Fact-checked by Grok 2 weeks ago

Potential vorticity

Potential vorticity (PV) is a conserved scalar in that quantifies the rotation of a element relative to its , defined generally as the of absolute and the of a conserved thermodynamic property (such as potential temperature) divided by . In the atmosphere, Ertel's potential vorticity takes the form q = \frac{\vec{\zeta}_a \cdot \nabla \theta}{\rho}, where \vec{\zeta}_a is absolute , \theta is potential temperature, and \rho is , making it materially conserved for adiabatic, frictionless along isentropic surfaces. This , first generalized by Hans Ertel in 1942 building on earlier work by Carl-Gustaf Rossby in 1940, allows PV to serve as a tracer for motion in rotating, stratified systems like the atmosphere and oceans. PV plays a central role in diagnosing large-scale atmospheric and dynamics, as its implies that changes in must balance alterations in or planetary effects, such as the Coriolis varying with . For instance, in the shallow-water approximation relevant to both geophysical contexts, PV is expressed as the absolute divided by fluid depth (q = \zeta_a / h), highlighting how vertical stretching of fluid columns amplifies relative . This interplay underpins phenomena like development, where intrusion of high-PV air from the into the can trigger , or gyre formation driven by and gradients. Non-conservative processes, including diabatic heating from or from surface drag, can generate or destroy PV, influencing patterns and variability. In practical applications, is mapped on isentropic surfaces to identify dynamical breaks or fronts, aiding meteorologists in forecasting mid-latitude storms and evolution. Similarly, in , PV conservation helps model mesoscale eddies and the meridional overturning circulation by tracing water masses with distinct PV signatures. The concept's versatility extends to numerical weather prediction models, where PV inversion techniques reconstruct balanced flows from PV distributions, underscoring its enduring importance since its formalization over eight decades ago.

Historical Foundations

Bjerknes Circulation Theorem

The Bjerknes circulation theorem, formulated by Norwegian meteorologist in 1898, represents a foundational principle in , extending Lord Kelvin's earlier circulation theorem to baroclinic fluids on a rotating . Published in the Proceedings of the Royal Swedish Academy of Sciences, Bjerknes' work marked a pivotal shift toward applying hydrodynamic principles to atmospheric and oceanic motions, emphasizing the role of density variations in generating circulation changes. The states that the material rate of change of circulation C = \oint \mathbf{v} \cdot d\mathbf{l} around a closed material in an inviscid is given by the of the non-conservative forces per unit mass acting on the elements along the . In the of geophysical fluids in a rotating frame, this simplifies to \frac{DC}{Dt} = \oint \left( -\frac{1}{\rho} \nabla p - \nabla \Phi \right) \cdot d\mathbf{l}, where \mathbf{v} is the , \rho is , p is , and \Phi is the . The does not contribute to this integral over a closed loop, as its form -2 \boldsymbol{\Omega} \times \mathbf{v} (with \boldsymbol{\Omega} as Earth's angular velocity vector) yields zero when dotted with d\mathbf{l} and integrated, due to its perpendicularity to the velocity direction. For absolute circulation C_a = \oint (\mathbf{v} + \boldsymbol{\Omega} \times \mathbf{r}) \cdot d\mathbf{l}, the theorem holds analogously, incorporating planetary rotation effects without altering the force contributions. This result derives from the Navier-Stokes momentum equations in a rotating frame, adapted for geophysical applications assuming inviscid, hydrostatic conditions: \frac{D\mathbf{v}}{Dt} = -2 \boldsymbol{\Omega} \times \mathbf{v} - \frac{1}{\rho} \nabla p - \nabla \Phi. To obtain the circulation theorem, consider the total time derivative of C following the material curve, which accounts for both the local change in velocity and the deformation of the curve itself (via ). This yields \frac{DC}{Dt} = \oint \frac{D\mathbf{v}}{Dt} \cdot d\mathbf{l}. Substituting the momentum equation and noting that the advective and Coriolis terms integrate to zero over the closed loop leaves only the and terms, as shown above. In baroclinic fluids where varies independently of (i.e., isobars and isopycnals do not coincide), the pressure term does not vanish and generates circulation through "solenoidal" fields—vector areas formed by gradients of and k = 1/\rho. Applying transforms the line integral to a : \frac{DC}{Dt} = \iint_S \left( \nabla k \times \nabla p \right) \cdot d\mathbf{A}, quantifying the net flux of these solenoids through the material surface S. The pressure gradient force -\frac{1}{\rho} \nabla p drives horizontal flows, while the buoyancy contribution arises from -\nabla \Phi, which integrates to zero over a closed loop in a conservative gravitational field but interacts with density stratification to influence vertical motions indirectly through hydrostatic balance. In Bjerknes' original acceleration form, the theorem equates the circulatory acceleration of fluid elements to the enclosed solenoids, each contributing a unit acceleration proportional to the misalignment of isobaric and isosteric surfaces, thus linking non-uniform density distributions to rotational tendencies in the atmosphere and oceans. By localizing the circulation theorem via (C \approx \iint \boldsymbol{\zeta} \cdot d\mathbf{A}, where \boldsymbol{\zeta} = \nabla \times \mathbf{v} is relative ), Bjerknes' result underpins the derivation of the in rotating, stratified fluids: \frac{D \boldsymbol{\zeta}}{Dt} = (\boldsymbol{\zeta} \cdot \nabla) \mathbf{v} - (\nabla \cdot \mathbf{v}) \boldsymbol{\zeta} + \frac{1}{\rho^2} (\nabla \rho \times \nabla p) + 2 (\boldsymbol{\Omega} \cdot \nabla) \mathbf{v} - 2 (\nabla \cdot \mathbf{v}) \boldsymbol{\Omega}, with the baroclinic term \frac{1}{\rho^2} \nabla \rho \times \nabla p (equivalent to -\nabla k \times \nabla p) directly mirroring the solenoid generation of circulation. This connection highlights how baroclinicity sources , essential for understanding large-scale geophysical circulations like cyclones.

Rossby's Contributions and Early Developments

Carl-Gustaf Rossby made seminal contributions to the development of potential vorticity (PV) during , introducing it as a conserved tracer in simplified models of geophysical flows. Building briefly on ' circulation theorem from the early 20th century, which described the evolution of circulation in fluid parcels, Rossby shifted focus toward material invariants that could track large-scale motions more effectively. In his 1936 paper "Dynamics of Steady Ocean Currents in the Light of Experimental ," he first formulated PV for barotropic shallow water systems, demonstrating its utility in analyzing steady circulations where fluid parcels conserve this quantity under adiabatic, frictionless conditions. Throughout the late 1930s and into the 1940s, Rossby's work evolved amid collaborations with prominent meteorologists, including Bernhard Haurwitz and members of the emerging Chicago school of meteorology, which he helped establish after joining the University of Chicago in 1940. His 1938 publication "On the Mutual Adjustment of Pressure and Velocity Distributions in Simple Current Systems" in the Journal of Marine Research extended PV concepts to continuously stratified atmospheres, emphasizing its role in displacements of circulation patterns. By 1940, in the influential paper "Planetary Flow Patterns in the Atmosphere" published in the Quarterly Journal of the Royal Meteorological Society, Rossby coined the term "potential vorticity" and applied it to explain the dynamics of long atmospheric waves, marking a transition from circulation-based analyses to PV as a fundamental conserved property in barotropic models. These efforts reflected a broader timeline of innovation, as Rossby integrated observational data from weather stations with theoretical insights to interpret hemispheric-scale flows. Early applications of PV centered on shallow water approximations, where Rossby conceptualized it as the sum of relative and planetary divided by the depth, serving as an that governed parcel trajectories in rotating, stratified systems. This approach proved effective for studying barotropic instabilities and the adjustment of flows to geostrophic balance, providing a tracer-like tool to predict how disturbances propagate without explicit tracking of fields. Rossby's formulations highlighted PV's dimensions akin to ordinary , facilitating analogies between oceanic and atmospheric circulations. Rossby's PV innovations profoundly influenced the origins of (NWP), laying theoretical groundwork for mid-20th-century advances in forecasting large-scale patterns. By framing atmospheric dynamics around PV conservation, his ideas enabled simplifications that his protégés, such as Jule Charney, incorporated into quasi-geostrophic models during efforts at . These developments culminated in the first operational NWP systems in the , where PV-based diagnostics improved the simulation of Rossby waves and evolutions, transforming subjective analysis into computational frameworks.

Core Concepts

Vorticity and Absolute Vorticity

In , quantifies the local of fluid elements around a point. The relative \boldsymbol{\zeta}, which describes the relative to a non-rotating , is defined as the of the velocity field \mathbf{v}: \boldsymbol{\zeta} = \nabla \times \mathbf{v}. This captures the infinitesimal circulation per unit area, derived from the antisymmetric part of the velocity gradient tensor. In three-dimensional flows, the components are given by \zeta_x = \frac{\partial w}{\partial y} - \frac{\partial v}{\partial z}, \quad \zeta_y = \frac{\partial u}{\partial z} - \frac{\partial w}{\partial x}, \quad \zeta_z = \frac{\partial v}{\partial x} - \frac{\partial u}{\partial y}, where u, v, and w are the velocity components in the x, y, and z directions, respectively. In many geophysical contexts, such as atmospheric and oceanic flows approximated as two-dimensional in the horizontal plane (with negligible vertical velocity), attention focuses on the vertical component \zeta = \zeta_z = \frac{\partial v}{\partial x} - \frac{\partial u}{\partial y}, which measures the rotation in the horizontal plane. This scalar form arises directly from the curl operator applied to horizontal velocity gradients, representing the net tendency for fluid parcels to rotate clockwise or counterclockwise. In rotating reference frames like Earth's, the planetary vorticity due to the planet's rotation must be accounted for. The Coriolis parameter f, which approximates the vertical component of twice the planetary angular velocity vector, is defined as f = 2 \Omega \sin \phi, where \Omega \approx 7.292 \times 10^{-5} s^{-1} is Earth's angular rotation rate and \phi is the latitude. At mid-latitudes (around 45°), \sin \phi \approx 0.707, yielding f \approx 10^{-4} s^{-1}. The absolute vorticity \eta, combining relative and planetary contributions, is then \eta = \zeta + f for the vertical component in shallow, geostrophically balanced flows. Vorticity has units of inverse time (s^{-1}), reflecting its role as an angular velocity measure. In geophysical scales, such as mid-latitude synoptic systems (e.g., weather fronts spanning ~ km), relative \zeta typically scales as $10^{-5} to $10^{-4} s^{-1}, often comparable to or smaller than f, highlighting the dominance of planetary in large-scale . This scaling underscores 's importance in balancing Coriolis effects against relative motions in rotating fluids.

Potential Temperature and Stratification

Potential temperature, denoted as \theta, is a thermodynamic variable that represents the temperature a parcel of dry air would attain if adiabatically compressed or expanded to a standard reference pressure p_0, typically 1000 hPa, without heat exchange with its surroundings. It is mathematically defined by the formula \theta = T \left( \frac{p_0}{p} \right)^{R / c_p}, where T is the actual of the air parcel, p is its , R is the specific gas constant for dry air (approximately 287 J kg⁻¹ K⁻¹), and c_p is the specific heat capacity of dry air at constant (approximately 1004 J kg⁻¹ K⁻¹). This quantity is conserved for reversible adiabatic processes in dry air, making it a fundamental tracer for analyzing atmospheric motion and stability. The derivation of potential temperature stems from the first law of thermodynamics applied to an adiabatic, reversible process for an ideal gas. For such a process, the heat added dq = 0, leading to c_p dT = \alpha dp, where \alpha = RT/p is the specific volume from the ideal gas law. Integrating this yields \ln \theta = \ln T + (R/c_p) \ln (p_0 / p) + \text{constant}, which simplifies to the Poisson equation form of \theta. Hydrostatic balance, expressed as dp/dz = -\rho g, underpins the vertical structure in the atmosphere, ensuring that pressure decreases with height and providing the context for adiabatic displacements along isentropic surfaces where \theta remains constant. This balance is crucial for interpreting \theta gradients in stably stratified flows, as it links pressure changes to gravitational forces without horizontal accelerations. In stratified fluids like the atmosphere, potential temperature governs static stability by indicating whether displaced parcels return to their original position. A positive vertical gradient d\theta/dz > 0 signifies stable stratification, where the atmosphere resists vertical motions. The Brunt-Väisälä frequency, a measure of this oscillatory stability, is given by N^2 = \frac{g}{\theta} \frac{d\theta}{dz}, where g is gravitational acceleration (approximately 9.81 m s⁻²). When N^2 > 0, parcels oscillate with frequency N; if N^2 < 0, the flow is unstable to convection. This frequency quantifies the restoring force due to buoyancy in a stratified medium, with typical tropospheric values ranging from 10⁻² to 10⁻¹ s⁻¹ in stable layers. Potential temperature is particularly important in dry atmospheres, where it serves as the primary conserved variable for parcel tracking, but its application extends to moist conditions with modifications like equivalent potential temperature \theta_e, which accounts for latent heat release during condensation. In moist air, \theta alone underestimates buoyancy due to water vapor's lower molecular weight, necessitating virtual potential temperature for accurate density assessments. In oceanography, the analog is potential density, referenced to a standard pressure (often surface level), defined as \sigma_\theta = \rho(S, \Theta, 0) - 1000 kg m⁻³, where \rho is seawater density, S is salinity, and \Theta is potential temperature; this preserves neutrality in adiabatic displacements and assesses oceanic stratification similarly to \theta in air./06%3A_Temperature_Salinity_and_Density/6.05%3A_Density)

Key Formulations

Shallow Water Potential Vorticity

The shallow water potential vorticity represents the foundational and simplest expression of potential vorticity, applicable to barotropic, rotating fluids where the horizontal scale greatly exceeds the vertical scale, such as in idealized models of ocean or atmospheric layers. Developed by , it quantifies the vertical component of absolute vorticity normalized by the fluid depth, capturing the interplay between rotation, relative motion, and layer thickness in conserving angular momentum. Absolute vorticity here refers to the sum of relative vorticity and the planetary vorticity due to Earth's rotation. The formulation is given by q = \frac{\zeta + f}{h}, where \zeta is the relative vorticity (the curl of the horizontal velocity), f = 2 \Omega \sin \phi is the (\Omega is Earth's angular velocity and \phi is latitude), and h is the total fluid depth. This expression arises from applying , which states that the circulation \Gamma around a closed material contour in a barotropic fluid is conserved in the absence of friction and diabatic heating (D\Gamma / Dt = 0). To derive q, consider a vertical material column of fluid with cross-sectional area A and height h; the theorem implies conservation of the circulation per unit area, but vertical stretching or compression of the column (via changes in h) modifies the vorticity, leading to the normalization by h for material invariance. Specifically, integrating the theorem over the column's base and dividing by A h yields the conservation of q. In this shallow water context, q carries units of s^{-1} m^{-1}, reflecting vorticity (s^{-1}) divided by depth (m), though it aligns conceptually with the more general potential vorticity unit (PVU) of 10^{-6} K m^{2} kg^{-1} s^{-1} when stratification is absent. Under frictionless and adiabatic conditions, q is materially conserved, satisfying the equation \frac{Dq}{Dt} = 0, or equivalently \partial q / \partial t + \mathbf{v} \cdot \nabla q = 0, where \mathbf{v} is the horizontal velocity; this ensures that fluid parcels retain their q value as they advect, enabling diagnostics of flow evolution from initial conditions.

Ertel's Potential Vorticity

Ertel's potential vorticity provides a general framework for understanding the dynamics of stratified, rotating fluids in three dimensions, applicable to both atmospheric and oceanic contexts. Defined as q = \frac{1}{\rho} \vec{\omega}_a \cdot \nabla \theta, where \rho denotes the fluid density, \vec{\omega}_a = \nabla \times \vec{u} + 2 \vec{\Omega} is the absolute vorticity vector (\vec{u} being the velocity and \vec{\Omega} the planetary rotation vector), and \theta the potential temperature, this quantity captures the interplay between rotation, stratification, and motion. In adiabatic, inviscid, barotropic flow, q is materially conserved, meaning its value remains constant following individual fluid parcels. The theorem originates from Hans Ertel's 1942 analysis, which leverages fundamental conservation laws via the chain rule applied to material derivatives. The derivation commences with the thermodynamic equation for potential temperature, \frac{D\theta}{Dt} = 0, under adiabatic conditions, and the absolute vorticity equation, \frac{D \vec{\omega}_a}{Dt} = \left( \vec{\omega}_a \cdot \nabla \right) \vec{u} + \frac{1}{\rho^2} \nabla \rho \times \nabla p + \frac{1}{\rho} \nabla \times \vec{F}, where \vec{F} represents body forces (often zero in standard derivations) and the baroclinic term \nabla \rho \times \nabla p arises from the momentum equations. The continuity equation, \frac{D\rho}{Dt} + \rho \nabla \cdot \vec{u} = 0 or equivalently \frac{D}{Dt} (\frac{1}{\rho}) = \frac{1}{\rho} \nabla \cdot \vec{u}, ensures mass conservation. To derive the evolution of q, consider the material derivative: \frac{Dq}{Dt} = \frac{1}{\rho} \frac{D}{Dt} (\vec{\omega}_a \cdot \nabla \theta) - \frac{q}{\rho} \frac{D\rho}{Dt}. Expanding the first term using the product rule and chain rule yields \frac{D}{Dt} (\vec{\omega}_a \cdot \nabla \theta) = \frac{D\vec{\omega}_a}{Dt} \cdot \nabla \theta + \vec{\omega}_a \cdot \frac{D}{Dt} (\nabla \theta) = \frac{D\vec{\omega}_a}{Dt} \cdot \nabla \theta + \vec{\omega}_a \cdot \left[ (\nabla \cdot \vec{u}) \nabla \theta - (\nabla \vec{u})^T \cdot \nabla \theta \right], where the second part follows from the transport theorem for gradients. Substituting the vorticity equation and simplifying, the stretching term (\vec{\omega}_a \cdot \nabla) \vec{u} \cdot \nabla \theta partially cancels with the transpose term, leaving the baroclinic contribution \frac{1}{\rho^2} (\nabla \rho \times \nabla p) \cdot \nabla \theta. Incorporating the continuity-derived term -\frac{q}{\rho} \frac{D\rho}{Dt} = -q \nabla \cdot \vec{u} further balances the divergence effects. For barotropic conditions (\nabla \rho \times \nabla p = 0) and no forcing (\vec{F} = 0), all terms vanish, yielding \frac{Dq}{Dt} = 0. This rigorous application of vector calculus to the primitive equations underscores the theorem's generality. In oceanic applications, where salinity influences density stratification, Ertel's theorem is generalized by replacing potential temperature with a conserved scalar incorporating salinity S, such as the entropy \eta or a potential density variable that accounts for both temperature and salinity effects. The form becomes q = \frac{1}{\rho} \vec{\omega}_a \cdot \nabla \eta, with conservation holding for volumes bounded by surfaces of constant \eta or S under moist adiabatic conditions without precipitation or friction; this adaptation preserves the theorem's utility for multi-component fluids. Non-conservative processes modify the evolution equation \frac{Dq}{Dt} \neq 0. Diabatic heating introduces a source term \frac{1}{\rho} \vec{\omega}_a \cdot \nabla \left( \frac{D\theta}{Dt} \right), reflecting changes in stratification due to heat addition or removal, while frictional effects contribute through the curl of viscous stresses in the , typically dissipating q in boundary layers. These terms highlight q's sensitivity to thermodynamic and dissipative influences, limiting strict conservation to idealized scenarios.

Quasi-Geostrophic Potential Vorticity

The quasi-geostrophic (QG) approximation to potential vorticity arises from scale analysis applied to Ertel's potential vorticity under the assumptions of nearly geostrophic balance and small Rossby number (Ro ≪ 1), where the Rossby number Ro = U/(f L) compares inertial to Coriolis forces, with U the characteristic velocity, f the Coriolis parameter, and L the horizontal length scale. This scaling neglects higher-order advective and ageostrophic terms, retaining leading-order contributions from planetary rotation, relative vorticity gradients, and vertical stretching due to stratification, while assuming hydrostatic balance and a basic state with constant f ≈ f₀ (the Coriolis parameter at a reference latitude). The result is a materially conserved quantity q that governs large-scale, slowly evolving flows, simplifying the full three-dimensional dynamics into a diagnostic equation suitable for numerical modeling and theoretical analysis. The QG potential vorticity is expressed as q = \nabla^2 \psi + \beta y + \frac{f_0}{\theta_0} \frac{\partial \theta}{\partial z}, where \psi is the geostrophic streamfunction (related to the geopotential via \psi = \Phi / f_0), \beta = \partial f / \partial y is the Rossby parameter (approximately 1.6 × 10⁻¹¹ m⁻¹ s⁻¹ at 45° latitude), y is the northward coordinate, f₀ is the reference Coriolis parameter, \theta_0 is the reference potential temperature, and \theta is the perturbation potential temperature. The stretching term \frac{f_0}{\theta_0} \frac{\partial \theta}{\partial z} accounts for baroclinic effects through the vertical gradient of potential temperature, which modulates the column thickness under hydrostatic balance. This form is conserved following the geostrophic flow: \frac{D_g q}{Dt} = 0, where \frac{D_g}{Dt} = \frac{\partial}{\partial t} + \mathbf{u}_g \cdot \nabla uses the geostrophic velocity \mathbf{u}_g = (-\partial \psi / \partial y, \partial \psi / \partial x). In its hybrid representation, q combines three primary contributions: relative vorticity \zeta_g = \nabla^2 \psi (from horizontal shears in the geostrophic wind), planetary vorticity \beta y (from the latitudinal variation of the , inducing meridional gradients), and the baroclinic term \frac{f_0}{\theta_0} \frac{\partial \theta}{\partial z} (reflecting vertical variations in density or temperature that alter vortex stretching). These terms capture the essential dynamics of balanced flows, where anomalies in q drive geopotential tendencies via elliptic inversion, linking surface pressure patterns to upper-level structures. This approximation holds for mid-latitude synoptic-scale motions (horizontal scales of 500–2000 km and timescales of 1–10 days), where Ro ≈ 0.1 and the deformation radius L_D = N H / f₀ (with buoyancy frequency N and height scale H) is comparable to L, ensuring geostrophic dominance. However, it introduces errors in the tropics, where small f and β lead to larger Ro (>0.3) and weaker geostrophic balance, necessitating extensions like equatorial or tropical approximations.

Interpretations and Properties

Physical and Dynamical Interpretation

Potential vorticity (PV) serves as a fundamental diagnostic tracer in atmospheric and oceanic dynamics, representing the intersection of surfaces with isentropic surfaces, which are surfaces of constant potential temperature. Physically, PV quantifies the density of vortex lines piercing a unit area of an isentropic surface, providing insight into the three-dimensional structure of motion. This interpretation arises because vortex lines, which are tangent to the , are "frozen" into the in the absence of diabatic processes, and their concentration on isentropic surfaces reflects the interplay between and . The of thickness further elucidates this: a thinner tube cross-section on the isentropic surface corresponds to higher PV, as the same circulation is concentrated over a smaller area, amplifying the effective density. Dynamically, PV encapsulates the balance between absolute vorticity and static stability, offering a qualitative measure of rotational and buoyant forces within fluid parcels. High PV values typically indicate regions of strong combined with stable stratification, where the positive alignment of and the vertical gradient of potential temperature enhances the overall dynamical activity. Conversely, low or negative PV signifies anticyclonic rotation or unstable stratification, often associated with weaker or opposing dynamical influences that can lead to baroclinic . This dual role allows PV to diagnose the potential for intense systems, as high PV parcels act as coherent structures that influence surrounding flow through their conserved nature./11:_General_Circulation/11.9:_Types_of_Vorticity) In frontal and upper-level contexts, PV anomalies provide a sharp indicator of dynamical boundaries, particularly at the tropopause where jet streams are prominent. Positive PV anomalies descending from the or forming at the tropopause interface often mark the poleward edge of jet streams, with steep PV gradients delineating the transition between tropospheric and stratospheric air masses. These anomalies drive ageostrophic circulations that intensify jets and fronts, as the intrusion of stratospheric air—characterized by high PV—enhances cyclonic shear and vertical motion. Such interpretations underscore PV's utility in mapping synoptic-scale features without relying on direct height or wind fields. To standardize comparisons across scales, PV is conventionally expressed in potential vorticity units (PVU), where $1 \, \text{PVU} = 10^{-6} \, \text{K} \, \text{m}^2 \, \text{kg}^{-1} \, \text{s}^{-1}. This unit arises from the dimensional form of Ertel's PV, \Pi = \frac{1}{\rho} \vec{\zeta}_a \cdot \nabla \theta, ensuring practical values around 1–2 PVU near the and lower in the ./11:_General_Circulation/11.9:_Types_of_Vorticity)

Conservation and Invertibility Principle

Potential vorticity (PV), as defined by Ertel, is materially conserved in inviscid, adiabatic flow, meaning that the material derivative of PV vanishes: \frac{Dq}{Dt} = 0, where q = \frac{\boldsymbol{\zeta} \cdot \nabla \theta}{\rho} with \boldsymbol{\zeta} the absolute , \theta the potential , and \rho the . This conservation follows from Ertel's , which generalizes earlier circulation theorems by considering a conserved scalar like potential temperature and projecting along its in a stratified . The proof relies on the vector identity for the material derivative of the and the adiabatic, inviscid and , ensuring that PV acts as a tracer advected by the flow without sources or sinks under these ideal conditions. In realistic atmospheric and oceanic models, PV is not strictly conserved due to diabatic and frictional processes, such as radiative heating, , and . Cumulus convection, for instance, introduces non-conservation through release during phase changes, which modifies the and thus PV, often leading to generation of positive PV anomalies in the mid-troposphere. These effects are typically parameterized in models using schemes like cumulus parameterizations that account for subgrid-scale heating and momentum redistribution, allowing for explicit diagnosis of PV tendencies from diabatic sources. The PV invertibility principle provides a mathematical framework to reconstruct balanced dynamical fields from a specified PV distribution, given suitable boundary conditions. In the quasi-geostrophic (QG) approximation, invertibility involves solving a linear for the geostrophic streamfunction \psi, such as \nabla^2 \psi = -q, where q incorporates relative , planetary , and stretching terms; this yields the balanced and fields. For global applications, Hoskins, McIntyre, and Robertson (1985) extended this to Ertel's PV on isentropic surfaces, stating that the balanced flow is uniquely determined by the global PV distribution on each surface together with the specification of the total mass (or at the surface) beneath it, enabling the solution via nonlinear elliptic equations that account for the full nonlinear balance. This formulation underpins PV-based diagnostic tools in , emphasizing the sufficiency of PV as a prognostic for balanced motions.

Applications and Extensions

Meteorological Forecasting

In meteorological forecasting, potential vorticity (PV) serves as a key diagnostic tool in operational numerical weather prediction models, such as those from the European Centre for Medium-Range Weather Forecasts (ECMWF) and the (NOAA). These models routinely compute PV fields to identify upper-level features, including the dynamic tropopause, which is defined as the isentropic surface where PV reaches 1.5–2 potential vorticity units (PVU), with 2 PVU commonly used as the threshold separating tropospheric and stratospheric air. This PV-based tropopause definition aids in diagnosing jet streaks, where strong PV gradients highlight regions of intense upper-level winds and potential cyclogenesis. Forecasters at ECMWF and NOAA use these PV contours on pressure or isentropic levels to track tropopause undulations and associated weather systems, enhancing the interpretation of model output for short- to medium-range predictions. PV streamers—elongated filaments of high PV descending from the —play a crucial role in forecasting blocking patterns, where they contribute to quasi-stationary high-pressure systems that disrupt typical westerly flow. These features, often identified in model analyses, signal the onset of persistent anticyclonic blocking over regions like the North Atlantic, leading to prolonged extremes such as outbreaks or domes. Similarly, PV cutoffs, where isolated pockets of stratospheric air become detached, are diagnostic of s, which form closed cyclones detached from the main westerly jet and are associated with heavy and in midlatitudes. Operational forecasters monitor the evolution of these PV structures in predictions to anticipate blocking persistence and cut-off low development, improving lead times for impacts like flooding in or . Case studies illustrate PV's diagnostic value in extreme events; during the , anomalous PV fluxes from eddy activity sustained an upper-level blocking ridge over the continent, amplifying surface temperatures and contributing to over 70,000 excess deaths. In tropical contexts, PV anomalies generated by diabatic heating in convective clusters have been linked to hurricane genesis, as seen in simulations of storms like Gert (1999), where low-level PV maxima from moist initiated vortex spin-up, though the principle extends to post-2000 cases analyzed in operational settings. These examples underscore how PV anomalies diagnose the dynamical precursors to or blocking, guiding forecaster alerts. Post-2000 advances have integrated PV into nowcasting and short-term forecasting tools, enhancing operational efficiency. The adoption of "PV thinking" in U.S. forecast offices since 2007 emphasizes PV diagnostics for real-time assessment of extratropical and coastal storms, using model-derived PV fields to blend observations with predictions up to ahead. Techniques like PV inversion, refined in the 2000s, allow forecasters to isolate PV perturbations and improve track predictions by incorporating localized modifications in numerical models. Additionally, three-dimensional PV tracking methods developed around 2020 enable nowcasting of cutoff life cycles, providing views of PV evolution for sub-daily updates on risks. These tools have been implemented in systems like ECMWF's Integrated System, contributing to forecasts of PV-influenced phenomena.

Oceanographic Dynamics

In oceanography, potential vorticity (PV) is adapted to account for the compressibility of seawater and the influence of salinity on density stratification, differing from atmospheric applications where temperature dominates. The oceanic form of Ertel's PV is defined as q = \frac{1}{\rho} \mathbf{\omega}_a \cdot \nabla \sigma_\theta, where \rho is the fluid density, \mathbf{\omega}_a is the absolute vorticity vector, and \sigma_\theta is the potential density referenced to the sea surface. This formulation incorporates salinity effects, which create baroclinic structures through thermohaline variations, leading to PV distributions that control large-scale currents and mixing unlike the primarily barotropic atmospheric jets. PV conservation governs the dynamics of the (), the strongest oceanic current, where meridional PV gradients drive eddy fluxes that homogenize PV in the interior, enabling the current's zonal extent across varying bottom topography. In subtropical gyres, such as the , curl inputs PV that balances interior dissipation, shaping the Sverdrup interior flow and recirculations along PV contours, with low-PV mode waters forming in the gyre centers due to . At submesoscale resolutions (1–10 km), frontal instabilities, including symmetric instability and frontogenesis, generate intense anomalies that facilitate vertical mixing and restratification, transferring from boundaries to the interior and modulating heat uptake in gyres. These processes enhance diapycnal , with destruction at fronts compensating planetary stretching. Observations from floats enable global mapping by providing temperature and salinity profiles to compute \sigma_\theta and , revealing homogenized in shadow zones and submesoscale variability in the upper . Climate models, such as high-resolution integrations, simulate budgets to assess long-term circulation changes, confirming eddy-driven fluxes as key to Southern Ocean overturning under climate forcing.

References

  1. [1]
    [PDF] Chapter 6 Circulation Theorem and Potential Vorticity
    The vorticity is related to. a quantity called the circulation, defined as the closed line integral of the fluid velocity.
  2. [2]
    [PDF] Chapter 8 Potential Vorticity 8.1 Ertel's theorem
    The vorticity equation describes the vector dynamics of the vorticity in a clear way. It is not, generally, a conservation statement.<|control11|><|separator|>
  3. [3]
    [PDF] Potential Vorticity
    Potential Vorticity. • If we consider atmospheric flow on surfaces of constant potential temperature instead of on constant height.Missing: dynamics | Show results with:dynamics
  4. [4]
    [PDF] Chapter 7: Circulation and Vorticity
    In a barotropic, inviscid fluid the circulation is constant!!! Bjerknes' Circulation theorem. Changes in circulation can arise due to friction or baroclinicity.Missing: geophysical | Show results with:geophysical
  5. [5]
    [PDF] THE BJERKNES' CIRCULATION THEOREM - electronic library -
    This essay is primarily about the Bjerknes circu- lation theorem (Bjerknes 1898). This theorem, formulated by Vilhelm Bjerknes, was published.
  6. [6]
    [PDF] Potential vorticity Michael E. McIntyre
    Synopsis: The significance of the potential vorticity (PV) for atmosphere– ocean dynamics was first explored by Carl-Gustaf Rossby in the 1930s. Re- viewed here ...
  7. [7]
    [PDF] Rossby, Ertel, and potential vorticity RM
    Oct 10, 2003 · ROSSBY (1938) generalized the potential vorticity conservation law to the case of a continuously stratified, hydrostatic fluid, again in the ...
  8. [8]
    (PDF) Rossby, Ertel, and potential vorticity - ResearchGate
    Feb 4, 2016 · ROSSBY (1940) does, however, mark the first appearance of the term "potential vorticity." ... 1933, for example, he published a new proof (ERTEL ...
  9. [9]
    The birth of numerical weather prediction - Tellus B
    The paper describes the major events leading gradually to operational, numerical, short-range predictions for the large-scale atmospheric flow.
  10. [10]
    [PDF] Chapter 4 Vorticity and Potential Vorticity
    4.4 Potential vorticity. We now develop an approximation to shallow water theory for flows in which the Rossby number is very much less than unity. This ...Missing: history | Show results with:history<|control11|><|separator|>
  11. [11]
    [PDF] Chapter 7 Fundamental Theorems: Vorticity and Circulation
    The vorticity is defined as the local rotation or spin of the fluid element about an axis through the element. Figure 7.1.1 shows the distinction. Figure 7.1.1 ...Missing: Bjerknes | Show results with:Bjerknes<|control11|><|separator|>
  12. [12]
    [PDF] Essentials of Atmospheric and Oceanic Dynamics
    If initially the absolute vortic- ity increases monotonically poleward, then the distur- bance will bring fluid with lower absolute vorticity into the cap ...
  13. [13]
    [PDF] (Potential) vorticity: the swirling motion of geophysical fluids
    Aug 22, 2018 · Planetary vorticity is the Coriolis parameter we used earlier to discuss flow in the ocean. It is greatest at the poles where it is twice ...
  14. [14]
    [PDF] a Coriolis tutorial - Woods Hole Oceanographic Institution
    The Coriolis terms that remain are those having the sine of the latitude, and the important combination f = 2Ω sin φ. (45) is dubbed the Coriolis parameter.
  15. [15]
    10.5 Are all the terms in these equations equally important? Let's ...
    The Coriolis Parameter. Define the Coriolis parameter as f≡2Ωsinϕ. At 45o N, f=(2)(7.27×10−5s−1)sin45o∼10−4s−1 . We will use this parameter in the scale ...
  16. [16]
    [PDF] Geophysical Fluid Dynamics II Lecture 5 Spring 2021 Vorticity
    It's the definition of vorticity. Vorticity. Page 9. 9. X i j r ⇥ u = i(. ∂w ... Absolute vorticity = Planetary vorticity + Relative vorticity k · (⇥ ua) ...
  17. [17]
    Vorticity - an overview | ScienceDirect Topics
    Vorticity is defined as a fundamental physical attribute that quantifies the local rotation or angular velocity of individual fluid particles as they move ...
  18. [18]
    [PDF] potential temperature - ACP
    Dec 15, 2020 · The application of the new reference potential tem- perature is discussed for computations of the Brunt–Väisälä frequency, Ertel's potential ...<|control11|><|separator|>
  19. [19]
    On Moist Potential Temperatures and Their Ability to Characterize ...
    Mar 29, 2022 · For dry air, approximated as an ideal gas, the potential temperature θ elegantly describes an air parcel's state. It does so by accounting for ...Terminology and definitions · Moist potential temperatures · Properties of the moist...
  20. [20]
    Potential Temperature - an overview | ScienceDirect Topics
    The rate at which temperature changes as a parcel moves up or down in the atmosphere without heating can be derived by using the hydrostatic equation,. (1.22) ...
  21. [21]
    PLANETARY FLOW PATTERNS IN THE ATMOSPHERE - 1940
    The general purpose of this paper is to discuss the factors which determine the character of the prevailing flow patterns in the atmosphere, ...
  22. [22]
    THE BJERKNES' CIRCULATION THEOREM - AMS Journals
    This theorem, formulated by Vilhelm Bjerknes, was published in a paper of 1898, in the Proceedings of the Royal. Swedish Academy of Sciences in Stockholm. The ...
  23. [23]
    [PDF] English translations of twenty-one of Ertel's papers on geophysical ...
    This collection includes the four classic 1942 papers on vorticity and potential vorticity conservation principles and also papers on generalized ...
  24. [24]
    [PDF] Conservation Laws for Potential Vorticity in a Salty Ocean or Cloudy ...
    The original derivation is approximately a century old, in the work of Rossby and. Ertel, and it is related to the celebrated circulation theorems of Kelvin and ...
  25. [25]
    On Ertel's Potential Vorticity Theorem. On the Impermeability ...
    The PV conservation theorem has been derived in several different ways since Ertel's original work in 1942. The derivation that appears in most text books ...Abstract · Introduction · Kinematical preliminaries · Derivations of the PV theorem
  26. [26]
    [PDF] Chapter 9 Geostrophy, Quasi-Geostrophy and the Potential Vorticity ...
    The governing equation for synoptic scale motions, (9.3.16) is the potential vorticity equation in which each term is evaluated in terms of the geostrophic ...
  27. [27]
    [PDF] 9. Quasi-geostrophic potential vorticity - MIT OpenCourseWare
    the quasi-geostrophic vorticity equation: 2ϕ + β0. ∂ϕ. − f2. ∂ω. − ˆ. Dg∇. 0 kf0∇ × F = 0. (9.3). ∂x. ∂p. The thermodynamic equation, (8.30), is first cast ...
  28. [28]
    [PDF] Quasi-Geostrophic Theory Chapter 4
    Definition: The quasi-geostrophic potential vorticity is defined as q = 1. /0 V. 2. Φ + / +. ∂. ∂p (. /0 σ. ∂Φ. ∂p ). Dgq. Dt. = 0. QG Potential Vorticity ...
  29. [29]
    On the use and significance of isentropic potential vorticity maps
    Hoskins, B. J. and McIntyre, M. E. 1985 On the role of nonlinear radiation, wave propagation and wave breaking in the life cycles of nonlinear baroclinic ...
  30. [30]
    A Breakdown in Potential Vorticity Estimation Delineates the ...
    Aug 10, 2020 · PV is conserved via material advection for adiabatic and inviscid flows (or nearly so for atmospheric and oceanic flows), is invertible in QG ...
  31. [31]
    Non‐conservation and conservation for different formulations of ...
    Apr 11, 2024 · Potential vorticity (PV) of each fluid parcel is known to be not conserved if clouds and phase changes are present.
  32. [32]
    Section 8.1.11 Potential vorticity charts - ECMWF Confluence Wiki
    Nov 12, 2024 · The height of the dynamic tropopause is commonly taken to be the level at which potential vorticity (PV) equals 2.0 PV units. In the troposphere ...
  33. [33]
    [PDF] Dynamic Tropopause and Potential Vorticity: An Introduction
    So, tropopause undulations PV anomalies are not just passively advected through the flow, they instead modify that and surrounding flow. Forecasters may get ...
  34. [34]
    Modification of Potential Vorticity near the Tropopause by ...
    Jun 1, 2019 · The features include the positive PV anomaly on the cyclonic-shear side of the upper-level front–jet system at the downstream flank of the ridge ...
  35. [35]
    Potential vorticity at various pressure levels - ECMWF | Charts
    It is measured in Potential Vorticity Units (PVU, where 1 PVU = m2.K.kg-1.s-1.10E-6). Typically air near the tropopause has a PV value of about 2 PVU; air ...Missing: definition | Show results with:definition
  36. [36]
    Identification and ERA-15 Climatology of Potential Vorticity ...
    A novel approach is introduced to identify potential vorticity (PV) streamers and cutoffs as indicators of Rossby wave breaking near the extratropical ...
  37. [37]
    Climatology of potential vorticity streamers and associated isentropic ...
    Apr 7, 2015 · The climatology of PV streamers reveals a seasonal cycle with a higher frequency in summer than in winter in both hemispheres on all isentropes.2 Data Sets And Methodology · 3 Results · 4 Discussion
  38. [38]
    The three-dimensional life cycles of potential vorticity cutoffs - WCD
    Jun 24, 2021 · The aim of this study is to explore the nature of potential vorticity (PV) cutoff life cycles. While climatological frequencies of such ...
  39. [39]
    Model Simulation and Projection of European Heat Waves in ...
    In an earlier diagnostic study, Illari and Marshall (1983) have also noted the role of eddy fluxes of potential vorticity in maintaining a blocking episode over ...
  40. [40]
    From Atmospheric Waves to Heatwaves: A Waveguide Perspective ...
    ... 2003 European heatwave (e.g., Weisheimer et al. 2011), although others, such ... potential vorticity (PV); PV is a function of both relative and ...<|control11|><|separator|>
  41. [41]
    Tropical Cyclone Evolution via Potential Vorticity Anomalies in a ...
    In the genesis case, as well as for an intensifying tropical storm, convection typically does not occur at the center of the vortex, confirming the importance ...
  42. [42]
    [PDF] simulation and interpretation of the genesis of tropical storm gert
    Jul 23, 2025 · Potential vorticity in convective regions was strongest at low levels while stratiform-region PV was strongest at mid levels, suggesting that ...<|control11|><|separator|>
  43. [43]
    Potential Vorticity (PV) Thinking in Operations
    The PV framework provides human forecasters with a means to evaluate NWP model forecasts in a way that facilitates recognition of when and how value may be ...
  44. [44]
    Improving the numerical prediction of a cyclone in the ...
    Apr 28, 2009 · This study explores the benefit that can be drawn from incorporating local potential vorticity (PV) modifications into a numerical forecast.
  45. [45]
    Potential Vorticity Structure across the Gulf Stream - AMS Journals
    Ertel's potential vorticity q is expressed (Mueller 1995) as. qρ−1ζΩ∇ψ (1) · In stream coordinates the downstream component ∂σθ/∂x is negligible and the first ...Missing: formula | Show results with:formula
  46. [46]
    None
    Summary of each segment:
  47. [47]
    Potential Vorticity Budgets in the North Atlantic Ocean in
    Jan 1, 2014 · Abstract. This paper focuses on potential vorticity (PV) budgets in the North Atlantic with an emphasis on the wind-driven subtropical gyre.
  48. [48]
    Effects of the Submesoscale on the Potential Vorticity Budget of ...
    Sep 1, 2018 · Submesoscale processes thus provide a direct link between small-scale boundary layer turbulence and the gyre-scale circulation, through their ...<|control11|><|separator|>
  49. [49]
    Deep Lagrangian Connectivity in the Global Ocean Inferred from ...
    If the coherent set detected by Argo floats corresponds with the shadow zone, we might then expect potential vorticity to be homogenized therein. To test this, ...
  50. [50]
    The Annual Cycle of Upper-Ocean Potential Vorticity and Its ...
    The evolution of upper-ocean potential vorticity (PV) over a full year in a typical midocean area of the northeast Atlantic is examined