Next Article in Journal
On Sufficiency Conditions for Some Robust Variational Control Problems
Previous Article in Journal
Geometric Properties of Planar and Spherical Interception Curves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Poisson Bracket Filter for the Effective Lagrangians

1
Department of Natural Sciences, Institute of Electrophysics, Kálmán Kandó Faculty of Electrical Engineering, Óbuda University, Tavaszmező u. 17, H-1084 Budapest, Hungary
2
Department of Natural Sciences, National University of Public Service, Ludovika tér 2, H-1083 Budapest, Hungary
3
Department of Physics, Institute of Physics, Budapest University of Technology and Economics, Muegyetem rkp. 3, H-1111 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Axioms 2023, 12(7), 706; https://doi.org/10.3390/axioms12070706
Submission received: 12 June 2023 / Revised: 16 July 2023 / Accepted: 19 July 2023 / Published: 20 July 2023
(This article belongs to the Section Mathematical Physics)

Abstract

:
One might think that a Lagrangian function of any form is suitable for a complete description of a process. Indeed, it does not matter in terms of the equations of motion, but it seems that this is not enough. Expressions with Poisson brackets are displayed as required fulfillment filters. In the case of the Schrödinger equation for a free particle, we show what we have to be careful about.

1. Introduction

The least action principle has an efficiency that is difficult to overestimate in the study of physical processes. The principle is axiomatic. Mathematically it is based on the variational calculus that pertains to the extremal problems of motion integrals [1,2]. Lagrange original idea was to deduce the Newton’s laws from a “higher principle”. Today, we consider that these desciptions are equivalent. The first step is to find the Lagrange function, to which we obtain the equation of motion of the physical problem by applying the extremization procedure. However, over the equations of motion or field equations, the Lagrangian formulation involves more desired physical relations. We can explore inner symmetries, conserved quantities of the process, the energy expressions, canonical variables and conjugated pairs [3]. The identifiability of these relationships and quantities enables the theory to become the basis of modern physics. The motivation of the present article is the following. There are such opinions in the literature that all Lagrangians are equivalent, producing the correct equations of motion [4,5]. We show that a careful examination of the Poisson bracket formulation is required to be sure of the right choice because deriving the correct equations of motion is only part of the solution.
A system can be described by a Lagrange function L V , by which the action S can be obtained
S = t 1 t 2 L V d t ,
and it is extremal for the real physical path during the time evolution ( t 1 t t 2 ) . Applying the calculus of variations we obtain the so-called Euler-Lagrange equations as equations of motions. If Lagrangian L V depends on the coordinate q and its time derivative q ˙ , i.e., L V = L V ( q , q ˙ ) , the Euler-Lagrange equation is
L V q d d t L V q ˙ = 0 ,
which is a second order ordinary differential equation. In the case of the systems with infinite degree of freedom (continua), the principle can be expressed by
δ S = δ t 1 t 2 L d V d t = 0 ,
where L is the Lagrange density function. The Lagrangian L may depend on the field variable Ψ = Ψ ( r , t ) as a generalized coordinate, and its time Ψ ˙ , Ψ ¨ and space Ψ , Ψ derivatives (∇ is the gradient, Δ is the Laplace operator), i.e., L = L ( Ψ , Ψ ˙ , Ψ ¨ , Ψ , Ψ ) . Then, the Euler-Lagrange equation is
L Ψ t L Ψ ˙ L Ψ + 2 t 2 L Ψ ¨ + L Ψ = 0 .
As a general rule, if the Lagrangian L contains a linear operator A acting on Ψ , then the term
A ˜ L ( A Ψ )
appears in the Euler-Lagrange equation, where A ˜ is the adjoint operator to A. The A = A ˜ means that the operator is self-adjoint. Here, we find the answer why it is impossible to calculate a Lagrangian for e.g., first time derivative of variables without any mathematical tricks. In general, we can say that for non-selfadjoint operators the Lagrangian can not be constructed directly. There are some useful ideas to override the difficulties: e.g., introducing the dissipation potential [6,7,8,9,10,11,12,13] in mechanical, and [14,15,16,17,18,19,20,21] in tranport problems. Many other modifications of variation methods exist to find relevant calculus to obtain the correct equations of motion [22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42]. The duplication of the variables—introduction the complex conjugted field variables—is also an applicable method in the case of Schrödinger field [43,44], other quantum fields [45], moreover the transport equations such as Fourier heat conduction [46,47,48]. A promising solution is usage of potential (generator) functions in different ways for electrodynamics [49,50], for the field theory of thermodynamics [51], or for dissipative mechanical systems [52]. In this method the potential functions generates the measurable fields. The potential based proceure is so effective that the canonical quantization of the dissipative harmonic oscillator becomes possible [53].
Regardless of the method used, it is true that not only the derived equations of motion (and their solutions) of the investigated system must be clear, but also that all related concepts must be part of the coherent description [54,55,56,57]. In this article, we present a well-known and interesting problem, which control points may be necessary for the correct construction of the theory.

2. Lagrange Density Functions for the Schrödinger Field of Free Particles

We show some examples for the possible Lagrange density functions in the classical quantum mechanics. It can be easily seen that the choice is not unique. The question is, are they all suitable for further description? We will use the physically-mathematically simplest example, the free quantum particle, to point out the background of the problem. It is enough because, from a logical point of view, even one case is enough to reveal a faulty construction. We can formulate different Lagrangians of a free quantum particle [4], e.g.,
L 1 = 2 2 m Ψ Ψ + i 2 ( Ψ Ψ ˙ Ψ ˙ Ψ ) ,
L 2 = 2 2 m Ψ Ψ i Ψ ˙ Ψ ,
L 3 = 2 2 m Ψ Ψ + i 2 ( Ψ Ψ ˙ Ψ ˙ Ψ ) ,
L 4 = 2 2 m Ψ Ψ i Ψ ˙ Ψ .
Here, Ψ is the complex conjugated variable to Ψ , and these are considered as independent of each other. All of these give the same two Euler-Lagrange equations
i Ψ ˙ = 2 2 m Ψ
i Ψ ˙ = 2 2 m Ψ ,
which are the Schrödinger equations. The real question is, are these Lagrangian functions equivalent? We show that it seems somehow contradictory to specify with Equations (6) and (8). It is enough to restrict our examination for the free particle because, in the case of any conservative potential of V, the required V Ψ Ψ does not influence the exposition below.

3. Canonical Momenta, Hamiltonian, and Poisson Bracket Expressions

To develop the description of behavior of the Schrödinger field, we shortly discuss the canonical formalism and the Poisson brackets. In general, the canonical momentum is
p = L q ˙ ,
when the Lagrangian is L ( q , q ˙ ) , it does not contain higher order time derivatives. In the case of fields when L ( Ψ , Ψ ˙ , Ψ , Ψ ) is without higher order time derivatives, the momentum can be similarly written by
p = L Ψ ˙ .
When the Lagrangian depends on two independent variables ( Ψ and Ψ in Equations (10) and (11)), there are two momenta
p = L Ψ ˙ , and p = L Ψ ˙ .
The Hamilton density function will be as it is usual
H = p Ψ ˙ + p Ψ ˙ L .
The time evolution of a physical quantity F can be obtained by the Poisson bracket expression, which means
F ˙ = [ F , H ] ,
where F and H may depend on φ i , φ i , φ i , φ i , φ i , φ i , p i and p i . Here, Ψ , p, Ψ and p are considered independent variables. The Poisson bracket is
[ F , H ] = δ F δ φ i δ H δ p i δ F δ p i δ H δ φ i ,
where we apply the notations φ 1 = Ψ , φ 2 = Ψ , p 1 = p and p 2 = p , and
δ δ φ i = φ i ( φ i ) + ( φ i ) , and δ δ p i = p i
are the functional derivatives. In our cases we may expect that
Ψ ˙ = [ Ψ , H ] , p ˙ = [ p , H ] ,
and
Ψ ˙ = [ Ψ , H ] , p ˙ = [ p , H ]
are completed.

4. Discussion of Lagrangians

Now, let us examine the different versions of Lagrangians given by Equations (6)–(9) from the viewpoint of the canonical formalism. We would expect that canonical variables ensures the same correct Hamiltonians. This is a strong requirement since the Hamiltonian relates to the energy of the system.
In the case of L 1 the momenta are
p 1 = i 2 Ψ , and p 1 = i 2 Ψ .
Applying Equation (15) the deduced Hamilton density is
H 1 = 2 2 m Ψ Ψ ,
by which we obtain the equation
Ψ ˙ = [ Ψ , H 1 ] = δ Ψ δ Ψ δ H 1 δ p 1 δ Ψ δ p 1 δ H 1 δ Ψ + δ Ψ δ Ψ δ H 1 δ p 1 δ Ψ δ p 1 δ H 1 δ Ψ = i m Ψ .
It can be seen that the result is not the Schrödinger equation, i.e., Equation (10) is not equal to Equation (23). Similarly,
Ψ ˙ = [ Ψ , H 1 ] = i m Ψ .
We can try it in another way. Substituting Ψ from Equation (21)
Ψ = 2 i p 1
into the Hamilton density (22), so we get
H 1 = i m p 1 Ψ .
Then we obtain
Ψ ˙ = [ Ψ , H 1 ] = δ Ψ δ Ψ δ H 1 δ p 1 δ Ψ δ p 1 δ H 1 δ Ψ + δ Ψ δ Ψ δ H 1 δ p 1 δ Ψ δ p 1 δ H 1 δ Ψ = i m Ψ ,
which is also incorrect, and of course this is the same as in Equation (23). The number 2 is missing from the denominator. It seems to us that there is no way to find the correct Schrödinger equation. Calculating the time evolution of p we obtain
p 1 ˙ = [ p 1 , H 1 ] = δ p 1 δ Ψ δ H 1 δ p 1 δ p 1 δ p 1 δ H 1 δ Ψ = i m p 1 ,
and using the form of p by Equation (21) we can write
Ψ ˙ = i m Ψ
Here, the number 2 is similarly missing from the denominator. This discrepancy is the previously mentioned motivation of the work.
In the case of L 2 the momenta are asymmetric
p 2 = 0 , and p 2 = i Ψ ,
the total momentum is in one of the canonical variable. The Hamiltonian is
H 2 = 2 2 m Ψ Ψ ,
which is equal to H 1 . Learning from the previous example, it is obvious to eliminate one of the variables, e.g., Ψ , by which we get
H 2 = 2 i m p 2 Ψ .
At this point we should calculate Ψ ˙
Ψ ˙ = [ Ψ , H 2 ] = δ Ψ δ Ψ δ H 2 δ p 2 δ Ψ δ p 2 δ H 2 δ Ψ = 2 i m Ψ ,
which is exactly the Schrödinger equation for Ψ
i Ψ ˙ = 2 2 m Ψ .
Let us calculate p ˙ 2
p ˙ 2 = [ p , H 2 ] = 2 i m p 2 .
Using the form of p 2 from Equation (29) we obtain
i Ψ ˙ = 2 2 m Ψ ,
which is also correct. It seems that the only non-zero momentum resolves the problem.
It is easy to check by a short calculation that the case of L 3 is similar to the ( L 1 ). The momenta are
p 3 = i 2 Ψ , and p 3 = i 2 Ψ .
The Hamilton density is
H 3 = 2 2 m Ψ Ψ ,
by which we obtain
Ψ ˙ = [ Ψ , H 3 ] = δ Ψ δ Ψ δ H 3 δ p 3 δ Ψ δ p 3 δ H 3 δ Ψ + δ Ψ δ Ψ δ H 3 δ p 3 δ Ψ δ p 3 δ H 3 δ Ψ = i m Ψ .
One can recognize that the term
i 2 ( Ψ Ψ ˙ Ψ ˙ Ψ )
causes the problem in both cases, because it results the devided momenta.
In the last case, L 4 , the canonical momenta are asymmetric
p 4 = 0 , and p 4 = i Ψ .
The Hamiltonian density is
H 4 = 2 2 m Ψ Ψ = H 4 = 2 i m p 4 Ψ .
Now, we can calculate the Poisson bracket
Ψ ˙ = [ Ψ , H 4 ] = 2 i m Ψ ,
which results the Schrödinger equation. Moreover, we obtain p 4 ˙
p ˙ 4 = [ p 4 , H 4 ] = 2 i m p 4 ,
which gives the Schrödinger equation for Ψ
Ψ ˙ = 2 i m Ψ .
That is correct, i.e., the there must be only one non-zero canonical momentum. The space derivative terms keep intact the description.

5. Resolution of the Problem

As we see, the problem is with the “symmetric” cases, with the time derivatives in the Lagrange term in Equations (6), (8) and (39). In contrast, the Poisson bracket of the “asymmetric” Lagrangian term results the correct equations of motion. It is because one of the canonically conjugate momenta, e.g., p, is zero, so the entire pulse of the space is in the other momentum p . This total pulse takes part in the generation of the time evolution of the relevant fields. From the point of view of the variational problem, the spaces Ψ and Ψ are considered independent. However, physically they are not. Similarly, the momenta p and p . They share the total impulse. On the other hand, it is also necessary to remember that Hamilton means total energy. Thus, when we calculate the Poisson brackets, the effect of Hamilton appears with a double weight. That is why division by two disappears. The solution is to enter only half of the Hamiltonian in the Poisson bracketed expression. We have shown this problem in a simple case. For complex tasks, the calculation of the Poisson brackets can ensure the selection of the correct Lagrange function.

6. Conclusions

We can deduce the equations of motion from different Lagrangian functions. In general, the canonical formulation and the construction of Hamiltonian allow a free choice since the Poisson brackets result from the correct equations of motion. In the present article, we point out a particular trap. We show the case of the free particle quantum motion, the Schrödinger equation, in which we must restrict the construction. The wave function, its complex conjugate, and derivatives formulate the time-dependent part. If both the canonical conjugated momenta (pulses) are non-zero, the Poisson brackets do not serve the required Scröndinger equations. If one of the momenta is zero, we obtain the correct Scrödinger equation. The reason for the faulty solution is the physical situation that the wave function and its complex conjugate are not independent fields. However, in the calculus of variation, we consider these fields independent. The requirement of physical independence is fulfilled if one of the momenta is zero. It means that the contradiction disappears by the asymmetric choice in the time-dependent term of the Lagrangian. The conclusion is that the Poisson brackets give a filter to get the correct Lagrangians.

Author Contributions

K.G. and F.M. contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

TKP2021-NVA-16 has been implemented with the support provided by the Ministry of Innovation and Technology of Hungary from the National Research, Development and Innovation Fund. This research was supported by the National Research, Development and Innovation Office (NKFIH) Grant Nr. K137852 and by the Ministry of Innovation and Technology and the NKFIH within the Quantum Information National Laboratory of Hungary. Supported by the V4-Japan Joint Research Program (BGapEng), financed by the National Research, Development and Innovation Office (NKFIH), under Grant Nr. 2019-2.1.7-ERA-NET-2021-00028.

Conflicts of Interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

References

  1. Morse, P.M.; Feshbach, H. Methods of Theoretical Physics I; McGraw-Hill: New York, NY, USA, 1953; pp. 275–347. [Google Scholar]
  2. Kristály, A.; Rǎdulescu, V.D.; Varga, C.G. Variational Principles in Mathematical Physics, Geometry, and Economics. Qualitative Analysis of Nonlinear Equations and Unilateral Problems; Encyclopedia of Mathematics and its Applications; Cambridge University Press: Cambridge, UK, 2010; Volume 136. [Google Scholar]
  3. Tar, J.K.; Rudas, I.J.; Bitó, J.F.; Kaynak, O.M. A New Utilization of the Hamiltonian Formalism in the Adaptive Control of Mechanical Systems Under External Perturbation. Intell. Autom. Soft Comput. 1999, 5, 303–312. [Google Scholar] [CrossRef]
  4. Brown, H.R.; Holland, P. Simple applications of Noether’s first theorem in quantum mechanics and electromagnetism. Am. J. Phys. 2004, 72, 34–39. [Google Scholar] [CrossRef] [Green Version]
  5. Anastasiou, B. Quantum Field Theory I; ETH Zürich: Zürich, Switzerland, 2020. [Google Scholar]
  6. Rayleigh, J.W.S. The Theory of Sound; Macmillan: London, UK, 1896. [Google Scholar]
  7. Onsager, L. Reciprocal Relations in Irreversible Processes I. Phys. Rev. 1931, 37, 405. [Google Scholar] [CrossRef] [Green Version]
  8. Onsager, L. Reciprocal Relations in Irreversible Processes II. Phys. Rev. 1931, 38, 2265. [Google Scholar] [CrossRef] [Green Version]
  9. Casimir, H.B.G. On Onsager’s Principle of Microscopic Reversibility. Rev. Mod. Phys. 1945, 17, 343. [Google Scholar] [CrossRef] [Green Version]
  10. Onsager, L.; Machlup, S. Fluctuations and Irreversible Processes. Phys. Rev. 1953, 91, 1505. [Google Scholar] [CrossRef]
  11. Machlup, S.; Onsager, L. Fluctuations and Irreversible Process. II. Systems with Kinetic Energy. Phys. Rev. 1953, 91, 1512. [Google Scholar] [CrossRef]
  12. Gyarmati, I. ; Non-Equilibrium Thermodynamics; Springer: Berlin/Heidelberg, Germany, 1970. [Google Scholar]
  13. Glansdorff, P.; Prigogine, I. Thermodynamics Theory of Structure Stability and Fluctuations; John Wiley and Sons: New York, NY, USA, 1971. [Google Scholar]
  14. Sieniutycz, S.; Berry, R.S. Conservation laws from Hamilton’s principle for nonlocal thermodynamic equilibrium fluids with heat flow. Phys. Rev. A 1989, 40, 348. [Google Scholar] [CrossRef]
  15. Sieniutycz, S.; Berry, R.S. Least-entropy generation: Variational principle of Onsager’s type for transient hyperbolic heat and mass transfer. Phys. Rev. A 1992, 46, 6359. [Google Scholar] [CrossRef]
  16. Sieniutycz, S.; Berry, R.S. Canonical formalism, fundamental equation, and generalized thermomechanics for irreversible fluids with heat transfer. Phys. Rev. E 1993, 47, 1765. [Google Scholar] [CrossRef]
  17. Sieniutycz, S. Conservation Laws in Variational Thermo-Hydrodynamics; Kluwer: Dordrecht, The Netherlands, 1994. [Google Scholar]
  18. Sieniutycz, S.; Shiner, J. Variational and extremum properties of homogeneous chemical kinetics. I. Lagrangian- and Hamiltonian-like formulations. Open Sys. Inf. Dyn. 1992, 1, 149. [Google Scholar] [CrossRef]
  19. Sieniutycz, S.; Shiner, J. Variational and extremum properties of homogeneous chemical kinetics. II. Minimum dissipation approaches. Open Sys. Inf. Dyn. 1992, 1, 327. [Google Scholar] [CrossRef]
  20. Sieniutycz, S.; Ratkje, S.K. Variational principle for entropy in electrochemical transport phenomena. Int. J. Eng. Sci. 1996, 34, 549. [Google Scholar] [CrossRef]
  21. Sieniutycz, S.; Ratkje, S.K. Perturbational thermodynamics of coupled electrochemical heat and mass transfer. Int. J. Heat Mass Trans. 1996, 39, 3293. [Google Scholar] [CrossRef]
  22. Meixner, J.; Reik, H.G. Thermodynamik der Irreversiblen Prozesse; Encyclopedia of Physics Volume III/2, Principles of Thermodynamics and Statistics; Springer: Berlin/Heidelberg, Germany, 1959. [Google Scholar]
  23. de Groot, S.R.; Mazur, P. Non-Equilibrium Thermodynamics; North-Holland: Amsterdam, The Netherlands, 1962. [Google Scholar]
  24. Prigogine, I. Introduction to Thermodynamics; Interscience: New York, NY, USA, 1969. [Google Scholar]
  25. Biot, M. Variational Principles in Heat Transfer; Oxford University Press: Oxford, UK, 1970. [Google Scholar]
  26. Djukic, D.; Vujanovic, B. On a New Variational Principle of Hamiltonian Type for Classical Field Theory. Z. Angew. Math. Mech. 1971, 51, 611. [Google Scholar] [CrossRef]
  27. Keizer, J. Variational principles in nonequilibrium thermodynamics. Biosystems 1977, 8, 219. [Google Scholar] [CrossRef]
  28. Lebon, G.; Perzyna, P. Variational Principles in Thermomechanics; Recent Developments in Thermomechanics of Solids; Lebon, G.; Perzyna, P. Springer: Wien, Austria, 1980. [Google Scholar]
  29. Kreuzer, H.J. Nonequilibrium Thermodynamics and its Statistical Foundations; Clarendon: Oxford, UK, 1981. [Google Scholar]
  30. Gerjuoy, E.; Rau, A.R.P.; Spruch, L. A unified formulation of the construction of variational principles. Rev. Mod. Phys. 1983, 55, 725. [Google Scholar] [CrossRef]
  31. Kupershmidt, B.A. Hamiltonian formalism for reversible non-equilibrium fluids with heat flow. J. Phys. A 1990, 23, L529. [Google Scholar] [CrossRef]
  32. Lampinen, M.J. A problem of the Principle of Minimum Entropy Production. J. Non-Equilib. Thermodyn. 1990, 15, 397. [Google Scholar]
  33. Grmela, M.; Lebon, G. Hamiltonian extended thermodynamics. J. Phys. A 1990, 23, 3341. [Google Scholar] [CrossRef]
  34. Grmela, M.; Jou, D. A Hamiltonian formulation for two hierarchies of thermodynamic evolution equations. J. Phys. A 1991, 24, 741. [Google Scholar] [CrossRef]
  35. Eu, B.C. Kinetic Theory and Irreversible Thermodynamics; Wiley: New York, NY, USA, 1992. [Google Scholar]
  36. Ichiyanagi, M. On a Variational Principle of Hamilton’s Type for Irreversible Processes. J. Phys. Soc. Japan 1993, 62, 2650. [Google Scholar] [CrossRef]
  37. Ichiyanagi, M. Variational principles of irreversible processes. Phys. Rep. 1994, 243, 125. [Google Scholar] [CrossRef]
  38. Ciancio, V.; Verhás, J. On the Nonlinear Generalizations of Onsager’s Reciprocal Relations. J. Non-Equilib. Thermodyn. 1994, 19, 184. [Google Scholar] [CrossRef]
  39. Ván, P.; Muschik, W. Structure of variational principles in nonequilibrium thermodynamics. Phys. Rev. E 1995, 52, 3584. [Google Scholar] [CrossRef] [PubMed]
  40. Verhás, J. Thermodynamics and Rheology; Kluwer: Dordrecht, The Netherlands, 1997. [Google Scholar]
  41. Kazinski, P.O. Stochastic deformation of a thermodynamic symplectic structure. Phys. Rev. E 2009, 79, 011105. [Google Scholar] [CrossRef] [Green Version]
  42. Kim, Eun-jin; Nicholson, S. B. Complementary relations in non-equilibrium stochastic processes. Phys. Lett. A 2015, 379, 1613. [Google Scholar] [CrossRef]
  43. Schiff, L.I. Quantum Mechanics; McGraw-Hill: New York, NY, USA, 1986. [Google Scholar]
  44. Arsenović, D.; Burić, N.; Davidović, D.M.; Prvanović, S. Lagrangian form of Schrödinger equation. Found Phys. 2014, 44, 725. [Google Scholar] [CrossRef] [Green Version]
  45. Weinberg, S. The Quantum Theory of Fields; Cambridge University Press: Cambridge, UK, 1995. [Google Scholar]
  46. Anthony, K.-H. Hamilton’s action principle and thermodynamics of irreversible processes—A unifying procedure for reversible and irreversible processes. J. Non-Newton. Fluid Mech. 2001, 96, 291. [Google Scholar] [CrossRef]
  47. Anthony, K.-H. Lagrange Formalism and Irreversible Thermodynamics: The Second Law of Thermodynamics and the Principle of Least Entropy Production. In Variational and Extremum Principles in Macroscopic Systems; Sieniutycz, S., Farkas, H., Eds.; Elsevier: Amsterdam, The Netherlands, 2005; pp. 25–56. [Google Scholar]
  48. Glavatskiy, K.S. Lagrangian formulation of irreversible thermodynamics and the second law of thermodynamics. J. Chem. Phys. 2015, 142, 204106. [Google Scholar] [CrossRef] [Green Version]
  49. Maxwell, J.C. Treatise on Electricity and Magnetism; Dover: New York, NY, USA, 1954. [Google Scholar]
  50. Jackson, J.D. Classical Electrodynamics; John Wiley and Sons: New York, NY, USA, 1999. [Google Scholar]
  51. Gambár, K.; Márkus, F. Hamilton–Lagrange formalism of nonequilibrium thermodynamics. Phys. Rev. E 1994, 50, 1227–1231. [Google Scholar] [CrossRef]
  52. Szegleti, A.; Márkus, F. Dissipation in Lagrangian formalism. Entropy 2020, 22, 930. [Google Scholar] [CrossRef]
  53. Márkus, F.; Gambár, K. A potential-based quantization procedure of the damped oscillator. Quantum Rep. 2022, 4, 390–400. [Google Scholar] [CrossRef]
  54. Kristály, A. Multiple solutions of a sublinear Schrödinger equation. NoDEA Nonlinear Differ. Eqs. Appl. 2007, 14, 291–301. [Google Scholar] [CrossRef] [Green Version]
  55. Kristály, A. A double eigenvalue problem for Schrödinger equations involving sublinear nonlinearities at infinity. Electron. J. Differ. Eqs. 2007, 42, 11. [Google Scholar]
  56. Kristály, A.; Repovš, D. On the Schrödinger-Maxwell system involving sublinear terms. Nonlinear Anal. Real World Appl. 2012, 13, 213–223. [Google Scholar] [CrossRef] [Green Version]
  57. Kristály, A.; Moroṣanu, G.; O’Regan, D. A dimension-depending multiplicity result for a perturbed Schrödinger equation. Dynam. Syst. Appl. 2013, 22, 325–335. [Google Scholar]
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.

Share and Cite

MDPI and ACS Style

Gambár, K.; Márkus, F. Poisson Bracket Filter for the Effective Lagrangians. Axioms 2023, 12, 706. https://doi.org/10.3390/axioms12070706

AMA Style

Gambár K, Márkus F. Poisson Bracket Filter for the Effective Lagrangians. Axioms. 2023; 12(7):706. https://doi.org/10.3390/axioms12070706

Chicago/Turabian Style

Gambár, Katalin, and Ferenc Márkus. 2023. "Poisson Bracket Filter for the Effective Lagrangians" Axioms 12, no. 7: 706. https://doi.org/10.3390/axioms12070706

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop