Quasi-neutral limit of the Euler-Poisson and Euler-Monge-Ampere systems
G. LoeperResearch supported by a doctoral grant at Laboratoire J.A. Dieudonne, Universite de Nice-Sophia-Antipolis
Abstract
This paper studies the pressureless Euler-Poisson system and its fully non-linear counterpart, the Euler-Monge-Ampere system, where the fully non-linear Monge-Ampere equation substitutes for the linear Poisson equation. While the first is a model of plasma physics, the second is derived as a geometric approximation to the Euler incompressible equations. Using energy estimates, convergence of both systems to the Euler incompressible equations is proved.
1 Introduction
In this paper we consider a model of a collisionless plasma where the ions are supposed to be at rest and create a neutralizing background field. The motion of the electrons can then be described by using either the kinetic formalism or the hydrodynamic equations of conservation of mass and momentum as we do here. The self-induced electric field is the gradient of a potential that depends on the electron’s density
either through the linear Poisson equation:
, or through the fully non-linear Monge-Ampere equation:
. This gives the Euler-Poisson (
) system and Euler-Monge-Ampere (
) system. The non-dimensional rescaled version of both systems is the following:
| |
| |
| |
| |
Note that the systems are pressureless, and the only force is due to electrostatic interaction. The energy of those systems is given by
| |
for Euler-Poisson and
| |
for Euler-Monge-Ampere. The asymptotic we look at consists in considering large scales compared to the Debye length (
).
At those scales the plasma appears to be electrically neutral. In this limit the plasma is expected to behave like an incompressible fluid, therefore governed by the incompressible Euler equation
.
We intend to rigorously justify those limits in the present work.
Physical interpretation of the quasi-neutral limit for Euler-Poisson The complete model of collisionless plasma describes the behavior of two species: the ions and the electrons. However the ratio of the electron’s mass and the ion’s mass is of several orders of magnitude, therefore we make the assumption that the ions are at rest, and distributed over a regular grid. This assumption will imply the neutralizing background electric field (the ’-1’ term in the Poisson equation
).
The parameter
comes from the vacuum permittivity, obtained after many rescalings of the equation.
The typical value of
is between
and
. We consider therefore
as a small parameter, and investigate the limit
of the Euler-Poisson system. Note that if
is a solution to the Euler-Poisson system with
,
is a solution of the Euler-Poisson system with parameter
. Therefore, the limit
can be interpreted as a study of the long time large scale behavior of the system.
Geometric interpretation of the quasi-neutral limit for Euler-Monge-Ampere Whereas the Euler-Poisson system relies on a well known physical model, the Euler-Monge-Ampere system, less famous, is a fully non-linear (but asymptotically close in the quasi-neutral regime) version of the Euler-Poisson system; it can be seen as a non-linear model of electrostatic interaction with the advantage of allowing finite electric field for point charges (see also [?] where the Born-Infeld system of electromagnetism is studied, a system that exhibits similar non-linear features). Apart from this interpretation, the main motivation for the study of the Euler-Monge-Ampere system is the following: it appears as a ’canonical’ relaxation of the geodesics on the group of measure preserving diffeomorphisms (therefore of the Euler incompressible equation, see [?] ). This interpretation will be developed more accurately in the section 3.1 devoted to the Euler-Monge-Ampere equation. This model was first introduced, in a discrete version, by Y. Brenier in [?] . Later, a kinetic version, the Vlasov-Monge-Ampere system, has been introduced and studied by Y. Brenier and the author in [?] .
The present work may be seen as a further step in this study.
To see why the
and the
systems should be asymptotically close in the quasi-neutral regime, notice that if
is close to 1 then
should be small, hence
and one recovers the Poisson equation. For this reason the proof of the convergence of both systems will be very close and this is why we present them altogether.Related results concerning singular perturbations This work is concerned with the motion of slightly compressible fluids seen as singular perturbations of the Euler incompressible equation; this field has been widely investigated using different techniques:
Traditional analysis and geometry on the group of volume preserving diffeomorphisms (see section 3.1 ) as done in [?] where the convergence holds in
norm,
large, for well prepared initial data, restricted to the case of barotropic fluids (i.e. when the pressure is a local function of the density
, a case different from the one studied here.) Energy estimates as done in [?] again for the case of barotropic fluids, where convergence holds in all
norms for well prepared initial data. The result has also been extended to the non-isentropic case by [?] .
Pseudo-differential energy estimates as done in [?] which can be seen as a pseudo-differential generalization of [?] and where the same convergence holds for a broader class of singular perturbations, including non-local dependence between density and pressure.
Modulated energy techniques for convergence of the Vlasov-Poisson system to the so-called dissipative solutions of the Euler equation, as done in [?] , and for the Vlasov-Monge-Ampere system (a kinetic version of the Euler-Monge-Ampere system) as done in [?] . The convergence result obtained there holds in weighted “
” norms, this method has the advantage to be valid for weak solutions, and does not require any smoothness of the solution. We give here more details on this result:
Modulated energy technique for Vlasov-Poisson and Vlasov-Monge-Ampere Those systems are the kinetic extensions of the
and
systems. They read as follows:
|
(1)
|
|
(2)
|
|
(3)
|
|
(4)
|
In [?] and [?] , the following results have been obtained:
Theorem 1.1
Let
be a weak solution of ( 1 , 3 , 4 ) (resp. of ( 1 , 2 , 4 )) with finite energy, let
be a smooth solution of the incompressible Euler equation ( 8 ) for
and
the corresponding pressure, let
| |
where
(resp.
in the Poisson case), then
| |
The constant
depends only of the
norm of
Results Here we shall obtain by energy estimates a convergence to the Euler incompressible system in
norm for any
large enough (the minimal smoothness will be made precise). The convergence of both systems holds on the range of time on which the solution of Euler is smooth enough (roughly speaking, we will need at least
to be bounded in
). Our work is based on the modulated energy techniques, restricted to the case of monokinetic velocity profiles. Indeed, the quasi-neutral limit is much more difficult without this assumption, and is even known to be false in some cases (the two-streams instability, see [?] ). We will mostly restrict ourselves to the case of well-prepared initial data, but we will investigate briefly the case of non-prepared initial data: in that case the divergence part of the initial velocity is not assumed to be small, and we only assume that the initial fluctuations of the electronic density
are of order
, so that the energy remains bounded.
The electric field is expected to oscillate at frequency
and with amplitude
. This case will be treated in section 2.3, and we will obtain that the divergence-free part of the velocity converges strongly to a solution of the Euler incompressible equation while its potential part stays bounded, but oscillates strongly with respect to time, and therefore converges weakly to 0.
We obtain also that both systems are closer to each other than they are close to the Euler incompressible system;
is thus a corrector in the convergence of
to
.
Although the operators that define the acceleration from the density are differential operators (and even fully non-linear in the second case), our proof does not use the pseudo-differential formalism.
Actually, we were not able to use the general theorem obtained by Grenier [?] for singular perturbations:
it might be because of the absence of pressure that changes the symmetrizers of the system. After a convenient change of variable however, the system appears under a form which is strongly reminiscent (at least for the highest order terms) of the rapidly rotating fluids. This limit has been treated in [?] .
Finally we also mention the work of Cordier & Grenier [?] , and Wang [?] , where the quasi-neutral limit ’with pressure’ is treated. The techniques used there do not apply here, and it is worth noting that the scaling obtained are not the same. The reader can also refer to [?] and [?] where different regimes of the Euler-Poisson system are studied.
We split the rest of the paper in two sections: the first one devoted to the study of the Euler-Poisson system, and the second devoted to the study of the Euler-Monge-Ampere system.
1.1 Notation
Hereafter
and
;
stands for the velocity and
is the macroscopic density of electrons;
is the electrostatic potential;
or
.
It is always assumed that
has total mass equal to 1.
The divergence of a vector field
will be denoted by
or
; its rotational (or curl) will be denoted by
or
.
The components of a vector will be denoted with superindices, i.e.
.
In all the paper,
will denote the integer part.
We denote respectively by
,
,
the systems Euler incompressible, Euler-Poisson, Euler-Monge-Ampere.
2 The Euler-Poisson system
We consider the following Euler-Poisson system denoted by
:
|
(5)
|
|
(6)
|
and consider the limit
going to 0. We recall also the incompressible Euler equation
:
We recall (see [?] for a reference on the topic) that in the periodic case, the Cauchy problem for ( 8 ) is well posed in
if
. More precisely, if
, for any divergence-free initial datum in
, there exists a unique global solution in
; if
one can only prove the existence of a smooth solution in finite time, belonging to
for some
. We will then prove the following:
2.1 Result
Theorem 2.1
Let
with
. Let
be a divergence-free vector field on
. Let
be a smooth solution of the Euler incompressible system ( 8 ) on
, with initial data
, satisfying
. Let
be a sequence of initial data such that
, and such that
is bounded in
.
Then there exists a sequence
of solutions to
with initial data
, belonging to
, with
. Moreover for any
and
small enough,
is bounded in
. Finally when
,
goes to infinity.
Remark. The models that we consider are valid in domains without boundaries, and although stated in the space periodic case, we believe that our results hold true, with some technical adaptation, in the case of the whole space.
2.2 Proof of Theorem 2.1
2.2.1 Heuristics.
Let us introduce
the solution of the Euler incompressible system ( 8 ) and corresponding pressure. Note that by taking the divergence of ( 8 ) the pressure is given by the following:
| |
We will all along the paper use the following notation: for two vector fields
,
| |
If
is solution to
, we introduce also
We suppose for now that
, and we take the curl and the divergence of the momentum equation, this yields
| |
| |
| |
If we assume for now that
and their spatial derivatives remain bounded (we do not specify in what sense yet),
are bounded terms. Hence the vorticity
is not affected by the electric field, and the vector
evolves through
| |
with
bounded, and
. Under this form, the system looks like a rapidly rotating fluid (up to the remainder
, and the fact that
) and the singular term
induces time oscillations of frequency
, but does not increase the energy of the perturbation, allowing energy estimates, as long as the remainder
is under control.
2.2.2 Reformulation of the system with new unknowns
For a vector
, we denote
its components. We define the new unknowns
as follows:
| |
| |
with
the solution of
, and
.
We will use the following observation:
Lemma 2.2
Let
be a solution to
. Then, the total momentum
does not depend on time.
Proof. The momentum equation can be rewritten
| |
We just have to show that the integral of the right hand side vanishes. For this, we use the identity
that yields 0 when integrated over
.
We note also
but
is not really an unknown since it can be obtained from the knowledge of
, and
: Indeed when
we have
|
(9)
|
|
(10)
|
In the 3 dimensional case we have equations ( 9 , 10 ) replaced by
| |
and thus
|
(11)
|
Note that when
the vorticity is scalar and when
it is a vector field of
. We show now that
can be estimated from
,
, and
.
Lemma 2.3
Let
be the solution at time
of
with initial datum
, let
be a solution at time
of
with initial datum
. Then we have, for
,
| |
| |
Proof. Let
be the unique vector field with zero average such that
| |
We have directly from identities ( 9 , 10 , 11 ) that
, which is the first point of the lemma. The difference
is a constant vector field, and we have
| |
| |
which yields
| |
and the result follows.
We immediately deduce the following corollary:
Corollary 2.4
Let
be a solution to
, let
be a solution to
. Then, for any
,
| |
| |
| |
Taking the curl of equation ( 6 ) we recall the following identities:
|
(12)
|
|
(13)
|
When
the
system then becomes:
|
(14)
|
|
(15)
|
|
(16)
|
whereas when
one would have to replace equation ( 14 ) by
|
(17)
|
Noticing that
, if we set
we get the following system for
:
|
(19)
|
When
, the first equation should be replaced by
| |
2.2.3 Energy estimates
We handle the energy estimates when
but the same result would hold when
, just with more terms. The system can be written in the following way:
|
(20)
|
where
is still the velocity, and where
| |
The ’source’ term
is given by
| |
We apply
to equation ( 20 ), where
, and
stands for
, with
.
We get
| |
where
| |
Then we have
Lemma 2.5
If
, then for
and
defined as above we have:
| |
and
| |
Proof. Point 1: the proof is a straightforward adaptation of the proof of [?] p.151. It is based on the following estimate: [?] Proposition 2.1.2 p. 100:
Proposition 2.6
If
, then for any
, one has
| |
Applying this result to
,
, we obtain
| |
We know that
if
. Thanks to ( 9 , 10 ) we have for any
|
(22)
|
thus
. Then using that
we conclude.
Point 2: We also know thanks to Proposition 2.6 that for
| |
It follows that for
, we have
| |
where
depends on the smoothness of the solution of ( 8 ).
Having applied
to ( 20 ), multiplied by
, and noticed that for any
,
, we obtain
| |
| |
Since
if
, using Lemma 2.5 , and integrating over
we have, for any
,
:
| |
By summing this over all multi-indexes
, we can conclude using a standard Gronwall’s lemma, that if the solution
of Euler is smooth (see below for the smoothness required) on the time interval
, for any
there exists
such that the sequence
is bounded in
.
Then we have
| |
We use Lemma 2.3 to get
from
, and from Corollary 2.4 , the bound obtained on
implies a bound on
(Note that from the assumption on the initial data, we have
.) This proves Theorem 2.1 .
Minimal regularity for the limiting field
In order to perform our computations, we need to have at least
in
and thus to have an estimate on their norms in
, with
. Therefore we need to control
. Then we need to apply
to 19 , with
, and we need to control
(this is the ’worst’ term). This implies to control
. which requires
to be bounded in
with
. If we take
integer, we ask
, and
must be bounded in
.
Remark. Usually, modulated energy techniques only require a bound on
. Here we need one more derivative, since the ’div-curl’ formulation of the system (performed in order to obtain energy estimates) is obtained by differentiation.
2.3 Non-prepared initial data
Here we obtain energy estimates in the general case of non-prepared initial data. What we mean by ’general case’ is the case of a generic smooth initial velocity, and smooth initial density, with finite energy, hence
will be of order
.
We will see that the energy estimates are the same as in the case of prepared initial data, the asymptotic
is then handled similarly to [?] , although the algebra in our case is quite simple.
The solution will exhibit a good space regularity and a strongly oscillating behavior with respect to time. As explained in [?] , the motion can be decomposed along a slow and a fast manifold; the slow manifold consists of divergence-free velocities with uniform density, and the fast manifold consists of potential velocities. Due to the rapid oscillations, the potential part of
will converge weakly to 0, and the divergence-free part will converge strongly to a smooth solution of the incompressible Euler equation.
We still consider
solution to
. For any vector field
, we introduce its solenoidal part and potential part, which is the pair
such that
with
periodic and
.
A priori estimates We first rescale the density fluctuation
We use the unknown
, given by
| |
(Note that we do not subtract
to
in this case.) We now restrict to the case
, but one can check easily that the same results hold when
.
We write the equation followed by
:
|
(23)
|
where
is a before, and the source term
is now given by
| |
When
, proceeding as in Lemma 2.5 , we have the estimate
| |
Arguing as in the previous case, we conclude that, given
and a sequence of initial data
, with
, and such that
| |
there exists a sequence
of solutions to
with initial data
, and
remains bounded in
for some
, independent of
.
Convergence We use the change of variable used for rotating fluids (see [?] ), that removes the time oscillations: Considering the pair
such that
| |
we have
| |
with
| |
| |
| |
from the a priori bounds ( we still have
). Hence,
is uniformly bounded at least in
, and we deduce classically that
converges weakly to 0 in
.
We assume now that
, the solenoidal part of
, converges weakly in
to some limit
, hence it converges strongly in
for
. We check that
is bounded uniformly on
under our assumptions. Hence from the a priori bound,
converges (if necessary passing to a subsequence, but see the remark below) in
for all
.
Let
be the unique vector field with zero average such that
. Then,
, the solenoidal part of
is equal to
with
a constant vector field. Since
, and since
is bounded in
, we have
. Hence
converges in
for all
.
Decomposing
as
, we have
that converges weakly to 0 in
. Hence, using the a priori bounds, in the vorticity equation
| |
we can pass to the weak limit in
and state that
satisfies in
,
where
is the limit of the solenoidal part of
. Moreover we have
, hence
is a solution to the incompressible Euler equation, with initial data
the limit of the solenoidal part of
.
Remark 1. From the regularity of
, the solution (in the distribution’s sense) to
with initial data
is unique. Therefore the whole sequence
is converging.
Remark 2. Here we did not introduce the solution of the limit equation (incompressible Euler in this case), and chose to argue by compactness. This method looks simpler, however, we obtain less informations concerning the ’rate of convergence’ of the sequence
with respect to
.
We gather those results in the following theorem:
Theorem 2.7
Let
, let
be a sequence of initial data, such that
.
Assume that
is bounded in
and that
is bounded in
. Then
-
1.
There exists
such that a sequence of solutions to
with initial data
exists on
, and the sequence
is bounded in
.
-
2.
The potential part of
, defined above, converges weakly to 0 in
.
-
3.
If
, the solenoidal part of
converges to some
weakly in
, then, for all
,
, the solenoidal part of
, converges in
to
the solution of
with initial data
.
3 The Euler-Monge-Ampere system
We consider here the following Euler-Monge-Ampere system denoted by
:
|
(25)
|
|
(26)
|
The last equation must be understood in the following weak sense:
is the only (up to a constant) convex function with
being
periodic such that
|
(28)
|
This definition will be made precise in Theorem 3.1 and Definition 3.2 .
3.1 Geometric derivation of the Euler-Monge-Ampere system
This derivation has been introduced in [?] , [?] , [?] . We reproduce it for sake of completeness, but the reader only interested in the proof of convergence may skip this section.
3.1.1 The Euler equations of an incompressible perfect fluid
The motion of an inviscid and incompressible fluid in a domain
is described by the Euler incompressible equation
that we recall here:
| |
Following Arnold (see [?] ), we have a formal interpretation of the Euler incompressible equations:
introducing
the group of all volume preserving diffeomorphisms of
with jacobian determinant equal to 1, the Euler equations describe the geodesics of
with length measured in the
sense.
3.1.2 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
|
(29)
|
In this equation, the unknown
is a curve in
,
is the distance (in
) of
to the manifold
, i.e. in our case when
,
|
(30)
|
finally,
denotes the gradient operator in
. This penalty approach has been introduced for the Euler equations by Brenier 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 [?] . Actually if
were a smooth manifold, the result would be exactly the one of [?] , Theorem 2.7, but this is not the case, here because the
metric is too weak. The penalized system is formally hamiltonian in variables
with hamiltonian (or energy) given by:
| |
Multiplying equation ( 29 ) by
, we get immediately that the energy is formally conserved.
Therefore it is plausible that the map
will remain close to
if it is close at
. A formal computation shows that, given a point
for which there is a unique closest point
to
in the
closure of
, we have:
|
(31)
|
Thus the equation ( 29 ) formally becomes:
| |
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 orthogonal to
thanks to ( 30 ) and
remains close to
.
Notation Since we intend to work on the flat torus
we might consider
additive mappings, i.e. mappings such that
as well as periodic mappings (i.e. mappings from
into itself ).
Then given
an additive mapping,we denote by
the naturally associated mapping on
. The following polar factorization Theorem is a periodic version of [?] , it has been discovered independently by [?] and [?] .
Theorem 3.1
Let
be
additive and assume that
has a density in
, then there exits an a.e. unique pair
satisfying
with
a convex function such that
is
periodic, and
additive such that
is measure preserving in
. Moreover we have
and if
is the Legendre transform of
then
Remark 1. The pair
is uniquely defined by the density
.
Remark 2. Important properties of the optimal potential: The periodicity of
implies that
and
are
additive, and that
is also
periodic. This allows the following definition:
Definition 3.2
Let
be a probability measure on
, then we denote
(resp.
) the unique up to a constant convex function such that
| |
| |
(resp. its Legendre transform).
Remark. We recover thus that
will be generalized solutions of the following Monge-Ampere equations
| |
3.1.3 Result
We are now ready to state the main result of this section.
Theorem 3.3
Let
with
. Let
be a divergence-free vector field on
, let
be a smooth solution of the Euler incompressible system ( 8 ) on
, with initial data
and satisfying
. Let
be a sequence of initial data, with
, and such that
is bounded in
. Then there exists a sequence
of solutions to
with initial data
belonging to
with
. Moreover for any
and
small enough,
is bounded in
. Finally when
,
goes to infinity.
Before entering the proof of this result, we need some preliminary results concerning the linearization of the Monge-Ampere operator.
3.2 Linearization of the Monge-Ampere operator in
norm
This section is devoted to the proof of the following Theorem:
Theorem 3.4
Let
be a probability measure on
,
, let
satisfy
in the sense of Definition 3.2 . Then, there exists
such that if
, for any
with
there exists
that satisfies
|
(32)
|
|
(33)
|
3.2.1 Preliminary results
We first state the following result obtained from [?] on the regularity of solutions to Monge-Ampere equation, adapted to the periodic case. This result will be the starting point of the proof of Theorem 3.4 .
Theorem 3.5
Let
for some
, with
, be a probability measure on
, let
in the sense of Definition 3.2 . Then
is a classical solution of
and satisfies for any
:
|
(34)
|
|
(35)
|
Then we state a classical result of elliptic regularity that we will need during the course of the proof. It can be found in [?] , Theorem 9.11.
Theorem 3.6
Let
be an open set in
,
, be a strong solution of the equation
| |
in
where the coefficients
satisfy
| |
| |
for
, with
. Then for any
,
| |
where
depends on
and the moduli of continuity of the coefficients
on
.
3.2.2 Proof of Theorem 3.4
Sketch of the proof We assume here
. We recall that
satisfies
. We first have to prove that
implies
. We will proceed by induction. We recall first that for
two
matrices, we have the following expansion:
| |
where
is the matrix whose elements are the minors of
, or co-matrix of
. Hence the elements of
are polynomials of degree
in the elements of
. When
is invertible, we have
.
Differentiating
times the Monge-Ampere equation, and denoting
the co-matrix of
, we will have
| |
where the first term contains the highest derivatives, and
will consist of products involving three derivatives of
. The order of each derivative will smaller or equal to
, and the sum of the three orders will be equal to
. By a careful analysis, this product will be controlled in
by
and
, using Sobolev injections. From Theorem 3.5 , assuming a minimal regularity for
(i.e.
the bound ( 37 )),
, and therefore
will be continuous elliptic matrices. Hence
solves an elliptic problem, with continuous coefficients, and we will use the Theorem 3.6 to obtain
.
This intermediate step will be done in Lemma 3.7 ; using this a priori estimate and the continuity method, we will obtain the estimate ( 32 ).
Then, the expansion
, where
is a polynomial in
whose terms are of degree two or three (when
), will yield ( 33 ).
Rigorous proof We recall that
satisfies
|
(37)
|
for some
to be chosen later. We suppose
and the proof can be reproduced in the case
with minor modifications. The parameter
is chosen such that ( 37 ) implies
for some
. Note also that thanks to ( 37 ),
is in
for
. Then from Theorem 3.5 ,
with
. Note also that since
and using equation ( 36 ),
implies that
, and thus
the co-matrix of
is uniformly elliptic and in
.
We first prove by induction that if
then
implies
. It can be checked during the proof that this bound will be uniform under the condition ( 37 ) for
small enough.
Lemma 3.7
Under assumption ( 37 ), for any
,
implies that
. If moreover
then
.
Proof. This lemma will be proved by induction. We first deal with the cases
.
The case
is a consequence of Theorem 3.5 .
For
we differentiate ( 36 ) with respect to
, to obtain
|
(38)
|
with
the co-matrix of
. Then if
, by Theorem 3.6 ,
. If
we also get that
.
For
differentiating once more with respect to
we obtain
|
(39)
|
still with
the co-matrix of
. Suppose that
, then
if
, and
. The term
is a sum of terms of the form
and the second term of the left hand side of ( 39 ) is thus bounded in
. Then once again by Theorem 3.6 one gets that
if
.
Moreover if
then
for some
. Using ( 38 ) and Schauder interior estimates (see [?] , Theorem 6.2.), we obtain
. Thus
. From ( 39 ) and Theorem 3.6 we obtain
.
As we just saw, Lemma 3.7 is true for
. We assume that it holds for all
with
for some
. Take now
,
, and apply
to ( 36 ):
|
(40)
|
with
some constant coefficients. We call
the second term of the left hand side of ( 40 ). Since
,
, and we have
for any
using the induction hypothesis.
Therefore
and since
we obtain
. Using Theorem 3.6 it follows that
.
Remember that
thus
, and
. Since
, we have
for some
and thus
,
are in
, moreover
and since
,
is in
. Therefore
is in
. If
we have
in
. Hence the lemma holds for all
. This achieves the proof of Lemma 3.7 .
Now by induction on
we prove ( 32 ) and ( 33 ). From ( 37 ) we have
small. Take
solution of
with
periodic and
. We begin to show that
for some
, is controlled by
. Indeed, the periodic solution of
can be built by the continuity method (see [?] ). Starting from
, we use the implicit function Theorem to obtain the solution
of
|
(41)
|
For this we differentiate ( 41 ) with respect to
, to obtain
|
(42)
|
for
, where
is the co-matrix of
. We know, from the a priori estimate of Theorem 3.5 , that for
,
and therefore
are
elliptic matrices for all
.
To see why ( 42 ) indeed admits a unique (up to a constant) periodic solution, we recall that
is the comatrix of a Hessian matrix, therefore it is ’divergence-free’:
| |
Hence equation ( 42 ) can be rewritten in divergence form
| |
and the operator
is a self adjoint operator on
and induces a bounded coercive bilinear form on
, where the subscript 0 means that we impose the mean value to be 0. Then the existence/uniqueness of a solution to ( 42 ) in
follows by Lax-Milgram Theorem. Hence,
is the unique (up to a constant) periodic solution of the above elliptic problem, and from Schauder interior estimates we obtain
| |
uniformly in
, and finally for
,
| |
Then we have
| |
where
is a symmetric matrix whose coefficients are polynomials in
of degree larger or equal to 1. The norms
are controlled by
, hence, for
small enough, the matrix
is uniformly bounded, elliptic, and
continuous. Since
satisfies
|
(43)
|
it follows from Theorem 3.6 that
and this proves ( 32 ) for
.
If
, we have
| |
with
uniformly bounded, elliptic and
continuous. For the same reasons we have
.
If
, we do as in the Proof of Lemma 3.7 : using ( 38 , 39 ) and keeping track of the bounds, we get
| |
with
uniform under the assumption of Lemma 3.7 .
If
, we go back to equation ( 40 ):
is a sum of terms which contain all a product of at least two derivatives of
of degree higher than 3. Since
, we have
| |
We assume by induction that
. Since
we also have that
, with a uniform bound thanks to Lemma 3.7 . We remember that when
, we have
, therefore, using the injection of
in
when
,
and
are uniformly bounded in
. We obtain that
| |
Then
satisfies
| |
thus, using Theorem 3.6 ,
| |
| |
and we conclude that
| |
for
and under condition ( 37 ); thus ( 32 ) is obtained.
Using Proposition 2.6 and the fact that
is uniformly bounded in
thanks to Lemma 3.7 , we can also obtain that
| |
for
. (When
, the estimate holds also, but not using Proposition 2.6 .) Therefore, for all
, we have
.
To conclude ( 33 ), we now write
under the form
| |
with
consisting of products of two or three second derivatives of
. Hence, under assumption ( 37 ), using Proposition 2.6 , we have for
,
. Using the bound ( 32 ), we conclude that for
,
| |
and Theorem 3.4 is proved.
3.3 Energy estimates and proof of the convergence
The proof of the energy estimates for Euler-Monge-Ampere is much inspired from the proof of Theorem 2.1 for the following reason: by taking the divergence of equation ( 26 ) one gets:
| |
Suppose that
is close to 1 at an order
as is the case for Euler-Poisson, we guess (from Theorem 3.4 ) that we have the following:
| |
| |
and thus
. Therefore we expect that
| |
and that the technique of Theorem 2.1 will apply.
Before performing the div-curl decomposition for the energy estimates, we need to establish the analog of Lemma 2.2 in the present case, so that Lemma 2.3 and its Corollary 2.4 hold.
Lemma 3.8
Let
be a solution to
. Then, the total momentum
does not depend on time.
Proof. We proceed similarly as in the proof of Lemma 2.2 . We need to establish that
For this we use Definition 3.2 . For
, we have
| |
| |
where we have used at the third line that, for
Legendre transform of each other,
, and at the last line that
is periodic.
Hence we have shown that one can retrieve
from the initial value of
, and
.
General framework We perform the same div-curl decomposition as in the Euler-Poisson case.
We then express the difference between the solution of
and the limiting solution : either the solution of
or the solution of
. After having applied a proper scaling to this difference, our solution is now described by a vector
whose first component (that can be a vector if
) is the rescaled vorticity, and whose last two components are a rescaled divergence and rescaled density fluctuation. For this perturbation we will obtain
| |
where we still use
| |
For the source term
, we have
where
goes to 0 when
goes to 0, and the constant
depends on the regularity of the limiting field.
Regularity of the limiting field The form of the source term will vary under the circumstances, but the general idea is that in order to bound
in
, we will need the limiting velocity to be bounded in
and the limiting density to be bounded in
. Remember that the
norm of
controls the norm of
in
, thus the limiting field must have one more derivative bounded than the order of the energy estimate.
A Gronwall’s lemma then yields a control on the perturbation that holds on a range of time
, where
,
being the time of existence of a smooth solution for the limiting equation.
Convergence to Euler, two dimensional case Doing the same change of variables as in the proof of Theorem 2.1
| |
we obtain:
|
(44)
|
|
(45)
|
|
(46)
|
Now we define
by
| |
and from Theorem 3.4 inequality ( 33 ), we have, if
,
. The system can here be written in the following way:
| |
still with
| |
with the same
as in the Euler-Poisson case,
, and with
We have
, for
large enough. Then the energy estimates are the same as in the first proof, the solution
satisfying a control of the form:
| |
and the same conclusion holds true. Then from Corollary 2.4 ,
can be retrieved from
, and we obtain the expected conclusion.
Convergence to Euler, three dimensional case In the 3-d case, equation ( 44 ) should be replaced by
| |
Note that the vorticity equation is the same as in the Euler-Poisson case. This change would not affect the energy estimates.
Higher order approximation Here we prove that the the Euler-Poisson system and the Euler-Monge-Ampere system are closer as
goes to 0 than Euler-Poisson and Euler. We fix
. For
a
smooth divergence-free vector field, we consider
a solution of the Euler incompressible system ( 8 ) such that
for some
. We consider also a sequence
of solutions of the
system with initial data
(with
) such that
is bounded in
, for any
, if
is small enough. Thanks to Theorem 2.1 , and from the regularity assumption made on
, such a sequence exists for any sequence of well prepared initial data.
Theorem 3.9
Let
with
. Let
be as above. Let
be a sequence of initial data such that
is bounded in
.
Then there exists a sequence
of solutions to
with initial data
, belonging to
, with
. Moreover for any
and
small enough, the sequence
is bounded in
.
Remark. We see here that the difference between solutions of
and
is of order
for the density and of order
for the velocity whereas the difference between solutions of
(or
) and Euler was of order
for the density and of order
for the velocity.
Proof. We give the proof when
, the proof would be the same when
, just with more terms. We introduce
solution to
, and
. Then we set
| |
| |
The system
now reads:
|
(47)
|
|
(48)
|
|
(49)
|
We still define
by
| |
and from Theorem 3.4 we will have
| |
(we use the notation
).
Setting
we obtain that
| |
with
as before and
defined by
| |
Using again Proposition 2.6 as in Lemma 2.5 we obtain that, for
,
| |
where the constant
is controlled by
(still with
). From Theorem 2.1 these quantities are controlled for
,
being the time on which the solution of
is smooth. Hence we have by Gronwall’s lemma a bound on
.
Arguing as in the previous proofs, and using Corollary 2.4 , we obtain that
remains bounded in
for any
and for
small enough. It follows that
remains bounded in
. This achieves the proof of Theorem 3.9 .
3.4 Non-prepared initial data
In this case, we obtain exactly the same result as for Euler-Poisson, using the same techniques.
We follow closely section 2.3, and we only have to estimate the additional source term that will appear in the equation followed by
, due to the Monge-Ampere coupling. We recall that
, and we will have to estimate the difference
| |
in
when we know that
is bounded in
. Thanks to Theorem 3.4 , we conclude that for
this term is controlled by
. Hence the energy estimate can be handled similarly just with an additional term, and the conclusion remains true.
Theorem 3.10
The Theorem 2.7 holds also when replacing the
system by the
system.
Gregoire Loeper EPFL, SB-IMA 1015 Lausanne, Switzerland gregoire.loeper@epfl.ch