise turning with height) alongside steep low-level θ_e gradients can enhance rotational organization and intensity, signaling heightened potential for severe storms such as supercells. Common approximations for θ_e use the , neglecting water loading from (assuming immediate removal), which is suitable for deep convection where falls out rapidly. The reversible moist , which retains , leads to a slightly different quantity but is rarely used due to .

Use in Weather Forecasting

In operational , equivalent potential temperature (θ_e) plays a key role in sounding analysis for identifying potential sites of initiation along mesoscale boundaries such as outflow boundaries and dry lines. Meteorologists examine vertical profiles from soundings to detect maxima in θ_e, which indicate regions of high and warmth conducive to convective updrafts when intersected by these boundaries, thereby focusing and triggering storms. Numerical weather prediction models like the European Centre for Medium-Range Weather Forecasts (ECMWF) Integrated Forecasting System and the (GFS) routinely output θ_e fields to monitor moisture transport and the buildup of over forecast periods. These θ_e forecasts help forecasters track the of warm, moist air masses into target areas, assessing the potential for convective development by evaluating changes in low-level θ_e gradients and ridges. For instance, increasing θ_e values in model projections signal enhanced conditional instability, aiding in the prediction of outbreaks. A notable example of θ_e's forecasting utility occurred during the 3 May , where very high surface θ_e values exceeding 360 K across central highlighted extreme and fueled the development of multiple long-track thunderstorms. Analyses of observed and modeled θ_e distributions revealed how these high values, combined with dynamic forcing along a , contributed to the event's intensity, allowing forecasters to anticipate widespread . On synoptic scales, θ_e ridges serve as indicators of warm, moist air , often preceding heavy rainfall events by delineating areas of elevated θ_e where and uplift can initiate deep convection. These ridges, typically at 850 , align with low-level jets transporting Gulf moisture northward, with heaviest favoring the apex or downstream flank of the ridge due to the resulting moist instability. Equivalent potential temperature (θ_e) differs from wet-bulb potential temperature (θ_w) in its treatment of moisture condensation during adiabatic processes. θ_w is the temperature a saturated air parcel at the would attain at 1000 after pseudo-adiabatic ascent along the moist adiabat, while θ_e assumes complete pseudoadiabatic condensation of all , releasing the full to heat the parcel before descent to 1000 . This makes θ_e more suitable for analyzing deep convective processes where total moisture contributes to , whereas θ_w better approximates near-surface evaporative cooling and partial saturation effects. In contrast to equivalent temperature (T_e), which is the temperature an air parcel would reach if all latent heat were released at constant pressure without accounting for pressure changes, θ_e incorporates the potential temperature adjustment to a reference pressure (typically 1000 hPa). T_e = T + (L_v w / c_p), where T is the actual temperature, L_v is the latent heat of vaporization, w is the mixing ratio, and c_p is the specific heat capacity at constant pressure, but it varies significantly during vertical motion due to non-conservation under pressure changes. θ_e, defined approximately as θ_e ≈ θ exp[(L_v w) / (c_p T_L)], with θ as dry potential temperature and T_L as the LCL temperature, remains nearly conserved during both dry and moist adiabatic processes, enabling reliable tracking of air parcels across pressure levels. Compared to the total totals index (TT), a composite stability parameter used primarily for forecasting severe thunderstorms, θ_e offers a more conserved vertical profile for assessing atmospheric structure. TT = T_{850} + T_{d_{850}} - 2 T_{500}, where T denotes temperature and T_d dew point at the specified pressures (hPa), evaluates low-level moisture and mid-level lapse rates but is surface-based and non-conserved, limiting its utility for three-dimensional air mass differentiation. θ_e, by contrast, provides a single conserved value per air mass, facilitating identification of boundaries and potential instability throughout the troposphere. The primary advantages of θ_e lie in its superior conservation during moist flows, with formulation errors typically below 1 K (e.g., 0.035 K for refined approximations), compared to 5–10 K deviations in older methods like the Rossby formula or non-potential indices like T_e in varying environments. This low error enables precise analysis, where θ_e gradients delineate fronts and convective potential more reliably than θ_w (which underestimates full release) or (which ignores vertical conservation).
ParameterKey FormulaConservation PropertiesPrimary Use Cases
Potential Temperature (θ)θ = T (p_0 / p)^{R_d / c_{p d}}Conserved in dry adiabatic processes; varies in moist conditionsDry atmospheric stability assessment; comparing parcels at different pressures without moisture effects
Equivalent Potential Temperature (θ_e)θ_e ≈ θ exp[(L_v w (1 + 0.448 w)) / (c_{p d} T_L)]Nearly conserved (~0.035 K error) in both dry and pseudoadiabatic moist processesAir mass identification; moist convective stability; tracking buoyancy in deep ascent
Wet-Bulb Potential Temperature (θ_w)Iterative solution along moist adiabat from wet-bulb T_w to 1000 hPa (no closed-form; ~0.002°C error in approximations)Conserved for saturated ascent to LCL; approximate for partial moist processesNear-surface moisture analysis; evaporative cooling; frontal boundaries in unsaturated air
Equivalent Temperature (T_e)T_e = T + (L_v w / c_p)Not conserved under pressure changes (varies 5–10 K in ascent); constant pressure onlyTotal heat content estimation at a fixed level; initial moisture impact without vertical motion

Historical Development

Origins in Atmospheric Science

The concept of equivalent potential temperature (θ_e) traces its roots to foundational work in atmospheric thermodynamics during the late 19th century. In 1888, Hermann von Helmholtz introduced the notion of potential temperature (θ), defined as the temperature an air parcel would attain if adiabatically compressed or expanded to a standard reference pressure, typically 1000 hPa, thereby serving as a conserved quantity for dry adiabatic processes. Concurrently, Wilhelm von Bezold explored the thermodynamics of moist air in his seminal series of papers, developing the idea of moist adiabats—curves representing the temperature path of saturated air undergoing adiabatic ascent with condensation—and extending potential temperature concepts to account for latent heat release, laying the groundwork for moist equivalents. The term "equivalent potential temperature" was coined independently in 1921 by Charles Normand and Wilhelm Schmidt, who formalized θ_e as a measure approximating the of moist air by incorporating the effects of into the dry potential temperature framework. Normand's approximation, θ_e ≈ θ [1 + (L_v q_v)/(c_{p_d} T)], where L_v is the of , q_v the mixing ratio, c_{p_d} the specific heat of dry air, and T the , generalized earlier ideas from von Bezold's pseudo-adiabats to quantify the total heat content including . This built on prior notions of equivalent temperature introduced by Heinrich von Schubert in 1904 and Ernst Knoche in 1905, which added contributions to the actual (T_e = T + (L_v q_v)/c_{p_d}). By , Carl-Gustaf Rossby advanced its application in analysis, devising the first practical formula for θ_e and integrating it into to assess stability and moisture in weather systems, particularly emphasizing its conservation during pseudo-adiabatic processes. In mid-20th-century synoptic , Sverre Petterssen's influential textbook adopted θ_e as a key diagnostic tool for analyzing atmospheric stability and frontal systems, promoting its routine use in weather charts and to trace moist air parcels across isentropic surfaces. As emerged in the 1950s and gained traction through the 1970s, refinements to θ_e computations addressed challenges in handling moist processes within models, improving accuracy for simulations of convective instability and tropical dynamics. The 1980 approximation by David Bolton standardized practical calculations, providing an efficient formula valid across a wide range of temperatures and pressures, which became widely implemented in operational systems.

Key Contributions and Evolutions

In the numerical era of the 1980s, equivalent potential temperature (θ_e) became integrated into operational models, enhancing the analysis of moist processes. The European Centre for Medium-Range Weather Forecasts (ECMWF) incorporated refined formulations of θ_e into its Integrated Forecasting System (IFS), drawing on advancements like Bolton's (1980) simplified computational procedure, which addressed challenges in tropical environments and ice-phase conditions for more accurate parcel tracking in simulations. This inclusion facilitated better representation of and stability assessments within the model's isentropic coordinates. During this period, applications extended to orographic studies, where θ_e proved instrumental in delineating air parcel trajectories and over complex terrain. Marwitz (1980) utilized θ_e cross-sections to infer trajectories in winter storms over the , revealing how low-level equivalent potential temperature gradients influenced upslope flow and efficiency in seeding potential evaluations. These analyses highlighted θ_e's role in quantifying convective instability and orographic enhancement, informing cloud strategies for mountainous regions. In recent developments, θ_e has been increasingly employed in climate models to analyze moist static energy (MSE) budgets, providing insights into energy transport and convective organization on global scales. Since MSE is thermodynamically linked to θ_e through hydrostatic approximations—where variations in θ_e reflect MSE perturbations—its use has improved simulations of tropical circulation and dynamics in models like GISS ModelE3. The exact derivations from Iribarne and Godson (1973), which emphasize rigorous thermodynamic consistency, were revisited in the and to support high-resolution simulations, enabling precise computation of θ_e in grid-scale moist processes without approximation errors. Ongoing research leverages θ_e in ensemble forecasting to quantify uncertainties in convective , particularly through gradients that signal thresholds. In convection-permitting ensembles, θ_e perturbations help assess the spread in low-level convergence and , improving probabilistic forecasts of mesoscale convective systems. For instance, ensemble-mean θ_e contours reveal sensitivities to initial conditions, aiding predictions of elevated where vertical θ_e profiles indicate potential for despite stable layers. Adaptations of θ_e concepts have emerged post-2010 for atmospheres, extending moist thermodynamic frameworks to model heat redistribution in habitable zones. In simulations of terrestrial , θ_e analogs track moist and baroclinic instabilities, influencing profiles and circulation patterns under varying stellar irradiation.

References

  1. [1]
    NOAA's National Weather Service - Glossary
    ### Summary of Definitions from https://forecast.weather.gov/glossary.php?word=potential%20temperature
  2. [2]
    Potential Temperature - an overview | ScienceDirect Topics
    The equivalent-potential temperature Θe is the potential temperature corresponding to the equivalent temperature Te, i.e.. (9.26) Θ e = T e ...
  3. [3]
    Meteorological explanation - Potential temperature - EUMeTrain
    The equivalent potential temperature is found by lowering an air parcel to the 1000 hPa level AND releasing the latent heat in the parcel. The lifting of a ...
  4. [4]
    Skew-T Parameters and Indices
    Equivalent Potential Temperature - Also known as THETA-E. Temperature of a parcel after all latent heat energy is released in a parcel then brought to the 1000 ...
  5. [5]
    Trends in surface equivalent potential temperature: A more ... - PNAS
    Jan 31, 2022 · Here we show that surface equivalent potential temperature, which combines the surface air temperature and humidity, is a more comprehensive ...
  6. [6]
    [PDF] Stratification
    1The buoyancy frequency is also known as the Brunt-Väisälä frequency. David ... Here we restrict ourselves to an analysis of an incompressible and inviscid fluid.
  7. [7]
    [PDF] Chapter 2 Buoyancy and Coriolis forces
    N = gΓ ρ0 1/2 . This is called the Brunt-Väisälä or buoyancy frequency, and it is the oscillation frequency of a parcel displaced vertically in an ...
  8. [8]
    [PDF] Chapter 5: Mixing processes, mixed layer and thermocline
    The stratification is stable. Often, we use "buoyancy frequency N" to represent the stratification of the ocean, where. N2 = -g¹d¹dº. It is also ...
  9. [9]
    [PDF] 6. Potential Temperature ( )
    In meteorology, it proves extremely useful to label air parcels with their potential temperature. In fact, the potential temperature is defined by θ κ. = æ è.
  10. [10]
    [PDF] Atmospheric Thermodynamics
    Typically this equation is applied to dry air, so R = Rd and cp = cpd. This gives: Example: What is the potential temperature at the ATOC weather station?
  11. [11]
    2.5 Adiabatic Processes: The Path of Least Resistance | METEO 300
    We call θ the potential temperature, which is the temperature that an air parcel would have if the air is brought to a pressure of po = 1000 hPa. Potential ...
  12. [12]
    [PDF] Notes on Potential Temperature - Weather and Climate Laboratory
    Indeed the atmosphere closely obeys the perfect gas law,. = p=RT , where R is the gas constant for dry air ... For a more complete derivation of potential ...Missing: meteorology | Show results with:meteorology
  13. [13]
    Chapter 5: Atmospheric Stability - UH Pressbooks
    The Lifting Condensation Level (LCL) is the level at which the water vapor in an air parcel that is lifted dry adiabatically will be saturated. The red dot is ...
  14. [14]
    Notes on Air Temperature and Clouds With Flow Over a Mountain
    This release of latent heat is what reduces the lapse rate to 6 C per kilometer (in other words, the release of heat due to condensation is added to the DALR ( ...
  15. [15]
    Chapter 4: Water Vapor – Atmospheric Processes and Phenomena
    Question 5. The moist adiabatic lapse rate (~4.5 K/km) is less than the dry adiabatic lapse rate (9.8 K/km) because of latent heat release that occurs within a ...
  16. [16]
    [PDF] Chapter 6: Stability - UCI ESS
    This moist adiabatic lapse rate is not a constant but determined by considering the combined effects of expansion cooling and latent heating.
  17. [17]
    [PDF] Moisture Variables (and Thermodynamic Variables)
    Saturation mixing ratio: (ws) The ratio of the mass of water vapor assuming a parcel is saturated with water vapor to the mass of dry air containing the vapor.Missing: content | Show results with:content
  18. [18]
    [PDF] MET 3502/5561 Synoptic Meteorology
    Pseudo-adiabatic means all the condensed water vapor is assumed to fall out immediately as the parcel rises adding its latent heat of condensation to the parcel ...
  19. [19]
    [PDF] Ooyama(1990).pdf
    Nov 1, 1990 · The pseudoadiabatic process, which assumes immediate removal of condensate, is also unrealistic in the opposite sense. If we may draw any ...
  20. [20]
    JetStream Max: Severe Weather - NOAA
    Jun 2, 2023 · Lifting Condensation Level (LCL) ... A parcel with a humidity of less than 100% will cool at the dry adiabatic rate of 9.8°C/Km as it lifts from ...
  21. [21]
    Assessing Strong Updrafts | Learning Weather at Penn State ...
    Recall from your previous studies that the lifting condensation level (LCL) is the level where net condensation begins in a lifted parcel. If the parcel ...
  22. [22]
    [PDF] MET 200 Lecture 8 Stability - SOEST Hawaii
    An adiabatic process assumes no heat exchange occurs across the parcel's boundary. 3.Moist adiabatic lapse rate occurs in a cloud because latent heat is ...<|control11|><|separator|>
  23. [23]
    [PDF] A GUIDE TO THE SKEW-T / LOG-P DIAGRAM
    Lifting Condensation Level (LCL): The level at which a parcel of air first becomes saturated when lifted dry adiabatically. This level can be found by finding ...
  24. [24]
    [PDF] How do Scientists Forecast Thunderstorms?
    Moist adiabatic lapse rate (MALR) - These lines represent the rate in which a rising saturated parcel of air will cool (depends on moisture content of air).Missing: reduced | Show results with:reduced
  25. [25]
    [PDF] SKEW-T, LOG-P DIAGRAM ANALYSIS PROCEDURES
    Sep 11, 2007 · Saturation adiabats tend to become parallel to the dry adiabats at low values of moisture, temperature, and pressure. They extend only to the ...Missing: ascent reduced
  26. [26]
    The Computation of Equivalent Potential Temperature in
    A simplified procedure is described for computation of equivalent potential temperature which remains valid in situations such as in the tropics.
  27. [27]
    On the Computation of Pseudoadiabatic Entropy and Equivalent ...
    Dec 1, 2008 · A set of approximate equations for pseudoadiabatic thermodynamics is developed. The equations are derived by neglecting the entropy of water vapor.
  28. [28]
    On Formulas for Equivalent Potential Temperature in - AMS Journals
    Sep 1, 2009 · The exact formula for θE is (3.7), but it is ... formula for saturation pseudo-adiabats and saturation equivalent potential temperature.
  29. [29]
    [PDF] Chapter 2 Approximate Thermodynamics
    The moist entropy and the equivalent potential temperature are conserved not only in processes without heating, but also in processes in which heating and ...
  30. [30]
    [PDF] Atmospheric Thermodynamics - ECMWF
    Equivalent potential temperature and moist static energy. The equivalent potential temperature is readily obtained by integration of (44). It is the temperature.<|control11|><|separator|>
  31. [31]
    wrf_eth
    wrf_eth calculates equivalent potential temperature from WRF model output, using water vapor mixing ratio, temperature, and pressure. It returns the equivalent ...
  32. [32]
    [PDF] using isentropic techniques to improve the
    When the θe difference is negative (posi- tive), θe is decreasing (increasing) with height through the layer, so the atmosphere is convectively unstable (stable) ...<|control11|><|separator|>
  33. [33]
    [PDF] 24th Conference on Severe Local Storms
    Past research has shown these low-level wind anomalies coupled with a sharp 25 K theta-e gradient or greater at the 1000, 925, or 850 hPa levels over ≤. 400 km ...
  34. [34]
    THUNDERSTORMS
    Frontal boundaries, dry lines, and outflow boundaries (low level convergence) ... *Important to look for thermal and dewpoint ridges (THETA-E) *For tornado ...
  35. [35]
    Glossary - NOAA's National Weather Service
    ... increasing with height). Tropopause Jet: Type of jet stream found near the ... It is best analyzed between 750-550 milibars using equivalent potential temperature ...
  36. [36]
    Parameter Database - ECMWF
    Oct 10, 2025 · Parameter database ; Equivalent potential temperature. eqpt. K · 4. GRIB2. GRIB1 ; Saturated equivalent potential temperature. sept. K · 5. GRIB2.Potential temperature · ECMWF (European Center for... · Parameter detail
  37. [37]
    Comparison of Evaporation and Cold Pool Development between ...
    ... 1999 Oklahoma tornado outbreak, with an emphasis on comparing ... Figure 1 shows an objective analysis of surface equivalent potential temperature θe ...
  38. [38]
    Synoptic Patterns for Heavy Rainfall in Minnesota
    The warm, moist air indicated by high θe is favorable for heavy rainfall, especially near the apex of the ridge of high θe values. Composites of divergence ...
  39. [39]
    [PDF] Heavy Convective Rainfall Patterns in the Midwest and Ohio Valley
    • Equivalent potential temperature (θe) and θe advection: Higher θe = higher temp and/or moisture at a level. Warm sector storms: heavy rain along/near 850 ...
  40. [40]
    RELATION BETWEEN EQUIVALENT POTENTIAL TEMPERATURE ...
    RELATION BETWEEN EQUIVALENT POTENTIAL TEMPERATURE AND WET-BULB POTENTIAL TEMPERATURE.
  41. [41]
    An Efficient and Accurate Method for Computing the Wet-Bulb ...
    In contrast, there are precise explicit formulas for equivalent potential temperature (EPT) θE (K) so we can easily calculate the parcel's equivalent ...Abstract · Introduction · Computation of temperature... · The Wobus method
  42. [42]
    On the conservatism of the equivalent potential and the wet‐bulb ...
    The use of the wet-bulb potential temperature may give a wrong impression of the gain of water by evaporation, and may lead to erroneous results in estimations ...
  43. [43]
    [PDF] A GUIDE TO THE SKEW-T / LOG-P DIAGRAM
    3. Equivalent Potential Temperature (Θe): The equivalent potential temperature is similar to the equivalent temperature however after the moisture has been ...
  44. [44]
  45. [45]
    Env Parameters and Indices - National Weather Service
    The DCI attempts to combine the properties of equivalent potential temperature (Qe) at 850 mb with instability. ... warm air advection (veering between 850 and ...
  46. [46]
    [PDF] NOAA Technical Memorandum NWS SR-109
    saturation mixing ratio The mixing ratio a sample of air would have if saturated ... used to determine the mean wet bulb temperature of the moist layer.
  47. [47]
    [PDF] A noniterative approach to modelling moist thermodynamics - ACP
    ... Bolton's formula (Bolton,. 20. 1980) for equivalent potential temperature valid for the pressure range 10 ≤ P ≤ 105 kPa and wet-bulb potential temperatures. − ...<|separator|>
  48. [48]
    [PDF] potential temperature - ACP
    Dec 15, 2020 · Starting from the equation for the moist spe- cific entropy, as derived from the first law of thermodynamics and the Gibbs equation, further ...
  49. [49]
    [PDF] definitions and applications in Meteorology and C - HAL Météo-France
    Sep 24, 2021 · Several other equivalent potential temperature formulations have been defined, but they all corre- ... by von Bezold (1888a) and Saunders (1957) ...
  50. [50]
    Thermodynamics applied to air mass analysis - Semantic Scholar
    The equivalent potential temperature θe, the liquid water potential temperature θℓ, and the entropy potential temperature θs are… Expand. 6 Citations · PDF. Add ...
  51. [51]
    [PDF] Part IV: Physical Processes - ECMWF
    the rhs of (6.72) using the updraught equivalent potential temperature θe = T ... Efficient prediction of ground surface temperature and moisture, with ...
  52. [52]
  53. [53]
    Intercomparison of different state-of-the-art formulations of the mass ...
    Dec 17, 2021 · The differences between one classical and three state-of-the-art formulations of the mass density of humid air were quantified.
  54. [54]
    [PDF] Formulations of moist thermodynamics for atmospheric modelling
    Sep 24, 2021 · The liquid-water and equivalent potential temperatures θl and θe defined in Betts (1973) or. Emanuel (1994) are conserved quantities for ...Missing: 2000s | Show results with:2000s
  55. [55]