Skip to main content

The Long Journey to the Schrödinger Equation

  • Chapter
  • First Online:
The Emerging Quantum

Abstract

This chapter marks the beginning of an itinerary that will take us to the quantum theory of matter. All along this exciting journey we will be accompanied by the zero-point radiation field introduced in Chap. 3, considered as a real, fluctuating field in permanent interaction with matter. Our point of departure is the (nonrelativistic) Abraham-Lorentz!equationAbraham-Lorentz equation governing the particle motion under the action of this field plus an arbitrary external (binding, conservative) force. A statistical treatment leads to a generalized phase-space Fokker-Planck equation!and laws of evolutionFokker-Planck equation. In the transition to configuration space through a partial averaging, a hierarchy of equations is obtained for the local moments of the momentum. When a balance is eventually reached in the mean between the energy lost by the particle through radiation reaction Radiation reactionand the energy gained by it from the background field, any remaining effect of the radiation terms becomes negligible. In this (time-asymptotic) limit, theFokker-Planck equation!generalized generalized Fokker-Planck equation!and laws of evolutionFokker-Planck equation transforms into a true Fokker-Planck equationFokker-Planck equation, describing a Markov processMarkov process. Further, in the Approximation!radiationlessradiationless approximationRadiationless approximation Linear sed!radiationless approximation the first two equations of the hierarchy decouple from the rest and are shown to be equivalent to Schrödinger’s equation. The main lessons and implications of these results are discussed. In particular, an explanation is given for the impossibility to go back from the Schrödinger description Quantum regime!and Schrödinger descriptionand retrieve a fullDetailed energy balance!and Schrödinger equation phase-space description that is consistent with quantum mechanics.

I am, in fact, rather firmly convinced that the Essentially statistical theory essentially statistical character of contemporary quantum theory is solely to be ascribed to the fact that this [theory] operates with an incomplete description of physical systems.

[In] a complete physical description, the statistical quantum theory would \(\ldots \) take an approximately analogous position to the statistical mechanics within the framework of classical mechanics \(\ldots \)

A. Einstein (1949)

\(\ldots \) I think that we cannot afford to neglect any possible point of view for looking at Quantum Mechanics and in particular its relation to Classical Mechanics. Any point of view which gives us any interesting feature and any novel idea should be closely examined to see whether they suggest any modification or any way of developing the theory along new lines.

P. A. M. Dirac (1951)

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 109.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 139.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 139.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    This chapter draws to a large extent from previous work, containedCetto, A. M. in the followingCetto, A. M. references: de la De la Peña, L. Peña and Cetto (1977a, b, 1995, 1996, 2005, 2006, 2007), de la Peña et al. (2009, 2012a, b), Cetto et al. (1984, 2012).

  2. 2.

    Instead of starting from the set of Hamilton equations for the entire system (particle plus field), we use as point of departure the (approximate) equation of motion for the particle. A detailed derivation of Eq. (4.1) from the Hamiltonian can be seen in many texts on electrodynamics. A particularly clear discussion is given by Cohen et al. (1989). See also de la Peña (1983), or The Dice.

  3. 3.

    In the usual derivations of Eq. (4.1), a retarded integral over time is written as a series expansion in terms of time derivatives of \(\varvec{x}.\) The lion’s share of this series pertains to the external (Lorentz) force. Then follow terms proportional to \(\varvec{\ddot{x}}\) and \(\varvec{\dddot{x}},\) and higher time derivatives are neglected, so the equation is approximate.

    The term proportional to \(\varvec{\ddot{x}}\) adds a ‘small’ electromagnetic correction \(\delta m\) to the mass. It happens that the integral that expresses this correction is divergent, because of the approximations made. In Eq. (4.1) this poses no problem, since \(m=m_{0}+\) \(\delta m\) is considered to correspond to the experimental (renormalized) mass, \(m_{0}\) being the mass that appears in the initial Hamiltonian. A formal procedure to solve this problem consists in adding to the initial Hamiltonian a mass counterterm that takes into account (with the opposite sign) the contribution to the mass of all neglected terms, and thus eliminates the infinite contribution (of course, it is infinite itself!). This clever cancellation of one infinity with another —a procedure that in qed (and more generally in quantum field theory) gives excellent results—represents a regularization Regularizationby renormalization.

    The radiation-reaction term, being proportional to \(\varvec{\dddot{x}},\) transforms the equation of motion into one of third order, thus demanding extra initial (or final) conditions. This term is known to lead to some awkward noncausal effects, such as preacceleration Preacceleration, i.e., response in advance to the external force (although in the present case the advanced times are of order \(\tau \sim 10^{-23}\) s, so the effect is in practice negligible). It should be clearly understood that this noncausal behaviour is also a result of the neglect of the higher-order terms, since the theory in its closed form is absolutely causal. A similar situation is met with the Lorentz-Dirac equation, which is the relativistic version of the Abraham-Lorentz equationRohrlich, F. (see Rohrlich 1965). A more extensive discussion of the radiation force and the problems connected with it in the context of the present theory, as well as a causal variant of it, can be seen in The Dice, Sect. 3.3. The current approximate form is more convenient in practical terms, provided one bears in mind that its noncausal features are an artifact of the approximation.

    Notice that \(m\tau \varvec{\dddot{x}}\) represents the electric component of the field radiated by the particle, which is of a nature similar to \(e\varvec{E}(t);\) this can be made explicit by writing \(m\tau \varvec{\dddot{x}}=e\varvec{E}_{\text {rad}}(\varvec{x},t)\), where \(\varvec{E}_{\text {rad}}(\varvec{x},t)\) stands for the radiated field.

  4. 4.

    As will become clear in Chap. 5, the relevant modes are those associated with the dominant response of the particle to the field (exhibited e.g. in the atomic transitions), which have indeed wavelengths much larger than the atomic dimensions.

  5. 5.

    This formula is the result of counting the number of field modes of frequency \(\omega =2\pi c/\lambda \) (with both polarizations)Correlations!photon polarization per unit volume within an interval of frequency \(\Delta \omega \). Assuming the distributionDistribution of radiation to be homogeneous and isotropic, this gives after integrating over the solid angle: \(\Delta n=(1/\pi ^{2}c^{3})\int _{\Delta \omega }\omega ^{2}d\omega \).

  6. 6.

    As remarked in note 5 of Chap. 3, the spectrum proportional to \(\omega ^{3} \) is the single one for which all inertial observers are equivalent. This can be confirmed by calculating the force exerted by a homogeneous and isotropic background field on a dipole moving with velocity \(v\), which is given by \(F=-(6/5)\pi ^{2}\tau c[\rho (\omega )-(\omega /3)(d\rho /d\omega )]v\) (see Einstein and Hopf 1910; The Dice, Chap. 4). Only for \(\rho (\omega )\sim \omega ^{3}\) this force becomes zero.

  7. 7.

    A Fokker-Planck equationFokker-Planck equation (fpe) is a differential equation of second order that describes the evolution of the probability density for the particle subject to a white noise White noise(an uncorrelated noise, with a flat power spectrumPower spectrum). Given that the present problem involves a colored noise, the corresponding equation for the probability density in phase space is not a true fpe, but a generalization of it that contains memory Memoryterms, leading to an integro-differential equation. It is to such equation that we refer as a gfpe. For conceptually rich, early introductions to the fpe for the study of Brownian motion see the papers by S. Chandrasekhar and by Ming Chen Wang and G. E. Uhlenbeck Uhlenbeck, G. E. Wax, N.in Wax (1954/1985). For a first-rate presentation of the subject seeStratonovich, R. L. Stratonovich 1963. For a more recent presentationRisken, H. see Risken (1984)Velasco, R. M.; see alsoCetto, A. M. Cetto et al. (1984).

  8. 8.

    The approximation consists in assuming that the field remains essentially unmodified. For a complete description one should write the continuity equation as

    $$\begin{aligned} \frac{\partial R}{\partial t}+\frac{\partial }{\partial x_{a}}\left( \dot{x}_{a}R\right) +\frac{\partial }{\partial p_{a}}\left( \dot{p}_{a}R\right) =0, \end{aligned}$$

    where \(R(\left\{ x_{a},p_{a}\right\} ,t)\) stands for the density of points in the entire phase space of the particle plus field system, so that \(\{x_{a}\}=\{x_{f},x_{i}\}\) and \(\{p_{a}\}=\{p_{f},p_{i}\}\), where the index \(f\) refers to the field quadratures and \(i\) to the particle’s variables.

  9. 9.

    Right after particle and field start to interact, the system is far from equilibrium. In this regime the main effect of the zpf Approximation!fixed zpf on the particle is due to the high-frequency modes, which produce violent accelerations and randomize the motion. Eventually, the interplay between the electric field force and radiation reaction Radiation reactionis expected to drive the system close to equilibrium; in this (time-reversible) regime the Markovian approximation Markovian approximationapplies. The duration of the transient period, i.e. the time \(t_{M}\) required by the system to reach the Markovian limit, is determined basically by the effect of the high-frequency modes. Since the particle is assumed to respond to modes of frequency up to \(mc^{2}/\hbar \) (see Chap. 6), \(t_{M}\) is estimated to be of of the order of \(\hbar /mc^{2}\simeq 10^{-20}\) s for an electron.

  10. 10.

    It was Born Born, M.who introduced the interpretation of \(q *q\) as a probability in quantum mechanics, though limited to the description of dispersionMomentum!dispersion states. The proposal of interpreting this quantity as a probability density more generally was put forward by Pauli in (1927). Here, Born’sBorn, M. rule ensues from the theory itself.

  11. 11.

    More generally, the condition under which Eq. (4.84) holds is

    $$\begin{aligned} \mathop {\displaystyle \int }\varvec{F}_{\varvec{\Sigma }}\cdot d\varvec{x}-\mathop {\displaystyle \int }\varvec{F}_{\varvec{g}}\cdot d\varvec{x}+h(t)=0\varvec{,} \end{aligned}$$

    with \(\varvec{F}_{\varvec{g}}=(\partial \varvec{g/}\partial t)-\varvec{v}\times \left( \varvec{\nabla }\times \varvec{g}\right) ,\) and \(\varvec{F}_{\varvec{\Sigma }}=-(1/m\rho )\varvec{\nabla }\cdot \left( \varvec{\tilde{\Sigma }}\rho \right) .\) The vector field \(\varvec{g}\) given by Eq. (4.74) determines the function \(\varvec{\tilde{\Sigma }}\) and should contain all electromagnetic contributions: the zpf, the self-field radiated by the particle, any external field, and any existing excitation. This leads to the territory of qed. Since the radiated field depends on the dynamics, it can be known only by solving the entire (matter-field) problem.

  12. 12.

    In Sect. 4.3.1, the transition from phase space to the momentum Momentumsubspace was shown to lead to an integro-differential equation that is explicitly nonlocal due to the integral transform. The nonlocal character of the reduced description becomes then obvious. A similar situation occurs in the transition to configuration space, but then the nonlocality of the description (even for a single-particle system) is manifested through the term containing \(\eta ^{2},\) which embodies information of the probability distribution of particles. This point is discussed more at length in Chap. 8.

  13. 13.

    In the present context Contextthese relations appear simply as a result of the calculations. Their physical meaning will become clear when they reappear in Chap. 5, in connection with the derivation of the Heisenberg descriptionLinear sed!Heisenberg description Harmonic oscillator!Heisenberg description of qm.

  14. 14.

    During the development of sed in the late ninetiteenseventies and until the eighties, use was made of the true fpe derived in Appendix B. That equation was applied to several problems, giving some correct results for linear problems, and wrong answers for the rest, particularly the H atom. From what we have just seen, it is clear that a mistake was being made by applying the classical fpe (with classical variables) to a system that is already following a nonclassical behavior. In particular, the fpe by itself does not guarantee that the quantum regimeQuantum regime Linear sed!and the quantum regime has been reached; hand in hand with it, the energy-balance conditionEnergy-balance condition must be in force. The erroneous results obtained, characterized by the violation of the energy-balance condition, led unfortunately to a quite extended belief that the stochastic approach to qm based on the zpf wasLozano, N. wrong. This was first shown inBoyer, T.H. Boyer (1976), (1980), and then by several other authorsAlcubierre, M. (see e.g. Marshall and Claverie 1980;Claverie, P. Alcubierre and Lozano 1988); a detailed exposition and further references are given in The Dice). In this context it is relevant to recall the statement by Claverie andDiner, S. Diner in (1977): “The relationship between quantum theory and sed, if it exists, is of a more subtle nature than [a] mere formal equivalence.”

  15. 15.

    AStochastic quantization related subject is stochastic quantization Stochastic quantization Parisi, G. Wu Y. S. (Parisi and Wu 1981; Masujima 2009), which makes use of an imaginaryMasujima, M. time \(\tau \) related to the real time by \(\tau \) \(=it\). This transforms the time-dependent Schrödinger equation into a diffusion-like equation, so the expansion in terms of eigenfunctions of the Hamiltonian takes the form of a partition function of statistical mechanics with \(\tau \) interpreted as the inverse temperature, \(\tau \rightarrow \beta =1/k_{B}T,\)

    $$\begin{aligned} \Psi (x,t)=\sum _{n}\psi _{n}\exp \left( -i\mathcal {E}_{n}t\right) =\sum _{n}\psi _{n}\exp \left( -\tau \mathcal {E}_{n}\right) =\sum _{n}\psi _{n}\exp \left( -\beta E_{n}\right) . \end{aligned}$$

    This procedure has proved to be of value in several applications, particularly in quantum field theory, by allowing for a treatment of quantum problems with the methods of statistical mechanics or stochastic processes. Of course, stochastic quantization Stochastic quantizationis just a formal method of calculation; it is not intended to improve the interpretation of qm.

  16. 16.

    There exists a profuse quantum literature in which the term related to the momentum fluctuations enters through one door or another (for a discussion and several examples seeCarroll, R. Carroll 2010). Their contribution is rarely identified as coming from Fluctuationsfluctuations in the momentum space, and almost never as due to the zpf. Most frequently they are simply taken as ‘quantum fluctuations’, a term that conveys the idea that they are spontaneous, i.e., causeless. As a result, the momentum fluctuations term appears in the literature under several guises. In the stochastic theory of qm it is identified as produced by the velocity \(\varvec{u}\); see Chap. 2 and references therein. In Olavo (2000)Olavo, L. S. F. it is interpreted as coming from a local Entropyentropy due to spontaneous local fluctuations in positions. In the Bohmian theory it is known as the quantum potential Quantum potential(see Chap. 8 and, e.g., Holland 1993)Holland, P. R., which in its turn is ‘explained’ as a Fisher information (Frieden 1998), Frieden, B.R. Roy, S.Roy (1986) relates it to fluctuations of the metric, and so on.

  17. 17.

    Though this result may be quite surprising, in fact it is not, since the continuity equation along with the factorization \(\rho =\psi ^{*}\psi \) readily leads to an equation having the general form of the Schrödinger equation, as has been knownDetailed balance!and Schrödinger equation for a long time (see e.g. de la Peña 1967; Jammer 1974; Kracklauer 1992)Kracklauer, A. F.. This shows that the structureStructure of the Schrödinger equation, rather than being specific to qm, is a kind of generic framework in the presence of fluctuations (seeCarroll, R. e.g. Carroll 2010). Of course, the term proportional to \(\psi \left( \mathbf {\nabla }^{2}R\right) /R\) in Eq. (4.150) disappears if it is subtracted from the potential in Eq. (4.145).

  18. 18.

    Entire books have been dedicated to analyze and discuss the problem of a quantum-mechanical phase-space distributionZachos, C. K.. See, e.g., Zachos et al. (2005)Curtright, T. L.

  19. 19.

    The first phase-space Phase-space distribution!Wigner functiondescription of a quantum system was madeWeyl H. in Weyl (1927); the Wigner functionHarmonic oscillator!Wigner function Wigner function was introduced and studied firstly in Dirac (1930). It was later proposed independently in Heisenberg, W.Heisenberg (1931) and Wigner (1932). The theory of the latter was substantially developed in Moyal, J. E.Moyal (1949). A discussion of the fundamentals of the distribution functions is given in Hillery, M.Hillery et al. (1984) (see alsoTatarskii V. I. Tatarskii 1983); for an introductory account of the Wigner functionPhase-space distribution!Wigner function see Case (2008)Case, W. B.. General formulas for quantum phase-spacePhase-space distribution!Wigner function Wigner function distributionsPhase-space distribution!Wigner function (which apply to the Wigner function as a special case)Harmonic oscillator!Wigner function Wigner function Oscillator!Wigner function are given in Cohen (1976) andZaparovanny, Y. I. Cohen and ZaparovannyCurtright, T. L. (1980). A general overview with selected papers is Zachos et al. (2005).

  20. 20.

    Some recent independent investigations are of particular relevance to the present theory. We recall the important numerical simulations inCole, D. C. Cole (2006) andZou, Y. Cole and Zou 2003–2004 leading to a correct statistical prediction of the ground state Oscillator!ground stateorbit for the H-atomH-atom (and some results for its excitations in ColeCole, D. C. and ZouZou, Y. 2009). In Huang, K. Batelaan, H. Huang and Batelaan (2012a, b), the modes of a classical one-dimensional harmonic oscillator immersed in the zpf and excited by an electromagnetic pulse are studied by numerical simulation, with results that are in excellent agreement with the quantum predictions. The authors are indebted to Khaled DechoumDechoum, Kh. and Emilio Santos Santos, E.for having drawn their attention to this work at an early stage.

  21. 21.

    It could be argued that the quantum description is indeterministic. This is obviously true, and is in consonance with any statistical description. The point is that the starting equation of motion, valid for an individual member of the ensemble, is deterministic. IndeterminismIndeterminism enters because the specific realization of the field in the individual case is unknown. Thus quantum indeterminism should not be understood as intrinsic to matter at the microscopic level, but rather of a nature similar to that of statistical physics. Something similar can be said about nonlocality: the initial theory is local; the final statistical and partial description is the one that acquires nonlocal properties. For related discussions on causalityCausality, determinismDeterminism, realism Realismand locality Localitysee Chap. 1.

  22. 22.

    Recently Huang, K. Huang and Batelaan (2012a, b) have proposed another form of visualizing the problem, by considering that the random motion gives rise to instantaneous multipolar moments that couple to the corresponding modes of the radiation field. Strictly speaking, this can be applied also to neutral (structured) particles.

  23. 23.

    The most obvious instance in which this is manifest is a system of \(N\) particles in three dimensions; then \(\psi \) lives in an abstract 3\(N\)-dimensional space, while physical waves (fields) live in three dimensions. This point, raised for the first time by Pauli, was a subject of much discussion during the early phase of qm; a detailed account can be seen inBacciagaluppi, G. Valentini, A. Bacciagaluppi and Valentini (1927). However, in the case of particles (fermions) the problem disappears using the number representation of the state vector. Then the 3\(N\) coordinates (if introduced at all) represent merely \(N\) points in three-dimensional space.

  24. 24.

    The fact that in qm the Born Born, M.rule is introduced as a postulate, is not a minor point. Indeed it is so important that serious efforts have been made for many years to demonstrate it from within qm. Probably the most far-reaching result of such attempts is due to Graham, R. N.Graham (1973), who proves that the probability of state \(n\) in the superposition Superpositionof states \(\mathop {\displaystyle \sum }\nolimits _{1}^{N}c_{n}\psi _{n}\) tends to \(\left| c_{n}\right| ^{2}\) for large values of \(N.\) See also Hartle (1968).

  25. 25.

    The factorization \(\rho =\psi ^{*}\psi \) taken as a minor mathematical liberty was apparently first introduced inCollins, R. E. Collins (1977) and has been repeatedly used by several authors. But in fact it is not inconsequential, since it opens the possibility to introduce a phase function, not present in the absence of such factorization. Only if a physical meaning and well-defined mathematical properties can be attributed to such function, so that it belongs naturally to the theory, the procedure can be considered acceptable. For other comments on this important matter see Sect. 2.6.

  26. 26.

    As shown in these references, it is correct to assume that the distribution is Gaussian in the case of the free field. Strictly speaking \(\varvec{E}(t)\) cannot be taken as a free field, since it is somehow modified by its interaction with the mechanical subsystem. However, the modifications affect only a very reduced set (which can be considered of measure zero) of the (averaged) modes of the field in the vicinity of the particle, by introducing correlations among the phases of some of these modes. For the rest, the field remains essentially unchanged. For details see Peña et al. (2009), or Chap. 5.

References

  • Alcubierre, M., Lozano N.: Tratamiento de Sistemas Multiperiódicos en la Electrodinámica Estocástica, Professional thesis (Universidad Nacional Autónoma de México, México) (1988)

    Google Scholar 

  • Bacciagaluppi, G., Valentini, A.: Quantum Theory at the Crossroads. Reconsidering the: Solvay Conference. Cambridge University Press, Cambridge (1927/2009)

    Google Scholar 

  • Balescu, R.: Equilibrium and Nonequilibrium Statistical Mechanics. Wiley, New York (1975)

    MATH  Google Scholar 

  • Ballentine, L.E.: Quantum Mechanics. Prentice Hall, New Jersey (1990)

    Google Scholar 

  • Ballentine, L.E.: Quantum Mechanics. A Modern Development. World Scientific, Singapore (1998)

    Book  MATH  Google Scholar 

  • **Boyer, T.H.: Random electrodynamics: The theory of classical electrodynamics with classical electromagnetic zero-point radiation. Phys. Rev. D 11, 790 (1975)

    Google Scholar 

  • *Boyer, T.H.: Equilibrium of random classical electromagnetic radiation in the presence of a nonrelativistic nonlinear electric dipole oscillator. Phys. Rev. D 13, 2832 (1976)

    Google Scholar 

  • Boyer, T.H.: A brief survey of stochastic electrodynamics. In: Barut, A.O. (ed.) Foundations of Radiation Theory and Quantum Electrodynamics. Plenum, New York (1980)

    Google Scholar 

  • Carroll, R.: On the Emergence Theme of Physics. World Scientific, Singapore (2010)

    Book  MATH  Google Scholar 

  • Case, W.B.: Wigner functions and Weyl transforms for pedestrians. Am. J. Phys. 76, 937 (2008)

    Article  ADS  Google Scholar 

  • Cetto, A.M., de la Peña, L., Velasco, R.M.: Generalized Fokker-Planck equations for coloured, multiplicative Gaussian noise. Rev. Mex. Fís. 31, 83 (1984)

    Google Scholar 

  • *Cetto, A.M., de la Peña, L., Valdés-Hernandez, A.: Quantization as an emergent phenomenon due to matter-zeropoint field interaction. J. Phys. JPCS 361, 012013 (2012)

    Google Scholar 

  • Claverie, P., Diner, S.: Stochastic electrodynamics and quantum theory. Int. J. Quantum Chem. 12, S1, 41 (1977)

    Google Scholar 

  • Cohen, L.: Quantization problem and variational principle in the phase-space formulation of quantum mechanics. J. Math. Phys. 17, 1863 (1976)

    Article  ADS  Google Scholar 

  • Cohen, L., Zaparovanny, Y.I.: Positive quantum joint distributions. J. Math. Phys. 21, 794 (1980)

    Article  ADS  MathSciNet  Google Scholar 

  • Cohen-Tannoudji, C., Dupont-Roc, J., Grynberg, G.: Photons and Atoms. Introduction to Quantum Electrodynamics. Wiley, New York (1989)

    Google Scholar 

  • Cole, D.C.: Simulation results related to stochastic electrodynamics. In: Adenier, G., Khrennikov, Y., Nieuwenhuizen, T.M. (eds.) Quantum Theory. Reconsideration of Foundations-3, AIP Conference Proceedings 810, New York, vol. 99 (2006)

    Google Scholar 

  • Cole, D.C., Zou, Y.: Quantum mechanical ground state of hydrogen obtained from classical electrodynamics. Phys. Lett. A 317, 14 (2003)

    Google Scholar 

  • Cole, D.C., Zou, Y.: Analysis of orbital decay time for the classical hydrogen atom interacting with circularly polarized electromagnetic radiation. Phys. Rev. E 69, 016601 (2004a)

    Article  ADS  Google Scholar 

  • Cole, D.C., Zou, Y.: Simulation study of aspects of the classical hydrogen atom interacting with electromagnetic radiation: circular orbits. J. Sci. Comput. 20, 43 (2004b)

    Article  MATH  MathSciNet  Google Scholar 

  • Cole, D.C., Zou, Y.: Simulation study of aspects of the classical hydrogen atom interacting with electromagnetic radiation: elliptical orbits. J. Sci. Comput. 20, 379 (2004c)

    Article  MATH  MathSciNet  Google Scholar 

  • Cole, D.C., Zou, Y.: Perturbation analysis and simulation study of the effects of phase on the classical hydrogen atom interacting with circularly polarized electromagnetic radiation. J. Sci. Comput. 21, 145 (2004d)

    Article  MATH  MathSciNet  Google Scholar 

  • Cole, D.C., Zou, Y.: Subharmonic resonance behavior for the classical hydrogen atomic system. J. Sci. Comput. 39(1) (2009)

    Google Scholar 

  • Collins, R.E.: Quantum theory: a Hilbert space formalism for probability theory. Found. Phys. 7, 475 (1977)

    Article  ADS  Google Scholar 

  • de la Peña, L.: A simple derivation of the Schrödinger equation from the theory of Markov processes. Phys. Lett. A 24, 603 (1967)

    ADS  Google Scholar 

  • **de la Peña, L.: Stochastic electrodynamics: its development, present situation and perspectives. In: Gmez, B., et al. (eds.) Stochastic Processes Applied to Physics and Other Related Fields. World Scientific, Singapore (1983)

    Google Scholar 

  • de la Peña, L., Cetto, A.M.: Stronger form for the position-momentum uncertainty relation. Phys. Lett. A 39, 65 (1972)

    Article  ADS  Google Scholar 

  • de la Peña, L., Cetto, A.M.: Derivation of quantum mechanics from stochastic electrodynamics. J. Math. Phys. 18, 1612 (1977a)

    Article  ADS  Google Scholar 

  • de la Peña, L., Cetto, A.M.: Why Schrödinger’s equation? Int. J. Quantum Chem. XII Supl. 1, 23 (1977b)

    Google Scholar 

  • de la Peña, L., Cetto, A.M.: Is quantum mechanics a limit cycle theory? In: Ferrero, M., van der Merwe, A. (eds.) Fundamental Problems in Quantum Physics, p. 47. Kluwer Academic, Dordrecht (1995)

    Google Scholar 

  • *de la Peña, L., Cetto, A.M.: The Quantum Dice An Introduction to Stochastic Electrodynamics. Kluwer Academic, Dordrecht (1996) (Referred to in the book as The Dice)

    Google Scholar 

  • *de la Peña, L., Cetto, A.M.: Contribution from stochastic electrodynamics to the understanding of quantum mechanics. http://arxiv.org/abs/quant-ph/0501011arXiv:quant-ph/0501011v2 (2005)

  • *de la Peña, L., Cetto, A.M.: Recent developments in linear stochastic electrodynamics. In: Adenier, G., Khrennikov, A.Y., Nieuwenhuizen, T. M. (eds.) Quantum Theory: Reconsideration of Foundations-3, AIP Conference Proceedings no. 810. AIP, New York. Extended version in arXivquant-ph0501011 (2006)

    Google Scholar 

  • *de la Peña, L., Cetto, A.M.: On the ergodic behaviour of atomic systems under the action of the zero-point radiation field. In: Nieuwenhuizen, T.M., et al. (eds.) Beyond the Quantum. World Scientific, Singapore (2007)

    Google Scholar 

  • *de la Peña, L., Valdés-Hernández, A., Cetto, A.M.: Quantum mechanics as an emergent property of ergodic systems embedded in the zero-point radiation field. Found. Phys. 39, 1240 (2009)

    Google Scholar 

  • *de la Peña, L., Cetto, A.M., Valdés-Hernández, A.: Quantum behavior derived as an essentially stochastic phenomenon. Phys. Scr. T151, 014008 (2012a)

    Google Scholar 

  • *de la Peña, L., Cetto A.M, Valdés-Hernández, A.: The emerging quantum. An invitation, Advanced School on Quantum Foundations and Open Quantum Systems, http://www.fisica.ufpb.br/asqf2012/index.php/ invited-lectures ( 2012b)

  • Dirac, P.A.M.: Note on exchange phenomena in the Thomas atom. Proc. Camb. Phil. Soc. 26, 376 (1930)

    Article  ADS  MATH  Google Scholar 

  • Dirac, P.A.M.: In: Proceedings of the second Canadian mathematical congress (University of Toronto), 10 (1951)

    Google Scholar 

  • Edwards, S.F., McComb, W.D.: Statistical mechanics far from equilibrium. J. Phys. A 2, 157 (1969)

    Article  ADS  MATH  Google Scholar 

  • Einstein, A.: Autobiographical notes and Einstein’s reply, In: Schilpp, P.A. (ed.) Albert Einstein: Philosopher-Scientist. Harper and Row, New York (1949)

    Google Scholar 

  • Einstein, A., Hopf, L.: Statistische Untersuchung der Bewegung eines Resonators in einem Strahlungsfeld. Ann. der Physik 33, 1096 (1910)

    Article  ADS  MATH  Google Scholar 

  • Frieden, B.R.: Physics from Fisher Information. Cambridge University Press, Cambridge (1998)

    Book  Google Scholar 

  • Frisch, U.: Wave propagation in random media. In: Bharucha-Reid, A.T. (ed.) Probabilistic methods in applied mathematics, vol. I. Academic, New York (1968)

    Google Scholar 

  • Fujita, S., Godoy, S.V.: Mathematical Physics. Wiley, Weinheim (2010)

    Google Scholar 

  • Gardiner, C.W.: Handbook of Stochastic Methods. Springer, Berlin (1983)

    Book  MATH  Google Scholar 

  • Graham, N.: The measurement of relative frequency. In: DeWitt, B.S., Graham, N. (eds.) The Many Worlds Interpretation of Quantum Mechanics. Princeton University Press, Princeton (1973)

    Google Scholar 

  • Hartle, J.B.: Quantum mechanics of individual systems. Am. J. Phys. 36, 704 (1968)

    Article  ADS  Google Scholar 

  • Hassani, S.: Mathematical Physics. Springer, New York (1999)

    Book  MATH  Google Scholar 

  • Heisenberg, W.: Über die inkohärente Streuung von Röntgenstrahlen. Zeit. Phys. 32, 737 (1931)

    Google Scholar 

  • Hillery, M., O’Connell, R.F., Scully, M.O., Wigner, E.P.: Distribution functions in physics fundamentals. Phys. Rep. 106, 121 (1984)

    Google Scholar 

  • Holland, P.R.: The Quantum Theory of Motion. Cambridge University Press, Cambridge (1993)

    Book  Google Scholar 

  • Huang, W.C.W., Batelaan, H.: Dynamics underlying the Gaussian distribution of the classical harmonic oscillator in zero-point radiation. arXiv/quant-ph:1206.5323v1 (2012a)

    Google Scholar 

  • Huang W.C.W., Batelaan H.: Quantized Excitation Spectrum of the Classical Harmonic Oscillator in Zero-Point Radiation. arXiv/quant-ph:1206.6891v1 (2012b)

    Google Scholar 

  • Hudson, R.L.: When is the Wigner quasi-probability density non-negative? Rep. Math. Phys. (Torun) 6, 249 (1974)

    Google Scholar 

  • Jammer, M.: The Philosophy of Quantum Mechanics. The interpretation of quantum mechanics in historical perspective. Wiley, New York (1974)

    Google Scholar 

  • Kracklauer, A.F.: An intuitive paradigm for quantum mechanics. Phys. Essays 5, 226 (1992)

    Article  ADS  MathSciNet  Google Scholar 

  • Landau, L., Lifshitz, E.: The Classical Theory of Fields. Addison-Wesley, Cambridge (1951)

    MATH  Google Scholar 

  • Madelung, E.: Quantentheorie in hydrodynamischer Form. Zeit. Phys. 40, 322 (1926)

    Google Scholar 

  • Marshall, T.W., Claverie, P.: Stochastic electrodynamics of nonlinear systems. I. Particle in a central field of force. J. Math. Phys. 21, 1819 (1980)

    Article  ADS  MathSciNet  Google Scholar 

  • Masujima, M.: Path Integral Quantization and Stochastic Quantization. Springer, Berlin (2009)

    MATH  Google Scholar 

  • Morse, P., Feshbach, H.: Methods of Theoretical Physics. McGraw-Hill, New York (1953)

    MATH  Google Scholar 

  • Moyal, J.E.: Quantum mechanics as a statistical theory. Proc. Camb. Phil. Soc. 45, 99 (1949)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  • Nernst, W.: Über einen Versuch, von quantentheoretischen Betrachtungen zur Annahme stetiger Energieänderungen zurückzukehren. Verh. Deutsch. Phys. Ges. 18, 83 (1916)

    Google Scholar 

  • Olavo, L.S.F.: Foundations of quantum mechanics: connection with stochastic processes. Phys. Rev. A 61, 052109 (2000)

    Article  ADS  Google Scholar 

  • Papoulis, A.: Probability, Random Variables, and Stochastic Processes, Chap. 6. McGraw-Hill, Boston (1991)

    Google Scholar 

  • Parisi, G., Wu, Y.S.: Perturbation theory without Gauge fixing. Sci. Sin. 24, 483 (1981)

    Google Scholar 

  • Pauli, W.: Über Gasentartung und Paramagnetismus. Zeit. Phys. 41, 81 (1927)

    Article  ADS  MATH  Google Scholar 

  • Piquet, C.: Fonctions de type positif associées a deux opérateurs hermitiens. C. R. Acad. Sci. Paris A 279, 107 (1974)

    MATH  MathSciNet  Google Scholar 

  • Risken, H.: The Fokker-Planck Equation. Methods of Solution and Applications. Springer, Berlin (1984)

    Book  MATH  Google Scholar 

  • Rohrlich, F.: Classical Charged Particles. Foundations of Their Theory. Addison-Wesley, Reading (1965)

    MATH  Google Scholar 

  • Rosen, N.: The relation between classical and quantum mechanics. Am. J. Phys. 32, 597 (1964)

    Article  ADS  Google Scholar 

  • Rosen, N.: Quantum particles and classical particles. Found. Phys. 16, 687 (1986)

    Article  ADS  MathSciNet  Google Scholar 

  • Roy, S.: Stochastic geometry and origin of quantum potential. Phys. Lett. A 115, 256 (1986)

    Article  ADS  MathSciNet  Google Scholar 

  • Schiller, R.: Quasi-classical theory of the nonspinning electron. Phys. Rev. 125, 1100 (1962a)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  • Schiller, R.: Quasi-classical transformation theory. Phys. Rev. 125, 1109 (1962b)

    Article  ADS  MathSciNet  Google Scholar 

  • Soto-Eguibar, F., Claverie, P.: When is the Wigner function of multi-dimensional systems nonnegative? J. Math. Phys. 24, 97 (1983a)

    Article  ADS  MathSciNet  Google Scholar 

  • Soto-Eguibar, F., Claverie, P.: Time evolution of the Wigner function. J. Math. Phys. 24, 1104 (1983b)

    Article  ADS  MathSciNet  Google Scholar 

  • Stratonovich, R.L.: Topics in the Theory of Random Noise, Vol. I: General Theory of Random Processes. Gordon and Breach, New York (1963)

    Google Scholar 

  • Surdin, M.: L’état fondamental de l’oscillateur harmonique est-il un cycle limite? Ann. Inst. Henri Poincaré 13, 363 (1970)

    Google Scholar 

  • Takabayasi, T.: On the formulation of quantum mechanics associated with classical pictures. Progr. Theor. Phys. 8, 143 (1952)

    Google Scholar 

  • Tatarskii, V.I.: The Wigner representation of quantum mechanics. Sov. Phys. Usp. 26, 311 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  • ’t Hooft, G.: Determinism beneath Quantum Mechanics. http://arxiv.org/abs/quant-ph/0212095v1arXiv:quant-ph/0212095v1 (2002)

  • Urbanik, K.: Joint probability distributions of observables in quantum mechanics. Studia Math. 21, 117 (1967)

    MathSciNet  Google Scholar 

  • van Kampen, N.G.: Stochastic differential equations. Phys. Rep. 24, 171 (1976)

    Google Scholar 

  • van Vleck, J.H.: The absorption of radiation by multiply periodic orbits, and its relation to the correspondence principle and the Rayleigh-Jeans law: Part II. Calculation of absorption by multiply periodic orbits. Phys. Rev. 24, 347 (1924)

    Article  ADS  Google Scholar 

  • van Vleck, J.H., Huber, D.L.: Absorption, emission, and linebreadths: a semihistorical perspective. Rev. Mod. Phys. 49, 939 (1977)

    Article  ADS  Google Scholar 

  • Wallstrom, T.C.: On the derivation of the Schrödinger equation from stochastic mechanics. Found. Phys. Lett. 2, 113 (1989)

    Article  MathSciNet  Google Scholar 

  • Wallstrom, T.C.: Inequivalence between the Schrödinger equation and the Madelung hydrodynamic equations. Phys. Rev. A 49, 1613 (1994)

    Article  ADS  MathSciNet  Google Scholar 

  • Wax, N. (ed.): Selected Papers on Noise and Stochastic Processes. Dover, New York (1954/1985)

    Google Scholar 

  • Weinberg, S.: The cosmological constant problem. Rev. Mod. Phys. 61, 1 (1989)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  • Wesson, P.S.: Cosmological constraints on the zero-point electromagnetic field. Astrophys. J. 387, 466 (1991)

    Article  ADS  Google Scholar 

  • Weyl, H.: Quantenmechanik und Gruppentheorie. Zeit. Phys. 46, 1 (1927)

    Google Scholar 

  • Wigner, E.: On the quantum correction for thermodynamic equilibrium. Phys. Rev. 40, 749 (1932)

    Article  ADS  Google Scholar 

  • Zachos, C.K., Fairlie, D.B., Curtright, T.L. (eds.) Quantum Mechanics in Phase Space. World Scientific, Singapore (2005)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ana María Cetto .

Appendices

Appendix A: Derivation of the Generalized Fokker-Planck Equation

In this appendix the generalized Fokker-Planck equation!and laws of evolutionFokker-Planck equation associated with the stochastic Eq. (4.3) is derived, borrowing from de laDe la Peña, L. Cetto, A. M. Peña and Cetto (1977a). For the sake of simplicity the derivation is presented in one dimension. The Fokker-Planck equation!generalizedgeneralization to three dimensions is made at the end of the appendix.

As discussed in the text, for any given realization of the field the density \(R(x,p,t)\) of points in the phase space of the particle satisfies the continuity Eq. (4.12),

$$\begin{aligned} \frac{\partial R}{\partial t}+\frac{1}{m}\frac{\partial }{\partial x}pR+\frac{\partial }{\partial p}\left( f+m\tau \dddot{x}+eE\right) R=0. \end{aligned}$$
(A.1)

The differential equation for \(Q,\) the mean value of \(R\) over the field realizations \(\left\{ (i)\right\} \), can be constructed by means of the smoothing method (see e.g. Frisch 1968)Bharucha-Reid, A. T.,Frisch, U. as follows. A smoothing operator \(\hat{P}\) is introduced, which acts on any phase function \(A(x,p,t)\) by giving its local (in the particle phase space) average,

$$\begin{aligned} \hat{P}A=\overline{A}^{(i)}, \quad \text {so} \quad A=\overline{A}^{(i)}+(1-\hat{P})A. \end{aligned}$$
(A.2)

Clearly \(\delta A=(1-\hat{P})A\) is the random component of \(A,\) which means that the second Equation in (A.2) is a decomposition of \(A\) into its average \(\overline{A}\) plus its fluctuating part. Further, \(\hat{P}=\hat{P}^{2},\) so \(\hat{P}\) is a projection (idempotent) operator. The application of this smoothing operator to the density \(R\) separates it into its average and its random parts \(Q\) and \(\delta Q\), respectively,

$$\begin{aligned} R=Q+\delta Q,\mathrm \quad Q=\hat{P}R,\quad \delta Q=(1-\hat{P})R. \end{aligned}$$
(A.3)

We are interested in constructing the differential equation for \(Q\). For this purpose we rewrite Eq. (A.1) in the form

$$\begin{aligned} \frac{\partial }{\partial t}\left( Q+\delta Q\right) +\hat{L}\left( Q+\delta Q\right) =-e\frac{\partial }{\partial p}E\left( Q+\delta Q\right) , \end{aligned}$$
(A.4)

where \(\hat{L}\) stands for the (nonrandom)Liouville operator for the particle, including the radiation-reaction force \(m\tau \dddot{x}\)—strictly speaking, the operator \(\hat{L}\) differs from a true Liouville operator due to the (small) radiation reaction term—,

$$\begin{aligned} \hat{L}=\frac{1}{m}\frac{\partial }{\partial x}p+\frac{\partial }{\partial p}\left( f+m\tau \dddot{x}\right) . \end{aligned}$$
(A.5)

Equation (A.4) becomes separated into its nonstochastic and fluctuating parts by applying to it the projection operators \(\hat{P}\) and \(1-\hat{P}\) in succession. Using that \(\hat{P}E=0\) [see Eq. (4.4)], one thus obtains the couple of equations

$$\begin{aligned} \left( \frac{\partial }{\partial t}+\hat{L}\right) Q=-e\frac{\partial }{\partial p}\hat{P}E\delta Q, \end{aligned}$$
(A.6)
$$\begin{aligned} \left( \frac{\partial }{\partial t}+\hat{L}\right) \delta Q=-e\frac{\partial }{\partial p}EQ-e\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\delta Q. \end{aligned}$$
(A.7)

The next step is to eliminate \(\delta Q\) from these equations. This can be achieved by introducing the operator \(\hat{G}=(\partial /\partial t+\hat{L})^{-1},\) which corresponds to the Green function of the differential operator \(\partial /\partial t+\hat{L},\) so for any phase function \(A(x,p,t)\) one has

$$\begin{aligned} \hat{G}A(x,p,t)=\int \nolimits _{-\infty }^{t}e^{-\hat{L}\left( t-t^{\prime }\right) }A(x,p,t^{\prime })dt^{\prime }. \end{aligned}$$
(A.8)

The operator \(\hat{G}\) is now used to invert Eq. (A.7),

$$\begin{aligned} \delta Q=-e\hat{G}\frac{\partial }{\partial p}EQ-e\hat{G}\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\delta Q, \end{aligned}$$
(A.9)

or, even better,

$$\begin{aligned} \left[ 1+e\hat{G}\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\right] \delta Q=-e\hat{G}\frac{\partial }{\partial p}EQ. \end{aligned}$$
(A.10)

Applying from the left the inverse of the operator in square brackets gives an expression for \(\delta Q,\) which combined with Eq. (A.6) gives a complicated integro-differential equation for \(Q,\)

$$\begin{aligned} \left( \frac{\partial }{\partial t}+\hat{L}\right) Q=e^{2}\frac{\partial }{\partial p}\hat{P}E\left[ 1+e\hat{G}\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\right] ^{-1}\hat{G}\frac{\partial }{\partial p}EQ. \end{aligned}$$
(A.11)

This is the gfpe for the problem. However, this is a formal expression in which the random field and the operator \(\hat{G}\) appear in a form which makes it quite impractical to use. A more manageable form is obtained by formally expanding the expression within square brackets into a power series,

$$\begin{aligned} \left( \frac{\partial }{\partial t}+\hat{L}\right) Q&=e^{2}\frac{\partial }{\partial p}\hat{P}E\sum _{k=0}^{\infty }\left[ -e\hat{G}\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\right] ^{k}\hat{G}\frac{\partial }{\partial p}EQ \\&=-e\frac{\partial }{\partial p}\hat{P}E\sum _{k=0}^{\infty }\left[ -e\hat{G}\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\right] ^{k+1}Q \end{aligned}$$
$$\begin{aligned}&=-e\frac{\partial }{\partial p}\hat{P}E\sum _{k=1}^{\infty }\left[ -e\hat{G}\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\right] ^{k}Q. \end{aligned}$$
(A.12)

To write the second equality we used the fact that \(EQ\) can also be written as \((1-\hat{P})EQ.\)

To somewhat simplify Eq. (A.12) we take into consideration some statistical properties of the random field. First of all, for the zpf the distribution is symmetric and centered around zero, so that the average of products of an odd number of factors vanishes,

$$\begin{aligned} \hat{P}E(t_{1})E(t_{2})\ldots E(t_{2n+1})\hat{P}A=0. \end{aligned}$$
(A.13a)

Assuming the distribution to be Gaussian, the average of products with an even number of factors take the form Gardiner, C. W.(Gardiner 1983, Sect. 2.8; Wang and Uhlenbeck Uhlenbeck, G. E. Wax, N.in Wax 1954/1985, Sect. 9a)Footnote 26

$$\begin{aligned} \hat{P}E(t_{1})E(t_{2})\ldots E(t_{2n})\hat{P}A=\sum \overline{E(t_{i})E(t_{j})}\ldots \overline{E(t_{r})E(t_{s})}^{(i)}\,\overline{A}^{(i)}, \end{aligned}$$
(A.13b)

where the sum is to be effected over all possible different pairs of factors.

Now from Eq. (A.13a) we note that all terms on the right-hand side of (A.12) with even \(k\) vanish, so the equation simplifies into

$$\begin{aligned} \left( \frac{\partial }{\partial t}+\hat{L}\right) Q=e\frac{\partial }{\partial p}\hat{P}E\sum _{k=0}^{\infty }\left[ e\hat{G}\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\right] ^{2k+1}Q, \end{aligned}$$
(A.14)

which is equivalent to

$$\begin{aligned} \left( \frac{\partial }{\partial t}+\hat{L}\right) Q=e^{2}\frac{\partial }{\partial p}\hat{P}E\hat{G}\frac{\partial }{\partial p}E\sum _{k=0}^{\infty }\left[ e\hat{G}\frac{\partial }{\partial p}\left( 1-\hat{P}\right) E\right] ^{2k}Q. \end{aligned}$$
(A.15)

This is the generalized Fokker-Planck equation!and laws of evolutionFokker-Planck equation (gfpe) that we use here. The generalization to three dimensions is straightforward and gives (summation over repeated indices is to be understood)

$$\begin{aligned} \left( \frac{\partial }{\partial t}+\hat{L}\right) Q=e^{2}\frac{\partial }{\partial p_{i}}\hat{P}E_{i}\hat{G}\frac{\partial }{\partial p_{j}} E_{j}\sum _{k=0}^{\infty }\left[ e\hat{G}\frac{\partial }{\partial p_{l}}\left( 1-\hat{P}\right) E_{l}\right] ^{2k}Q, \end{aligned}$$
(A.16)

where

$$\begin{aligned} \hat{L}=\frac{1}{m}\frac{\partial }{\partial x_{i}}p_{i}+\frac{\partial }{\partial p_{i}}\left( f_{i}+m\tau \dddot{x}_{i}\right) \end{aligned}$$
(A.17)

is the Liouville operator with the radiation reactionRadiation reaction force added as an ‘external’ force. Substitution of this expression into (A.16) gives Eq. (4.13).

The above derivations have been made in terms of operators; a more common form involves the Green function, as follows. The differential equation for the Green function \(\mathcal {G}\) of the Liouville equationLiouville equation is

$$\begin{aligned} \left( \frac{\partial }{\partial t}+L\right) \mathcal {G}=\delta (\varvec{x,x}^{\prime })\delta (\varvec{p,p}^{\prime }), \end{aligned}$$
(A.18)

in terms of which the evolution of a dynamical variable \(A(\varvec{x},\varvec{p},t),\) described above with the aid of the operator \(e^{-\hat{L}\left( t-t^{\prime }\right) },\) is given by the expression

$$\begin{aligned} e^{-\hat{L}(t-t^{\prime })}A(\varvec{x},\varvec{p},t^{\prime })=\int dx^{\prime }dp^{\prime }\mathcal {G}(\varvec{x},\varvec{p};\varvec{x}^{\prime },\varvec{p}^{\prime };t-t^{\prime })A(\varvec{x}^{\prime },\varvec{p}^{\prime },t^{\prime }), \end{aligned}$$
(A.19)

where the prime refers to the values of the dynamical variables at \(t^{\prime }<t,\) subject to the condition \(\left. A(\varvec{x}^{\prime },\varvec{p}^{\prime },t^{\prime })\right| _{t}=A(\varvec{x},\varvec{p},t).\) The evolution from (\(\varvec{x}^{\prime },\varvec{p}^{\prime }\)) to (\(\varvec{x},\varvec{p}\)) is deterministic, as follows from the (modified) Liouville operator (A.17).

Appendix B: Diffusion Coefficients in the Markovian Approximation

We now proceed to the derivation of a simplerDiffusion coefficient!Markovian approximation version of Eq. (4.13), by considering the approximation to second order of the gfpe, known as Approximation!MarkovianMarkovian approximation. The results obtained are suitable to develop the sed theory further, once the system has reached a reversible condition, which is called quantum regime Quantum regimefor the reasons mentioned in Sect. (4.4.4). It is important to remark that between the gfpe with an infinity of terms and the fpe written to order \(n=2\) in the derivatives, there is no intermediate approximation for a positive probability density. In fact, any truncation of the gfpe above \(n=2\) will automatically revert the expansion to second order for a nonnegative \(Q\) (all higher-order coefficients vanish). Thus, there are only three nontrivial possibilities: truncation at \(n=1,\) which corresponds to deterministic (Newtonian) processes (described by the Liouville equationLiouville equation); truncation at \(n=2,\) which corresponds to diffusions (Markovian) processes (described by a true fpe, which may be only approximate); and, finally, no truncation at all, which corresponds to the gfpe of infinite order (A.16). This is the essential content of Pawula’s theorem, a detailed discussion of which can be seen in Risken, H.Risken (1984).

To get the (true, but approximate) fpe (4.19),

$$\begin{aligned}&\frac{\partial Q}{\partial t}+\frac{1}{m}\frac{\partial }{\partial x_{i}}p_{i}Q+\frac{\partial }{\partial p_{i}}\left( f_{i}+m\tau \dddot{x}_{i}\right) Q \end{aligned}$$
$$\begin{aligned}&\quad =\frac{\partial }{\partial p_{i}}D_{ij}^{pp}\frac{\partial Q}{\partial p_{j}}+\frac{\partial }{\partial p_{i}}D_{ij}^{px}\frac{\partial Q}{\partial x_{j}}, \end{aligned}$$
(B.1)

we start from the gfpe (4.13). The first step is to write Eq. (4.14) to first order in \(e^{2}\) (i.e., to take only the term of the sum with \(k=0\)), which is

$$\begin{aligned} \left. \mathcal {\hat{D}}_{i}Q\right| _{k=0}=\hat{P}E_{i}\hat{G}\frac{\partial }{\partial p_{j}}E_{j}Q=\hat{P}E_{i}(t)\int \nolimits _{-\infty }^{t}dt^{\prime }e^{-\hat{L}\left( t-t^{\prime }\right) }\frac{\partial }{\partial p_{j}}(E_{j}Q)(t^{\prime }), \end{aligned}$$
(B.2)

where Eq. (4.15) has been used. Since \(E_{j}\) depends on time only, this can be written in the form

$$\begin{aligned} \left. \mathcal {\hat{D}}_{i}Q\right| _{k=0}=\mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\overline{E_{i}(t)E_{j}(t^{\prime })}^{(i)}e^{-\hat{L}\left( t-t^{\prime }\right) }\frac{\partial }{\partial p_{j}}Q(t^{\prime }). \end{aligned}$$
(B.3)

Now, the evolution law

$$\begin{aligned} Q(t)=e^{-\hat{L}\left( t-t^{\prime }\right) }Q(t^{\prime }) \end{aligned}$$
(B.4)

allows us to write, inserting the identity operator \(e^{\hat{L}\left( t-t^{\prime }\right) }e^{-\hat{L}\left( t-t^{\prime }\right) },\)

$$\begin{aligned} e^{-\hat{L}\left( t-t^{\prime }\right) }\frac{\partial }{\partial p_{j}}Q(t^{\prime })&=\left( e^{-\hat{L}\left( t-t^{\prime }\right) }\frac{\partial }{\partial p_{j}}e^{\hat{L}\left( t-t^{\prime }\right) }\right) \left( e^{-\hat{L}\left( t-t^{\prime }\right) }Q(t^{\prime })\right) \end{aligned}$$
$$\begin{aligned}&=\frac{\partial }{\partial p_{j}^{\prime }}Q(t), \end{aligned}$$
(B.5)

where, as explained in Appendix A, the prime refers to the values of the dynamical variables at \(t^{\prime }\le t,\) from where they follow a deterministic evolution to their end values at time \(t\). Substitution into Eq. (B.3) using Eq. (4.9) gives

$$\begin{aligned} \left. \mathcal {\hat{D}}_{i}Q\right| _{k=0}=\mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\delta _{ij}\frac{\partial }{\partial p_{j}^{\prime }}Q(x,p,t). \end{aligned}$$
(B.6)

From the chain rule for the derivation,

$$\begin{aligned} \frac{\partial Q}{\partial p_{j}^{\prime }}=\frac{\partial p_{k}}{\partial p_{j}^{\prime }}\frac{\partial Q}{\partial p_{k}}+\frac{\partial x_{k}}{\partial p_{j}^{\prime }}\frac{\partial Q}{\partial x_{k}}, \end{aligned}$$
(B.7)

it follows that

$$\begin{aligned} \left. \mathcal {\hat{D}}_{i}Q\right| _{k=0}=\left( \mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\frac{\partial p_{j}}{\partial p_{i}^{\prime }}\right) \frac{\partial Q}{\partial p_{j}}+\left( \mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\frac{\partial x_{j}}{\partial p_{i}^{\prime }}\right) \frac{\partial Q}{\partial x_{j}}. \end{aligned}$$
(B.8)

Direct substitution into (4.13) gives the (approximate) equation

$$\begin{aligned} \frac{\partial Q}{\partial t}+\frac{1}{m}\frac{\partial }{\partial x_{i}}p_{i}Q+\frac{\partial }{\partial p_{i}}f_{i}Q+m\tau \frac{\partial }{\partial p_{i}}\dddot{x}_{i}Q=\frac{\partial }{\partial p_{i}}\left( D_{ij}^{pp}\frac{\partial Q}{\partial p_{j}}+D_{ij}^{px}\frac{\partial Q}{\partial x_{j}}\right) , \end{aligned}$$
(B.9)

with the diffusion coefficients in the Markovian approximation Markovian approximationgiven by

$$\begin{aligned} D_{ij}^{pp}&=e^{2}\mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\frac{\partial p_{j}}{\partial p_{i}^{\prime }},\end{aligned}$$
(B.10a)
$$\begin{aligned} D_{ij}^{px}&=e^{2}\mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\frac{\partial x_{j}}{\partial p_{i}^{\prime }}. \end{aligned}$$
(B.10b)

Now, the dominant contribution to these time integrals comes from times \(t^{\prime }\) close to \(t\), when the system is already in the regime in which Eq. (4.78) controls the dynamics. This means that the dynamical variables are now represented by operators (see Sect. 4.4.2), and the factors \(\partial p_{j}/\partial p_{i}^{\prime },\) \(\partial x_{j}/\partial p_{i}^{\prime }\) should be properly expressed in terms of them. By noticing that for the corresponding classical variables the equalities

$$\begin{aligned} \frac{\partial p_{j}}{\partial p_{i}^{\prime }}&=\left[ x_{i}^{\prime },p_{j}\right] _{\text {PB}},\end{aligned}$$
(B.11a)
$$\begin{aligned} \frac{\partial x_{j}}{\partial p_{i}^{\prime }}&=\left[ x_{i}^{\prime },x_{j}\right] _{\text {PB}} \end{aligned}$$
(B.11b)

apply, in terms of operators the following substitutions must be made,

$$\begin{aligned} \frac{\partial p_{j}}{\partial p_{i}^{\prime }}&\Rightarrow \frac{1}{2i\eta }\left[ \hat{x}_{i}^{\prime },\hat{p}_{j}\right] ,\end{aligned}$$
(B.12a)
$$\begin{aligned} \frac{\partial x_{j}}{\partial p_{i}^{\prime }}&\Rightarrow \frac{1}{2i\eta }\left[ \hat{x}_{i}^{\prime },\hat{x}_{j}\right] . \end{aligned}$$
(B.12b)

These can be obtained from the general rules

$$\begin{aligned} \frac{\partial }{\partial p_{j}}\Rightarrow \frac{1}{2i\eta }\left[ \hat{x}_{j},\quad \right] ,\quad \frac{\partial }{\partial x_{j}}\Rightarrow \frac{1}{2i\eta }\left[ \quad ,\hat{p}_{j}\right] . \end{aligned}$$
(B.13)

With (B.12a, B.12b), Eqs. (B.10a), (B.10b) take the form

$$\begin{aligned} D_{ij}^{pp}&=\frac{e^{2}}{2i\eta }\mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\left[ \hat{x}_{i}^{\prime },\hat{p}_{j}\right] ,\end{aligned}$$
(B.14a)
$$\begin{aligned} D_{ij}^{px}&=\frac{e^{2}}{2i\eta }\mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\left[ \hat{x}_{i}^{\prime },\hat{x}_{j}\right] . \end{aligned}$$
(B.14b)

It is now straightforward to show that the diffusion coefficients [Eqs. (B.10a), (B.10b)] comply with the relation

$$\begin{aligned} \frac{\partial D_{ij}^{pp}}{\partial p_{j}}+\frac{\partial D_{ij}^{px}}{\partial x_{j}}=0. \end{aligned}$$
(B.15)

For this purpose we apply Eqs. (B.13) to (Eqs. B.14a), (B.14b), thus obtaining

$$\begin{aligned} \frac{\partial D_{ij}^{pp}}{\partial p_{j}}+\frac{\partial D_{ij}^{px}}{\partial x_{j}}=\frac{e^{2}}{(2i\eta )^{2}}\mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\left\{ \left[ \hat{x}_{j},\left[ \hat{x}_{i}^{\prime },\hat{p}_{j}\right] \right] +\left[ \left[ \hat{x}_{i}^{\prime },\hat{x}_{j}\right] ,\hat{p}_{j}\right] \right\} . \end{aligned}$$
(B.16)

Resorting to the Jacobi identity (which is valid both for Poisson brackets and for commutators)

$$\begin{aligned} \left[ x_{j},\left[ x_{i}^{\prime },p_{j}\right] \right] +\left[ x_{i}^{\prime },\left[ p_{j},x_{j}\right] \right] +\left[ p_{j},\left[ x_{j},x_{i}^{\prime }\right] \right] =0, \end{aligned}$$
(B.17)

and noticing that the second term vanishes, we find

$$\begin{aligned} \left[ x_{j},\left[ x_{i}^{\prime },p_{j}\right] \right] =-\left[ p_{j},\left[ x_{j},x_{i}^{\prime }\right] \right] =-\left[ \left[ x_{i}^{\prime },x_{j}\right] ,p_{j}\right] , \end{aligned}$$
(B.18)

and therefore the right-hand side of (B.16) is null, which proves Eq. (B.15). A direct consequence of this result is that Eq. (4.44) reduces to (4.45).

Appendix C: Detailed Derivation of the ‘Generalized’ Schrödinger Equation

The starting point for the present derivation is the couple of Eq. (4.68)

$$\begin{aligned}&\frac{\partial \rho }{\partial t}+\frac{\partial }{\partial x_{j}}\left( v_{j}\rho \right) =0,\end{aligned}$$
(C.1a)
$$\begin{aligned}&m\frac{\partial }{\partial t}\!\left( v_{i}\rho \right) +m\frac{\partial }{\partial x_{j}}\!\left( v_{i}v_{j}\rho \right) \!-\!\frac{\eta ^{2}}{m}\frac{\partial }{\partial x_{j}}\!\left( \!\rho \frac{\partial ^{2}}{\partial x_{i}\partial x_{j}}\ln \rho \!\!\right) \!+\frac{1}{m}\frac{\partial }{\partial x_{j}}\Sigma _{ij}\rho -\!f_{i}\rho \\&=\tau v_{j}\frac{\partial f_{i}}{\partial x_{j}}\rho -e^{2}\!\left. (\widetilde{\mathcal {\hat{D}}Q})_{i}\right| _{\varvec{z}=0}. \end{aligned}$$
(C.1b)

Inserting the continuity equation into (C.1b) and multiplying by \(\rho ^{-1}\) results in

$$\begin{aligned} m\frac{\partial \varvec{v}}{\partial t}\!\!+\!\frac{m}{2}\varvec{\nabla v}^{2}-m\varvec{v}\times \left( \varvec{\nabla }\times \varvec{v}\right) +\varvec{\nabla }\left( -\frac{2\eta ^{2}}{m}\frac{\varvec{\nabla }^{2}\sqrt{\rho }}{\sqrt{\rho }}+V\right) =\varvec{F}_{\text {rad}}+\varvec{F}_{\Sigma }, \end{aligned}$$
(C.2)

since \(\left( \varvec{v}\cdot \varvec{\nabla }\right) \varvec{v}=\tfrac{1}{2}\varvec{\nabla v}^{2}-\varvec{v}\times \left( \varvec{\nabla }\times \varvec{v}\right) \). The \(i\) components of the vectors \(\varvec{F}_{\text {rad}}\) and \(\varvec{F}_{\Sigma }\) are, respectively,

$$\begin{aligned} F_{i\text {rad}}&=\tau v_{j}\frac{\partial f_{i}}{\partial x_{j}}-\frac{e^{2}\!}{\rho }\left. (\widetilde{\mathcal {\hat{D}}Q})_{i}\right| _{\varvec{z}=0},\end{aligned}$$
(C.3)
$$\begin{aligned} F_{i\Sigma }&=-\frac{1}{m\rho }\frac{\partial }{\partial x_{j}}\!\Sigma _{ji}\rho . \end{aligned}$$
(C.4)

Now, according to Eq. (4.66a) \(m\varvec{v}\) decomposes as

$$\begin{aligned} m\varvec{v}=-i\eta \varvec{\nabla }\ln \frac{q(\varvec{x},t)}{q^{*}(\varvec{x},t)}+\varvec{g}=2\eta \varvec{\nabla }S+\varvec{g}, \end{aligned}$$
(C.5)

where the last equality follows from writing \(q(\varvec{x},t)\) in its polar form

$$\begin{aligned} q(\varvec{x},t)=\sqrt{\rho }e^{iS(\varvec{x},t)}. \end{aligned}$$
(C.6)

Substitution of Eqs. (C.5) into (C.2) gives (using \(\varvec{\nabla }\times m\varvec{v}=\varvec{\nabla }\times \varvec{g})\)

$$\begin{aligned} \!\!\varvec{\nabla }M=\varvec{F}_{\text {rad}}+\varvec{F}_{\Sigma }-\frac{\partial \varvec{g}}{\partial t}+\varvec{v}\times \left( \varvec{\nabla }\times \varvec{g}\right) , \end{aligned}$$
(C.7)

with

$$\begin{aligned} M=2\eta \frac{\partial S}{\partial t}+\frac{1}{2}m\varvec{v}^{2}-\frac{2\eta ^{2}}{m}\frac{\varvec{\nabla }^{2}\sqrt{\rho }}{\sqrt{\rho }}+V. \end{aligned}$$
(C.8)

This is basically the Hamilton-Jacobi-type equation of Bohm’s theory [see Sect. (4.4.1)], when \(M=0\) (and \(\eta =\hbar /2\)), which means that it should be possible to arrive at the Schrödinger equation from the above expressions. For this purpose we proceed as follows.

From Eq. (C.6) it follows \(\ln q=(1/2)\ln \rho +iS\), which combined with the continuity Eq. (C.1a) gives

$$\begin{aligned} \frac{\partial S}{\partial t}=-\frac{i}{q}\frac{\partial q}{\partial t}-\frac{i}{2}\nabla \cdot \varvec{v}-i\varvec{v}\cdot \frac{\varvec{\nabla }q}{q}-\varvec{v}\cdot \varvec{\nabla }S. \end{aligned}$$
(C.9)

Using here Eq. (C.5) leads to

$$\begin{aligned} \frac{\partial S}{\partial t}&=-\frac{i}{q}\frac{\partial q}{\partial t}-\frac{i\eta }{m}\varvec{\nabla }^{2}S-\frac{i}{2m}\nabla \cdot \varvec{g}-\frac{2i\eta }{m}\varvec{\nabla }S\cdot \frac{\varvec{\nabla }q}{q}- \end{aligned}$$
$$\begin{aligned}&\quad -\frac{i}{m}\varvec{g}\cdot \frac{\varvec{\nabla }q}{q}-\frac{2\eta }{m}\left( \varvec{\nabla }S\right) ^{2}-\frac{1}{m}\varvec{g}\cdot \varvec{\nabla }S. \end{aligned}$$
(C.10)

This expression, together with \(m\varvec{v}^{2}=(1/m)\left( 2\eta \varvec{\nabla }S+\varvec{g}\right) ^{2}\), allows to recast Eq. (C.8) in the form

$$\begin{aligned} M&=\frac{1}{q}\left[ -2i\eta \frac{\partial q}{\partial t}+\frac{1}{2m}\left( -2i\eta \varvec{\nabla }+\varvec{g}\right) ^{2}q+Vq\right] -\\&\quad -\frac{2\eta ^{2}}{m}\left[ i\varvec{\nabla }^{2}S+\frac{\varvec{\nabla }^{2}\sqrt{\rho }}{\sqrt{\rho }}+\left( \varvec{\nabla }S\right) ^{2}+2i\varvec{\nabla }S\cdot \frac{\varvec{\nabla }q}{q}-\frac{\varvec{\nabla }^{2}q}{q}\right] . \end{aligned}$$
(C.11)

Now we notice that as a consequence of Eq. (C.6),

$$\begin{aligned} \frac{\varvec{\nabla }q}{q}=\frac{1}{2}\frac{\varvec{\nabla }\rho }{\rho }+i\varvec{\nabla }S, \end{aligned}$$
(C.12)

whence

$$\begin{aligned} \frac{\varvec{\nabla }^{2}q}{q}=\frac{\varvec{\nabla }^{2}\sqrt{\rho }}{\sqrt{\rho }}+i\varvec{\nabla }^{2}S+i\varvec{\nabla }S\cdot \frac{\varvec{\nabla }\rho }{\rho }-\left( \varvec{\nabla }S\right) ^{2}. \end{aligned}$$
(C.13)

From here it follows that the term in the second line of Eq. (C.11) vanishes. Consequently \(M\) reduces to

$$\begin{aligned} M=\frac{1}{q}\left[ -2i\eta \frac{\partial q}{\partial t}+\frac{1}{2m}\left( -2i\eta \varvec{\nabla }+\varvec{g}\right) ^{2}q+Vq\right] =\frac{1}{ q}\hat{M}q, \end{aligned}$$
(C.14)

with \(\hat{M}\) the linear operator

$$\begin{aligned} \hat{M}=-2i\eta \frac{\partial }{\partial t}+\frac{1}{2m}\left( -2i\eta \varvec{\nabla }+\varvec{g}\right) ^{2}+V. \end{aligned}$$
(C.15)

Finally, inserting Eq. (C.14) into (C.7) we arrive at

$$\begin{aligned} \!\!\varvec{\nabla }\left( \frac{1}{q}\hat{M}q\right) =\varvec{F}_{\text {rad}}+\varvec{F}_{\Sigma }-\frac{\partial \varvec{g}}{\partial t}+\varvec{v}\times \left( \varvec{\nabla }\times \varvec{g}\right) . \end{aligned}$$
(C.16)

We call this the generalized Schrödinger equation, since as is shown in Sect. 4.4.1, it reduces to the Schrödinger equation in the radiationless approximationRadiationless approximation. Equation (C.16) and its adjoint are equivalent to the first two equations of the hierarchy, (C.1a, C.1b).

Appendix D: Diffusive Contribution to the Energy Balance

This appendix is devoted to the calculation of theDetailed balance!frequency mixing right-hand sideDetailed balance!condition of the energy-balance condition (4.37). The particle is considered in equilibrium with the zpf, which means that it must be in its ground state, represented by the solution \(\psi _{0}(x)\) of the Schrödinger equation (4.78) (still in terms of \(\eta \)). We recall that in the time-asymptotic limit, the Markovian approximation holds, described by the fpe (B.1). This means that one may use the simpler Eq. (4.103), which in one dimension reads

$$\begin{aligned} \tau \left\langle \dddot{x}p\right\rangle _{0}=-\frac{1}{m}\left\langle D^{pp}\right\rangle _{0}. \end{aligned}$$
(D.1)

Introducing the diffusion coefficient \(D^{pp}\) given by Eq. (B.14a), one obtains

$$\begin{aligned} -\frac{1}{m}\left\langle D^{pp}\right\rangle _{0}=\frac{ie^{2}}{2\eta m}\mathop {\displaystyle \int }\nolimits _{-\infty }^{t}dt^{\prime }\varphi (t-t^{\prime })\left\langle \left[ \hat{x}_{i}^{\prime },\hat{p}_{j}\right] \right\rangle _{0}, \end{aligned}$$
(D.2)

With \(\varphi (t-t^{\prime })\) given by (4.10) and \(\rho =\rho _{0}\) given by (4.101), this becomes (recall that \(\tau =2e^{2}/3mc^{3}\))

$$\begin{aligned} -\frac{1}{m}\left\langle D^{pp}\right\rangle _{0}=\frac{i\hbar \tau }{2\eta \pi }\mathop {\displaystyle \int }_{0}^{\infty }d\omega \ \omega ^{3}\int \nolimits _{-\infty }^{t}dt^{\prime }\cos \omega (t-t^{\prime })\ \left\langle \left[ \hat{x}_{i}^{\prime },\hat{p}_{j}\right] \right\rangle _{0}, \end{aligned}$$
(D.3)

with

$$\begin{aligned} \left\langle \left[ \hat{x}^{\prime },\hat{p}\right] \right\rangle _{0}=\sum _{k}\left( \hat{x}_{0k}^{\prime }\hat{p}_{k0}-\hat{p}_{0k}\hat{x}_{k0}^{\prime }\right) , \end{aligned}$$
(D.4)

Equations (4.104) and (4.105) can be used to write \(x_{kn}(t)=e^{i\omega _{kn}t}x_{kn}\) and \(p_{nm}(t)=im\omega _{nm}e^{i\omega _{nm}t}x_{nm},\) which gives

$$\begin{aligned} \left\langle \left[ \hat{x}^{\prime },\hat{p}\right] \right\rangle _{0}=2im\sum _{k}\omega _{k0}\left| x_{0k}\right| ^{2}\cos \omega _{k0}(t-t^{\prime }). \end{aligned}$$
(D.5)

Introducing this into Eq. (D.3) leads to

$$\begin{aligned} -\frac{1}{m}\left\langle D^{pp}\right\rangle _{0}&=-\frac{\hbar m\tau }{\pi \eta }\sum _{k}\omega _{k0}\left| x_{0k}\right| ^{2}\\&\times \mathop {\displaystyle \int }_{0}^{\infty }d\omega \ \omega ^{3}\int \nolimits _{-\infty }^{t}dt^{\prime }\cos \omega (t-t^{\prime })\cos \omega _{k0}(t-t^{\prime }). \end{aligned}$$
(D.6)

The integral over time can be calculated with the formula

$$\begin{aligned} \int \nolimits _{-\infty }^{\infty }dke^{ikx}=\int \nolimits _{-\infty }^{\infty }dk\cos kx=2\pi \delta (x). \end{aligned}$$
(D.7)

This gives

$$\begin{aligned} \int \nolimits _{-\infty }^{t}dt^{\prime }\cos \omega (t-t^{\prime })\cos \omega _{k0}(t-t^{\prime })=\frac{\pi }{2}[\delta (\omega -\omega _{k0})+\delta (\omega +\omega _{k0})]. \end{aligned}$$
(D.8)

For the ground state there are no negative frequencies, i.e. all \(\omega _{k0}>0,\) whence the second term in the right-hand side does not contribute to (D.6), and therefore

$$\begin{aligned} -\frac{1}{m}\left\langle D^{pp}\right\rangle _{0}=-\frac{\hbar m\tau }{2\eta }\sum _{k}\omega _{k0}^{4}\left| x_{0k}\right| ^{2}. \end{aligned}$$
(D.9)

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

de la Peña, L., Cetto, A.M., Valdés Hernández, A. (2015). The Long Journey to the Schrödinger Equation. In: The Emerging Quantum. Springer, Cham. https://doi.org/10.1007/978-3-319-07893-9_4

Download citation

Publish with us

Policies and ethics