Skip to main content
Log in

Discrete Excitation Spectrum of a Classical Harmonic Oscillator in Zero-Point Radiation

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

We report that upon excitation by a single pulse, a classical harmonic oscillator immersed in the classical electromagnetic zero-point radiation exhibits a discrete harmonic spectrum in agreement with that of its quantum counterpart. This result is interesting in view of the fact that the vacuum field is needed in the classical calculation to obtain the agreement.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. The mass value is chosen to keep the integration time manageable without losing the physical characteristics of the problem.

References

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

    Article  ADS  Google Scholar 

  2. Boyer, T.H.: General connection between random electrodynamics and quantum electrodynamics for free electromagnetic fields and for dipole oscillator systems. Phys. Rev. D 11, 809 (1975)

    Article  ADS  Google Scholar 

  3. Milonni, P.W.: The Quantum Vacuum: An Introduction to Quantum Electrodynamics. Academic Press, Boston (1994)

    Google Scholar 

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

    Google Scholar 

  5. Cetto, A.M., de la Peña, L.: The Quantum Dice, an Introduction to Stochastic Electrodynamics. Kluwer, Dordrecht (1996)

    Google Scholar 

  6. Boyer, T.H.: Asymptotic retarded van der Waals forces derived from classical electrodynamics with classical electromagnetic zero-point radiation. Phys. Rev. A 5, 1799 (1972)

    Article  ADS  Google Scholar 

  7. Dressel, J., Bliokh, K.Y., Nori, F.: Classical field approach to quantum weak measurements. Phys. Rev. Lett. 112, 110407 (2014)

    Article  ADS  Google Scholar 

  8. Milonni, P.W.: The Quantum Vacuum: An Introduction to Quantum Electrodynamics. Academic Press, Boston (1994)

    Google Scholar 

  9. P. W. Milonni, The Quantum Vacuum: An Introduction to Quantum Electrodynamics, pp. 51–54, 123–128, 487–488. Academic Press, Boston (1994)

  10. Griffiths, D.J.: Introduction to Electrodynamics, 3rd edn. Prentice-Hall, Upper Saddle River (1999)

    Google Scholar 

  11. Landau, L.D., Lifshitz, E.M.: The Classical Theory of Fields, 4th edn, p. 207. Pergamon Press, New York (1987). (Eq. 75.10)

  12. Jackson, J.D.: Classical Electrodynamics, 3rd edn, p. 749. Wiley, New York (1998). (Eq. 16.10)

    Google Scholar 

  13. Huang, W., Batelaan, H.: Dynamics underlying the Gaussian distribution of the classical harmonic oscillator in zero-point radiation. J. Comput. Methods Phys. 2013, 308538 (2013)

    Article  Google Scholar 

  14. Thornton, S.T., Marion, J.B.: Classical Dynamics of Particles and Systems, 5th edn, pp. 117–128. Brooks/Cole, Belmont (2004)

    Google Scholar 

  15. Landau, L.D., Lifshitz, E.M.: Mechanics, 3rd edn, p. 80. Butterworth-Heinemann, Oxford (1976)

    Google Scholar 

  16. Yariv, A.: Quantum Electronics, 3rd edn, p. 407. Wiley, New York (1988)

    Google Scholar 

  17. Heisenberg, W.: Über quantentheoretische Umdeutung kinematischer und mechanischer Besiehungen. Z. Phys. 33, 879 (1925)

    Article  ADS  MATH  Google Scholar 

  18. Aitchison, I.J.R., MacManus, D.A., Snyder, T.M.: Understanding Heisenbergs magical paper of July 1925: a new look at the calculational details. Am. J. Phys. 72, 11 (2004)

    Article  MathSciNet  Google Scholar 

  19. Averbukh, V., Moiseyev, N.: Classical versus quantum harmonic-generation spectrum of a driven anharmonic oscillator in the high-frequency regime. Phys. Rev. A 57, 1345 (1998)

    Article  ADS  Google Scholar 

  20. Griffiths, D.J.: Introduction to Quantum Mechanics, 2nd edn, p. 366. Upper Saddle River, Pearson Prentice Hall (2005)

    Google Scholar 

Download references

Acknowledgments

We gratefully acknowledge comments from Prof. Peter W. Milonni. The funding support comes from NSF Grant No. 0969506. This work was completed utilizing the Holland Computing Center of the University of Nebraska and the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by NSF Grant No. OCI-1053575.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Herman Batelaan.

Appendices

The Derivation of \(\langle W_{1\omega _{0}} \rangle \)

In this appendix, the value of \(\langle W_{1\omega _{0}} \rangle \) is calculated. This expounds the steps leading from Eq. (36) to Eq. (37). We will focus here on \(\langle W_{1\omega _{0}} \rangle \) and indicate how the steps are different in notation for \(\langle W_{2\omega _{0}} \rangle \) and \(\langle W_{3\omega _{0}} \rangle \). In Eq. (36), the energy change at drive frequency \(\omega _{p} \simeq 1\omega _{0}\) is

$$\begin{aligned} W_{1\omega _{0}} = \frac{\sqrt{\pi }}{2} (D_{0} \omega _{0})(qE_{0}\Delta t) \times \exp {\left[ - \left( \frac{\omega _{p} - \omega _{0}}{2/\Delta t} \right) ^2 \right] } \cos {(\omega _{0}t_{c} + \varphi _{0})}. \end{aligned}$$
(39)

The initial conditions come in by substituting \(D_{0}\) and \(\varphi _{0}\), using

$$\begin{aligned} \cos {(\varphi _{0})}&= (x_{0}-x_{1p}(0))/D_{0}, \nonumber \\ \sin {(\varphi _{0})}&= -(v_{0}-v_{1p}(0))/D_{0}\omega _{0}, \end{aligned}$$
(40)

as defined in Eq. (26), and expanding \(\cos {(\omega _{0}t_{c} + \varphi _{0})}\) to \(\cos {(\omega _{0}t_{c})}\cos {(\varphi _{0})}-\sin {(\omega _{0}t_{c})}\sin {(\varphi _{0})}\) so that

$$\begin{aligned} D_{0}\cos {(\omega _{0}t_{c} + \varphi _{0})} = \cos {(\omega _{0}t_{c})}(x_{0}-x_{1p}(0)) + \sin {(\omega _{0}t_{c})}(v_{0}-v_{1p}(0))/\omega _{0}. \end{aligned}$$
(41)

The energy change \(W_{1\omega _{0}}\) now depends on \(x_{0}\), \(v_{0}\), \(x_{1p}(0)\), and \(v_{1p}(0)\). Similar procedure is also used to evaluate \(\langle W_{2\omega _{0}} \rangle \) and \(\langle W_{3\omega _{0}}\rangle \), where the sinusoidal functions are expanded and substitutions are made using Eqs. (30) and (34),

$$\begin{aligned}&\cos {(\phi _{0})} = x_{0}/A_{0}, \nonumber \\&\sin {(\phi _{0})} = -v_{0}/A_{0}\omega _{0}, \nonumber \\&\cos {(\xi _{0})} = -{x_{2p}^{(1)}(0)}/B_{0}, \nonumber \\&\sin {(\xi _{0})} = {v_{2p}^{(1)}(0)}/B_{0}\omega _{0}, \nonumber \\&\cos {(\zeta _{0})} = -{x_{3p}^{(1)}(0)}/C_{0}, \nonumber \\&\sin {(\zeta _{0})} = {v_{3p}^{(1)}(0)}/C_{0}\omega _{0}. \end{aligned}$$
(42)

To compute the ensemble average of Eq. (39), the values of \(\langle x_{0}\rangle \), \(\langle v_{0}\rangle \), \(\langle x_{1p}(0) \rangle \), and \(\langle v_{1p}(0) \rangle \) are needed. The statistical moments, \(\langle x_{0}\rangle \) and \(\langle v_{0}\rangle \), can be evaluated given the stationary state of the harmonic oscillator in the vacuum field [2],

$$\begin{aligned} \langle x_{0}\rangle = 0, \quad \langle v_{0}\rangle = 0. \end{aligned}$$
(43)

As the particular solution (\(x_{1p}(0)\) and \(v_{1p}(0)\)) does not depend on \(x_{0}\) or \(v_{0}\), its ensemble average is equal to itself,

$$\begin{aligned} \langle x_{1p}(0) \rangle = x_{1p}(0), \nonumber \\ \langle v_{1p}(0) \rangle = v_{1p}(0). \end{aligned}$$
(44)

The particular solutions are evaluated at \(t=0\) according to Eq. (25),

$$\begin{aligned} x_{1p}(0)&= -\int _{0}^{\infty } \!d\omega f_{1}(\omega ) \sin {(-\omega t_{c})}, \nonumber \\ v_{1p}(0)&= -\int _{0}^{\infty } \!d\omega f_{1}(\omega ) \omega \cos {(-\omega t_{c})}, \end{aligned}$$
(45)

where

$$\begin{aligned} f_{1}(\omega ) \equiv \frac{\Delta t}{2\sqrt{\pi }} \frac{qE_{0}}{m(\omega _{0}^2-\omega ^2)} \exp { \left[ -\left( \frac{\omega -\omega _{p}}{2/\Delta t}\right) ^2 \right] }. \end{aligned}$$
(46)

Using the change of variables, \(u \equiv \omega /\omega _{0}\), \(u_{p} \equiv \omega _{p}/\omega _{0}\), \(\Delta u \equiv 2/\omega _{0}\Delta t\), and \(\kappa \equiv \omega _{0}t_{c}\), the particular solutions in Eq. (45) can be written as

$$\begin{aligned} x_{1p}(0)&= \left( \frac{qE_{0}\Delta t}{m\omega _{0}}\frac{1}{2\sqrt{\pi }} \right) D_{1}, \nonumber \\ v_{1p}(0)&= \left( \frac{-qE_{0}\Delta t}{m}\frac{1}{2\sqrt{\pi }} \right) D_{2}, \end{aligned}$$
(47)

where

$$\begin{aligned} D_{1}&\equiv \int _{0}^{\infty } \!du \frac{1}{1-u^{2}} \exp { \left[ -\left( \frac{u-u_{p}}{\Delta u}\right) ^2 \right] } \sin {(\kappa u)}, \nonumber \\ D_{2}&\equiv \int _{0}^{\infty } \!du \frac{u}{1-u^{2}} \exp { \left[ -\left( \frac{u-u_{p}}{\Delta u}\right) ^2 \right] } \cos {(\kappa u)}. \end{aligned}$$
(48)

Therefore, the ensemble average of Eq. (39) is

$$\begin{aligned} \langle W_{1\omega _{0}} \rangle&= \frac{\sqrt{\pi }}{2} \omega _{0}(qE_{0}\Delta t) \exp {\left[ - \left( \frac{\omega _{p} - \omega _{0}}{2/\Delta t} \right) ^2 \right] } \langle D_{0}\cos {(\omega _{0}t_{c} + \varphi _{0})} \rangle \nonumber \\&= \frac{-1}{4} \frac{(qE_{0}\Delta t)^{2}}{m} \exp {\left[ - \left( \frac{\omega _{p} - \omega _{0}}{2/\Delta t} \right) ^2 \right] } \left( D_{1}\cos {(\kappa )} - D_{2}\sin {(\kappa )} \right) .\qquad \quad \end{aligned}$$
(49)

The integral \(D_{1}\cos {(\kappa )} - D_{2}\sin {(\kappa )}\) can be further evaluated,

$$\begin{aligned}&D_{1}\cos {(\kappa )} - D_{2}\sin {(\kappa )} \nonumber \\&= \int _{0}^{\infty } \!du \frac{1}{1-u^{2}} \exp { \left[ -\left( \frac{u-u_{p}}{\Delta u}\right) ^2 \right] } \left( \sin {(\kappa u)}\cos {(\kappa )} - u\cos {(\kappa u)}\sin {(\kappa )} \right) \nonumber \\&= \int _{0}^{\infty } \!du \frac{1}{2(1+u)} \exp { \left[ -\left( \frac{u-u_{p}}{\Delta u}\right) ^2 \right] } \sin {(\kappa (u+1))} \nonumber \\&\quad +\,\int _{0}^{\infty } \!du \frac{1}{2(1-u)} \exp {\left[ -\left( \frac{u-u_{p}}{\Delta u}\right) ^2 \right] } \sin {(\kappa (u-1))}. \end{aligned}$$
(50)

Because \(u_{p} \simeq 1\), the first term

$$\begin{aligned} \int _{0}^{\infty } \!du \frac{1}{2(1+u)} \exp { \left[ -\left( \frac{u-u_{p}}{\Delta u}\right) ^2 \right] } \sin {(\kappa (u+1))} \end{aligned}$$
(51)

is effectively zero. Let \(u_{p} = 1-\epsilon \Delta u\), where \(\epsilon \ll 1 \) is a small number, we obtain

$$\begin{aligned} D_{1}\cos {(\kappa )} - D_{2}\sin {(\kappa )} = \frac{-1}{2}\int _{0}^{\infty } \!du \frac{\sin {(\kappa (u-1))}}{u-1} \exp { \left[ -\left( \frac{u-1}{\Delta u} + \epsilon \right) ^2 \right] }. \end{aligned}$$
(52)

By the change of variables \(x \equiv (u-1)/\Delta u\) and \(\alpha \equiv \kappa \Delta u\), the above integral can be rewritten as

$$\begin{aligned} \int _{0}^{\infty } \!du \frac{\sin {(\kappa (u\!-\!1))}}{u\!-\!1} \exp { \left[ \!-\!\left( \frac{u\!-\!1}{\Delta u} \!+\! \epsilon \right) ^2 \right] } = \int _{-1/\Delta u}^{\infty } \! dx \frac{\sin {(\alpha x)}}{x} \exp { \left[ -\left( x \!+\! \epsilon \right) ^2 \right] }. \end{aligned}$$
(53)

Because the width of the integrand is much smaller than the lower integral limit, \(\pi /\alpha \ll 1/\Delta u\), the integral can be approximated by extending the lower limit to the negative infinity,

$$\begin{aligned} \int _{-1/\Delta u}^{\infty } \! dx \frac{\sin {(\alpha x)}}{x} \exp { \left[ -\left( x + \epsilon \right) ^2 \right] } \simeq \int _{-\infty }^{+\infty } \!\! dx \frac{\sin {(\alpha x)}}{x} \exp { \left[ -\left( x + \epsilon \right) ^2 \right] }, \end{aligned}$$
(54)

which can be written in the complex form,

$$\begin{aligned} \int _{-\infty }^{+\infty } \!\! dx \frac{\sin {(\alpha x)}}{x} \exp { \left[ -\left( x + \epsilon \right) ^2 \right] } = \mathrm Im \left( \int _{-\infty }^{+\infty } \!\! dx \frac{e^{i\alpha x}}{x} \exp { \left[ -\left( x + \epsilon \right) ^2 \right] } \right) . \end{aligned}$$
(55)

We will use the contour integral to evaluate this complex integral,

$$\begin{aligned} \oint _{C} \! dz \frac{e^{i\alpha z}}{z} \exp { \left[ -\left| z + \epsilon \right| ^2 \right] }&= \int _{C_{S}} \! dz \frac{e^{i\alpha z}}{z} \exp { \left[ -\left| z + \epsilon \right| ^2 \right] } \nonumber \\&\quad +\,\int _{C_{L}} \! dz \frac{e^{i\alpha z}}{z} \exp { \left[ -\left| z\,+\,\epsilon \right| ^2 \right] } \nonumber \\&\quad +\,\int _{-\infty }^{+\infty } \!\! dx \frac{e^{i\alpha x}}{x} \exp { \left[ -\left( x + \epsilon \right) ^2 \right] }. \end{aligned}$$
(56)

The contour \(C\) consists of one large hemicircles \(C_{L}\) on the upper-half of the complex plane, one small hemicircle \(C_{S}\) on the lower-half complex plane around the pole \(z=0\), and a line on the real axis from the negative infinity to the positive infinity. The integral along \(C_{L}\) is zero according to Jordan’s lemma,

$$\begin{aligned} \int _{C_{L}} \! dz \frac{e^{i\alpha z}}{z} \exp { \left[ -\left| z + \epsilon \right| ^2 \right] } = 0. \end{aligned}$$
(57)

The integral along \(C_{S}\) can be evaluated as

$$\begin{aligned} \int _{C_{S}} \! dz \frac{e^{i\alpha z}}{z} \exp { \left[ -\left| z + \epsilon \right| ^2 \right] } = \pi i \left( e^{i\alpha z} \exp { \left[ -\left| z + \epsilon \right| ^2 \right] } \right) \Big |_{z=0} = \pi i e^{-\epsilon ^2}. \end{aligned}$$
(58)

The contour integral is evaluated accordingly to the Cauchy integral formula,

$$\begin{aligned} \oint _{C} \! dz \frac{e^{i\alpha z}}{z} \exp { \left[ -\left| z + \epsilon \right| ^2 \right] } = 2 \pi i \left( e^{i\alpha z} \exp { \left[ -\left| z + \epsilon \right| ^2 \right] } \right) \Big |_{z=0} = 2 \pi i e^{-\epsilon ^2}. \end{aligned}$$
(59)

Therefore, we obtain for values of \(\alpha \gg 1\) the value of the integral in Eq. (55),

$$\begin{aligned} \int _{-\infty }^{+\infty } \!\! dx \frac{\sin {(\alpha x)}}{x} \exp { \left[ -\left( x + \epsilon \right) ^2 \right] } = \mathrm Im \left( \pi i e^{-\epsilon ^2} \right) = \pi e^{-\epsilon ^2}. \end{aligned}$$
(60)

Combining Eqs. (52),  (53), (54), and (55), the integral \(D_{1}\cos {(\kappa )} - D_{2}\sin {(\kappa )}\) in Eq. (49) can be evaluated,

$$\begin{aligned} D_{1}\cos {(\kappa )} - D_{2}\sin {(\kappa )} = \frac{-\pi }{2} \exp { \left[ -\left( \frac{\omega _{p}-\omega _{0}}{2/\Delta t}\right) ^2 \right] }. \end{aligned}$$
(61)

Given Eqs. (49) and (61), the ensemble average of the energy change at drive frequency \(\omega _{p} \simeq 1\omega _{0}\) is

$$\begin{aligned} \langle W_{1\omega _{0}} \rangle = \frac{\pi }{8}\frac{(qE_{0}\Delta t)^2}{m}\exp {\left[ - 2\left( \frac{\omega _{p} - \omega _{0}}{2/\Delta t} \right) ^2 \right] }. \end{aligned}$$
(62)

Quantum Perturbation Analysis

In this appendix, we use the second-order quantum perturbation to obtain the energy expectation value for an harmonic oscillator that is excited by a propagating Gaussian pulse. The Gaussian pulse is calculated beyond dipole approximation. In other words, the pulse field has spatial dependence in addition to its temporal dependence. The agreement between quantum perturbation analysis and quantum simulation is about \(80~\%\). The result of the quantum analysis is similar to that of the classical analysis given in Eq. (37). A brief summary for the derivation of the quantum analysis is given in the following. The quantum state of an harmonic oscillator is

$$\begin{aligned} |\psi \rangle (t) = \sum _{n=0}^{\infty } c_{n}(t)e^{-i\omega _{n}t}|n\rangle , \end{aligned}$$
(63)

where \(|n\rangle \) is the unperturbed eigenstate, \(\omega _{n} = \omega _{0}(n+1/2)\) is the eigenfrequency, and \(c_{n}(t)\) is the probability amplitude. In the interaction picture, the Schrödinger equation with the unperturbed Hamiltonian \(\widehat{H}_{0}\) and the perturbative Hamiltonian \(\widehat{H}^{'}\) can be generally written as

$$\begin{aligned} i\hbar \frac{d}{dt}\left( \sum _{n=0}^{\infty } c_{n}(t)|n\rangle \right) = \widehat{\fancyscript{H}} \left( \sum _{n=0}^{\infty } c_{m}(t)|m\rangle \right) , \end{aligned}$$
(64)

where \(\displaystyle {\widehat{\fancyscript{H}} = e^{\frac{i}{\hbar } \widehat{H}_{0} t} \widehat{H}^{'} e^{-\frac{i}{\hbar } \widehat{H}_{0} t} }\). Using the second-order perturbation theory, the perturbative expansion of the probability amplitude \(c_{n}(t) \simeq c_{n}^{(0)}(t) + \lambda _{n} c_{n}^{(1)}(t) + \lambda ^{2}_{n} c_{n}^{(2)}(t)\) turns the Schrödinger equation into a system of equations,

$$\begin{aligned} \left\{ \begin{array}{ll} i\hbar \frac{d}{dt}\left( \sum _{n=0}^{\infty } c^{(0)}_{n}(t)|n\rangle \right) &{}= 0 \\ i\hbar \frac{d}{dt}\left( \sum _{n=0}^{\infty } \lambda _{n}c^{(1)}_{n}(t)|n\rangle \right) &{}= \widehat{H}^{'} \left( \sum _{n=0}^{\infty } c^{(0)}_{m}(t)|m\rangle \right) \\ i\hbar \frac{d}{dt}\left( \sum _{n=0}^{\infty } \lambda ^{2}_{n}c^{(2)}_{n}(t)|n\rangle \right) &{}= \widehat{H}^{'} \left( \sum _{n=0}^{\infty } \lambda _{n}c^{(1)}_{m}(t)|m\rangle \right) , \end{array} \right. \end{aligned}$$
(65)

where \(\lambda _{n}\) denotes the expansion factor for \(c_{n}(t) \). In our study, the perturbative Hamiltonian \(\widehat{H}^{'}\) is provided by the interaction between a charged quantum particle and a classical field,

$$\begin{aligned} \widehat{H}^{'} = -\frac{q}{2m}\left( 2\mathbf {A}_{p}\cdot \hat{\mathbf {p}}- q\mathbf {A}_{p}^2 \right) ,\end{aligned}$$
(66)

where \(q\) and \(m\) are the charge and the mass of the particle. The driving field \(\mathbf {A}_{p}\) is a propagating Gaussian pulse,

$$\begin{aligned} \mathbf {A}_{p}= A_{p} \cos {( \mathbf {k}_{p}\cdot \hat{\mathbf {x}}- \omega _{p} \tau )} \exp {\left[ - \left( \frac{\mathbf {k}_{p}\cdot \hat{\mathbf {x}}}{|\mathbf {k}_{p}|\Delta x} - \frac{\tau }{\Delta t} \right) ^2 \right] }\varvec{\varepsilon }_{p}, \end{aligned}$$
(67)

where \(\tau = t - t_{c}\), \(\mathbf {k}_{p} = \omega _{p}/c\left( \sin {\theta _{p}}, 0, \cos {\theta _{p}} \right) \) is the wave vector of the field and \(\varvec{\varepsilon }_{p}= \left( \cos {\theta _{p}}, 0, -\sin {\theta _{p}} \right) \) is the field polarization. Note that in order to take the calculation beyond the dipole approximation, we will keep the operator \(\hat{\mathbf {x}}\) in the function form of the driving field \(\mathbf {A}_{p}\). Using \(\mathbf {k}_{p}\cdot \hat{\mathbf {x}}\) as an expansion factor for \(\mathbf {A}_{p}\), the perturbative Hamiltonian can be expanded. In the interaction picture, the expanded perturbative Hamiltonian is

$$\begin{aligned} \widehat{\fancyscript{H}} \simeq \widehat{\fancyscript{H}}_{1\omega _{0}}(t) + \widehat{\fancyscript{H}}_{2\omega _{0}}(t) + \widehat{\fancyscript{H}}_{3\omega _{0}}(t), \end{aligned}$$
(68)

where

$$\begin{aligned} \widehat{\fancyscript{H}}_{1\omega _{0}}(t)&= f_{1}(t) \left( \hat{b}^{\dagger }e^{i\omega _{0}t} - \hat{b}e^{-i\omega _{0}t} \right) , \nonumber \\ \widehat{\fancyscript{H}}_{2\omega _{0}}(t)&= f_{2}(t) \left( \hat{b}^{\dagger 2}e^{i2\omega _{0}t} + \hat{1\!\!1} - \hat{b}^{2}e^{-i2\omega _{0}t} \right) , \nonumber \\ \widehat{\fancyscript{H}}_{3\omega _{0}}(t)&= f_{3}(t) \left[ \hat{b}^{\dagger 3}e^{i3\omega _{0}t} + \left( \hat{b}\hat{b}^{\dagger 2}e^{i\omega _{0}t} - \hat{b}^{\dagger }\hat{b}^{2}e^{-i\omega _{0}t} \right) \right. \nonumber \\&\left. + \left( \hat{b}^{\dagger }e^{i\omega _{0}t} + \hat{b}e^{-i\omega _{0}t} \right) - \hat{b}^{3}e^{-i3\omega _{0}t} \right] + g_{3}(t) \hat{1\!\!1}. \end{aligned}$$
(69)

The symbol \(\hat{b}^{\dagger }\) and \(\hat{b}\) are the raising and the lowing operators for the harmonic oscillator. The symbol \(\hat{1\!\!1}\) denotes the identity operator.

The time-dependent functions \(f_{1}(t)\), \(f_{2}(t)\), \(f_{3}(t)\), and \(g_{3}(t)\) are defined as

$$\begin{aligned} f_{1}(t)&= \left[ \left( \frac{-qA_{p}\varepsilon _{x}}{m} \right) i \sqrt{\frac{\hbar m \omega _{0}}{2}} \right] \cos {( \omega _{p} \tau )}\exp {(- \tau ^{2}/\Delta t^{2})}, \nonumber \\ f_{2}(t)&= \left[ \left( \frac{-qA_{p}\varepsilon _{x}}{m} \right) k_{x} \sqrt{\frac{\hbar }{2 m \omega _{0}}} i \sqrt{\frac{\hbar m \omega _{0}}{2}}\right] \sin {( \omega _{p} \tau )}\exp {(- \tau ^{2}/\Delta t^{2})} , \nonumber \\ f_{3}(t)&= \left[ \left( \frac{qA_{p}\varepsilon _{x}}{m} \right) \frac{k^{2}_{x}}{2} \left( \sqrt{\frac{\hbar }{2 m \omega _{0}}} \right) ^{2} i \sqrt{\frac{\hbar m \omega _{0}}{2}}\right] \cos {( \omega _{p} \tau )}\exp {(- \tau ^{2}/\Delta t^{2})}, \nonumber \\ g_{3}(t)&= \left( \frac{-q^{2}A^2_{p}}{m} \right) \cos ^{2}{( \omega _{p} \tau )}\exp {(- \tau ^{2}/\Delta t^{2})}. \end{aligned}$$
(70)

In the case of \(1\omega _{0}\)-excitation with \(\omega _{p} \simeq 1\omega _{0}\), only the \(\widehat{\fancyscript{H}}_{1\omega _{0}}(t)\) term is effective. In the cases of \(2\omega _{0}\)-excitation with \(\omega _{p} \simeq 2\omega _{0}\), only the \(\widehat{\fancyscript{H}}_{2\omega _{0}}(t)\) term is effective. In the case of \(3\omega _{0}\)-excitation with \(\omega _{p} \simeq 3\omega _{0}\), only the \(\widehat{\fancyscript{H}}_{3\omega _{0}}(t)\) term is effective. Assuming that the harmonic oscillator is initially at the ground state, the expectation value of the energy change can be calculated up to \(\lambda ^{2}_{n}\),

$$\begin{aligned} \langle \Delta E_{1\omega _{0}} \rangle&= \frac{\pi }{8}\left( \frac{q^{2}A^{2}_{p}\omega _{0}^{2}\Delta t^{2}}{m} \right) \exp {\left[ -2\left( \frac{\omega _{p} - \omega _{0}}{2/\Delta t} \right) ^{2} \right] } \cos ^2{(\theta _{p})} , \nonumber \\ \langle \Delta E_{2\omega _{0}} \rangle&= \frac{\pi }{16}\left( \frac{4q^{2}A^{2}_{p}\omega _{0}^{2}\Delta t^{2}}{m} \right) \left( \frac{\hbar \omega _{p}}{mc^{2}} \right) \left( \frac{\omega _{p}}{\omega _{0}}\right) \sin ^2{(\theta _{p})} \nonumber \\&\times \exp {\left[ -2\left( \frac{\omega _{p} - 2\omega _{0}}{2/\Delta t} \right) ^{2} \right] } \cos ^2{(\theta _{p})}, \nonumber \\ \langle \Delta E_{3\omega _{0}} \rangle&= \frac{3\pi }{16^2}\left( \frac{12q^{2}A^{2}_{p}\omega _{0}^{2}\Delta t^{2}}{m} \right) \left[ \left( \frac{\hbar \omega _{p}}{mc^{2}} \right) \left( \frac{\omega _{p}}{\omega _{0}}\right) \sin ^2{(\theta _{p})} \right] ^{2} \nonumber \\&\times \exp {\left[ -2\left( \frac{\omega _{p} - 3\omega _{0}}{2/\Delta t} \right) ^{2} \right] } \cos ^2{(\theta _{p})}. \end{aligned}$$
(71)

The result of quantum perturbation analysis, Eq. (71), is compared with quantum simulation in Fig. 3. The agreement is about \(80~\%\). In the same figure, the quantum analysis is also compared with the classical analysis, Eq. (37). The quantum analysis agrees well with the classical analysis, as evidenced by the similarity between Eqs. (71) and (37).

Fig. 3
figure 3

Comparison between quantum analysis, quantum simulation, and classical analysis. Up the overall agreement between quantum perturbation and quantum simulation is within \(80~\%\). Bottom the quantum analysis and the classical analysis shows identical resonance structures at \(\omega _{p} \simeq 1\omega _{0}\) and \(\omega _{p} \simeq 2\omega _{0}\), while the agreement at \(\omega _{p} \simeq 3\omega _{0}\) is about \(65~\%\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Huang, W.CW., Batelaan, H. Discrete Excitation Spectrum of a Classical Harmonic Oscillator in Zero-Point Radiation. Found Phys 45, 333–353 (2015). https://doi.org/10.1007/s10701-015-9866-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-015-9866-9

Keywords

Navigation