Skip to main content
Log in

Open quantum dynamics theory on the basis of periodical system-bath model for discrete Wigner function

  • Published:
Journal of Computational Electronics Aims and scope Submit manuscript

Abstract

Discretizing a distribution function in a phase space for an efficient quantum dynamics simulation is a non-trivial challenge, in particular for a case in which a system is further coupled to environmental degrees of freedom. Such open quantum dynamics is described by a reduced equation of motion (REOM), most notably by a quantum Fokker–Planck equation (QFPE) for a Wigner distribution function (WDF). To develop a discretization scheme that is stable for numerical simulations from the REOM approach, we employ a two-dimensional (2D) periodically invariant system-bath (PISB) model with two heat baths. This model is an ideal platform not only for a periodic system but also for a non-periodic system confined by a potential. We then derive the numerically “exact” hierarchical equations of motion (HEOM) for a discrete WDF in terms of periodically invariant operators in both coordinate and momentum spaces. The obtained equations can treat non-Markovian heat-bath in a non-perturbative manner at finite temperatures regardless of the mesh size. As demonstrations, we numerically integrate the discrete QFPE for a 2D free rotor and harmonic potential systems in a high-temperature Markovian case using a coarse mesh with initial conditions that involve singularity.

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
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon a reasonable request.

References

  1. Frensley, W.R.: Boundary conditions for open quantum systems driven far from equilibrium. Rev. Mod. Phys. 62, 745 (1990). https://doi.org/10.1103/RevModPhys.62.745

    Article  Google Scholar 

  2. Jacoboni, C., Bordone, P.: The Wigner-function approach non-equilibrium electron transport. Rep. Prog. Phys. 67, 1033 (2004). https://doi.org/10.1088/0034-4885/67/7/R01

    Article  Google Scholar 

  3. Grossmann, F., Koch, W.: A finite-difference implementation of the Caldeira-Leggett master equation. J. Chem. Phys. 130, 034105 (2009). https://doi.org/10.1063/1.3059006

    Article  Google Scholar 

  4. Kim, K.Y.: A discrete formulation of the Wigner transport equation. J. Appl. Phys. 102, 113705 (2007). https://doi.org/10.1063/1.2818363

    Article  Google Scholar 

  5. Weinbub, J., Ferry, D.K.: Recent advances in Wigner function approaches. Apply. Phys. Rev. 5, 041104 (2018). https://doi.org/10.1063/1.5046663

    Article  Google Scholar 

  6. Schwinger, J.: In: P84, Englert, B.-G. , (ed.) Quantum Mechanics: Symbolism of atomic measurements. Springer, Berlin (2001). (ISBN 13:9783662045893)

  7. Caldeira, A.O., Leggett, A.J.: Quantum tunneling in a dissipative system. Ann. Phys. 149, 374 (1983). https://doi.org/10.1016/0003-4916(83)90202-6

    Article  Google Scholar 

  8. Weiss, U.: Quantum Dissipative Systems, 4th edn. World Scientific, Singapore (2012)

    Book  Google Scholar 

  9. Breuer, H.P., Petruccione, F.: The Theory of Open Quantum Systems. Oxford University Press, New York (2002)

    MATH  Google Scholar 

  10. Caldeira, A.O., Leggett, A.J.: Path integral approach to quantum Brownian motion. Physica 121A, 587 (1983). https://doi.org/10.1016/0378-4371(83)90013

    Article  MathSciNet  MATH  Google Scholar 

  11. Waxman, D., Leggett, A.J.: Dissipative quantum tunneling at finite temperatures. Phys. Rev. B 32, 4450 (1985). https://doi.org/10.1103/PhysRevB.32.4450

    Article  Google Scholar 

  12. Tanimura, Y., Kubo, R.: Time evolution of a quantum system in contact with a nearly Gaussian-Markoffian noise bath. J. Phys. Soc. Jpn. 58, 101 (1989). https://doi.org/10.1143/JPSJ.58.101

    Article  MathSciNet  Google Scholar 

  13. Tanimura, Y., Wolynes, P.G.: Quantum and classical Fokker-Planck equations for a Gaussian-Markovian noise bath. Phys. Rev. A 43, 4131 (1991). https://doi.org/10.1103/PhysRevA.43.4131

    Article  Google Scholar 

  14. Tanimura, Y., Wolynes, P.G.: The interplay of tunneling, resonance, and dissipation in quantum barrier crossing: A numerical study. J. Chem. Phys. 96, 8485 (1992). https://doi.org/10.1063/1.462301

    Article  Google Scholar 

  15. Tanimura, Y.: Stochastic Liouville, Langevin, Fokker-Planck, and master equation approaches to quantum dissipative systems. J. Phys. Soc. Jpn. 75, 082001 (2006). https://doi.org/10.1143/JPSJ.75.082001

    Article  Google Scholar 

  16. Tanimura, Y.: Numerically "exact" approach to open quantum dynamics: the hierarchical equations of motion (HEOM). J. Chem. Phys. 153, 020901 (2020). https://doi.org/10.1063/5.0011599

    Article  Google Scholar 

  17. Tanimura, Y.: Reduced hierarchical equations of motion in real and imaginary time: Correlated initial states and thermodynamic quantities. J. Chem. Phys. 141, 044114 (2014). https://doi.org/10.1063/1.4890441

    Article  Google Scholar 

  18. Tanimura, Y.: Real-time and imaginary-time quantum hierarchical Fokker-Planck equations. J. Chem. Phys. 142, 144110 (2015). https://doi.org/10.1063/1.4916647

    Article  Google Scholar 

  19. Jensen, K.L., Buot, F.A.: Numerical simulation of intrinsic bistability and high-frequency current oscillations in resonant tunneling structures. Phys. Rev. Lett. 66, 1078 (1991)

    Article  Google Scholar 

  20. Jensen, K.L., Buot, F.A.: The methodology of simulating particle trajectories through tunneling structures using a Wigner distribution approach. IEEE Trans. Electron 38, 2337 (1991). https://doi.org/10.1109/16.88522

    Article  Google Scholar 

  21. Zhan, Z., Colomes, E., Oriols, X.: Unphysical features in the application of the Boltzmann collision operator in the time-dependent modeling of quantum transport. J. Comput. Electron. 15, 1206 (2016)

    Article  Google Scholar 

  22. Sakurai, A., Tanimura, Y.: An approach to quantum transport based on reduced hierarchy equations of motion: Application to a resonant tunneling diode. J. Phys. Soc. Jpn. 82, 033707 (2013). https://doi.org/10.7566/JPSJ.82.033707

    Article  Google Scholar 

  23. Sakurai, A., Tanimura, Y.: Self-excited current oscillations in a resonant tunneling diode described by a model based on the Caldeira-Leggett Hamiltonian. New J. Phys. 16, 015002 (2014). https://doi.org/10.1088/1367-2630/16/1/015002

    Article  Google Scholar 

  24. Grossmann, F., Sakurai, A., Tanimura, Y.: Electron pumping under non-Markovian dissipation: The role of the self-consistent field. J. Phys. Soc. Jpn. 85, 034803 (2016). https://doi.org/10.7566/JPSJ.85.034803

    Article  Google Scholar 

  25. Ringhofer, C., Ferry, D.K., Kluksdahl, N.: Absorbing boundary conditions for the simulation of quantum transport phenomena. Transp. Theory Stat. Phys. 18, 331 (1989)

    Article  MathSciNet  Google Scholar 

  26. Jiang, H., Lu, T., Cai, W.: A device adaptive inflow boundary condition for Wigner equations of quantum transport. J. Comput. Phys. 258, 773 (2014)

    Article  MathSciNet  Google Scholar 

  27. Schulz, D., Mahmood, A.: Approximation of a phase space operator for the numerical solution of the Wigner equation. IEEE J. Quant. Electron. 52, 1 (2016). https://doi.org/10.1109/JQE.2015.2504086

    Article  Google Scholar 

  28. Yamada, Y., Tsuchiya, H., Ogawa, M.: Quantum transport simulation of silicon-nanowire transistors based on direct solution approach of the Wigner transport equation. IEEE Trans. Electron Dev. 56, 1396 (2009). https://doi.org/10.1109/TED.2009.2021355

    Article  Google Scholar 

  29. Morandi, O., Schurrer, F.: Wigner model for quantum transport in graphene. J. Phys. A: Math. Theor. 44, 265301 (2011)

    Article  Google Scholar 

  30. Barraud, S.: Dissipative quantum transport in silicon nanowires based on Wigner transport equation. J. Appl. Phys. 110, 093710 (2011). https://doi.org/10.1063/1.3654143

    Article  Google Scholar 

  31. Jonasson, O., Knezevic, I.: Dissipative transport in superlattices within the Wigner function formalism. J. Comput. Electron. 14, 879 (2015)

    Article  Google Scholar 

  32. Tilma, T., Everitt, M.J., Samson, J.H., Munro, W.J., Nemoto, K.: Wigner functions for arbitrary quantum systems. Phys. Rev. Lett. 117, 180401 (2016). https://doi.org/10.1103/PhysRevLett.117.180401

    Article  Google Scholar 

  33. Ivanov, A., Breuer, H.P.: Quantum corrections of the truncated Wigner approximation applied to an exciton transport model. Phys. Rev. E 95, 042115 (2017). https://doi.org/10.1103/PhysRevE.95.042115

    Article  MathSciNet  Google Scholar 

  34. Kim, K.Y., Kim, J., Kim, S.: An efficient numerical scheme for the discrete Wigner transport equation via the momentum domain narrowing. AIP Adv. 6, 065314 (2016). https://doi.org/10.1063/1.4954237

    Article  Google Scholar 

  35. Kim, K.Y., Kim, S., Tang, T.: Accuracy balancing for the finite-difference-based solution of the discrete Wigner transport equation. J. Comput. Electron. 16, 148 (2017). https://doi.org/10.1007/s10825-016-0944-9

    Article  Google Scholar 

  36. Frensley, W.R.: Transient response of a tunneling device obtained from the Wigner function. Phys. Rev. Lett. 57, 2853 (1986). https://doi.org/10.1103/PhysRevLett.57.2853

    Article  Google Scholar 

  37. Frensley, W.R.: Wigner-function model of a resonant-tunneling semiconductor device. Phys. Rev. B 36, 1570 (1987). https://doi.org/10.1103/PhysRevB.36.1570

    Article  Google Scholar 

  38. Kluksdahl, N.C., Kriman, A.M., Ferry, D.K.: Self-consistent study of the resonant-tunneling diode. Phys. Rev. B 39, 7720 (1989)

    Article  Google Scholar 

  39. Shifren, L., Ringhofer, C., Ferry, D.K.: A Wigner function-based quantum ensemble Monte Carlo study of a resonant tunneling diode 50, 769 (2003). https://doi.org/10.1109/TED.2003.809434

    Article  Google Scholar 

  40. Jensen, K.L., Buot, F.A.: The methodology of simulating particle trajectories through tunneling structures using a Wigner distribution approach. IEEE Trans. Electron Dev. 38, 2337 (1991). https://doi.org/10.1109/16.88522

    Article  Google Scholar 

  41. Jensen, K.L., Buot, F.A.: Numerical simulation of intrinsic bistability and high-frequency current oscillations in resonant tunneling structures. Phys. Rev. Lett. 66, 1078 (1991). https://doi.org/10.1103/PhysRevLett.66.1078

    Article  Google Scholar 

  42. Zhao, P., Cui, H.L., Woolard, D., Jensen, K.L., Buot, F.A.: Simulation of resonant tunneling structures: origin of the I–V hysteresis and plateau-like structure. J. App. Phys. 87, 1337 (2010). https://doi.org/10.1063/1.372019

    Article  Google Scholar 

  43. Biegel, B.A., Plummer, J.D.: Comparison of self-consistency iteration options for the Wigner function method of quantum device simulation. Phys. Rev. B 54, 8070 (1996). https://doi.org/10.1103/PhysRevB.54.8070

    Article  Google Scholar 

  44. Yoder, P.D., Grupen, M., Smith, R.K.: Demonstration of intrinsic tristability in Double-Barrier resonant tunneling diodes with the Wigner transport equation. IEEE Trans. Electron Dev. 57, 3265 (2010). https://doi.org/10.1109/TED.2010.2081672

    Article  Google Scholar 

  45. Schulz, L., Schulz, D.: Application of a slowly varying envelope function onto the analysis of the Wigner transport equation. IEEE Trans. Nanotechnol. 15(5), 801–809 (2016). https://doi.org/10.1109/TNANO.2016.2581880

    Article  Google Scholar 

  46. Dorda, A., Schurrer, F.: A WENO-solver combined with adaptive momentum discretization for the Wigner transport equation and its application to resonant tunneling diodes. J. Comput. Phys. 284, 95 (2015)

    Article  MathSciNet  Google Scholar 

  47. Zueco, D., Garcıa-Palacios, J.L.: Quantum ratchets at high temperatures. Phys. E. 29, 435 (2005). https://doi.org/10.1016/j.physe.2005.05.043

    Article  Google Scholar 

  48. Cleary, L., Coffey, W.T., Kalmykov, Y.P., Titov, S.: Semiclassical treatment of a Brownian ratchet using the quantum Smoluchowski equation. Phys. Rev. E. 80, 051106 (2009). https://doi.org/10.1103/PhysRevE.80.051106

    Article  Google Scholar 

  49. Kato, A., Tanimura, Y.: Quantum suppression of ratchet rectification in a Brownian system driven by a biharmonic force. J. Phys. Chem. B 117, 13132 (2013). https://doi.org/10.1021/jp403056h

    Article  Google Scholar 

  50. Tanimura, Y., Mukamel, S.: Multistate quantum Fokker-Planck approach to nonadiabatic wave packet dynamics in pump-probe spectroscopy. J. Chem. Phys. 101, 3049 (1994). https://doi.org/10.1063/1.467618

    Article  Google Scholar 

  51. Chernyak, V., Mukamel, S.: Collective coordinates for nuclear spectral densities in energy transfer and femtosecond spectroscopy of molecular aggregates. J. Chem. Phys. 105, 4565 (1996). https://doi.org/10.1063/1.472302

    Article  Google Scholar 

  52. Tanimura, Y., Maruyama, Y.: Gaussian-Markovian quantum Fokker-Planck approach to nonlinear spectroscopy of a displaced Morse potentials system: Dissociation, predissociation, and optical Stark effects. J. Chem. Phys. 107, 1779 (1997). https://doi.org/10.1063/1.474531

    Article  Google Scholar 

  53. Maruyama, Y., Tanimura, Y.: Pump-probe spectra and nuclear dynamics for a dissipative molecular system in a strong laser field: predissociation dynamics. Chem. Phys. Lett. 292, 28 (1998). https://doi.org/10.1016/S0009-2614(98)00634-4

    Article  Google Scholar 

  54. Ikeda, T., Tanimura, Y.: Low-temperature quantum Fokker-Planck and Smoluchowski equations and their extension to multistate systems. J. Chem. Theory Comput. 15, 2517 (2019). https://doi.org/10.1021/acs.jctc.8b01195

    Article  Google Scholar 

  55. Ikeda, T., Tanimura, Y.: Probing photoisomerization processes by means of multi-dimensional electronic spectroscopy: The multi-state quantum hierarchical Fokker-Planck equation approach. J. Chem. Phys. 147, 014102 (2017). https://doi.org/10.1063/1.4989537

    Article  Google Scholar 

  56. Ikeda, T., Tanimura, Y.: Phase-space wavepacket dynamics of internal conversion via conical intersection: Multi-state quantum Fokker-Planck equation approach. Chem. Phys. 515, 203 (2018). https://doi.org/10.1016/j.chemphys.2018.07.013

    Article  Google Scholar 

  57. Ikeda, T., Tanimura, Y., Dijkstra, A.: Modeling and analyzing a photo-driven molecular motor system: Ratchet dynamics and non-linear optical spectra. J. Chem. Phys. 150, 114103 (2019). https://doi.org/10.1063/1.5086948

    Article  Google Scholar 

  58. Sakurai, A., Tanimura, Y.: Does \(\hbar\) play a role in multidimensional spectroscopy? Reduced hierarchy equations of motion approach to molecular vibrations. J. Phys. Chem. A 115, 4009 (2011). https://doi.org/10.1021/jp1095618

    Article  Google Scholar 

  59. Ikeda, T., Ito, H., Tanimura, Y.: Analysis of 2D THz-Raman spectroscopy using a non-Markovian Brownian oscillator model with nonlinear system-bath interactions. J. Chem. Phys. 142, 212421 (2015). https://doi.org/10.1063/1.4917033

    Article  Google Scholar 

  60. Ito, H., Tanimura, Y.: Simulating two-dimensional infrared-Raman and Raman spectroscopies for intermolecular and intramolecular modes of liquid water. J. Chem. Phys. 144, 074201 (2016). https://doi.org/10.1063/1.4941842

    Article  Google Scholar 

  61. Wootters, W.K.: A Wigner-function formulation of finite-state quantum mechanics. Ann. Phys. N.Y.) 176, 1 (1987)

    Article  MathSciNet  Google Scholar 

  62. Iwamoto, Y., Tanimura, Y.: Linear absorption spectrum of a quantum two-dimensional rotator calculated using a rotationally invariant system-bath Hamiltonian. J. Chem. Phys. 149, 084110 (2018). https://doi.org/10.1063/1.5044585

    Article  Google Scholar 

  63. Iwamoto, Y., Tanimura, Y.: Open quantum dynamics of a three-dimensional rotor calculated using a rotationally invariant system-bath Hamiltonian: Linear and two-dimensional rotational spectra. J. Chem. Phys. 151, 044105 (2019). https://doi.org/10.1063/1.5108609

    Article  Google Scholar 

  64. Suzuki, Y., Tanimura, Y.: Two-dimensional spectroscopy for a two-dimensional rotator coupled to a Gaussian-Markovian noise bath. J. Chem. Phys. 119, 1650 (2003). https://doi.org/10.1063/1.1578630

    Article  Google Scholar 

  65. Vourdas, A.: Quantum systems with finite Hilbert space. Rep. Prog. Phys. 67, 267 (2004). https://doi.org/10.1088/0034-4885/67/3/r03

    Article  MathSciNet  Google Scholar 

  66. Pegg, D.T., Barnett, S.M.: Phase properties of the quantized single-mode electromagnetic field. Phys. Rev. A 39, 1665 (1989). https://doi.org/10.1103/PhysRevA.39.1665

    Article  Google Scholar 

  67. Galetti, D., de Toledo Piza, A.: An extended Weyl-Wigner transformation for special finite spaces. Phys. A 149, 267 (1988). https://doi.org/10.1016/0378-4371(88)90219-1

    Article  MathSciNet  Google Scholar 

  68. Luis, A., Perina, J.: Discrete Wigner function for finite-dimensional systems. J. Phys. A: Math. Gen. 31, 1423 (1998)

    Article  MathSciNet  Google Scholar 

  69. Klimov, A.B., Munoz, C.: Discrete Wigner function dynamics. J. Opt. B. 7, S588 (2005)

    Article  MathSciNet  Google Scholar 

  70. Carruthers, P., Nieto, M.M.: Phase and angle variables in quantum mechanics. Rev. Mod. Phys. 40, 411 (1968)

    Article  Google Scholar 

Download references

Acknowledgements

Y. T. is supported by JSPS KAKENHI Grant Number B 21H01884.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yoshitaka Tanimura.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Y. T. is supported by JSPS KAKENHI Grant Number B 21H01884.

Appendices

A Canonical commutation relation in the large N limit

In this Appendix, we show that our coordinate and momentum operators satisfy the canonical commutation relation in the large N limit.

First we consider a non-periodic case, \(dx = x_0 \sqrt{2 \pi /(2N + 1)}\) and \(dp = p_0 \sqrt{2 \pi /(2N + 1)}\) with \(x_0 p_0 = \hbar\). We employ the relationship between the displaced operator, \({\hat{U}}_x {\hat{U}}_p - {\hat{U}}_p {\hat{U}}_x \omega ^{-1} = 0\). Assuming large N, we express \({\hat{U}}_x\) and \({\hat{U}}_p\) in Taylor expansion forms as

$$\begin{aligned}&\left[ 1 + \frac{i dp {\hat{x}}}{\hbar } + \frac{(i dp)^2}{2 \hbar ^2} {\hat{x}}^2\right] \left[ 1 + \frac{i {\hat{p}} dx}{\hbar } + \frac{(i dx)^2}{2 \hbar ^2} {\hat{p}}^2 \right] \nonumber \\&\quad - \left[ 1 + \frac{i {\hat{p}} dx}{\hbar } + \frac{(i dx)^2}{2 \hbar ^2} {\hat{p}}^2\right] \left[ 1 + \frac{i dp {\hat{x}}}{\hbar } + \frac{(i dp)^2}{2 \hbar ^2} {\hat{x}}^2\right] \left[ 1 - \frac{i dx dp}{\hbar }) + O( (N^{\frac{-3}{2}})\right] \nonumber \\&\quad = \frac{dx dp}{\hbar ^2}({\hat{x}} {\hat{p}} - {\hat{p}} {\hat{x}}) - \frac{i dx dp}{\hbar } + O( N^{\frac{-3}{2}} ). \end{aligned}$$
(35)

This indicates that the canonical commutation relation \([{\hat{x}} , {\hat{p}}] = i \hbar\) satisfies to an accuracy of \(O( N^{\frac{-3}{2}} )\).

In the \(2 \pi\)-periodic case, we set \(dx = {2 \pi }/{(2 N + 1)}\) and \(dp = \hbar\). Then, we obtain

$$\begin{aligned}&(\cos {\hat{x}} + i \sin {\hat{x}}) \left( 1 - \frac{i dx}{\hbar } {\hat{p}} \right) - \left( 1 - \frac{i dx}{\hbar } {\hat{p}}\right) (\cos {\hat{x}} + i \sin {\hat{x}}) (1 - i dx) + O(N^{-2}) \nonumber \\&\quad = \frac{dx}{\hbar }(\sin {\hat{x}} {\hat{p}} - {\hat{p}} \sin {\hat{x}} - i \hbar \cos {\hat{x}}) + \frac{i dx}{\hbar }(\cos {\hat{x}} {\hat{p}} - {\hat{p}} \cos {\hat{x}} + i \hbar \sin {\hat{x}}) + O(N^{-2}). \end{aligned}$$
(36)

The first and second terms of the RHS in Eq. (36) are the anti-Hermite and Hermite operators. Therefore, the contributions from these terms are zero. Thus, for large N, we obtain the canonical commutation relations for a periodic case as [70]

$$\begin{aligned} {[}\sin {\hat{x}} , {\hat{p}}]&= i \hbar \cos {\hat{x}}, \end{aligned}$$
(37)

and

$$\begin{aligned} {[}\cos {\hat{x}} , {\hat{p}}]&= - i \hbar \sin {\hat{x}}, \end{aligned}$$
(38)

to an accuracy of \(O(N^{-2})\).

B QME for 2D PISB model and counter term

To demonstrate a role of the counter term, here we employ the QME for the 2D PISB model. As shown in [62], the QME for the reduced density matrix of the system, \({\hat{\rho }}(t)\), is derived from the second-order perturbation approach as

(39)

where

$$\begin{aligned} {{\hat{\varGamma }}}_\alpha (\tau ) {\hat{\rho }}(t - \tau )&\equiv C(\tau ) [{\hat{V}}_{\alpha } , {\hat{G}}_S(\tau ) {\hat{V}}_{\alpha } {\hat{\rho }}(t - \tau ) {\hat{G}}_S^{\dagger }(\tau )] \nonumber \\&- C(-\tau )[ {\hat{V}}_{\alpha } , {\hat{G}}_S(\tau ) {\hat{\rho }}(t - \tau ) {\hat{V}}_{\alpha } {\hat{G}}_S^{\dagger }(\tau )] \end{aligned}$$
(40)

is the damping operator for \(\alpha = x\) or y, in which

$$\begin{aligned} C(\tau )&= \hbar \int ^{\infty }_0 \frac{d \omega }{\pi } J (\omega ) \left[ \coth \left( \frac{\beta \hbar \omega }{2}\right) \cos (\omega \tau ) - i \sin (\omega \tau )\right] \end{aligned}$$
(41)

is the bath correlation function and \({\hat{G}}_S(\tau )\) is the time-evolution operator of the system. For the Ohmic SDF \(J(\omega )=\eta \omega\), \(C(\tau )\) reduces to the Markovian form as

$$\begin{aligned} C(\tau ) = \eta \left( \frac{2 }{\beta } + i \hbar \frac{d}{d\tau } \right) \delta (\tau ). \end{aligned}$$
(42)

Using the relation \(\int ^{t}_{0} d \tau {{\hat{\varGamma }}}_{\alpha }(\tau ) {\hat{\rho }}(t - \tau ) = \hat{{{\bar{\varGamma }}}}_{\alpha } {\hat{\rho }}(t)+{i \hbar \eta }\delta (0) [({\hat{V}}_{\alpha })^2, {\hat{\rho }}(t)]\), we can rewrite the damping operator, Eq. (40), as

$$\begin{aligned} \hat{{{\bar{\varGamma }}}}_{\alpha } {\hat{\rho }}(t)&= \frac{\eta }{\beta } \left( [{\hat{V}}_{\alpha }, {\hat{V}}_{\alpha } {\hat{\rho }}(t) ] - [ {\hat{V}}_{\alpha } , {\hat{\rho }}(t ) {\hat{V}}_{\alpha }] \right) +\frac{i\hbar \eta }{2 } \left[ ({\hat{V}}_{\alpha })^2 , \frac{d {\hat{\rho }}(t - \tau )}{d \tau }|_{\tau = 0} \right] \nonumber \\&- \frac{\eta }{2 } \left( [{\hat{V}}_{\alpha } , {\hat{H}}_S {\hat{V}}_{\alpha } {\hat{\rho }}(t ) ] + [ {\hat{V}}_{\alpha } , {\hat{H}}_S {\hat{\rho }}(t ) {\hat{V}}_{\alpha } ] - [{\hat{V}}_{\alpha } , {\hat{V}}_{\alpha } {\hat{\rho }}(t ) {\hat{H}}_S] - [ {\hat{V}}_{\alpha } , {\hat{\rho }}(t ) {\hat{V}}_{\alpha } {\hat{H}}_S] \right) . \end{aligned}$$
(43)

In the case if there is only \({{\hat{V}}}_{y} =\hbar \sin ({{\hat{x}} dp}/{\hbar } )/dp\) interaction in the PISB model, (i.e., \({\hat{V}}_{x}=0\)), we encounter the divergent term \({i \hbar \eta }\delta (0) [({\hat{V}}_{y})^2, {\hat{\rho }}(t)]\) that arises from the second term in the RHS of Eq. (43). Because \({\hat{V}}_{y}\) reduces to the linear operator of the coordinate \({\hat{V}}_{y} \approx {{\hat{x}}}\) in the large N limit, the PISB model under this condition corresponds to the Caldeira–Leggett model without the counter term: Divergent term arises because we exclude the counter term in the bath Hamiltonian, Eq. (3). (See also [64].) If we include \({{\hat{V}}}_{x} =\hbar \cos ({{\hat{x}} dp}/{\hbar } )/dp\), this divergent term vanishes, because, by using the relation \(\sin ^2({{\hat{x}} dp}/{\hbar } ) + \cos ^2({{\hat{x}} dp}/{\hbar } )=1\), we have

$$\begin{aligned} {i \hbar \eta }\delta (0) [({\hat{V}}_{x})^2, {\hat{\rho }}(t)] + {i \hbar \eta }\delta (0) [({\hat{V}}_{y})^2, {\hat{\rho }}(t)]&= {i \hbar \eta }\delta (0)[{\hat{I}}, {\hat{\rho }}(t)] \nonumber \\&= 0. \end{aligned}$$
(44)

This implies that the interaction \({\hat{V}}_{y}\) plays the same role as the counter term. This fact indicates the significance of constructing a system-bath model with keeping the same symmetry as the system itself. If we ignore this point, the system dynamics are seriously altered by the bath even if the system-bath interaction is feeble [64].

C Discrete Moyal bracket

Using the kinetic term (the first term in the RHS of Eq. (32)) as an example, here we demonstrate the evaluation of the discrete Moyal bracket defined as Eq. (30). The kinetic energy in a finite Hilbert space representation is expressed as

$$\begin{aligned} \varvec{T}(p_j, q_k)&= \frac{\hbar ^2}{dx^2} \left( 1 - \sum _{l = -N}^{N} \exp \left( i \frac{-2 q_k (p_j - p_l)}{\hbar } \right) \langle P, l | \cos \left( \frac{{\hat{p}} dx}{\hbar } \right) | P, 2 j - l \rangle \right) \nonumber \\&= \frac{\hbar ^2}{dx^2} \left( 1 - \sum _{l = -N}^{N} \exp \left( i \frac{-2 q_k (p_j - p_l)}{\hbar } \right) \cos \left( \frac{p_l dx}{\hbar } \right) \delta _{l, 2j - l } \right) \nonumber \\&= \frac{\hbar ^2}{dx^2} \left( 1 - \cos \left( \frac{p_j dx}{\hbar } \right) \right) . \end{aligned}$$
(45)

Because the Moyal bracket with \(\varvec{A_1} = \hbar ^2/dx^2\) and \(\varvec{A_2} = \varvec{W}\) is zero, we focus on the \(\cos \left( {p_j dx}/{\hbar } \right)\) term. Let \(\varvec{A_1} = \exp \left( \pm i{p_j dx}/{\hbar }\right)\) and \(\varvec{A_2} = \varvec{W}\) in Eq. (30). Then we have

$$\begin{aligned}&\left[ \exp \left( \pm i\frac{p_j dx}{\hbar }\right) \star \varvec{W} \right] (p_j, q_k) = \frac{1}{(2 N + 1)^2} \sum _{j_1, j_2, k_1, k_2 = -N}^N \exp \left( i \frac{2 p_{j_2} q_{k_1} - 2 p_{j_1} q_{k_2}}{\hbar } \right) \nonumber \\&\quad \times \exp \left( \pm i\frac{(p_j + p_{j_1}) dx}{\hbar }\right) \varvec{W}( p_j + p_{j_2}, q_k + q_{k_2})\nonumber \\&\quad = \frac{1}{(2 N + 1)} \sum _{j_1, k_2 = -N}^N \exp \left( i \frac{(\pm 1 - 2 k_2) p_{j_1} dx}{\hbar } \right) \exp \left( \pm i\frac{p_j dx}{\hbar }\right) \varvec{W}( p_j , q_k + q_{k_2}) \nonumber \\&\quad = \sum _{k_2 = -N}^N \delta '_{\pm 1 - 2 k_2, 0} \exp \left( \pm i\frac{p_j dx}{\hbar }\right) \varvec{W}( p_j , q_k + q_{k_2}) \nonumber \\&\quad = \exp \left( \pm i\frac{p_j dx}{\hbar } \right) \varvec{W}( p_j , q_{k \pm (N + 1)}). \end{aligned}$$
(46)

Similarly, for \(\varvec{A_1} = \varvec{W}\) and \(\varvec{A_2} = \exp \left( \pm i{p_j dx}/{\hbar }\right)\), we have

$$\begin{aligned}&\left[ \varvec{W} \star \exp \left( \pm i\frac{p_j dx}{\hbar }\right) \right] (p_j, q_k) = \exp \left( \mp i\frac{p_j dx}{\hbar }\right) \varvec{W}( p_j , q_{k \pm (N + 1)}). \end{aligned}$$
(47)

Thus, the discrete Moyal product of the kinetic energy is expressed as

$$\begin{aligned}&-\frac{i}{\hbar }[ \varvec{T} \star \varvec{W}](p_j, q_k) = - \hbar \sin \left( \frac{p_j dx}{\hbar } \right) \frac{\varvec{W}( p_j , q_{k + N + 1}) - \varvec{W}( p_j , q_{k - N - 1})}{dx^2}. \end{aligned}$$
(48)

For example, for \(q_0\), the above expression involves the contributions from \(q_{N + 1 \equiv -N (mod \ 2N + 1)}\) and \(q_{-N - 1 \equiv N (mod \ 2N + 1)}\), which are the elements near the boundary of the periodic state. Note that \(N + 1\) arises from \(\delta '_{1 - 2 k_2, 0}\) that is the inverse element of 2 modulo \(2N + 1\). For large N, the above expression reduces to the kinetic term of the conventional QFPE by regarding the finite difference near the boundary as the derivative of the coordinate.

D Discrete quantum Fokker–Planck equation for large N

For a large N, Eq. (33) reduces to

$$\begin{aligned} \frac{\partial }{\partial t} W(p, q)&= -p \frac{\partial }{\partial q} W(p, q) -\frac{i}{\hbar } \{ \varvec{U} , \varvec{W} \}_M + \frac{\eta }{ \beta } \frac{\partial ^2}{\partial p^2}W(p, q) \nonumber \\&+\frac{\eta }{2} \left( {\hat{M}}^2_p {\hat{M}}_x W(p, q) + p {\hat{M}}_x \frac{\partial }{\partial p} W(p, q) \right) , \end{aligned}$$
(49)

where

$$\begin{aligned}&\frac{\partial W(p, q)}{\partial q} \equiv \frac{W(p_j, q_{k + N + 1)} - W(p_j, q_{k - N - 1})}{dx}, \end{aligned}$$
(50)
$$\begin{aligned}&\frac{\partial W(p, q)}{\partial p} \equiv \frac{W(p_{j + N + 1}, q_k) - W(p_{j - N - 1}, q_k)}{dp}, \end{aligned}$$
(51)
$$\begin{aligned}&{\hat{M}}_x W(p, q) \equiv \frac{W(p_j, q_{k + N + 1)} + W(p_j, q_{k - N - 1})}{2} , \end{aligned}$$
(52)

and

$$\begin{aligned} {\hat{M}}_p W(p, q) \equiv \frac{W(p_{j + N + 1}, q_k) + W(p_{j - N - 1}, q_k)}{2}. \end{aligned}$$
(53)

Although the above expression has a similar form to the QFPE, the finite difference operators for the discrete WDF are defined by the elements near the periodic boundary, i.e., for \(W(p_0, q_0)\), \(\partial /\partial q\) is evaluated from \(W(p_0, q_{-(N+1)} )\), and \(W(p_0, q_{N+1})\). Thus, the appearance of the discrete WDF can be different from the regular WDF, as depicted in Fig. 5 even for large N.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Iwamoto, Y., Tanimura, Y. Open quantum dynamics theory on the basis of periodical system-bath model for discrete Wigner function. J Comput Electron 20, 2091–2103 (2021). https://doi.org/10.1007/s10825-021-01754-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10825-021-01754-z

Keywords

Navigation