Fact-checked by Grok 2 weeks ago

Kelvin's circulation theorem

Kelvin's circulation theorem, named after (William Thomson, 1824–1907), asserts that in an inviscid, subject to conservative body forces, the circulation of the velocity field around any closed material —meaning a composed of the same particles that advects with the flow—remains constant over time. This theorem, first formulated by in 1815 and independently rediscovered by Kelvin in 1869, provides a key for motion and is mathematically expressed as \frac{D\Gamma}{Dt} = 0, where \Gamma = \oint_C \mathbf{v} \cdot d\mathbf{l} is the circulation and C is the material . The theorem arises from the Euler equations of motion for ideal fluids, where the material derivative of the velocity satisfies \frac{D\mathbf{v}}{Dt} = -\frac{1}{\rho} \nabla p + \mathbf{f}, with \mathbf{f} being a conservative force per unit mass (e.g., , \mathbf{f} = -\nabla \Psi) and the barotropic condition ensuring that p = p(\rho) allows \frac{\nabla p}{\rho} to be expressed as the gradient of a potential. Under these assumptions—no viscosity, density as a function of pressure only, and reversible body forces—the rate of change of circulation vanishes, as shown by applying and the curl of the momentum equation, yielding \frac{d\Gamma}{dt} = \int_S \frac{\nabla \rho \times \nabla p}{\rho^2} \cdot d\mathbf{A} = 0 for barotropic conditions where \nabla \rho \times \nabla p = 0. Lord Kelvin derived this result in his seminal paper "On Vortex Motion," emphasizing its role in understanding vortex persistence in perfect fluids. This principle has profound implications for analysis, implying that if a is initially irrotational (\boldsymbol{\omega} = \nabla \times \mathbf{v} = 0), it remains so, enabling theory for applications like and . It also underpins on vortex lines, which move materially and conserve strength, influencing phenomena such as behind bluff bodies and the stability of geophysical flows. In baroclinic fluids (where \nabla \rho \times \nabla p \neq 0), such as those with heating or salinity gradients, circulation can change, leading to extensions like the Bjerknes circulation theorem in . Despite its idealizations, the theorem remains a for approximating real viscous flows at high Reynolds numbers.

Introduction

Historical Development

The development of Kelvin's circulation theorem emerged within the broader 19th-century advancements in hydrodynamics, particularly the study of vortex motion in inviscid, incompressible fluids. Earlier foundational work by in the late laid groundwork through his , including treatments of fluid equilibrium and motion that influenced subsequent investigations into rotational flows. A circulation conservation principle was first formulated by in 1815 for incompressible inviscid fluids. Building directly on this and on Hermann von Helmholtz's advancements, who in presented seminal theorems on vortex dynamics, demonstrating that vortex lines behave as material lines in ideal fluids—transported and deformed with the flow—and that the circulation along a vortex filament remains constant. These ideas provided a for understanding conserved quantities in fluid rotation, stimulating further exploration of steady, irrotational, and vortical flows. William Thomson, later ennobled as , entered this discourse amid his extensive contributions to physics, including the establishment of the absolute temperature scale and early work in . Motivated by the need to analyze vortex dynamics in ideal fluids—particularly the persistence and evolution of rotational motion in steady inviscid flows—Thomson independently rediscovered and extended the circulation theorem to barotropic flows in his 1869 paper "On Vortex Motion." This work, published in the Transactions of the Royal Society of , emphasized the invariance of circulation around material contours, building on Helmholtz's insights. Thomson's motivation was rooted in resolving questions about the stability and topological conservation of vortices, which he later applied to his "vortex theory of atoms," positing knotted vortex filaments as models for stable atomic structures in an ether-like medium. The theorem's formulation marked a pivotal synthesis of variational principles and Helmholtz's representations of , though circulation as a scalar had roots in Cauchy's earlier work. Kelvin's focused on practical implications for hydrodynamics, such as the behavior of vortex rings and filaments, without delving into modern extensions. This contribution solidified Thomson's role as a bridge between and emerging fluid theories, influencing subsequent geophysical and engineering applications while remaining tied to the ideal fluid assumptions of the time.

Physical Interpretation

Kelvin's circulation theorem provides a fundamental into the rotational dynamics of by characterizing circulation as a measure of the net rotation or enclosed within a closed of fluid particles, often visualized as a deformable . This quantity encapsulates the collective "swirling" motion of the elements bounded by the loop, distinguishing rotational flows from purely translational ones. In essence, it quantifies how much the tends to rotate as a whole, akin to the average flux through the surface spanned by the . The theorem asserts that, in ideal inviscid and barotropic fluids subject only to conservative body forces, this circulation remains invariant as the material contour evolves with the flow. Consequently, fluid parcels preserve their inherent circulatory character, preventing the spontaneous creation or destruction of rotational motion and leading to the long-term persistence of vortices and eddies. This conservation underscores the "frozen-in" nature of vorticity lines within the fluid, where initial rotational structures endure without dissipation or amplification from internal pressure differences. Physically, this invariance parallels the conservation of in rigid body mechanics, but extends to flexible, deformable loops that can elongate, contract, or contort while maintaining their total circulatory strength. Unlike fixed to a central , circuits adapt to the , yet their rotational integrity is safeguarded, highlighting the theorem's role in describing sustained coherent structures in motion. A key implication for flow stability arises from the prohibition of circulation generation in these ideal conditions: no new vorticity can emerge from baroclinic torques or viscous effects, ensuring that smooth, irrotational flows remain free of rotational disturbances. This explains the absence of starting vortices in idealized scenarios, such as the initial motion of a body through an inviscid fluid, where flow remains attached and stable without shedding rotational wakes. Such principles reveal why certain fluid configurations resist the onset of instability, preserving orderly motion until non-ideal effects intervene.

Mathematical Framework

Circulation in Fluid Dynamics

In , circulation serves as a fundamental measure of the rotational motion associated with a , quantifying the net tendency of the to rotate about a closed path. The circulation \Gamma around a closed curve C in the is rigorously defined as the of the \mathbf{u} tangent to the curve: \Gamma = \oint_C \mathbf{u} \cdot d\mathbf{l}, where d\mathbf{l} is the infinitesimal line element along C. This integral captures the cumulative contribution of the fluid velocity components aligned with the path, providing a scalar quantity that reflects the overall circulatory strength enclosed by the curve. By Stokes' theorem, the circulation can be equivalently expressed as the surface integral of the vorticity over any surface S bounded by the curve C: \Gamma = \iint_S (\nabla \times \mathbf{u}) \cdot d\mathbf{A}, where \nabla \times \mathbf{u} is the vorticity vector and d\mathbf{A} is the vector area element normal to S. This relation establishes circulation as the total flux of vorticity through the enclosed surface, linking the macroscopic rotation along the curve to the local rotational tendencies distributed across the area. For a closed that moves and deforms with the —known as a material curve—the temporal evolution of circulation is described by its \frac{D\Gamma}{Dt}, which accounts for both the local of fluid particles and the stretching or contraction of the curve itself. This derivative quantifies how the circulation changes as the curve is advected by the flow, incorporating contributions from the field's variations along and across the path. In two-dimensional (2D) flows, where the velocity varies only in the plane and vorticity reduces to a scalar \omega = \frac{\partial u_y}{\partial x} - \frac{\partial u_x}{\partial y} perpendicular to that plane, circulation simplifies to \Gamma = \iint_S \omega \, dA, emphasizing its role as the integrated scalar vorticity over the area. In contrast, three-dimensional (3D) flows feature a vector vorticity \boldsymbol{\omega} = \nabla \times \mathbf{u}, making circulation the flux of this vector through the surface, which allows for more complex orientations and distributions of rotation not confined to a single direction. This vectorial nature in 3D enables circulation to probe rotations in arbitrary planes, whereas 2D treatments inherently scalarize the concept for planar motions.

Formal Statement of the Theorem

Kelvin's circulation theorem asserts that, in an inviscid () fluid that is barotropic and subject only to conservative body forces, the circulation around any closed material —meaning a that moves and deforms with the fluid—remains constant over time. The circulation \Gamma is defined as the \Gamma = \oint_C \mathbf{v} \cdot d\mathbf{l}, where \mathbf{v} is the fluid velocity and the integral is taken around the closed C. Mathematically, the theorem is expressed as \frac{D\Gamma}{Dt} = 0, where \frac{D}{Dt} denotes the material derivative following the fluid motion, indicating that \Gamma is conserved for the material curve. The key conditions for the theorem's validity include barotropicity, where pressure p is a function solely of density \rho (i.e., p = p(\rho)), ensuring surfaces of constant pressure and density coincide; the absence of viscosity or frictional effects; and body forces that are conservative, such as gravity, which can be derived from a scalar potential. These conditions must hold along the material curve for the conservation to apply. The theorem applies to any co-moving closed loop in three-dimensional flows of barotropic fluids, encompassing both incompressible cases (constant \rho) and compressible cases where the barotropic relation is satisfied, provided the fluid remains inviscid and the forces are conservative.

Derivation and Proof

Key Assumptions

Kelvin's circulation theorem, which states that the material derivative of the circulation around a closed material curve is zero (D\Gamma/Dt = 0), relies on several fundamental assumptions about the fluid and its motion. The theorem assumes an inviscid flow, where viscous effects are neglected, corresponding to zero viscosity (\mu = 0 or \nu = 0) in the Navier-Stokes equations. This eliminates diffusive terms that would otherwise transport vorticity across fluid elements, allowing circulation to remain conserved along material paths. A key prerequisite is the barotropic condition, under which density is a solely of (\rho = \rho(p)), meaning isobaric and isosteric surfaces coincide. This enables the term in the momentum equation to be expressed as the of a potential, preventing baroclinic torques that would generate from misaligned and s. The fails in baroclinic flows, where such gradients do not align, leading to non-zero solenoidal contributions to circulation change. Body forces must be conservative, derivable from a (e.g., as \mathbf{F} = -\nabla \Phi), ensuring their around any closed curve vanishes. Non-conservative forces, such as the in rotating frames, violate this and introduce changes in circulation. The fluid is idealized, with no of or scalars, and perfect slip at boundaries, implying no generation at surfaces. The theorem does not hold in the presence of shocks, where discontinuities violate the smooth flow assumption, or with multi-valued pressures in unsteady or separated flows. Additionally, it assumes a simply connected to avoid complications from multiply connected regions, such as those enclosing obstacles, where circulation may not be uniquely defined or conserved for all material curves.

Detailed Derivation

The derivation of Kelvin's circulation theorem begins with the definition of circulation \Gamma around a closed material curve C(t) in the fluid, given by the line integral \Gamma = \oint_{C(t)} \mathbf{u} \cdot d\mathbf{l}, where \mathbf{u} is the fluid velocity and d\mathbf{l} is the infinitesimal line element along the curve. To find the rate of change of circulation following the fluid motion, the material derivative is applied: \frac{D\Gamma}{Dt} = \frac{D}{Dt} \oint_{C(t)} \mathbf{u} \cdot d\mathbf{l}. The Reynolds transport theorem for line integrals along a material curve yields \frac{D\Gamma}{Dt} = \oint_{C} \frac{D\mathbf{u}}{Dt} \cdot d\mathbf{l}, where the contribution from the deformation of the line element integrates to zero over the closed loop. Substituting the material derivative of the velocity from the Euler equation for an inviscid fluid, \frac{D\mathbf{u}}{Dt} = -\frac{1}{\rho} \nabla p + \mathbf{g}, where \rho is the , p is the , and \mathbf{g} is the per unit (assumed conservative), gives \frac{D\Gamma}{Dt} = \oint_{C} \left( -\frac{1}{\rho} \nabla p + \mathbf{g} \right) \cdot d\mathbf{l} = -\oint_{C} \frac{\nabla p}{\rho} \cdot d\mathbf{l} + \oint_{C} \mathbf{g} \cdot d\mathbf{l}. Since \mathbf{g} = -\nabla \Phi for a conservative potential \Phi, the \oint_{C} \mathbf{g} \cdot d\mathbf{l} = -\oint_{C} \nabla \Phi \cdot d\mathbf{l} = 0 over any closed curve. For the pressure term, under the barotropic assumption where density is a function of pressure alone (\rho = \rho(p)), the expression \frac{\nabla p}{\rho} is the gradient of a scalar potential: \frac{\nabla p}{\rho} = \nabla \int \frac{dp}{\rho(p)}. Thus, \oint_{C} \frac{\nabla p}{\rho} \cdot d\mathbf{l} = \oint_{C} \nabla \left( \int \frac{dp}{\rho} \right) \cdot d\mathbf{l} = 0 for a closed curve in a simply connected domain. This leaves \frac{D\Gamma}{Dt} = 0, proving that the circulation is conserved along the material curve. An alternative expansion explicitly accounts for the convective term before substitution. The material derivative expands as \frac{D}{Dt} \oint_{C} \mathbf{u} \cdot d\mathbf{l} = \oint_{C} \frac{\partial \mathbf{u}}{\partial t} \cdot d\mathbf{l} + \oint_{C} (\mathbf{u} \cdot \nabla) \mathbf{u} \cdot d\mathbf{l}, where the line element evolution contributes to the second term. Using the vector identity (\mathbf{u} \cdot \nabla) \mathbf{u} = \nabla \left( \frac{u^2}{2} \right) - \mathbf{u} \times (\nabla \times \mathbf{u}), with vorticity \boldsymbol{\omega} = \nabla \times \mathbf{u}, the convective integral becomes \oint_{C} \nabla \left( \frac{u^2}{2} \right) \cdot d\mathbf{l} - \oint_{C} (\mathbf{u} \times \boldsymbol{\omega}) \cdot d\mathbf{l}. The first part vanishes as a closed gradient integral, while the second, by Stokes' theorem, equals \iint_{S} \nabla \times (\mathbf{u} \times \boldsymbol{\omega}) \cdot d\mathbf{A} over a surface S bounded by C. However, combining with the local acceleration term from Euler's equation recovers the earlier form, confirming \frac{D\Gamma}{Dt} = 0 under the stated assumptions. This completes the proof, originally formulated by William Thomson in 1869.

Implications and Applications

In Irrotational and Potential Flows

In irrotational flows, Kelvin's circulation theorem implies that if the fluid is initially irrotational—meaning the circulation \Gamma = 0 around any closed material contour—then the flow remains irrotational for all subsequent times under the theorem's assumptions of inviscid, barotropic conditions and conservative body forces. This conservation of zero circulation ensures that the \boldsymbol{\omega} = \nabla \times \mathbf{u} = \mathbf{0} everywhere, allowing the velocity field \mathbf{u} to be expressed as the of a scalar velocity \phi, such that \mathbf{u} = \nabla \phi. The existence of this simplifies the governing equations to \nabla^2 \phi = 0 for incompressible flows, facilitating analytical solutions in theory. A key engineering application arises in the analysis of lift on airfoils in inviscid steady flow. According to the Kutta-Joukowski theorem, the lift force L per unit span on a two-dimensional airfoil is given by L = \rho_\infty U_\infty \Gamma, where \rho_\infty is the freestream density, U_\infty is the freestream speed, and \Gamma is the circulation around the airfoil. Kelvin's theorem ensures that once circulation is established, it remains constant, but in potential flow, the value of \Gamma is uniquely determined by the Kutta condition at the trailing edge, which requires finite and equal velocities on both sides of the sharp edge to model smooth flow departure. This condition prevents singularities in the inviscid solution and links the bound circulation on the airfoil to an equal and opposite starting vortex shed into the wake during flow initiation, maintaining overall circulation conservation. In steady irrotational flows past bodies such as or , the theorem's conservation of circulation enables the in theory. Complex potential functions can be constructed by adding uniform flow to singularity distributions (e.g., sources, sinks, or dipoles) that satisfy boundary conditions on the body surface, with any added circulatory component remaining fixed due to the theorem. For a circular , the non-circulatory solution yields zero lift (), but including a constant circulation term produces the , where lift is perpendicular to the oncoming flow. Similarly, for a , the superposition predicts no drag but allows for circulatory modifications in rotating cases, all while preserving the irrotational nature. Regarding boundary conditions, Kelvin's theorem indicates that no circulation is generated at smooth, impermeable boundaries in inviscid flows, as the tangential velocity is continuous and the normal component satisfies impermeability without introducing vorticity. This contrasts with real viscous flows, where the no-slip condition at walls creates boundary layers that can generate starting vortices during transient startup, shedding circulation into the flow despite the theorem's prediction of conservation in the inviscid limit. In practice, viscosity enforces the Kutta condition at airfoil trailing edges but introduces dissipative effects absent in pure potential flow models.

In Geophysical Fluid Dynamics

Kelvin's circulation theorem plays a crucial role in , particularly in understanding the evolution of large-scale flows in the atmosphere and oceans under approximately inviscid and barotropic conditions. In these systems, the theorem implies that circulation around a closed contour remains conserved in the absence of non-conservative forces, providing a foundation for analyzing balanced motions where the Coriolis effect dominates. This conservation principle helps explain how initial imbalances in rotating fluids adjust to geostrophic equilibrium through the propagation of inertial-gravity waves, during which the circulation is preserved, leading to the establishment of balanced states in atmospheric and disturbances. A key application arises in vorticity dynamics, where the theorem connects to the conservation of in shallow water approximations, which are widely used to model systems such as mid-latitude cyclones. In these approximations, the theorem's invariance ensures that , defined as the vertical component of divided by fluid depth, is materially conserved, influencing the intensification and propagation of Rossby waves and synoptic-scale features in the atmosphere. This linkage underscores how Kelvin's theorem underpins the predictability of patterns by maintaining the integrity of fields over time scales relevant to . In contexts, the theorem accounts for the persistence of mesoscale eddies and jets, such as those observed in the , where approximate inviscid barotropic conditions allow circulation to remain nearly conserved, sustaining coherent structures against dissipation. These features, with scales of 10–100 km, owe their longevity to the theorem's implications, as weak friction and baroclinicity minimally disrupt the circulation balance, enabling eddies to transport heat and momentum across ocean basins. However, in real geophysical fluids, the theorem's validity is approximate due to inherent baroclinicity and ; nonetheless, it remains instrumental in numerical modeling of cyclones and western boundary currents like the , where simplified inviscid assumptions yield accurate representations of circulation-driven dynamics.

Poincaré–Bjerknes Circulation Theorem

The Poincaré–Bjerknes circulation theorem extends Kelvin's circulation theorem to fluid motion in a , accounting for the effects of planetary rotation such as in Earth's atmosphere and oceans. Formulated initially by in 1893 and developed further by in 1898, it addresses the influence of non-conservative forces like the Coriolis effect on circulation dynamics. In a barotropic, inviscid within the rotating frame, the states that the absolute circulation—the sum of the fluid-relative circulation \Gamma = \oint_C \mathbf{u} \cdot d\mathbf{l} and the planetary circulation $2 \mathbf{\Omega} \cdot \mathbf{A}, where \mathbf{\Omega} is the vector of the frame and \mathbf{A} is the oriented area enclosed by the material curve C—is materially conserved: \frac{D}{Dt} \left( \Gamma + 2 \mathbf{\Omega} \cdot \mathbf{A} \right) = 0. This conservation arises from the material derivative of the circulation equation in the rotating frame, where the Coriolis acceleration -2 \mathbf{\Omega} \times \mathbf{u} contributes a term equivalent to -2 \mathbf{\Omega} \cdot \frac{D\mathbf{A}}{Dt} to the rate of change of relative circulation \frac{D\Gamma}{Dt}. The key addition relative to is the planetary term involving $2\mathbf{\Omega}, often described in historical contexts as a "solenoideal" contribution from the of the reference frame, which modifies circulation evolution through the changing of the . For baroclinic fluids, an additional solenoidal term \iint_S \frac{1}{\rho^2} (\nabla \rho \times \nabla p) \cdot d\mathbf{A} accounts for due to misaligned and surfaces, but the rotational extension emphasizes the Coriolis-induced planetary effects. This theorem underpins the conservation of absolute circulation in geophysical contexts, enabling explanations of large-scale atmospheric and oceanic phenomena where rotation is dominant. For instance, it elucidates the meridional propagation of Rossby waves, where northward-moving air parcels experience a decrease in relative vorticity compensated by an increase in planetary vorticity to maintain absolute circulation constancy. Similarly, in tropical cyclones, the theorem highlights how rotational effects amplify cyclonic circulation through interactions with the varying Coriolis parameter. In contrast to Kelvin's theorem, which asserts \frac{D\Gamma}{Dt} = 0 for non-rotating, barotropic, inviscid flows, the Poincaré–Bjerknes version incorporates source terms from and, when applicable, baroclinicity; it recovers Kelvin's result precisely when \mathbf{\Omega} = 0 and the fluid is barotropic.

Generalizations to Baroclinic and Viscous Fluids

Kelvin's circulation theorem, originally stated for barotropic and inviscid fluids, requires modification when these assumptions are relaxed. In baroclinic fluids, where \rho is not a solely of p (i.e., isobars and isopycnals do not coincide), circulation is no longer conserved due to a solenoidal arising from the misalignment of and gradients. The of circulation \Gamma becomes \frac{D\Gamma}{Dt} = \iint_S \frac{\nabla \rho \times \nabla p}{\rho^2} \cdot d\mathbf{A}, where the surface integral vanishes only in the barotropic limit since \nabla \rho \parallel \nabla p. This baroclinic generation of circulation is crucial in stratified flows, such as those in the atmosphere and oceans, where it drives phenomena like frontogenesis. For viscous fluids, the Navier-Stokes equations introduce a diffusive term that erodes circulation over time. The theorem generalizes to \frac{D\Gamma}{Dt} = \iint_S \nu \nabla^2 \boldsymbol{\omega} \cdot d\mathbf{A}, where \nu is the kinematic , \boldsymbol{\omega} = \nabla \times \mathbf{u} is the , and the integral is over the surface S enclosed by the material curve (for constant \nu); this term typically leads to decay of \Gamma as vorticity diffuses across the material contour. In low-viscosity regimes, such as high-Reynolds-number flows, this effect is small, but it becomes significant near boundaries or in turbulent . In (CFD), these generalizations inform numerical schemes for high-Reynolds-number simulations, where inviscid approximations invoke circulation conservation to model large-scale dynamics while incorporating viscous boundary layers or baroclinic effects through subgrid models. Circulation-preserving discretizations, such as those based on simplicial complexes, ensure by mimicking Kelvin's theorem in the inviscid limit. For instance, in or geophysical simulations, high-Re flows approximate barotropic inviscid behavior away from shear layers, with viscous and baroclinic terms added for accuracy. A key modern extension ties Kelvin's theorem to Ertel's potential vorticity theorem, which provides a three-dimensional generalization for stratified, baroclinic flows. Ertel's theorem conserves q = \frac{(\boldsymbol{\omega}_a \cdot \nabla \theta)}{\rho} (with \boldsymbol{\omega}_a the absolute and \theta a conserved scalar like potential temperature) along material trajectories in inviscid, adiabatic conditions, effectively extending circulation conservation to surfaces of constant \theta where baroclinic torques vanish. This connection underpins applications in for diagnosing vortex dynamics in rotating, stratified environments.

References

  1. [1]
    Kelvin Circulation Theorem - Richard Fitzpatrick
    According to the Kelvin circulation theorem, which is named after Lord Kelvin (1824-1907), the circulation around any co-moving loop in an inviscid fluid is ...
  2. [2]
    Kelvins Circulation Theorem - an overview | ScienceDirect Topics
    Kelvin's circulation theorem is defined as the principle that for a barotropic fluid with no frictional forces, the absolute circulation remains conserved ...Missing: original | Show results with:original
  3. [3]
    Classic and Historical Papers Papers on Geophysical Fluid Dynamics
    The Circulation Theorem. The original circulation theorem, valid in a barotropic fluid, is due to Kelvin. If a fluid is baroclinic (i.e., surfaces density and ...The Coriolis Force · Ekman Layers · Potential Vorticity EtcMissing: explanation | Show results with:explanation
  4. [4]
    Kelvin's Theorem - an overview | ScienceDirect Topics
    Kelvin's theorem is defined as the principle stating that the rate of change of circulation around a closed curve comprising the same fluid element is zero ...Missing: original | Show results with:original
  5. [5]
    [PDF] VORTEX DYNAMICS: THE LEGACY OF HELMHOLTZ AND KELVIN
    Kelvin conceived his “Vortex theory of Atoms” (1867–1875) on the basis that, since vortex lines are frozen in the flow of an ideal fluid, their topology should ...
  6. [6]
    [PDF] LESSON 05A: KELVIN'S CIRCULATION THEOREM JM Cimbala
    • Define and discuss Kelvin's circulation theorem for a material contour. • Discuss the physical significance of Kelvin's circulation theorem. Circulation and ...Missing: interpretation | Show results with:interpretation
  7. [7]
    [PDF] 1214.2.K.pdf - Caltech PMA
    The qualitative content of Kelvin's theorem is that vorticity in a fluid is long-lived. A fluid's vorticity and circulation (or lack thereof) will persist, ...
  8. [8]
    [PDF] Lecture 4: Circulation and Vorticity - UCI ESS
    Circulation is the total “push” you get when going along a path, such as a circle. The circulation theorem is obtained by taking the line integral of Newton's ...
  9. [9]
    [PDF] Lecture 3: Circulation and Vorticity
    Here we introduce the basic concepts. Circulation C is defined as the line integral about a closed contour within a fluid of the local velocity of elements. C ...
  10. [10]
    [PDF] Circulation and Vorticity Atmos 5110 Synoptic–Dynamic Meteorology I
    Circulation Mathematical definition: The line integral about a contour of the component of the velocity vector that is locally tangent to the contour. Although ...
  11. [11]
    [PDF] Circulation and Vorticity
    Kelvin's circulation theorem states that circulation is constant in a barotropic fluid. Can we expand this result to apply for less restrictive conditions ...
  12. [12]
    [PDF] 3 Vorticity, Circulation and Potential Vorticity. - Staff
    where we have made use of Stokes' theorem. The circulation around the loop can also be approximated as the mean tangential velocity times the length of the ...
  13. [13]
    [PDF] Well-posedness for the 2D Vorticity Equation with Measure-Valued ...
    Feb 18, 2019 · Note, in 2D, the curl is a scalar quantity, and the equation for the vorticity is quite simple. In fact, it looks almost like a linear ...
  14. [14]
    [PDF] Chapter 4 Vorticity and Potential Vorticity
    4.1 Circulation and vorticity. The vorticity of a fluid flow is defined as the curl of the velocity field: ζ ≡∇× v. (4.1)<|control11|><|separator|>
  15. [15]
    [PDF] Vortex Dynamics - Department of Mathematics & Statistics
    A vortex is typically defined by a region in the fluid of concentrated vorticity. A simple model is a point vortex in 2D flow, which corresponds to a straight.
  16. [16]
  17. [17]
    [PDF] Chapter 7 Fundamental Theorems: Vorticity and Circulation
    In Geophysical Fluid Dynamics, especially the study of the atmosphere and the ... For a barotropic, inviscid fluid Kelvin's theorem then becomes,. dΓ dt.
  18. [18]
    [PDF] Lecture 13
    The circulation of a vortex tube is constant in time. All of these follow from Kelvin's theorem. The first theorem is the most far-reaching. It says that vortex ...
  19. [19]
  20. [20]
    [PDF] On Vortex Motion. By Sir W. THOMSON.
    The mathematical work of the present paper has been performed to illus- trate the hypothesis, that space is continuously occupied by an incompressible.Missing: Kelvin | Show results with:Kelvin
  21. [21]
    [PDF] 2.25 9 Vorticity and Circulation - MIT
    9.4 Three vortex theorems for inviscid, barotropic flow in an irrotational force field. (corollaries of Kelvin's theorem):. (a) Vortex lines move with the fluid ...Missing: implications | Show results with:implications
  22. [22]
    Potential Flow Theory – Introduction to Aerospace Flight Vehicles
    A potential flow assumes the aggregate of an incompressible, irrotational, and inviscid fluid motion, ie, which is called an “ideal” flow.Missing: implications | Show results with:implications<|control11|><|separator|>
  23. [23]
    [PDF] Incomp Flow over Airfoils.key - UTRGV Faculty Web
    Kelvin's Circulation Theorem: Time rate of change of circulation around ... Lift can then be calculated using the Kutta-Joukowski theorem. Page 24. The ...
  24. [24]
    Classic Airfoil Theory – Introduction to Aerospace Flight Vehicles
    The Kutta-Joukowski theorem has already been explained for determining lift; however, lift can also be obtained by integrating the pressure around the airfoil ...
  25. [25]
  26. [26]
    [PDF] 6 Fundamental Theorems: Vorticity and Circulation - UBC EOAS
    3 and 4 are familiar from Kelvin's circulation theorem, but 1 and 2 require more thought. We will state them now and derive an interpretation immediately after.<|control11|><|separator|>
  27. [27]
    [PDF] THE BJERKNES' CIRCULATION THEOREM - electronic library -
    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 ...
  28. [28]
    [PDF] Stable, Circulation-Preserving, Simplicial Fluids
    The domain may have non-trivial topology, e.g., it can contain tunnels and voids (3D) or holes (2D), but is assumed to be compact. To ensure good numerical ...