Fact-checked by Grok 2 weeks ago

Enstrophy

Enstrophy is a fundamental quantity in fluid dynamics that quantifies the intensity of rotational motion within a fluid, defined as half the square of the L² norm of the vorticity field, or equivalently, the volume integral of half the squared vorticity magnitude. This measure arises naturally in the study of incompressible flows, where vorticity \boldsymbol{\omega} = \nabla \times \mathbf{u} represents the local curl of the velocity field \mathbf{u}, and enstrophy Z = \frac{1}{2} \int_V |\boldsymbol{\omega}|^2 \, dV provides insight into the distribution and amplification of vortices. Unlike kinetic energy, which focuses on translational motion, enstrophy emphasizes the rotational components essential to turbulent structures. In two-dimensional incompressible , enstrophy conservation in inviscid flows—alongside —leads to a distinctive dual-cascade : cascades upscale to form large coherent structures like vortices, while enstrophy cascades downscale toward dissipative scales, enabling a forward transfer of rotational intensity without accumulation at small scales. This behavior, originally proposed by Kraichnan (1967), contrasts sharply with three-dimensional , where vortex stretching prevents enstrophy conservation and allows to dissipate directly at small scales. The enstrophy cascade in flows predicts a logarithmic correction to the at high wavenumbers, E(k) \sim \eta^{2/3} k^{-3} [\ln (k/k_l)]^{-1/3}, where \eta is the enstrophy dissipation rate and k_l is the lower wavenumber bound of the range, influencing models of atmospheric and . Beyond theoretical turbulence, enstrophy is pivotal in numerical simulations and analyses of systems, serving as a diagnostic for dissipation mechanisms and a in structure-preserving schemes for solving the Navier-Stokes equations. For instance, in direct numerical simulations of homogeneous , enstrophy balances reveal the interplay between production at large scales and viscous destruction at small scales, aiding predictions of and laws. In geophysical applications, such as stratified or rotating flows, variants like potential enstrophy further extend its utility to quasi-geostrophic models, conserving angular momentum proxies in planetary atmospheres.

Definition and Fundamentals

Core Definition

Enstrophy is a scalar quantity that quantifies the intensity of rotation in a fluid flow, serving as an aggregate measure of the vorticity field across the entire domain. Vorticity, defined as the curl of the velocity field, represents the local rotation rate of fluid elements, and enstrophy captures the overall "vortational content" by integrating the square of its magnitude. Mathematically, enstrophy Z is expressed as the volume integral Z = \frac{1}{2} \int_V |\boldsymbol{\omega}|^2 \, dV, where \boldsymbol{\omega} is the vector and V denotes the volume. This formulation emphasizes enstrophy as a global diagnostic, analogous to but focused on al rather than translational motion; it effectively measures the total squared rates, providing insight into the flow's vortical structure without regard to the direction of . In SI units, the local enstrophy density \frac{1}{2} |\boldsymbol{\omega}|^2 has dimensions of s^{-2}, reflecting the squared frequency of rotation, while the total enstrophy incorporates and thus has dimensions of m^3 s^{-2}. In non-dimensional analyses, such as those involving Reynolds or Rossby numbers, enstrophy is often normalized to be dimensionless, facilitating comparisons across scales and flow regimes. The integral of squared vorticity, now known as enstrophy, became central in studies, particularly in two-dimensional flows where it is conserved in inviscid conditions, as analyzed in Kraichnan's work on inertial ranges. The term "enstrophy" was coined around 1968 by C.E. , from the Greek words for "" and "energy," referring to the mean-square .

Relation to Vorticity

, denoted as \boldsymbol{\omega}, is defined as the of the \mathbf{u} in , mathematically expressed as \boldsymbol{\omega} = \nabla \times \mathbf{u}. This captures the rotational component of the motion at each point in the flow. Physically, represents the local rate of elements, where the magnitude |\boldsymbol{\omega}| corresponds to twice the of for a . Thus, |\boldsymbol{\omega}|/2 quantifies the speed, rotational motion from pure or deformation. In Cartesian coordinates, the components of are given by \omega_x = \frac{\partial u_z}{\partial y} - \frac{\partial u_y}{\partial z}, \omega_y = \frac{\partial u_x}{\partial z} - \frac{\partial u_z}{\partial x}, and \omega_z = \frac{\partial u_y}{\partial x} - \frac{\partial u_x}{\partial y}, illustrating how spatial variations in gradients produce . Enstrophy depends directly on , as it is proportional to the volume of the squared magnitude of the field, Z \propto \int \boldsymbol{\omega} \cdot \boldsymbol{\omega} \, dV. This relation underscores enstrophy's role in measuring the cumulative intensity of rotational motion across the domain, contrasting with quantities like that emphasize linear . Enstrophy can thus be interpreted as the L^2 norm of the , providing a scalar aggregate of rotational strength.

Mathematical Formulation

Two-Dimensional Case

In two-dimensional incompressible flows, the vorticity \omega is a scalar quantity directed perpendicular to the plane of motion, arising from the curl of the velocity field \mathbf{u} = (u, v) as \omega = \partial u / \partial y - \partial v / \partial x. This scalar nature simplifies the analysis compared to three-dimensional cases, where vorticity is a vector. The enstrophy Z, a measure of the total squared vorticity, is defined as Z = \frac{1}{2} \int_\Omega \omega^2 \, dA, where \Omega denotes the flow domain and dA is the area element. This integral quantifies the overall intensity of rotational motion in the fluid. For ideal (inviscid) two-dimensional incompressible flows, the Navier-Stokes equations reduce to the Euler equations, leading to the transport equation \frac{\partial \omega}{\partial t} + \mathbf{u} \cdot \nabla \omega = 0. This form implies that is conserved along fluid particle trajectories. To obtain the enstrophy evolution, multiply the transport equation by \omega and integrate over \Omega: \int_\Omega \omega \left( \frac{\partial \omega}{\partial t} + \mathbf{u} \cdot \nabla \omega \right) dA = 0. The first term yields \frac{1}{2} \frac{d}{dt} \int_\Omega \omega^2 \, dA = \frac{dZ}{dt}, while the second term, using incompressibility \nabla \cdot \mathbf{u} = 0 and suitable boundary conditions (e.g., periodic or no-flux), integrates to zero via the . Thus, \frac{dZ}{dt} = 0, demonstrating strict conservation of enstrophy in inviscid 2D flows. In the presence of viscosity, the full two-dimensional Navier-Stokes equations yield the vorticity equation \frac{\partial \omega}{\partial t} + \mathbf{u} \cdot \nabla \omega = \nu \nabla^2 \omega, where \nu > 0 is the kinematic . Repeating the by \omega and procedure gives \frac{dZ}{dt} = \nu \int_\Omega \omega \nabla^2 \omega \, dA. , assuming boundary terms vanish, transforms this to \frac{dZ}{dt} = -\nu \int_\Omega |\nabla \omega|^2 \, dA \leq 0. The negative definite integrand highlights that enstrophy undergoes monotonic dissipation through viscous diffusion of gradients, with no mechanism present in two dimensions—unlike higher dimensions where vortex stretching can amplify enstrophy. The conservation of in ideal flows underpins key dynamical features of two-dimensional . Specifically, it enables a forward of enstrophy to smaller scales (higher wavenumbers), characterized by a k^{-3} in the inertial , while cascades inversely to larger scales. This -cascade phenomenology, first predicted theoretically, distinguishes from its three-dimensional counterpart and arises directly from the inviscid invariants of and enstrophy.

Three-Dimensional Case

In three-dimensional incompressible flows, enstrophy is defined as Z = \frac{1}{2} \int_V |\boldsymbol{\omega}|^2 \, dV, where \boldsymbol{\omega} = \nabla \times \mathbf{u} is the vector and V denotes the flow domain volume. This quantity measures the overall intensity of rotation in the fluid, with the factor of \frac{1}{2} chosen for convenience in derivations. Unlike the two-dimensional case, where enstrophy is conserved for flows, the three-dimensional vector structure of permits amplification through nonlinear interactions. For ideal (inviscid) incompressible flows governed by the Euler equations, the time rate of change of enstrophy follows from the vorticity transport equation \frac{D \boldsymbol{\omega}}{Dt} = (\boldsymbol{\omega} \cdot \nabla) \mathbf{u}, yielding \frac{dZ}{dt} = \int_V \boldsymbol{\omega} \cdot (\boldsymbol{\omega} \cdot \nabla) \mathbf{u} \, dV. The integrand \boldsymbol{\omega} \cdot (\boldsymbol{\omega} \cdot \nabla) \mathbf{u} captures the vortex stretching effect, where velocity gradients tilt and elongate vorticity lines, potentially increasing local vorticity magnitude when the alignment is favorable. This term is absent in two dimensions due to the scalar nature of vorticity there, highlighting a key distinction in flow dynamics. The integral can be positive overall, allowing unbounded enstrophy growth in finite time for some initial conditions, though global regularity remains an open question for the Euler equations. In the presence of viscosity, as described by the incompressible Navier-Stokes equations, the vorticity equation becomes \frac{D \boldsymbol{\omega}}{Dt} = (\boldsymbol{\omega} \cdot \nabla) \mathbf{u} + \nu \Delta \boldsymbol{\omega}, leading to the enstrophy evolution \frac{dZ}{dt} = \int_V \boldsymbol{\omega} \cdot (\boldsymbol{\omega} \cdot \nabla) \mathbf{u} \, dV - \nu \int_V |\nabla \boldsymbol{\omega}|^2 \, dV + \text{boundary terms}. The second integral arises from integration by parts of the viscous contribution \nu \int_V \boldsymbol{\omega} \cdot \Delta \boldsymbol{\omega} \, dV and is non-positive, representing irreversible dissipation that smooths small-scale vorticity structures. Boundary terms, such as surface integrals involving vorticity flux, vanish under periodic boundary conditions or decay sufficiently fast at infinity in whole-space problems. Positive contributions from vortex stretching can thus amplify enstrophy despite dissipation, promoting the cascade of energy to smaller scales and intensifying turbulence; for instance, numerical studies show enstrophy growth rates scaling with initial conditions in high-Reynolds-number flows. This amplification via has profound implications for flow stability, as sustained positive production can lead to rapid fine-scale development and potential formation in the inviscid limit, though regularizes solutions for smooth initial data. In turbulent regimes, the balance between stretching-induced production and viscous governs the , with stretching dominating at large Reynolds numbers to drive multiscale dynamics.

Physical Properties and Dynamics

Conservation in Ideal Flows

In ideal two-dimensional (2D) incompressible flows governed by the Euler equations, enstrophy is strictly conserved due to the absence of vortex stretching mechanisms. The time derivative of the total enstrophy Z = \frac{1}{2} \int |\boldsymbol{\omega}|^2 \, dV vanishes, yielding \frac{dZ}{dt} = 0, as the vorticity transport equation lacks terms that amplify vorticity magnitude. This conservation arises fundamentally from the two-dimensional geometry, where vorticity lines cannot stretch or tilt in the third dimension, preserving the L^2-norm of vorticity under advective dynamics. In contrast, ideal three-dimensional (3D) incompressible flows exhibit non-conservation of enstrophy owing to vortex stretching and tilting effects inherent in the Euler equations. The rate of change of enstrophy is given by \frac{dZ}{dt} = \int \omega_i \omega_j \frac{\partial u_i}{\partial x_j} \, dV, where \boldsymbol{\omega} is the vorticity vector and \mathbf{u} is the velocity field; this integral captures the alignment between vorticity and the velocity gradient tensor, leading to amplification or reorientation of vorticity. Positive contributions from stretching dominate in turbulent-like states, resulting in net enstrophy production, while incompressibility ensures no dilatation term appears. This dimensional distinction underpins the selective decay hypothesis in relaxed turbulent states, where enstrophy minimizes subject to conserved quantities like , particularly in flows approaching coherent structures such as monopoles or dipoles. In such scenarios, the hypothesis posits that dissipation selectively reduces enstrophy more rapidly than , driving the system toward minimum-enstrophy equilibria that stabilize large-scale features. Enstrophy conservation in ideal flows facilitates an inverse , transferring upscale while enstrophy cascades forward to small scales for , inverting the direct observed in flows where enstrophy is not conserved. This highlights how dynamics suppress the vortex stretching that enables efficient energy at small scales in .

Production and Dissipation Mechanisms

In viscous fluid flows, enstrophy production arises primarily from the nonlinear interactions of vorticity with the velocity gradient tensor, particularly through vortex stretching and tilting mechanisms in three-dimensional (3D) flows. The production term in the enstrophy evolution equation is given by the volume integral \int_V \boldsymbol{\omega} \cdot (\boldsymbol{\omega} \cdot \nabla) \mathbf{u} \, dV, where \boldsymbol{\omega} is the vorticity vector and \mathbf{u} is the velocity field; this term is generally positive in 3D turbulence due to the amplification of vorticity magnitudes by stretching along principal strain directions. Tilting contributes by reorienting vorticity vectors, further enhancing local enstrophy growth, though stretching dominates the net positive contribution. In contrast, this production mechanism is negligible in two-dimensional (2D) flows, as the vorticity is perpendicular to the velocity plane, rendering the stretching term zero. Enstrophy dissipation occurs through viscous effects that convert rotational kinetic energy into heat at small scales, represented by the negative definite term -\nu \int_V |\nabla \boldsymbol{\omega}|^2 \, dV, where \nu is the kinematic ; this integral quantifies the irreversible destruction of enstrophy via gradients in the field. The process is always dissipative, with the magnitude increasing at higher Reynolds numbers due to steeper small-scale gradients. In statistically steady 3D turbulence at high Reynolds numbers, the mean enstrophy production balances , maintaining a constant enstrophy level, with the enstrophy rate \chi \approx \nu \langle |\nabla \boldsymbol{\omega}|^2 \rangle serving as the through the inertial range. This equilibrium contrasts with ideal inviscid flows, where enstrophy is conserved absent . In bounded domains, boundary effects introduce additional contributions to enstrophy dynamics, particularly at no-slip walls where high gradients form in thin s, significantly enhancing local . These wall-generated fluxes can increase overall enstrophy levels by factors of up to three compared to periodic domains, as detachment and rollover of amplify small-scale gradients.

Applications in Fluid Dynamics

Role in Turbulence

In three-dimensional isotropic turbulence, enstrophy undergoes a forward cascade in the inertial , transferring from larger to smaller scales alongside , ultimately leading to viscous at high s. The canonical energy spectrum in this range follows Kolmogorov's scaling, E(k) \sim \epsilon^{2/3} k^{-5/3}, where \epsilon is the energy dissipation rate and k is the wavenumber. The enstrophy dissipation rate \chi, defined as \chi = \nu \langle |\nabla \boldsymbol{\omega}|^2 \rangle, quantifies this process and scales as \chi \sim \epsilon^{3/2} \nu^{-3/2} under high-Reynolds-number assumptions in the inertial range, highlighting the small-scale amplification driven by nonlinear interactions. In contrast, two-dimensional turbulence exhibits a distinct dual-cascade structure: an inverse to larger scales with E(k) \sim \epsilon^{2/3} k^{-5/3}, and a forward enstrophy cascade to smaller scales governed by a constant enstrophy \eta. This enstrophy inertial range yields E(k) \sim \eta^{2/3} k^{-3}, where enstrophy is conserved in the inviscid limit and dissipates primarily at small scales, leading to a steeper that emphasizes concentration in coherent structures like vortices. The enstrophy in the incompressible Navier-Stokes equations decomposes the of enstrophy into terms: vortex (), \omega_i \omega_j S_{ij} where S_{ij} is the ; viscous , \nu \nabla^2 (\omega^2 / 2); and viscous , -\nu |\nabla \boldsymbol{\omega}|^2. In turbulent flows, the stretching term dominates in three dimensions, amplifying enstrophy through three-dimensionality, while and balance it at small scales; a pressure-Hessian term may also contribute indirectly via the transport equation. Enstrophy norms, such as the global \langle \omega^2 \rangle / 2, serve as diagnostics for intensity in simulations, capturing amplification and small-scale activity more sensitively than alone, particularly in resolving extreme events or scales. In direct numerical simulations of isotropic , elevated enstrophy levels signal transitions to regimes by quantifying vortex filamentation and straining, aiding validation of convergence and Reynolds-number effects without relying solely on metrics.

Use in Geophysical and Astrophysical Flows

In geophysical flows, enstrophy conservation plays a key role in quasi-two-dimensional shallow water models, where it constrains potential vorticity fluxes to simulate the dynamics of atmospheric jets and ocean eddies accurately. These models exhibit a forward cascade of potential enstrophy to small scales, dissipated by viscosity, while kinetic energy undergoes an inverse cascade to larger scales, limited by bottom friction or planetary vorticity gradients. Parameterizations that enforce both energy and enstrophy budgets enable coarse-resolution simulations to replicate eddy-resolving outcomes, producing realistic topography-following zonal flows in barotropic systems. In geostrophic turbulence, prevalent in the atmosphere and oceans, enstrophy acts as a conserved quadratic invariant alongside energy, driving a forward enstrophy cascade to dissipative scales and an inverse energy cascade that forms large-scale structures like jets. The planetary β-effect arrests these cascades at characteristic scales, such as L_\beta \approx \sqrt{U / \beta}, where U is a typical velocity and \beta is the meridional gradient of the Coriolis parameter, leading to anisotropic turbulence with zonal alignment. This framework explains the upscale energy transfer observed in mid-latitude flows, distinct from three-dimensional isotropic cases. The Coriolis force in rapidly rotating systems modifies vortex stretching, suppressing three-dimensional effects and promoting two-dimensional-like enstrophy conservation, as seen in flows relevant to planetary interiors like Earth's core. In such regimes, partial enstrophy invariance facilitates inverse energy cascades, supporting geodynamo action where helical flows sustain magnetic fields against ohmic dissipation. Numerical schemes incorporating the full Coriolis terms ensure discrete conservation of potential enstrophy, enhancing stability in simulations of rotating convection. In astrophysical contexts, particularly (MHD), magnetic enstrophy—defined as \int |\nabla \times \mathbf{B}|^2 \, dV, where \mathbf{B} is the —analogizes fluid enstrophy and governs turbulence cascades in . Enstrophy-like partial invariants drive inverse transfers of magnetic helicity, enabling large-scale amplification in stellar interiors and accretion disks. In turbulence, ion enstrophy undergoes a direct cascade toward smaller scales, contributing to anisotropic MHD spectra observed at 1 , with forward transfers dominating in Alfvénic fluctuations. Observationally, satellite altimetry, such as from the dataset (1993–2013), measures surface enstrophy in ocean currents by deriving geostrophic velocities from sea-surface height anomalies, revealing forward cascades at scales of 100–200 km consistent with quasi-geostrophic theory. More recent analyses using satellite altimetry data up to 2023, such as from the Copernicus Marine Service, confirm forward enstrophy cascades at submesoscales (10–100 km) in global ocean currents, consistent with quasi-geostrophic theory. In numerical models like general circulation models (GCMs), enstrophy budgets track production, transfer, and , with subgrid schemes removing enstrophy near truncation scales to mimic forward cascades while conserving resolved energy. These budgets highlight the need for scale-selective to prevent spurious upscale enstrophy accumulation in atmospheric and oceanic simulations.

References

  1. [1]
    [PDF] ICASE REPORT NO. 88-8 r*
    The enstrophy is defined as the square of the &norm of the vorticity. From dimensional arguments it follows that. Page 4. - 2- where is the total rate of ...
  2. [2]
    [PDF] Enstrophy dynamics for flow past a solid body with no-slip boundary ...
    Jun 19, 2025 · The enstrophy is defined as the square of the L2-norm of the vortex function w: E(t) = ∫. Ω w2(t, x)dx. Under periodic boundary conditions ...
  3. [3]
    [PDF] Dependence of enstrophy transport and mixed mass on ... - OSTI.GOV
    Nov 8, 2019 · Moreover, the transfer equation of enstrophy across scales of turbulence and the definition of ... turbulent fluid dynamics. Cambridge University ...<|control11|><|separator|>
  4. [4]
    [PDF] DNS, Enstrophy Balance, and the Dissipation Equation in a ...
    Jun 27, 2013 · It is known that mean-square vorticity fluctuations (enstrophy) equal the energy dissipation in homogeneous flows. Hence an equation for the ...
  5. [5]
    [PDF] 2D Homogeneous Turbulence
    1 These two, co-existing tendencies are referred to, respectively, as the inverse energy cascade and the forward enstrophy cascade of 2D turbulence. The ...
  6. [6]
    [PDF] ON THE DIFFERENCES BETWEEN 2D AND QG TURBULENCE
    In section 2 the conservation of energy and enstrophy is derived for the case of 2D turbulence, and previous arguments for the direction of energy cascade are.
  7. [7]
    [PDF] 2-D Turbulence
    Thus enstrophy (can cascade, since there is a source terrn for vorticity gratlierlts. These results suggest that we could derive a spectrurn frorn an enstrophy ...
  8. [8]
    [PDF] A mass, energy, vorticity, and potential enstrophy conserving lateral ...
    1. Introduction. The use of energy, enstrophy, and/or potential enstrophy conserving numerical schemes for the solution of the governing. equations of fluid ...
  9. [9]
    [PDF] Potential Enstrophy in Stratified Turbulence
    Jun 11, 2013 · I V-conservation important in QG turbulence (enstrophy cascade, inverse energy cascade). I what happens at larger Ro – atmospheric mesoscale ...<|control11|><|separator|>
  10. [10]
    A novel compressible enstrophy transport equation-based analysis ...
    Apr 22, 2022 · Enstrophy, defined as Ω = ω → · ω → ⁠, represents a measure of rotational energy of the flow and it is an important quantity for vortex ...
  11. [11]
    Inertial Ranges in Two‐Dimensional Turbulence - AIP Publishing
    Two‐dimensional turbulence has both kinetic energy and mean‐square vorticity as inviscid constants of motion. Consequently it admits two formal inertial ranges.
  12. [12]
    The Lyman–Huggins interpretation of enstrophy transport
    Mar 6, 2023 · The Lighthill–Panton and Lyman–Huggins interpretations of vorticity dynamics are extended to the dynamics of enstrophy.
  13. [13]
    Phenomena connected with turbulence in the lower atmosphere
    The object of the present paper is to bring together some of the meteorological phenomena which depend on the turbulence of the lower atmosphere, ...
  14. [14]
    [PDF] Vorticity - MIT
    o. The vorticity is defined as the curl of the velocity vec- tor: @=VX V. Thus each point in the fluid has an as- sociated vector vorticity, and the whole fluid ...
  15. [15]
    [PDF] 3 Vorticity, Circulation and Potential Vorticity. - Staff
    3.1 Definitions. • Vorticity is a measure of the local spin of a fluid element given by. ~ω = ∇ × v. (1) So, if the flow is two dimensional the vorticity will ...
  16. [16]
    [PDF] translation, rotation, and vorticity
    It turns out that vorticity is equal to twice the angular velocity of a fluid particle,. 2 ζ ω. =.. Thus, vorticity is a measure of rotation of a fluid ...
  17. [17]
    [PDF] Lecture 4: Circulation and Vorticity - UCI ESS
    • Vorticity may thus be regarded as a measure of the local angular velocity of the fluid. ESS227. Prof. Jin-Yi Yu angular velocity of the fluid. Page 18 ...
  18. [18]
    [PDF] Circulation and Vorticity
    Vorticity – a microscopic measure of rotation at any point in a fluid. Vorticity is defined as the curl of the velocity ( ) Absolute vorticity:Missing: dynamics | Show results with:dynamics<|separator|>
  19. [19]
    Weak solutions, renormalized solutions and enstrophy defects in 2D ...
    Feb 2, 2004 · Enstrophy, half the integral of the square of vorticity, plays a role in 2D turbulence theory analogous to that played by kinetic energy in the ...
  20. [20]
    [PDF] Stability of Shear Flow - WHOI GFD
    of the enstrophy equation, i.e. an equation for the integral of vorticity squared. Consider. R y2 y1. (D2 − α2)v∗ · (20) + (D2 − α2)v · (21) dy = 0 to ...
  21. [21]
    [PDF] Epi-Two-Dimensional Flow and Generalized Enstrophy - UT Physics
    In this paper, we make a deeper inquiry into the helicity-enstrophy interplay: the ideal fluid mechanics is cast into a Hamiltonian form in the phase space ...
  22. [22]
    [PDF] Enstrophy dissipation for 2D incompressible flows - CSCAMM
    Vorticity -velocity formulation to 2D Navier-Stokes: ∂tω + u · ∇ω = ν∆ω,. (2a) u = K ∗ ω,. (2b) where ν is the viscosity coefficient.Missing: squared | Show results with:squared
  23. [23]
    None
    Nothing is retrieved...<|control11|><|separator|>
  24. [24]
    [PDF] Euler Equations, Navier-Stokes Equations and Turbulence
    . The vorticity ω = ∇ × u obeys an equation similar to (3):. Dνω = ω · ∇u. (12). The Eulerian-Lagrangian equations (7) and (9) have also viscous counterparts.
  25. [25]
    Maximum amplification of enstrophy in three-dimensional Navier ...
    Apr 23, 2020 · The goal of this study is to assess the largest growth of enstrophy possible in finite time in viscous incompressible flows in three dimensions.
  26. [26]
    [PDF] arXiv:1201.0102v1 [nlin.CD] 30 Dec 2011
    Dec 30, 2011 · The cascades follow from inviscid conservation of energy and enstrophy (in 2D), separation between forcing and dissipation scales and the ...<|separator|>
  27. [27]
    Vortex stretching and enstrophy production in stationary ...
    Jan 4, 2022 · One of the most basic phenomena of three-dimensional turbulence is the net production of enstrophy due to vortex stretching. Because it is easy ...
  28. [28]
    Nonperturbative mean-field theory for minimum enstrophy relaxation
    May 29, 2015 · The selective decay hypothesis states that 2D turbulence relaxes to a minimum enstrophy state. ... G. K. Vallis, Atmospheric and Oceanic Fluid ...
  29. [29]
    [PDF] Inertial-range transfer in two- and three-dimensional turbulence
    The enstrophy transfer is then analogous to spectral transfer of a passive scalar in the k-l range (Batchelor 1959;. Kraichnan 1968)) where the dominant ...Missing: coin | Show results with:coin
  30. [30]
    None
    ### Summary of Enstrophy Production and Dissipation in 2D Turbulence
  31. [31]
    Enstrophy production and dissipation in developing grid-generated ...
    Feb 19, 2016 · Direct numerical simulations are performed to investigate the spatial evolution of small-scale motions in turbulence behind a single square grid.
  32. [32]
    [PDF] Two-dimensional turbulence in square and circular domains with no ...
    The decrease of the enstrophy (figure 2(b)) indicates that self-organisation of the flow, and the associated appearance of large scale vortices (note that the ...
  33. [33]
    Dissipation and enstrophy in isotropic turbulence: Resolution effects ...
    Apr 30, 2008 · The turbulence energy dissipation and enstrophy (ie, the square of vorticity) possess different scaling properties, while available theory suggests that there ...
  34. [34]
    Space-local Navier--Stokes turbulence | Phys. Rev. Fluids
    Jan 23, 2024 · Using the enstrophy budget equation, we develop a theoretical argument explaining the alternative scaling of the large scales. We conclude the ...
  35. [35]
    Extreme events in computational turbulence - PNAS
    We focus on “extreme events” in energy dissipation and squared vorticity (enstrophy). For the Reynolds numbers of these simulations, events as large as ...Extreme Events In... · Spatial Structure · Temporal Evolution
  36. [36]
    and Enstrophy-Constrained Parameterization of Barotropic Eddy ...
    Here, we present a method for representing ocean eddies in climate models, which uses the conservation of energy, and of a similar quantity that measures the ...
  37. [37]
  38. [38]
    [PDF] Geostrophic Turbulence - ucla.edu
    This is the simplest model for the advective dynamics of geostrophic turbulence, with inverse energy cascade, forward enstrophy cascade and dissipation, ...
  39. [39]
    [PDF] Turbulent diffusion in the geostrophic inverse cascade
    This paper studies the transport of a passive tracer in turbulent flow, relevant to heat transport in Earth's atmosphere and oceans, where energy cascades to ...
  40. [40]
    [PDF] An energy and potential enstrophy conserving numerical scheme for ...
    Abstract. We present an energy- and potential enstrophy-conserving scheme for the non-traditional shallow water equations that include the complete Coriolis ...Missing: 2D | Show results with:2D
  41. [41]
  42. [42]
    A high-order finite-difference solver for direct numerical simulations ...
    The magnetic enstrophy reaches its maximal value at around t=1.0, when the magnetic energy starts to be converted back into kinetic energy. A good agreement ...
  43. [43]
    Dispersive Fast Magnetosonic Waves and Shock‐Driven ...
    Sep 24, 2020 · On the other hand, the enstrophy of solar wind ions and PUIs develops a direct energy cascade in time, where enstrophy moves toward small scales ...
  44. [44]
    Surface Ocean Enstrophy, Kinetic Energy Fluxes, and Spectra From ...
    May 4, 2018 · Surface Ocean Enstrophy, Kinetic Energy Fluxes, and Spectra From Satellite Altimetry · 1 Introduction · 2 Data Analysis and Methodology · 3 Results.
  45. [45]
    [PDF] Energy and Enstrophy Cascades in Numerical Models - ECMWF
    Jun 24, 2011 · The nonlinearity of the equations governing atmospheric flow implies interscale transfers of energy and potential enstrophy. It is important to understand ...