2.1. Generalities on Classical Symmetries, Canonical Structure, and Integrals of Motion
As generally admitted, the classical state of a non-moving physical system is described by a set of functions in a coordinate system. Under a coordinate transformation, these functions may take different forms. But if they remain invariant, the system is said to admit a symmetry.
Now, for a physical system in motion, its states are specified by functions of time
t and dynamical coordinates in
phase space. If under a transformation of dynamical coordinates its state functions remain invariant, we say that the system admits a
dynamical symmetry. The set of symmetry transformations may have a group structure, which is called
dynamical symmetry group. Such a dynamical symmetry group (or equivalently, the algebra of its generators) is of highest interest in the search for solutions of its equation of motion (The term
dynamical symmetry was coined by A. O. Barut [
5]. Earlier, such symmetry was known as
hidden or
accidental [
6]).
In this paper, we are concerned with the dynamical symmetries of a system in the framework of Hamilton’s canonical formalism of mechanics, with the time-independent Hamiltonian function H. We now recall some useful main points of this framework.
- A system with
n degrees of freedom is described by
canonical variables
with
in phase space, verifying the following fundamental Poisson bracket commutation relations:
where the Poisson bracket between two functions in phase space
and
is defined by the following:
and the canonical equations of motion:
the solutions of which depend on the
initial conditions
.
Consequently a dynamical symmetry exists if it originates from a canonical coordinate transformation in phase space, leaving both the Hamiltonian H and the fundamental Poisson bracket relations invariant. In differential geometry, this is called a symplectic structure. Non-trivial dynamical symmetry exists only for .
- Canonical transformations (see Chapters 10–12 of [
7]) are parts of a wider class of coordinate transforms in phase space called
contact transformations. A differentiable mapping
is called a contact transformation if the differential form
is the exact differential
of a function
.
This can be equivalently expressed by requiring the fundamental Poisson brackets in the variables
to be valid:
The Poisson bracket of two functions is invariant under the contact transformations . The lower index (resp. ) refers to the variables in the partial derivatives of the respective Poisson brackets. A contact transformation , which preserves the equations of motion of the system is a canonical transformation if is an integral invariant of the system (Jacobi’s theorem). Consequently, a canonical transformation implements a dynamical symmetry.
An
infinitesimal canonical transformation of parameter
has the following form:
where
is an arbitrary function of
and
is a small increment of the parameter
.
is called the
the generating function of the contact transformation. The variation
of an arbitrary function
f in phase space under an infinitesimal contact transformation of generator
is given by
.
can be an
integral of the motion, a function which takes a constant value and does not explicitly depend on time
t, see [
7]. A system of integrals of motion
is said to be in involution when
for all
. If
F and
are two integrals of the motion, their Poisson bracket is also an integral of the motion (
Poisson’s theorem). The set of all
forms a
Lie algebra with respect to the Poisson bracket.
The determination of all canonical transformations for a dynamical system is at the core of finding the dynamical symmetries of this system.
- For a system with an
n degree of freedom, with a time-independent Hamiltonian
H, total energy is
conserved and
. If there exists
n functionally
independent integrals of motion
, where
in involution (or
for all
), the system is said to be Liouville
integrable and its solution can be given up to
quadratures. If there are further
m functionally independent integrals of motion
, where
, the system is called
super-integrable [
8]. For
(resp.
) it is called
minimal super-integrable (resp.
maximal super-integrable). These
m extra integrals of motion usually build a Lie algebra. Maximal super-integrable systems are also known as
exactly solvable systems, and their properties can be derived algebraically. Each integral of motion
may originate from Noether’s conservation law or from a coordinate variable separation. The Kepler–Coulomb (or inverse distance potential) and the isotropic harmonic oscillator problems are known to be super-integrable systems in two dimensions [
9].
2.3. Dynamical Symmetries
As in two dimensions, the maximal number of possible symmetries is three. Iwai and Rew [
4] were the first to obtain dynamical symmetries by considering linear inhomogeneous transformations
in the phase space dependent on three parameters
:
Proposition 1. is a canonical transformation.
Proof. By the simple substitution of the
in terms of the
, as given by Equation (
6), it appears that the Hamiltonian function form is invariant:
Moreover, using the expression of the Poisson bracket and Equation (
6), one can also check the following:
Hence, is a bona fide canonical transformation. □
Proposition 2. For all triplet, the canonical transforms form a group—the dynamical symmetry group of the free fall problem in two dimensions.
Proof. Composition law. Let another canonical transform
with parameters
act on the previous one
and compute the resulting product
, which is expressed by the set of new parameters
Then, the substitution of Equation (
6) into Equation (
9) yields the following equations:
Hence, we conclude that the composition of two operations is as follows:
Because of the extra term , the previous formula does not correspond to the additive structure of a group operation with respect to the parameters . But since the v variable is a cyclic variable, this term is physically irrelevant’; in this way, the group structure is physically restored.
Iwai and Rew proposed to represent each element of the group with fifth-order, upper-triangular matrices, as can be checked by the matrix multiplication rule:
- The inverse transform is clearly given by since the extra term vanishes at and .
-The identity matrix is obviously the neutral group element. □
2.4. Infinitesimal Iwai–Rew Canonical Linear Transforms and Integrals of Motion
The infinitesimal form of the Iwai–Rew canonical linear transformations is written in the form of Equation (
8), with infinitesimal
as:
Proposition 3. The infinitesimal transformations in phase space defined by Equation (12) are canonical transformations: H and the Poisson bracketsare invariant.
Proof. By working out the Poisson brackets using Equation (
12), they will appear to remain in the canonical form of Equation (
8), while the form of the Hamiltonian in the variables
is the same as the form in variables
. □
We now determine the three generating functions
from the three sub-groups of canonical transformations, respectively defined by the following equations:
Then, partial derivatives of
with respect to
can be deduced as follows:
Therefore, we obtain the exact differentials
in terms of
(Schwarz’s theorem is trivially verified for all pairs of variables). Their integration yields the sought integrals of motion:
Proposition 4. The free fall system in two dimensions is super-integrable
since its has three integrals of the motion: Proof. Compute the three Poisson brackets and observe that they are zero. □
Proposition 5. The three generating functionsbuild a Lie algebra structure, the dynamical symmetry algebraof the two-dimensional free fall with respect to the Poisson bracket: Proof. Work out the Poisson brackets and use the expressions of the
from Equation (
15). They are identical to those of Iwai–Rew [
4]. If we make the substitutions
in our Equation (
15), we recover the Iwai–Rew commutation relations given by their Equation (2.21). □
Remark 1. Observe that the classical trajectory data (two initial position coordinates, two initial momentum coordinates) can be used to compute the values ofand vice versa.
Proposition 6. Origins of the integrals of motion
(a)is due to translational invariance of H in thedirection;
(b)is due the separability of H in a-rotated coordinate system;
(c)is due to the manifest separability of H in v and u variables.
Proof. (a) For this case, the proof is trivial because v is a cyclic variable in H.
(b) Consider the change of variables
, from which one deduces
. Then, since
, we have
. Substitution into the expressions of
H and
yields the following equation:
where
H and
are both separable in the new coordinate system obtained by a
rotation of the
coordinate system around the origin. In fact, they are the sum and difference of the one-dimensional free fall Hamiltonian in the
and
directions. Hence,
. In reference [
10], it was claimed that at the quantum level,
is due to separability in
translated parabolic coordinates. But so far, no proof has appeared in print.
(c) As H is the sum of a free-motion Hamiltonian in the v-direction, and the Hamiltonian is that of a one-dimensional free fall in the u-direction, it is then clear that , which is the difference of these two Hamiltonians, should verify . Since the total energy is conserved as , it may be simpler to write . □
Remark 2. We notice that v-parity is also a dynamical symmetry and a discrete canonical transformation since (which also implies ) leaves H as well as and all other Poisson brackets invariant.
2.5. The Free Fall Problem as a Special Limiting Case of the Kepler–Coulomb Problem
In this section, we show that the free fall problem can emerge from a special limit of the Kepler–Coulomb (KC) problem or the problem of the inverse distance potential, occurring either in gravitational or electrical interaction. This problem is known to be super-integrable and in two dimensions, and the set of its three integrals of motion makes up the components of the so-called Runge–Lenz vector [
11]. Here, we adopt the notations of [
12], with the particle mass set equal to one.
The idea is simple. The inverse distance potential arising from a source of strength is rotation-symmetric, where r is the distance from the source to the observation site. If the source recedes to infinity in the u-direction, the potential at the observation site tends toward zero. But one may compensate this potential decrease by taking a source strength , which increases with the distance. Thus, we may choose an increasing functional dependence so that in the limit of infinite source-observation separation, a linear potential appears.
In our Cartesian coordinate system, if the source of strength
is placed at the coordinate origin, the Hamiltonian of the inverse distance problem is the following:
and the three integrals of motion are given by [
12]:
where
is the angular momentum around
, which is orthogonal to the plane
.
They verify the Poisson bracket relations of the dynamical symmetry algebra
Now, if the source is no longer at the origin
O but situated on the
axis at a distance
l from the origin, the Hamiltonian and the components of the Runge–Lenz vector have new expressions
and
, which are deduced from the previous expressions in which
u is replaced by
and
by
:
and
Now, as
, the asymptotic behaviour of the inverse distance potential is as follows:
Hence, if the source strength increases as
, then the inverse distance potential reaches the limiting form of a linear potential in
u:
In this limit, the inverse distance potential problem tends toward the free fall problem up to a negative infinite constant:
and the components of the Runge–Lenz vector take the following asymptotic forms:
Theorem 1. The dynamical symmetry algebra of the free fall problemis a contraction of the dynamical symmetry algebra of the Kepler–Coulomb problem.
Proof. We now rewrite the Poisson brackets of the Kepler–Coulomb problem when the potential source is at a large distance l from the origin, and then replace the generators by their asymptotic expansions for . Therefore:
(a)
becomes
Extracting the leading order in
on the left-hand side and on the right-hand side, we obtain
, as expected; see Equation (
17).
(b)
becomes
Collecting terms of leading order in on both sides of this equation, we get since there is no term in on the right-hand side.
(c)
becomes
Equating terms of order on both sides of this equation yields precisely . Hence, we reproduce all the Poisson brackets of . □
2.6. A “Higher” Order Integral of the Motion
We now raise the question whether there exists a “higher” order integral of the motion as a construct of the dynamical symmetry algebra generators. What comes to mind is a weighted sum of squares of the
, an object similar to the square of the angular momentum in a rotation algebra. Instead of a tedious systematic search, a more astute way of finding such an integral of the motion would start by observing that the symmetry algebra of the Kepler–Coulomb problem does have a non-trivial limit when the source strength is turned off, i.e.,
. Then, the generators take the following form:
They fulfil the same Poisson bracket relations as those with
:
where
is the Hamiltonian for free particle motion in two dimensions.
Now, if we allow a linear potential to “
grow” in the
u-direction, rotational symmetry disappears. Hence,
must be discarded as a possible integral of motion in the presence of a linear potential
u. Thus, from the two remaining
, only one can survive under a modified form as a “higher” order integral of the motion, because of Poisson ’s theorem in
Section 2.1.
Let
be this hypothetical “higher” integral of the motion. As the introduced linear potential is in the
u direction, we may assume
to be of the simple form:
where
is an unknown function in phase space.
Proposition 7. There exists a second order integral of the motiongiven by the following equation:
Proof. The function
is determined by the condition
Since from explicit computation one gets the following:
it is obvious that one should require that
. □
Remark 3. The same search procedure for another quadratic integral of the motion does not work withbecause the commutativity of the Poisson bracket with H leads to impossible conditions having to be satisfied.
Proposition 8. The expression ofin terms of the generatorsis as follows: Proof. Substitute the expressions into the expression of . □
Corollary 1. The Poisson brackets ofwith the generatorsare easily obtained: Proposition 9. The infinitesimal transformation generated bywith parameter z is canonical.
Proof. The infinitesimal canonical transform generated by
with the parameter
z is given by the following equations:
We can check that the Hamiltonian remains invariant if the terms of order
are ignored:
Next, we can verify that the six canonical Poisson brackets are preserved, but the details are not presented here. Moreover, it does possess the additive abelian group property with respect to z, i.e., , as can be checked explicitly. □
2.7. Passage to Parabolic Coordinates and the Physical Meaning of
The issue here is to understand how
arises as a dynamical symmetry. The comprehensive work of Miller et al. [
13] on quantum separability has revealed that parabolic coordinates do play a central role. Following this indication, we make a passage to parabolic coordinates
from our Cartesian coordinates
, as defined by
. The Lagrangian
L then changes to a new expression:
From the following definitions of conjugate momenta:
we deduce a relation between the
and the
momenta:
This allows for the acquisition of new expressions of
H and
in parabolic coordinates:
As such, these expressions do not show any obvious x and y variable separation. The three integrals of motion also do not display any obvious separation into an x part and a y part when re-expressed in the parabolic coordinates. Next, we recall that the initial conditions fully determine the four integrals of motion . Hence, has a fixed value because it is a construct of these four integrals of motion.
Proposition 10. For, takes a constant value in the range of energy valuesof a confining quartic oscillator with an angular frequency square.
Proof. As total energy is conserved,
implies that the following relation must be verified for all
:
This means that for a given
E, the sum of the Hamiltonians of a confining quartic
x-oscillator and a non-confining quartic
y-oscillator must be equal to zero for all
. Since these one dimensional quartic oscillators are time-independent, their respective Hamiltonians have fixed values, i.e.:
however, subjected to the condition
. Note that for a given
E,
takes all real values above the minimum value of the quartic
x-oscillator Hamiltonian polynomial in phase space.
On the other hand,
may be rewritten as follows:
Since , one gets . This is due to this new aspect of separation of variables called the Stäckel separation of variables. □