ArticlePDF Available

Long time confinement of vorticity around a stable stationary point vortex in a bounded planar domain

Authors:

Abstract and Figures

In this paper we consider the incompressible Euler equation in a simply-connected bounded planar domain. We study the confinement of the vorticity around a stationary point vortex. We show that the power law confinement around the center of the unit disk obtained in [2] remains true in the case of a stationary point vortex in a simply-connected bounded domain. The domain and the stationary point vortex must satisfy a condition expressed in terms of the conformal mapping from the domain to the unit disk. Explicit examples are discussed at the end.
Content may be subject to copyright.
Long time confinement of vorticity around a stable
stationary point vortex in a bounded planar domain
Martin Donati and Dragos
,Iftimie
Abstract
In this paper we consider the incompressible Euler equation in a simply-connected
bounded planar domain. We study the confinement of the vorticity around a stationary
point vortex. We show that the power law confinement around the center of the unit
disk obtained in [2] remains true in the case of a stationary point vortex in a simply-
connected bounded domain. The domain and the stationary point vortex must satisfy
a condition expressed in terms of the conformal mapping from the domain to the unit
disk. Explicit examples are discussed at the end.
1 Introduction and main result
To study the behavior of an incompressible inviscid fluid, we consider the planar Euler equa-
tions in a bounded domain R2:
tu(x, t) + u(x, t)· ∇u(x, t) = −∇p(x, t),(x, t)×R
+
u(x, 0) = u0(x),x
∇ · u(x, t) = 0,(x, t)×R+
u(x, t)·~n = 0,(x, t)×R+
(1.1)
where udenotes the velocity of the fluid, pits internal pressure, and nis the exterior normal
to . If Ω = R2the boundary condition should be changed into a vanishing condition at
infinity. We define the fluid’s vorticity by ω=1u22u1, which satisfies the equation:
tω(x, t) + u(x, t)· ∇ω(x, t) = 0.(1.2)
If the domain is smooth, then we have a unique global smooth solution of (1.1), see
[17]. In addition, if C1,1we have the following result due to Yudovitch (see [18]): for
any ω0L1L, there exists a unique solution to (1.1), with uL(R+, W 1,p), for every
1<p<, and ωL(R+, L1L). The result is true in more general domains, in
particular in domains with finite number of corners with angle strictly less than πand when
the vorticity is compactly supported in , see [13] and [6]. In particular, one could take
to be a convex polygon.
1
Let us denote by Gthe Green’s function of . Since the domain is supposed to be
simply-connected, the velocity can be recovered from the vorticity through the following
Biot-Savart law
u(x, t) = Z
xG(x, y)ω(y, t) dy, (1.3)
where x= (x2, x1).
The point vortex system is a simplified version of the Euler equations where the vorticity is
assumed to be a finite sum of Dirac masses ω0=PN
i=1 aiδzi. It was introduced by Helmholtz
in [8], see also [15]. Since (1.2) is a transport equation, one expects the vorticity to remain
a sum of Dirac masses at some points zi(t). The Biot-Savart law (1.3) reads in this case
u(x, t) =
N
X
i=1
ai
xG(x, zi(t)).
However, if xis one of the points zi(t), then this velocity is not defined as G(x, y)is singular
at y=x. But as xapproaches zi(t), the singular part of the velocity defined above is given
by fast rotation around that point. More precisely, since the map GGR2is harmonic in
both its variable on , the function γ: Ω ×R,γ=GGR2=G1
2πln |xy|is
smooth. So the singular part of
xG(x, zi(t)) is given by (xzi)
2π|xzi|2. The point vortex system
consists in ignoring this singular part which should have no influence on the motion of zi
itself. Denoting by eγ(x) = γ(x, x)the Robin function of the domain we obtain then the
following point vortex dynamic:
1iN, dzi(t)
dt=
N
X
j=1
j6=i
aj
xG(zi(t), zj(t)) + ai
1
2eγ(zi(t)),(1.4)
where the (ai)1iNR\ {0}are the masses of the point vortices (zi(t))1iN. Equations
(1.4) are called Kirchhoff-Routh equations.
Due to the singularities of the Green’s function, these equations are valid only while the
points zi(t)stay distinct and do not leave the domain . There exist configurations leading
to collapse of the point vortices, but they are exceptional, see [15] for the case of R2, and [14]
for the case of the unit disk (we will discuss this more in detail later).
An important question in fluid dynamics is whether the point vortex system is a good
approximation of the Euler equations. There are convergence results in both ways.
Let us first mention that the point vortex system was used as a numerical approximation
of the Euler system. More precisely, consider a smooth solution of the Euler equations and
construct an initial discrete vorticity which is a sum of Dirac masses located on a grid (hj)jZ2
where hRis the length of the grid, with masses h2ω0(hj). Solve then the point vortex
system with this initial vorticity. In [4], the authors proved that this point vortex method is
consistent, stable, and converges to the smooth solution of the Euler equation.
The convergence from the Euler system to the point vortex system was also proved in
[16]. More precisely, consider a smooth initial vorticity that is sharply concentrated around
some initial point vortices zi: the support of ω0is included in the union of the disks of radius
εaround the points zi, with ε > 0being small. The authors of [16] proved that for any
2
time τ, and for any δ > 0, if ε=ε(τ, δ)>0is small enough then the solution stays sharply
concentrated within disks of radius δfrom the points zi(t)up to time τ. This statement can
be seen as a fixed time confinement result.
In this paper we are interested in the so-called long time confinement problem, that is we
want to know for how long confinement around point vortices remains true. More precisely,
we want to understand how ε,δand τare linked together and we wish to obtain a confinement
time τas large as possible. We already know from [16] that τgoes to infinity when εgoes
to 0, but we would like to obtain an explicit rate as good as possible.
This problem was already studied by Buttà and Marchioro [2]. These authors assumed
that δ=εβ, with β < 1/2, and made the following assumptions on the initial vorticity.
Assume that ω0L1Land there exists νsuch that
|ω0| ≤ εν
ω0=
N
X
i=1
ω0,i,supp ω0,i D(zi, ε)
ω0,i has a definite sign
Z
ω0,i dx=ai.
(1.5)
Let ω(x, t)the solution of (1.2). We denote by τε,β the exit time of the vorticity from the
disks of radius εβ:
τε,β = sup (t0,s[0, t],supp ω(·, s)
N
[
i=1
D(zi(s), εβ)).(1.6)
For any N-tuple of distinct points (zi), there exists εsmall enough, such that the disks
D(zi(0), εβ)are disjoints, and therefore this exit time is well defined and strictly positive.
The aim is to obtain a lower bound on τε,β depending explicitly on ε. Two results have been
obtained in [2]. The first is a logarithmic confinement for the whole plane.
Theorem 1.1 ([2]).Assume that Ω = R2, that the initial vorticity satisfies (1.5) and that
the point vortex system with initial data Pn
i=1 aiδzihas a global solution. Then for every
β < 1/2there exists ε0>0and C > 0such that
ε<ε0, τε,β > C|ln(ε)|.
The second result is more restrictive, it holds true for the unit disk and for a single point
vortex located at the center, but the conclusion is much stronger since it gives a power-law
confinement.
Theorem 1.2 ([2]).Let Ω = Dand ω0satisfying (1.5) with N= 1 and z1= 0, so that it is
compactly supported within the disk D(0, ε). Then for every β < 1/2there exists ε0>0and
α > 0such that:
ε<ε0, τε,β > εα.
3
The aim of this paper is to extend Theorem 1.2 to more general domains. We also observe
that Theorem 1.1 can also be extended to bounded domains; we will discuss this problem in
a forthcoming paper.
We consider a single point vortex in a simply-connected bounded domain. We assume for
simplicity that the mass of the point vortex is 1, but the results below hold true for a general
mass. The first question that arises is the location of the point vortex. We will show that
there are special points that allow us to obtain the power-law lower bound while for others
the logarithmic bound is probably optimal. The dynamic of a single point vortex of mass a
reduces to the following ODE
d
dtz(t) = a1
2eγ(z(t)).
It is obvious from this ODE that a point vortex is stationary if and only if it is a critical
point of the Robin function eγ. The Robin function has been studied, see [5], and we know
that such critical points always exist in a bounded domain. Let x0be a critical point of the
Robin function which is fixed for the rest of this paper. From the Riemann mapping theorem
we know that there exists a biholomorphic map Tfrom to the unit disk D. We can chose
Tsuch that it maps x0to 0: T(x0) = 0. We shall see below that x0is a critical point of
the Robin function if and only if T00(x0)=0(see Proposition 2.4). This condition therefore
characterizes the fact that x0is a stationary point for the point vortex system. We call such
points stationary points.
Our main result is the following.
Theorem 1.3. Let be a simply connected bounded domain of R2with C1,1boundary. Let
x0be a stationary point such that T000(x0) = 0 where Tis a biholomorphism from to the
unit disk, mapping x0to 0. Assume that ω0satisfies (1.5) with N= 1 and z1=x0. Then
for every β < 1/2and for any α < min(β , 24β), there exists ε0>0such that
ε<ε0, τε,β > εα.
This extends Theorem 1.2 to more general bounded domains. Indeed, in the case of the
unit disk we can choose T(z) = zso the hypothesis given above is verified for the center of
the disk. Let us also observe that the hypothesis that x0is stationary induces no restriction
on the domain . Indeed, we recall that Gustafsson [5] proved that every simply-connected
smooth domain has at least a stationary point. However, the hypothesis that T000(x0)=0is
a condition that not all domains satisfy. We will comment on this in the last section. We
will see in particular that any domain which is invariant by some rotation of angle θ(0, π)
around x0satisfies the condition T000(x0) = 0.
In order to understand better the significance of the condition T000(x0)=0, one could
assume that the vorticity ωitself is a point vortex. We study in detail this perturbation
problem in Section 3. We will prove there that if |T000(x0)|<2|T0(x0)|3then τε,β =if εis
small enough while if |T000(x0)|>2|T0(x0)|3then τε,β is in general not better than C|ln ε|, see
Theorem 3.1. In other words, in this particular case we have long time confinement better
than C|ln ε|if and only if |T000(x0)|<2|T0(x0)|3. However, when ωis smooth we require the
stronger assumption T000(x0)=0.
4
The plan of the paper is the following. In Section 2 we introduce some notation and
discuss some facts about the Green’s function and the point vortex system. In section 3 we
consider the particular case when ωis a point vortex itself. In Section 4 we prove Theorem
1.3. The last section contains some final remarks and some examples of domains for which
our theorem applies.
2 Preliminary tools
List of notation:
is a C1,1bounded and simply connected domain of R2;
D(x0, r)is the disk of center x0and of radius rand D=D(0,1);
uis the velocity of the fluid and pits pressure, satisfying equations (1.1);
ω=1u22u1is the vorticity of the fluid;
supp fis the support of the function f, namely the closure of the set {x, f (x)6= 0};
δzis the Dirac mass in z;
Gor Gis the Green’s function of the domain ;
γor γis the regular part of G, see relation (2.1);
eγ(x) = γ(x, x)is the Robin function;
C, C1, C2, . . . ;K, K1, K2, . . . , L, are strictly positive constants that may vary from one line to
another, when their value is not important to the result;
a·bis the scalar product of vectors in R2;
• ∇f,D2fand ∇ · gare respectively the gradient of f, its Hessian matrix, and the divergence
of g.
2.1 Green’s Function
We recall that the Green’s function of a domain is the solution of
xG(x, y) = δ(xy)
vanishing at the boundary, and at infinity if is unbounded. It is a symmetric function on
2that satisfies for x6=y
x(G(x, y)GR2(x, y)) = 0,
which means that GGR2is a function, denoted by γ, which is harmonic in both of its
variable. Therefore, we have:
G(x, y) = 1
2πln |xy|+γ(x, y)(2.1)
5
where γis symmetric and smooth.
In the particular case of Ω = D, we have that
GD(x, y) = ln |xy|
2πln |xy||y|
2π(2.2)
where y=y
|y|2is the inverse relative to the unit circle. In particular,
γD(x, y) = ln |xy||y|
2π.(2.3)
Let x0. From the Riemann Mapping Theorem (see for instance [1], chapter 6) and
recalling that is simply-connected, we know that there exists a biholomorphic map Tfrom
to the unit disk D. The map Tis unique up to compositions with the biholomorphisms of
the unit disk which are given by
φz0(z) = λz0z
1z0z,
with z0Dand |λ|= 1 arbitrary. Choosing z0and λconveniently, we can assume without
loss of generality that T(x0) = 0 and that T0(x0)is a strictly positive real number. These two
conditions insure the uniqueness of the conformal map T. Let us observe that in Theorem
1.3 the mapping Tis not unique since we only assume that it maps x0to 0. However, the
condition T000(x0) = 0 does not depend on the choice of T(once we assumed that it maps
x0to 0). Indeed, if T1and T2are two biholomorphisms from to Dmapping x0to 0,
then T1T1
2is a biholomorphism from Dto Dmapping 0 to 0. So it must be a rotation:
there exists some λof modulus 1 such that T1T1
2(z) = λz. So T1=λT2and therefore
T000
1(x0) = λT 000
2(x0). Then T000
1(x0) = 0 if and only if T000
2(x0)=0.
We will assume from now on that T(x0) = 0. The assumption that T0(x0)>0can be
made but is not necessary. In the following, T0(x0)is a complex number.
The properties of the conformal map Timply a precise description of the Green’s function
of . Indeed, a Green’s function composed with a conformal mapping is another Green’s
function, see for example [1] chapter 6. Therefore the formula (2.2) yields the following
proposition.
Proposition 2.1. Let Tbe the biholomorphic mapping introduced above. Then
G(x, y) = GD(T(x), T (y)) = ln |T(x)T(y)|
2πln |T(x)T(y)||T(y)|
2π.
In the following, we will have to use both characterizations of the map T, as a CCmap,
and as a R2R2map. In particular, we recall that T0(x) = 1T(x) = 1T1(x) + i∂1T2(x),
and that 1T1=2T2and 2T1=1T2. So for any map fC1(R2,R), we have that
(fT)(x) = 1T1(x)1f(T(x)) + 1T2(x)2f(T(x))
1T2(x)1f(T(x)) + 1T1(x)2f(T(x)),
6
or as complex numbers:
1(fT)(x) + i∂2(fT)(x) = Re(T0(x))1f(T(x)) + Im(T0(x))2f(T(x))
+i[Im(T0(x))1f(T(x)) + Re(T0(x))2f(T(x))]
=T0(x)(1f(T(x)) + i∂2f(T(x))).
Identifying f=1f+i∂2f, we have:
(fT)(x) = T0(x)f(T(x)),(2.4)
where the product above must be understood as the product of two complex numbers. We
will frequently use this property in the following.
From Proposition 2.1 and from relation (2.1) applied for and for Dwe get that
γ(x, y) = γD(T(x), T (y)) + 1
2πln |T(x)T(y)|
|xy|,
and letting yxwe obtain at the level of Robin functions
x,eγ(x) = eγD(T(x)) + 1
2πln |T0(x)|.(2.5)
Using the explicit expression of γD, see relation (2.3), we can compute the gradient of γ:
xγ(x, y) = T0(x)(T(x)T(y))
2π|T(x)T(y)|2T0(x)(T(x)T(y))
2π|T(x)T(y)|2(xy)
2π|xy|2
=T0(x)
2π(T(x)T(y)) T0(x)
2π(T(x)T(y))1
2π(xy)
(2.6)
where we used relation (2.4).
The mapping eγ: Ω R2, that will be named only eγwhen there is no ambiguity, has
been studied extensively in [5]. In particular, it was shown in that paper that eγis a super-
harmonic function, that is that eγ>0, and that it goes to infinity near the boundary like
1
2πln(d(x, ∂Ω)). This implies that we have the following proposition (see [5]).
Proposition 2.2 ([5]).For every bounded and simply connected open set , there exists at
least one point x0where eγreaches its minimum.
Critical points of eγwill be of special interest in the following. The proposition above
implies the existence of a critical point of the Robin function. Moreover, if is convex, then
the critical point is also unique (see [5]). Though we will not use this result here, it may be
interesting to keep this in mind, especially when looking for explicit examples.
2.2 Point vortex system
The point vortex dynamic, which is described by equations (1.4), can exhibit finite time
blow-up of solutions. One scenario of blow-up is when two point vortices hit each other
in finite time, meaning that there exists i6=jand t < such that zi(t) = zj(t). This
7
phenomena can happen, see for instance [15] or [10] for an example of finite-time collapse of
a self similar evolution of point vortices. Another scenario for blow-up is when a point vortex
hits the boundary. However, the finite time blow-up is exceptional in the case of the whole
plane, see [15], and for the unit disk, see [14]. In those cases, the N-dimensional Lebesgue
measure of the set of initial positions that lead to a collapse is 0. The case of a more general
bounded domain is an ongoing work.
Let us recall the convergence theorem obtained in [16], for configurations of point vortices
that do not lead to finite-time blow-up:
Theorem 2.3 ([16]).Let (ai, zi(t)) be a global solution of the point vortex system (1.4). For
every δ > 0and for every time τ, there exists ε > 0such that if ω0satisfies (1.5) for some
0< ν < 8/3, then the vorticity stays confined up to the time τin disks of radius δcentered
on zi(t).
In addition, the authors of [3] proved that τε,β +as εgoes to 0 for any β < 1/3
(recall that τε,β was defined in relation (1.6)). However, these theorems don’t say anything
about how δdepends on the time τ, or conversely, for how long this confinement remains
true, depending on ε.
For some explicit examples of point vortices we refer to [9].
As mentioned in the introduction, the stationary point vortices are the critical points of
the Robin function. Indeed, the relation (1.4) with N= 1 reduces to
z0(t) = a
2eγ(z(t)).
We refer to such a critical point x0by saying that it is a stationary point vortex.
Those points can also be characterized in terms of the conformal mapping Tas zeroes of
T00.
Proposition 2.4. The following conditions are equivalent:
(i) A single point vortex placed in x0is stationary,
(ii) eγ(x0) = 0,
(iii) T00(x0) = 0.
Proof. This proposition was already proved in [5]. We recall the proof for the convenience of
the reader.
We already observed that (i)and (ii)are equivalent. We only need to prove that (ii)is
equivalent to (iii). From relations (2.5) and (2.4), and recalling that T0can’t vanish in ,
we deduce that
eγ(x) = eγD(T(x))T0(x) + 1
2π
T0(x)T00(x)
|T0(x)|2.
One can easily check that 0is a stationary point for the unit disk, so eγD(0) = 0. Thus
eγ(x0) = 1
2πT00(x0)
T0(x0).
Clearly eγ(x0) = 0 is equivalent to T00(x0) = 0.
8
3 Confinement for Dirac mass around a stationary vortex
The aim of this section is to study the case where the vorticity ωitself is a Dirac mass. The
reason we consider this simpler case is because it is easier to have a complete description of
what happens. And this in turn gives us an indication of what to expect in the smooth case
considered in Theorem 1.3.
We consider a single point vortex located at z(t)which is close to the stationary point
vortex x0. Rescaling time if needed, we can assume without loss of generality that the mass
of the point-vortex is 1. The question of knowing if z(t)remains close to x0is closely related
to the notion of stability of x0that we discuss in what follows.
3.1 Stability of a stationary point vortex
Since the mass of the point vortex z(t)is 1, its equation of motion is the following
z0(t) = 1
2eγ(z(t)).(3.1)
We have that d
dteγ(z(t)) = z0(t)· ∇eγ(z(t)) = 0
which means that the point vortex is evolving on the level set eγ(x) = cst. We know from [5]
that the Robin function eγ(x)goes to infinity as xapproaches the boundary of . Therefore,
the point vortex z(t)can never reach the boundary so (3.1) has a global solution.
Since eγis super-harmonic, the eigenvalues of the real symmetric matrix D2eγ(x0)have
positive sum, meaning that at least one of them is positive. So two main cases can occur:
either both eigenvalues are positive, either one is positive and one is negative. We skip the
study of the degenerate case when one eigenvalue is 0.
So let λ+>0and λbe the two eigenvalues of D2eγ(x0). If λ>0, the Morse Lemma
implies that there exists a change of variables y=φ(x)in the neighborhood of x0such that
in this neighborhood of x0,
eγ(x) = eγ(φ1(y)) = eγ(x0)+(yx0)2
1+ (yx0)2
2.
In particular, in these new local coordinates, the level sets of eγare circles, so in the real
coordinates, they are diffeomorphic to circles. More precisely, the level set eγ(φ1(y)) =
eγ(x0) + ris a circle of radius r, provided that r > 0is small enough. Because φis a
Cdiffeomorphism, and φ(x0) = x0, there exist constants k, K > 0such that k|yx0|<
|xx0|< K|yx0|and thus the level set eγ(x) = eγ(x0) + ris contained in the annulus
{kr < |xx0|< Kr}.
We conclude from these observations that if |z(0) x0| ≤ εwith εsmall enough, then
we have that |z(t)x0| ≤ εK
kfor every time t0. This a stability property: if εis small
enough and if we assume that the point vortex starts at distance εof x0, then it remains at
distance of order εfor any time.
Assume now that λ<0. We have in this case that
eγ(x) = eγ(φ1(y)) = eγ(x0)(yx0)2
1+ (yx0)2
2,
9
so the level sets in a neighborhood of x0are in the local coordinates yhyperbolas, with the
special level set eγ(x) = eγ(x0)being a union of two line segments. In particular, we will see
later that one line segment is repulsive, meaning that when a point vortex evolves on that
segment, it moves away from the point x0. We call this situation unstable. In that case,
no matter how close the point vortex starts from the critical point x0, it goes away from x0
exponentially fast, as we will see in section 3.3.
Let us summarize what we observed above. If the eigenvalues of the matrix D2eγ(x0)
are both positive, x0is said to be stable, and a point vortex close enough to x0remains
indefinitely close to it. If one of the eigenvalues is negative, then x0is said to be unstable,
because there exists a point vortex evolving on a level set going away from x0. This notion
of stability is the same as the one introduced in [15], Chapter 3. Figure 1 shows examples of
the two situations.
Figure 1: Trajectories of a single point vortex in a domain with two stable points and an unstable
one.
3.2 Taylor expansion near a stationary vortex
The aim is now to express the stability of a stationary point vortex in terms of the conformal
map T. We already know from Proposition 2.4 that x0is stationary if and only if T00(x0)=0.
We will now make a Taylor expansion of Taround x0in order to have a better understanding
of the point vortex dynamic near a critical point. We therefore need to compute the Hessian
matrix D2eγin that point. By relation (2.5) we have that
1
2D2eγ(x0) = 1
2D2(eγDT)(x0) + 1
4πD2(ln |T0|)(x0).
This expression is actually explicit in terms of the conformal map T; indeed, we know
that eγD(x) = 1
2πln(1 − |x|2). We recall that T00(x0)=0, so 1T0(x0) = 2T0(x0)=0. We
recall also that for an holomorphic function ϕwe have that 1ϕ=ϕ0and 2ϕ=0. Some
direct computations give that
D2(eγDT)(x0) = 1
π|T0(x0)|20
0|T0(x0)|2
10
and
ij(ln |T0|)(x0) = ijT0(x0)·T0(x0)
|T0(x0)|2
so that
D2(ln |T0|)(x0) = 1
|T0(x0)|2T000(x0)·T0(x0) (T000 (x0))·T0(x0)
(T000(x0))·T0(x0)T000(x0)·T0(x0).
We conclude that 1
2D2eγ(x0) = 1
4π2µ2+p q
q2µ2p,(3.2)
where
µ2=|T0(x0)|2
p=1
µ2T000(x0)·T0(x0)
q=1
µ2(T000(x0))·T0(x0).
We compute the characteristic polynomial of the matrix 1
2D2eγ(x0):
det(1
2D2eγ(x0)X I2) = X2µ2
πX+1
16π24µ4p2q2
and the eigenvalues are
λ±=2µ2±pp2+q2
4π.
Furthermore, we notice that p2+q2=|T000(x0)|2
|T0(x0)|2. We deduce from the expression of the
determinant that x0is stable when λ>0, that is when 2|T0(x0)|3>|T000(x0)|and unstable
when 2|T0(x0)|3<|T000(x0)|.
Another interesting thing to notice is that when T000(x0) = 0, the matrix D2eγ(x0) =
µ2
2πI2, with I2the identity matrix, and thus the trajectories of the linearized system z0(t) =
(D2eγ(x0)(z(t)x0))are circles of center x0.
3.3 Exit time around an unstable stationary point
We now consider the unstable case, meaning λ+λ<0. We now go back to (3.1). Using
relation (3.2) and the subsequent calculations, one can check that the eigenvalues of the
Jacobian matrix of 1
2eγin x0are ξ=pλ+λand ξ.
We thus know from [7] Theorem 6.1 of Chapter 9, and the corollary and remarks associated
to this theorem that there exist two local invariant manifolds for the equation, and thus there
exist two special solutions to equation (3.1) associated with each eigenvalue ±ξ. We denote
by z1the solution that goes away from x0, corresponding to the eigenvalue ξ. It satisfies that
0< c < ξ, lim
t→−∞ |z1(t)x0|ect = 0.(3.3)
11
We denote by t1the first time when |z1(t1)x0|=εβand by t2the last time when
|z1(t2)x0|=ε. Provided εis small enough, we have that 0> t1> t2. We define
˜z(t) = z1(t+t2). This is a solution of equation (3.1) such that |˜z(0) x0|=εand thus we
can set
τε,β = max t0,s[0, t],|˜z(t)x0| ≤ εβ
which is the exit time as defined in relation 1.6 associated to a point vortex starting at
distance εof x0and evolving along the same trajectory as z1. By construction, it satisfies
that τε,β =t1t2≤ −t2.
We also know from relation (3.3) that there exist a constant Mand a constant 0< c < ξ,
such that for every t < 0, we have that |z1(t)x0|ect M. Therefore,
ε=|z1(t2)x0| ≤ Mect2.
Thus, ln εln M+ct2which means that τε,β ≤ −t22
c|ln ε|for εsmall enough.
This result empowers the conjecture that one cannot expect any general confinement
result, as in Theorem 1.1, better than C|ln ε|, since we have an example of such an exit time
in the case where the vorticity is itself a point vortex.
We can summarize the results of this section in the following theorem:
Theorem 3.1. Assume that T(x0) = T00 (x0) = 0, so that x0is a stationary point. If
|T000(x0)|<2|T0(x0)|3then any single point vortex starting close to x0will remain indefinitely
close to it. More precisely, there exists some small constant ε0>0such that if εε0and
z(0) D(x0, ε)then z(t)D(x0, Cε)for all times tand τε,β = +.
Conversely, the condition |T000(x0)|>2|T0(x0)|3insures that for every ε > 0, there exists
a starting position z(0) such that |z(0) x0| ≤ ε, but the point vortex exits the disk D(x0, εβ)
with β < 1in finite time τε,β C|ln ε|.
4 Power-law confinement near a stationary vortex
The present section is devoted to the proof of Theorem 1.3. To simplify the notation, we
denote from now on γ=γ,eγ=eγand G=G.
Recall that is a simply connected and bounded domain, that Tis a biholomorphism
from to Dsuch that T(x0) = 0. We have that x0is a stationary point which, according to
Proposition 2.4, means that T00(x0)=0.
We assume that ω0satisfies (1.5) with N= 1 and z1=x0. Without loss of generality we
can assume that the mass a1is 1 (changing the mass means rescaling the time). We therefore
have that ω0is non negative, is supported in D(x0, ε)and kωkL1= 1.
Let us introduce
F(x, t) = Z
xγ(x, y)ω(y, t) dy, (4.1)
which is the influence of the blob over itself through the boundary. The key point of the
proof of Theorem 1.2 from [2] in the case of the disk is the following property:
x, z D(0, δ),t0,|F(x, t)F(z, t)| ≤ C|xz|δ2.(4.2)
12
To prove that the result remains true for more general domains, we will need a similar
inequality. It is important to notice that unlike the case when ωis a point vortex itself, we
don’t necessarily have that F(x0, t) = 0 a priori. According to (2.6) we have that
2πxγ(x0, y) = T0(x0)
T(y)+T0(x0)
T(y)1
x0y.
Recalling that T(y)= 1/T (y)and making the Taylor expansion
T(y) = T0(x0)(yx0) + T000 (x0)
6(yx0)3+O(|yx0|4)
we get that
2πxγ(x0, y) = 1
yx0
1
1 + T000 (x0)
6T0(x0)(yx0)2+O(|yx0|3)+T0(x0)T(y)1
x0y
=1
yx01T000(x0)
6T0(x0)(yx0)2+O(|yx0|3)
+T0(x0)T0(x0)(yx0) + O(|yx0|2)1
x0y.
We conclude that
2πxγ(x0, y) = T000 (x0)
6T0(x0)(yx0) + |T0(x0)|2(yx0) + O(|yx0|2).(4.3)
4.1 Estimate of the influence of the boundary
The first step of the proof is a technical lemma giving a Lipschitz inequality for F. This is
the counterpart in of the relation (4.2) valid for the unit disk.
Lemma 4.1. We have the following estimate, for δsufficiently small and for every x, y, z
D(x0, δ)
xγ(x, y)− ∇xγ(z, y) = (xz)T000(x0)
6πT 0(x0)+O(δ),(4.4)
where the term O(δ)is bounded by with Ca constant depending only on and x0. In
particular, if we assume that T000(x0) = 0, there exists a constant K1=K1(Ω, x0)such that
x, y, z D(x0, δ),|∇xγ(x, y)− ∇xγ(z, y)| ≤ K1|xz|δ. (4.5)
Furthermore, if we assume that supp ωD(x0, δ)we have that
x, z D(x0, δ),|F(x, t)F(z, t)| ≤ K1|xz|δ. (4.6)
Proof. Let us define
R(x, y, z) = 2π(xγ(x, y)− ∇xγ(z, y)).
13
Using (2.6) we can write
R(x, y, z) = T0(x)
T(x)T(y)T0(x)
T(x)T(y)1
xyT0(z)
T(z)T(y)T0(z)
T(z)T(y)1
zy
so
R(x, y, z) = T0(x)(T(z)T(y)) T0(z)(T(x)T(y))
(T(x)T(y))(T(z)T(y)) +xz
(xy)(zy)
T0(x)(T(z)T(y))T0(z)(T(x)T(y))
(T(x)T(y))(T(z)T(y)).
We decompose Rinto two parts:
R=R1R2
with
R1(x, y, z) = T0(x)(T(z)T(y)) T0(z)(T(x)T(y))
(T(x)T(y))(T(z)T(y)) +xz
(xy)(zy)
and
R2(x, y, z) = T0(x)(T(z)T(y))T0(z)(T(x)T(y))
(T(x)T(y))(T(z)T(y))
=T0(x)T(z)T0(z)T(x)
(T(x)T(y))(T(z)T(y))+T(y)(T0(z)T0(x))
(T(x)T(y))(T(z)T(y))
R21 +R22.
Since Tis smooth and T(x0) = 0, we clearly have that |T(x)| ≤ and |T(y)| ≤ Cδ. So
we can estimate
|T(x)T(y)| ≥ |T(y)|−|T(x)|=1
|T(y)|− |T(x)| ≥ 1
Cδ 1
2Cδ
if δis sufficiently small. We can therefore bound R21 as follows:
|R21| ≤ 4C2δ2|T0(x)T(z)T0(z)T(x)| ≤ Cδ2|xz|.
Similarly
|T(x)T(y)| ≥ |T(y)|−|T(x)|=1
|T(y)|− |T(x)| ≥ 1
2|T(y)|
so |T(y)|
|T(x)T(y)|2|T(y)||T(y)|= 2.
Then
|R22|=|T(y)|
|T(x)T(y)||T0(z)T0(x)|
|T(z)T(y)|4Cδ|T0(z)T0(x)| ≤ Cδ|xz|2C δ2|xz|.
14
We conclude from the estimates above that
|R2(x, y, z)| ≤ Cδ2|xz|.(4.7)
To estimate R1, we write it under the form
R1(x, y, z) = N(x, y, z)
(T(x)T(y))(T(z)T(y))(xy)(zy)
with
N(x, y, z)=[T0(x)(T(z)T(y)) T0(z)(T(x)T(y))](xy)(zy)
+ (xz)(T(x)T(y))(T(z)T(y)).(4.8)
As Tis holomorphic we observe that Nis holomorphic in the variables x,yet z. One
can notice that Nis 0 if x=z, so it can be factorized by xz. Let us also recall that γis
smooth everywhere, so Ris smooth too. We proved above that R2is bounded in D(x0, δ)3,
so R1=R+R2is also bounded in this set. But the denominator of R1has a factor
(xy)2(yz)2, so it follows that N(x, y, z)can be factorized by (xy)2(yz)2. But we
observed that Ncan also be factorized by xz. This implies that there exists a holomorphic
function N1(x, y, z)such that:
N(x, y, z) = (xy)2(zy)2(xz)N1(x, y, z).(4.9)
Therefore R1(x, y, z)
xz=(xy)(zy)N1(x, y, z)
(T(x)T(y))(T(z)T(y)).(4.10)
We need now to compute N1(x0, x0, x0). To do that, we will differentiate five times
relation (4.9) and evaluate it in (x0, x0, x0). It is clear that derivatives up to order 4 of
(xy)2(zy)2(xz)all vanish at (x0, x0, x0). We can therefore write that
3
x2
zN(x0, x0, x0) = 3
x2
z(xy)2(zy)2(xz)(x0, x0, x0)N1(x0, x0, x0) = 12N1(x0, x0, x0).
Differentiating relation (4.8) allows to find after some computations that
3
x2
zN(x0, x0, x0) = 4T0(x0)T000(x0).
Therefore,
N1(x0, x0, x0) = T0(x0)T000(x0)
3
so
N1(x, y, z) = N1(x0, x0, x0) + O(δ) = T0(x0)T000(x0)
3+O(δ).
We now go back to (4.10). We observe that
xy
T(x)T(y)
15
is smooth on D(x0, δ)3with value 1/T 0(x)when x=y. Therefore
xy
T(x)T(y)=1
T0(x0)+O(δ).
Similarly zy
T(z)T(y)=1
T0(x0)+O(δ).
Combining the previous relations results in
R1(x, y, z)
xz=1
T0(x0)+O(δ)2T0(x0)T000(x0)
3+O(δ)=T000(x0)
3T0(x0)+O(δ)
so
R1(x, y, z)=(xz)T000(x0)
3T0(x0)+O(δ).
Recalling (4.7) finally implies that
R(x, y, z) = (xz)T000(x0)
3T0(x0)+O(δ),
which proves (4.4). Clearly (4.5) follows from (4.4) if T000(x0) = 0. Finally, relation (4.6)
follows from (4.5) after integrating and recalling that the mass of ωis 1.
Comparing (4.6) and (4.2), we see that we lose the factor δ2. But in the case T000 (x0)=0,
we still get a factor δand this is enough to make our argument work. Actually the proof of
[2] would still be correct assuming only that |F(x, t)F(z, t)| ≤ C|xz|δ. In our proof, a
factor δ2would improve the restriction over the power αin Theorem 1.3. However, please
notice that if T000(x0)6= 0 we lose the factor δand our proof does not work anymore.
4.2 Estimates of the trajectories
From now on we assume that T000(x0) = 0.
Let us introduce the center of vorticity:
B(t) = Z
(x, t) dx, (4.11)
and the moment of inertia:
I(t) = Z|xB|2ω(x, t) dx.
For future needs, let us compute the time derivative of B. Recall that ωsatisfies the
16
equation (1.2) in the sense of distributions and that it is compactly supported. We have that
d
dtB(t) = Zx∂tω(x, t) dx
=Zxu(x, t)· ∇ω(x, t) dx
=Z(u(x, t)· ∇)x ω(x, t) dx
=Zu(x, t)ω(x, t) dx
=ZZ (xy)
2π|xy|2+
xγ(x, y)ω(y, t)ω(x, t) dxdy
where we used (1.3) and (2.1).
Observing that (xy)
2π|xy|2is antisymmetric when exchanging xand yand recalling the
definition of Fgiven in (4.1), we infer that
d
dtB(t) = ZF(x, t)ω(x, t) dx. (4.12)
Let us define
Rt= max{|xB(t)|;xsupp ω(t)}
and choose some X(t)supp ω(t)such that |X(t)B(t)|=Rt. We denote by s7→ Xt(s)
the trajectory passing through X(t)at time tso that Xt(t) = X(t).
We have the following lemma which allows us to control the time evolution of Rt:
Lemma 4.2. For any tτε,β we have that
d
ds|Xt(s)B(s)|s=t2K1εβRt+5
πR3
t
I(t) + K2ενZ|xB|>Rt/2
ω(x, t) dx1/2
where νis the constant from relation (1.5),K1is the constant from Lemma 4.1 and K2is a
universal constant.
This lemma shows that in order to obtain upper bounds for the growth of the support
of ω, one needs estimates for I(t),B(t)and for the mass of vorticity far from the center of
mass.
Proof. We have that for any s0and t0,X0
t(s) = u(Xt(s), s), so
d
ds|Xt(s)B(s)|= (u(Xt(s), s)B0(s)) ·Xt(s)B(s)
|Xt(s)B(s)|.
We fix now the time t0, we take s=t, and we write Xinstead of Xt(t). By the Biot-Savart
law (1.3), the relation (4.1) and recalling that the vorticity is assumed to be of integral 1, we
have that
u(X, t) = F(X, t) + Z(Xy)
2π|Xy|2ω(y, t) dy=ZF(X, t) + (Xy)
2π|Xy|2ω(y, t) dy.
17
Relation (4.12) now implies that
d
ds|Xt(s)B(s)|s=t=ZF(X, t) + (Xy)
2π|Xy|2F(y, t)ω(y, t) dy·XB(t)
|XB(t)|
=H1+H2
where
H1=Z(F(X, t)F(y, t)) ω(y, t) dy·XB(t)
|XB(t)|,
and
H2=Z(Xy)
2π|Xy|2ω(y, t) dy·XB(t)
|XB(t)|,
Thanks to Lemma 4.1 and recalling that tτε,β, we have that
|H1|=Z(F(X, t)F(y, t)) ω(y, t) dyK1εβZ|Xy|ω(y, t) dy
K1εβ2Rt
where we used that supp ωD(B, R). The second term H2is the same as the left hand side
of relation (2.28) in [2], and its estimate is the same:
|H2| ≤ 5
πR3
t
I(t) + 1
πενZ|xB|>Rt/2
ω(x, t) dx1/2
The lemma is now proved.
4.3 Estimates of the moments of the vorticity
We have the following lemma:
Lemma 4.3. For every t < min(τε,β, εβ), we have:
I(t)K3ε2.
and:
|B(t)x0| ≤ K3ε,
where K3is a positive constant depending only on and x0.
Proof. We differentiate I(t):
I0(t) = Z|xB|2tω(x, t)2B0(t)·(xB)ω(x, t)dx=Z|xB|2tω(x, t) dx
18
where we used relation (4.11). Next, we use the equation of ωgiven in (1.2) and write
I0(t) = Z−|xB|2u(x, t)· ∇ω(x, t)dx
=Z2(xB)·u(x, t)ω(x, t) dx
=ZZ 2(xB)· ∇
xG(x, y)ω(x, t)ω(y, t) dxdy
=ZZ 2(xB)·
xγ(x, y) + (xy)
2π|xy|2ω(x, t)ω(y, t) dxdy
=ZZ 2(xB)· ∇
xγ(x, y)ω(x, t)ω(y, t) dxdy
+ZZ x·yB·(xy)
π|xy|2ω(x, t)ω(y, t) dxdy
Exchanging xand yshows that the last term above vanishes. So
I0(t) = ZZ 2(xB)· ∇
xγ(x, y)ω(x, t)ω(y, t) dxdy
=ZZ 2(xB)·[
xγ(x, y)− ∇
xγ(B, y)]ω(x, t)ω(y, t) dxdy
2K1εβZZ |xB|2ω(x, t)ω(y, t) dxdy
= 2K1εβI(t)
where we used (4.5) with δ=εβbecause (x, y, z)B(x0, εβ)(see the definition (1.6) of τε,β).
The Gronwall lemma now implies that I(t)I(0) exp(2K1εβt). Since at the initial time we
have that supp ω0and B(0) are in the disk D(x0, ε)we infer that I(0) 4ε2. We assumed
that t<εβso we finally obtain that
tmin(τε,β, εβ), I(t)4e2K1ε2.
We estimate now the center of vorticity. By relations (4.12) and (4.1), we have that
d
dt|B(t)x0|2= 2B0(t)·(B(t)x0)
= 2 ZZ
xγ(x, y)ω(y, t)ω(x, t) dxdy·(B(t)x0).
To estimate this, we use Lemma 4.1 and relation (4.3) and we recall that we assume that
T000(x0) = 0 and supp ωD(x0, εβ)to obtain
2π
xγ(x, y) = 2π(
xγ(x, y)− ∇
xγ(x0, y) +
xγ(x0, y))
= 2π(
xγ(x, y)− ∇
xγ(x0, y)) + |T0(x0)|2(yx0)+O|yx0|2
=|T0(x0)|2(yx0)+εβ(|xx0|+|yx0|)O(1).
19
Putting the estimates above together we have that
d
dt|B(t)x0|2= 2 ZZ |T0|2(x0)(yx0)
2π+εβ(|xx0|+|yx0|)O(1)
ω(y, t)ω(x, t) dxdy·(B(t)x0).
But we have the following cancellation:
ZZ |T0|2(x0)(yx0)ω(y, t)ω(x, t) dxdy·(B(t)x0)
=|T0|2(x0)ZZ (yx0)ω(y, t)ω(x, t) dxdy·(B(t)x0)
=|T0|2(x0)(B(t)x0)·(B(t)x0)
= 0.
Therefore, d
dt|B(t)x0|2C|B(t)x0|εβZ|xx0|ω(x, t) dx.
We notice now that
Z|xx0|2ω(x, t) dx=Z|xB(t)|2ω(x, t) dx
Z(x0B(t)) ·(xx0+xB(t))ω(x, t) dx
=Z|xB(t)|2ω(x, t) dx(x0B(t)) ·(B(t)x0+B(t)B(t))
=I(t) + |x0B(t)|2.
By the Cauchy-Schwarz inequality
Z|xx0|ω(x, t) dxZ|xx0|2ω(x, t) dx1
2= (I+|B(t)x0|2)1
2,
so d
dt|B(t)x0| ≤ Cεβ(I1
2+|B(t)x0|).
By the Gronwall lemma we infer that
|B(t)x0| ≤ exp Cεβt|B(0) x0|+CεβZt
0
I(s)1
2ds.
We already know that I(s)2, and |B(0) x0| ≤ ε, so
|B(t)x0| ≤ exp(Cεβt)(ε+Cεβ).
As we assumed that εβt1we conclude that
|B(t)x0| ≤ Cε.
This completes the proof of the lemma.
20
We have just shown that up to the time min(τε,β, εβ)the center of mass Bstays close
to x0and the moment of inertia Iremains small. We need a last technical lemma, which is
inspired by the appendix of [11].
Lemma 4.4. For every k1and tmin(τε,β, εβ)there exists a small constant ε0=ε0(k),
a large constant C(k)and a constant K4which depends only on and x0, such that if
εε0and r4K4ε2(1 + kt ln(2 + t)) ,
then Z|xB|>r
ω(x, t) dxC(k)εk/2
rk.
Proof. Let us introduce the moment of vorticity of order 4n
mn(t) = Z|xB(t)|4nω(x, t) dx.
One can differentiate to obtain, by using relations (1.3), (4.12) and recalling that ∇ · u= 0
m0
n(t) = Z|xB(t)|4nu(x, t)· ∇ω(x, t) dx
4nZB0(t)·(xB(t))|xB(t)|4n2ω(x, t) dx
= 4nZu(x, t)·(xB(t))|xB(t)|4n2ω(x, t) dx
4nB0(t)·Z(xB(t))|xB(t)|4n2ω(x, t) dx
= 4nZZ
xG(x, y)·(xB(t))|xB(t)|4n2ω(x, t)ω(y, t) dxdy
4nZZ
xγ(z, y)ω(z, t)ω(y, t) dzdy·Z(xB(t))|xB(t)|4n2ω(x, t) dx.
Recalling that G(x, y) = ln |xy|
2π+γ(x, y)and that eω(x0, t) = ω(x0+B(t), t)), we can further
decompose
m0
n(t) = an(t) + bn(t)cn(t)
where
an(t)=4nZZ (x0y0)
2π|x0y0|2·x0|x0|4n2eω(x0, t)eω(y0, t) dy0dx0,
bn(t) = 4nZZ
xγ(x, y)·(xB(t))|xB(t)|4n2ω(x, t)ω(y, t) dxdy,
cn(t)=4nZZZ
xγ(z, y)·(xB(t))|xB(t)|4n2ω(x, t)ω(y, t)ω(z, t) dxdydz.
We observe that anis exactly the same quantity that appears in [11] on the last line of
page 1726. Observing that the center of mass of eω(x, t)is in 0, we deduce that the estimates
21
given in [11] are true for an. More precisely, ansatisfies the estimate given on the line 5, page
1729 of [11]:
|an(t)| ≤ Cn2I(t)mn1(t).
Applying Lemma 4.3 we obtain that
|an(t)| ≤ Cn2ε2mn1(t).(4.13)
The terms bnand cnare similar. We decompose
xγ(x, y) =
xγ(x, y)− ∇
xγ(x0, y) +
xγ(x0, y).
We apply Lemma 4.1 with δ=εβto deduce that
|∇
xγ(x, y)− ∇
xγ(x0, y)| ≤ K1|xx0|εβK1ε2β
for all x, y supp ω(t). Moreover, we also have that |xB(t)| ≤ βat least for εsmall
enough, since |xx0| ≤ εβand |B(t)x0| ≤ K3ε. Using these estimates we infer that
|bn(t)| ≤ C2βZ|xB(t)|4n1ω(x, t) dx+dnCnε5βmn1(t) + dn
where
dn= 4nZZ
xγ(x0, y)·(xB(t))|xB(t)|4n2ω(x, t)ω(y, t) dxdy.
Decomposing
xγ(z, y) =
xγ(z, y)
xγ(x0, y) +
xγ(x0, y)we obtain that the same
estimate holds true for cn:
|cn(t)| ≤ C5βmn1(t) + dn.
We estimate now dn. From relation (4.3) we know that there exists a bounded function
c(y)such that
xγ(x0, y)=[a(yx0) + c(y)|yx0|2]where a=|T0(x0)|2
2π. We thus have
that
dn= 4nZZ[a(yx0) + c(y)|yx0|2]·(xB(t))|xB(t)|4n2ω(x, t)ω(y, t) dxdy
4na Z(B(t)x0)·(xB(t))|xB(t)|4n2ω(x, t) dx
+Cn ZZ |yx0|2|xB(t)|4n1ω(x, t)ω(y, t) dxdy
4na Z(B(t)x0)·(xB(t))|xB(t)|4n2ω(x, t)) dx+C5βmn1(t)
Cn(|B(t)x0|ε3β+ε5β)mn1(t).
Let us recall that Lemma 4.3 gives that |B(t)x0| ≤ K3εand therefore
|bn(t)|+|cn(t)| ≤ C5βmn1(t)+2dnCn(ε1+3β+ε5β)mn1(t).
Together with relation (4.13) and recalling that m0
n(t) = an(t)+bn(t)cn(t), the last estimate
yields that
|m0
n(t)| ≤ Cn2(ε2+ε1+3β+ε5β)mn1(t).
22
Let us observe now that it suffices to assume that β > 2/5. Indeed, assume that Theorem
1.3 is proved for any β]2/5,1/2[. Let β02/5and α < min(β0,24β0). Since α < 2/5,
one can easily check that there exists β]2/5,1/2[ such that α < min(β, 24β)(one can
choose βclose to 2/5). Since β0< β, we also have that τε,β τε,β0so τε,β0> εα.
We assume in the sequel that β > 2/5. Due to this additional assumption, we have that
|m0
n(t)| ≤ Cn2ε2mn1(t).(4.14)
Using Hölder’s inequality on f(x) = ω1/n(x)and g(x) = |xB(t)|4n4ω11/n(x)with p=n
and q=n
n1, we have that
mn1(t) = Z|xB(t)|4n4ω(x, t) dx
Zω(x, t) dx1/n Z|xB(t)|4nω(t, x) dx(n1)/n
=mn(t)(n1)/n.
This last inequality combined with relation (4.14) gives that
m0
n(t)Cn2ε2mn(t)(n1)/n.
We integrate to obtain that
mn(t)mn(0)1/n +C2tn.
Clearly, mn(0) (2ε)4nso, assuming that ε1,
mn(t)2(1 + nt)n
for some constant Lwhich depends only on and x0. Let us choose any k1,rand nsuch
that
r4221 + kln(2 + t)
ln 2 t
and
kln(2 + t)
ln 2 1< n kln(2 + t)
ln 2 .
This defines n1since k1and ln(2+t)
ln 2 1. It also implies that 2n+1 >(2 + t)k. Thus we
have that
Z|xB|>r
ω(x, t) dx=Z|xB|>r
ω(x, t)|xB(t)|4n
|xB(t)|4ndx
mn(t)
r4n
(2(1 + nt))n
rkr4nk
23
1
rk2(1 + kln(2+t)
ln 2 t)n
22(1 + kln(2+t)
ln 2 t)nk/4
=1
rk2(1 + kln(2+t)
ln 2 t)k/4
2nk/4
=2k/4+1
rkεk/2L(1 + kln(2+t)
ln 2 t)k/4
2n+1
2k/4+1
rkεk/2L(1 + kln(2+t)
ln 2 t)k/4
(2 + t)k.
The function t7→ L(1 + kln(2+t)
ln 2 t)k/4
(2 + t)kis bounded on R+for every k, so there exists a
constant C(k)such that for εsmall enough:
Z|xB|>r
ω(x, t) dxC(k)εk/2
rk.
This completes the proof of the lemma.
4.4 End of the proof of Theorem 1.3
We can now finish the proof of Theorem 1.3.
We recall that, according to Lemma 4.2, we have that for each particle such that |X(t)
B(t)|=Rt,
d
ds|Xt(s)B(s)|s=t2K1εβRt+5
πR3
t
I(t) + K2ενZ|xB|>Rt/2
ω(x, t) dx1/2
.
Due to the estimates obtained in Lemma 4.3, taking t < min(τε,β, εβ), we infer that
d
ds|Xt(s)B(s)|s=t2K1εβRt+5K3ε2
πR3
t
+K2ενZ|xB|>Rt/2
ω(x, t) dx1/2
.(4.15)
Let us introduce fthe solution of the ODE:
f0(t)=4K1εβf(t) + 4 max 5K3ε2
πf 3(t), K2ενZ|xB|>f (t)/2
ω(x, t) dx1/2!
f(0) = 4ε.
(4.16)
We want to show that for every t[0,min(τε,β, εβ)],Rt< f (t). Assume that this assertion
is false, and let t2be the first time when it breaks down. Since f(0) = 4εand R02ε, we
24
infer that t2>0. Let s7→ Xt2(s)be a trajectory such that |Xt2(t2)B(t2)|=Rt2=f(t2).
From (4.15) and (4.16), we see that
d
ds|Xt2(s)B(s)|s=t2< f0(t2).(4.17)
However, we have that for every 0< h < t2,
|Xt2(t2h)B(t2h)| ≤ Rt2h< f(t2h)
which implies that
|Xt2(t2h)B(t2h)|−|Xt2(t2)B(t2)|
h>f(t2h)f(t2)
h
since |Xt2(t2)B(t2)|=Rt2=f(t2). Taking the limit as h0, we get a contradiction with
(4.17).
We choose now αand k > 6such that
0< α < min(β, 24β)
and
k(1/2β)+6β4ν > 0.
We define
t2= inf{t > 0, f(t) = εβ}
and
t1= sup{t<t2, f (t) = εβ/2}
so that t1< t2and
f(t1) = εβ
2, f(t2) = εβand εβ
2f(t)εβt[t1, t2].
If t2εαthen Rt< f(t)εβfor all t[0,min(τε,β , εβ, εα)] = [0,min(τε,β , εα)]. By
definition of τε,β we know that Rτε,β =εβ. So necessarily τε,β εαwhich completes the
proof of Theorem 1.3.
We assume from now on that t2< εα.
We have the following inequality:
t[t1, t2],f(t)
24
K4ε2(1 + kt ln(2 + t)) (4.18)
which implies that Lemma 4.4 can be applied with r=f(t)/2for t[t1, t2]. Indeed, relation
(4.18) is true since
K4ε2(1 + kt ln(2 + t)) K4ε2(1 + kεαln(2 + εα)) εβ
44
f(t)
24
25
for εsmall enough, as we chose α < 24β. Lemma 4.4 yields that
ενZ|xB|>f(t)/2
ω(x, t) dx1/2
C(k)εk/2ν
(f(t)/2)k1/2
t[t1, t2].(4.19)
Since we chose ksuch that k(1/2β)+6β4ν > 0, for εsmall enough we have that
ε(k6)βεk/24ν=εk(1/2β)+6β4ν1
C(k)22k6.
Recalling that f(t)εβ/2for every t[t1, t2], we infer that
fk6(t)εβ
2k6C(k)2kεk/24ν
which in turns gives that
C(k)εk/2ν
(f(t)/2)kε4
f6(t).(4.20)
Using relations (4.19) and (4.20) in (4.16) yields that
f0(t)Cεβf(t) + Cε2
f3(t).
which implies that
(f4)0(t)C1εβf4(t) + C1ε2
for some constant C1. The end of the argument is now straightforward. We use the Gronwall
lemma to obtain that
f4(t2)f4(t1)eC1εβ(t2t1)+ε2β(eC1εβ(t2t1)1).
Since t2t1εαwe have that C1εβ(t2t1)C1εβα<1for εsmall enough. Using the
inequality ex1 + 2xfor 0x1and recalling that f(t1) = εβ/2and f(t2) = εβwe have
that
ε4β(εβ/2)4eC1εβα+ 2C1ε2α.
which implies that
1eC1εβα
16 + 2C1ε24βα.
Since α < min(β, 24β), the right hand side of this inequality goes to 1/16 as ε0. So
we obtain a contradiction if εis small enough. We thus proved that t2εαif εis small
enough.
In conclusion, for any β < 1/2, and for any α < min(β, 24β), there exists ε0small
enough such that for every ε(0, ε0), we have that τε,β > εα. This completes the proof of
Theorem 1.3.
26
5 Final remarks
A natural question is the signification of the hypothesis T000(x0)=0. As we already discussed,
the condition T00(x0)=0means that x0is a critical point of the map eγ. Such critical
points always exist in a bounded simply connected domain . But to have a strong result
of confinement as we proved, we need more than just a critical point of eγ. Indeed, we
observed in Section 3 that we should not expect a confinement time better than |ln ε|around
unstable points. So we need the stationary point x0to be at least stable. But the hypothesis
T000(x0) = 0 is stronger than the stability. Indeed, the stability is characterized by the
condition |T000(x0)|<2|T0(x0)|3which is significantly weaker than T000(x0) = 0.
So the hypothesis T000(x0) = 0 is more than just stability. Because we obtained the
explicit value of D2eγ(x0), see relation (3.2), we see that this condition is equivalent to the
fact D2eγ(x0)is a multiple of the identity. This means that the orbit of a single point vortex
in the neighborhood of x0is almost a circle. We don’t know whether this condition is indeed
necessary to have a strong confinement as proved in Theorem 1.3. In our proof we use some
crucial cancellations to prove the estimate of the moment of inertia from Lemma 4.3 that we
can’t reproduce without the hypothesis T000(x0)=0.
One can wonder about the existence of domains satisfying the condition T000(x0) = 0 in a
stationary point x0. We call these domains valid domains. Let us notice that if T(x0) = 0 and
T00(x0) = 0, then the condition T000 (x0) = 0 is equivalent to (T1)000(0) = 0. This means that
the image Ω = f(D(0,1)) by any biholomorphic map of the form f(z) = x0+a1z+P
k=4 akzk
of the unit disk is a valid domain: there exists a biholomorphic map T: Ω D(0,1) such
that T(x0) = T00(x0) = T000 (x0)=0(one can choose T=f1). However, checking that a
map fof the form given above is indeed biholomorphic may not be an easy task.
Let us observe now that regular convex polygons are valid domains. Indeed, by the
Schwarz-Christoffel formula (see [1]), there exists a conformal map ffrom the unit disk to
the nsided regular polygon with vertices at ωn=ei2π/n, and its derivative has the form
f0(z) = c
n
Y
k=1 1k
n12
n.
We can therefore obtain an explicit value for the second and third derivatives, and check that
they vanish at 0. So regular convex polygons are valid domains.
On the other hand, an ellipse which is not a circle is not a valid domain. Indeed, we know
from [12] that a conformal mapping from the disk to an ellipse of foci ±1mapping 0 to 0 has
a Taylor expansion f(z) = z+A3z3+. . . near 0 with A3>0.
Another method to obtain valid domains is to analyze the effect of rotational invariance.
Assume that the domain is invariant by rotation of angle θ(0,2π)around the point
x0= 0. This means that eγ(x) = eγ(ex)for every x. Differentiating this relation and
using relation (2.4) implies that
eγ(e x) = cos θsin θ
sin θcos θeγ(x)
and thus eγ(0) = 0 implying that 0is a stationary point. Similarly, differentiating again
27
yields that
2
1eγ(x) = 2
1[eγ(ex)] = cos2θ∂2
1eγ(e x) + sin2θ∂2
2eγ(e x) + 2 cos θsin θ12eγ(e x)
2
2eγ(x) = 2
2[eγ(e x)] = sin2θ∂2
1eγ(e x) + cos2θ∂2
2eγ(e x)2 cos θsin θ∂12eγ(e x)
12eγ(x) = 12[eγ(e x)] = (cos2θsin2θ)12eγ(e x)
+ cos θsin θ(2
2eγ(e x)2
1eγ(e x)).
We set x= 0 and we subtract the first equation above from the second equation. We get
(1 cos2θ+ sin2θ)(2
2eγ(0) 2
1eγ(0)) = 4 cos θsin θ12eγ(0).
Using this in the third equation yields after some calculations that
θ=πor 12eγ(0) = 0.
If 12eγ(0) = 0 and θ6=πwe observe that 2
1eγ(0) = 2
2eγ(0). Therefore we have that
θ=πor λ, D2eγ(0) = λI2.
From (3.2) we know that
D2eγ(0) = µ2
πI2+1
2πp q
qp.
Clearly D2eγ(0) is a multiple of the identity if an only if p=q= 0. Therefore we have that
either θ=πor p=q= 0. From the definition of pand qgiven after relation (3.2) we see that
for any conformal map Tmapping 0 to 0 the condition p=q= 0 is equivalent to T000(0) = 0.
So if θ6=π, then T000(0) = 0 and the domain is therefore valid. This is another proof of the
fact that regular polygons are valid domains, and it also gives us many other valid domains.
We used Mathematica to plot the boundary of some valid domains. We consider different
functions f(z) = a1z+PN
k=4 akzk, with N4and |a1|>PN
k=4 k|ak|in order to ensure
that fis injective on the unit disk. We obtain a large class of valid domains, with the unit
disk of course and small variations of it (figure 2), but also larger perturbation of the disk
(figure 3), and even some very erratic domains (figure 4). Notice that these domains don’t
necessarily have symmetry properties. Finally, by using the rotational invariance property,
we can plot more complicated boundaries without knowing the biholomorphic map, like in
figure 5. We plotted images of the interval [0,1] by maps of the form b(x) = r(x)ei2π(x+θ(x)),
with r(x)>0. We choose θand rto be 1/p-periodic functions with p > 2an integer. This
way, bplots a closed curve in Cthat is invariant by rotation of angle 2π/p (0, π). If this
curve does not self intersect and is smooth, then its interior is a valid domain.
28
Figure 2: Plot for f(z) = z(left), and f(z) = 40z+z4(right).
Figure 3: Plot for f(z) = 40z+z7(left), and f(z) = 50z+ (i+ 1)z4+z23 (right).
Figure 4: Plot for f(z) = 20z+ (2i+ 1)z4+z7(left), and f(z) = 19z+iz7+z10 (right).
29
Figure 5: Plot for b(x) = (2 + cos(8πx))ei2πx (left), and b(x) = exp 2(x+1
8cos(6πx)
3
2+1
4cos(6πx)(right),
with x[0,1].
Acknowledgments. Dragos
,Iftimie has been partially funded by the LABEX MILYON
(ANR-10-LABX-0070) of Université de Lyon, within the program “Investissements d’Avenir”
(ANR-11-IDEX-0007) operated by the French National Research Agency (ANR). Martin
Donati thanks the Dipartimento di Matematica, SAPIENZA Università di Roma for its
hospitality. The authors wish to acknowledge helpful discussions with Paolo Buttà and
Carlo Marchioro.
References
[1] L. Ahlfors. Complex Analysis. McGraw Hill Publishing Co., New York., 1966.
[2] P. Buttà and C. Marchioro. Long time evolution of concentrated Euler flows with planar
symmetry. SIAM J. Math. Anal., 50(1):735–760, 2018.
[3] D. Cao and G. Wang. Euler evolution of a concentrated vortex in planar bounded
domains. arXiv:1801.01629 [math.AP], 2018.
[4] J. Goodman, T. Hou, and J. Lowengrub. Convergence of the point vortex method for
2-D Euler equations. Communications on Pure and Applied Mathematics, 43(3):415 –
430, 1990.
[5] B. Gustafsson. On the Motion of a Vortex in Two-dimensional Flow of an Ideal Fluid in
Simply and Multiply Connected Domains. Trita-MAT-1979-7. Royal Institute of Tech-
nology, 1979.
[6] Z. Han and A. Zlatos. Euler equations on general planar domains. arXiv:2002.05092
[math.AP], 2020.
30
[7] P. Hartman. Ordinary Differential Equations: Second Edition. Classics in Applied Math-
ematics. Society for Industrial and Applied Mathematics (SIAM, 3600 Market Street,
Floor 6, Philadelphia, PA 19104), 1982.
[8] H. Helmholtz. Über integrale der hydrodynamischen gleichungen, welche den wirbel-
bewegungen entsprechen. Journal für die reine und angewandte Mathematik, 55:25–55,
1858.
[9] D. Iftimie. Large time behavior in perfect incompressible flows. In Partial differential
equations and applications, volume 15 of Sémin. Congr., pages 119–179. Soc. Math.
France, Paris, 2007.
[10] D. Iftimie and C. Marchioro. Self-similar point vortices and confinement of vorticity.
Communications in Partial Differential Equations, 43(3):347–363, 2018.
[11] D. Iftimie, T. C. Sideris, and P. Gamblin. On the evolution of compactly supported
planar vorticity. Communications in Partial Differential Equations, 24(9-10):159–182,
1999.
[12] S. Kanas and T. Sugawa. On conformal representations of the interior of an ellipse.
Annales Academiae Scientiarum Fennicae. Mathematica, 31(2):329–348, 2006.
[13] C. Lacave and A. Zlatoš. The Euler Equations in planar domains with corners. Archive
for Rational Mechanics and Analysis, 234(1):57–79, 2019.
[14] C. Marchioro and M. Pulvirenti. Vortex methods in two-dimensional fluid dynamics.
Lecture notes in physics. Springer-Verlag, 1984.
[15] C. Marchioro and M. Pulvirenti. Mathematical Theory of Incompressible Nonviscous
Fluids. Applied Mathematical Sciences. Springer New York, 1993.
[16] C. Marchioro and M. Pulvirenti. Vortices and localization in euler flows. Comm. Math.
Phys., 154(1):49–61, 1993.
[17] W. Wolibner. Un theorème sur l’existence du mouvement plan d’un fluide parfait, ho-
mogène, incompressible, pendant un temps infiniment long. Mathematische Zeitschrift,
37(1):698–726, 1933.
[18] V. Yudovich. Non-stationary flow of an ideal incompressible liquid. USSR Computational
Mathematics and Mathematical Physics, 3(6):1407 – 1456, 1963.
Martin Donati: Université de Lyon, CNRS, Université Lyon 1, Institut Camille Jordan, 43
bd. du 11 novembre, Villeurbanne Cedex F-69622, France.
Email: donati@math.univ-lyon1.fr
Dragos
,Iftimie: Université de Lyon, CNRS, Université Lyon 1, Institut Camille Jordan, 43
bd. du 11 novembre, Villeurbanne Cedex F-69622, France.
Email: iftimie@math.univ-lyon1.fr
Web page: http://math.univ-lyon1.fr/˜iftimie
31
... In the present paper, we try to push further the analysis and prove that any solution to the Euler equation concentrated near this type of configuration remains concentrated for a very long time following the vortex crystal prediction. More precisely, if the support of the initial vorticity lies within disks of size ε, we prove that it remains within disks of size ε β , with β < 1/2, for a time at least of order ε −α for some α > 0. This extends previous results, see [3,7], where such long time confinement was obtained for configurations satisfying either N = 1 (only one vortex), or that point-vortices had to move away from each other very fast. Here, we prove that we need neither of those assumptions, but instead that the vortices are disposed according to the previously described polygonal configuration. ...
... There are other configurations found in [3] that satisfy this improved bound on the time of confinement when one only considers a single blob of vorticity (N = 1), in the plane, or at the center of a disk. In [7], it is proved that, for N = 1 in some special bounded and simply connected domains with a suitable choice of z 1 , the solution satisfies that τ ε,β ≥ ε −α for any α < min(β, 2 − 4β). ...
... Let us formulate a few remarks about Theorem 3.4 and its hypotheses. As mentioned in the introduction, the big difference with the similar long-time confinement results obtained in [3] and [7] is that we neither assume that N = 1, nor that all the distances between the point-vortex grow quickly. ...
Preprint
Full-text available
In this paper, we control the growth of the support of particular solutions to the Euler two-dimensional equations, whose vorticity is concentrated near special vortex crystals. These vortex crystals belong to the classical family of regular polygons with a central vortex, where we choose a particular intensity for the central vortex to have strong stability properties. A special case is the regular pentagon with no central vortex which also satisfies the stability properties required for the long-time confinement to work.
... The long time confinement problem consists of obtaining a lower-bound on τ ε,β in order to describe how long the approximation of a concentrated solution of equations (1) by the point-vortex model (α-PVS) remains valid. Results have been obtained in [2,7,5]. In the following, we recall some of them and state our main results, starting with the case α = 1. ...
... In that same article, the authors proved that when the initial vorticity is concentrated near the center of a disk, namely that Ω = D(0, 1), N = 1, and z 1 = 0, then we obtain the same power-law lower-bound τ ε,β ≥ ε −ξ0 than with expanding self-similar configurations. This result has been generalized to other bounded domains in [7]. This is due to a strong stability property induced by the shape of the boundary. ...
... Known results on critical points of the Robin's function. In this section we refer to [7] and recall the following results. For more details on the Robin's function we refer the reader to [15,11]. ...
... The long time confinement problem consists in obtaining a lower-bound on τ ε,β in order to describe how long the approximation of a concentrated solution of equations (1) by the point-vortex model (α-PVS) remains valid. Results have been obtained in [2,8,5]. In the following, we recall some of them and state our main results, starting with the case α = 1. ...
... In that same article, the authors proved that when the initial vorticity is concentrated near the center of a disk, namely that Ω = D(0, 1), N = 1 and z 1 = 0, then we obtain the same powerlaw lower-bound τ ε,β ≥ ε −ξ 0 than with expanding self-similar configurations. This result has been generalized to other bounded domains in [8]. This is due to a strong stability property induced by the shape of the boundary. ...
... In this section we refer to [8] and recall the following results. For more details on the Robin's function we refer the reader to [16,12]. ...
Preprint
Full-text available
In this paper we construct solutions to the Euler and gSQG equations that are concentrated near unstable stationary configurations of point-vortices. Those solutions are themselves unstable, in the sense that their localization radius grows from order \eps to order \eps^\beta (with β<1\beta < 1) in a time of order |\ln\eps|. This proves in particular that the logarithmic lower-bound obtained in previous papers (in particular [P. Buttà and C. Marchioro, \emph{Long time evolution of concentrated Euler flows with planar symmetry}, SIAM J. Math. Anal., 50(1):735–760, 2018]) about vorticity localization in Euler and gSQG equations is optimal. In addition we construct unstable solutions of the Euler equations in bounded domains concentrated around a single unstable stationary point. To achieve this we construct a domain whose Robin's function has a saddle point.
... In the literature, (P1) and (P2) are often referred to as desingularization of vortices. The study of (P1) was initiated by Marchioro and Pulvirenti [26], then followed by many authors [8,14,15,17,[25][26][27]33]. Roughly speaking, the answer to (P1) is positive, although some open problems still remain, such as the optimal concentration rate. ...
Preprint
In this paper, we construct a family of global solutions to the incompressible Euler equation on a standard 2-sphere. These solutions are odd-symmetric with respect to the equatorial plane and rotate with a constant angular speed around the polar axis. More importantly, these solutions ``converges" to a pair of point vortices with equal strength and opposite signs. The construction is achieved by maximizing the energy-impulse functional relative to a family of suitable rearrangement classes and analyzing the asymptotic behavior of the maximizers. Based on their variational characterization, we also prove the stability of these rotating solutions with respect to odd-symmetric perturbations.
... In (2.23) the condition a > |b| is required. This corresponds to the ellipticity of the critical point ξ 0 and it is necessary for the dynamics to be confined locally near ξ 0 , as explained in [31]. ...
Preprint
Full-text available
We examine the Euler equations within a simply-connected bounded domain. The dynamics of a single point vortex are governed by a Hamiltonian system, with most of its energy levels corresponding to time-periodic motion. We show that for the single point vortex, under certain non-degeneracy conditions, it is possible to desingularize most of these trajectories into time-periodic concentrated vortex patches. We provide concrete examples of these non-degeneracy conditions, which are satisfied by a broad class of domains, including convex ones. The proof uses Nash-Moser scheme and KAM techniques, in the spirit of the recent work of Hassainia-Hmidi-Masmoudi on the leapfrogging motion, combined with complex geometry tools. Additionally, we employ a vortex duplication mechanism to generate synchronized time-periodic motion of multiple vortices. This approach can be, for instance, applied to desingularize the motion of two symmetric dipoles (with four vortices) in a disc or a rectangle. To our knowledge, this is the first result showing the existence of non-rigid time-periodic motion for Euler equations in generic simply-connected bounded domain. This answers an open problem that has been pointed in the literature, for example by Bartsch-Sacchet.
... The point vortex model is only an approximate model, and rigorous analysis on its connection with the vorticity equation with concentrated vorticity is an interesting research topic. We refer the interested readers to [18,[30][31][32][33][34][35]39] for some deep results in this respect. ...
Article
Full-text available
In this paper, we study the stability of two-dimensional steady Euler flows with sharply concentrated vorticity in a bounded domain. These flows are obtained as maximizers of the kinetic energy on some isovortical surface, under the constraint that the vorticity is compactly supported in a finite number of disjoint regions of small diameter. We prove the nonlinear stability of these flows when the vorticity is sufficiently concentrated in one small region, or in two small regions with opposite signs. The proof is achieved by showing that these flows constitute a compact and isolated set of local maximizers of the kinetic energy on the isovortical surface. The separation property of the stream function plays a crucial role in validating isolatedness.
... For such initial vorticities, Buttà and Marchioro proved in [10] that the solution ω ǫ (x, t) remains concentrated near p(t) during a time scale t ≈ | log ǫ|. They also show that when M = 1, p 0 1 = 0 (a single vortex concentrated near origin) and if the space domain R 2 is replaced by B R (0), then the solution remains concentrated near origin during a time scale t ≈ ǫ −a for some a > 0. In [19], Donati and Iftimie generalized the result to simply connected bounded domains. ...
Preprint
We consider the two-dimensional incompressible Euler equation {tω+uω=0ω(0,x)=ω0(x).\begin{cases} \partial_t \omega + u\cdot \nabla \omega=0 \\ \omega(0,x)=\omega_0(x). \end{cases} We are interested in the cases when the initial vorticity has the form ω0=ω0,ϵ+ω0p,ϵ\omega_0=\omega_{0,\epsilon}+\omega_{0p,\epsilon}, where ω0,ϵ\omega_{0,\epsilon} is concentrated near M disjoint points pm0p_m^0 and ω0p,ϵ\omega_{0p,\epsilon} is a small perturbation term. First, we prove that for such initial vorticities, the solution ω(x,t)\omega(x,t) admits a decomposition ω(x,t)=ωϵ(x,t)+ωp,ϵ(x,t)\omega(x,t)=\omega_{\epsilon}(x,t)+\omega_{p,\epsilon}(x,t), where ωϵ(x,t)\omega_{\epsilon}(x,t) remains concentrated near M points pm(t)p_m(t) and ωp,ϵ(x,t)\omega_{p,\epsilon}(x,t) remains small for t[0,T]t \in [0,T]. Second, we give a quantitative description when the initial vorticity has the form ω0(x)=m=1Mγmϵ2η(xpm0ϵ)\omega_0(x)=\sum_{m=1}^M \frac{\gamma_m}{\epsilon^2}\eta(\frac{x-p_m^0}{\epsilon}), where we do not assume η\eta to have compact support. Finally, we prove that if pm(t)p_m(t) remains separated for all t[0,+)t\in[0,+\infty), then ω(x,t)\omega(x,t) remains concentrated near M points at least for tc0logAϵt \le c_0 |\log A_{\epsilon}|, where AϵA_{\epsilon} is small and converges to 0 as ϵ0\epsilon \to 0.
Article
Full-text available
The first part of this article studies the collapses of point-vortices for the Euler equation in the plane and for surface quasi-geostrophic equations in the general setting of α models. In these models the kernel of the Biot–Savart law is a power function of exponent − α . It is proved that, under a standard non-degeneracy hypothesis, the trajectories of the point-vortices have a Hölder regularity up to the time of collapse. The Hölder exponent obtained is 1 / ( α + 1 ) and this exponent is proved to be optimal for all α by exhibiting an example of a three-vortex collapse. The same question is then addressed for the Euler point-vortex system in smooth bounded connected domains. It is proved that if a given point-vortex has an accumulation point in the interior of the domain as t → T , then it converges towards this point and displays the same Hölder continuity property. A partial result for point-vortices that collapse with the boundary is also established: we prove that their distance to the boundary is Hölder regular.
Article
In this paper, we prove that in bounded planar domains with C^{2,\alpha} boundary, for almost every initial condition in the sense of the Lebesgue measure, the point-vortex system has a global solution, meaning that there is no collision between two point-vortices or with the boundary. This extends the work previously done in [C. Marchioro and M. Pulvirenti, Vortex Methods in Two-Dimensional Fluid Dynamics, Springer-Verlag, 1984] for the disk. The proof requires the construction of a regularized dynamics that approximates the real dynamics and some strong inequalities for the Green's function of the domain. The estab- lishment of some useful estimates is discussed and the details of the proof are given in the original article [M. Donati, SIAM J. Math. Anal., 54 (2022), pp. 79--113].
Preprint
We construct a family of steady solutions to the two-dimensional incompressible Euler equation in a general bounded domain, such that the vorticity is supported in two well-separated regions of small diameter and converges to a pair of point vortices with opposite signs. Compared with previous results, we do not need to assume the existence of an isolated local minimum point of the Kirchhoff-Routh function. Moreover, due to their variational nature, the solutions obtained are Lyapunov stable in LpL^p norm of the vorticity. The proofs are achieved by maximizing the kinetic energy over an appropriate family of rearrangement classes of sign-changing functions and studying the limiting behavior of the maximizers.
Article
Full-text available
When the velocity field is not a priori known to be globally almost Lipschitz, global uniqueness of solutions to the two-dimensional Euler equations has been established only in some special cases, and the solutions to which these results apply share the property that the diffuse part of the vorticity is constant near the points where the velocity is insufficiently regular. Assuming that the latter holds initially, the challenge is then to propagate this property along the Euler dynamic via an appropriate control of the Lagrangian trajectories. In domains with obtuse corners and sufficiently smooth elsewhere, Yudovich solutions fail to be almost Lipschitz only near these corners, and we investigate the necessary and sufficient conditions for the vorticity to remain constant there. We show that if the vorticity is initially constant near the whole boundary, then it remains so forever (and global weak solutions are unique), provided that no corner has angle greater than π\pi . We also show that this fails in general for domains that do have such corners.
Article
Full-text available
We study the time evolution of an incompressible Euler fluid with planar symmetry when the vorticity is initially concentrated in small disks. We discuss about how long this concentration persists, showing that in some cases this happens for quite long times. Moreover, we analyze a toy model that shows a similar behavior and gives some hints on the original problem.
Article
Full-text available
We consider the conformal mappings f and g of the unit disk onto the inside of an ellipse with foci at 1 so that f(0) = 0, f'(0) > 0, g(0) = -1 and g'(0) > 0. The main purpose of this article is to show positivity of the Taylor coefficients of f and g about the origin. To this end, we use a special relation between f and g and the fact that f satisfies a second-order linear ODE. Some applications are given to the class of k-uniformly convex functions.
Article
This papers deals with the large time behavior of solutions of the incompressible Euler equations in dimension 2. We consider a self-similar configuration of point vortices which grows like the square root of the time. We study the confinement properties of a blob of vorticity initially located around the first point vortex and moving in the velocity field produced by itself and by the other point vortices. We find a sufficient condition on the point vortices such that the vorticity stays confined around the first point vortex at a rate better than the square root of the time. The relevance to the large time behavior of the Euler equations is discussed.
Article
An overview is given on information based on lecture notes of courses on differential equations. Focus is on the theory of differential equations and its many applications to other disciplines. Exercises are roughly graded into three types according to difficulty. Based on some results, it is emphasized that the theory of differential equations depends heavily on the "integration of differential inequalities".
Research
The study of the motion of a point vortex in a planar domain, along with related studies in geometric function theory. (When writing this paper I was unaware of the related work, on the motion of several point vortices, by C.C. Lin from 1943.)