Abstract
We consider stability of solitons of 3D Maxwell–Lorentz system
with extended charged spinning particle.
The solitons are solutions which correspond to a particle
moving with a constant velocity with
and rotating with a constant angular velocity .
Our main results are
the orbital stability of moving solitons with
and a linear orbital stability of
rotating solitons with .
The Hamilton–Poisson structure of the Maxwell–Lorentz system
is degenerate and
admits the Casimir invariants.
We construct the Lyapunov function as a linear combination of the
Hamiltonian with a suitable Casimir invariant.
The key point is
a lower bound for this function.
The proof of the bound in the case
relies on angular momentum conservation
and
suitable spectral arguments including the Heinz inequality and closed graph theorem.
1 Introduction
The paper addresses
the stability of solitons of 3D Maxwell–Lorentz system
with an extended charged spinning particle.
We choose the units such that the speed
of light is .
The equations read as follows
(see [37, 40]):
|
|
|
(1.1) |
where
is the charge distribution of the extended
particle centered at the point ,
is the mass of the particle and is its moment of inertia,
is the angular velocity
of the particle rotation and is the velocity field.
The brackets here and below
denote the
integral over or inner product in the Hilbert spaces
with suitable .
This model
was introduced by Abraham in 1903–1905 (see [1, 2])
for the description of the classical extended electron coupled to its own Maxwell field.
The model allows one to avoid the
“ultraviolet divergence”, that is,
the infiniteness of the own energy and mass of electrons
which takes place in the case .
Using this model,
Abraham was the first to discover the mass-energy equivalence (up to the doubtful factor ; see [1, 2] and also
[23, Section 12.9]), which was
a precursor of
the
Einstein theory.
This system also served as the classical Landé model of spin in Old Quantum Mechanics (1900–1924): see [6] (and also
[23, Chapter 14]
and [24, Appendix A]).
Various approximations of such systems were used to explain
the famous radiation damping in classical electrodynamics: the Lorentz–Dirac equation
with runaway solutions introduced by Dirac in [9] and many other approximations; see
[20, Chapter 16].
The detailed account on the genesis and early investigations
of the system
by Dirac, Poincaré, Sommerfeld, and others
can be found in [40, Chapter 3].
The first results on the corresponding long-time behavior
for the system (1.1)
were obtained in the 1990s.
This system
plays a crucial role in a rigorous analysis of radiation by moving particles; see
[26, 40].
The mathematical analysis of this system
is useful in connection to the related problems of
nonrelativistic QED.
In particular, the similarity in the renormalization of mass
was pointed out
by Hiroshima and Spohn [12].
We assume that the charge density
is smooth, spherically-symmetric, and not identically zero:
|
|
|
(1.2) |
One of the basic peculiarities of
the system (1.1) is its invariance under
the Euclidean group generated by
translations, rotations and reflections of the space .
Moreover, the system
admits solitons
discovered by Spohn [40].
The solitons are solutions
moving with constant speed
and rotating with constant angular velocity .
The solitons are trajectories of
one-parametric subgroups of the
Euclidean
symmetry group.
As is pointed out in [40],
for ,
the solitons with finite energy exist only
for , where
|
|
|
(1.3) |
We justified
the restrictions (1.3) on the parameters of solitons
and calculated effective moment of inertia of solitons (6.9).
Our main results are i)
the orbital stability of solitons with
and ii) the linear orbital stability
of solitons with
in the class of solutions with zero total momentum.
In particular, the total momentum is zero for solutions
with “symmetric” initial data:
|
|
|
(1.4) |
Our approach essentially relies on the Hamilton–Poisson structure of the
system (1.1) which was constructed in [27].
The structure is degenerate, so the system admits the Casimir invariants,
and the methods of [11] are not applicable to the system.
The proof of the orbital stability of solitons relies on the reduction to the comoving frame and lower bound
for the Hamiltonian as the Lyapunov function.
The Hamilton functional of the reduced system depends on the parameter which is the conserved total momentum. Thus, we restrict the Hamiltonian to the states with the fixed value of .
To prove the stability of solitons , we employ conservation of
angular momentum . The angular momentum is a functional on the phase space
of the reduced system only for the total momentum . So, we
consider the stability in the class of solutions with .
In particular,
all solutions with initial states (1.4) have zero total momentum.
The linear orbital stability of the
solitons refers to
the stability of the
linearized reduced dynamics
in the tangent space to the manifold of states with constant angular momentum.
The manifold is not well-defined, but the tangent space is;
see Section 9.
So, the nonlinear stability problem on the manifold is not correct,
as opposed to
the linearized problem on the tangent space.
To prove the linear orbital stability,
we
construct the Lyapunov function
for the linearized equations
as the sum of the
quadratic part of the
Hamiltonian with the Casimir functional.
Such a strategy is an
analog of the energy–Casimir method
[13, 36]; see Remark 10.4.
We prove the conservation of in the linearized dynamics
and
establish a suitable lower bound for .
A lower bound provides a priori estimate
and existence of global solutions, and at the same time
the stability of the linearized dynamics.
The bound is proved
by suitable spectral arguments including the Heinz inequality from the theory
of interpolation [21] and the closed graph theorem.
Further comments on our methods and open questions can be found in
Appendix
A.
Let us comment on related work.
The mathematical theory of Maxwell–Lorentz system (1.1) started after
Nodvick’s paper [37], where the Hamilton least action principle
and conservation laws
have been established
in the Euler angles representation.
The coordinate-free proof of the conservation laws
was given by Kiessling [22].
The coordinate-free proof of the Hamilton principle
was given in
[18] using the technique of [4, 38] which relies on the
Lie algebras and the Poincaré equation.
This technique was developed
in [19], where the general theory of invariants was constructed
for the Poincaré equations on manifolds and applied to the system (1.1).
In 1990’–2000’, the stability of solitons was studied
for the Maxwell–Lorentz system (1.1)
without spinning
(the first three lines with ).
In [5], Bambusi and Galgani proved the
existence and orbital stability of solitons for
such a system.
The proof relies on the transition to a comoving frame
and a lower bound for the reduced Hamiltonian.
The
global convergence to solitons for the same system
with relativistic kinetic energy of the particle
was proved in
[15].
In [32], the global attraction to stationary states
was established in presence of an external
confining potential.
All these results
were obtained
under the Wiener condition on the Fourier transform of the charge density:
|
|
|
(1.5) |
which allows us to apply the Wiener Tauberian theorem.
This method
was introduced first in [31, 33] for the system
of the particle coupled to the scalar wave field
instead of the Maxwell field.
The condition means a strong coupling of the charged particle to the
eigenfunctions of the continuous spectrum of the free field.
The results [15, 32]
provide the first rigorous proof of the radiation damping
in classical electrodynamics.
For the survey, see [40] and [25, 26].
In [17, 29],
the asymptotic stability of the solitary manifold
was proved for the 3D and 2D
Maxwell–Lorentz systems without spinning:
for large times,
any solution with
initial data close to the solitary manifold
is close to the sum of a soliton and a dispersive wave.
The proof relies on the Buslaev–Perelman–Sulem method of
symplectic projection onto solitary manifold [7, 8] and uses
the Wiener
condition (1.5).
The adiabatic effective dynamics of solitons for the 3D system was proved
in [35].
For the Maxwell–Lorentz system with spinning particle
at rest
(),
the global convergence to rotating solitons
was established in [16] and [34].
In [16],
the convergence is proved
in the case of sufficiently small charge density .
In [34], Kunze extended the result to
the same system without the smallness condition.
This condition is replaced by the assumption that
the mass of the particle does not belong to a discrete set of resonances.
The stability of solitons for the Maxwell–Lorentz system (1.1) with moving and spinning particle was not considered previously.
2 The Maxwell potentials and the well-posedness
As is well-known [20], the second line of the system (1.1) implies
the following expressions of the Maxwell fields in terms of the potentials
and :
|
|
|
(2.1) |
We choose the Coulomb gauge
|
|
|
(2.2) |
Then the first line of the system (1.1) becomes
|
|
|
(2.3) |
Hence, is the Coulomb potential .
The gradient
in the first equation in (2.3)
can be eliminated
by the application of the
orthogonal projection in onto the
divergence-free vector fields.
Substituting the expressions (2.1) into
the system (1.1)
and using the notation for the current density,
we obtain the system
|
|
|
(2.4) |
In the last two equations, the terms with cancel since
|
|
|
Indeed,
the first equality holds
since the function is even.
The second equality follows from the rotation invariance (1.2).
Denote the spaces of real functions
with ,
the Sobolev spaces
with ,
and
being the completion of with the norm
.
Definition 2.1.
Denote i)
the Hilbert space
endowed with the norm of , and ii) the Hilbert space which is the completion of with respect to the norm .
Denote the Hilbert spaces
|
|
|
(2.5) |
For
let us denote
|
|
|
(2.6) |
The notations are suitable versions of
the basic operation [3, (10)] from the
theory of coadjoint representation
for the groups of translations and rotations, respectively.
Denote by and
the momentum and angular momentum
of the particle in the Maxwell field,
|
|
|
(2.7) |
Proposition 2.3.
i) For any initial state , the system (2.4) admits a unique
solution
|
|
|
(2.8) |
ii) The map
is continuous in for every , and
the energy and momentum are conserved:
|
|
|
|
|
(2.9) |
|
|
|
|
|
(2.10) |
iii) Let for and ,
|
|
|
(2.11) |
Then for any , , and
, and
|
|
|
(2.12) |
For such solutions,
|
|
|
(2.13) |
The items i) and ii) were proved in [27], and
(2.13) is proved in
[22] by a partial integration; see also [28].
The bounds (2.12) justify the calculations.
Let us comment on the proof of (2.12).
In the first equation in (2.4), the current
reads
|
|
|
In the Fourier transform, .
Hence, as . Therefore,
for , so
|
|
|
Now the decay (2.12) follows from the Kirchhoff formula for solutions
to the wave equation.
3 The Hamilton–Poisson structure
In this section we recall
the Hamilton–Poisson structure of the system (2.4)
which has been
constructed in [27]. The construction
relies on
the Hamilton least action principle for the system (2.4)
established
in [18, 37].
Using this principle and the Lie–Poincaré technique [3, 14],
we have shown in
[27] that
the system (2.4)
admits the following representation:
|
|
|
(3.1) |
Let us rewrite
the energy (2.9)
as the Hamiltonian functional:
|
|
|
|
|
(3.2) |
|
|
|
|
|
The Hamiltonian is well-defined and Fréchet-differentiable
on the Hilbert phase space defined in (2.5).
The system (3.1)
admits the
Hamilton–Poisson
representation
|
|
|
(3.3) |
Note that the operator is not invertible for all .
In detail,
the system
(3.3) reads as
|
|
|
The system admits the Casimir invariants
|
|
|
(3.4) |
Indeed,
, and so
|
|
|
|
|
(3.5) |
|
|
|
|
|
since the structural operator is skew-symmetric by (3.3).
5 The reduced system
The Hamiltonian (4.4) does not depend
on the variable . Hence, the equations for and
in the system (4.5)
reduce to
|
|
|
(5.1) |
which correspond to the conservation
of the conjugate momenta .
As a result,
the system
(4.5) reduces to the following family of the
Hamiltonian systems with the parameter :
|
|
|
(5.2) |
where
|
|
|
(5.3) |
and
denotes the reduced Hamiltonian
.
According to (4.4) and (3.2),
|
|
|
|
|
(5.4) |
|
|
|
|
|
In detail, the system (5.2) reads as
|
|
|
(5.5) |
where
while
and are given by
|
|
|
(5.6) |
The Hamiltonian (5.4) is
well-defined and Fréchet-differentiable on the reduced
phase space
|
|
|
(5.7) |
and it is conserved along trajectories of the system (5.5).
6 The solitons
Here
we calculate the solitons
of the system (3.1), which
are solutions of the form
|
|
|
(6.1) |
The second line of (3.1) and the definition (2.7) imply that
|
|
|
(6.2) |
The corresponding conserved (total) momentum is given by (2.10):
|
|
|
(6.3) |
In the comoving frame,
the solitons
become stationary solutions
of the reduced system
(5.5):
|
|
|
(6.4) |
The first two equations imply that
|
|
|
(6.5) |
Note that due to the spherical symmetry (1.2),
while on the Fourier transform side one has
.
Hence,
solving (6.5) on the Fourier transform side, we obtain:
|
|
|
(6.6) |
Now
the first equation of (6.4) gives
The conditions
(1.2) imply that
|
|
|
(6.7) |
It remains to check the last equation in (3.1). It is equivalent to the relation
|
|
|
(6.8) |
Lemma 6.2.
Let all conditions (1.2) hold.
Then the relation (6.8) holds
if and only if . In this case,
|
|
|
(6.9) |
Above, is the effective moment of inertia of the soliton.
We will prove this lemma in Appendix B.
By (1.2), we have
|
|
|
(6.10) |
Corollary 6.3.
The soliton exists and is unique
if . The solitons do not exist for .
The following lemma is a particular case of Lemma 5.2 from [27].
Lemma 6.4.
The total momentum (6.3)
admits the representation
|
|
|
(6.11) |
where the effective mass is a srictly increasing continuous
function of and .
Now we formulate the definition of the orbital stability
of solitons (6.1)
in terms of reduced
trajectories
corresponding to solutions
of the system (3.1).
Introduce the distance
|
|
|
(6.12) |
Here
where
, are defined according to (4.1), while
, are expressed by (5.6).
Suppose that
the initial data is close to in the following sense:
|
|
|
(6.13) |
Definition 6.5.
The soliton (6.1)
is orbitally stable if
for any there
is such that, for any solution
of the system (3.1)
the inequality (6.13)
implies that
|
|
|
(6.14) |
Obviously, for all solitons are orbitally stable; see
Remark 1.1.
7 The Lyapunov function
To prove the stability of a soliton , we must construct a Lyapunov
function which is an invariant and admits a lower bound:
|
|
|
(7.1) |
with some .
We will construct the Lyapunov function as a perturbation
of the Hamiltonian by a suitable Casimir
invariant.
In this section we calculate
for ,
where
|
|
|
(7.2) |
according to (2.2).
We have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
After rearrangements, we obtain
|
|
|
|
|
|
(7.3) |
where
|
|
|
(7.4) |
due to (6.3) and (6.2).
Using the first two equations of (6.4) and taking into account (7.2), we obtain:
|
|
|
|
|
|
|
|
|
Substituting into (7), after rearrangements, we get
where
|
|
|
|
|
(7.5) |
|
|
|
|
|
(7.6) |
|
|
|
|
|
As a result,
|
|
|
(7.7) |
The first term admits a lower bound
|
|
|
(7.8) |
9 Angular momentum conservation for reduced system with
For , the solitons are not critical points of the reduced Hamiltonian
because of
(8.2). In this case
we will consider the Lyapunov function which is a perturbation
of the Hamiltonian by a Casimir invariant:
|
|
|
(9.1) |
This function is an invariant for the reduced system (5.5),
and the solitons are its critical points since
|
|
|
by (8.2) and (6.9).
Now
(7.7) leads to
|
|
|
|
|
(9.2) |
|
|
|
|
|
The presence of the last “negative” term makes problematic
a lower bound of type (7.1),
and
perturbations of
by other Casimir invariants do not help; see Appendix A.
That is why we need to involve an
additional invariant of our system: the angular momentum.
However, the expression (2.13) is a functional on the phase space
of the reduced system only in the case of zero total momentum .
This is why we consider
below
solutions and solitons
of the system (3.1)
with the total momentum .
In this case,
the reduced system (5.5) reads
|
|
|
(9.3) |
where
and are given by (5.6) with .
The angular momentum conservation (2.13) with becomes
|
|
|
(9.4) |
Stationary solutions of this system are the solitons with .
This follows from the formula
established in [27, (5.17)].
We denote the solutions as .
Due to the first equation of (6.4) with ,
|
|
|
(9.5) |
Finally, the Hamiltonian
(5.4) with
becomes
|
|
|
|
|
(9.6) |
|
|
|
|
|
The set is dense in by Remark 2.2.
Proposition 2.3 iii) implies the following lemma.
Lemma 9.1.
For solutions to (9.3) with
initial state , we have
for any , and
,
that
, and
|
|
|
(9.7) |
and the angular momentum conservation (9.4) holds.
The conservation (9.4) suggests
that
the stability of solitons for the nonlinear system (9.3)
follows from a lower bound
of type (7.1)
on the manifold
|
|
|
(9.8) |
for the function (9.2) with .
However, the expression (9.4) is
not well-defined on the phase space , so
the manifold is also not well-defined.
On the other hand,
the “tangent space” can be defined as
|
|
|
(9.9) |
where . Indeed,
according to
(9.4),
|
|
|
|
|
(9.10) |
|
|
|
|
|
where and belong to
by (6.6) with . Moreover,
|
|
|
(9.11) |
Indeed, (1.2) and
(6.6) with imply
that
|
|
|
(9.12) |
Definition 9.2.
Denote by the Hilbert space which is the closure of in endowed with the norm of .
The space admits also the representation (9.9) with .
So,
is a closed linear subspace of codimension three in .
Denote by the orthogonal
projection of onto :
|
|
|
(9.13) |
where are suitable normalization factors.
10 The stability of the linearized dynamics in the tangent space
Let us linearize the reduced system
(9.3)
at the soliton .
We write the system as
|
|
|
(10.1) |
The variational equation is obtained
by replacing
with
and
keeping the terms linear in . Formally, the variational equation reads as
|
|
|
(10.2) |
Accordingly,
the linearization of (10.1) at the soliton is
|
|
|
(10.3) |
Let us calculate the corresponding Hamilton–Poisson structure for the linearized
dynamics.
Substitute
into (5.2). Then (5.3) implies that
|
|
|
(10.4) |
We have to keep in the right-hand side of (10.4)
only linear terms in .
Thus,
it suffices to keep in only linear and quadratic terms.
Hence,
we can replace with by
its quadratic part: according to (7.7) and (7.4),
|
|
|
(10.5) |
Note that the last linear term contributes only to the last component
of the differential,
|
|
|
Moreover, due to the structures of the matrices and
,
this component contributes only to the equation for :
|
|
|
|
|
(10.6) |
|
|
|
|
|
since by (6.9).
The corresponding linearized equation is
|
|
|
(10.7) |
Now it is clear that
the linearized equation (10.3) reads as
|
|
|
(10.8) |
Now we are going to prove that the tangent space is invariant with respect
to the linearized dynamics (10.3). As the first step, we prove
that the vector field is tangent to .
Lemma 10.1.
For , one has
|
|
|
(10.9) |
Proof.
Let us consider solutions to (10.1)
with initial states . Then
the angular momentum conservation (9.4) holds by Lemma 9.1.
Differentiating (9.4) in time, we obtain by (9.11) that
for
Therefore,
|
|
|
(10.10) |
Taking into account (9.11) and the
density of in ,
we
conclude that the identity (10.10)
holds
for
with any vector
.
Differentiating
the obtained identity
in at ,
we get
|
|
|
(10.11) |
where stands for the operator of
differentiation.
However, , so
(10.9) follows since
.
∎
The following theorem on stability is the main result of the present paper.
Theorem 10.2.
Let conditions (1.2)
hold. Then
i)
For every there exists a unique solution
to (10.8), and the map
is continuous in
for .
ii) There is the uniform bound
|
|
|
(10.12) |
Proof.
It suffices to prove (10.12) as an priori estimate;
then
the theorem will follow by the same methods as Proposition 2.3.
The estimate holds due to the conservation of an appropriate “energy functional”. Let us show that the functional
can be chosen as
the quadratic part of the Lyapunov function (9.1) with :
|
|
|
(10.13) |
Indeed,
(11.1) shows that the needed lower bound holds.
It suffices to prove that the functional is conserved
under the linearized dynamics (10.8).
Assuming “a priori” the existence of the solution and
differentiating , we obtain by (10.8) and (10.7) that
|
|
|
|
|
(10.20) |
|
|
|
|
|
|
|
|
|
|
(10.21) |
Finally,
the conservation of provides (10.12) by
the following proposition.
Proposition 10.3.
The Lyapunov function admits a lower bound: for some ,
|
|
|
(10.22) |
This theorem is proved in the remaining part of the paper.
11 Lower bound for the Lyapunov function on the tangent space
In this section we prove
the lower bound (10.22).
We will prove a stronger bound than (10.22):
|
|
|
|
|
(11.1) |
|
|
|
|
|
The quadratic form is continuous
on the Hilbert space endowed with the norm of . Hence,
it suffices to prove the bound (11.1) on the space of smooth functions
which is dense in .
For the smooth functions, the bound (11.1) can be written as
|
|
|
(11.2) |
where
the brackets denote the inner product in the Hilbert space ,
and
|
|
|
(11.3) |
Here
by (6.10), and
denotes the operator ,
, with
the corresponding skew-symmetric matrix
|
|
|
The operator
is positive-definite and selfadjoint in with domain .
Hence,
is also a selfadjoint operator in with the same domain since the difference
is a bounded finite rank operator in
by (11.3) and
(1.2).
The common domain of both operators and contains the space
The space is dense in by Remark 2.2.
In the next section, we will prove the following proposition.
Proposition 11.1.
Let conditions (1.2) hold.
Then
i)
the operator is nonnegative:
|
|
|
(11.4) |
ii) and .
Proof of Proposition 10.3.
Consider the restrictions of the quadratic forms
and onto and
denote by and the corresponding nonnegative
selfadjoint operators in :
|
|
|
(11.5) |
The space is dense in and
the operators can be expressed as
|
|
|
(11.6) |
where is the projection (9.13).
Lemma 11.2.
Let conditions (1.2) hold.
Then
with sufficiently small , i.e.
|
|
|
(11.7) |
Proof.
We have . Hence,
it suffices to prove that
the operator is bounded
since then
for
which is equivalent to (11.7).
The difference is a finite rank bounded operator in . Hence,
is a finite rank bounded operator in
by (11.6) and (9.13) since by (9.9). Therefore, the selfadjoint operators
and have common domain
. Moreover, Proposition 11.1 ii)
implies that is injective, and
hence it is a bijection
of onto . Therefore,
the operator is well-defined on the entire Hilbert
space , so the operator is bounded by the closed graph theorem.
∎
By Proposition 11.1, is a selfadjoint nonnegative operator in
, similarly to .
Hence, (11.7) and
the Heinz inequality [21, Theorem 2] imply that
for all , i.e.,
|
|
|
(11.8) |
Finally,
this bound with obviously implies (11.2) with
.
12 The positivity of the Lyapunov function on the tangent space
Here we prove Proposition 11.1.
It suffices to prove the nonnegativity (11.4) for :
|
|
|
This form can be written as
for
where the brackets now denote the inner product
in the Hilbert space
and
|
|
|
(12.1) |
by (11.3).
Proposition 11.1 i) follows from the next lemma.
Lemma 12.1.
Let conditions (1.2) hold.
Then
the operator is nonnegative:
|
|
|
(12.2) |
Proof.
For , the action of
the Laplacian in the sense of distributions coincide with
the Friedrichs closure
of in the invariant space .
Hence, in (12.2) can be considered as a
selfadjoint operator in with domain .
Therefore,
it suffices to show that for the resolvent is a bounded operator.
The operator is selfadjoint in .
Hence, it remains to check that
|
|
|
(12.3) |
The calculation of
the kernel reduces to a finite-dimensional
problem in a standard way. Indeed,
the equation
with reads as
|
|
|
(12.4) |
Using the last equation, we rewrite the first one as
|
|
|
|
|
(12.5) |
Hence, on the Fourier transform side,
|
|
|
(12.6) |
Let us emphasise that the formula holds for since
.
Substituting (12.6) into the second equation of (12.4),
we obtain:
|
|
|
(12.7) |
The matrix is Hermitian for , so it is defined uniquely by its quadratic form:
|
|
|
|
|
|
|
|
|
|
Thus, the matrix is the scalar , and (12.3)
is equivalent to
|
|
|
(12.8) |
We note that the function is strictly increasing
while is strictly decreasing for , and
that both functions are continuous. Hence, (12.8)
holds since
. Indeed, the last equality is equivalent to
|
|
|
(12.9) |
which
holds since
by (6.9) with .
∎
Proof of Proposition 11.1.
The nonnegativity (11.4) follows from Lemma 12.1.
By (11.3), , where , and
is defined by (12.6) with :
|
|
|
(12.10) |
It is
important that since
.
Hence, .
Finally,
(9.9) and (9.10) imply that for
.
Then also by (12.10).
As a result,
.
Appendix B On the soliton parameters
ad i).
Substituting (6.6) into (6.2), we obtain:
|
|
|
|
|
(B.1) |
|
|
|
|
|
where we denote .
We consider the two cases and separately.
The case .
We may assume that and . Then
|
|
|
(B.2) |
Substituting into (B.1), we obtain:
|
|
|
(B.3) |
The case .
We may assume that and . In this case,
|
|
|
(B.4) |
by (B.1). Now Lemma 6.2 i) is proved.
ad ii). Now we consider the case . We may assume that with and , where and . In this case,
|
|
|
(B.11) |
It remains to show that
since then ,
so (6.8) breaks down by (B.1). Calculating in spherical coordinates, we obtain:
|
|
|
where due to the last condition of (1.2).
Hence, for , we obtain:
|
|
|
Appendix C Density of in the space of divergence-free vector fields
Here we prove that is dense in the Hilbert space . For any ,
its Fourier transform satisfies for . Hence,
with . For ,
let us denote
|
|
|
Then
|
|
|
(C.1) |
Note that .
Hence,
, so there exist such that as for every .
Therefore,
|
|
|
Hence,
there exists a subsequence as . It remains to note that
and .