A geometric approximation to the Euler equations : the Vlasov-Monge-Ampere system
Yann Brenier, Gregoire Loeper, UMR 6621, Parc Valrose, 06108 Nice, France
November 27, 2006
Abstract
This paper studies the Vlasov-Monge-Ampere system (
), a fully non-linear version of the Vlasov-Poisson system (
) where the (real) Monge-Ampere equation
substitutes for the usual Poisson equation. This system can be derived as a geometric approximation of the Euler equations of incompressible fluid mechanics in the spirit of Arnold and Ebin. Global existence of weak solutions and local existence of smooth solutions are obtained. Links between the
system, the
system and the Euler equations are established through rigorous asymptotic analysis.
1 Introduction
The classical Vlasov-Poisson (
) system describes the evolution of an electronic cloud in a neutralizing uniform background through the following equations
|
(1)
|
where
denotes the electronic density at time
, point
, velocity
(usually
),
denotes the ’macroscopic’ density
|
(3)
|
and
denotes the electric potential at time
and point
generated, through the Poisson equation ( 2 ), where
is a coupling constant, by the difference between the electronic density
and the neutralizing background density, which is supposed to be uniform and normalized to unity. Standard notations
and
have been used and
stands for the inner product in
. The mathematical theory of the
system is now well understood. In particular, existence of global smooth solutions in three space dimensions has been proved in [?] (see also [?] , [?] ). In the present paper, a fully nonlinear version of the
system is addressed :
|
(4)
|
|
(5)
|
where the (real) Monge-Ampere equation ( 5 ) substitutes for the Poisson equation ( 2 ). Here,
stands for the
symmetric matrix made of all second order
partial derivatives of
,
stands for the
identity matrix and
for the determinant of a square matrix. The occurrence of the Monge-Ampere equation in mathematical modeling is not very common.
Notice, however, that a very similar system can be found in meteorology with Hoskins’ semi-geostrophic equations (cf. [?] , [?] and the included references).
In a simplified two dimensional setting, the semi-geostrophic equations read
|
(6)
|
|
(7)
|
where
denotes the usual Poisson bracket.
Formally, as the coupling constant
is small, the
and
equations asymptotically approach each other up to order
. Indeed, linearizing the determinant about the identity matrix leads to
|
(8)
|
The formal limit, as
, reads
|
(9)
|
where constraint ( 10 ) substitutes for both the Poisson and the Monge-Ampere equations. The limit system ( 9 , 10 ), that we call constrained Vlasov system, can be seen as a ’kinetic’ extension of the Euler equations of classical incompressible fluid mechanics,
|
(11)
|
where
and
respectively are the velocity and the pressure of the fluid at time
and position
. Indeed, any smooth solution
provides a ’monokinetic’ solution to the constrained Vlasov system ( 9 , 10 ), defined by
Here a monokinetic solution means a delta-valued solution in the
variable.
In addition, the constrained Vlasov system ( 9 , 10 ) turns out to be a natural extension (or
limit) of the Euler equations from both geometrical and variational reasons, as explained in section 2 In a similar way, there is a monokinetic version of the
system, the so-called (pressureless) Euler-Poisson (
) system, which reads
|
(13)
|
|
(14)
|
A rigorous asymptotic analysis of the
system as
will be provided (sections 5.1 and 5.2 ), in the case when the initial electronic density
|
(16)
|
is asymptotically monokinetic, namely approaching
for some smooth divergence free velocity field
, as
tends to zero. Before this asymptotic analysis, we want to explain the geometric origin of the
system. It has been known, since Arnold’s celebrated work (cf. [?] ), that the Euler equations (formally) describe geodesics curves along a suitable group of volume preserving maps, lengths being measured in the
sense.
We will show (section 2 ) that the
system just describes approximate geodesics obtained through a very natural penalty method, where
stands for the penalty parameter. For this geometric interpretation to be valid, the Monge-Ampere equation ( 5 ) must be understood in the following weak sense:
for each fixed
,
is the unique (up to an additive constant) function such that
is convex in
and
|
(17)
|
where
is a fixed bounded open convex set where the neutralizing background of the electrons is assumed to be located. (This definition is made precise in section 2.3 .) Notice that, by construction,
must be valued in the closure of
and, therefore, the potential
enjoys the following property
There is no similar bound for the electrostatic potential of the classical
system. Thus, in some sense, the
system can be seen as a nonlinearly saturated version of the
system.
Beyond the geometric derivation of the
system, our main analytic results are as follows:
-
∙
The
system admits global energy preserving weak solutions.
-
∙
The
system admits local strong solutions in periodic domains.
-
∙
For well prepared, nearly monokinetic initial data, the solutions of the
system converge when
goes to 0 to those of the Euler equations.
-
∙
In this asymptotic, the
system is a higher order approximation of the
system.
The paper is organized as follows: in section 2, we first recall the geometric nature of the Euler equations, then we explain why the constrained Vlasov system ( 9 , 10 ) is a natural extension of the Euler equations from a variational point of view, finally we introduce the concept of approximate geodesics for volume preserving maps, and derive the
system. Section 3 is devoted to the proof of existence of global energy preserving weak solutions. In section 4, we prove existence of local strong solutions, in the case of a periodic domain.
Finally, in section 5, we study the asymptotic behavior of the
system as
goes to 0.
2 The geometric origin of the Vlasov-Monge-Ampere system
2.1 The Euler equations
The motion of an incompressible fluid in a domain
is classically described by the Euler equations
:
|
(18)
|
with
,
, where
stands for the velocity field and
for the scalar pressure field. These equations have a nice geometrical interpretation going back to Arnold (see [?] ). Introducing
the group of all volume preserving diffeomorphisms of
with jacobian determinant equal to 1, and measuring lengths in the
sense, we may define (at least formally) geodesic curves along
. It turns out that the Euler equations just describe these curves. For the same reasons, the Euler equations can be seen as the optimality equations for the corresponding minimization problem:
given two maps chosen in
, find an
shortest path between them along
. It was shown by Shnirelman [?] (see also [?] and [?] ) that, in the case when
is the unit cube in
, there are many maps for which there are no such shortest paths. Beyond this negative result, [?] established that minimizing paths are more appropriately described by doubly stochastic measures. These measures (also called polymorphisms) generalize volume preserving maps in the following way: a doubly stochastic measure
is a (Borel) probability measure on
with two projections on each copy of
both equal to the (normalized) Lebesgue measure. It is known -see [?] , for instancethat any such
can be weakly approximated by a sequence
where each
is a volume preserving map of
.
In [?] it was shown that, in the case considered by Shnirelman for which there is no classical shortest path, minimizing paths along
converge to paths of doubly stochastic measures
governed by the following extension of the Euler equations
|
(20)
|
|
(21)
|
where
and
can be respectively seen as the velocity field and the pressure field attached to
. (Notice that the velocity field
generally depends on the extra variable
and is not a classical but rather a multivalued velocity field.) These equations are just a reformulation of the constrained Vlasov system ( 9 , 10 ). Indeed, it can be checked, under appropriate regularity assumptions, that the kinetic measure
defined by
|
(22)
|
solves ( 9 , 10 ) when
solves ( 20 , 21 ). Thus we conclude that the constrained Vlasov system ( 9 , 10 ) is a natural variational extension of the Euler equations.
2.2 Approximate geodesics
A general strategy to define approximate geodesics along a manifold
(in our case
) embedded in a Hilbert space
(here
) is to introduce a penalty parameter
and the following
dynamical system in
|
(23)
|
In this equation, the unknown
is a curve in
,
is the distance (in
) of
to the manifold
, i.e. in our case as
,
|
(24)
|
and, finally,
denotes the gradient operator in
. This penalty approach has been used for the Euler equations by the first author in [?] . It is similar-but not identicalto Ebin’s slightly compressible flow theory [?] , and is a natural extension of the theory of constrained finite dimensional mechanical systems [?] . The penalized system is formally hamiltonian in variables
with Hamiltonian (or energy) given by:
| |
(Multiplying equation ( 23 ) by
, we formally get that the energy is conserved.) Therefore it is plausible that the map
will remain close to
if properly initialized at
. A formal computation shows that, given a point
for which there is a unique closest point
to
in the
closure of
, we have:
|
(25)
|
Thus the equation ( 23 ) formally becomes:
|
(26)
|
To understand why solutions to such a system may approach geodesics along
as
goes to 0, just recall that, in the simple framework of a surface
embedded in the 3 dimensional Euclidean space, a geodesic
along
is characterized by the fact that for every
, the plane defined by
is orthogonal to
. In our case,
is nearly orthogonal to
thanks to ( 26 ), meanwhile
remains close to
.
The approximate geodesic equation was introduced in [?] in order to allow a spatial approximation of
by the group of permutations of
points
chosen to form a discrete grid on
. On such a discrete group, the concept of geodesics becomes unclear meanwhile approximate geodesics still make sense.
They can be interpreted as trajectories of a cloud of
particles
moving in the Euclidean space
, which substitutes for
. These particles solve the following coupled system of harmonic oscillators
where
is a time dependent permutation minimizing, at each fixed time
,
among all other permutations of the first
integers. The convergence of this discrete model to the incompressible Euler equations for well prepared initial data was proved in [?] . In order to study the continuous version ( 26 ), a specific study of the projection problem ( 24 ) is needed.
2.3 The polar decomposition Theorem
Let us first recall a general measure theoretic definition:
Definition 2.1
Let
and
be two topological spaces, let
be a Borel finite measure of
and
a Borel map
, we call the push-forward of
by
and note
the Borel measure
on
defined by
| |
Let us now consider the case of a bounded open subset
of the Euclidean space
equipped with the Lebesgue measure that we denote
. We say that a Borel map
is volume (or Lebesgue measure) preserving if
, i.e. if for all
one has
or equivalently, for any Borel subset
of
one has
. The set of all measure preserving maps of
is a closed subset of the Hilbert space
and will be denoted by
. Notice that
is only a semi-group for the composition rule and contains the group of volume preserving diffeomorphisms
. It is known [?] that, at least in the case when
is convex and
,
is exactly the closure of
in
, which implies
The polar decomposition Theorem for maps [?] (extended to Riemannian manifolds in [?] ) will be crucial for our analysis of the
system:
Theorem 2.2
Let
be a bounded convex open subset of
, let
and
where
is the Lebesgue measure on
Assume
to be a Lebesgue integrable function, or, equivalently,
to satisfy the non-degeneracy condition:
|
(27)
|
Then there exists a unique pair
where
is a convex function and
, such that
In this ’polar decomposition’,
is also characterized as the unique closest point to
on
in the
sense and
is characterized to be (up to an additive constant) the unique convex function on
satisfying
|
(29)
|
for any
such that
.
In addition, the Legendre-Fenchel transform
of
defined by
|
(30)
|
is Lipschitz continuous on
, with Lipschitz constant bounded by
and has the following properties :
holds true for
a.e.
,
|
(31)
|
for any
, and
|
(32)
|
|
(33)
|
|
(34)
|
We make here several remarks on Theorem 2.2 :
Link with the Monge-Ampere equation We can interpret ( 29 ) as a weak version of the Monge-Ampere equation:
| |
and ( 31 ) can be seen as a weak version of another Monge-Ampere equation:
| |
The pair
depends in fact only of
and the measure
and if condition ( 27 ) fails, then existence and uniqueness of the projection
may fail, but existence and uniqueness of
remain true.
Theorem 2.2 and the subsequent remarks allow us to introduce the following notation that will be used throughout the paper:
Definition 2.3
Let
be a fixed bounded convex open set of
, let
be a positive measure on
of total mass
, absolutely continuous w.r.t the Lebesgue measure and such that
. We call
, or, in short,
, the unique up to a constant convex function on
satisfying
|
(35)
|
We call
its Legendre-Fenchel transform satisfying
|
(36)
|
If no confusion is possible we may write
(resp.
) instead of
(resp.
We will use some additional results from [?] . The first one establishes the continuity of the polar decomposition:
Theorem 2.4
Let
be a Lebesgue integrable positive measure on
, with total mass
, such that
. Let
be a sequence of Lebesgue integrable positive measures on
, with total mass
, such that
,
. Let
and
be as in Definition 2.3 . If for any
such that
converges to
, then
-
∙
converges to
uniformly on each compact set of
and strongly in
-
∙
converges to
uniformly on each compact set of
and strongly in
for every
compact in
.
The second one provides a ’dual’ definition of the distance between a map
and the semi-group
:
Theorem 2.5
Let
and
where
is the Lebesgue measure on
Assume
to be a Lebesgue integrable function. Then
where the supremum if performed over all pairs
of continuous functions on
such that
pointwise.
2.4 The Vlasov-Monge-Ampere system
Let us now derive the
system as the kinetic formulation of the approximate geodesic equation ( 26 ). First, from the polar decomposition Theorem 2.2 , equation ( 26 ) reads
|
(37)
|
where
|
(38)
|
and
is as in Definition ( 2.3 ). This means that
satisfies ( 5 ) in a weak form with the additional condition that the range of
is contained in
.
Next, let
be a given initial density function, that we assume to be in
, compactly supported and satisfying the compatibility condition
|
(39)
|
For each
, let us define
to be
pushed forward by the following ODE
|
(40)
|
|
(41)
|
|
(42)
|
Then
satisfies the following kinetic (or Liouville) equation
|
(43)
|
|
(44)
|
which must be understood in the following weak sense
| |
| |
|
(45)
|
This linear Liouville equation is nonlinearly coupled to equation ( 38 ), where
is linked to
by equation ( ?? ). Finally, we have defined, through ( 38 , 43 , 44 ), the weak formulation of the
initial value problem.
The energy of the system is defined by
| |
|
(46)
|
3 Existence of global renormalized weak solutions
The main result of this section is as follows:
Theorem 3.1
Let
be in
, with compact support in both
and
, satisfying condition ( 39 ).
Then the
system ( 38 , 43 , 44 ) admits a global weak solution
, with
and
for all
. In addition, each such weak solution enjoys the following properties:
-
∙
is a continuous function of
, valued in
, for every
-
∙
the density
is a continuous function of
, valued in
, for every
,
-
∙
the support of
in
is compact, with a diameter growing no more than linearly in
.
-
∙
the total energy defined by ( 46 ) is conserved,
-
∙
the ’renormalization’ property (in the sense of [?] )
holds true for all
,
-
∙
the trajectories of ( 41 , 42 ) are uniquely defined for almost every initial condition
,
-
∙
is just
pushed forward along the trajectories of ( 41 , 42 ).
Proof of Theorem 3.1 :
We build a sequence of approximate solutions
by time discretization and let the time step
go to zero. To handle the limiting process, the non-linear terms will be treated with the help of Theorem 2.4 . More precisely if one can extract a subsequence such that, for every
,
converges weakly, then we can deduce from Theorem 2.4 that the corresponding sequence
will converge strongly, and this will allow us pass to the limit in the nonlinear term.
3.1 Construction of a sequence of approximate solutions
We consider
such that
,
and
.
We then seek approximate solutions as solutions of the approximate problem
|
(47)
|
|
(48)
|
|
(49)
|
being a smooth function of space this regularized equation admits a unique solution that one builds by the method of characteristics. Since the flow is divergence-free in the phase space, the solution
satisfies
|
(50)
|
By construction (through Theorem 2.2 ),
is valued in the convex bounded set
. Suppose that
vanishes outside of the set
for some constant
fixed and denote
. Then we have
Lemma 3.2
is supported in
.
Proof : We just write
in complex notation
, where
and
, which is bounded by
. This leads to
ant the desired bound easily follows. Notice here a sharp contrast with the classical
system, for which the
support of the solutions cannot be controlled so easily (except in the one dimensional case).
Convergence of the sequence of approximate solutions Using ( 50 ) and Lemma 3.2 there exists, for any
, up to the extraction of a subsequence,
such that
converges weakly to
as
.
It remains to show that the product
converges to the good limit.
For this we need strong convergence of
We already know that
We claim that for all
,
converges strongly to
in
,
. Indeed, such a strong convergence of
follows from Theorem 2.4 provided that we have for all
,
|
(51)
|
for any
such that
. Note first that from Lemma 3.2 , we can restrict ourselves here to test functions
that are compactly supported. Then we show that the sequence
is relatively compact in
This is done by the following lemma:
Lemma 3.3
For all
, for all
with
the sequence
(resp.
) satisfies
-
∙
(resp.
) is a bounded sequence in
(resp.
in
,
-
∙
(resp.
) is a bounded sequence in
, (resp. in
,
and one can extract from
(resp. from
) a subsequence converging in
(resp. in
).
Proof: the first point uses equation ( 50 ) and Lemma 3.2 . The second point uses equation ( 43 ) and the identity:
with the fact that the
are uniformly compactly supported in
and
(Lemma 3.2 ); the last point is a classical result of functional analysis (see [?] for example).
This lemma and Lemma 3.2 yield ( 51 ). Then using Theorem 2.4 , with
the limit of a subsequence of
, we have convergence of the sequence
to
in
. We have extracted a subsequence
such that
-
∙
converges in
for every
.
-
∙
converges in
for every
.
-
∙
converges in
for every
and for every
.
Thus the limit
satisfies equations ( 43 - 44 ) and the first part of Theorem 3.1 is proved.
3.2 Conservation of energy
We now give a rigorous proof of the conservation of energy following an argument going back to F. Otto (in an unpublished work on the semi-geostrophic equations). We recall the definition of the energy as
| |
We call the first term the kinetic energy
and the second term, multiplied by
, the (normalized) potential energy
We have
Proposition 3.4
Let
be any solution of ( 43 ) such that on every interval
,
is uniformly compactly supported in
for some function
. Then the energy of the solution
is conserved.
Proof: From Theorem 2.5 , we know that
where the supremum if performed over all pairs
of continuous functions on
such that
pointwise. Thus for each
, we have
| |
and this implies
| |
| |
| |
Notice that the product in the second line is licit since
is in
for any
,
and therefore
are compactly supported in space uniformly on
, and
is in
. Exchanging
and
we would have found
| |
moreover we have for the kinetic energy
| |
Dividing by
we find
| |
| |
and
| |
| |
We know from 3.1 that
converges strongly in
to
as
goes to
, and so the right hand sides of the above inequalities converges to 0 and we conclude that
We could take
and find the same result. Finally we conclude that
3.3 Renormalized solutions and existence of characteristics
The study of renormalized solutions for transport equations has been introduced in [?] for vector fields in
with bounded divergence. These results have been extended by Bouchut [?] to the case of Vlasov-type equations with acceleration field in
(A recent result of L. Ambrosio, [?] , has extended the existence of renormalized solutions to transport equations with vector fields in
and with bounded divergence). The fact that solutions of ( 43 , 44 ) are renormalized solutions is an immediate consequence of the following theorem:
Theorem 3.5
F. Bouchut
Let
satisfy
| |
with
, then, for any
,
and for every
,
belongs to
.
In our case the BV bound on the acceleration
is a direct consequence of the fact that
is a globally Lipschitz convex function. This result implies the strong time continuity results for
and
in Theorem 3.1 . Finally, as in [?] , it can be deduced from the renormalization property that 1) for almost every initial condition
, there is a unique trajectory solving ( 41 , 42 ), 2)
is just
pushed forward along these trajectories.
A complete proof is given in appendix.
Remark: From the renormalization property it follows that, once the potential
is known, there exists a unique solution to ( 43 ) in
.
Of course, this does not imply at all the uniqueness of weak solutions to the Vlasov-Monge-Ampere system! This paragraph ends the proof of Theorem 3.1 .
4 Strong solutions
In this section we show existence of strong solutions over a finite time intervall. To do so, we need regularity estimates for solutions of Monge-Ampere equation. We will get rid of the difficulties that may arise at the free boundary of the set
by considering the periodic case. Note that for the Vlasov-Poisson system, existence of global smooth solutions has been proved (see [?] ); in the present case, due to the non-linearity of the Monge-Ampere equation, we were only able to obtain a result for finite time.
4.1 The periodic Vlasov-Monge-Ampere system
Polar factorization of maps in a periodic domain The polar decomposition Theorem has been generalized in [?] to general Riemannian manifolds, while the particular case of the flat torus
had been addressed in [?] .
Definition 4.1
We say that a mapping
is
additive if the mapping
is
periodic. The set of all measurable
additive mappings is denoted by
. For each
we call
the class of
in
, and for any
,
the mapping of
into itself defined by
We may say if no confusion is possible additive instead of
additive. Then the following theorem can be deduced from the results of [?] and [?] :
Theorem 4.2
Let
be additive and assume that
has a density in
. Then there exists a unique pair
such that
where
is a convex function and
is
periodic,
is additive and
is Lebesgue measure preserving in
. Moreover we have
and,
denoting the Legendre transform of
, we have
Remark : The pair
is uniquely defined by the density
.
Notice that the periodicity of
implies that
and
are
additive, and that
is also
periodic. As in the previous case, we introduce the following notation:
Definition 4.3
Let
be a probability measure on
, with density in
. We denote
(resp.
) the unique up to a constant convex function such that
|
(52)
|
|
(53)
|
(resp. its Legendre fenchel transform).
will thus be a generalized solution of the Monge-Ampere equation
Next the results of Caffarelli ([?] , [?] , [?] ) on the regularity of solutions of the Monge-Ampere equation yield the following theorem:
Theorem 4.4
Let
be a
probability density on
, for some
.
Then
(see Definition 4.3 ) is a classical solution of
and satisfies:
| |
| |
where
and
.
This theorem is an adaptation of the regularity results stated above, whose complete proof is given in appendix.
The periodic Vlasov-Monge-Ampere system We now seek
, for some
, solution of the initial value problem for the periodic Vlasov-Monge-Ampere
system
|
(54)
|
|
(55)
|
for a given
satisfying the compatibility condition
|
(56)
|
The macroscopic density
is still related to
by equation ( 3 ), and
is as in Definition 4.3 .
4.2 Existence of local strong solutions
We mention first that the proof of existence of global weak solutions adapts with minor changes to the periodic case, and that the obtain for the periodic
system the same result as Theorem 3.1 .
Our result in this section is the following:
Theorem 4.5
Let
be such that:
|
(57)
|
|
(58)
|
then there exists
and a solution
to the
system ( 54 , 55 ), in the space
.
Proof of Theorem 4.5 : First we deduce from Theorem 4.4 :
Corollary 4.6
Let
be as in Theorem 4.4 . Then, we have
and we can define
We see that in order to use Theorem 4.4 we need
to be bounded away from 0. In the following lemma, we show that under suitable assumptions on the initial data, it is possible to enforce locally in time the condition
Lemma 4.7
Let
satisfy
| |
with
and
| |
let the initial condition
be such that
with
and
and
satisfying
| |
Then there exists a constant
depending on
, such that
| |
The proof of the lemma is given in appendix.
4.2.1 Construction of approximate solutions
Let us consider
a smooth vector-field on
, and write
| |
If
satisfies
, we have
| |
| |
and therefore
|
(59)
|
which implies
| |
Now let
be given as in Theorem 4.5 , satisfying ( 57 , 58 ). Thanks to Lemma 4.7 it is possible to find
such that for any
satisfying
for any field
satisfying
, we have
|
(60)
|
|
(61)
|
with
supported in
and with
as in Theorem 4.5 , so that we have for
:
|
(62)
|
being the volume of the unit ball of
. Then we construct a family of approximate solutions
to ( 54 ), in the same spirit as we did for weak solutions, by solving
| |
| |
Note that we have neither mollified the term
nor the initial condition and that
. Now choose
, if for some
we have
| |
this implies, thanks to ( 62 ), that
| |
and conditions ( 60 , 61 ) are satisfied because
. Then if we denote
as in Corollary 4.6 , we have for
| |
and then
| |
So if we define
as
with
, we have for
,
| |
Thus we can extract a subsequence converging to a strong solution of ( 54 , 55 ). Then we argue as in section 2 to show that all terms converge to the correct limit. This ends the proof of Theorem 4.5 .
5 Asymptotic analysis
5.1 Convergence to the Euler equation
In this section we justify that the Vlasov-Monge-Ampere system describes approximate geodesics on volume preserving transformations: indeed we will show that weak solutions of this system converge to a solution of the incompressible Euler equations
as the parameter
goes to 0, at least for well prepared initial data. We restrict ourselves to the space periodic case, the macroscopic density
is still defined by ( 3 ) and the convex potentials
are still as in Definition 4.3 .
For sake of simplicity, we slightly modify our notations and introduce the following rescaled potentials
| |
| |
so that
| |
and the
system takes the following form:
|
(63)
|
|
(64)
|
The energy is given by
|
(65)
|
| |
It has been shown in section 3.2 that the energy is conserved. The Euler equations for incompressible fluids
reads:
|
(66)
|
We shall here consider a smooth solution of
and a weak solution of
, with ‘well prepared initial data’, meaning that the initial data of both systems are close a time 0. Then we will show that as time evolves, both solutions stay close to each other.
Theorem 5.1
Let
be a weak solution of ( 63 , 64 ) with finite energy, let
be a smooth
solution of ( 66 , 67 ) for
and
the corresponding pressure, let
| |
then
| |
depends only on
.
Remark 1: This estimate is enough to compare the weak solutions
to the
system (for well prepared initial data) and the smooth solutions
of the Euler equations. For instance,
implies
and therefore,
implies
where
depends only on
,
and
.
Remark 2: We see that we consider nearly monokinetic initial data for the
system.
Proof of Theorem 5.1 We shall show that
| |
| |
|
(68)
|
where we will use the notation
| |
The proof of this identity is postponed to the end of the section.
Now we look at all terms of the right hand side. All the constants that we denote by
are controlled as in Theorem 5.1 . We set
| |
| |
| |
First we have
For
we have
| |
| |
| |
| |
we have used that
is divergence-free thus its integral against any gradient is zero. Next we have the following lemma:
Lemma 5.2
Let
be Lipschitz continuous such that
, then for all
, one has
Proof: We just write a Taylor expansion of
:
| |
| |
Again, since
is divergence-free,
thus from Lemma 5.2 we have
| |
We remind that
Since it costs no generality to suppose that for all
,
, we obtain that
| |
| |
| |
again using Lemma 5.2 , where
Thus
| |
and we have the following inequality:
|
(69)
|
Moreover, using Lemma 5.2 ,
|
(70)
|
thus
|
(71)
|
and we can transform ( 69 ) in
|
(72)
|
Gronwall’s lemma then yields
Using again ( 70 ) we obtain
|
(73)
|
which achieves the proof of Theorem 5.1 .
Proof of identity ( 68 ):
We first notice that, for all
, we have:
We also use the conservation of energy defined by ( 65 ). Then we get
| |
| |
| |
Integrating by part, we get
| |
| |
The first two terms can be rewritten as
| |
| |
| |
| |
and finally using equation ( 66 ) we conclude.
5.2 Comparison with the Euler-Poisson system
Here we show that, as mentioned in the introduction, the Euler-Poisson (
) system is a more accurate approximation to the Vlasov Monge-Ampere system than the Euler equations, as
goes to zero.
The
system Let us recall that the (pressureless) Euler-Poisson system describes the motion of a continuum of electrons on a neutralizing background of ions through electrostatic interaction. Let
and
be the velocity and density of electrons. Let
be the (rescaled) electric potential. Under proper scaling, these functions of
and
satisfy the Euler-Poisson system:
|
(74)
|
|
(75)
|
The so-called ’quasi-neutral’ limit
of similar systems has been studied for example in [?] and [?] , and convergence results have been established using pseudo-differentials energy estimates. For well-prepared initial data, solutions of
are expected to behave as solutions of Euler incompressible equations.
This fact is proved by the second author in his PhD thesis ([?] , Chap 2), see also [?] . We give here the complete result that we will use herafter. We will denote by
(resp.
the solutions of the
(resp.
) system with parameter
.
Theorem 5.3
Let
be a solution of ( 66 , 67 ) on
, with initial data
, and satisfying
for some
.
There for some
,
, if
is such that the sequences
| |
are bounded in
, then there exists
with
and a sequence
of solutions to the
system on
with initial data
, belonging to
. Moreover, for
small enough, the sequences
| |
are bounded in
. Finally,
goes to
as
goes to
.
Assumptions Here we consider
a solution to
( 66 , 67 ) with initial data
, a sequence
of solutions of
( 63 , 64 ) with initial data
, and a sequence
solutions of
( 74 , 75 , 76 ) with initial data
. We still define
as in Theorem 5.1 :
| |
We introduce the following assumptions:
-
H0
solution of
satisfies, for some
,
, and
is large enough so that
in Theorem 5.3 satisfies
.
-
H1
The sequence
of initial data of
is such that, for some
,
| |
-
H2
The sequence
satisfies
for some
.
H0, H1, H2 imply that
-
1.
There exists
such that
|
(77)
|
-
2.
From Theorem 5.1 , there exists
such that
|
(78)
|
-
3.
From Theorem 5.4 and Sobolev imbeddings, there exists
such that
|
(79)
|
We are now ready to prove the following result:
Theorem 5.4
Let
be as above, satisfying assumptions H0, H1, H2. Define
| |
Then there exists
such that
| |
where C depends on
.
Remark: the theorem shows that the distance between solutions of the
system and the
system measured with
is like
whereas Theorem 5.1 showed that the distance between the solution of the Euler equation and the
system was like
. Note also that
and
can both be interpreted as the square of a distance.
Proof of Theorem 5.4 : For notational simplicity, we drop most
’s. Proceeding as in ( 68 ) and noticing that:
| |
we obtain the following identity:
| |
| |
| |
|
(80)
|
Then we notice
| |
Next, we have the following lemma:
Lemma 5.5
Define for any
| |
| |
Then
| |
| |
Proof: The proof just uses the Taylor expansion and the identity
.
Using Lemma 5.5 , we get
| |
We claim that, under our assumptions, we have
Proof: from ( 75 ), we have
| |
Using ( 79 ), we obtain that
. Since
is continuously embedded in
,
is continuously embedded in
.
Then, using ( 76 ) and classical elliptic regularity, we have
and the desired result follows.
This implies, using ( 78 ), that
| |
Next,
| |
| |
Using again Lemma 5.5 , we get
| |
and for the same reasons we have
. This yields
Moreover, it does not cost to set
and deduce
| |
Thus the remaining terms are
| |
Calculations that we postpone to the end of the proof show that
| |
|
(81)
|
with
depending on
and
, therefore uniformly bounded thanks to ( 79 ). Finally we obtain
| |
| |
with
for
. From ( 79 ) we have
and
, whereas ( 78 ) yields
. Note that we also have
| |
| |
We conclude that
| |
and the conclusion of Theorem 5.4 follows by Gronwall’s lemma.
Proof of identity ( 81 ): Here we have to compute:
| |
Using Lemma 5.5 we have:
| |
| |
| |
We see that the first line cancels. Then we show that the last line is bounded by
.
This is obvious for the last term since from ( 77 , 78 ) we have
, and
.
Then for the first term we have the following lemma:
Lemma 5.6
We define
then one has:
Proof: First we show that if
,
Proof:
, implies that
Since
for
| |
| |
Then we have
| |
| |
| |
| |
| |
for all R, so for
one obtains:
| |
Thus we have shown that
, and
where each
is given by:
| |
| |
| |
| |
| |
| |
where we have used Einstein’s convention for repeated indices. First we have
| |
| |
with
.
| |
and
| |
thus
| |
with
. Then
| |
| |
and finally we obtain
| |
| |
and the identity ( 81 ) is proved.
6 Appendix
Existence and uniqueness of solutions to second order ODE’s with BV field In this section we prove existence and a.e. uniqueness for ordinary differential equations of the form:
|
(82)
|
for
, and where the field
belongs to
. We work in the flat torus for simplicity, but our results are still valid in an open subset of
. This result is an adaptation of the proof of [?] that uses the result of [?] on renormalized solutions of transport equations.
Remark: After this proof was written, the authors learned of a result by L. Ambrosio ([?] ) that extends the results of [?] to transport equations when the vector field is in
with bounded divergence.
Renormalized solutions for Vlasov equations with BV field Theorem 3.4 in [?] adapted to the periodic case sates that if
satisfies:
|
(83)
|
with
, then for all
Lipschitz continuous we have
| |
The property of renormalization implies that
-
∙
solutions to ( 83 ) with initial data in
belong to
for any
,
-
∙
solutions to ( 83 ) with prescribed initial data in
are a.e.
unique,
-
∙
if
converges to
in
then the solutions of ( 83 ) with
instead of
converge to the solution of ( 83 ).
We notice that equation ( 54 ) satisfies the assumptions of the Theorem, and thus will have the renormalization property. This renormalization property was used in [?] to obtain a.e. uniqueness for solutions of the corresponding ODE’s. Indeed, the ODE
| |
is associated to the transport equation:
| |
whose solutions satisfy for all
| |
We extend this consequence to the case of second order equations, with
acceleration field. To the kinetic equation
|
(84)
|
we associate the second order ODE ( 82 ) which can rewritten as
. The result is then the following:
Theorem 6.1
Let
, then the ODE
|
(85)
|
|
(86)
|
admits an a.e. unique solution.
Remark: Here almost everywhere must be understood for the Lebesgue measure of
.
Proof of Theorem 6.1 : We know that through equation ( 82 ) equation ( 85 ) can be considered as a first order differential equation. Let us first consider the case where
is smooth. Note
(resp.
) for
(resp.
for
) and
for
. Then for all
,
solves:
|
(87)
|
Then for all
we have the following:
| |
| |
Differentiating the last equation with respect to
yields:
|
(89)
|
|
(90)
|
thus solves a transport equation which is nothing but equation ( 84 ). Using Theorem 3.5 we know that for all
Lipschitz continuous,
is the unique solution of
|
(91)
|
|
(92)
|
Now we show existence and uniqueness for solutions of ( 87 , 88 ). Let
and
be fixed. Let us consider a regularization
of the the field
and set
. We have
-
∙
that satisfies ( 87 , 88 )
-
∙
that satisfies ( 89 , 90 ).
From the stability Theorem 2.4 in [?] we know that the whole sequence
converges in
to
, the unique renormalized solution of ( 89 , 90 ). Thus for fixed
the whole sequence
converges strongly in
. Now since for every
we have
the same property holds for
Now we can pass to the limit in the term
. Indeed, by density of
functions in
if we have
approximating
then
| |
| |
| |
| |
The second term goes to 0 because of the strong convergence of
, the first and the third go to 0 because
and
are measure preserving mappings, and so for example
.
So finally we have
| |
| |
| |
that goes to 0 and we can pass to the limit in the equation ( 87 , 88 ) and the existence of a solution to ( 87 , 88 ) is proved.
To obtain uniqueness, we argue as in [?] . Any function of the form
is a solution of ( 91 , 92 ), thus by uniqueness of the solution of the transport equation we obtain uniqueness of the ODE.
A remark on ODE’s of second order In this section, we want to solve the Cauchy problem for:
| |
| |
with
as above. We are thus interested in monokinetic initial data.
Theorem 6.2
for all
vector field on
, and for Lebesgue almost every
, there exists an a.e. unique solution to
| |
| |
Proof: Let
be the indicator function of the set of those
such that the trajectory coming from x is not well defined. We just have to prove that for a.e.
we have
which is true because
Stability
Using the fact that for
converging to
in
with
, we have
in
, we have then, for all t, for almost every
in
. Thus we have
Theorem 6.3
If
converges to
in
let
be solution of
| |
| |
then for all t, for almost every
converges in
to a solution (unique for almost every
) of
| |
| |
Control of macroscopic density in kinetic equations We prove here Lemma 4.7 :
Lemma 6.4
Let
satisfy
|
(93)
|
|
(94)
|
with
and
|
(95)
|
Let the initial condition
be such that:
with
and
and
satisfying
|
(96)
|
Then there exists a constant
such that
| |
Proof: First suppose that the force field and the initial data are smooth.
For equation ( 93 , 94 ) we can exhibit characteristics
giving the evolution of the particles in the phase space. We have
Since the initial data is compactly supported and the force field is bounded in the
norm, we have
If for
we have
then
| |
| |
| |
Using ( 96 ) and integrating in
we find thus a constant
such that for
we have:
| |
Next we need to show that the solution of the regularized equation converges to the solution we are studying: this result comes from the uniqueness of the solution to ( 93 , 94 ) which is a consequence of the renormalization property. Indeed since
is bounded in
the system ( 93 , 94 ) admits a unique renormalized solution and the sequence of approximate solutions converge in
for
thus the bounds obtained above are preserved.
Regularity of the polar factorization on the flat torus Here we deduce from [?] , [?] and [?] , [?] , [?] the Theorem 4.4 .
Theorem 6.5
If
with
is a probability measure on
then
(see Definition 4.3 ) is a classical solution of
and satisfies:
|
(98)
|
|
(99)
|
Proof of Theorem 6.5 : Consider
a
periodic probability measure, satisfying
and
as in Definition 4.3 . First it is shown in [?] that
|
(101)
|
It follows that the strict convexity argument of [?] applies: indeed if
is not strictly convex its graph contains a line and this contradicts ( 101 ).
Moreover since
is globally Lipschitz and periodic there exists
such that
. It follows then that there exists
and
such that
|
(102)
|
It remains to show that our solution is a solution in the Aleksandrov sense of the Monge-Ampere equation
This is a direct consequence with minor changes (to adapt to the periodic case) of Lemma 2 of [?] . Then, normalizing
to
it follows that
is a solution of
| |
| |
Thus the interior regularity results of [?] apply uniformly to all
with
satisfying ( 100 ) and
bounded and Theorem 4.4 follows.
Acknowledgments The authors acknowledge the support of the European IHP network ”HYKE” HPRN-CT-2002-00282 and the LRC CEA-Cadarache/UNSA.