9
Vortex stripping and the erosion of coherent structures in two-dimensional flows Annarita Mariotti and Bernard Legras Laborafoire de M%orologie Dynamique du C.N.R.S., .??cole Normale Supirieure, 24 rue Lhomond, 75231 Paris Cedex 05, France David G. Dritschel Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Silver Street, Cambridge CB3 9Ew United Kingdom (Received 16 March 1994, accepted 5 August 1994) This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid stripping, induced by an externally imposed adverse shear, and viscous diffusion, either in the form of Newtonian viscosity or hyperviscosity. It is shown that vortex erosion is greatly amplified by the presence of diffusion; abrupt vortex breakup or gradual quasi-equilibrium evolution depend crucially on the strain to peak vorticity ratio and on the Reynolds number. Peculiar, unexpected effects are observed when hyperviscosity is used in place of Newtonian viscosity. 0 1994 American Institute of Physics. I. INTRODUCTION When a two-dimensional vortex flow is submitted to the action of an external straining field, due either to other vor- tices or prescribed body forces, the vorticity initially located at the periphery of the vortex is torn away, leaving behind an eroded vortex with a strongly intensified vorticity gradient on its edge. When the strain is strong enough, the whole vortex breaks into filamentary structures which mix them- selves within the ambient flow. This phenomenon, termed as “stripping,“’ is particularly relevant to the dynamics of geo- physical vortices like the polar stratospheric vortex, blocking anticyclones and oceanic eddies.2 Various approaches have been made to this problem, starting from models in which a uniform, initially elliptical vorticity patch is submitted to a rotating straining field3 (for which partial erosion is not possible but complete breaking may still occur), to more complicated and realistic situations, describing the effect of strain on monotonically distributed vorticity profiles.4,5 For an initially axisymmetric monotonic vorticity distri- bution, the flow evolution depends on (1) the ratio of the maximum absolute value of vorticity within the vortex to the strain, (2) the background rotation/strain ratio, (3) the time- dependence of the applied strain (i.e., progressive or abrupt) and (4) the isolation, or lack thereof, of the vortex. Stripping occurs because the combined velocity field of the vortex and the external strain possesses stagnation points (in a fixed or rotating frame of reference) capable of pen- etrating the interior of the vortex. An external strain is not essential if the vortex is initially sufficiently eIongated6 be- cause the vortex itself may then initially contain the stagna- tion points. However, the effects of stripping are more pro- nounced in the presence of external strain. For instance, the abrupt application of adverse shear leads to a strong deformation of the vortex and the extrusion of two filaments or tongues of vorticity, which are subse- quently advected away by the external shear flow. Figure 1 shows the case of a street of vortices submitted to a uniform adverse shear. Here the vorticity filaments expelled from one vortex collide with neighboring vortices along the street and, by iterative folding and stretching, generate a complex multi- layer structure. An important consequence of stripping is the steepening of vorticity gradients on the edge of the vortex-see the cross-section on Fig. 1. The shear selectively removes the outer part of the vortex, leaving the inner part intact. Under weakly diffusive conditions, the resulting sharp gradient is preserved for many vortex rotation periods even if the exter- nal strain is turned off. When the external strain is main- tained, the sharp gradient is permanently regenerated by stripping. This effect is readily observed in the atmosphere, where for example the polar stratospheric vortex exhibits a well-defined edge, both in potential vorticity and in chemical tracer concentration, which persists throughout the winter. The generation of vorticity filaments from the polar vortex edge has also been lent observational support by a recent analysis of Waugh and Plumb (see Ref. 7 and references therein). The piling of vorticity filaments along the edge of the vortex seen in Fig. 1 (and also in the stratosphere7) eventu- ally generates a well-mixed region when the width of the 6laments reaches the diffusive scale. In cases when the fila- ments reconnect with the vortex, the edge gradients may be diminished and a smoother profile restored. In many practical situations, the edge of the vortex is smoothed by diffusive processes; This may occur naturally when the Reynolds number is not large but usually happens as a consequence of not resolving the small scales of the motion in numerical experiments. This is particularly true in planetary geophysical fluid dynamics where the grid scale of the models employed is about 10s times bigger than the dis- sipation length. None of the currently used global weather- forecasting models are able to resolve the polar vortex edge and the encircling filaments. The lack of resolution is palli- ated by using artificial devices grouped under the name of turbulent viscosity which tend to diffuse down the large- scale resolved gradients. There is no doubt that such proce- 3954 Phys. Fluids 6 (12), December 1994 1070-6631/94/6(12)/3954/9/$6.00 0 1994 American Institute of Physics Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

Vortex stripping and the erosion of coherent structures in two-dimensional flows

Annarita Mariotti and Bernard Legras Laborafoire de M%orologie Dynamique du C.N.R.S., .??cole Normale Supirieure, 24 rue Lhomond, 75231 Paris Cedex 05, France

David G. Dritschel Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Silver Street, Cambridge CB3 9Ew United Kingdom

(Received 16 March 1994, accepted 5 August 1994)

This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid stripping, induced by an externally imposed adverse shear, and viscous diffusion, either in the form of Newtonian viscosity or hyperviscosity. It is shown that vortex erosion is greatly amplified by the presence of diffusion; abrupt vortex breakup or gradual quasi-equilibrium evolution depend crucially on the strain to peak vorticity ratio and on the Reynolds number. Peculiar, unexpected effects are observed when hyperviscosity is used in place of Newtonian viscosity. 0 1994 American Institute of Physics.

I. INTRODUCTION

When a two-dimensional vortex flow is submitted to the action of an external straining field, due either to other vor- tices or prescribed body forces, the vorticity initially located at the periphery of the vortex is torn away, leaving behind an eroded vortex with a strongly intensified vorticity gradient on its edge. When the strain is strong enough, the whole vortex breaks into filamentary structures which mix them- selves within the ambient flow. This phenomenon, termed as “stripping,“’ is particularly relevant to the dynamics of geo- physical vortices like the polar stratospheric vortex, blocking anticyclones and oceanic eddies.2

Various approaches have been made to this problem, starting from models in which a uniform, initially elliptical vorticity patch is submitted to a rotating straining field3 (for which partial erosion is not possible but complete breaking may still occur), to more complicated and realistic situations, describing the effect of strain on monotonically distributed vorticity profiles.4,5

For an initially axisymmetric monotonic vorticity distri- bution, the flow evolution depends on (1) the ratio of the maximum absolute value of vorticity within the vortex to the strain, (2) the background rotation/strain ratio, (3) the time- dependence of the applied strain (i.e., progressive or abrupt) and (4) the isolation, or lack thereof, of the vortex.

Stripping occurs because the combined velocity field of the vortex and the external strain possesses stagnation points (in a fixed or rotating frame of reference) capable of pen- etrating the interior of the vortex. An external strain is not essential if the vortex is initially sufficiently eIongated6 be- cause the vortex itself may then initially contain the stagna- tion points. However, the effects of stripping are more pro- nounced in the presence of external strain.

For instance, the abrupt application of adverse shear leads to a strong deformation of the vortex and the extrusion of two filaments or tongues of vorticity, which are subse- quently advected away by the external shear flow. Figure 1 shows the case of a street of vortices submitted to a uniform

adverse shear. Here the vorticity filaments expelled from one vortex collide with neighboring vortices along the street and, by iterative folding and stretching, generate a complex multi- layer structure.

An important consequence of stripping is the steepening of vorticity gradients on the edge of the vortex-see the cross-section on Fig. 1. The shear selectively removes the outer part of the vortex, leaving the inner part intact. Under weakly diffusive conditions, the resulting sharp gradient is preserved for many vortex rotation periods even if the exter- nal strain is turned off. When the external strain is main- tained, the sharp gradient is permanently regenerated by stripping. This effect is readily observed in the atmosphere, where for example the polar stratospheric vortex exhibits a well-defined edge, both in potential vorticity and in chemical tracer concentration, which persists throughout the winter. The generation of vorticity filaments from the polar vortex edge has also been lent observational support by a recent analysis of Waugh and Plumb (see Ref. 7 and references therein).

The piling of vorticity filaments along the edge of the vortex seen in Fig. 1 (and also in the stratosphere7) eventu- ally generates a well-mixed region when the width of the 6laments reaches the diffusive scale. In cases when the fila- ments reconnect with the vortex, the edge gradients may be diminished and a smoother profile restored.

In many practical situations, the edge of the vortex is smoothed by diffusive processes; This may occur naturally when the Reynolds number is not large but usually happens as a consequence of not resolving the small scales of the motion in numerical experiments. This is particularly true in planetary geophysical fluid dynamics where the grid scale of the models employed is about 10s times bigger than the dis- sipation length. None of the currently used global weather- forecasting models are able to resolve the polar vortex edge and the encircling filaments. The lack of resolution is palli- ated by using artificial devices grouped under the name of turbulent viscosity which tend to diffuse down the large- scale resolved gradients. There is no doubt that such proce-

3954 Phys. Fluids 6 (12), December 1994 1070-6631/94/6(12)/3954/9/$6.00 0 1994 American Institute of Physics

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Page 2: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

FIG. 1. The vorticity tield of a vortex submitted to a uniform adverse shear at a late stage after stripping has started; filaments surround the vortex and a sharp boundary develops as is shown in the vorticity cross section. This figure shows a part of the computational domain.

dures are necessary to perform calculations within the present state-of-the-art computers and models. The calcula- tions presented above shows, however, how far mixing can be from a down-gradient process ‘in reality.

We have seen that there is a pronounced tendency to generate high-vorticity gradients on the edge of vortices. This tendency persists under diffusive conditions, and one may expect that diffusion balances it and, therefore, that dif- fusion will be strongly enhanced across the edge of vortices. As a result, fast vortex-edge erosion may ensue. In the case of under-resolved structures, this enhanced diffusion shortens considerably the lifetime of atmospheric eddies in numerical models with respect to their counterparts in the real world. This effect has been demonstrated in a simple test problem beginning with two unequal, like-signed, distributed v0rtices.s The reference evolution was calculated using “contour surgery” (CS),’ a numerical method in which the small scales of the flow are much better taken into account than in conventional grid-based models. This reference evo- lution shows the development of an essentially discontinuous edge profile along the periphery of the smaller vortex which appears to persist indefinitely thereafter. For the present pur- pose, this calculation can be considered as effectively invis- cid. A number of pseudospectral (PS) calculations were per- formed of the same interaction, using ordinary (Newtonian) viscosity and hyperviscosity, and illustrate the rapid erosion and subsequent breakup of the smaller vortex at resolutions comparable to what is commonly used in studies of two- dimensional turbulence.‘”

The study presented here is devoted to a quantitative analysis of the influence of diffusion on the erosion of a vortex in a simple external straining field. We vary both the

strain and the level and type of viscosity. Section II presents the experiments and gives a preliminary qualitative discus- sion. In Sec. III, we study the quasi-adiabatic stage of strip- ping during which the vortex is slowly eroded while retain- ing an almost elliptical shape. Section IV discusses the final vortex break up and its universal features. In Sec. V, we show that hyperviscosity leads to spurious behavior during erosive stripping. Section VI offers a summary and further discussion on the efficiency of the diffusion-strain combina- tion in eroding coherent vortices. In addition, some remarks are given on the side effects produced by hyperviscosity in two-dimensional turbulence simulations.

II. QUASI-ADIABATIC EROSION

We model here the physical situation of a coherent ro- tating vortex entering a flow region dominated by a large- scale shear, defined by

u ,f=2Y(t)y; U,,t=O.

Unlike the example discussed in Sec. I, the external straining flow is not applied abruptly but, more realistically, y(t) grows slowly in time (for simplicity, linearly) from zero to a maximum value by a characteristic time 7 and then remains constant. An external flow of this type produces a spatially uniform shear in which the initially axisymmetric vortex is gradually embedded.

The inviscid case has been studied by Legras and Dritschel,s using contour dynamics to calculate the evolution of a discretized vortex in which the continuous profile is replaced by a series of finite jumps, an approximation which is known to give excellent results.4 It was shown that, pro- vided the strain grows slowly with respect to the vortex turn- over time, the evolution can be considered “quasi- adiabatic,” that is, the vortex remains in approximate equilibrium with the instantaneous value of the strain. A slight angular departure from the equilibrium shape is just enough to elongate the vortex in response to the increase in the strain. At any time, the edge of the vortex coincides, approximately, with the two intersecting separatrices of the streamfunction field, and stripping is produced at the tip of the vortex as the stagnation points slowly penetrate the inte- rior of the vortex. A steep gradient, which is numerically indistinguishable from an in&rite gradient, is produced on this boundary. Quite surprisingly, this mechanism is ame- nable to an analytical calculation using a simplified model, the “elliptical model,“‘* in which all the interior contours are taken as pure ellipses. Quasi-adiabatic evolution contin- ues until the strain reaches a critical “breaking” value, and then the vortex is suddenly torn apart.

It was shown in this and in other calculations’ that the critical breaking value depends on the ratio rotation/strain and on the way the strain is applied (abrupt or progressive), but not on the vorticity profile of the vortex (for the general class of vortices having a leading-order parabolic profile around the vorticity extremum; this class excludes the vortex patch). The critical strain is a fixed proportion of the maxi- mum vorticity (y=O.O67 w,, for pure adverse shear in the quasi-adiabatic casej. This result may be understood by not- ing that breaking is the final step of the erosion process and

Phys. Fluids, Vol. 6, No. 12, December 1994 Mariotti, Legras, and Dritschel 3955

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Page 3: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

occurs when only a small fraction of the initial profile re- mains near the center. This central region is generally well approximated by a parabolic profile, the size of which is irrelevant, and therefore the critical value is universal.

Here we study the same problem under the additional effect of viscous diffusion, modeled within the two- dimensional Navier-Stokes equations:

In all experiments but one, the initial vorticity profile is axi- symmetric and Gaussian:

2 o(r,t=O)=wo exp i 1 $ .

0

2 16

0

a 0

\

62

In the absence of external strain, the viscous decay of this vortex is self similar, i.e., its evolution is described by

t..“.----- I I I

(3) The other initial profile to be used in the remaining case is the so-called cosine profile: w(r,t=O)=$do X (1 +cos(~/uo)), for rCau.

erodes more quickly, as expected, and rapidly leads to break- ing by t=82, when it is overcome by the strain. As in the inviscid case, the breaking is preceded by a sudden rotation of the vortex towards the axis of extensional strain.

The external effects of strain and diffusion are measured by the strain parameter, q= y/w0 (with wo= 27r), and the Reynolds number, Re=aiwo/v. In all experiments, the ramping time is r= 30 and uo= l/a. The evolution of the system is calculated using the standard pseudo-spectral method with 128X 128 and 256X 256 spatial collocation grid points within a 27rX2rr doubly periodic box. The Fourier modes are truncated outside a circle (with radius 63 at reso- lution 128X128 and 127 at resolution 256X256), and there is no de-aliasing. The basic shear is made periodic in the y-direction by introducing an opposite intense shear in a thin boundary layer centered on y= t rr. Unless vorticity ex- pelled from the vortex hits this boundary layer, it has no dynamical effect. To stop vorticity filaments from passing through the periodic boundary at x= f r and re-interacting with the vortex, a sponge layer is introduced in the vicinity of x= + rr which absorbs any incoming vorticity (details may be found in the Appendix).

As has been anticipated, the smooth growth of the exter- nal straining field and its subsequent maintenance at a con- stant value induces, in a certain range of experimental pa- rameters, a slow, quasi-adiabatic evolution of the vortex. For ?= 0.05 and viscosity ~65 X 10m4 (Res6000), the vortex elongation closely follows the growth of the strain during the Initial stage and the subsequent evolution carries on smoothly without any significant wobbling of the vortex body.

III. DIFFUSIVE EROSION

To quantify the process of diffusive erosion, Fig. 4 shows, for several Reynolds numbers, the area bounded by selected vorticity contours as a function of time. The figure also shows the corresponding results for pure viscous decay described by Eq. (3), i.e., for no strain. The illustrated con- tours- are chosen within the vortex interior; the top curves

Figure 2 shows the case ?=0.05 and Re=8000 (v=3.7X10m4). The initial axisymmetric shape elongates in the y-direction, then two filaments emerge and later detach from the vortex at the location of the critical points of the streamfunction field and are swept away by the shear flow. The expulsion of the peripheral, low-vorticity layers is ac- companied by an intensification of the vorticity gradient on the edge of the vortex. Unlike in the inviscid case, the vortex cannot reach a final stationary equilibrium; instead, the high vorticity gradients induce an outward diffusive flux, thereby causing a continuous leakage of vorticity across the separa- trices. This vorticity is then carried rapidly towards the criti- cal points (in less than a characteristic vortex turnover time) where it feeds the expelled filament. Notably, while this is occurring, the vortex core retains an almost elliptical shape.

18 :) Q 8 0

66

We next examine what happens when the Reynolds number is reduced from 8000 to 3000 in Fig. 3. The vortex

FIG. 3. Evolution to breakup of the Gaussian vortex submitted to external adverse shear. Same as Fig. 2, except Re=3000.

FIG. 2. Quasr-adiabatic evolution of the Gaussian vortex submitted to ex- ternal adverse shear (strain parameter eO.O.5, Re=8000). This figure shows a 4.32X4.32 box centered in the 2rrX2rr computational domain. Times 2, 18, 34, 50, 66, and 82 as indicated in the frames.

3956 Phys. Fluids, Vol. 6, No. 12, December 1994 Mariotti, Legras, and Dritschel

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Page 4: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

0.0 1 --.. I 1 / I I I I

20 40 60 SD 100 120 140 Time

FIG. 4. Area within vorticity contours as a function of time for += 0.05 for various values of Re. Vorticity contours are o/0,=0.5 and 0.7. Light curves show the purely diffusive evolution described by Eq. (3). Heavy curves show the evolution in the presence of strain.

correspond to a contour just inside the vortex edge (0.5 wo), and the bottom curves correspond to a more interior contour (0.7 coo).

Initially, when the external strain is small, the dynamics of the internal regions is driven primarily by diffusion, and at this stage our experimental curves coincide with the pure viscous prediction of Eq. (3). As the strain approaches its final constant value (at t =30), the two families of curves diverge. The divergence becomes noticeable after t-20, when the external strain reaches two-thirds of its final value. During this stage, the erosion produced by the continually growing external strain increases the vorticity gradient at the edge of the vortex, thereby enhancing the outward diffusive transport of vorticity. Then there follows a phase of smooth, quasi-adiabatic evolution during which the area decays ap- proximately linearly in time. The peak vorticity also begins to decrease by this time, as seen in Fig. 5. The vorticity contours at the edge of the vortex are progressively and rap- idly eroded by the dual action of strain and diffusion; the strain readily advects away the low-level vorticity at the vor- tex periphery, low-level vorticity which is constantly being

T 20 40 50 60 100

Time

Voriicity contour -a- 0.5 I -o- 0.6 -a- 0.7 -cl- 0.6

Re

FIG. 6. Area decay rate as a function of Re calculated for various vorticity contours with i=O.OS.

regenerated by the outward diffusive flux. This action would eventually lead to the straining-out of the vortex if the ex- periment lasted long enough. Estimates of the area decay rate for selected vorticity contours, for the strain parameter +=o.os, and for various viscosities ranging from 1.3X 10m4 to 1.5X 10-s are reported in Fig. 6; increased diffusion clearly hastens the erosion rate.

The high vorticity gradients at the vortex edge clearly play a key role in the erosion process. The evolution of the maximum vorticity gradient (calculated on a cross section along the x axis) is shown in Fig. 7. During the early diffusion-dominated stage, the vorticity gradients remain weak. As soon as the strain becomes significant (around t = 20), the gradient grows rapidly to a peak value (this peak increases with decreasing viscosityj. The initial peak is formed as a direct effect of erosion; it is reached at the end of the strain growth and is limited by the outward diffusive flux of vorticity. As the vortex edge erodes, this flux is instanta- neously determined by the vorticity profile just inside the vortex edge (the contour associated with the maximum cross-sectional vorticity gradient) established prior to ero- sion. We do have to assume that, between the onset of ero- sion and the arrival at the peak vorticity gradient, diffusion

\\ \ I I I I I 1 1

20 40 60 80 100 120 140 Time

FIG. 5. Evolution of o,,,, for various values of Re with i=O.OS. Dotted FIG. 7. Maximum vorticity gradient as a function of time for various values line: the critical value of o,, for breakup. of Re and FO.05 calculated on the mid-transversal cross section.

Phys. Fluids, Vol. 6, No. 12, December 1994 Mariotti, Legras, and Dritschel 3957

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Page 5: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

has not had a chance to significantly modify the interior of the vortex. We must also assume that the vorticity which is transferred across the separatrices does not accumulate in the vicinity of the stagnation points and generate additional per- turbations of the vortex. That is, the viscosity cannot be too large.

In the subsequent stage of evolution, after the strain has stopped growing, the vortex edge can only retreat inward by diffusion, There is a rapid modification of the vorticity pro- file just inside the vortex edge and, since this profile deter- mines the diffusive flux, there is a rapid change in the diffu- sive flux (in fact, there are oscillations due to the delay in the diffusive response). The flux at fist increases strongly, thereby reducing the maximum vorticity gradient, and hence reducing the flux until an approximate balance is reached in which the maximum gradient remains roughly constant. We are then close to the inviscid situation in which the vortex reaches a final stationary state without further erosion. With viscosity, there is further erosion, and the interesting result is that this is accompanied by a growth of the maximum vor- ticity gradient. This apparently paradoxical behavior is actu- ally necessary to maintain the almost constant area decay rate shown in Fig. 4, since an increasing outward flux is required to compensate for the decreasing contour perimeter. The mechanism involved is exactly the same as that which produced the initial peak vorticity gradient. On the one hand, the steepness of the profile at the vortex edge determines the outward diffusive flux and thus the progression of the ero- sion towards the interior. On the other, the steepness is de- termined by the divergence of the diffusive flux, that is by the flux inside the edge which is itself determined by the interior profile of the vortex. Since for any smooth mono- tonic profile, the diffusive flux decreases as we approach the center, the maximum vorticity gradient should increase as diffusive erosion progresses. The largest value of the maxi- mum gradient is reached just prior to the final breaking, as shown on Fig. 7 for Res6000. During this stage, the small oscillations which are visible on the graph are due to a weak lateral wobbling of the vortex.

Briefly, consider the opposite parameter variation, namely fixing the Reynolds number (we illustrate Re=6000 here) and increasing the strain. Figure 8 shows results for 9 between 0.0125 and 0.05. The maximum gradient pro- duced at a given time increases with 9, which is sensible since the critical points of the combined vortex-external strain system penetrate deeper into the vortex and the advec- tive transport is larger. The production of stronger gradients enhances the erosion at a fixed value of viscosity, thus pro- ducing a faster decay of the vortex.

IV. VORTEX BREAKUP

Additionally, we have found that the profile of the vortex does not significantly influence breakup. For instance, in the decay of the cosine profile (reported in Fig. 5), we see the same maximum vorticity when the decay turns from linear to exponential as in the Gaussian cases. This is consistent with the inviscid results and with the argument in Sec. I.

We turn next to the conditions which lead to vortex While y/o,,= breakup. The inviscid calculations of Legras and Dritschel’

Jo, fixes the condition for breakup, at a given external strain, dissipation controls how fast the vortex

for a slowly increasing pure adverse shear showed that will break up. As the Reynolds number decreases, the life- breakup is determined uniquely by the strain to maximum time of the vortex shortens, as one would expect. It can be vorticity ratio r; that is, vortex breakup occurs for shown that for a given value of the strain, the lifetime varies r> Jo,= 0.067. We investigate here how breakup occurs in almost Iinearly with the initial Reynolds number. The life- the presence of diffusion. time is, however, dependent on the doubly periodic geom-

35

30 E .a, j$ 25

1; z 20 B

4 15

10

5

20 40 60 80 100 120 r40 Ttme

FIG. 8. Maximum vorticity gradient as a function of time for various values of F and Re=6000.

In the experiments, the strain parameter grows and then levels off at the value q= 0.05, which is less than $, . Unlike the inviscid case, the maximum vorticity o,, varies as a function of time. Its evolution, for Re ranging from 3000 to 1000, is reported in Fig. 5. The almost linear decay by dif- fusive erosion abruptly gives way to a fast exponential decay (at a time t, which depends on Re), leading to the complete destruction of the vortex. When the linear decay gives way to exponential, one may notice that the maximum vorticity is the same in all of the experiments. Taking the ratio -5/= y/a,,(tb) of the strain to the maximum vorticity at this time gives the value ~=0.0675+0.0025 over the range of experiments, a value which is more than coincidentally close to the inviscid value, ?,=0.067. The effects of the doubly periodic geometry can be shown to be negligible here (see the discussion in the Appendix). The same result has been found in the experiments with ;i=O.O45 and 0.0375. Hence, the mechanism for vortex breakup is an inviscid one.

Notice that the critical value $, is significantly smaller than the limit value y,=O.O855 obtained by Moore and Saffman12 (see also Ref. 13) for the existence of elliptical- patch equilibria in adverse shear. It is conjectured, on the basis of inviscid calculations and results from the elliptical model (cf. Ref. 5), that a distributed vortex with a monotonic vorticity profile cannot be eroded until its area vanishes (when y/w,,= r,) but must break at a finite area (when YI%ax= yb<yC) corresponding to the limiting state for equilibria with critical points on, or close to, the boundary.

3958 Phys. Fluids, Vol. 6, No. 12, December 1994 Mariotti, Legras, and Dritschel

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Page 6: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

fi 1 +-I T5 J “~,.A*,, .---... ..+._. . . ,_~._, pg-j

. .-.-.---.- ._._._._._._.-. _.__ ._,__.,_,_,_,_,_, _._. 2.0

f 1.5

ii

8

1.0

--~-‘-r-.~-...-.-- .-.-.-...-..... _ ._._.___._._,_, _ _,_,.____,_,_,_,_,_,-,-,-,-, _,_, 1-I--- *

0.7

‘t-.. _ . .-I*” _,_. _ .-.-.

h

-. I .,c..- -

\

l--J -

0.6 ;-' .L..c_.e .--...-._.-.-.-.-.-.-...~.~ ,_,___,_,_._._,_..,_, _._._. _._.

0.51 , , , ( , , ,

0 20 40 60 60 100 120 140 Time

~Hyperviscous ReI :

l-

ii0 260 so0 4ao Time

FIG. 9. Area within vorticity contours for the Gaussian vortex as a function of time using hyperviscnsity with ?=O.OS and orders p =2 and ~‘4. Re=L76XlO’ for p=2 and Re=1.91X10n for p=4. The vorticity con- tours shown are o/o,,=OS, 0.6, 0.7, and 0.8.

FIG. 10. Maximum vorticity versus time for ~=O.OS and for various values of the hyperviscous Re with p = 2.

etry, that is of the ratio between the radius of the vortex and the size of the box (see the discussion in the Appendix).

progressively earlier in time after an initial period of de- crease. The maximum vorticity later decreases again when the vortex enters the final decay stage.

It is clear that this “overshoot” depends on the value of

V. HYPERVISCOSITY EFFECTS

The hyperviscosity parametrization of unresolved turbu- lence consists of replacing the right hand side of Eq. (1) by the hyperdiffusion operator ( - 1)Pv,V2Po. This has the effect of compressing the enstrophy dissipation range rela- tive to that for normal diffusivity. This scheme has been used in many numerical simulations of two-dimensional turbu- lence and of large-scale atmospheric and oceanic flows. It is thus important to determine whether its use is harmless, or whether it may be able to alter results when a flow develops high vorticity gradients during stripping events. Additional results can be found in Ref. 8 as discussed in the introduc- tion.

We have studied the combined effect of strain and hy- perviscosity by employing the hyperdiffusion operator with p = 2 (i.e., bi-Laplacian) and p = 4 (quad-Laplacian) in oth- erwise identical simulations of vortex stripping. The Rey- nolds number for the flow is then defined as Re =a;%+, .

The vortex evolution is tist analyzed in terms of the area decay of selected vorticity contours; plots for both the p = 2 and p=4 cases are reported in Fig 9. A comparison with normal viscosity, see Fig. 4, reveals that hyperviscosity leads to spurious behavior: the areas do not regularly decrease. Signiiicant temporary area growth occurs for some of the contours, even for the innermost ones, and we find therefore that the evolution of the vortex is substantially different from that observed with normal viscosity. Contrary to physical sense, there is an inward “pumping” of vorticity. This in- ward pumping furthermore causes the vorticity near the cen- ter of the vortex to increase-see Fig. 10, which shows the maximum vorticity inside the vortex versus time and for various values of the hyperviscosity coefficients. As the hy- perviscous Reynolds number is decreased from 2.4X lo7 to 2.5X 104, the initial maximum vorticity value is exceeded

the hyperviscosity vp. For sufficiently large values of vp, one may avoid the overshoot altogether (the vortex rapidly decays). But, for small values of vr, (values which are how- ever still adequate to ensure numerical stability), the over- shoot can reach 12% of the initial maximum vorticity value. Paradoxically perhaps, the smaller the hyperdiffusion, the greater the overshoot.

To rule out the possibility that the observed overshoot could be due to the lack of numerical resolution, calculations on a 256X 256 grid were carried out (twice resolution in each direction) maintaining the previous values of the hypervis- cosity. At all but the smallest values of hyperviscosity, the above results are nearly perfectly reproduced. When the hy- perviscosity is too small, weak numerical instabilities are observed following the fist peak in the maximum vorticity gradient; however, as demonstrated in Fig. 11, these insta- bilities do not modify the gross evolution of the system.

1 I L f 50 100 150 200

Time

FIG. 11. Comparison of the evolution of the maximum vorticity using hy- perviscosity with p=2 and Re=9.9X104 at two different resolutions 128 Xl28 and 256x256.

Phys. Fluids, Vol. 6, No. 12, December 1994 Mariotti, Legras, and Dritschel 3959

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Page 7: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

35

30 mi

z 25 .; .’ s ,,’ : - 20 -’ ; .2 ----.

2 . . . . ..______._.

--- .-

g 1.5 _______------- __/---

I a 10

5

: 0 ._____

1 0 100 200 300 400

Time

FIG. 12. Maximum vorticity gradient as a function of time using hypervis- cosity with p = 2 and +O.OS for various values of the hyperviscous Re.

The maximum gradient calculated along the x-axis cross section evolves very much like in the case of normal viscos- ity, see Fig. 12 (compare with Fig. 7) except for the fact that the value of the first peak now depends strongly on the Rey- nolds number. The steep growth to the first peak is followed by a more gradual evolution towards the second peak, whose value is nearly constant for small hyperviscosity but rapidly increases for larger hyperviscosity. This ‘figure shows that the overshoot in maximum vorticity is not accompanied by a peculiar signature in the maximum gradient evolution. The strong dependence of the maximum vorticity gradient on the Reynolds number suggests that experiments using hypervis- cosity are much more prone to numerical instabilities than those using normal diffusion. In many cases, such as that shown in Fig. 11, these instabilities may damp out rapidly and remain unnoticed.

Insight into the cause of the overshoot can be obtained by looking at a mid cross section of the vortex profile, Fig. 13, at various key times in the evolution. This figure displays curves for t = 0, the initial condition, t = 28, during the first stage of the hyperviscous erosion characterized by decreas- ing maximum vorticity, t = 120, at the beginning of the stage of inward pumping characterized by a widening of the zone

FIG. 13. Mid-transversal cross section of the Gaussian vortex at different stages of its hyperviscous eroSion with F-0.05, p =2, and Re= 1.76X 10’.

l!IG. 14. Profile (i)p(s) of the self-similar solution to Eq. (4). Dotted line: p= 1 (normal diffusion); solid line: p =2; dashed line: p = 4.

of high vorticity gradients at the vortex edge, t = 200, at the end of the pumping stage when the overshoot has reached its maximum, and t =300, during the second stage of erosion and decreasing maximum vorticity in which the vorticity profile tends to collapse onto a spike-like structure before the final breaking.

The pumping effect responsible for the overshoot can be understood from a simple one-dimensional model based on the following equation:

do r?po -g=(-Wq+ ’ (4)

If we start from an initial step profile, the solution to Rq. (4) is self-similar in time, that is w(x,t)=OJxlu(t)) with a(t)=(2pt)l”J’ and

Ij,(s)=;+T/;~exp( - $)dk, (5)

Figure 14 shows GP(s) for p= 1, 2, and 4. It demonstrates that, unlike normal diffusion, hyperviscosity induces spuri- ous convergences and divergences near the zone of high vor- ticity gradients which act to pump vorticity across the edge, i.e., up-gradient. The number of bumps on the two sides of the step increases with p. Applied to the two-dimensional case of an eroding vortex core, this pumping leads to the growth of the vorticity at the vortex center in advance of the inward moving front of high vorticity gradients.

VI. DISCUSSION AND CONCLUSIONS

The stripping of a vortex by an external straining field causes very steep vorticity gradients to be formed on the vortex periphery, gradients which in turn enhance the out- ward diffusive flux across this boundary (under normal vis- cosity). By these means, stripping and diffusion combine ef- ficiently to accelerate the erosive decay of a vortex. For a strain growing from zero and saturating at T=O.OS, and for an initial Reynolds number larger than about 6000, the vor- tex exhibits a well-defined stage of quasi-adiabatic evolution during which the vortex core takes on a nearly stationary elliptical form and thin filaments slowly leak from the vicin- ity of the separatrix crossings. For lower values of the Rey-

3960 Phys. Fluids, Vol. 6, No. 12, December 1994 Mariotti, Legras, and Dritschel

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Page 8: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

nolds number, the quasi-equilibrium stage is skipped and the vortex rapidly succumbs to the external straining field.

Vortex breaking occurs when the maximum vorticity at the core of the vortex has been reduced to 14.5 times the value of the strain. This criterion does not depend on the initial profile nor on the Reynolds number (it does depend on the external straining flow being a pure adverse shear). The Reynolds number does however decide the lifetime .of the vortex. The lifetime is also dependent on the size of the vortex relative to the size of the doubly periodic box (see the Appendix).

As for atmospheric and oceanic models that use eddy viscosity to represent unresolved scales, the accelerated vor- tex decay due to the strain-viscosity combination implies a premature or simply inaccurate disappearance of coherent structures. It is tempting to propose that this phenomenon is one of the possible causes of the well-known tendency of atmospheric models to under-predict both the intensity and lifetime of intense cyclones or blocking anticyclones. This effect could be even more important than illustrated in the present work, since the usual number of points employed to resolve vortices is far less than used here.

Often the parametrization of subgrid-scale turbulence is performed with the hyperviscosity operator. Our calculations have shown that this operator produces serious numerical artifacts in the process of vortex erosion by stripping; namely, the artificial oscillation of the vorticity profile near the vortex edge; the inward, up-gradient pumping of vortic- ity, a consequent delayed growth in the maximum vorticity (overtaking the initial value by as much as 12%), and the formation of a spike-like profile at the vortex center (particu- larly at low resolution) preceding the final breakup of the vortex. Since breakup depends on the ratio of maximum vor- ticity to strain, this overshoot increases the lifetime of small vortices, an effect which may be beneficial in some circum- stances but is paid for by a lack of accuracy in the profile, which tends to exhibit a spike-like structure near its unre- solved center.

In turbulent flow simulations, in which hyperviscosity is usually preferred to simple viscosity, it is therefore possible that the maintenance of “spurious” small scales could have an undesirable side-effect on the fluid’s statistics. These re- sults, if not conclusive, at least question the use of hypervis- cosity and many results concerning two-dimensional (2-D) turbulence obtained by its use.

In a wider context, the observed effects associated with the use of dissipative operators, both of the viscosity and hyperviscosity types, to achieve numerical stability appear to introduce significant artificial changes to the dynamics of the system one wishes to model. The outlook should be towards the use of high resolution schemes and towards the develop- ment of alternative methods to control numerical instabili- ties.

ACKNOWLEDGMENTS

AM is supported by a fellowship of the Fondazione San Paolo di Torino. DGD is supported by the UK Natural Envi- ronmental Research Council (NERC). Additional support

was provided by the Innovative Science and Technology Pro- gram through the U.S. Naval Research Laboratory, by NATO and the European Community.

APPENDIX: SPONGE-LAYER AND DOUBLY PERIODIC GEOMETRY

The sponge-layer which absorbs the incoming filaments at the lateral boundary is obtained by relaxing the vorticity to a ground value in a small strip in the vicinity of x= ? v. At each time step, the vorticity field o(x,y) is multiplied by a function F(t), where 8 stands for x + rr near the left bound- ary and for rr - x near the right boundary, and F is defined by: F@)=O for @?r<O.OS, F@=1/2(1-cos((&rr-0.05)l 0.30)) for 0.05<yr<O.2, F(a=l for @rr>Od.

Introducing a sponge layer is not enough, however, to permit a direct comparison with the dynamics in an infinite plane. We must also take into account the presence of neigh- boring vortices in the adjacent cells as well as the fact that the average vorticity in the domain is automatically zero (the area integral of a Laplacian is necessarily zero). The first condition does not prove to be serious here since, owing to the second condition and the symmetry of our problem, the contribution of neighboring cells to the shear at the center of the vortex arises only at fourth order in an expansion where the small parameter is the ratio between the size of the vortex and the size of the box.

The second condition proves to be more serious. It means that we must subtract from the vorticity distribution [-defined by Eq. (2)] a uniform background rotation equal to half of the box-averaged vorticity of the vortex. This correc- tion does not modify the axisymmetric decay described by Eq. (3) but should be added ‘to the external imposed shear flow during the diffusive erosion process.

The sponge layer, as it relaxes vorticity to the ground value, modifies the average vorticity with respect to this ground value and in turn the background rotation. It turns out that the background rotation practically vanishes when the vortex undergoes final breaking because the average circula- tion of the vortex is then much smaller than initially. As a result, the breaking criterion for a pure shear in an infinite plane does not require any correction to fit the results shown in Fig. 5.

The breaking time, however, depends on the details of the evolution and is thus dependent on the sponge-layer. In other words, it is dependent on the ratio of the size of the vortex to the size of the box. We must recall that in the absence of a sponge-layer, the filaments reinteract with the vortex after crossing the boundary. The limit of a vortex in an infinite domain is only reached as the above ratio tends to zero. Effects due to the doubly periodic geometry are other- wise expected.

In order to compare different ratios we have modified the sponge layer by redistributing uniformly, for each value of x within the sponge layer, the vorticity which is destroyed. The average vorticity is then conserved by this artificial mecha- nism but the ground value, that is the vorticity value far from the vortex, is evolving with time.

Figure 15 shows a comparison of the erosion in two cases at different ratio. The first case is for a resolution

Phys. Fluids, Vol. 6, No. 12, December 1994 Mariotti, Legras, and Dritschel 3961

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Page 9: Vortex stripping and the erosion of coherent structures in ...dgd/papers/mld94.pdf · This paper studies the erosion of a monotonically distributed vortex by the joint action of inviscid

an excellent agreement between the two cases, which cor- roborates our analysis.

0 20 40 SO 80 100 120 140

Tiie

PIG. 15. Maximum vorticity versus time for q=O.OS, Re=3100 comparing two experiments within a standard 2~x2~ domain (resolution 128X128) and within a doubled 4~x4~ domain (resolution 256X256). See the text for the initial conditions.

128X 128 which is standard in this study, ;jl= 0.05 and Re =3100. The second case is obtained by doubling the size of the box (4 TX 4 rr now) and keeping the same resolution with a 256X256 grid. All other parameters are left identical. The initial vorticity distribution is given by Eq. (2) in the central 2rrX 27r region, where it is identical to the initial distribu- tion in the frrst case. The initial vorticity is set to a constant value, equal to the average within the central region, in the remaining part of the 45-X 49-r domain. In this way, we im- pose the same background rotation and the same maximum vorticity in the two cases. Figure 1.5 then shows that there is

‘D. Dritschel, “Strain-induced vortex stripping,” in Mathematical Aspects of Vortex Dynamics, edited by R. Caflisch (Society of Industrial and Ap- plied Mathematics, Philadelphia, PA, (1989j, p. 107.

‘D. Dritschel and B. Legras, “Modeling oceanic and atmospheric vortices,” Phys. Today 46, 44 (1993).

3S. Kida, “Motion of an elliptical vortex in a uniform shear flow,” J. Phys. Sot. Jpn. 50, 3517 (1981).

‘B. Legras and D. Dritschel, “A comparison of the contour surgery and pseudo-spectral methods,“J. Comput. Phys. 104, 287 (1993).

‘B. Lcgras and D. Dritschel, “Vortex stripping and the generation of high vorticity gradients in two-dimensional flows,” Appl. Sci. Res. 5l, 445 (1993).

6M. Melander, J. McWiIlliams, and N. Zabusky, “Axisymmetrization and vorticity-gradient intensification of an isolated two-dimensional vortex through filamentation,” J. Fluid Mech. 178, 137 (1987).

7D. Waugh, R. Plumb, R. Atkinson, M. Schoeberl, L. Lait, P. Newman, M. Lowenstein, D. Toohey, L. Avallone, C. Webster, and R. May, “Transport of material out of the stratospheric Arctic vortex by Rossby wave break- ing,” J. Geophys. Res. D 99, 1071 (1994).

‘H. Yao, D. Dritschel, and N. Zabusky, “The simulation of high gradients in 2-D vortex interactions,” to appear in Phys. Fluids (1994).

‘D. Dritschel, “Contour dynamics and contour surgery: numerical aigo- rithms for extended, high-resolution modelling of vortex dynamics in two- dimensional, inviscid, incompressible ~Iows,” Comput. Phys. Rep. 10, 77 (1989).

“J. McWilliams, “The vortices of two-dimensional turbulence,” J. Fluid Mech. 219, 361 (1990).

“B. Legras and D. Dritschel, “The elliptical model of two-dimensional vortex dynamics. Part I : The basic state,” Phys. Fluids A 3, 845 (1991).

“D. Moore and P. Saffman, “Structure of a line vortex in an imposed strain,” in Aircrafr TKzke L%rbulence and ifs Detection, edited by J. Olsen, A. Goldberg, and N. Rogers (Plenum, New York, 1971).

13P. Saffman, Vortex Dynamics (Cambridge University Press, Cambridge, 1992).

3962 Phys. Fluids, Vol. 6, No. 12, December 1994 Mariotti, Legras, and Dritschel

Downloaded 19 Jan 2009 to 138.251.95.1. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp