Skip to main content
Log in

Coherent dynamics of a qubit–oscillator system in a noisy environment

  • Published:
Quantum Information Processing Aims and scope Submit manuscript

Abstract

We investigate the non-Markovian dynamics of a qubit–oscillator system embedded in a noisy environment by employing the hierarchical equations of motion approach. It is shown that the decoherence rate of the whole qubit–oscillator–bath system can be significantly suppressed by enhancing the coupling strength between the qubit and the harmonic oscillator. Moreover, we find that the non-Markovian memory character of the bath is able to facilitate a robust quantum coherent dynamics in this qubit–oscillator–bath system. Our findings may be used to engineer some tunable coherent manipulations in mesoscopic quantum circuits.

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

Similar content being viewed by others

References

  1. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information. Cambridge University Press, Cambridge (2000)

    MATH  Google Scholar 

  2. Georgescu, I.M., Ashhab, S., Nori, F.: Quantum simulation. Rev. Mod. Phys. 86, 153 (2014)

    Article  ADS  Google Scholar 

  3. Horodecki, R., Horodecki, P., Horodecki, M., Horodecki, K.: Quantum entanglement. Rev. Mod. Phys. 81, 865 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  4. Modi, K., Brodutch, A., Cable, H., Paterek, T., Vedral, V.: The classical-quantum boundary for correlations: discord and related measures. Rev. Mod. Phys. 84, 1655 (2012)

    Article  ADS  Google Scholar 

  5. Baumgratz, T., Cramer, M., Plenio, M.B.: Quantifying coherence. Phys. Rev. Lett. 113, 140401 (2014)

    Article  ADS  Google Scholar 

  6. Streltsov, A., Rana, S., Bera, M.N., Lewenstein, M.: Towards resource theory of coherence in distributed scenarios. Phys. Rev. X 7, 011024 (2017)

    Google Scholar 

  7. Weiss, U.: Quantum Dissipative Systems. World Scientific, Singapore (1999)

    Book  Google Scholar 

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

    MATH  Google Scholar 

  9. Leggett, A.J., Chakravarty, S., Dorsey, A.T., Fisher, M.P.A., Garg, A., Zwerger, W.: Dynamics of the dissipative two-state system. Rev. Mod. Phys. 59, 1 (1987)

    Article  ADS  Google Scholar 

  10. Breuer, H.-P., Laine, E.-M., Piilo, J., Vacchini, B.: Non-Markovian dynamics in open quantum systems. Rev. Mod. Phys. 88, 021002 (2016)

    Article  ADS  Google Scholar 

  11. de Vega, I., Alonso, D.: Dynamics of non-Markovian open quantum systems. Rev. Mod. Phys. 89, 015001 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  12. Viola, L., Lloyd, S.: Dynamical suppression of decoherence in two-state quantum systems. Phys. Rev. A 58, 2733 (1998)

    Article  ADS  MathSciNet  Google Scholar 

  13. Shiokawa, K., Lidar, D.A.: Dynamical decoupling using slow pulses: efficient suppression of \(1/f\) noise. Phys. Rev. A 69, 030302 (2004)

    Article  ADS  Google Scholar 

  14. Xu, H.-S., Xu, J.-B.: Protecting quantum correlations of two qubits in independent non-Markovian environments by bang-bang pulses. J. Opt. Soc. Am. B 29, 2074 (2012)

    Article  ADS  Google Scholar 

  15. Wu, W., Luo, D.-W., Xu, J.-B.: Double sudden transitions of geometric discord at finite-temperature in the framework of stochastic description. J. Appl. Phys. 115, 244906 (2014)

    Article  ADS  Google Scholar 

  16. Huang, P., Zheng, H.: Effect of bath temperature on the quantum decoherence. Chem. Phys. Lett. 500, 256 (2010)

    Article  ADS  Google Scholar 

  17. Wu, W., Lin, H.-Q.: Effect of bath temperature on the decoherence of quantum dissipative systems. Phys. Rev. A 94, 062116 (2016)

    Article  ADS  Google Scholar 

  18. Vierheilig, C., Bercioux, D., Grifoni, M.: Dynamics of a qubit coupled to a dissipative nonlinear quantum oscillator: an effective-bath approach. Phys. Rev. A 83, 012106 (2011)

    Article  ADS  Google Scholar 

  19. Yao, Y.: Coherent dynamics of singlet fission controlled by nonlocal electron–phonon coupling. Phys. Rev. B 93, 115426 (2016)

    Article  ADS  Google Scholar 

  20. Ma, T., Chen, Y., Chen, T., Hedemann, S.R., Yu, T.: Crossover between non-Markovian and Markovian dynamics induced by a hierarchical environment. Phys. Rev. A 90, 042108 (2014)

    Article  ADS  Google Scholar 

  21. Man, Z.-X., Xia, Y.-J., Franco, R.L.: Cavity-based architecture to preserve quantum coherence and entanglement. Sci. Rep. 5, 13843 (2015)

    Article  ADS  Google Scholar 

  22. Levi, E.K., Irish, E.K., Lovett, B.W.: Coherent exciton dynamics in a dissipative environment maintained by an off-resonant vibrational mode. Phys. Rev. A 93, 042109 (2016)

    Article  ADS  Google Scholar 

  23. Lv, Z., Zheng, H.: Engineered tunable decay rate and controllable dissipative dynamics. J. Chem. Phys. 136, 121103 (2012)

    Article  ADS  Google Scholar 

  24. Chen, H.-B., Lien, J.-Y., Hwang, C.-C., Chen, Y.-N.: Long-lived quantum coherence and non-Markovianity of photosynthetic complexes. Phys. Rev. E 89, 042147 (2014)

    Article  ADS  Google Scholar 

  25. Man, Z.-X., An, N.B., Xia, Y.-J.: Non-Markovian dynamics of a two-level system in the presence of hierarchical environments. Opt. Express 23, 5763 (2015)

    Article  ADS  Google Scholar 

  26. Tanimura, Y., Kubo, R.K.: Time evolution of a quantum system in contact with a nearly Gaussian–Markoffian noise bath. J. Phys. Soc. Jpn. 58, 101 (1989)

    Article  ADS  MathSciNet  Google Scholar 

  27. Xu, R.X., Yan, Y.J.: Dynamics of quantum dissipation systems interacting with bosonic canonical bath: hierarchical equations of motion approach. Phys. Rev. E 75, 031107 (2007)

    Article  ADS  Google Scholar 

  28. Makarov, D.E., Makri, N.: Path integrals for dissipative systems by tensor multiplication. Condensed phase quantum dynamics for arbitrarily long time. Chem. Phys. Lett. 221, 482 (1994)

    Article  ADS  Google Scholar 

  29. Kast, D., Ankerhold, J.: Persistence of coherent quantum dynamics at strong dissipation. Phys. Rev. Lett. 110, 010402 (2013)

    Article  ADS  Google Scholar 

  30. Wang, H., Thoss, M.: Multilayer formulation of the multiconfiguration time-dependent Hartree theory. J. Chem. Phys. 119, 1289 (2003)

    Article  ADS  Google Scholar 

  31. White, S.R., Feiguin, A.E.: Real-time evolution using the density matrix renormalization group. Phys. Rev. Lett. 93, 076401 (2004)

    Article  ADS  Google Scholar 

  32. Wu, N., Duan, L., Li, X., Zhao, Y.: Dynamics of the sub-Ohmic spin-boson model: a time-dependent variational study. J. Chem. Phys. 138, 084111 (2013)

    Article  ADS  Google Scholar 

  33. Breuer, H.-P., Laine, E.-M., Piilo, J.: Measure for the degree of non-Markovian behavior of quantum processes in open systems. Phys. Rev. Lett. 103, 210401 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  34. Lu, X.-M., Wang, X., Sun, C.P.: Quantum Fisher information flow and non-Markovian processes of open systems. Phys. Rev. A 82, 042103 (2010)

    Article  ADS  Google Scholar 

  35. Luo, S., Fu, S., Song, H.: Quantifying non-Markovianity via correlations. Phys. Rev. A 86, 044101 (2012)

    Article  ADS  Google Scholar 

  36. Rajagopal, A.K., Devi, A.R.U., Rendell, R.W.: Kraus representation of quantum evolution and fidelity as manifestations of Markovian and non-Markovian forms. Phys. Rev. A 82, 042107 (2010)

    Article  ADS  Google Scholar 

  37. Bylicka, B., Chruscinski, D., Maniscalco, S.: Non-Markovianity and reservoir memory of quantum channels: a quantum information theory perspective. Sci. Rep. 4, 5720 (2014)

    Article  ADS  Google Scholar 

  38. Chruscinski, D., Maniscalco, S.: Degree of non-Markovianity of quantum evolution. Phys. Rev. Lett. 112, 120404 (2014)

    Article  ADS  Google Scholar 

  39. Liu, B.-H., Li, L., Huang, Y.-F., Li, C.-F., Guo, G.-C., Laine, E.-M., Breuer, H.-P., Piilo, J.: Experimental control of the transition from Markovian to non-Markovian dynamics of open quantum systems. Nat. Phys. 7, 931 (2011)

    Article  Google Scholar 

  40. Huelga, S.F., Rivas, A., Plenio, M.B.: Non-Markovianity-assisted steady state entanglement. Phys. Rev. Lett. 108, 160402 (2012)

    Article  ADS  Google Scholar 

  41. Ma, J., Yadin, B., Girolami, D., Vedral, V., Gu, M.: Converting coherence to quantum correlations. Phys. Rev. Lett. 116, 160407 (2016)

    Article  ADS  Google Scholar 

  42. Rabi, I.I.: Space quantization in a gyrating magnetic field. Phys. Rev. 51, 652 (1937)

    Article  ADS  Google Scholar 

  43. Scully, M.O., Zubairy, M.S.: Quantum Optics. Cambridge University Press, Cambridge (1997)

    Book  Google Scholar 

  44. Xiang, Z.-L., Ashhab, S., You, J.Q., Nori, F.: Hybrid quantum circuits: superconducting circuits interacting with other quantum systems. Rev. Mod. Phys. 85, 623 (2013)

    Article  ADS  Google Scholar 

  45. Koppens, F.H.L., Buizert, C., Tielrooij, K.J., Vink, I.T., Nowack, K.C., Meunier, T., Kouwenhoven, L.P., Vandersypen, L.M.K.: Driven coherent oscillations of a single electron spin in a quantum dot. Nature (London) 442, 766 (2006)

    Article  ADS  Google Scholar 

  46. Nowack, K.C., Koppens, F.H.L., Nazarov, Y.V., Vandersypen, L.M.K.: Coherent control of a single electron spin with electric fields. Science 318, 1430 (2007)

    Article  ADS  Google Scholar 

  47. Gan, C., Zheng, H.: Non-Markovian dynamics of a dissipative two-level system: nonzero bias and sub-Ohmic bath. Phys. Rev. E 80, 041106 (2009)

    Article  ADS  Google Scholar 

  48. Zhao, C., Lv, Z., Zheng, H.: Entanglement evolution and quantum phase transition of biased \(s=1/2\) spin-boson model. Phys. Rev. E 84, 011114 (2011)

    Article  ADS  Google Scholar 

  49. Nazir, A., McCutcheon, D.P.S., Chin, A.W.: Ground state and dynamics of the biased dissipative two-state system: beyond variational polaron theory. Phys. Rev. B 85, 224301 (2012)

    Article  ADS  Google Scholar 

  50. Wu, W.: Realization of hierarchical equations of motion from stochastic perspectives. Phys. Rev. A 98, 012110 (2018)

    Article  ADS  Google Scholar 

  51. Wu, M.-H., Lei, C.U., Zhang, W.-M., Xiong, H.-N.: Non-Markovian dynamics of a microcavity coupled to a waveguide in photonic crystals. Opt. Express 18, 18407 (2010)

    Article  ADS  Google Scholar 

  52. Zhu, L., Liu, H., Shi, Q.: A new method to account for the difference between classical and quantum baths in quantum dissipative dynamics. New J. Phys. 15, 095020 (2013)

    Article  ADS  Google Scholar 

  53. Thoss, M., Wang, H., Miller, W.H.: Self-consistent hybrid approach for complex systems: application to the spin-boson model with Debye spectral density. J. Chem. Phys. 115, 2991 (2001)

    Article  ADS  Google Scholar 

  54. Jing, J., Yu, T.: Non-Markovian relaxation of a three-level system: quantum trajectory approach. Phys. Rev. Lett. 105, 240403 (2010)

    Article  ADS  Google Scholar 

  55. Yoshihara, F., Fuse, T., Ashhab, S., Kakuyanagi, K., Saito, S., Semba, K.: Characteristic spectra of circuit quantum electrodynamics systems from the ultrastrong-to the deep-strong-coupling regime. Phys. Rev. A 95, 053824 (2017)

    Article  ADS  Google Scholar 

  56. Yoshihara, F., Fuse, T., Ashhab, S., Kakuyanagi, K., Saito, S., Semba, K.: Superconducting qubit–oscillator circuit beyond the ultrastrong-coupling regime. Nat. Phys. 13, 44 (2017)

    Article  Google Scholar 

  57. Pineda, C., Gorin, T., Davalos, D., Wisniacki, D.A., Garcia-Mata, I.: Measuring and using non-Markovianity. Phys. Rev. A 93, 022117 (2016)

    Article  ADS  Google Scholar 

  58. Irish, E.K., Gea-Banacloche, J., Martin, I., Schwab, K.C.: Dynamics of a two-level system strongly coupled to a high-frequency quantum oscillator. Phys. Rev. B 72, 195410 (2005)

    Article  ADS  Google Scholar 

  59. Lee, C.K., Moix, J., Cao, J.: Accuracy of second order perturbation theory in the polaron and variational polaron frames. J. Chem. Phys. 136, 204120 (2012)

    Article  ADS  Google Scholar 

  60. Wu, W., Liu, M.: Effects of counter-rotating-wave terms on the non-Markovianity in quantum open systems. Phys. Rev. A 96, 032125 (2017)

    Article  ADS  Google Scholar 

  61. Makela, H., Mottonen, M.: Effects of the rotating-wave and secular approximations on non-Markovianity. Phys. Rev. A 88, 052111 (2013)

    Article  ADS  Google Scholar 

  62. Wibmann, S., Karlsson, A., Laine, E.M., Piilo, J., Breuer, H.P.: Optimal state pairs for non-Markovian quantum dynamics. Phys. Rev. A 86, 062108 (2012)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

W. Wu wishes to thank Dr. Yi-Nan Fang, Dr. Li-Jing Jin, Dr. Maoxin Liu, Professor Zhiguo Lv and Professor Hai-Qing Lin for many useful discussions. This project is supported by the NSFC (Grant No. 11704025), the NSAF (Grant No. U1530401) and the China Postdoctoral Science Foundation (Grant Nos. 2017M610753 and 2018T110038).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wei Wu.

Appendices

Appendix A: The HEOM approach

In this paper, we adopt the HEOM method to investigate the non-Markovian coherent dynamics. For the Ornstein–Uhlenbeck-type bath correlation function given by Eq. 3, the hierarchical equations can be derived from the ordinary Schr\(\ddot{\mathrm {o}}\)dinger equation \(\partial _{t}|\psi (t)\rangle =-i{\hat{H}}|\psi (t)\rangle \) or the quantum Liouville equation \(\partial _{t}{\hat{\rho }}(t)=-i[{\hat{H}},{\hat{\rho }}(t)]\) by making use of the Feynman–Vernon influence functional approach [26, 27] or stochastic perspectives [17, 50]. Following procedures shown in Refs. [17, 50], one can obtain the hierarchy equations of the reduced qubit–oscillator system (the degrees of freedom of the multi-mode bosonic bath has been partially traced out) as follows:

$$\begin{aligned} \begin{aligned} \frac{\mathrm{d}}{\mathrm{d}t}{\hat{\rho }}_{{\varvec{\ell }}}(t)&=\left( -i{\hat{H}}_{\mathrm {qo}}^{\times }-{\varvec{\ell }}\cdot {\varvec{\mu }}\right) {\hat{\rho }}_{{\varvec{\ell }}}(t)+{\hat{\Phi }}\sum _{p=1}^{2}{\hat{\rho }}_{{\varvec{\ell }}+\mathbf {e}_{p}}(t)+\sum _{p=1}^{2}\ell _{p}{\hat{\Psi }}_{p}{\hat{\rho }}_{{\varvec{\ell }}-\mathbf {e}_{p}}(t), \end{aligned} \end{aligned}$$
(5)

where vector \({\varvec{\ell }}=(\ell _{1},\ell _{2})\) is a two-dimensional index, \(\mathbf {e}_{1}=(1,0)\), \(\mathbf {e}_{2}=(0,1)\) and \({\varvec{\mu }}=(\lambda +i\Delta ,\lambda -i\Delta )\) are two-dimensional vectors, and the superoperators \({\hat{\Phi }}\) and \({\hat{\Psi }}_{p}\) are defined as follows:

$$\begin{aligned} {\hat{\Phi }}=-\,i{\hat{\sigma }}_{z}^{\times },\quad {\hat{\Psi }}_{p}=\frac{i}{4}\gamma \lambda \left[ (-1)^{p}{\hat{\sigma }}_{z}^{\circ }-{\hat{\sigma }}_{z}^{\times }\right] , \end{aligned}$$

with \({\hat{X}}^{\times }{\hat{Y}}\equiv [{\hat{X}},{\hat{Y}}]={\hat{X}}{\hat{Y}}-{\hat{Y}}{\hat{X}}\) and \({\hat{X}}^{\circ }{\hat{Y}}\equiv \{{\hat{X}},{\hat{Y}}\}= {\hat{X}}{\hat{Y}}+{\hat{Y}}{\hat{X}}\).

The initial-state conditions of the auxiliary matrices are \({\hat{\rho }}_{{\varvec{\ell }}=\mathbf {0}}(0)={\hat{\rho }}_{\mathrm {q}}(0)\otimes {\hat{\rho }}_{\mathrm {o}}(0)\) and \({\hat{\rho }}_{{\varvec{\ell }}\ne \mathbf {0}}(0)=0\), where \(\mathbf {0}=(0,0)\) is a two-dimensional zero vector. In our numerical simulations, we need to truncate the size of Hilbert space of the single-mode harmonic oscillator as well as number of the hierarchical equations. Unless otherwise stated, we restrict \({\hat{H}}_{\mathrm {qo}}\) in a \(16\times 16\) matrix (8 bosonic modes of the bath are involved). We also set a sufficiently large integer L as the cutoff order of the HEOM, which means all the terms of \({\hat{\rho }}_{{\varvec{\ell }}}(t)\) with \(\ell _{1}+\ell _{2}>L\) are set to be zero, and all the terms of \({\hat{\rho }}_{{\varvec{\ell }}}(t)\) with \(\ell _{1}+\ell _{2}\le L\) form a closed set of differential equations. These differential equations can be easily solved directly by using the traditional Runge–Kutta method. One can keep on increasing the hierarchy order L until the final result converges. The reduced density matrix of the qubit can be obtained by simply partially tracing out of the degrees of freedom of the single-mode harmonic oscillator: \({\hat{\rho }}_{\mathrm {q}}(t)=\mathrm {tr}_{\mathrm {o}}[{\hat{\rho }}_{{\varvec{\ell }}=\mathbf {0}}(t)]\).

Appendix B: The population difference in the longtime limit

The equilibrium steady value of the population difference, i.e., \(\langle {\hat{\sigma }}_{z}(t\rightarrow \infty )\rangle \), can be approximately obtained by making use of the unitary transformation method, which is extensively used in the studies of the Rabi model [58] and the spin-boson model [16, 23, 47, 48].

First, we apply a unitary transformation to the original Hamiltonian \({\hat{H}}\) as \({\hat{H}}^{\prime }=e^{{\hat{S}}}{\hat{H}}e^{-{\hat{S}}}\), where the generator \({\hat{S}}\) is given by \({\hat{S}}=\zeta {\hat{\sigma }}_{x}({\hat{a}}^{\dag }-{\hat{a}})\) with a undetermined parameter \(\zeta \). By using the Baker–Campbell–Hausdorff formula, the transformed Hamiltonian \({\hat{H}}^{\prime }\) can be exactly derived as follows

$$\begin{aligned} \begin{aligned} {\hat{H}}^{\prime }&=\frac{1}{2}\epsilon \left\{ \cosh \left( {\hat{\Xi }}\right) {\hat{\sigma }}_{z}-i\sinh \left( {\hat{\Xi }}\right) {\hat{\sigma }}_{y}\right\} +\frac{1}{2}\Delta {\hat{\sigma }}_{x}+\omega _{0}{\hat{a}}^{\dag }{\hat{a}}\\&\quad +\,(g_{0}-\zeta \omega _{0}){\hat{\sigma }}_{x}\left( {\hat{a}}^{\dag }+{\hat{a}}\right) +\zeta (\zeta \omega _{0}-2 g_{0})+\sum _{k}\omega _{k}{\hat{b}}_{k}^{\dag }{\hat{b}}_{k}\\&\quad +\,\left\{ \cosh \left( {\hat{\Xi }}\right) {\hat{\sigma }}_{z}-i\sinh \left( {\hat{\Xi }}\right) {\hat{\sigma }}_{y}\right\} \sum _{k}g_{k}\left( {\hat{b}}_{k}^{\dag }+{\hat{b}}_{k}\right) , \end{aligned} \end{aligned}$$
(6)

where \({\hat{\Xi }}\equiv 2\zeta ({\hat{a}}^{\dagger }-{\hat{a}})\). If we choose \(\zeta =g_{0}/\omega _{0}\) [23, 58], the qubit–oscillator coupling term vanishes. In this case, the transformed Hamiltonian can be rewritten as \({\hat{H}}^{\prime }={\hat{H}}^{\prime }_{0}+{\hat{H}}^{\prime }_{1}\), where

$$\begin{aligned} {\hat{H}}^{\prime }_{0}= & {} \frac{1}{2}\eta \epsilon {\hat{\sigma }}_{z}+\frac{1}{2}\Delta {\hat{\sigma }}_{x}+\sum _{k}\omega _{k}{\hat{b}}_{k}^{\dag }{\hat{b}}_{k}\nonumber \\&+\,\eta {\hat{\sigma }}_{z}\sum _{k}g_{k}\left( {\hat{b}}_{k}^{\dag }+{\hat{b}}_{k}\right) +\omega _{0}|\alpha |^{2}-\frac{g_{0}^{2}}{\omega _{0}}, \end{aligned}$$
(7)
$$\begin{aligned} {\hat{H}}^{\prime }_{1}= & {} \frac{1}{2}\epsilon \left\{ \cosh \left( {\hat{\Xi }}\right) -\eta \right\} {\hat{\sigma }}_{z}+\omega _{0}{\hat{a}}^{\dag }{\hat{a}}-\omega _{0}|\alpha |^{2}\nonumber \\&+\,\left\{ \cosh \left( {\hat{\Xi }}\right) -\eta \right\} {\hat{\sigma }}_{z}\sum _{k}g_{k}\left( {\hat{b}}_{k}^{\dag }+{\hat{b}}_{k}\right) \nonumber \\&-\,i\epsilon \sinh \left( {\hat{\Xi }}\right) {\hat{\sigma }}_{y}-i\sinh \left( {\hat{\Xi }}\right) {\hat{\sigma }}_{y}\sum _{k}g_{k}\left( {\hat{b}}_{k}^{\dag }+{\hat{b}}_{k}\right) . \end{aligned}$$
(8)

In this unitary transformation treatment, the parameter \(\eta \) is constructed such that \(\langle {\hat{H}}^{\prime }_{1}\rangle _{\mathrm {o}}=0\) [47, 48, 59], where \(\langle {\hat{H}}^{\prime }_{1}\rangle _{\mathrm {o}}\equiv \mathrm {tr}_{\mathrm {o}}({\hat{\rho }}_{\mathrm {o}}{\hat{H}}^{\prime }_{1})/\mathrm {tr}_{\mathrm {o}}({\hat{\rho }}_{\mathrm {o}})\) denotes the thermodynamic average value with respect to the equilibrium state of the single-mode harmonic oscillator. In an approximate estimation, one can regard \({\hat{\rho }}_{\mathrm {o}}={\hat{\rho }}_{\mathrm {o}}(0)=|\alpha \rangle \langle \alpha |\); then, the condition \(\langle {\hat{H}}^{\prime }_{1}\rangle _{\mathrm {o}}=0\) reduces to \(\langle \alpha |{\hat{H}}^{\prime }_{1}|\alpha \rangle =0\). It is easy to demonstrate that

$$\begin{aligned} \langle \alpha |\sinh [2\zeta ({\hat{a}}^{\dag }-{\hat{a}})]|\alpha \rangle =0;\quad \langle \alpha |{\hat{a}}^{\dag }{\hat{a}}|\alpha \rangle =|\alpha |^{2}. \end{aligned}$$

Then, if we choose the parameter \(\eta \) as follows

$$\begin{aligned} \begin{aligned} \eta&=\langle 0_{a}|e^{-\alpha ({\hat{a}}^{\dag }-{\hat{a}})}\cosh [2\zeta ({\hat{a}}^{\dag }-{\hat{a}})]e^{\alpha ({\hat{a}}^{\dag }-{\hat{a}})}|0_{a}\rangle \\&=\langle 0_{a}|\cosh [2\zeta ({\hat{a}}^{\dag }-{\hat{a}})]|0_{a}\rangle \\&=\exp \left[ -2\left( \frac{g_{0}}{\omega _{0}}\right) ^{2}\right] , \end{aligned} \end{aligned}$$
(9)

one can find that the condition \(\langle \alpha |{\hat{H}}^{\prime }_{1}|\alpha \rangle =0\) automatically satisfied. In a perturbative treatment, \({\hat{H}}^{\prime }_{1}\) can be neglected; then, the effective system is given by \({\hat{H}}^{\prime }_{\mathrm {eff}}={\hat{H}}^{\prime }_{0}\). It is quite clear to see the effect of the parameter \(\eta \) is a renormalization factor of \(\epsilon \), which renormalizes \(\epsilon \) to a smaller value (because \(\eta \) is a real number and smaller than one).

Then, the effective system \({\hat{H}}^{\prime }_{\mathrm {eff}}\) reduces to the well-known biased spin-boson model and has been widely studied by a variety of different methods. Due to the fact that the last two terms in Eq. 7 are constant which have no influences on the dynamical results, one can safely drop them. Following the detailed exposition in Refs. [47, 48, 59], one can obtain the equilibrium steady value of the population difference in this biased spin-boson model, which is given by

$$\begin{aligned} \begin{aligned} \langle {\hat{\sigma }}_{z}(t\rightarrow \infty )\rangle&\simeq \,-\,\frac{\eta \epsilon }{\sqrt{\Delta ^{2}+(\eta \epsilon )^{2}}}\\&=-\,\frac{\epsilon }{\sqrt{\left( \frac{\Delta }{\eta }\right) ^{2}+\epsilon ^{2}}}. \end{aligned} \end{aligned}$$
(10)

From the above expression, it is easy to check that a larger value of \(\zeta \) makes a larger steady value of the population difference \(\langle {\hat{\sigma }}_{z}(t\rightarrow \infty )\rangle \) which is in agreement with our numerical simulation in Fig. 2.

Appendix C: The measure of non-Markovianity

According to Ref. [33], the non-Markovianity can be described by the rate of the trace distance between two physical initial states. For a initial-state pair \({\hat{\rho }}_{1,2}(0)\) and a non-Markovian dynamical map \({\hat{\Lambda }}_{t}\) which can generate the mapping \({\hat{\rho }}(t)={\hat{\Lambda }}_{t}[{\hat{\rho }}(0)]\), the rate of change of the trace distance is defined by \(\varpi [t;{\hat{\rho }}_{1,2}(0)]\equiv \frac{\mathrm{d}}{\mathrm{d}t}D[{\hat{\rho }}_{1}(t),{\hat{\rho }}_{2}(t)]\) ,where \(D({\hat{\rho }}_{1},{\hat{\rho }}_{2})\equiv \frac{1}{2}\Vert {\hat{\rho }}_{1}-{\hat{\rho }}_{2}\Vert _{1}\) with \(\Vert {\hat{X}}\Vert _{1}\equiv \mathrm {tr}\sqrt{{\hat{X}}^{\dagger }{\hat{X}}}\) being the trace norm or the Schatten one-norm of an arbitrary operator \({\hat{X}}\). Then, a measure for the non-Markovianity of a given quantum process is written by [60, 61]

$$\begin{aligned} {\mathcal {N}}\equiv \max _{{\hat{\rho }}_{1,2}(0)}\frac{1}{2}\int _{0}^{t_{c}}\mathrm{d}t\left\{ |\varpi [t;{\hat{\rho }}_{1,2}(0)]|+\varpi [t;{\hat{\rho }}_{1,2}(0)]\right\} . \end{aligned}$$
(11)

where the time integration is extended over all time intervals \(t\in [0,t_{c}]\), during which \(\varpi [t;{\hat{\rho }}_{1,2}(0)]\) is positive, and the maximum runs over all possible initial-state pairs \({\hat{\rho }}_{1,2}(0)\).

According to Refs. [10, 25, 60,61,62], the numerical simulation of the non-Markovianity \({\mathcal {N}}\) can be further simplified by choosing a pair of initial state as two orthogonal pure states for the qubit-system case. Thus, in our numerical simulations, we assume the expressions of these initial-state pairs of the qubit are given by \({\hat{\rho }}_{\mathrm {q}1}(0)=|\varphi _{\mathrm {q}}(0)\rangle \langle \varphi _{\mathrm {q}}(0)|\) and \({\hat{\rho }}_{\mathrm {q}2}(0)=|\varphi _{\mathrm {q}}^{\perp }(0)\rangle \langle \varphi _{\mathrm {q}}^{\perp }(0)|\) with

$$\begin{aligned} |\varphi _{\mathrm {q}}(0)\rangle =\cos \left( \frac{\theta }{2}\right) |e\rangle +e^{i\phi }\sin \left( \frac{\theta }{2}\right) |g\rangle , \end{aligned}$$
(12)

and

$$\begin{aligned} |\varphi _{\mathrm {q}}^{\perp }(0)\rangle =\sin \left( \frac{\theta }{2}\right) |e\rangle -e^{i\phi }\cos \left( \frac{\theta }{2}\right) |g\rangle , \end{aligned}$$
(13)

where \(|e\rangle \) and \(|g\rangle \) are the excited and the ground states of Pauli z operator \({\hat{\sigma }}_{z}\), respectively. The ranges of the two initial-state parameters are given by \(\theta \in [0,\pi ]\) and \(\phi \in [0,2\pi ]\). By randomly generating a sufficiently large sample of initial-state parameter combinations \((\theta _{i},\phi _{i})\), one can find the optimal initial-state pair for the non-Markovianity. According to the definition in Eq. 11, the measure \({\mathcal {N}}\) is nonnegative, and we have \({\mathcal {N}}=0\) if and only if the process is Markovian. A nonzero value \({\mathcal {N}}>0\) implies a non-Markovian process. In our simulation, the upper bound of the time integration is \(t_{c}=50~\mathrm {ps}\), and the single-mode harmonic oscillator is approximately regarded as a \(4\times 4\) matrix due to the limitation of our computation resources.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wu, W., Cheng, JQ. Coherent dynamics of a qubit–oscillator system in a noisy environment. Quantum Inf Process 17, 300 (2018). https://doi.org/10.1007/s11128-018-2071-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11128-018-2071-y

Keywords

Navigation