It is well known that Equation (
4) presents solitons under suitable assumptions on
W. It was largely studied during the 1970s and the 1980s. The first rigorous result for the finite energy solution was due to Strauss [
14], and later Beresticky and Lions [
15] gave sufficient and “almost necessary” condition for the existence. In [
10], there is a detailed analysis of the case in which
. If we couple (
4) with the Maxwell equation via the interaction (
2), the solitons usually are called
Q-balls (Coleman [
16]). The first rigorous result about the existence of
Q-balls was established in 2002 [
17]. Afterwards, their stability was proved in [
18]. A detailed analysis of
Q-balls and the references to the large literature can be found in [
10]; in all these papers, the interaction between the solitons and the e.m. field is established by the Lagrangian (
2).
3.1. Existence of Stationary Waves
Let us prove the existence of some particular solution of Equations (
11)–(
14); first, we look for stationary solutions, namely solutions where
is a stationary wave, i.e.,
We make the following ansaz
Replacing these variables in (
11)–(
14), Equations (
12) and (
14) are identically satisfied, while Equations (
11) and (
13) become
These two equations have nontrivial solutions, provided that suitable conditions on
are satisfied: we write
W as follows,
In the model of our interest, N must be considered as a small perturbation of the parabola However, in order to get an existence result, it is sufficient to make the following assumptions on N
(N-1)
(N-2) ( is allowed also to be );
(N-3) There exist and , such that
We will show that, at least for
small, the above assumptions guarantee the existence of nontrivial solutions to Equations (
46) and (
47). In most of the literature relative to (
4), we usually have the following choice of
This assumption implies the existence of nontrivial solutions also for Equations (
46) and (
47) for every
. However, in our model, it is more interesting (see Theorem 3) to choose a “bump-like”
N, such as
or a “bell” function such as
where
is a small parameter which makes
Its relevance will be discussed in Theorem 3.
We define the following bilinear form
Notice that
where
is the Green function relative to the Poisson equation
namely
.
Now, let us introduce a number
, which is very relevant in this study of solitons
depends on
and the shape of
For example, if
is given by (
49), then it is immediately necessary to check that
for every
. If
is given by (
50),
depends on
(see Corollary 1).
We have the following theorem.
Theorem 2. If (N-1) and (N-3) hold and if , then for every Equations (46) and (47) have nontrivial solutions in . Proof. The couple of Equations (
46) and (
47) can be easily solved by standard variational methods; we will here provide a sketch of the proof, avoiding standard estimates which are well known among people working in nonlinear analysis.
Since
V is a Hilbert space equipped with scalar product
and norm
Moreover, by the definition of
Then, using the Gagliardo–Nirenberg–Sobolev estimate, we can see that
Now, we define, on
V, the following functional:
By (N-3), (
55) and standard arguments,
J is a differentiable functional in
So we have to prove two facts: (1) the critical points of
J solve Equations (
46) and (
47) and (2) if
J has at least a nontrivial critical point.
(1) We have that
and by well known arguments, we have that
Taking account of (
48)
finally, setting
we get Equation (
46) while Equation (
47) follows from the definition of
(2) The simplest way to prove the existence of critical points of
J is the use of the Mountain Pass theorem of Ambrosetti and Rabinowitz [
19]. Following standard arguments, it is easy to prove that
J satisfies the Palais–Smale condition (for a very similar result, see [
17], Lemma 4.3). The interesting fact is to check the conditions, which guarantee the geometry of the Mountain Pass theorem, namely, that
and
If
is sufficiently small, by (
48), (
N-1), (
N-3) and standard computations,
such that
then, if
Then if
r is sufficiently small
and (
57) is proved. (
58) holds by the definition (
54) of
. □
Corollary 1. If (N-1), (N-2) and (N-3) hold, then there exists , such that for every Equations (46) and (47) have nontrivial solutions in . Proof. By Theorem 2, it is sufficient to prove that
By (
N-2), we can choose a point
such that
The first part can be estimated as follows:
For the second part, we have that
Therefore, we can choose a
so large that
and
so small that
□
3.2. Stationary q-Solitons
Using the equivariant Mountain Pass theorem and exploiting the fact that the functional (
56) is even, it is possible to prove that Equations (
11)–(
14), have an infinite number of radially symmetric solutions of the form (
44) and (
45), namely solitary waves. We call
ground state solution, the radially symmetric solution
, which minimizes the following quantity
in
Clearly, at least for a generic
W, this solution is unique and it corresponds to the critical value determined by the Mountain Pass Theorem, having chosen
, which minimizes
. Notice that
is the ratio of the matter energy (
29) and the hylenic charge (
35). If
the ground state solution, is a soliton in the sense that it is orbitally stable (see, e.g., [
18] or [
10]).
By Theorem 1, Equations (
11)–(
14) define a dynamical system whose phase space is given by (
43). A generic point of the phase space, at time
can be represented as follows:
From now on,
we will denote the ground state solution of Equations (
46) and (
47), namely
where
is a possible phase shift which is not relevant and, from now on, it will be neglected. Such a function will be called a
q-soliton. We have chosen this name to emphasize the comparison with the
Q-balls which are stable configurations of the equations determined by the action
(see (
2) and (
3) and the discussion at the beginning of
Section 3). Roughly speaking, a
Q-ball behaves like a swarm of charged particles kept close to each other by the gluing force determined by
(see [
9] or [
10] Section 5.1.5). Instead, as we will see in this and the following sections, a
q-soliton behaves like a single particle of “matter” condensed by the gluing force determined by
.
The
q-soliton, has a positive electric charge
(and hence, by (
47),
). However, the Equations (
11)–(
14), have a solution with negative charge, given by
Then, (
11)–(
14), have at least two orbitally stable solutions determined by a
q-soliton
and a
q-antisoliton
:
Generally, they are unique up to space–time translations and phase shift. Rotations do not produce new solutions, since is radially symmetric.
The “shape” of a soliton is determined by the nonlinear
. In [
20], there is a detailed analysis of this topic in the case
. Clearly, this analysis can be extended to the
q-soliton when
is small. The next theorem examines some properties of the
q-solitons in the case, in which
is a small bump such as (
51)
Theorem 3. For every we can choose N, such that
if is the Mountain Pass solution of Equations (46) and (47), then
Proof. We choose
N to be a bell function such as (
51), so that
and
Then, the first inequality is trivially verified. In order to prove the second inequality, we see that
and so, if we put
we have that
Therefore, by the definition of
(
54), we have that
and hence
.
The third inequality follows applying to Equation (
46) the maximum principle. The details of the proof can bo found in [
20]. □
Remark 3. The picture which comes out from Corollary 1 and Theorem 3 is the following: given the free electromagnetic field and the free matter field relative to KG, we obtain q-solitons, provided that
If we want to analyze the properties of a
q-soliton considered as a model for physical particles, it is useful to rewrite Equation (
11) with dimensional constants. We get the following equation (which is satisfied by
if suitably rescaled)
In this equation,
c is the speed of light which makes the equation invariant for the Lorentz transformations with the parameter c;
u has the dimension of
this fact can be deduced, e.g., by the fact that, by Theorem 1,
has the dimension of energy;
has the dimension of a frequency; if we linearize Equation (
61) with
, we get KG
which has the following dispersion relations
where
and
are the frequency and the wave number of the small perturbations of the matter field. Since
, the oscillations of the
q-soliton, having frequency
, do not excite dispersive waves in the surrounding matter field. This fact partially explains the stability of the soliton;
If we give to
the same dimension of
u,
ℓ has the dimension of a length and it is of the order of the radius of the soliton in the sense that
where ≅ means that the quantity is exponentially small and
k is a dimensionless variable which depends on
Here, S is supposed to be dimensionless;
represents the strength of the interaction of the matter field with the electromagnetic field; by (
17)
and if we give to
the dimension of an electric field, i.e.,
then
using these variables, then
defined by (
52) becomes
hence, if
is too large with respect to the other constants, then, by (
54), if
and there are no solitons. Actually, there is a competition between the gluing force which increases with
and the electric force, which increases with
The gluing force tends to concentrate the matter field while the electric force tends to spread it.
By this discussion, it shows that
represents the potential of the “nuclear force”, which is repellent when
u is small and attractive when the values of
u are in range, where
is negative.
is responsible for the nonlinear behavior of the matter field, and hence of the existence of
q-solitons. By these considerations, Theorem 3 and Remark 3, a
q-soliton is a good model for physical particles if, in the dimensionless equation
Additionally, the condition
is suitable for a physical model.
We denote by
the equilibrium configuration containing a
q-soliton. The energy of this configuration is given by
where
All these terms are positive;
is positive by (
29) and (
64);
is positive since, by Equation (
47)
and
have same sign. Thus, this term is also positive for anti-solitons. The energy
is concentrated around 0 in a region of radius
In fact, since
u decays exponentially, from the physical perspective, it can be considered null for
larger that a suitable
The field energy
is not concentrated; by Equation (
47), it decays as
. Finally, notice that the e.m. field energy of a soliton does not diverge as the energy of a pointwise particle would.
3.4. Mechanical Properties of q-Solitons
First, we will investigate the intrinsic quantities of a moving q-soliton. Since these properties are independent of R and , we will just consider with and
The simplest quantity to describe of a
q-soliton is the electric charge. It is defined by (
34) and, in this case, is
This depends only on the soliton and not on the configuration of the surrounding field. Moreover, it has the following property:
Proposition 3. The electric charge of a moving soliton is independent of the motion: Proof. By (
34) and (
71), making a change in variable
we have that
□
From now on, q will denote the charge of a q-soliton. Next, let us consider the mass:
Definition 2. If the momentum of the matter field, (see Proposition 2) is proportional to the constant of proportionality is called the mass the moving q-soliton, namely Remark 4. Notice that this definition of mass is intrinsic to the Equations (11)–(14) and it is independent of any physical interpretation; it can be interpreted as a “physical” mass whenever x and t are interpreted as variables of the physical space-time. Let us explicitly compute the momentum:
Theorem 4. The momentum of a q-soliton takes the following form: Proof. By Proposition 2,
assuming
by (
68)–(
71)
Making a change of variable
we get
Since
is radially symmetric,
then
It is immediate to see that , and hence we obtain the conclusion. □
Therefore, we have obtained the following result.
Corollary 2. The mass of a q-soliton is well defined and it takes the following value: From now on, m will denote the rest mass of a q-soliton.
We define the energy of a moving soliton as follows:
The next proposition describes how the energy transforms in a moving soliton:
Theorem 5. The energy of a q-soliton is given bywhere (see (34)). Remark 5. In a theory with the energy of a soliton coincides with its mass , and hence it transforms as the time-component of a time-like vector. If , part of the energy transforms differently. This fact is not so surprising since the energy of a q-soliton includes the energy of the self-interaction of the soliton with the e.m. field generated by itself. The energy-momentum of the e.m. field does not transform as the energy of a space-like vector since it is a light-like vector. Hence, there is a small term of the order β, which transforms differently. Since this term is related to the interaction of the matter field with the e.m. field it might be related to a sort of not quantistic counterpart of the fine-structure constant; however, this point needs further investigation.
In order to prove Theorem 5, we need the following lemma, which is a variant of the Pohozaev–Derrik theorem [
21,
22]:
Lemma 1. If u is any solution of Equations (46) and (47), then Proof. Let
be the functional
J defined by (
56). Then, if
u is a solution of Equations (
46) and (
47), we have that
Now, let us consider the “curve”
in
defined by
Making the change in variable
we get that
□
Corollary 3. Given a stationary q-soliton we have that Proof. Replacing
W in (
66), and using Lemma 1, we get
□
Proof of Theorem 5. By Proposition 1 and (
68)–(
73) we have that
making the change of the integration variable
, we get
Let us compute each piece individually
Since
u is radially symmetric,
Let us compute the second piece using (
79) again
In order to compute the third piece, we need (
70) and (
71):
The computation of the fourth piece uses Lemma 1:
Concluding, using Corollary 3, we have that
□
Placing a stationary
q-soliton in an generic electromagnetic field with gauge potential
and using the notation (
60), we get the following configuration,
By Proposition 1, the energy of this configuration is
If the soliton is small with respect to
(namely, if
is small), then, by (
75),
where “≅” means that the accuracy of this approximation is good if the quantities involved are large with respect to
(and to the radius of the soliton). In fact, the field
produced by the
q-soliton, is of the order of
and hence, if
we have that
and
Then
Therefore, thanks to Proposition 1 and our analysis if a soliton is placed in an e.m. field, we can distinguish the
soliton energy , the
potential energy and the
e.m. field energy . This distinction is crucial for the study of the dynamics of the soliton (see
Section 3.5). Finally, we remark that, the potential energy
is localized within the radius of the soliton. This fact eliminates one of the difficulties posed by the dualism particle-field where the localization of the potential energy of a particle is a meaningless problem.
If the q-soliton is moving, extending the above arguments, we have the following result:
Proposition 4. If the q-soliton is small with respect to and and , then Proof. By Theorem 5,
Then, by Proposition 1 and (
70), we get
and, using (
67), (
68), (
72), (
73) and causing a change in variables, we have that
where
L denotes the Lorentz boost defined by (
67), namely
If the soliton is small with respect to
and
then, using the definition (
75) of
q,
and similarly
□
Notice that (
80) is the energy of the soliton, namely the matter field energy plus the interaction energy contained in the radius of the soliton; the total energy of a configuration which contains a soliton also depends on
and, by Proposition 1 and Proposition 4, it takes the following form:
Now, let us examine a configuration containing several solitons
where
was defined by Definition 1. We assume that
where
denote the radius of the solitons. We remember that
u decays exponentially, so the matter field is essentially null out of a neighborhood of each soliton, and hence
where
Notice that, in the configuration (
82), the
q-antisolitons can be included. They have the same mass of solitons, but opposite electric charges.
If we embed this configuration in an external e.m. field, the total energy takes the following form
3.5. Dynamics of q-Solitons
Now, let us examine the dynamics of a solitons in the presence of an “external” electromagnetic field. More exactly, we want to examine the behavior of the solution of the Cauchy problem with the following initial conditions
where
is such that
It is well known that, thanks to the invariance of the hylenic ratio, the soliton is orbitally stable (see, e.g., [
10]). This means that if the perturbation field generated by
is small (with respect to
around the soliton, then the solution of the Cauchy problem has the following form
where
are essentially null thanks to the orbital stability of the soliton and they will be neglected;
is the configuration of the q-solitons and its structure is determined by a N function such that
Our aimis to investigate the dynamics of the q-solitons under the following assumptions:
(A-1)—as we have seen, this condition implies that the Cauchy problem is well posed and that the energy of a q-solitons equals its mass (Theorem 5);
(A-2) The solitons are far from each other (i.e., (
81) holds) during the time interval considered; this happens if
- -
(i) This assumption is satisfied by the initial condition (
83);
- -
(ii) All the q-solitons have the same charge (namely, there are not q-antisolitons), so that, during the evolution, the q-solitons repel each other;
- -
(iii) The e.m. field is not locally too strong, so that the q-solitons cannot collide;
(A-3) this fact avoids the q-soliton to produce a strong radiation and, from the technical point of view, it simplifies the computations. Clearly, this happens if the e.m. field is not too strong.
We will show, that, under these assumptions, the
q-solitons behave as classical particles. To this aim, we analyze the action functional relative to the configuration (
84)
Since we have assumed
(A-2), then
and
Let us compute each piece of the action separately
Lemma 2. Under the assumptions (A-1),(A-2),(A-3), we have that Proof. Since the Lagrangian
does not depend explicitly on
t and
we can choose a reference frame where, for a fixed
t,
so that
We recall that by (
68) and (
69),
where we have set
Then, by (
7), and (
68)–(
71)
If we assume that
is not too large (i.e.,
(A-2)),
then, arguing as in the proof of Theorem 5 and using similar notations for each
k,
where
Continuing with our computation,
The term
will be ignored, since we have assumed
(i.e.,
(A-2). Then, by (
77),
□
Now let us compute .
Lemma 3. If (A-1) and (A-2) hold, then Proof. As in the previous lemma, we can choose a reference frame where, for a fixed
t,
and
. Then, following the same arguments as used in the proof of Proposition 4, we have that
□
The above lemmas give the following result
Theorem 6. Let be the solution of the Cauchy problem relative to Equations (11)–(14) with the initial condition (83). Then If (A-1), (A-2) and (A-3) hold, we have that Proof. Making the variation in
with respect to
, we get the Lorentz Equation (
86); making the variation in
given by (
25), we obtain the Maxwell equations. □
Remark 6. Theorem 6 states that Equations (11)–(14) provide a model for material particles which, at low energies, agree with the well known physics. It is interesting to investigate the predictions of this model when the assumptions (A-2), (A-3) are violated. If (A-2)-(ii) is violated, there are antisolitons which attract solitons since they have opposite charges; then (A-2)-(i) will eventually be violated and the two particles will annihilate. Since our equation is invariant for time-reversal, the creation of an particle–antiparticle couple might occur; of course, this can happen only if there is sufficient energy, namely if (A-2),(iii) does not hold. If (A-3) does not hold, a numerical computation of the radiation when is large gives a spectrum which can probably be compared with the experimental data.