+ All Categories
Home > Documents > Direct Statistical Simulation of Jets and Vortices in 2D Flows · the statistics of turbulent...

Direct Statistical Simulation of Jets and Vortices in 2D Flows · the statistics of turbulent...

Date post: 26-Jan-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
10
This is an author produced version of Direct Statistical Simulation of Jets and Vortices in 2D Flows. White Rose Research Online URL for this paper: http://eprints.whiterose.ac.uk/121266/ Article: Tobias, SM and Marston, JB (2017) Direct Statistical Simulation of Jets and Vortices in 2D Flows. Physics of Fluids. ISSN 1070-6631 (In Press) This article may be downloaded for personal use only. Any other use requires prior permission of the author and AIP Publishing. The following article has been accepted by Physics of Fluids. After it is published, it will be found at http://aip.scitation.org/journal/phf/. Uploaded in accordance with the publisher's self-archiving policy. promoting access to White Rose research papers [email protected] http://eprints.whiterose.ac.uk/
Transcript
  • This is an author produced version of Direct Statistical Simulation of Jets and Vortices in 2D Flows.

    White Rose Research Online URL for this paper:http://eprints.whiterose.ac.uk/121266/

    Article:

    Tobias, SM and Marston, JB (2017) Direct Statistical Simulation of Jets and Vortices in 2D Flows. Physics of Fluids. ISSN 1070-6631 (In Press)

    This article may be downloaded for personal use only. Any other use requires prior permission of the author and AIP Publishing. The following article has been accepted by Physics of Fluids. After it is published, it will be found at http://aip.scitation.org/journal/phf/. Uploaded in accordance with the publisher's self-archiving policy.

    promoting access toWhite Rose research papers

    [email protected]://eprints.whiterose.ac.uk/

    mailto:[email protected]://eprints.whiterose.ac.uk/

  • DSS of Jets to Vortices

    Direct Statistical Simulation of Jets and Vortices in 2D FlowsS. M. Tobias1, a) and J. B. Marston2, b)

    1)Department of Applied Mathematics, University of Leeds. Leeds, LS2 9JT,

    UK2)Department of Physics, Box 1843, Brown University, Providence, RI 02912-1843 USA

    (Dated: 19 August 2017)

    In this paper we perform Direct Statistical Simulations of a model of two-dimensional flow that exhibitsa transition from jets to vortices. The model employs two-scale Kolmogorov forcing, with energy injecteddirectly into the zonal mean of the flow. We compare these results with those from Direct Numerical Simula-tions. For square domains the solution takes the form of jets, but as the aspect ratio is increased a transitionto isolated coherent vortices is found. We find that a truncation at second order in the equal-time but nonlocalcumulants that employs zonal averaging (zonal CE2) is capable of capturing the form of the jets for a rangeof Reynolds numbers as well as the transition to the vortex state, but, unsurprisingly, is unable to reproducethe correlations found for the fully nonlinear (non-zonally symmetric) vortex state. This result continues theprogram of promising advances in statistical theories of turbulence championed by Kraichnan.

    PACS numbers: 47.27.De, 47.32.cd, 47.27.eb, 94.05.JqKeywords: two-dimensional turbulence, coherent structures, statistical theories

    I. DIRECT STATISTICAL SIMULATION: THE LEGACY

    OF KRAICHNAN

    Robert Kraichnan’s vision of a statistical mechanics ofturbulence notably emphasized essential differences be-tween flows in two and three spatial dimensions1. Twodimensional flows are striking for the frequent emergenceof coherent structures. The structures are of two basictypes: Vortices and jets2,3. The Juno mission to Jupiterhas recently returned beautiful images of both types ofstructures, with jets dominating at low latitudes, and aproliferation of vortices near the poles4. In this paperwe investigate a simple model of two dimensional fluidflow that exhibits a transition between jets and vortices.We employ both Direct Numerical Simulation (DNS) andDirect Statistical Simulation (DSS). DSS is a rapidly de-veloping set of tools that attempt to describe, directly,the statistics of turbulent flows, bypassing the traditionalway of accumulation of those statistics (for example meanflows and two-point correlation functions) by DNS. Thesestatistical methods lead to a deeper understanding offluid flows that should guide researchers to regimes notaccessible through DNS.The pioneering work of Kraichnan largely focused on

    flows with isotropic and homogeneous statistics5. Thestatistical description of forward and/or inverse cascadesof energy between different scales, the topic explored inhis seminal 1967 paper1, is particularly clear in this con-text. Equally important, translational and rotationalsymmetries reduce the technical complexity of statisti-cal theories. Most fluid flows in nature, however, areboth anisotropic and heterogeneous. In DSS this is seenas a feature, rather than a defect, as anisotropy and

    a)Electronic mail: [email protected])Electronic mail: [email protected]

    inhomogeneity can lessen nonlinearity of the flows andmake the statistics accessible to perturbative computa-tion. We show that a particularly simple version of DSS,one in which the equations of motion for the spatially-averaged statistics are closed at the level of second-ordermoments or cumulants6,7, is already able to reproducemany features of the model two-dimensional flow. TheKolmogorov forcing we employ is purely deterministicand no parameters are tuned at the level of the second-order cumulant expansion (CE2), permitting a fair andunbiased comparison between DNS and DSS.We introduce the model in Section II. Results from

    DNS and DSS are presented in Section III. Compari-son between the two approaches is made in Section IVand some conclusions and possible directions for furtherexploration are discussed in Section V.

    II. SET-UP OF THE MODEL: FORMULATION,

    EQUATIONS, AND FORCING.

    The models we study are of incompressible fluid mov-ing on a two-torus — i.e a two dimensional Cartesiandomain (0 ≤ x < Lx, 0 ≤ y < Ly = 2π) with peri-odic boundary conditions in both directions. The fluidmotion is damped by viscosity ν and driven by a time-independent forcing (described below). Owing to thetwo-dimensionality of the system the dynamics is com-pletely described by the time-evolution of the vorticity

    ζ ≡ ẑ · (~∇× ~v), which (in dimensional units) is given by

    ζ̇ + J(ψ, ζ) = ν∇2ζ + g(y), (1)

    where g(y) is the forcing term and J(A,B) is the Jaco-bian operator given by J(A,B) = AxBy − AyBx. Wenote that, in contrast to some earlier models, here we donot consider the effects of rotation via a β-effect.The forcing g(y) is a generalisation of the Kolmogorov

    forcing (see e.g. 8) to two meridional wavenumbers; that

  • DSS of Jets to Vortices 2

    is we set

    g(y) = A1 cos(y) + 4.0 ∗A4 cos(4y). (2)

    We set A1 = −1 and A4 = −2, which leads to non-trivialdynamics in the fluid system. We note here that thischoice of deterministic forcing injects energy directly intothe zonal flow, and should be contrasted with previousstudies that impose stochastic forcing in the small zonalscales. Once the forcing and the length scale in y (say)is fixed then the dynamics (and indeed statistics) of theflow is determined by the choice of the viscosity ν (whichcontrols the Reynolds number of the flow, which may becalculated a posteriori) and the aspect ratio (determinedby Lx).The aim of this paper is to determine how successful

    Direct Statistical Simulation truncated at second order(sometimes termed CE2) is at describing the transitionsthat occur as the parameters are varied. In particular weshall investigate how well DSS reproduces the strengthof the mean shear flows and the transition from solutionsdominated by jets to those dominated by coherent vortexpairs.

    III. RESULTS

    In this section we describe results from DNS (in sub-section IIIA) and those obtained from DSS using CE2(in subsection III B) before comparing them in SectionIV.

    A. Direct Numerical Simulation

    Direct Numerical simulation is performed using apseudo-spectral code optimised for use on parallel archi-tectures with typical resolutions of 5122. In all cases theresolution is increased to this level to obtain until con-vergent results. Initially we integrate the equations forthree different values of the viscosity in a square domainwith Lx = Ly = 2π. The time series for the result-

    ing spatially-averaged enstrophy ζ2 and kinetic energyψ2y + ψ

    2x where

    A ≡1

    LxLy

    ∫∫

    Adxdy, (3)

    are shown in Figure 1 for three values of the viscosity.This figure clearly shows that, as expected, as the viscos-ity is decreased (with the forcing fixed) both the enstro-phy (top panel) and kinetic energy (bottom panel) of thesolutions increases.Figure 2 (multimedia view) shows snapshots from

    movies of the evolution of the vorticity for the cases withLx = 2π (which are included in the supplementary ma-terial). The flow is reasonably laminar although the so-lution has already undergone a bifurcation from a steadystate. After some initial transients the solution becomes

    0 100 200 300 400 500 600 700 800 900

    t

    0

    200

    400

    600

    800

    1000

    1200

    1400

    ζ2

    ν = 0.03

    ν = 0.022

    ν = 0.02

    (a)

    0 100 200 300 400 500 600 700 800 900

    t

    0

    200

    400

    600

    800

    1000

    1200

    u2+v2

    ν = 0.03

    ν = 0.022

    ν = 0.02

    (b)

    FIG. 1. Enstrophy and Kinetic Energy timeseries for the so-lutions at three different viscosities for Lx = Ly = 2π. Bothof these increase with decreasing viscosity.

    time periodic, with a strong band/jet of positive vorticityin the domain and weaker negative vorticity (in the formof a vortex) at the edges. This corresponds to a right-ward jet in the upper half of the domain and a reversejet in the lower half (see later). Both the jet and the vor-tex remain fixed in space though pulse in time. Thoughtime-dependence is present, these vortex regions possessa well defined zonal mean, which can be calculated by av-eraging over a suitably long time. The average Reynoldsnumber for this flow is given by Re ≈ 730.

    As the viscosity is decreased from this solution the dy-namics becomes more irregular and time-dependent, asshown in the two movies. Decrease of the viscosity leadsto stronger patches of vorticity and faster flows. Forν = 0.022, Re ≈ 1370, whilst for ν = 0.02, Re ≈ 1650.For both of these solutions the non-zonally symmetricpart i.e. the part of the solution with kx 6= 0 and thevortex travel in space, rather than remaining fixed as forthe earlier case. The strength of the vortex patches and

  • DSS of Jets to Vortices 3

    jets increases with decreasing ν (as does the correspond-ing mean flows and vorticities — see later) as the inertialterms play an increasingly important role.When the aspect ratio is increased so that Lx = 4π, the

    nature of the solution changes. The driving which is in-dependent of x no longer puts substantial power into thekx = 0 modes, but instead drives a fully nonlinear quasi-steady vortex pair solution as shown in Figure 4 (multi-media view) and the corresponding movie. This state isreminiscent of the localised states analysed extensively inRef. 8. For this state the average enstrophy and vorticityare significantly lower than for the jet states (as shownin the time series in Figure 3). Moreover, as we shallsee, this state has little energy in the zonally-averagedvorticity and flow and so can not be characterised as ajet state. This remarkable transition appears to be theopposite of a zonostrophic instability (see Ref. 9); therea small-scale forcing with zero zonal mean drives flowthat interacts with rotation to put significant amount ofenergy into a zonally averaged jet. Here the forcing isdesigned to drive strong zonal flows, but nonlinear in-teractions prefer to put energy into vortex states withweak zonal flows, and may therefore be termed a “vor-tostrophic instability.” We note that the aspect ratiocontrols a similar transition between jets and vortices inother two-dimensional models10–13. The non-trivial non-linear dynamics provides an interesting testing groundfor the types of statistical theories favoured by Kraichnanand so, in the next section, we compare the results ob-tained here via Direct Numerical Simulation, with thoseobtained by Direct Statistical Simulation truncated atsecond order (CE2).

    B. Direct Statistical Simulation: The Cumulant

    Equations

    In this section we perform DSS for the system for thesame range of parameters as above. The approach wetake is based upon truncating the hierarchy of equationsof motion for the equal-time cumulants at low order. It isrelated to stochastic structural stability theory (S3T)14,15

    and other approaches16,17 that do not assume spatial ho-mogeneity or isotropy in the statistics. Here we definethe cumulants in terms of zonal averages over the x-direction (see Refs. 6, 7, 18–20) as opposed to ensembleaverages21–23. Thus

    cζ(y) = 〈ζ〉, (4)

    where 〈〉 indicates a zonal average, is the first cumulantand

    cζζ(y, y′, ξ) = 〈ζ ′(x, y) ζ ′(x+ ξ, y′)〉, (5)

    is the second cumulant (or two-point correlation func-tion). We note that owing to the translational symmetryof the system (including the forcing) the first cumulantis a function only of y and the second cumulant is a func-tion of three rather than four dimensions18,19. There are

    similar definitions for the first and second cumulants in-volving the streamfunction (i.e. cψ and cψζ), but thesecan be related straightforwardly to the cumulants for thevorticity.The cumulant hierarchy can be derived in a number of

    ways19. Truncated at second order (CE2) this takes theform of evolution equations for cζζ(y, y

    ′) and cζ(y) (seeRef. 18):

    ∂tcζ(y) =

    [

    (

    ∂y+

    ∂y′

    )

    ∂ξcψζ(y,y′,ξ)

    ]

    y′=y,ξ=0

    + g(y) + ν∂2

    ∂y2cζ(y), (6)

    together with

    ∂tcζζ =

    ∂ycψ(y)

    ∂ξcζζ(y, y

    ′, ξ)

    −∂

    ∂y(cζ(y))

    ∂ξcψζ(y, y

    ′, ξ)

    −∂

    ∂y′cψ(y

    ′)∂

    ∂ξcζζ(y, y

    ′, ξ)

    −∂

    ∂y′(cζ(y

    ′))∂

    ∂ξcζψ(y, y

    ′, ξ)

    + ν

    (

    2∂2

    ∂ξ2+

    ∂2

    ∂y2+

    ∂2

    ∂y′2

    )

    cζζ . (7)

    In the limit of no forcing or dissipation, equations (6–7) conserve energy, enstrophy and the Kelvin impulse;thus the CE2 is a conservative approximation6,19. Theequations are integrated forward in time numerically us-ing a pseudospectral code with typical resolutions of(ny, ny′ , nξ) = (64, 64, 16) until the statistics have set-tled down to a statistically steady state and means andtwo-point correlation functions are averaged in time.

    IV. COMPARISON OF DNS AND DSS

    The solutions from the cumulant equations are com-pared with the corresponding statistics obtained fromDNS by averaging in x and time. In all cases the DNS so-lutions are averaged over the final third of the evolutionand the statistics are well converged. The DSS solutionswere averaged over the final 10% of the evolution, thoughthe averages rapidly converge in all cases.Figure 5 shows the comparison between DNS and CE2

    for a domain of length 2π and decreasing viscosity. Asthe viscosity is decreased the mean vorticity amplitude〈ζ〉 and the mean zonal flow 〈u〉 increase in amplitudeas expected. It is clear that the comparison of the meanflows between CE2 and DNS is excellent. CE2 has a ten-dency to emphasise turning points in the vorticity thatare washed out by eddy + eddy → eddy interactions inthe DNS 6. However the agreement in the amplitude andform of the solution is very good. One might expect CE2

  • DSS of Jets to Vortices 4

    (a) (b) (c)

    FIG. 2. Snapshots of vorticity for (a) ν = 0.03 (b) ν = 0.022 (c) ν = 0.02, with Lx = Ly = 2π. For these figures the coloursare scaled between (a) [−61, 34], (b) [−67, 46], (c) [−75, 52]. Movies showing the dynamics for these parameters are containedin the supplementary material (multimedia view).

    0 200 400 600 800 1000

    t

    0

    100

    200

    300

    400

    500

    600

    ζ2

    Xmx = 2π

    Xmx = 4π

    (a)

    0 200 400 600 800 1000

    t

    0

    100

    200

    300

    400

    500

    u2+v2

    Xmx = 2π

    Xmx = 4π

    (b)

    FIG. 3. Enstrophy and Kinetic Energy timeseries for the so-lutions at ν = 0.03 and Lx = 2π and Lx = 4π. The solutionsfor the larger aspect ratio are significantly weaker in energyand have less enstrophy.

    to improve as the ratio of energy in the zonal mean flowto that in the fluctuations increases. This expectation islargely met, though in all cases agreement is good.

    What is remarkable is that CE2 is capable of capturingthe transition to vortices as shown in Figure 6. When thezonal flow is switched off in DNS by changing the aspectratio, CE2 predicts the same behaviour. Although CE2does not get the form of the weak zonal flow completelycorrect, it predicts the amplitude very well. This is un-expected since zonally averaged CE2 is expected to workpoorly in a case where the zonal means are small andsubdominant to the fluctuations.

    A more stringent test of the accuracy of statistical rep-resentation involves a comparison of not only the zonalmeans (first cumulants) but also the two-point correla-tion functions (second cumulants). These are given inFigures 7 and 8. These show the two-point correlationfunction cζζ(y, 3π/4, ξ), i.e. the correlation in space withthe point three-eighths of the way up on the left handside of each plot. In all the cases for the jet solutionsthis is dominated by a kx = 1, ky = 2 solution, thoughthis is modulated in y. DNS and CE2 can be seen tobe in good agreement here as they should be for flowswith such strong mean. Figure 8 shows less good agree-ment between DNS and CE2, with CE2 failing to matchboth the amplitude and spatial form of the second cu-mulant (underestimating the ky = 0 component) for thecase where the flow is dominated by strong vortices —the DNS velocity correlation function has a near k = 0symmetry in the meridional direction and a near reflec-tion symmetry in the zonal direction as the solution isdominated by a kx = 1, ky = 0 vortex mode. This isclearly picked up by the correlation function, which isalso dominated by these wavenumbers. The failure ofCE2 to match the two-point correlation function for thevortex state is not surprising as the form of this correla-tion function is presumably determined by eddy + eddy→ eddy interactions that are discarded from the quasilin-ear CE2 description. The importance of cubic terms for

  • DSS of Jets to Vortices 5

    FIG. 4. Snapshot of vorticity for ν = 0.03 (b) Lx = 4π and Ly = 2π. For this figures the colours are scaled between [−23, 18].A movie showing the dynamics for these parameters is contained in the supplementary material (multimedia view).

    the form of the solution is determined by the amplitudeof their projection onto the second cumulant. This canbe seen by comparing the second cumulants from CE2and DNS. What is clear is that this projection is smallfor the cases with a significant zonal mean and large forthe case of the vortex We discuss possible strategies forimproving the agreement between DNS and DSS for thiscase next in the Discussion.

    V. DISCUSSION

    We have shown that an expansion in equal-time andzonally averaged cumulants, truncated at second order,is able to describe both the jet- and vortex-dominatedphases of a two-dimensional flow driven by deterministicKolmogorov forcing. Zonal mean flows in both phasesare accurately reproduced, and two-point correlations ofthe vorticity are also captured in the jet phase, but notin the vortex-dominated phase.

    It is remarkable that DSS with such a simple closurecan capture much of the behavior exhibited by DNS. Weexpect that more sophisticated closures will be able todescribe the vortex phase accurately. Higher-order clo-sures such as CE3∗ and CE2.5 (see Ref. 19) include eddy+ eddy → eddy interactions and can improve qualita-tive agreement in two-point correlations found by DSSin comparison to DNS. The replacement of zonal aver-ages with ensemble averages has been demonstrated todescribe non-zonal structures22, likely including the vor-tices seen here at aspect ratio Ly/Lx = 2. Finally thegeneralized quasi-linear approximation (GQL)24,25 andits associated generalized cumulant expansion (GCE2),by allowing long-wavelength non-zonal structures to in-teract fully nonlinearly, should also be able to describethe vortex-dominated phase. Each of these variantsis more computationally demanding than simple zonal-

    average CE2, but we plan to test these other forms ofDSS for the Kolmogorov forced model.All of these different forms of DSS respect the real-

    izability inequalities studied by Kraichnan26,27. Theygeneralize the program of understanding the statisticsof turbulence, greatly advanced by Kraichnan, to encom-pass anisotropy and heterogeneity. Statistical theoriesof turbulence thus continue to extend their reach, per-mitting a deeper understanding of fluid flows that maysomeday allow us to access regimes not currently reach-able by DNS.

    ACKNOWLEDGMENTS

    We wish to acknowledge the help of Mark Dixon of TheUniversity of Leeds HPC facility team. All calculationswere performed on the Arc2 machine at The Universityof Leeds.

    1R. H. Kraichnan, “Inertial Ranges in Two-Dimensional Turbu-lence,” Phys. Fluids 10, 1417–8 (1967).

    2R. H. Kraichnan and D. Montgomery, “Two-dimensional turbu-lence,” Rep. Prog. Phys. 43, 547–619 (1980).

    3J. Y.-K. Cho and L. M. Polvani, “The emergence of jets andvortices in freely evolving, shallow-water turbulence on a sphere,”Physics of Fluids (1994-present) 8, 1531–1552 (1996).

    4S. J. Bolton, A. Adriani, V. Adumitroaie, M. Allison, J. Ander-son, S. Atreya, J. Bloxham, S. Brown, J. E. P. Connerney, E. De-Jong, W. Folkner, D. Gautier, D. Grassi, S. Gulkis, T. Guillot,C. Hansen, W. B. Hubbard, L. Iess, A. Ingersoll, M. Janssen,J. Jorgensen, Y. Kaspi, S. M. Levin, C. Li, J. Lunine, Y. Miguel,A. Mura, G. Orton, T. Owen, M. Ravine, E. Smith, P. Steffes,E. Stone, D. Stevenson, R. Thorne, J. Waite, D. Durante, R. W.Ebert, T. K. Greathouse, V. Hue, M. Parisi, J. R. Szalay, andR. Wilson, “Jupiter’s interior and deep atmosphere: The initialpole-to-pole passes with the Juno spacecraft,” Science 356, 821–825 (2017).

    5J. R. Herring and R. H. Kraichnan, “Comparison of some ap-proximations for isotropic turbulence,” in Statistical Models andTurbulence: Proceedings of a Symposium held at the University

    of California (Goddard Space Flight Center, 1972) pp. 148–194.

  • DSS of Jets to Vortices 6

    −30 −20 −10 0 10 20 30 40〈ζ〉

    0

    1

    2

    3

    4

    5

    6

    y

    DNS

    CE2

    −40 −30 −20 −10 0 10 20 30 40〈u〉

    0

    1

    2

    3

    4

    5

    6

    y

    DNS

    CE2

    (a) (b)

    −60 −40 −20 0 20 40 60〈ζ〉

    0

    1

    2

    3

    4

    5

    6

    y

    DNS

    CE2

    −60 −40 −20 0 20 40 60〈u〉

    0

    1

    2

    3

    4

    5

    6

    y

    DNS

    CE2

    (c) (d)

    −40 −30 −20 −10 0 10 20 30 40 50〈ζ〉

    0

    1

    2

    3

    4

    5

    6

    y

    DNS

    CE2

    −60 −40 −20 0 20 40 60〈u〉

    0

    1

    2

    3

    4

    5

    6

    y

    DNS

    CE2

    (e) (f)

    FIG. 5. Comparison of DNS and CE2. Mean vorticity and zonal flow for ν = 0.03 (top row) ν = 0.022 (middle row) ν = 0.02(bottom row).

  • DSS of Jets to Vortices 7

    −30 −20 −10 0 10 20 30 40〈ζ〉

    0

    1

    2

    3

    4

    5

    6

    y

    DNS

    CE2

    DNS

    CE2

    −40 −30 −20 −10 0 10 20 30 40〈u〉

    0

    1

    2

    3

    4

    5

    6

    y

    DNS

    CE2

    DNS

    CE2

    (a) (b)

    FIG. 6. Comparison of DNS and CE2. Mean vorticity and zonal flow for ν = 0.03 and Lx = 2π (red and cyan) and Lx = 4π(blue and green).

    6J. Marston, E. Conover, and T. Schneider, “Statistics of anunstable barotropic jet from a cumulant expansion,” Journal ofthe Atmospheric Sciences 65, 1955–1966 (2008).

    7S. Tobias, K. Dagon, and J. Marston, “Astrophysical fluid dy-namics via direct statistical simulation,” The Astrophysical Jour-nal 727, 127 (2011).

    8D. Lucas and R. Kerswell, “Spatiotemporal dynamics intwo-dimensional Kolmogorov flow over large domains,” Jour-nal of Fluid Mechanics 750, 518–554 (2014), arXiv:1308.3356[physics.flu-dyn].

    9K. Srinivasan andW. R. Young, “Zonostrophic Instability,” Jour-nal of Atmospheric Sciences 69, 1633–1656 (2012).

    10F. Bouchet and E. Simonnet, “Random Changes of Flow Topol-ogy in Two-Dimensional and Geophysical Turbulence,” PhysicalReview Letters 102, 094504–4 (2009).

    11F. Bouchet and A. Venaille, “Statistical mechanics of two-dimensional and geophysical flows,” Physics Reports 515, 227–295 (2012).

    12J. Laurie and F. Bouchet, “Computation of rare transitions in thebarotropic quasi-geostrophic equations,” New Journal of Physics17, 1–25 (2015).

    13A. Frishman, J. Laurie, and G. Falkovich, “Jets or vor-tices—What flows are generated by an inverse turbulent cas-cade?” Physical Review Fluids 2, 032602–8 (2017).

    14B. F. Farrell and P. J. Ioannou, “Structure and Spacing of Jetsin Barotropic Turbulence,” Journal of the Atmospheric Sciences64, 3652–3665 (2007).

    15N. C. Constantinou, B. F. Farrell, and P. J. Ioannou, “Emer-gence and equilibration of jets in beta-plane turbulence: applica-tions of Stochastic Structural Stability Theory,” Journal of theAtmospheric Sciences , 131121143600001 (2013).

    16J. Laurie, G. Boffetta, G. Falkovich, I. Kolokolov, and V. Lebe-dev, “Universal Profile of the Vortex Condensate in Two-Dimensional Turbulence,” Physical Review Letters 113, 254503–5 (2014).

    17G. Falkovich, “Interaction between mean flow and turbulence intwo dimensions,” Proceedings of the Royal Society of LondonSeries a-Mathematical Physical and Engineering Sciences 472,20160287–14 (2016).

    18S. Tobias and J. Marston, “Direct statistical simulation of out-of-equilibrium jets,” Physical review letters 110, 104502 (2013).

    19J. Marston, W. Qi, and S. Tobias, “Direct Statistical Simulationof a Jet,” arXiv preprint arXiv:1412.0381 (2014).

    20F. Ait-Chaalal, T. Schneider, B. Meyer, and J. Marston, “Cumu-

    lant expansions for atmospheric flows,” New Journal of Physics18, 025019 (2016).

    21N. A. Bakas and P. J. Ioannou, “Emergence of large scale struc-ture in barotropic β-plane turbulence,” Physical review letters110, 224501 (2013).

    22N. A. Bakas, N. C. Constantinou, and P. J. Ioannou, “S3T Sta-bility of the Homogeneous State of Barotropic Beta-Plane Tur-bulence,” Journal of the Atmospheric Sciences 72, 1689–1712(2015).

    23A. Allawala, S. M. Tobias, and J. B. Marston, “Model reductionof direct statistical simulation model reduction of direct statisti-cal simulation,” (2017), (in preparation).

    24J. B. Marston, G. P. Chini, and S. M. Tobias, “GeneralizedQuasilinear Approximation: Application to Zonal Jets,” PhysicalReview Letters 116, 214501 (2016).

    25A. Child, R. Hollerbach, B. Marston, and S. Tobias, “Gener-alised quasilinear approximation of the helical magnetorotationalinstability,” Journal of Plasma Physics 82, 905820302 (2016).

    26R. H. Kraichnan, “Realizability Inequalities and Closed MomentEquations,” Annals of the New York Academy of Sciences 357,37–46 (1980).

    27R. H. Kraichnan, “Decimated Amplitude Equations in Tur-bulence Dynamics,” in Theoretical Approaches to Turbulence(Springer New York, New York, NY, 1985) pp. 91–135.

  • DSS of Jets to Vortices 8

    0 1 2 3 4 5 6

    ξ

    0

    1

    2

    3

    4

    5

    6

    y

    −72

    −54

    −36

    −18

    0

    18

    36

    54

    72

    0 1 2 3 4 5 6

    ξ

    0

    1

    2

    3

    4

    5

    6

    y

    −102.0

    −76.5

    −51.0

    −25.5

    0.0

    25.5

    51.0

    76.5

    102.0

    (a) (b)

    0 1 2 3 4 5 6

    ξ

    0

    1

    2

    3

    4

    5

    6

    y

    −84

    −63

    −42

    −21

    0

    21

    42

    63

    84

    0 1 2 3 4 5 6

    ξ

    0

    1

    2

    3

    4

    5

    6

    y

    −180

    −135

    −90

    −45

    0

    45

    90

    135

    180

    (c) (d)

    0 1 2 3 4 5 6

    ξ

    0

    1

    2

    3

    4

    5

    6

    y

    −96

    −72

    −48

    −24

    0

    24

    48

    72

    96

    0 1 2 3 4 5 6

    ξ

    0

    1

    2

    3

    4

    5

    6

    y

    −180

    −135

    −90

    −45

    0

    45

    90

    135

    180

    (e) (f)

    FIG. 7. cζζ(y, 3π/4, ξ) for DNS (left) and CE2 (right) for the case ν = 0.03 (top row) ν = 0.022 (middle row) and ν = 0.02(bottom row). In all cases Lx = 2π.

  • DSS of Jets to Vortices 9

    0 2 4 6 8 10 12

    ξ

    0

    1

    2

    3

    4

    5

    6

    y

    −34.0

    −27.2

    −20.4

    −13.6

    −6.8

    0.0

    6.8

    13.6

    20.4

    27.2

    0 2 4 6 8 10 12

    ξ

    0

    1

    2

    3

    4

    5

    6

    y

    −3.40

    −2.55

    −1.70

    −0.85

    0.00

    0.85

    1.70

    2.55

    3.40

    (a) (b)

    FIG. 8. cζζ(y, 3π/4, ξ) for DNS (left) and CE2 (right) for the case ν = 0.03, Lx = 4π.


Recommended