Three-Dimensional Stability of Burgers Vortices: the Low Reynolds Number Case.
Thierry Gallay Institut Fourier Université de Grenoble I BP 74 38402 Saint-Martin-d'Hères France
C. Eugene Wayne Department of Mathematics and Center for BioDynamics Boston University 111 Cummington St. Boston, MA 02215, USA
March 15, 2005
Abstract
In this paper we establish rigorously that the family of Burgers vortices of the three-dimensional Navier-Stokes equation is stable for small Reynolds numbers. More precisely, we prove that any solution whose initial condition is a small perturbation of a Burgers vortex will converge toward another Burgers vortex as time goes to infinity, and we give an explicit formula for computing the change in the circulation number (which characterizes the limiting vortex completely.) We also give a rigorous proof of the existence and stability of non-axisymmetric Burgers vortices provided the Reynolds number is sufficiently small, depending on the asymmetry parameter.
1 Introduction
Numerical simulations of turbulent flows have lead to the general conclusion that vortex tubes serve as important organizing structures for such flows – in the memorable phrase of [9] they form the “sinews of turbulence”. After the discovery by Burgers [1] of the explicit vortex solutions of the three-dimensional Navier-Stokes equation which now bear his name, these solutions have been used to model various aspects of turbulent flows [17] . It was also observed in numerical computations of fluid flows that the vortex tubes present in these simulations usually did not exhibit the axial symmetry of the explicit Burgers solution, but rather an elliptical core region. This lead to a search for non-axisymmetric vortices [13] , [9] , [6] . While no rigorous proof of their existence was available until recently, perturbative calculations and extensive numerical simulations have lead to the expectation that stationary vortical solutions of the three-dimensional Navier-Stokes equation do exist for any Reynolds number and all values of the asymmetry parameter (which we define below) between zero and one.
When addressing the stability of Burgers vortices, it is very important to specify the class of allowed perturbations. If we consider just two-dimensional perturbations (i.e., perturbations which do not depend on the axial variable), then fairly complete answers are known. Robinson and Saffman [13] computed perturbatively the eigenvalues of the linearized operator at the Burgers vortex and proved its stability for sufficiently small Reynolds numbers. Numerical computations of these eigenvalues were performed by Prochazka and Pullin [10] , and no instability was found up to
. A similar conclusion was drawn for non-symmetric vortices [11] . The first mathematical work is [5] , where we proved that the axisymmetric Burgers vortex is globally stable with respect to integrable, two-dimensional perturbations, for any value of the Reynolds number. Decay rates in time of spatially localized perturbations were also computed, explaining partially the numerical results of [10] . Building on this work the existence and local stability of slightly asymmetric vortices with respect to two dimensional perturbations was proved in [4] for arbitrary Reynolds numbers.
The stability issue is much more difficult if we allow for perturbations which depend on the axial variable too, and very few results have been obtained so far in this truly three-dimensional case. One early study by Leibovich and Holmes [7] concluded that one could not prove global stability for any Reynolds number solely by means of energy methods. Using a kind of Fourier expansion in the axial variable, Rossi and Le Dizès [14] showed that the point spectrum of the linearized operator is associated with purely two-dimensional perturbations. Crowdy [2] obtained a formal asymptotic expansion of the eigenfunctions in the axial variable. In an important recent work, Schmid and Rossi [16] rewrote the linearized equations in a form which allowed them to compute numerically the evolution of various Fourier modes, from which they concluded that eventually all perturbative modes will be damped out.
In this paper we address rigorously the existence of non-axisymmetric vortices and the stability with respect to three-dimensional perturbations of both the symmetric and non-symmetric vortex solutions. More precisely we will prove that, for all values of the asymmetry parameter between zero and one, non-axisymmetric vortices exist at least for small Reynolds numbers. In addition, we show that this family of vortex solutions is, in the language of dynamical systems theory, asymptotically stable with shift. That is to say, if we take initial conditions that are small perturbations of a vortex solution, the resultingsolution of the Navier-Stokes equation will converge toward a vortex solution, but not, in general, the one which we initially perturbed. We also give a formula for computing the limiting vortex toward which the solution converges.
We now state our results more precisely. The three-dimensional Navier-Stokes equations for an incompressible fluid with constant density
and kinematic viscosity
are the partial differential equations:
|
(1)
|
Here
is the velocity of the fluid and
its pressure. Equation ( 1 ) will be considered in the whole space
. Burgers vortices are particular solutions of ( 1 ) which are perturbations of the background straining flow
|
(2)
|
where
are real constants satisfying
. We restrict ourselves to the case of an axial strain aligned with the vertical axis, namely we assume
and
. Setting
, we obtain the following evolution equation for the vorticity
:
|
(3)
|
Under reasonable assumptions which will be satisfied for the solutions we consider, the rotational part
of the velocity can be recovered from the vorticity
by means of the Biot-Savart law:
|
(4)
|
For notational simplicity we begin by discussing the axisymmetric case where
. In this situation, it is well-known [1] that ( 3 ) has a family of explicit stationary solutions of the form
, where
is a parameter and
|
(5)
|
Here
,
, and
. The velocity field corresponding to
is
, where
|
(6)
|
These solutions are called the axisymmetric Burgers vortices. Observe that
, so that the parameter
represents the circulation of the velocity field
at infinity (in the horizontal plane
). Following [9] , we define the Reynolds number associated with the Burgers vortex
as
Our principal result concerns the evolution of solutions of ( 3 ) with initial conditions that are close to a Burgers vortex. Unlike in much previous work the perturbations we consider do not merely depend on the transverse variables
, but also on
. We prove that any solution of ( 3 ) starting sufficiently close to the Burgers vortex with circulation
converges as
toward a Burgers vortex with circulation
close to
, and we give an explicit formula for computing the difference
in terms of the initial perturbation.
We now introduce some function spaces to measure the size of our perturbations.
Roughly speaking, we require the perturbations to decay as inverse powers of
as
, but need only boundedness in
. To be specific, we use the following weight function
|
(8)
|
Given any
, we define
, where
|
(9)
|
In other words, a function
belongs to
if and only if
,
,
, and
are square integrable over
. For later use, we observe that
is continuously embedded into
if
, i.e. there exists
such that
for all
.
Our main space
will be the set of all
such that
for all
, and such that the map
is bounded and continuous from
into
. As is easily verified,
is a Banach space equipped with the norm
|
(10)
|
Remark 1.1
If
is a vector field whose components are elements of
, we shall often write
instead of
, and
instead of
. A similar abuse of notation will occur for other function spaces too.
Consider initial conditions for the vorticity equation ( 3 ) which are a perturbation of the Burgers vortex:
|
(11)
|
with
and
. Define
|
(12)
|
Just to make sure the notation is clear, in the integrand
refers to the third component of the initial perturbation
.
Theorem 1.2
Fix
, and assume that
. For any
, there exist
and
such that if
and
, then the solution
of ( 3 ) with initial condition ( 11 ) converges as
tends to infinity to the Burgers vortex with circulation number
, where
is given by ( 12 ).
Convergence is with respect to the
norm in
and uniformly on compact sets in
. More explicitly, if
is any compact interval we have
|
(13)
|
Remark 1.3
Here and in the sequel, all constants are independent of the strain
and the viscosity
. In fact, both parameters will shortly be eliminated by a rescaling.
The proof of Theorem 1.2 uses ideas from our analysis of the stability of the two-dimensional Oseen vortices in [5] . The main observation is that, if we linearize equation ( 3 ) at the Burgers vortex
for small
, we obtain a small perturbation of a non-constant coefficient differential operator for which we can explicitly compute an integral representation of the associated semigroup. This semigroup decays exponentially when acting on functions
provided
, where
|
(14)
|
Thus an important step in the proof consists in decomposing the perturbation as
, where
By construction
, hence
will decay exponentially to zero by the remark above. Now, the crucial point is that
satisfies the amazingly simple equation
which can be solved explicitly, see ( 51 ) below. From the solution formula we see that
converges uniformly on compact sets to the value
as
, and ( 13 ) follows.
In other words,
is a “zero mode” which is responsible for the fact that the family of Burgers vortices is only asymptotically stable with shift.
While the axisymmetric vortex solution has been extensively studied because of the explicit formulas for its velocity and vorticity fields, numerical experiments on turbulent flows seem to indicate that the vortex tubes that are prominent in these flows are not symmetric, but rather elliptical in cross section. A natural way to obtain such vortices is to assume that the straining flow is not axisymmetric, namely
|
(15)
|
where
and
is an additional parameter which measures the asymmetry of the strain. While no explicit formulas for the vortex are known when
, extensive perturbative and numerical investigations indicate that there do exist stationary solutions of ( 3 ) for
, which for
close to zero are small perturbations of the vorticity field of the axisymmetric Burgers vortex [13] , [9] , [11] . As in the symmetric case, there is in fact a family of vortices for each value of
parametrized by the total circulation
, but when
these solutions are not just multiples of one another.
In Section 2 we give a simple but rigorous proof of the existence of these non-axisymmetric vortex solutions for all values of
, provided the circulation Reynolds number of the flow is sufficiently small (depending on
). A complementary result is obtained in [4] where we prove that, if
is sufficiently small, non-axisymmetric vortex solutions exist for all values of the Reynolds number.
The construction of these non-axisymmetric vortices requires some work, so as a first step we rewrite the Navier-Stokes and associated vorticity equation in non-dimensional form. This simplifies the expressions for the solutions and also the equations themselves.
Fix
and assume that the straining flow is given by ( 2 ), ( 15 ) for some
. We replace the variables
and the functions
with the dimensionless quantities
|
(16)
|
where
is the kinematic viscosity. Dropping the tildes for simplicity, we see that the new functions
satisfy the Navier-Stokes equation ( 1 ) with
. Similarly the new straining flow
is given by ( 2 ), ( 15 ) with
. Setting
, we obtain that
satisfies ( 3 ) with
, namely
|
(17)
|
Thus the main effect of the change of variables ( 16 ) is to set
everywhere.
In particular, in the dimensionless variables the weight function ( 8 ) becomes
, and the norm ( 9 ) reduces to
.
To formulate our result, we define
|
(18)
|
If
,
is just the vorticity field ( 5 ) of the symmetric Burgers vortex written in the new coordinates, and the family of these vortices is indexed by the non-dimensionalized circulation number
. As we show below, for any
,
is still the leading order approximation to the vorticity of the non-axisymmetric Burgers vortex, for small Reynolds number
. Our precise result is:
Theorem 1.4
Fix
,
, and assume that
is given by ( 15 ) with
. There exist
and
such that, for
, the vorticity equation ( 17 ) has a stationary solution
which satisfies
|
(19)
|
and
|
(20)
|
Furthermore,
is a smooth function of
and
, and there is no other stationary solution of ( 3 ) of the form ( 19 ) satisfying
.
Remark 1.5
The proof shows that
and
as
. On the other hand,
and
as
. In particular, setting
in ( 20 ), we recover that
.
Remark 1.6
Theorem 2 shows that the asymmetric Burgers vortex
decays rapidly as
, since the parameter
is arbitrary (note, however, that the constants
depend on
). In fact, proceeding as in [
4]
, it is possible to show that
has a Gaussian decay as
. Moreover,
is also a smooth function of
, see Remark 2.3 below.
Finally, we prove that these families of non-symmetric vortices are stable in the same sense as the symmetric Burgers vortices are.
Theorem 1.7
Fix
,
, and assume that
is given by ( 15 ) with
. For any
, there exist
and
such that, if
and if
with
satisfying
|
(21)
|
then the solution
of ( 17 ) with initial data
converges as
to the vortex solution
, where
and
, see ( 12 ).
More precisely, for any compact interval
, we have
|
(22)
|
The symmetric case
is included in Theorem 1.7 , which therefore subsumes Theorem 1.2 . Note however that the assumptions on the initial data are more restrictive when
, because we then need a condition on
. This is due to the fact that non-axisymmetric Burgers vortices with different circulation numbers are not multiples of one another.
The rest of the text is organized as follows. In Section 2 , we prove the existence of non-axisymmetric Burgers vortices for small Reynolds numbers. The core of the paper is Section 3 , where we show that these families of vortices are asymptotically stable with shift. Section 4 is an appendix where we collect various estimates on the semigroup associated to the linearized vorticity equation, together with a few remarks concerning the Biot-Savart law.
2 Existence of non-axisymmetric Burgers vortices
The properties of non-axisymmetric Burgers vortices seem first to have been studied by Robinson and Saffman [13] who used perturbative methods to investigate their existence for small values of the Reynolds number. There were many further investigations in the intervening years we mention particularly the perturbative study of the large Reynolds number limit of these vortices by Moffatt, Kida and Ohkitani [9] , and the numerical work of Prochazka and Pullin [11] . However, as far as we know there has been no rigorous proof of the existence of these types of solutions and so in this section we present a simple argument which proves the existence of non-symmetric vortices in the case of small Reynolds number.
Fix
and assume that
are given by ( 15 ) with
. Motivated by the perturbative calculations of [13] we look for stationary solutions of ( 17 ) of the form
for some
(recall that
is the Reynolds number). Since
depends only on the horizontal variable
and has only the third component nonzero, the associated velocity field
depends only on
and has only the first two components nonzero. Thus
is naturally identified with a two-dimensional velocity field
which can be computed using the two-dimensional version of the Biot-Savart law:
|
(23)
|
Inserting these expressions into ( 17 ), we see that
satisfies the scalar equation
|
(24)
|
where
and
are the differential operators
|
(25)
|
Here we have used the natural notations
and
.
We shall solve ( 24 ) in the weighted space
defined by ( 9 ) (with
). Our approach rests on the fact that the spectrum of the linear operator
in
can be explicitly computed, see Section 4.2 . If
, this operator turns out to be invertible on the invariant subspace
defined by
|
(26)
|
This allows to rewrite ( 24 ) as a fixed point problem which is easily solved by a contraction argument.
As a preliminary step, let
be the two-dimensional velocity field obtained from
by the Biot-Savart law ( 23 ). Using ( 18 ) and ( 25 ) one can easily verify that
If we are given
with
and if
, we can decompose
|
(27)
|
where
and
is the velocity obtained from
by the Biot-Savart law ( 23 ).
With these notations, finding a solution to ( 24 ) is equivalent to solving
|
(28)
|
Note that
since
and
are divergence-free. Thus the right-hand side of the ( 28 ) has zero mean as expected.
The next proposition ensures that the operator
is invertible on
and that
defines a bounded operator from
into
, where
is the weighted space defined in ( 72 ).
Proposition 2.1
Fix
and
. There exists
such that, for all
,
|
(29)
|
Moreover, if
, there exists
such that, for all
,
|
(30)
|
Proof: Let
denote the strongly continuous semigroup generated by
, the properties of which are studied in the Section 4.2 . If
, we know from ( 73 ) that
|
(31)
|
Thus
is invertible on
and we have the Laplace formula
|
(32)
|
Combining ( 31 ), ( 32 ), we easily obtain ( 29 ). Assume now that
for some
and some
. Using ( 32 ) and ( 74 ) or ( 75 ), we obtain an estimate of the form
where
. Since
, the singularity in the integral at
is integrable and ( 30 ) follows.
We can now rewrite ( 28 ) as
, where
is defined by
|
(33)
|
For any
, let
denote the closed ball of radius
centered at the origin in
. The main result of this section is:
Proposition 2.2
Fix
and
. There exist
and
such that, if
, then
has a unique fixed point
in
. Moreover
is contained in
and
is a smooth function of both
and
.
Proof: Let
be the bilinear map defined by
where
is the velocity field obtained from
by the Biot-Savart law ( 23 ). If
, then
by Corollary 4.5 . Using in addition ( 30 ), we see that there exists
such that
Since
, we obtain, for all
,
|
(34)
|
where
and
. Similarly, for all
,
|
(35)
|
When
,
is radially symmetric and
is azimuthal, hence
. Thus
, so that
as
.
Now, choose
sufficiently small so that
If
and
, estimates ( 34 ) and ( 35 ) imply that
maps the ball
into itself and is a strict contraction there. More precisely, if
, then
By the contraction mapping theorem,
has a unique fixed point
in
.
Choosing
, we obtain the existence and uniqueness claim in Proposition 2.2 . Then setting
with
, we see that
. Finally, the smoothness property is a immediate consequence of the implicit function theorem. Indeed, the map
is obviously
from
into
, and the partial differential
satisfies
whenever
and
. Thus
is invertible at
, and the implicit function theorem implies that
is a smooth function of both
and
.
Theorem 1.4 is a direct consequence of Proposition 2.2 . Indeed, if
, we set
, where
is as in Proposition 2.2 , and we denote by
the two-dimensional velocity field obtained from
by the Biot-Savart law ( 23 ). Then
|
(36)
|
is a stationary solution of ( 17 ) which has all the desired properties. In particular, since
, we have
|
(37)
|
while the fact that
implies that ( 20 ) holds. For later use, we observe that there exists
such that, for
,
|
(38)
|
Moreover
, because in the symmetric case
so that
.
Remark 2.3
We chose to solve ( 24 ) in
because this is basically the space we shall use in Section 3 to study the stability of the vortices. But it is clear from the proof of Proposition 2.2 that nothing important changes if we replace
with the corresponding Sobolev space
for any
. This shows that the asymmetric Burgers vortex
is a smooth function of
too. In particular, by Sobolev embedding,
(the space of all continuous and bounded functions on
) and we have the analogue of ( 38 ):
|
(39)
|
Moreover, since
for all
, Proposition 4.4 implies that
for all
, and there exists
such that
|
(40)
|
3 Stability with respect to three-dimensional perturbations
We now prove that the family of vortices constructed in the previous section is asymptotically stable with shift, provided the circulation Reynolds number is sufficiently small, depending on the asymmetry parameter
. In particular, our result applies to the classical family of symmetric Burgers vortices (
).
Throughout this section we fix some
. For
sufficiently small we denote by
,
the asymmetric vortex ( 36 ) with total circulation
(to simplify the notation, we omit the dependence on
). As we mentioned in the introduction, if we slightly perturb the vortex
the solution of the vorticity equation will converge toward another vortex with a possibly different circulation. This means that we must allow the parameter
to depend on time. Also, since the perturbations we consider may depend on the axial variable
, it turns out to be convenient to approximate the solutions by vortices with different circulation numbers in different
sections. In other words, we will consider solutions of ( 17 ) of the form
|
(41)
|
where
is determined so that
for all
and
. In view of ( 37 ), it is obvious that any pertubation of
that is integrable with respect to the transverse variables
can be decomposed in a unique way as in ( 41 ). Similarly, we write the rotational part of the velocity field as
|
(42)
|
where
is the velocity field obtained from the vorticity
by the Biot-Savart law ( 4 ). It will be shown in Proposition 4.11 that
is a small perturbation of
if
varies slowly in the
direction, namely there exists
such that
|
(43)
|
Let
and
denote the remainder terms in ( 41 ) and ( 42 ) respectively. By construction,
is the velocity field obtained from
by the Biot-Savart law ( 4 ). Remark that
, but
, hence
.
In broadest terms, our strategy is to show that
and
converge to zero as time tends to infinity, so that the vorticity
approaches one of the vortices
constructed in Section 2 . With that in mind, we now write out the evolution equations for
and
.
Inserting ( 41 ), ( 42 ) into ( 17 ) and using the identity
, we obtain after straightforward calculations:
|
(44)
|
where the various terms in the right-hand side are defined as follows.
The linear operator
is the leading order part of the equation, which takes into account the diffusion and the effects of the background strain:
Here
are given by ( 15 ) with
, and
|
(45)
|
The term
describes the linear interaction between the perturbation and the modulated vortex, namely:
|
(46)
|
Here and in the sequel, to simplify the notation, we write
instead of
and
instead of
.
The term
collects all the nonlinear contributions in
, specifically:
|
(47)
|
Finally,
is an inhomogeneous term which is due to the fact that
fails to be a solution of ( 17 ) if
is not identically constant.
A simple calculation shows that
for
. The third component of
has a more complicated expression:
| |
|
(48)
|
where
and
are defined in ( 25 ).
Equation ( 44 ) governs the evolution of both
and
. To separate out the evolution equation for
, we recall that
satisfies the constraint
. If we integrate the third component of the vectorial equation ( 44 ) with respect to the transverse variables
, the first three terms in the right-hand side give no contribution, as can be seen from the formulas ( 45 ), ( 46 ), ( 47 ). So we must impose
|
(49)
|
As is clear from ( 24 ), the first term in the right-hand side of ( 48 ) has zero mean with respect to
, and so does the second term because it is explicitly in divergence form. On the other hand, differentiating ( 37 ) with respect to
, we obtain the identities
and
. Thus ( 49 ) gives the evolution equation for
:
|
(50)
|
Remarkably, this equation is linear and completely decoupled from the rest of the system. As is easily verified, the solution of ( 50 ) with initial data
is given by the explicit formula
|
(51)
|
where
|
(52)
|
The following simple estimates will be useful:
Proposition 3.1
If
, the solution of ( 50 ) with initial data
satisfies
|
(53)
|
If moreover
, we also have
|
(54)
|
Proof: Since
, it follows immediately from ( 51 ) that
. If
, the same argument gives ( 54 ), because
|
(55)
|
To prove the second estimate in ( 53 ), we use the last expression in ( 55 ) and observe that
, where
.
Remark 3.2
Proposition 3.1 shows in particular that ( 51 ) defines a semigroup of bounded linear operators on
, the space of all bounded and continuous functions on
equipped with the
norm. It is easy to verify that this semigroup is not strongly continuous in time, due to the dilation factor
in ( 51 ) which in turn originates in the unbounded advection term
in ( 50 ). However, if we equip
with the (weaker) topology of uniform convergence on compact sets, then ( 51 ) defines a continuous function of time.
We now return to the evolution equation for
. Using ( 48 ), equation ( 50 ) for
, and equation ( 24 ) satisfied by the asymmetric vortex
, we obtain for the inhomogeneous term
the simpler expression
|
(56)
|
where as usual
. Before starting the rigorous analysis, let us briefly comment here on why we expect solutions of ( 44 ) to go to zero as
goes to infinity.
Given
, we assume that
for
, where
is the space defined in ( 10 ). By construction,
then belongs to the subspace
given by ( 14 ).
As we show in Section 4.3 , the linear operator
has spectrum that lies in the half-plane
when acting on
. Thus, the semigroup generated by this operator can be expected to decay like
. The remaining linear terms in the equation, namely
, contain a factor of the vortex solution which is proportional to
(see ( 39 ) and ( 100 )) and hence, for small Reynolds number, they will be a small perturbation of
and will not destroy the exponential decay. The same is true for the nonlinear terms
provided we restrict ourselves to sufficiently small perturbations. Finally, the inhomogeneous term
decays at least like
by ( 43 ) and Proposition 3.1 , so we expect the solution
of ( 44 ) to converge exponentially to zero if the initial data are sufficiently small.
We now put these heuristic arguments into a rigorous form. Let
be the Banach space
equipped with the norm
. As is shown in Proposition 4.6 , the linear operator
is the generator of a semigroup
of bounded operators on
, hence the same is true for the operator
acting on
. A natural idea is then to use Duhamel's formula to rewrite ( 44 ) as an integral equation:
|
(57)
|
which can then be solved by a fixed point argument. A problem with this approach is that the semigroup
fails to be strongly continuous on
, essentially for the reason mentioned in Remark 3.2 . To restore continuity, it is thus necessary to equip
with a weaker topology. For any
we define the seminorm
|
(58)
|
and we denote by
the space
equipped with the topology defined by the family of seminorms ( 58 ) for
, i.e. the topology of the Fréchet space
.
In other words, a sequence
converges to zero in
if and only if
as
for all
, namely if
converges to zero in
uniformly on compact sets in
. We define the product space
in a similar way. Then Proposition 4.6 shows that the semigroup
is strongly continuous on
, and the integrals in ( 57 ) can be defined as
-valued Riemann integrals, see Corollary 4.7 and Remark 4.8 .
Since we expect
to converge exponentially to zero as
, we shall solve ( 57 ) in the Banach space
for some
, where
Given initial data
and
, we first define
by ( 51 ), and then use the integral equation ( 57 ) to determine
for all
. Our main result is:
Proposition 3.3
Fix
,
, and
. There exist positive constants
,
, and
such that, if
,
, and if
satisfies
, then for all
with
equation ( 57 ) has a unique solution
with
.
Proof: Fix
,
, and
. To simplify the notations, we shall write
instead of
and
instead of
. For any
, we denote by
(respectively,
) the closed ball of radius
centered at the origin in
(respectively,
). Let
and denote by
the solution of ( 50 ) with initial data
. Given
,
, and
, we estimate the various terms in the right-hand side of ( 57 ).
We begin with the linear term
. From Proposition 4.6 we know that the linear operator
generates a semigroup
which is strongly continuous on
, uniformly bounded on
, and which decays like
on
.
Since
where
are given by ( 15 ) with
, we deduce that
is continuous in
and satisfies
| |
|
(59)
|
Note that it is crucial here that
, otherwise we do not get any decay at all.
We next consider the linear term
. For
and
, we know from Proposition 4.11 that
and
by Proposition 3.1 . Since
for
, it follows that
and
where
. Moreover, it is not difficult to verify that
is continuous in
. Similarly,
for
by Proposition 4.9 and
for
by Remark 2.3 , hence
by Hölder's inequality and
Again
is continuous in
. Thus Corollary 4.7 and Remark 4.8 show that the three components of the vector
are well defined and continuous in
for
. Using Proposition 4.6 , we can estimate the first component as follows:
| |
| |
| |
where
. (This estimate could be sharpened somewhat by using the functions
,
, and
which appear in Proposition 4.6 , but they would lead to no qualitative improvement in the final result and so we use this somewhat simpler form.) The other two components can be estimated in exactly the same way except for a slower exponential decay of the linear semigroup, see ( 59 ). Summarizing, we obtain:
|
(60)
|
We now consider the nonlinear term
. Let
. For
and
, we know from Corollary 4.10 that
with
Moreover
is continuous in
. Thus, by Remark 4.8 , the integral
is well defined and continuous in
for
. Proceeding as above we can estimate the first component as follows:
| |
| |
| |
Repeating the same arguments for the other two components, we thus find
|
(61)
|
Finally, we turn our attention to the inhomogeneous term
. To bound the first two components, we observe that
with
|
(62)
|
Indeed,
. Since
, one has
, hence by ( 38 ), ( 100 )
Moreover, as
is the velocity field obtained from
by the Biot-Savart law, the proof of Proposition 4.11 shows that
, hence
which proves ( 62 ). As usual, one checks that
is continuous in
for
. Using ( 53 ), ( 62 ), the first component of the inhomogeneous term can be estimated as follows:
| |
| |
and the second one is bounded in exactly the same way. To bound the third component, we first remark that
belongs to
for
and depends continuously on
in
. By ( 38 ), ( 43 ), ( 53 ),
hence using ( 53 ) we find
| |
| |
On the other hand,
lies in
and depends continuously on
in
. As was mentioned before Remark 2.3 ,
. If
, we assume that
and using ( 54 ) we obtain
| |
| |
Thus we have shown that
is well defined and continuous in
for
. Moreover,
|
(63)
|
Given
, we denote by
the right-hand side of ( 57 ). Estimates ( 59 ), ( 60 ), ( 61 ), ( 63 ) show that
. Thus
maps
into itself and
|
(64)
|
where
. Moreover, if
, the same estimates show that
|
(65)
|
because the linear term
and the inhomogeneous term depending on
drop out when we consider the difference
. Now, choose
and
small enough so that
where
is as in Proposition 2.2 . Assume that
,
,
, and
. If
, then ( 64 ) shows that
maps the ball
into itself. Indeed, under the assumptions above we have
and
, hence if
then ( 64 ) implies
Similarly,
if
. By the contraction mapping theorem,
has thus a unique fixed point
in
. Choosing
with
, we see that
is the unique solution of ( 57 ) such that
.
Theorem 1.7 is a direct consequence of Proposition 3.3 . Indeed, suppose that the initial condition for the vorticity is
, where
satisfies ( 21 ). Then we can decompose
, where
is as in ( 21 ) and
belongs to
, namely
. Moreover, there exists
such that
and so the smallness conditions on the perturbation in Theorem 1.7 imply those in Proposition 3.3 . We deduce that the solution of ( 17 ) with initial data
satisfies
for some
, hence
converges exponentially in
toward the modulated vortex
. On the other hand, from ( 51 ), ( 52 ), we see that, for any
,
converges toward the limiting value
and that
for any compact interval
. Thus the difference
will converge exponentially to zero as
, uniformly for
in any compact interval. Combining these estimates, we obtain ( 22 ).
4 Appendix
In this appendix we collect a number of technical estimates used in the main body of the paper. They relate mostly to the behavior of the semigroup generated by the linearization of the vorticity equation around the Burgers vortex. We also prove some estimates relating the vorticity field to the corresponding velocity field defined by the Biot-Savart law.
4.1 The one-dimensional Fokker-Planck operator
Fix
, and consider the one-dimensional linear equation
|
(66)
|
where
and
. If
, then
hence we could assume without loss of generality that
. However for our purposes it is more convenient to keep
arbitrary.
The linear operator
is formally conjugated to the Hamiltonian of the harmonic operator in quantum mechanics:
As is well-known, the spectrum of
in
is a sequence of simple eigenvalues:
and the associated eigenfunctions are the Hermite functions
.
This observation, however, is not sufficient to determine the whole spectrum of
because we want to consider this operator acting on a space of functions with algebraic (rather than Gaussian) decay at infinity.
For any
and
we define the space
, where
. This Banach space is equipped with the natural norm
The parameter
determines the decay rate at infinity of functions in
. For instance, it is easy to verify that
if (and only if )
, because in that case
so that any
satisfies
|
(67)
|
by Hölder's inequality. If
, we thus define
This closed subspace of
is clearly invariant under the evolution defined by ( 66 ).
In ([3] , Appendix A) it is shown that the spectrum of
in
is
|
(68)
|
Thus, in addition to the discrete spectrum of the harmonic oscillator, the operator
also has essential spectrum due to the slow spatial decay of functions in
. Note however that this essential spectrum can be pushed far away from the imaginary axis by taking
sufficiently large. Therefore, if
is large, the relevant part of the spectrum of
is still given by the first few eigenvalues of
. In particular zero is an isolated eigenvalue of
if
, and the rest of the spectrum is strictly contained in the left-half plane. If we restrict ourselves to the invariant subspace
, the spectrum of
is unchanged except for the zero eigenvalue (which is absent).
Equation ( 66 ) can be explicitly solved as
, where
and
. Using this expression, it is straightforward to verify that
defines a strongly continuous semigroup in
for any
. Moreover,
maps
into
if
, and the following estimates hold (see [3] , Appendix A):
Proposition 4.1
If
, the semigroup
is uniformly bounded in
for all
. Moreover, if
, there exists
such that, for all
,
|
(69)
|
Finally, if
and
, then
defines a bounded operator from
into
and there exists
such that, for all
,
|
(70)
|
where
.
4.2 Two-dimensional estimates
We next consider the two-dimensional equation
|
(71)
|
where
,
, and
. In the particular case where
and
for some
, we see that
, where
are defined in ( 25 ). Note that the parameters
cannot be eliminated by a rescaling, unless
.
We study the operator
in the weighted space
|
(72)
|
where
,
, and
. It is clear that
, where
is the one-dimensional space defined in the previous paragraph and
denotes the tensor product of Hilbert spaces, see [12] .
Comparing the definitions ( 66 ), ( 71 ), we see that our operator can be decomposed as
, where
denotes the identity operator. It follows that the spectrum of
in
is just the sum
where
is given by ( 68 ) for
. In particular, zero is an isolated eigenvalue of
if
, and if
there exists
such that
(Recall that we assumed
.) Moreover, if
, the subspace
defined by ( 26 ) is invariant under the action of
, and the restriction of
to
has spectrum
. Thus
is invertible in
if
, with bounded inverse.
The semigroup generated by
satisfies
. Thus, using Proposition 4.1 , we immediately obtain the following estimates:
Proposition 4.2
If
, the semigroup
is uniformly bounded in
for all
. Moreover, if
, there exists
such that, for all
,
|
(73)
|
Finally, if
and
, then
defines a bounded operator from
into
for
, and there exists
such that, for all
,
|
(74)
|
|
(75)
|
where
Remark 4.3
For
and
, we also have the following bound:
We conclude this paragraph with a short discussion of the two-dimensional Biot-Savart law:
|
(76)
|
Proposition 4.4
Let
be the velocity field defined from
via the Biot-Savart law ( 76 ).
i) If
for some
, then
where
, and there exists
such that
.
ii) If
for some
and some
then
and there exists
such that
Proof: Assertion i) is a direct consequence of the Hardy-Littlewood-Sobolev inequality [8] . For a proof of ii), see for instance ([3] , Lemma 2.1).
We deduce from Proposition 4.4 the following useful bound on the product
:
Corollary 4.5
Assume that
for some
, and let
be the velocity field obtained from
via the Biot-Savart law ( 76 ). Then
for all
, and there exists
such that
Proof: Assume that
. By Hölder's inequality
Now
by Proposition 4.4 , and
because
for
if
. This gives the desired result.
4.3 The three-dimensional semigroup
This section is devoted to the three-dimensional equation
|
(77)
|
where
,
, and
. In the particular case where
and
, we have
where
is defined in ( 45 ).
It is important to realize that the evolution defined by ( 77 ) is essentially contracting in the transverse variables
and expanding in the axial variable
. This is due to the signs of the advection terms, which in turn originate in our choice of the straining flow ( 2 ). For this reason we can assume that the solutions of ( 77 ) decay to zero as
, but we cannot impose any decay in the
variable (otherwise the solutions will not stay uniformly bounded for all times in the corresponding norm). This motivates the following choice of our function space. For
and
, we introduce the Banach space
|
(78)
|
equipped with the norm
For any
we also define the seminorm
|
(79)
|
and we denote by
the space
equipped with the topology defined by the family of seminorms ( 79 ) for
. For later use, we observe that the ball
is closed in
for any
.
At least formally, the space
can be thought of as the tensor product
, i.e. the space generated by linear combinations of elements of the form
, with
and
. In this picture, the linear operator defined by ( 77 ) can be decomposed as
, where
is defined in ( 71 ) and
is the one-dimensional operator
. It is readily verified that
generates a semigroup in
given by the explicit formula
|
(80)
|
where
is defined in ( 52 ), and we know from Section 4.2 that
generates a strongly continuous semigroup in
. Thus we expect that
will generate a semigroup
in
given by
, or explicitly
|
(81)
|
We shall prove that these heuristic considerations are indeed correct in the sense that ( 81 ) defines a semigroup of bounded operators in
with the property that
is the solution of ( 77 ) with initial data
. However, the map
is not continuous in the topology of
, but only in the (weaker) topology of
. This is due to the fact that equation ( 77 ) has “infinite speed of propagation” in the sense that the advection term in the vertical variable is unbounded, see Remark 3.2 .
Proposition 4.6
For any
, the family
defined by ( 81 ) and
is a semigroup of bounded linear operators on
. If
and
, then
is continuous, and
solves ( 77 ) for
. For any
, if
is equipped with the topology of
, then
is continuous, uniformly in time on compact intervals. Moreover: i) If
then
is uniformly bounded on
for all
.
ii) If
there exists
such that, for all
in the subspace
defined in ( 14 ),
|
(82)
|
iii) If
and
,
defines a bounded operator from
into
for
and
, and there exists
such that
|
(83)
|
|
(84)
|
|
(85)
|
where
are as in Proposition 4.2 .
Proof: We first rewrite ( 81 ) in a slightly more convenient form. By ( 52 ) we have
, where
and
. Thus setting
in ( 81 ), we obtain the equivalent formula
|
(86)
|
Fix
. If
, then for any
the map
also belongs to
, because
is a bounded operator on
by Proposition 4.2 . Thus it follows immediately from ( 86 ) that
and
|
(87)
|
where
. The semigroup formula
is easily verified using ( 81 ), Fubini's theorem, and the identity
Thus
is a semigroup of bounded operators in
.
On the other hand, by ( 86 ), we have for all
:
where
as
. Since
whenever
and
, we deduce that for all
:
|
(88)
|
This bound implies that
is continuous, uniformly in time on compact intervals.
Furthermore, if
and
, we have
| |
| |
hence proceeding as above we find for all
:
| |
| |
The last term goes to zero as
because
is a strongly continuous semigroup on
and
is continuous (hence has compact range). Similarly, for each
, the integral goes to zero as
by Lebesgue's dominated convergence theorem, because
is uniformly continuous. Since
as
, we conclude that
in
as
. Then, using the semigroup property, we deduce that the map
is continuous to the right at any
. Finally, if
, we have
, hence by ( 88 )
Since
as
for all
, this shows that
is also continuous to the left at any
.
Next, using ( 86 ), ( 80 ), and the explicit formula for
, it is rather straightforward to verify that, for any
, the map
is smooth and satisfies ( 77 ) for
. Thus
is indeed the solution of ( 77 ) with initial data
.
It remains to establish the decay properties of
:
i) If
, we know from Proposition 4.2 that
for all
, hence
is uniformly bounded on
by ( 87 ).
ii) If
and
, then
for all
and ( 82 ) follows immediately from ( 87 ) and ( 73 ).
iii) If
, we define
by
|
(89)
|
If
and
, we know from Proposition 4.2 that
is a bounded operator from
into
satisfying ( 74 ) or ( 75 ). Thus the formula ( 89 ) defines a bounded operator from
into
and
Thus ( 83 ), ( 84 ) follow immediately from ( 74 ), ( 75 ). Finally we define
by
|
(90)
|
We know that
is a bounded operator from
into
, see Remark 4.3 .
Since
, we thus find
| |
| |
which is ( 85 ). This concludes the proof.
Corollary 4.7
Let
and let
be a bounded function satisfying
. Then the map
defined by
satisfies
, and
, where
is as in ( 87 ).
Proof: For any
, we define
by
. Since
and since the semigroup
is continuous on
as described in Proposition 4.6 , it is easy to verify that the map
is also continuous.
As
is a subspace of the Fréchet space
, the integral
can be defined as in ([15] , Theorem 3.17). However, in the present case, we can also use the following “pedestrian” construction (which agrees with the general one). For any
, we define
where
denotes the map
. Clearly
is a continuous function of
with values in the Banach space
, hence
can be defined for any
as a Banach-valued Riemann integral. Using again the continuity of the semigroup
one finds that
is continuous and satisfies
(Note that
and
are lower semicontinuous, hence measurable.) Now, for each
, it is clear that
if
, hence there is a unique
such that
whenever
. By construction,
for all
, and
.
Remark 4.8
Similarly, if
and
, Proposition 4.6 implies that, if
is bounded in
and continuous in
, the map
defined by
is bounded in
and continuous in
. In that case, for each
and each
,
is defined by a “generalized” Riemann integral, because the integrand has a singularity at
.
4.4 The three-dimensional Biot-Savart law
In this final section, we discuss the three-dimensional Biot-Savart law, namely
|
(91)
|
We first prove the analogue of Proposition 4.4 in the spaces
defined by ( 78 ).
Proposition 4.9
Fix
. If
, the velocity field given by ( 91 ) satisfies
for all
, and there exists
such that
.
Proof: Assume that
for some
. For all
, we have by Fubini's theorem
where
By Minkowski's inequality, it follows that
|
(92)
|
If
, we shall show that there exists
and
such that
|
(93)
|
Together with ( 92 ), this gives
for all
, which is the desired bound. Since the Biot-Savart law is invariant under spatial translations, the same arguments show that, for all
,
As
the right-hand side converges to zero by Lebesgue's dominated convergence theorem, thus
.
To prove ( 93 ), for any
we define
. If
, then
for all
, and there exists
such that
Since
, Young's inequality implies
|
(94)
|
where
, hence
. On the other hand, by Hölder's inequality,
| |
| |
Using Young's inequality again, we obtain
| |
|
(95)
|
where
. Combining ( 94 ), ( 95 ), we obtain ( 93 ).
An easy consequence is the analogue of Corollary 4.5 :
Corollary 4.10
Assume that
for some
, and let
be the velocity field obtained from
via the Biot-Savart law ( 91 ). Then
for all
, and there exists
such that
|
(96)
|
Proof: Assume that
and
. By Hölder's inequality, we have for all
Taking the supremum over
and using Proposition 4.9 , we obtain ( 96 ). Moreover, since
and
, it is clear that
is continuous from
into
.
In the rest of this section, we fix some
. Given
and
, our goal is to compare the velocity field
defined by ( 36 ) with the velocity field
obtained from
via the Biot-Savart law. (As in Section 3 we omit the dependence on
for simplicity.) Since
has only the third component nonzero, ( 91 ) implies that
has only the first two components nonzero:
|
(97)
|
On the other hand, for any
,
is obtained from
via the two-dimensional Biot-Savart law ( 76 ), which can be written in the form
|
(98)
|
because
|
(99)
|
Using these representation formulas, it is easy to show that the velocity fields
,
are close if the function
varies sufficiently slowly.
Proposition 4.11
Fix
, and assume that
and
satisfy
, where
is defined in Proposition 2.2 . Then
, and there exists
such that
|
(100)
|
|
(101)
|
Proof: Since
is a continuous function of
which decays rapidly as
, uniformly in
, it is not difficult to verify that the velocity field
defined by ( 97 ) depends continuously on
. Next using ( 99 ) and ( 39 ) with
, we find
Since
for all
, the above integral is uniformly bounded for all
(see Proposition 4.4 ), and we obtain ( 100 ).
Finally, taking the difference of ( 97 ) and ( 98 ), we see that
, where
But
| |
| |
Since
we thus find
|
(102)
|
Using now ( 39 ) with
, we see that the integrand is bounded by
, hence the integral in ( 102 ) is finite. This gives ( 101 ).
Acknowledgements. A part of this work was completed when CEW was a visitor at Institut Fourier, Université de Grenoble I, whose hospitality is gratefully acknowledged.
The research of CEW is supported in part by the NSF through grant DMS-0405724, and the work of ThG is supported by the ACI “Structure and dynamics of nonlinear waves” of the French Ministry of Research. References
-
J. M. Burgers. A mathematical model illustrating the theory of turbulence. Adv. Appl. Mech., 1:171–199, 1948.
-
D. G. Crowdy. A note on the linear stability of Burgers vortex. Stud. Appl. Math., 100(2):107–126, 1998.
-
Th. Gallay and C. E. Wayne. Invariant manifolds and the long-time asymptotics of the Navier-Stokes and vorticity equations on
. Arch. Ration. Mech. Anal., 163(3):209–258, 2002.
-
Th. Gallay and C. E. Wayne. Existence and stability of asymmetric Burgers vortices. Preprint, 2005.
-
Th. Gallay and C. E. Wayne. Global stability of vortex solutions of the two-dimensional Navier-Stokes equation. Comm. Math. Phys., 255(1):97–129, 2005.
-
J. Jiménez, H. K. Moffatt, and C. Vasco. The structure of the vortices in freely decaying two-dimensional turbulence. J. Fluid Mech., 313:209–222, 1996.
-
S. Leibovich and Ph. Holmes. Global stability of the Burgers vortex. Phys. Fluids, 24(3):548–549, 1981.
-
E. H. Lieb and M. Loss. Analysis, volume 14 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 1997.
-
H. K. Moffatt, S. Kida, and K. Ohkitani. Stretched vortices—the sinews of turbulence; large-Reynolds-number asymptotics. J. Fluid Mech., 259:241–264, 1994.
-
A. Prochazka and D. I. Pullin. On the two-dimensional stability of the axisymmetric Burgers vortex. Phys. Fluids, 7(7):1788–1790, 1995.
-
A. Prochazka and D. I. Pullin. Structure and stability of non-symmetric Burgers vortices. J. Fluid Mech., 363:199–228, 1998.
-
M. Reed and B. Simon. Methods of modern mathematical physics. I. Functional analysis. Academic Press, New York, 1972.
-
A. C. Robinson and P. G. Saffman. Stability and structure of stretched vortices. Stud. Appl. Math., 70(2):163–181, 1984.
-
M. Rossi and S. Le Dizès. Three-dimensional temporal spectrum of stretched vortices. Phys. Rev. Lett., 78:2567–2569, 1997.
-
W. Rudin. Functional analysis. International Series in Pure and Applied Mathematics. McGraw-Hill Inc., New York, second edition, 1991.
-
P. J. Schmid and M. Rossi. Three-dimensional stability of a Burgers vortex. J. Fluid Mech., 500:103–112, 2004.
-
A. A. Townsend. On the fine-scale structure of turbulence. Proc. R. Soc. Lond. A, 208:534–542, 1951.