ArticlePDF Available

Abstract and Figures

Time and again, non-conventional forms of Lagrangians with non-quadratic velocity dependence have received attention in the literature. For one thing, such Lagrangians have deep connections with several aspects of nonlinear dynamics including specifically the types of the Liénard class; for another, very often, the problem of their quantization opens up multiple branches of the corresponding Hamiltonians, ending up with the presence of singularities in the associated eigenfunctions. In this article, we furnish a brief review of the classical theory of such Lagrangians and the associated branched Hamiltonians, starting with the example of Liénard-type systems. We then take up other cases where the Lagrangians depend on velocity with powers greater than two while still having a tractable mathematical structure, while also describing the associated branched Hamiltonians for such systems. For various examples, we emphasize the emergence of the notion of momentum-dependent mass in the theory of branched Hamiltonians.
This content is subject to copyright.
Citation: Bagchi, B.; Ghosh, A.; Znojil,
M. A Reappraisal of Lagrangians with
Non-Quadratic Velocity Dependence
and Branched Hamiltonians.
Symmetry 2024,16, 860. https://
doi.org/10.3390/sym16070860
Academic Editor: Fernando Haas
Received: 1 June 2024
Revised: 26 June 2024
Accepted: 2 July 2024
Published: 7 July 2024
Copyright: © 2024 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
symmetry
S
S
Article
A Reappraisal of Lagrangians with Non-Quadratic Velocity
Dependence and Branched Hamiltonians
Bijan Bagchi 1, Aritra Ghosh 2and Miloslav Znojil 3,4,5,*
1Department of Mathematics, Brainware University, Kolkata 700125, West Bengal, India;
bbagchi123@gmail.com
2School of Basic Sciences, Indian Institute of Technology Bhubaneswar, Jatni 752050, Odisha, India;
ag34@iitbbs.ac.in
3The Czech Academy of Sciences, Nuclear Physics Institute, Hlavní 130, 25068 ˇ
Rež, Czech Republic
4Department of Physics, Faculty of Science, University of Hradec Králové, Rokitanského 62,
50003 Hradec Králové, Czech Republic
5Institute of System Science, Durban University of Technology, Durban 4001, South Africa
*Correspondence: znojil@ujf.cas.cz
Abstract: Time and again, non-conventional forms of Lagrangians with non-quadratic velocity
dependence have received attention in the literature. For one thing, such Lagrangians have deep
connections with several aspects of nonlinear dynamics including specifically the types of the Liénard
class; for another, very often, the problem of their quantization opens up multiple branches of
the corresponding Hamiltonians, ending up with the presence of singularities in the associated
eigenfunctions. In this article, we furnish a brief review of the classical theory of such Lagrangians
and the associated branched Hamiltonians, starting with the example of Liénard-type systems. We
then take up other cases where the Lagrangians depend on velocity with powers greater than two
while still having a tractable mathematical structure, while also describing the associated branched
Hamiltonians for such systems. For various examples, we emphasize the emergence of the notion of
momentum-dependent mass in the theory of branched Hamiltonians.
Keywords: nonstandard Lagrangians; branched Hamiltonians; Liénard systems; momentum-
dependent mass
1. Introduction
During the past few decades, the study of non-conventional types of dynamical
systems, in particular those that are controlled by Lagrangians that are not quadratic
in velocity, has entered a new phase of intense development [
1
4
]. Such Lagrangians
lead to certain exotic Hamiltonians, commonly termed as branched Hamiltonians , which
have relevance in their applicability to problems of nonlinear dynamics pertaining to
autonomous differential equations [
5
,
6
] and to certain exotic quantum mechanical models,
especially in the context of non-Hermitian parity-time (
PT
)-symmetric schemes [
7
], along
with their relativistic counterparts [8].
A simple way to see how Lagrangians that are not quadratic in velocity can lead
to meaningful dynamical systems is to consider the following toy model [
9
,
10
] (see also
Refs. [11,12]):
L(x,˙
x) = (αx+β˙
x)1, (1)
where
α
and
β
are real numbers satisfying
αβ >
0. We may also require that
αx+β˙
x=
0,
i.e., that the velocity phase space accessible to the system is defined as a subset of R2.
Notice that the Lagrangian cannot be expressed as the difference between the kinetic
and potential energies; such Lagrangians shall be referred to as nonstandard, i.e., in this
paper, we will be adopting such nomenclature in which the term ‘nonstandard Lagrangian’
Symmetry 2024,16, 860. https://doi.org/10.3390/sym16070860 https://www.mdpi.com/journal/symmetry
Symmetry 2024,16, 860 2 of 16
would refer to a Lagrangian with a non-quadratic velocity dependence. (A linear depen-
dence on velocity makes the Hessian matrix singular, resulting in a singular Legendre
transform while passing from the Lagrangian to the Hamiltonian formalism (see, for ex-
ample, Ref. [13]). We do not address such cases here and deal with Lagrangians that have
velocity dependence either in excess of quadratic powers or in inverse powers.)
A direct computation reveals that the Euler–Lagrange equation is
¨
x+γ˙
x+ω2
0x=0, (2)
where
γ=3α
2β
and
ω0=α
2β
.
(2)
is just the harmonic oscillator in the presence of linear
damping. We remind the reader that there is no time-independent Lagrangian of the ‘stan-
dard’ kind from which one can reproduce
(2)
upon invoking the Euler–Lagrange equation.
(One could recover the damped oscillator from a standard Lagrangian by using a Rayleigh
dissipation function [
14
]. Alternatively, one can consider the modified Euler–Lagrange
equations from the Herglotz variational problem to describe the damped oscillator [
15
].
We do not consider such situations here.) There exist various other families of nonstan-
dard Lagrangians (giving rise to different dynamical systems), which look quite different
from (1); each family is endowed with its own intriguing features. However, the common
theme is the existence of Lagrangians that are not quadratic in velocity, thereby leading to a
nonlinear relationship between the velocity and the momentum. It may be emphasized
that a Lagrangian
L=L(x
,
˙
x)
defined for a system whose configuration space is a subset
of Ris called regular if its Hessian with respect to the velocity is non-vanishing, i.e.,
2L(x,˙
x)
˙
x2=0, (3)
and is of constant sign, allowing us to solve for the velocity
˙
x
in favor of the momentum
p(x
,
˙
x) = L(x,˙
x)
˙
x
, i.e., we can write
˙
x(x
,
p)
. Thus, Lagrangians with a quadratic velocity
dependence are regular, and one can formulate a Hamiltonian description by means of a
Legendre transform. The condition (3) fails for Lagrangians that are linear in velocity (see,
for instance, Ref. [
13
]), but as mentioned earlier, they will not be our concern here. Instead,
we shall be looking at Lagrangians for which solving the equation
p= (x
,
˙
x)
in favor of
˙
x
leads to a non-unique solution, e.g., the appearance of a square root, which gives rise to
what will be called branching. Such Lagrangians would not permit the construction of a
Hamiltonian function in a unique way.
In the classical context, the problems associated with branched Hamiltonians and the
ones that are inevitably posed after their quantization were addressed by Shapere and
Wilczek [
1
3
]. This has triggered a series of papers by Curtright and
Zachos [1621],
which
were subsequently followed up by other works in a similar direction (see, for example,
Refs. [
5
,
22
,
23
]). It bears mentioning that local branching is not so sufficient to ensure
integrability. In particular, finding an integrable differential equation having solutions
that are not locally finitely branched with a finitely sheeted Riemann surface, but not yet
identified through Painlevé analysis, is in itself an interesting open problem [16].
Against this background, a new class of innovations on the description and simu-
lations of quantum dynamics emerged in relation to the specific role played by certain
models constructed appropriately. Not quite unexpectedly, Hamiltonians that are multi-
valued functions of momenta confront us with some typical insurmountable ambiguities
of quantization. In such cases, the underlying Lagrangian possesses time derivatives in
excess of quadratic powers (or sometimes, inverse powers). The use of these models leads,
on both classical and quantum grounds, to the necessity of a re-evaluation of the dynamical
interpretation of the momentum, which, in principle, becomes a multi-valued function of
the velocity. It also needs to be pointed out that the traditional approaches often do not
always work as is the case with certain
PT
-symmetric complex potentials possessing real
spectra [24] or upon employing tractable non-local generalizations [25].
In the context of nonlinear models, certain Liénard-class systems present an intriguing
feature of the Hamiltonian in which the roles of the position and momentum variables are
Symmetry 2024,16, 860 3 of 16
exchanged with the emergence of the notion of a momentum-dependent
mass [23,2631].
Naturally, the presence of the damping as is the case for Liénard systems poses a problem
whenever one tries to contemplate a quantization of the model. It is important to realize
that the quantization is hard to tackle in the coordinate representation of the Schrödinger
equation, but can be straightforwardly carried out in the momentum
space [26,30]
(see also
Ref. [32]).
Although much has been said about the quantum mechanical formalisms, in this
paper, we focus on the classical theory, (briefly) reviewing some aspects of nonstandard
Lagrangians and the associated branched Hamiltonians. The theory is exemplified by
focusing on various examples, which include some systems of the Liénard class, which are
of great interest in the theory of dynamical systems. Apart from Liénard systems, we discuss
some interesting toy Lagrangians, which contain time derivatives in excess of quadratic
powers, leading to branched Hamiltonians. The basic features of the theory are discussed in
light of these examples. However, we begin with a discussion on some simple nonstandard
Lagrangians, which can be figured out via some guesswork, in Section 2. Following this,
in Section 3, we discuss nonstandard Lagrangians and branched Hamiltonians in the
context of Liénard systems, wherein we outline a systematic derivation of the Lagrangians,
provided the system admits a certain integrability condition. In
Sections 4and 5,
we
analyze various intriguing examples of Lagrangians in which time derivatives occur in
excess of quadratic powers, while also discussing the associated Hamiltonians. We conclude
with some remarks in Section 6and in Appendix A, where a few further aspects of the
problem of quantization are also discussed.
2. Some Illustrative Examples
Example 1. Consider the following Lagrangian [9,10]:
L(x,˙
x) = 1
αµ(x) + β˙
x,α˙
x+βµ(x)=0, (4)
where
µ(x)
is a well-behaved function (typically a polynomial), while
α
and
β
are real-valued
and non-zero constant numbers. Obviously, it does not reveal the ‘standard’ form as the differ-
ence between the kinetic and potential energies. However, the Euler–Lagrange equation gives
¨
x+f(x)˙
x+g(x) = 0,
with
f(x) = 3αµ(x)
2β
and
g(x) = α2µ(x)µ(x)
2β2
, where, for instance, picking
µ(x) = x
gives the linearly damped harmonic oscillator, while the choice
µ(x) = x2
implies
f(x)x and g(x)x3. Lagrangians of this type (4)are termed as reciprocal Lagrangians.
Example 2. Consider another form of Lagrangians classified by [9]:
L(x,˙
x) = ln[γµ(x) + δ˙
x],γµ(x) + δ˙
x>0, (5)
where
δ
and
β
are real-valued and non-zero constant numbers. The Euler–Lagrange equation goes
as
¨
x+f(x)˙
x+g(x) =
0, with
f(x) = 2γµ(x)
δ
and
g(x) = γ2µ(x)µ(x)
δ2
. Lagrangians that look
like (5)
are termed as logarithmic Lagrangians. The relation between logarithmic and reciprocal
classes of Lagrangians has been explored in [
11
] (see also Ref. [
12
]). As with the system described by
the Lagrangian (1), the systems given by (4)and (5)are defined only on appropriate regions of R2.
Example 3. As another example, we point out that some equations that go as
¨
x+A(x,˙
x)˙
x+B(x,˙
x) = 0,
where
A(x
,
˙
x)
and
B(x
,
˙
x)
are suitable functions of
(x
,
˙
x)
can be
derived from (reciprocal) Lagrangians that read
L(x,˙
x) = 1
αµ(x) + βρ(˙
x), (6)
such that βρ′′(˙
x)[αµ(x) + βρ(˙
x)] =2β2ρ(˙
x)2. Specifically, the functions A(x,˙
x)and B(x,˙
x)are
A(x,˙
x) = 2αβρ(˙
x)µ(x)
2β2ρ(˙
x)2βρ′′(˙
x)[αµ(x) + βρ(˙
x)] , (7)
Symmetry 2024,16, 860 4 of 16
B(x,˙
x) = αµ(x)αµ(x) + βρ(˙
x)
2β2ρ(˙
x)2βρ′′(˙
x)[αµ(x) + βρ(˙
x)] . (8)
However, there is a limited variety of differential equations that can be described by Lagrangians,
which may be guessed; in general, it is often not possible to systematically derive a Lagrangian from
which a given differential equation may emerge as the Euler–Lagrange equation. In what follows,
we describe Liénard systems and demonstrate that, if a certain integrability condition is satisfied,
then one may systematically find nonstandard Lagrangians describing such systems.
3. Liénard Systems
A Liénard system is a second-order ordinary differential equation that goes as
¨
x+f(x)˙
x+g(x) = 0 (9)
(often, it is sufficient to have
f(x)
,
g(x)C2(U
,
R)
, where
UR
).
f(x)
,
g(x)C(R
,
R)
can
be suitably chosen. Interesting choices for
f(x)
and
g(x)
include
f(x) =
1 and
g(x) = x
, which
is just the damped linear oscillator, while the choice
f(x) = (
1
x2)
and
g(x) = x
gives the
van der Pol oscillator [
33
], known to admit limit-cycle behavior due to the particular choice of
f(x)
[
34
]. Another choice is
f(x) =
1 and
g(x) = x3
, for which we have the linearly damped
(nonlinear) Duffing oscillator (see, for example, Refs. [
35
,
36
]). It is noteworthy that, in any
case with
f(x)=
0, the system exhibits non-conservative dynamics because
(9)
does not stay
invariant under the transformation
t t
, namely time reversal. Furthermore, oscillatory
dynamics can be obtained if
f(x)
is an even function and if
g(x)
is odd; this follows from the
fact that the overall force (the second and third terms of
(9)
) should be odd under
x x
in
order to support oscillations [35].
3.1. Chiellini Condition and Nonstandard Lagrangians
Given a second-order differential equation, the inverse problem of finding the La-
grangian has been the subject of much investigation [
37
42
] (see also Ref. [
43
]). In particular,
for Liénard systems satisfying a certain integrability condition, one can find nonstan-
dard Lagrangians from which they emerge as the Euler–Lagrange equation [
41
] (see also
Refs. [
23
,
26
]). The idea is to make use of the so-called Jacobi last multiplier, which may be
defined as in [37] (see Appendix A).
In this manner, starting from an ordinary differential equation:
¨
x=F(x,˙
x), (10)
one defines the last multiplier Mas that which satisfies
dln M
dt +F(x,˙
x)
˙
x=0. (11)
As has been discussed in Whittaker’s classic textbook [
37
], if a second-order differential
equation such as
(10)
follows from the Euler–Lagrange equations, then the Lagrangian is
related to the latter multiplier as
M=2L(x,˙
x)
˙
x2. (12)
This allows one to determine the Lagrangian function for a given second-order differential
equation, provided it admits a Lagrangian formalism.
For the Liénard system, a formal solution for the multiplier is found to be
M(t,x) = expZf(x)dt. (13)
We may define a new nonlocal variable uas [41]
u=˙
xg(x)
f(x), (14)
Symmetry 2024,16, 860 5 of 16
with =1, which is determined from
d
dx g(x)
f(x)+(+1)f(x) = 0. (15)
Then, using Equations
(14)
and
(15)
, we have
˙
u=u f (x)
. In other words, if the
condition (15)
is true, the Liénard system (9) can be expressed as the following system of first-order equations:
˙
u=u f (x),˙
x=u+W(x), (16)
where
W(x) = 1g(x)/f(x)
. Since we have
˙
u=u f (x)
, using
(13)
, we can write
M=
u1/
, which, upon using
(12)
, gives us (up to a gauge function, which can be ignored here)
Lu2+1
. In terms of xand ˙
x, the Lagrangian reads as [23]
L(x,˙
x) = 2
(+1)(2+1)˙
x1
g(x)
f(x)2+1
, (17)
which is of the nonstandard kind. (It may be noted that one cannot at this stage set
f(x) =
0
in
(14)
(16)
or (17) to recover the conservative case. However, one can set
f(x) =
0 in
(13), which gives
M=
1, and consequently, from (12), one has
L(x,˙
x) = ˙
x2
2+J(x)˙
x+K(x),
where
K(x)
is the (conservative) scalar potential, while
J(x)
may be interpreted as a
vector potential.)
It is noteworthy that (15) represents what is known as the Chiellini condition, allowing
one to recast the Liénard system in the form (16). Specifically, if
f(x) = axα
, then (15)
dictates that g(x)must satisfy the following differential equation (see also Refs. [44,45]):
d
dx xαg(x)+(+1)a2xα=0. (18)
A simple integration gives
g(x) = kxα(+1)a2
α+1x2α+1, (19)
where
kR
is an integration constant. This gives the functional form of
g(x)
so as to
satisfy the Chiellini condition. In the cases where the Chiellini condition is satisfied, the
Lagrangian is given by (17).
3.2. Hamiltonian Aspects
With the Lagrangian that describes the Liénard system, we can move on to its Hamil-
tonian aspects. The conjugate momentum is found to be
p=L
˙
x=
+1˙
x1
g(x)
f(x)(+1)/
. (20)
Thus, the expression ˙
x=˙
x(p)may be multi-valued, depending on . It goes as
˙
x=K(l)p/(+1)+1
g(x)
f(x), (21)
where
K()
is some function of
and is a constant. Using the above form, the Hamiltonian
is found to be
H(x,p) = K()p2+1
+1g(x)
f(x)p. (22)
Notice that, if (21) admits branching, then so does the Hamiltonian (22). Below, we discuss
a concrete example.
Symmetry 2024,16, 860 6 of 16
3.3. A Concrete Example
An illustrative example will help us understand the framework discussed above.
In particular, this will also enable us to turn attention to the notion of momentum-dependent
mass [
26
28
], a concept that has generated some interest in recent times, especially within
the quantum mechanical physics-oriented model-building approaches.
We will consider the case where
f(x) = x
and
g(x) = xx3
[
23
]; (15) is satisfied for
=1, 2. For the sake of clarity, we will consider every such case separately.
3.3.1. Case with =1
In the case of =1, we have
p=p(x,˙
x) = 1
2˙
x+x212=˙
x=˙
x(x,p) = 1x2±p2p. (23)
This points towards branching. Notice that branching originates from the nonlinear depen-
dence between
p
and
˙
x
in the equation
p=L
˙
x
. The corresponding branched Hamiltonians
turn out to be
H±(x,p) = p1x2±2
3p2p, (24)
exhibiting two distinct branches, where
p
0. We plot the function
˙
x=˙
x(p
,
x)
in
Figure 1, while Figure 2shows a plot of the branched pair of Hamiltonians,
H±=H±(x,p).
The branches coalesce at p=0.
Figure 1. Plot of
˙
x±=˙
x±(x
,
p)
for the case
=
1. We have
p
0 with the two branches meeting
at p=0.
Figure 2. Plot of the branched Hamiltonian
H±=H±(x
,
p)
arising for the case
=
1, with
p
0,
and the two branches coalesce at p=0.
Symmetry 2024,16, 860 7 of 16
3.3.2. Case with =2
In the case where =2, we have
p=p(x,˙
x) = 2˙
x+1x2
21/2
=˙
x=˙
x(x,p) = p2
4(1x2)
2, (25)
implying that there is no branching, because
˙
x
can be extracted, uniquely, as a function of
the momentum. A straightforward calculation reveals that the Hamiltonian turns out to be
H(x,p) = p3
12 p(1x2)
2, (26)
wherein there is only one branch.
In Figures 3and 4, we plot
˙
x=˙
x(x
,
p)
and
H=H(x
,
p)
. An intriguing aspect of the
Hamiltonian (26) is that it may be expressed as
H(x,p) = x2
2p1+U(p),U(p) = p3
12 p
2. (27)
This resembles a standard Hamiltonian, only with the roles of the coordinate and momen-
tum being interchanged.
Figure 3. Plot of ˙
x=˙
x(x,p)for the case =2. There are no branches.
Figure 4. Plot of the Hamiltonian H=H(x,p)arising for the case =2, showing no branches.
It is then certainly tempting to interpret
m(p) = p1
as a momentum-dependent mass.
Also, the quantization of such systems proceeds, in the momentum space, often in the con-
text referring to the notion of momentum-dependent mass (see, for example, Ref. [
26
]). Still,
in our particular model, such a mass becomes singular (infinite) in the zero-momentum
limit. In the same limit, moreover, also the potential itself is vanishing, i.e., a consis-
tent physical interpretation of the system would require a suitable regularization of the
limiting process.
Marginally, let us add that one of the regularization recipes that appeared applicable
to the quantum version of our very specific model (27) has been proposed and tested in an
Symmetry 2024,16, 860 8 of 16
older paper [
46
]. Based on an ad hoc complexification of the momentum and on a certain
rather sophisticated strong-coupling perturbation expansion technique, the recipe has been
even found to provide the numerically fairly reliable spectra, in certain ranges of couplings
at least.
4. A Generalized Class of Lagrangians Yielding Branched Hamiltonians
Let us note that, if one is given a single-valued Lagrangian
L(x
,
v)
and defines it
according to formula
L(x
,
v) = x2V(v)
, rather than according to the usual recipe
L(x
,
v) = v2V(x)
and moreover, if the
p
or
v
dependence is non-convex, then, as a result
of employing the Legendre transformation, the branched functions are always encountered
despite our having started from a single-valued Lagrangian or Hamiltonian function.
4.1. The v4Model
Shapere and Wilczek have discussed a concrete model depicting the non-convex nature
of the Lagrangian, which reads [1]
L(v) = 1
4v4κ
2v2, (28)
where
v
is the velocity (now and in subsequent discussions, we will denote
˙
x=v
) and
κ>
0 is a coupling parameter. Corresponding to (28), the conjugate momentum is a cubic
function in vthat is given by
p(v) = v3κv. (29)
Clearly,
p
is not monotonic in velocity, which may lead to branching. The corresponding
Hamiltonian is obtained as
H(p) = 3
4v4κ
2v2,v=v(p), (30)
which, like L(v), is also a multi-valued function (with cusps) in the conjugate momentum
p, since each given pcorresponds to one or three values of v, as shown in (29).
For systems with a non-convex Lagrangian as sampled by (28), the routine construction
of the corresponding Hamiltonian in the conjugate momentum variable is not unique. An
analogous incertitude is encountered in cosmology models [
47
,
48
], in generalized schemes
of Einstein gravity, which involve topological invariants, and in theories of higher curvature
gravity [49].
4.2. Velocity-Independent Potentials
Curtright and Zachos [
20
] extended the analysis of [
1
] by considering a generalized
class of non-quadratic Lagrangians that go as
L(x,v) = C(v1)2k1
2k+1V(x),C=2k+1
2k11
42
2k+1, (31)
where the traditional kinetic energy term is replaced by a fractional function of the velocity
variable
v
and
V(x)
represents a convenient local interaction potential. The fractional
powers facilitate the derivation of supersymmetric partner forms of the potential á la
Witten [
50
]. We remark that the
(
2
k+
1
)
st root of the first term in
L(x
,
v)
is required to be
real, and >0 or <0 for v>1 or v<1, respectively.
Let us focus on the case with
k=
1. Performing a Taylor expansion for
v
near zero,
we can write
L(x
,
v)C(
1
+v
3+v2
9+O(v3)) V(x)
. While the first term is merely a
constant and the second term contributes to the boundary of the action and, therefore, does
not influence the equations of motion, the third term yields the kinetic structure:
A=Zt2
t1
L(x,v)dt Ct2t1+1
3(x(t2)x(t1)) + 1
9Zt2
t1
v2dt +Zt2
t1
O(v3)dtZt2
t1
V(x)dt. (32)
Symmetry 2024,16, 860 9 of 16
Thus, for small velocities, the action results in the usual Newtonian form of the equations
of motion.
For the large velocities, on the other hand, we have a less trivial scenario, which leads
(for finite, positive-integer values of
k
) to a non-convex function of
v
. The curvature term
corresponding to the quantity
2L
v2
changes sign at the point
v=
1. Thus,
L(x
,
v)
may be
interpreted as a single pair of convex functions that have been judiciously pieced together.
Now, from the Lagrangian (31), the canonical momentum can be calculated:
p=p(v) = 1
42
2k+11
(v1)2
2k+1
. (33)
Inverting the relation, we observe that the velocity variable
v(p)
emerges as a double-
valued function of p:
v=v±(p) = 11
41
p(2k+1)
. (34)
Corresponding to the two signs above, a pair of branches of the Hamiltonian, namely
H±(x
,
p)
, will appear. Specifically, for any positive-integer value of
k
, these may be identi-
fied to be
H±(x,p) = p±1
4k21
p2k1
+V(x). (35)
From a classical perspective, in order to avoid an imaginary
v(p)
, one needs to address a
non-negative
p
. This in turn implies that the slope
L
v
is always positive. It is interesting
to note that, for the
k=
1 case, we are led to the quantum mechanical supersymmetric
structure for the difference
H±(x
,
p)V(x)
, which reads
p±1
2p
, in the momentum space.
The associated spectral properties have been analyzed in the literature [26,30].
We end our discussion on this example by noting that, in the special case where
V(x) = x2
, the branched Hamiltonian is
H±(x
,
p) = x2+U±(p)
, wherein it appears as if
the roles of the coordinate and the momentum have been interchanged, with
U±(p)
being
a momentum-dependent potential that exhibits two branches.
4.3. Velocity-Dependent Potentials
Lines of force can be ascertained with the help of velocity-dependent potentials, which
ensure that particles take certain specified paths [
14
,
51
]. In electrodynamics, the field
vectors
E
and
B
can be determined given such a potential function when the trajectories of
a charged particle’s motion are specified. In the present context, we proceed to set up an
extended scheme where the Lagrangian depends on a velocity-dependent potential
V(x
,
v)
in the manner as given by [22,29]:
L(x,v) = C(v1)2k1
2k+1V(x,v),C=2k+1
2k11
42
2k+1, (36)
where
V(x
,
v)
is assumed to be given in a separable form, i.e.,
V(x
,
v) = U(v) + V(x)
; here,
U(v)
and
V(x)
are well-behaved functions of
v
and
x
, respectively. Using the standard
definition of the canonical momentum, we find its form to be
p=p(v) = 1
42
2k+1(v1)2
2k+1U(v). (37)
The complexity of the right side does not facilitate an easy inversion of the above relation
that would reveal the multi-valued nature of velocity in a closed, tractable form. Neverthe-
less, the associated branches of the Hamiltonian can be straightforwardly written down
upon employing the Legendre transform as
H±(x,p) = p±1
4[p+U(v)]2k1
22k+1
2k1p[p+U(v)]1+V(x,v),v=v(p). (38)
Symmetry 2024,16, 860 10 of 16
Unfortunately, since a Hamiltonian has to be a function of the coordinate and its corre-
sponding canonical momentum, the generality of the form of
H±(x
,
p)
as derived above
is of little use unless we have an explicit inversion of (37) giving
v=v(p)
. We, therefore,
have to go for the specific cases of kand U(v).
A Special Case
Indeed, the case
k=
1 proves to be particularly worthwhile to understand the spectral
properties of the Hamiltonian. It corresponds to the Lagrangian as given by
L(x,v) = 31
42
3(v1)1
3U(v)V(x). (39)
A sample choice for U(v)could be [22]
U(v) = λv+3δ(v1)1
3, (40)
in which
λ(
0
)
and
δ(<
4
2
3)
are suitable real constants. The presence of the parameter
δ
scales the kinetic energy term in the Lagrangian. The canonical momentum
p
is now
given by
p=p(v) = µ(v1)2
3λ, (41)
where the quantity
µ=
4
2
3δ>
0. We are, therefore, led to a pair of relations for the
velocity depending on p:
v=v±(p) = 1µ3
2(p+λ)3
2. (42)
As a consequence, we find two branches of the Hamiltonian, which are expressible as
H±(x,p) = (p+λ)±2γ
pp+λ+V(x), (43)
where µ3/2 has been replaced by γ. As a final comment, the special case corresponding to
λ=0 and γ=1/4 conforms to the Hamiltonian (35) advanced in [20].
5. Three More Forms of Hamiltonians
5.1. Higher Power Lagrangians
As an extension of (31), the following higher power Lagrangian was proposed in [
5
]:
L(x,v) = C(v+σ(x)) 2m+1
2m1δ,Λ=12m
1+2m(δ)2
12m,δ>0, (44)
where we notice that the coefficient
Λ
is non-negative for 0
m<1
2
. The main difference
from (31) is in the choice of a general function
σ(x)
in place of
σ(x) =
1 as in (31).
The other point is that the inverse exponent with respect to the model of Curtright and
Zachos [
20
] has been taken for the convenience of calculus. We have omitted the explicit
potential function assuming that the interaction re-appears in a more natural manner via
a suitable choice of an auxiliary free parameter
δ
and that of a nontrivial function
σ(x)
.
As long as our Lagrangian
L(x
,
v)
is of a nonstandard type, we will not feel disturbed by
the absence of the explicit potential V(x).
For this particular model, the canonical momentum reads as
p=p(x,v) = (δ)2
12m(v+σ(x)) 2
2m1, (45)
and a simple inversion yields
v=v±(x,p) = σ(x) + δ±pp2m1. (46)
This means the Hamiltonian is obtained to be
Symmetry 2024,16, 860 11 of 16
H±(x,p) = (p)σ(x)2δ
2m+1±pp2m+1+δ. (47)
Special Case
The specific case with
m=
0 is of interest as it allows us to easily derive the (double-
valued) velocity profile, which reads as
v=v±(x,p) = σ(x)±δ
p, (48)
implying that the Hamiltonian branches out into components:
H±(x,p) = (p)σ(x)2δpp+δ. (49)
The nature of the two Hamiltonians depends on the sign of
p
. Once we specify the following
choice of σ(x), namely
σ(x) = λ
2x2+9λ2
2k2,λ>0, (50)
together with the choice
δ=9λ2
2k2
, then upon imposing a simple translation
p2k
3λp
1,
the Hamiltonians H±acquire the forms that go as
H±(x,p) = 9λ2
2k2"2212kp
3λ1
2+k2x2
9λ2kp
3λ2k3x2p
27λ2#. (51)
These are readily identifiable as a set of plausible Hamiltonians representing a nonlinear
Liénard system [
23
,
26
,
28
]. The appearance of the coordinate–momentum coupling is
noteworthy and leads us to the notion of a momentum-dependent mass as
H±(x,p) = x2
2m(p)+U±(p),m(p) = λ2kp
31
,U±(p) = 9λ2
2k2"2212kp
3λ1
22kp
3λ#. (52)
From a classical perspective, the momentum
p
needs to be restricted to the range
<
p3λ
2k
to account for the physical properties of the system in the real space; this also
ensures that the momentum-dependent mass is positive and finite. However, because of a
branch-point singularity at
p=3λ
2k
, a thorough analytical study of
H±(x
,
p)
becomes greatly
involved. Observe that, when
p=3λ
2k
, we find the coincidence of the two Hamiltonians
H±(x,p).
5.2. Rational Function Lagrangians
In another characteristic example, let us pick up an illustration where
L(x
,
v)
is of the
reciprocal kind and is defined to be [5]
L(x,v) = 1
s1
3sx2+3
sλv1
, (53)
where sis a real parameter. The canonical momentum comes out as
p=p(x,v) = 1
s1
3sx2+3
sλv2
, (54)
which, when inverted, yields
v=v±(x,p) = 1
3sx2+3
sλ±1
sp . (55)
The accompanying Hamiltonian corresponding to the above Lagrangian has two branches:
H±(x,p) = s
3x2p+3
sλp±2rp
s. (56)
Symmetry 2024,16, 860 12 of 16
It should be remarked that, as
λ
0, (53) is just the trial Lagrangian (6), for the choice
µ(x) = ax2
and under a suitable identification of the constant parameters. We end by
noting that (56) can be re-expressed as a model:
H±(x,p) = x2
2m(p)+U±(p),m(p) = 3
2s p ,U±(p) = 3
sλp±2rp
s. (57)
with a momentum-dependent mass.
5.3. Relativistic Free Particle
As a final example of the dynamics due to non-quadratic Lagrangians, let us re-examine
the much-studied problem of a relativistic (free) particle, which is described by the Lagrangian:
L(v) = mcpc2v2,v<c. (58)
The conjugate momentum is obtained as
p=p(v) = L(v)
v=mcv
c2v2. (59)
This implies that p2c2p2v2m2c2v2=0. Solving for the velocity gives
v=v±(p) = ±pc
pp2+m2c2. (60)
Thus, the Hamiltonian reads
H±(p) = c(±p2m2c2)
pp2+m2c2. (61)
In particular, one may consider the ultrarelativistic limit (
vc
) in which
H±(p) = ±pc
.
Related to the above example, the reader is referred to Ref. [
52
] for the example of a spinning
particle with the Lagrangian being non-quadratic both in the position and spin variables.
Another physically interesting example is that of an axially symmetric charged body in an
electromagnetic field, which is governed by Euler–Poisson equations [53].
6. Concluding Remarks
In the present treatment on the existence of nonstandard Lagrangians, we empha-
sized, first of all, the existence of certain unusual aspects of their relationship with the
associated branched Hamiltonians. Various different examples were discussed; in all of
them, the velocity dependence of the Lagrangian was not of (homogenous) degree two, but
contained either powers larger than two or negative powers. This resulted in a nonlinear
relationship between the generalized velocity and the conjugate momentum, leading to a
multi-valued behavior of the velocity when solved as a function of the momentum (and,
perhaps, the coordinate).
We observed that, in the description of Hamiltonians emerging from nonstandard
Lagrangians, the notion of momentum-dependent mass is often encountered. It is then as if
the coordinate of the particle played the role of momentum and vice versa, with a function
of the momentum variable appearing as an ‘effective mass’ describing the system. Such
systems can be quantized straightforwardly in the momentum space [
26
,
30
,
32
]. Naturally,
this reopens a few mathematically deeper questions concerning their quantization. Indeed,
the technicalities of canonical quantization can be perceived as widely assessed in the
literature (see, for example, Ref. [
54
]), wherein it is not infrequent to encounter certain
fundamental difficulties. For example, in certain ‘anomalous’ quantum systems with non-
Hermitian Hamiltonians supporting real eigenvalues, it has been shown that the quantum
wave functions themselves could still, in finite time, diverge [
55
]. Moreover, after one
admits the unusual forms of the Hamiltonians characterized, typically, by the popular
parity-time-symmetry (PT -symmetry; see, for example, Refs. [56,57] for a pedagogic and
Symmetry 2024,16, 860 13 of 16
introductory discussion on such specific variants of non-self-adjoint models), the anomalies
may occur even when the PT -symmetry itself remains unbroken.
Several unusual forms of the latter anomalies may appear in both the spectra and
eigenfunctions, materialized as Kato’s exceptional points [
58
,
59
] or the so-called spectral
singularities [
60
]. In particular, exceptional points can be regarded as a typical feature of
non-Hermitian systems related to a branch-point singularity where two or more discrete
eigenvalues, real or complex, and corresponding to two different quantum states, along
with their accompanying eigenfunctions, coalesce [6163].
Naturally, the possible relevance of the latter anomalies in the quantum systems
controlled by the branched Hamiltonians is more than obvious. One only has to emphasize
the difference between the systems characterized by the unitary and non-unitary evolution.
In the former case, indeed, one is mainly interested in the description of the systems of
stable bound states. In the latter setting, the scope of the theory is broader; the states are
resonant and unstable in general. In the related models, one deals with Hamiltonians
that are manifestly non-Hermitian and that undergo non-unitary quantum evolution;
they generally represent open systems with balanced gain and loss [
64
,
65
]. Exceptional
points occur there as experimentally measurable phenomena. In this connection, it is
also relevant to point out the occurrence of certain theoretical anomalies like the possible
breakdown of the adiabatic theorem [
66
] or the feature of stability-loss delay [
67
], etc. In all
of these contexts, one encounters the possibility of interpreting branched Hamiltonians as
an innovative theoretical tool admitting a coalescence of the branched pairs of operators
at an exceptional point. Thus, preliminarily, let us conclude that the (related) possible
innovative paths towards quantization look truly promising.
Author Contributions: Conceptualization, B.B. and M.Z.; methodology, B.B. and A.G. and M.Z.;
software, A.G.; validation, M.Z.; formal analysis, B.B.; investigation, A.G. and M.Z.; resources,
B.B. and M.Z.; data curation, A.G.; writing and original draft preparation, B.B. and A.G. and M.Z.;
review and editing, B.B. and A.G. and M.Z.; visualization, A.G.; supervision, B.B. and M.Z.; project
administration, B.B.; funding acquisition, A.G. All authors have read and agreed to the published
version of the manuscript.
Funding: This research was funded by the Ministry of Education (MoE), Government of India, by the
Prime Minister’s Research Fellowship grant number 1200454, and from the budget of the Brainware
University and of the University of Hradec Kralove.
Data Availability Statement: Data are contained within the article.
Acknowledgments: We thank Anindya Ghose Choudhury for discussions and for his interest in
this work. B.B. thanks Brainware University for infrastructural support. A.G. thanks the Ministry of
Education (MoE), the Government of India, for financial support in the form of a Prime Minister’s
Research Fellowship (ID: 1200454). M.Z. is financially supported by the Faculty of Science of UHK.
Conflicts of Interest: The authors declare no conflict of interest.
Appendix A. Jacobi Last Multiplier
The Jacobi last multiplier [
37
] is a very useful tool in the mathematics of dynamical
systems. On the one hand, given a dynamical system on an
m
-dimensional phase space
with
(m
2
)
linearly independent and known first integrals, it allows one to determine
the ‘last’, i.e.,
(m
1
)
th independent first integral; on the other hand, given a second-order
(ordinary) differential equation, it facilitates the computation of a suitable Lagrangian
function that describes the dynamics via the Euler–Lagrange equation [
37
40
]. Thus,
the last multiplier is intimately related to the integrability properties of a dynamical system.
Naively, one can define it as follows. Given a vector field
X
on the phase space, the Jacobi
last multiplier
M
is a factor such that
MX
has zero divergence. For Hamiltonian systems
for which Liouville’s theorem holds, i.e.,
div ·X=
0, the last multiplier is just a constant
number as
X
is already divergence-free. On the other hand,
M
assumes a more non-trivial
Symmetry 2024,16, 860 14 of 16
form if
X
has non-vanishing divergence, i.e., if
div ·X=
0, say, for the Liénard equation,
which is dissipative and, hence, is not volume-conserving (see, for example, Refs. [44,45]).
Consider a vector field
X
; in local coordinates
{xj}
, where
j=
1, 2,
. . .
,
m
, one can write
its components as
˙
xj=Xj({xj})
, which define the dynamical system as a system of first-
order equations. Further, let there be a certain number of first integrals
(F1
,
F2
,
. . .
,
Fk)
, where
k<m
. For any open subset
Rm
, we define the Jacobi last multiplier to be a function
M:RmR, which is non-negative and defines an invariant measure RMdmx, i.e.,
ZMdx1dx2. . . dxm=Zϕt()M(x1,x2, . . . , xk,xk+1, . . . , xm)
(F1,F2, . . . , Fk,xk+1, . . . , xm)dF1dF2. . . dFkdxk+1. . . dxm, (A1)
where
ϕt()
is the transformation of the region
under the flow of
X
. Notice that it
is necessary that
(F1
,
F2
,
. . .
,
Fk)
be independent, ensuring that
dF1dF2. . . dFk=
0.
The above-mentioned invariance condition leads to the equation:
d
dt M(x1,x2, . . . , xk,xk+1, . . . , xm)
(F1,F2, . . . , Fk,xk+1, . . . , xm)!=0, (A2)
which, upon employing the chain rule, gives
dM
dt
(x1,x2, . . . , xk,xk+1, . . . , xm)
(F1,F2, . . . , Fk,xk+1, . . . , xm)+M
m
j=1
Xj
xj
(x1,x2, . . . , xk,xk+1, . . . , xm)
(F1,F2, . . . , Fk,xk+1, . . . , xm)=0. (A3)
This is just equivalent to
d
dt ln M+
m
j=1
Xj
xj
=0, (A4)
which coincides with (11) as appropriate for the system (10). Notice that
m
j=1
Xj
xj
is just the
divergence of
X
, and therefore,
M=
1 or some constant if the vector field has zero divergence.
Now, having defined the Jacobi last multiplier, let us demonstrate its use in deriving
Lagrangians for a second-order ordinary differential equation with a two-dimensional
phase space. Consider the system (10). If it is derivable from the Euler–Lagrange equation:
d
dt L(x,˙
x)
˙
x=L(x,˙
x)
x, (A5)
or 2L(x,˙
x)
˙
x2¨
x+2L(x,˙
x)
˙
xx˙
xL(x,˙
x)
x=0, (A6)
then, from (10), one should have
2L(x,˙
x)
˙
x2F(x,˙
x) + 2L(x,˙
x)
˙
xx˙
xL(x,˙
x)
x=0. (A7)
Differentiating both sides with respect to ˙
xgives
˙
x2L(x,˙
x)
˙
x2F(x,˙
x)+3L(x,˙
x)
˙
x2x˙
x=0. (A8)
Defining Σ(x,˙
x) = 2L(x,˙
x)
˙
x2, (A8) gives
˙
x(Σ(x,˙
x)F(x,˙
x)) +
x(Σ(x,˙
x)˙
x) = 0. (A9)
But, this is just Equation (A4) of the last multiplier (if
M
has no explicit time dependence),
i.e., we should identify
Σ(x
,
˙
x) = M(x
,
˙
x)
, and this readily gives (12). See Refs. [
37
40
] for
more details.
Symmetry 2024,16, 860 15 of 16
References
1. Shapere, A.; Wilczek, F. Branched Quantization. Phys. Rev. Lett. 2012,109, 200402. [CrossRef]
2. Shapere, A.; Wilczek, F. Classical Time Crystals. Phys. Rev. Lett. 2012,109, 160402. [CrossRef]
3. Wilczek, F. Quantum Time Crystals. Phys. Rev. Lett. 2012,109, 160401. [CrossRef]
4.
Henneaux, M.; Teitelboim, C.; Zanelli, J. Quantum mechanics for multivalued Hamiltonians. Phys. Rev. A 1987,36, 4417.
[CrossRef]
5.
Bagchi, B.; Modak, S.; Panigrahi, P.K.; Ruzicka, F.; Znojil, M. Exploring branched Hamiltonians for a class of nonlinear systems.
Mod. Phys. Lett. A 2015,30, 1550213. [CrossRef]
6.
Mitsopoulos, A.; Tsamparlis, M. Cubic first integrals of autonomous dynamical systems in E2 by an algorithmic approach.
J. Math. Phys. 2023,64, 012701. [CrossRef]
7.
Bender, C.M.; Dorey, P.E.; Dunning, C.; Fring, A.; Hook, D.W.; Jones, H.F.; Kuzhel, S.; Lévai, G.; Tateo, R. PT Symmetry: In Quantum
and Classical Physics ; World Scientific: Hackensack, NJ, USA, 2019.
8.
Mandal, B.P.; Mourya, B.K.; Ali, K.; Ghatak, A. PT phase transition in a (2 + 1)-d relativistic system. Ann. Phys. 2015,363, 185–193.
[CrossRef]
9. Saha, A.; Talukdar, B. On the non-standard Lagrangian equations. arXiv 2013, arXiv:1301.2667.
10.
Cariñena, J.F.C.;Rañada, M.F.R.; Santander, M. Lagrangian formalism for nonlinear second-order Riccati systems: One-
dimensional integrability and two-dimensional superintegrability. J. Math. Phys. 2005,46, 062703. [CrossRef]
11.
Cariñena, J.F.C.; Nunez, J.F. Geometric approach to dynamics obtained by deformation of Lagrangians. Nonlinear Dyn. 2016,
83, 457–461. [CrossRef]
12.
Cariñena, J.F.C.; Guha, P.; Rañada, M.F.R. A geometric approach to higher-order Riccati chain: Darboux polynomials and constants
of the motion. J. Phys. Conf. Ser. 2009,175, 012009. [CrossRef]
13. Dirac, P.A.M. Generalized Hamiltonian dynamics. Proc. R. Soc. Lond. A 1958,246, 326–332. [CrossRef]
14. Goldstein, H.; Poole, C.; Safko, J. Classical Mechanics, 3rd ed.; Addison-Wesley: Glenview, IL, USA, 2001.
15. de León, M.; Laínz, M. A review on contact Hamiltonian and Lagrangian systems. Rev. Acad. Canar. Cienc. 2019,31, 1.
16. Curtright, T.; Zachos, C. Evolution profiles and functional equations. J. Phys. A Math. Theor. 2009,42, 485208. [CrossRef]
17.
Curtright, T.L.; Zachos, C.K. Chaotic maps, Hamiltonian flows and holographic methods. J. Phys. A Math. Theor. 2010,43, 445101.
[CrossRef]
18. Curtright, T.; Veitia, A. Logistic map potentials. Phys. Lett. A 2011,375, 276–282. [CrossRef]
19. Curtright, T. Potentials Unbounded Below. SIGMA 2011,7, 042. [CrossRef]
20. Curtright, T.L.; Zachos, C.K. Branched Hamiltonians and supersymmetry. J. Phys. A Math. Theor. 2014,47, 145201. [CrossRef]
21. Curtright, T. The BASICs of Branched Hamiltonians. Bulg. J. Phys. 2018,45, 102–113.
22.
Bagchi, B.; Kamil, S.M.; Tummuru, T.R.; Semorádová, I.; Znojil, M. Branched Hamiltonians for a Class of Velocity Dependent
Potentials. J. Phys. Conf. Ser. 2017,839, 012011. [CrossRef]
23.
Choudhury, A.G.; Guha, P. Branched Hamiltonians and time translation symmetry breaking in equations of the Liénard type.
Mod. Phys. Lett. A 2019,34, 1950263. [CrossRef]
24.
Bagarello, F.; Gazeau, J.P.; Szafraniec, F.H.; Znojil, M. (Eds.) Non-Selfadjoint Operators in Quantum Physics: Mathematical Aspects;
John Wiley & Sons: Hoboken, NJ, USA, 2015.
25.
Znojil, M.
PT
-symmetric model with an interplay between kinematical and dynamical non-localities. J. Phys. A Math. Theor.
2015,48, 195303. [CrossRef]
26.
Bagchi, B.; Choudhury, A.G.; Guha, P. On quantized Liénard oscillator and momentum dependent mass. J. Math. Phys. 2015,
56, 012105. [CrossRef]
27.
Bagchi, B.; Ghosh, R.; Goswami, P. Generalized Uncertainty Principle and Momentum-Dependent Effective Mass Schrödinger
Equation. J. Phys. Conf. Ser. 2020,1540, 012004. [CrossRef]
28.
Chandrasekar, V.K.; Senthilvelan, M.; Lakshmanan, M. Unusual Liénard-type nonlinear oscillator. Phys. Rev. E 2005,72, 066203.
[CrossRef]
29.
Bagchi, B.; Ghosh, D.; Tummuru, T.R. Branched Hamiltonians for a quadratic type Liénard oscillator. J. Nonlinear Evol. Equ. Appl.
2020,2018, 101–106.
30.
Ruby, V.C.; Senthilvelan, M.; Lakshmanan, M. Exact quantization of a PT-symmetric (reversible) Liénard-type nonlinear oscillator.
J. Phys. A Math. Theor. 2012,45, 382002. [CrossRef]
31. Bagchi, B.; Ghosh, D.; Modak, S.; Panigrahi, P.K. Nonstandard Lagrangians and branching: The case of some nonlinear Liénard
systems. Mod. Phys. Lett. A 2019,34, 1950110. [CrossRef]
32. von Roos, O. Position-dependent effective masses in semiconductor theory. Phys. Rev. B 1983,27, 7547. [CrossRef]
33. van der Pol, B. LXXXVIII. On “relaxation-oscillations”. Philos. Mag. 1926,2, 978–992. [CrossRef]
34.
Strogatz, S.H. Nonlinear Dynamics and Chaos: With Applications to Physics, Biology, Chemistry, and Engineering, 2nd ed.; CRC Press:
Boca Raton, FL, USA, 2014.
35. Mickens, R.E. Truly Nonlinear Oscillations; World Scientific: Hackensack, NJ, USA, 2009.
36. Demina, M.V. Liouvillian integrability of the generalized Duffing oscillators. Anal. Math. Phys. 2021,11, 25. [CrossRef]
37. Whittaker, E.T. A Treatise on the Analytical Dynamics of Particles and Rigid Bodies; Cambridge University Press: Cambridge, UK, 1988.
38.
Yan, C.C. Construction of Lagrangians and Hamiltonians from the equation of motion. Am. J. Phys. 1978,46, 671–675. [CrossRef]
Symmetry 2024,16, 860 16 of 16
39. Nucci, M.C.; Leach, P.G.L. The Jacobi Last Multiplier and its applications in mechanics. Phys. Scr. 2008,78, 065011. [CrossRef]
40. Nucci, M.C.; Leach, P.G.L. An Old Method of Jacobi to Find Lagrangians. J. Nonlinear Math. Phys. 2009,16, 431–441. [CrossRef]
41.
Nucci, M.C.; Tamizhmani, K.M. Lagrangians for Dissipative Nonlinear Oscillators: The Method of Jacobi Last Multiplier.
J. Nonlinear Math. Phys. 2010,17, 167–178. [CrossRef]
42.
Mitra, S.; Ghose-Choudhury, A.; Poddar, S.; Garai, S.; Guha, P. The Jacobi Last Multiplier, Lagrangian and Hamiltonian for
Levinson–Smith type equations. Phys. Scr. 2024,99, 015237. [CrossRef]
43.
Cariñena, J.F.; Fernández–Núñez, J. Jacobi Multipliers in Integrability and the Inverse Problem of Mechanics. Symmetry 2021,
13,1413. [CrossRef]
44.
Cariñena, J.F.; Guha, P. Non-standard Hamiltonian structures of Liénard equation and contact geometry. Int. J. Geom. Methods
Mod. Phys. 2019,16, 1940001. [CrossRef]
45.
Cariñena, J.F.; Guha, P. Geometry of non-standard Hamiltonian structures of Liénard equations and contact structure. Int. J. Geom.
Methods Mod. Phys. 2024,21, 2440005. [CrossRef]
46.
Fernández, F.M.; Guardiola, R.; Ros, J.; Znojil, M. Strong-coupling expansions for the PT-symmetric oscillators
V(x) = aix +b(i x)2+c(ix)3.J. Phys. A: Math. Gen. 1998,31, 10105. [CrossRef]
47. Armendáriz-Picón, C.; Damour, T.; Mukhanov, V. k-Inflation. Phys. Lett. B 1999,458, 209–218. [CrossRef]
48.
Arkani-Hamed, N.; Cheng, H.-C.; Luty, M.; Mukohyama, S. Ghost condensation and a consistent infrared modification of gravity.
J. High Energy Phys. 2004,0405, 074. [CrossRef]
49.
Teitelboim, C.; Zanelli, J. Dimensionally continued topological gravitation theory in Hamiltonian form. Class. Quantum Gravity
1987,4, L125. [CrossRef]
50. Witten, E. Dynamical breaking of supersymmetry. Nucl. Phys. B 1981,188, 513–554. [CrossRef]
51. Coffman, M.L. Velocity-Dependent Potentials for Particles Moving in Given Orbits. Am. J. Phys. 1952,20, 195–199. [CrossRef]
52.
Deriglazov, A.A.; Ramírez, W.G. Recent Progress on the Description of Relativistic Spin: Vector Model of Spinning Particle and
Rotating Body with Gravimagnetic Moment in General Relativity. Adv. Math. Phys. 2017,2017, 7397159. [CrossRef]
53.
Deriglazov, A.A. Rotation Matrix of a Charged Symmetrical Body: One-Parameter Family of Solutions in Elementary Functions.
Universe 2024,10, 250. [CrossRef]
54. Klauder, J.R. Valid Quantization: The Next Step. J. High Energy Phys. Gravit. Cosmol. 2022,8, 628–634. [CrossRef]
55.
Graefe, E.-M.; Korsch, H.J.; Rush, A.; Schubert, R. Classical and quantum dynamics in the (non-Hermitian) Swanson oscillator.
J. Phys. A Math. Theor. 2015,48, 055301. [CrossRef]
56.
Novák, R. On the Pseudospectrum of the Harmonic Oscillator with Imaginary Cubic Potential. Int. J. Theor. Phys. 2015,54,
4142–4153. [CrossRef]
57.
R ˚užiˇcka, F. Hilbert Space Inner Products for
PT
-symmetric Su-Schrieffer-Heeger Models. Int. J. Theor. Phys. 2015,54, 4154–4163.
[CrossRef]
58. Kato, T. Perturbation Theory for Linear Operators: Classics in Mathematics; Springer: Berlin/Heidelberg, Germany, 1995.
59. Heiss, W.D. The physics of exceptional points. J. Phys. A Math. Theor. 2012,45, 444016. [CrossRef]
60.
Correa, F.; Plyushchay, M.S. Spectral singularities in
PT
-symmetric periodic finite-gap systems. Phys. Rev. D 2012,86, 085028.
[CrossRef]
61.
Znojil, M. Exceptional points and domains of unitarity for a class of strongly non-Hermitian real-matrix Hamiltonians. J. Math.
Phys. 2021,62, 052103. [CrossRef]
62.
Bagarello, F.; Gargano, F. Model pseudofermionic systems: Connections with exceptional points. Phys. Rev. A 2014,89, 032113.
[CrossRef]
63. Bagchi, B.; Ghosh, R.; Sen, S. Exceptional point in a coupled Swanson system. EPL 2022,137, 50004. [CrossRef]
64. Moiseyev, N. Non-Hermitian Quantum Mechanics; Cambridge University Press: Cambridge, UK, 2011.
65.
Rotter, I. A non-Hermitian Hamilton operator and the physics of open quantum systems. J. Phys. A Math. Theor. 2019,42, 153001.
[CrossRef]
66.
Hashimoto, K.; Kanki, K.; Hayakawa, H.; Petrosky, T. Non-divergent representation of a non-Hermitian operator near the
exceptional point with application to a quantum Lorentz gas. Prog. Theor. Exp. Phys. 2015,2015, 023A02. [CrossRef]
67.
Milburn, T.J.; Doppler, J.; Holmes, C.A.; Portolan, S.; Rotter, S.; Rabl, P. General description of quasiadiabatic dynamical
phenomena near exceptional points. Phys. Rev. A 2015,92, 052124. [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
... The observed correspondence between isochronous systems and equally-spaced quantum spectra [8] suggests that the Liénard system (1.2) may admit a harmonic-oscillator-like spectrum [9] (see also, [10,11]). The important point here is that due to the nonstandard form of the Hamiltonian function, the 'quantum mechanics' is worked out in the momentum representation in which one also encounters the notion of a momentum-dependent mass [9,10,11,12]. ...
... Then, we will demonstrate the existence of a bi-Hamiltonian structure in the sense that there are two distinct classes of Hamiltonians which can describe the system consistently. Such Hamiltonians would turn out to be of the nonstandard kind in which there is no direct identification of kinetic and potential energies, although they admit functional forms in which the roles of the coordinate and the momentum appear interchanged with the appearance of momentum-dependent masses and potentials [9,12]. For the derivation of such aspects, we will follow the preceding works [15,16,17,18,19,20] which have involved the use of the Jacobi last multiplier to derive Lagrangians (and hence, Hamiltonians) for Liénard-type systems. ...
... In this section, we will make use of the Jacobi last multiplier to derive Lagrangian (and subsequently, Hamiltonian) functions appropriate for the Liénard system (1.2). The reader is referred to [12,15,16,17,18,19,20] for the technical details. ...
Preprint
In this paper, we explore some classical and quantum aspects of the nonlinear Li\'enard equation x¨+kxx˙+ω2x+(k2/9)x3=0\ddot{x} + k x \dot{x} + \omega^2 x + (k^2/9) x^3 = 0, where x=x(t) is a real variable and k,ωRk, \omega \in \mathbb{R}. We demonstrate that such an equation could be derived from an equation of the Levinson-Smith kind which is of the form z¨+J(z)z˙2+F(z)z˙+G(z)=0\ddot{z} + J(z) \dot{z}^2 + F(z) \dot{z} + G(z) = 0, where z=z(t) is a real variable and {J(z),F(z),G(z)}\{J(z), F(z), G(z)\} are suitable functions to be specified. It can further be mapped to the harmonic oscillator by making use of a nonlocal transformation, establishing its isochronicity. Computations employing the Jacobi last multiplier reveal that the system exhibits a bi-Hamiltonian character, i.e., there are two distinct types of Hamiltonians describing the system. For each of these, we perform a canonical quantization in the momentum representation and explore the possibility of bound states. While one of the Hamiltonians is seen to exhibit an equispaced spectrum with an infinite tower of states, the other one exhibits branching but can be solved exactly for certain choices of the parameters.
Preprint
Full-text available
Euler-Poisson equations of a charged symmetrical body in external constant and homogeneous electric and magnetic fields are deduced starting from the variational problem, where the body is considered as a system of charged point particles subject to holonomic constraints. The final equations are written for the center-of mass-coordinate, rotation matrix and angular velocity. General solution to the equations of motion is obtained for the case of a charged ball. For the case of a symmetrical charged body (solenoid), the task of obtaining the general solution is reduced to the problem of a one-dimensional cubic pseudo-oscillator. Besides, we present a one-parametric family of solutions to the problem in elementary functions.
Article
Full-text available
Euler–Poisson equations of a charged symmetrical body in external constant and homogeneous electric and magnetic fields are deduced starting from the variational problem, where the body is considered as a system of charged point particles subject to holonomic constraints. The final equations are written for the center-of-mass coordinate, rotation matrix and angular velocity. A general solution to the equations of motion is obtained for the case of a charged ball. For the case of a symmetrical charged body (solenoid), the task of obtaining the general solution is reduced to the problem of a one-dimensional cubic pseudo-oscillator. In addition, we present a one-parametric family of solutions to the problem in elementary functions.
Article
Full-text available
We derive the Jacobi last multiplier for second-order ordinary differential equations of the Levinson--Smith type by using a combination of previous techniques employed for the Lieˊ\acute{e}nard-I and II classes of equations. This opens up the possibility for a Lagrangian or Hamiltonian description of the systems governed by the Levinson--Smith type of equations as well as simplifying the problem of finding first integrals of motion. The procedure has been illustrated by a number of suitable examples alongwith Kamke's equation with explicit time dependent coefficients.
Article
Full-text available
In a recent paper (A. Mitsopoulos and M. Tsamparlis, J. Geom. Phys. 170, 104383, 2021), a general theorem is given which provides an algorithmic method for the computation of first integrals (FIs) of autonomous dynamical systems in terms of the symmetries of the kinetic metric defined by the dynamical equations of the system. In the present work, we apply this theorem to compute the cubic FIs of autonomous conservative Newtonian dynamical systems with two degrees of freedom. We show that the known results on this topic, which have been obtained by means of various different methods, and additional ones derived in this work can be obtained by the single algorithmic method provided by this theorem. The results are collected in four Tables which can be used as an updated reference of this type of integrable and superintegrable potentials. The results we find are for special values of free parameters; therefore, using the methods developed here, other researchers by different suitable choice of the parameters will be able to find new integrable and superintegrable potentials.
Article
Full-text available
We propose an interacting nonhermitian model described by a two-mode quadratic Hamiltonian along with an interaction term to locate and analyse the presence of an exceptional point in the system. Each mode is guided by a Swanson-like quadratic Hamiltonian and a suitable choice is made for the interaction term. The parity-time symmetric transformation is adopted in the standard way relevant for a coupled system.
Article
Full-text available
We review the general theory of the Jacobi last multipliers in geometric terms and then apply the theory to different problems in integrability and the inverse problem for one-dimensional mechanical systems. Within this unified framework, we derive the explicit form of a Lagrangian obtained by several authors for a given dynamical system in terms of known constants of the motion via a Jacobi multiplier for both autonomous and nonautonomous systems, and some examples are used to illustrate the general theory. Finally, some geometric results on Jacobi multipliers and their use in the study of Hojman symmetry are given.
Article
Full-text available
The problem of Liouvillian integrability for the classical force-free generalized Duffing oscillators is solved completely. All the cases when the generalized Duffing oscillators possess Liouvillian first integrals are classified. It is shown that the general solutions in integrable cases are expressible via elliptic and hyperelliptic functions. The relationship between the generalized Duffing systems and the Newell–Whitehead–Segel equation is used to characterize algebraically invariant traveling waves of the latter.
Article
We start with a self-contained brief review of the construction of non-standard Lagrangian and Hamiltonian structures for the Liénard equations satisfying Chiellini condition, we apply it to a special polynomial class of Liénard equation and explore the integrability structure. Then after a brief exposition of the contact geometry and its connection with the non-standard Hamiltonian structures, we describe the time evolution of the contact Hamiltonian and Liénard equation. We present the formulation of the Liénard equation in terms of General Equation for the Non-Equilibrium Reversible–Irreversible Coupling (GENERIC) method and also study the gradient-type flow. Finally, we describe the generalized Liénard equation by using the conformal Hamiltonian mechanics and illustrate our construction using spatio-temporal and autocatalysis systems.
Article
A family of non-Hermitian real and tridiagonal-matrix candidates H(N)(λ)=H0(N)+λW(N)(λ) for a hiddenly Hermitian (a.k.a. quasi-Hermitian) quantum Hamiltonian is proposed and studied. Fairly weak assumptions are imposed upon the unperturbed matrix [the square-well-simulating spectrum of H0(N) is not assumed equidistant)] and upon its maximally non-Hermitian N-parametric antisymmetric-matrix perturbations [matrix W(N)(λ) is not even required to be PT-symmetric]. Despite that, the “physical” parametric domain D[N] is (constructively) shown to exist, guaranteeing that in its interior, the spectrum remains real and non-degenerate, rendering the quantum evolution unitary. Among the non-Hermitian degeneracies occurring at the boundary ∂D[N] of the domain of stability, our main attention is paid to their extreme version corresponding to Kato’s exceptional point of order N (EPN). The localization of the EPNs and, in their vicinity, of the quantum-phase-transition boundaries ∂D[N] is found feasible, at the not too large N, using computer-assisted symbolic manipulations, including, in particular, the Gröbner-basis elimination and the high-precision arithmetics.