1. Introduction
The Hamiltonian viewpoint offers a theoretical framework in lots of physical fields. In classical mechanics, it constitutes the basis for further developments, including the Hamilton-Jacobi theory, the perturbation approaches and the chaos [
1]. The canonical equations of Hamilton (CEH) in the field of classical mechanics are expressed as [
1]
where
and
are respectively the generalized coordinate and momentum, while
is the generalized velocity. The generalized momentum
is defined as
with
L being the Lagrangian. By the Legendre transformation
, the Hamiltonian
H is then obtained.
The CEH (
1) can be extended to the continuous system [
1]
where the subscript
s indicates the components of the quantity of the continuous system [
1],
and
denote the functional derivatives of
h with respect to
and
with
and
,
and
are respectively the generalized coordinate and momentum, and
h is the Hamiltonian density. The generalized momentum
is defined as
The Hamiltonian density
h is acquired by the following Legendre transformation
where
l is the Lagrangian density. But it is significantly different for the continuous system that
and
are now the functions of both the time coordinate
t and the spatial coordinate
x. It should be noted that the spatial coordinate
x is not the generalized coordinate, but is only the continuous index replacing the discrete
i in Equation (
1). To avoid confusion, we refer to time
t as the evolution coordinate.
h is a function of
and
but not
[
1], so
, then the first equation of Equation (
2) can be also expressed by
To our knowledge, the CEH in all current literatures are of the form (
2), which are constructed on the basis of the second-order differential system (SODS). Besides the SODS, there are a lot of the first-order differential systems (FODS) to describe physical phenomena. Among them, the nonlinear Schrödinger equation (NLSE) is just the universal FODS. A question is raised in the nature of things: what is the form of the CEH applicable for the FODS? Are the conventional CEH, Equation (
2), still valid for the FODS? In this paper, we gain a new CEH of the formal symmetry valid for the FODS, from which the NLSE can be expressed in a consistent manner. We also prove that the symmetric CEH is equivalent to the conventional CEH for the SODS. However, the conventional CEH can not model the FODS.
It is well known that the symmetry plays an important role in theoretical physics [
2]. The search for and the discovery of new symmetries promote the exploration of fundamental laws of physics. Based on the idea, the CEH with the symmetry found by us, although this symmetry is only formal, might find their appropriate position in modern theoretical physics.
2. The CEH for the FODS
The Newton’s second law, the base of the Hamiltonian formulation, is modeled by a second-order differential equation of the evolution coordinate (the time coordinate). Here, we define the system governed by second-order partial differential equations of evolution coordinates as the SODS. Similarly, the FODS is the system governed by the first-order partial differential equations of evolution coordinates. The Lagrangian density of the SODS of the continuous systems is expressed in general as [
1]
where
depend on not only
but also
in general. The generalized momentum can be obtained by the definition (
3) as
which is the function of
,
and
. Equation (
7) in fact have
N equations and contain
variables, which are
,
,
and
. So the degree of freedom of Equation (
7) is
. Then we take
and
as independent variables, and express the generalized velocities
by these independent variables.
Besides the SODSs, there are a number of the FODSs. The nonlinear Schrödinger equation (NLSE)
is just a universal model that can be applied to hydrodynamics [
3], nonlinear optics [
4,
5,
6], nonlinear acoustics [
7], Bose-Einstein condensates [
8]. In nonlinear optics [
4,
5,
6], the NLSE (
8) governs the propagation of the slowly-varying light-envelope with the evolution coordinate
t being the propagation direction coordinate. The light-envelope
is a cw paraxial beam in a planar waveguide [
5] or a narrow spectral-width pulse in optical fibers [
4,
6].
x is a transverse space coordinate for the beam and a frame moving at the group velocity (the so-called retarded frame) for the pulse, respectively.
For the FODS, the Lagrangian density should be a linear function of the generalized velocities
. If the Lagrangian density is a quadratic function of the generalized velocities like Equation (
6), the equation of motion, i.e., the Euler-Lagrange equation
will be the second-order partial differential equation of the evolution coordinate
t, which is in contradiction with the definition of the FODS. Therefore, the Lagrangian density of the FODS can only be expressed as
Besides,
in Equation (
10) is not the function of
. If
is also the function of
, there will be such terms as
appearing in Equation (
10). Substitution of Equation (
10) into Equation (
9) leads to the appearance of the mixed partial derivative terms
. Via the coordinate rotation transform, terms
and
will appear instead. Therefore, the Euler-Lagrange equation (
9) expressed with the canonical form of the second-order partial differential equation [
9] is the second-order partial differential equation about the evolution coordinate
t. According to our definition, the system is the SODS. Consequently, the generalized momentum
, obtained by definition (
3)
is only a function of
. This is of significant difference from the case of the SODS, where the generalized momentum
is the function of not only
, but also
and
, as shown in Equation (
7). Equation (
11) have
N equations, but contain
variables,
and
. Therefore, the degree of freedom of the system described by Equation (
11) is
N. We can take
and
as the independent variables, where
. The rest of generalized coordinates and momenta can be expressed by the independent variables
and
We now derive the CEH for the FODS. The total differential of the Hamiltonian density
h can be obtained by using Equation (
4)
while the total differential of the Lagrangian density
with respect to its arguments is
Substitution of Equation (
13) into Equation (
12) yields
Since the total differential of
with respect to its arguments can be written as
by comparing this equation with Equation (
14), we obtain
equations
where
. From the Euler-Lagrange Equation (
9), we obtain
substitution of which into Equations (
15) and (
16) yields
Then substituting Equation (
17) into Equations (
18) and (
19), we obtain
N CEHs for the FODS
To obtain Equation (
21), we have used
, because
h is not a function of
. The CEH, Equations (
20) and (
21), can be expressed in a symmetric form as
(
,
, and
), because
,
and
. The CEH above can be extended to the discrete system
where
,
, and
.
We will prove that the symmetric CEH (
22) and (
23) can also describe the SODS. In Equations (
22) and (
23), if all the generalized coordinates
and momenta
are independent, we can obtain that
(
), then the CEH will be reduced to Equation (
2). It does be the case of the SODS, where all the generalized coordinates and momenta are independent. So, the symmetric CEH obtained by us can express both the FODS and the SODS. In other words, the new CEH and the conventional CEH are equivalent when describing the SODS, but the former are of some formally symmetry. The conventional CEH, Equation (
2), can only be used to expresses the SODS.
3. Application of the Symmetric CEH for Continuous Systems to the NLSE
In this section, we will apply the new CEH with symmetry, Equations (
22) and (
23), to the NLSE. The Lagrangian density for the NLSE is stated as [
10]
. The NLSE is complex, so it is in fact an equation of two real functions, one is of its real part and the other is of its imaginary part. Alternatively, we can take the fields
and
as two independent functions. In this sense, the components of the quantity
N for the NLSE is equal to two. In other words, for the NLSE there are two generalized coordinates,
and
, and two generalized momenta
By using Equation (
4) [
4], we can obtain the Hamiltonian density
If we take
and
as the independent variables,
and
can be expressed by the relations (
25) as
and
, respectively. It also should be noted that
h is also the function of
. Then, we can express the Hamiltonian density (
26) with independent variables
and
as
For the NLSE,
and
, therefore the Equation (
22) in fact have only one equation. It is the same case for Equation (
23). Consequently, for the NLSE, the CEH (
22) and (
23) can produce two equations. From the left side of Equation (
22), we can obtain
From its right side, we can have
Then the NLSE (
8) is obtained. While, for the other CEH (
23), the left side is
and the right side is
which results in the generation of the complex conjugate of the NLSE. As a result, the CEHs (
22) and (
23) are consistent. From one of the two CEHs, the NLSE is expressed; from the other, the complex conjugate of the NLSE is expressed.
We now demonstrate that the conventional CEH (
2) can not be used to express the NLSE. By Equation (
3), we have
. Substituting the Hamiltonian density (
26) into the second equation of Equation (
2) only yields
which does not be the NLSE (
8). By substituting the Hamiltonian density (
26) into Equation (
5), we can obtain the left side
, and the right side
, where
. Then the equation
can be obtained. It is surely not the complex conjugate of the NLSE.
4. Application of the Symmetric CEH for Discrete Systems to Light-Envelope Propagations
In this section, we use the symmetric CEH for discrete systems, i.e., Equation (
24), to discuss light-envelope propagations in nonlocally nonlinear media, which is modeled by the following (1+D)-dimensional nonlocal nonlinear Schrödinger equation (NNLSE) [
11,
12,
13,
14]
is the nonlinear refractive index, which can phenomenologically be expressed as a convolution between the response function
and the light intensity
When
R is the Dirac delta function, the NNLSE (
29) is reduced to the NLSE (
8).
We assume the trial solution of Equation (
29)
where
are the amplitude and phase of the complex amplitude
, respectively,
is the width,
is the phase-front curvature, and they all vary with
z. We consider the response function
Substituting the trial solution (
31) into the Lagrangian density
and performing the integration
, we have
which is the function of generalized coordinates
and velocities
.
Then the generalized momenta can be obtained
By Legendre transformation, the Hamiltonian can be determined
which will be proved to be a constant, that is
Four generalized coordinates and four generalized momenta are contained in the four equations (
34). It indicates that Equation (
34) have four degrees of freedom. Here, we can take
and
as independent variables. By solving Equation (
34), we can express the generalized coordinates
and
by generalized momenta
and
as
and
Then, the Hamiltonian (
35) is rewritten as
By using the CEH (
24) with
and
, we can obtain four equations as
One can find that the generalized coordinate
does not appear in the Hamiltonian (
36), therefore
is a cyclic coordinate. Because the generalized momentum conjugate to a cyclic coordinate is conserved [
1], the generalized momentum
conjugate to the generalized coordinate
is a constant, which can be confirmed by Equation (
40). In fact, this represents that the power
is conservative. Then we can obtain
Taking derivative with respect to
z on two sides of the third equation of (
34), then comparing it with Equation (
39) we obtain
the substitution of which into the Hamiltonian (
35) yields
where
are the generalized kinetic energy and potential, respectively.
From the Hamiltonian point of view, the dynamics of light-envelopes in nonlinear media can be regarded as a problem of small oscillations of a Hamiltonian system about its equilibrium position. The equilibrium state of the system described by the Hamiltonian
H corresponds to the soliton solutions of the NNLSE, which can be gained as the extremum points of generalized potential
V. The equilibrium position is stable if a small disturbance from equilibrium leads to small bounded motion about the rest position. While, if an infinitesimal disturbance produces unbounded motion, the equilibrium is unstable [
1]. In other words, the equilibrium must be stable when the extremum of the generalized potential is a minimum, and unstable otherwise. In this sense, therefore, the viewpoint in a few literatures [
15,
16,
17,
18], where solitons were taken for the extremum of the Hamiltonian rather than the generalized potential, might be somewhat ambiguous. In these literatures [
15,
16,
17,
18] the trial solution has an invariable profile (soliton profile), and the soliton state is the static state in fact. In this case, the kinetic energy is zero, and the Hamiltonian is just equal to the potential. In this connection, for the static system the extrema of the Hamiltonian and the generalized potential are equal only in value. Although in such literatures [
15,
16,
17,
18] the obtained soliton solutions are correct, it is more reasonable to regard the soliton solutions as the extremum points of the generalized potentials rather than the Hamiltonian.
To find the equilibrium state (soliton solution), we let
, then we obtain
The critical power is then obtained as
with which the light can propagate with a changeless profile. Besides,
also leads to
which indicates that the soliton wavefront is a plane.
Then we address the soliton stability by discussing the generalized potential
V. Performing the second-order derivative of
V with respect to
, and substituting
into it, we have
where
represents the degree of nonlocality. When
, generalized potential
V has a minimum, then the soliton is stable. From Equation (
44) we can get the stability criterion of solitons
which also agrees with the Vakhitov-Kolokolov criterion [
19].
4.1. The Local Case
When
,
, and the NNLSE is reduced to the NLSE (
8). Then, Equations (
43) and (
44) will be reduced to
In the case of
,
, it is the same as Equation (42) of Ref. [
10]. In the case of
,
, it is the same as Equation (16a) of Ref. [
20]. We can obtain
if
,
if
, and
if
. Therefore, in the local case, the soliton is stable if
, but unstable if
. In the case of
, it needs further analysis because
. In the case of
, the generalized potential
which does not have extreme when
. When
, we obtain that
. It is the extreme rather than the minimum. Hence, (1+2)-dimensional local solitons are always unstable. When
, the potential
is constant, the light-envelope without the external disturbance will stay in its initial state. But the ideal condition can not occur in experiment. If the external disturbance makes
, the beam will become narrower and narrower, and the optical beam will eventually collapse. If the external disturbance makes
, the optical beam will diffract at last. These conclusions all agree with those of Refs. [
21,
22,
23].
4.2. The Nonlocal Case
When
and
, the condition (
45) is satisfied automatically. It means that the (1+1)-dimensional and the (1+2)-dimensional solitons are always stable in nonlocally nonlinear media of Gaussian response. It is consistent with the conclusion of Ref. [
24]. When
the stabe solitons can exist when the degree of nonlocality should be strong enough that satisfies the criterion (
45), which is also the same as Ref. [
24].