Skip to main content
Log in

Excess Entropy Production in Quantum System: Quantum Master Equation Approach

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

For open systems described by the quantum master equation (QME), we investigate the excess entropy production under quasistatic operations between nonequilibrium steady states. The average entropy production is composed of the time integral of the instantaneous steady entropy production rate and the excess entropy production. We propose to define average entropy production rate using the average energy and particle currents, which are calculated by using the full counting statistics with QME. The excess entropy production is given by a line integral in the control parameter space and its integrand is called the Berry–Sinitsyn–Nemenman (BSN) vector. In the weakly nonequilibrium regime, we show that BSN vector is described by \(\ln \breve{\rho }_0\) and \(\rho _0\) where \(\rho _0\) is the instantaneous steady state of the QME and \(\breve{\rho }_0\) is that of the QME which is given by reversing the sign of the Lamb shift term. If the system Hamiltonian is non-degenerate or the Lamb shift term is negligible, the excess entropy production approximately reduces to the difference between the von Neumann entropies of the system. Additionally, we point out that the expression of the entropy production obtained in the classical Markov jump process is different from our result and show that these are approximately equivalent only in the weakly nonequilibrium regime.

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

Similar content being viewed by others

Notes

  1. For non-interacting system, \(A_n^O(\alpha )\) is calculated from the Brouwer formula, Brouwer [31], using the scattering matrix. Recently, the quantum pump in interacting systems has been actively researched. There are three theoretical approaches. The first is the Green’s function approach, Splettstoesser et al. [32]. The second is the generalized master equation approach [34, 35]. The third is the FCS-QME approach. Reference [25] showed the equivalence between the second and the third approaches for all orders of pumping frequency (see also Pluecker et al. [33]).

  2. In the research of adiabatic pumping, the expression of (41) is essential. In Refs. [23,24,25], (41) with (42) was used to study the quantum pump. On the other hand, in Ref. [35], (41) was derived using the generalized master equation [34] and without using the FCS. In Ref. [33], \(A_n^{O_\mu }(\alpha )\) was described by the quantity corresponding to the current operator and the pseudoinverse of the Liouvillian, as shown in (43). Reference [25] showed the equivalence between the FCS-QME approach and the generalized master equation approach for all orders of pumping frequency.

  3. Here, we supposed \(\frac{d}{dt}{\langle } O {\rangle }_{t} \approx i^O(t)\) for \(O=H_b,N_b\). However, because the thermodynamic parameters \(\beta _b\) and \(\mu _b\) are modulated, \(\frac{d}{dt}{\langle } H_b {\rangle }_t\) and \(\frac{d}{dt}{\langle } N_b {\rangle }_t\) also include the currents from the outside of the total system to the bath b.

References

  1. Landauer, R.: \(dQ=T dS\) far from equilibrium. Phys. Rev. A 18, 255 (1978)

    Article  ADS  Google Scholar 

  2. Oono, Y., Paniconi, M.: Steady state thermodynamics. Prog. Theor. Phys. Suppl. 130, 29–44 (1998)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  3. Sasa, S., Tasaki, H.: Steady state thermodynamics. J. Stat. Phys. 125, 125–227 (2006)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  4. Speck, T., Seifert, U.: Integral fluctuation theorem for the housekeeping heat. J. Phys. A: Math. Gen. 38, L581–L588 (2005)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  5. Esposito, M., Harbola, U., Mukamel, S.: Entropy fluctuation theorems in driven open systems: application to electron counting statistics. Phys. Rev. E 76, 031132 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  6. Esposito, M., Van den Broeck, C.: Three detailed fluctuation theorems. Phys. Rev. Lett. 104, 090601 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  7. Pradhan, P., Ramsperger, R., Seifert, U.: Approximate thermodynamic structure for driven lattice gases in contact. Phys. Rev. E 84, 041104 (2011)

    Article  ADS  Google Scholar 

  8. Crooks, G.E.: Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences. Phys. Rev. E 60, 2721 (1999)

    Article  ADS  Google Scholar 

  9. Seifert, U.: Entropy production along a stochastic trajectory and an integral fluctuation theorem. Phys. Rev. Lett. 95, 040602 (2005)

    Article  ADS  Google Scholar 

  10. Komatsu, T.S., Nakagawa, N., Sasa, S., Tasaki, H.: Steady-state thermodynamics for heat conduction: microscopic derivation. Phys. Rev. Lett. 100, 230602 (2008)

    Article  ADS  Google Scholar 

  11. Komatsu, T.S., Nakagawa, N., Sasa, S., Tasaki, H.: Entropy and nonlinear nonequilibrium thermodynamic relation for heat conducting steady states. J. Stat. Phys. 142, 127–153 (2011)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  12. Saito, K., Tasaki, H.: Extended Clausius relation and entropy for nonequilibrium steady states in heat conducting quantum systems. J. Stat. Phys. 145, 1275–1290 (2011)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  13. Ruelle, D.P.: Extending the definition of entropy to nonequilibrium steady states. Proc. Natl. Acad. Sci. USA 100, 3054–3058 (2003)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  14. Bertini, L., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Thermodynamic transformations of nonequilibrium states. J. Stat. Phys. 149, 773–802 (2012)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  15. Bertini, L., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Clausius inequality and optimality of quasistatic transformations for nonequilibrium stationary states. Phys. Rev. Lett. 110, 020601 (2013)

    Article  ADS  Google Scholar 

  16. Hatano, T., Sasa, S.: Steady-state thermodynamics of Langevin systems. Phys. Rev. Lett. 86, 3463 (2001)

    Article  ADS  Google Scholar 

  17. Sasa, S.: Possible extended forms of thermodynamic entropy. J. Stat. Mech. 2014, P01004 (2014)

    Article  MathSciNet  Google Scholar 

  18. Maes, C., Netočný, K.: A nonequilibrium extension of the Clausius heat theorem. J. Stat. Phys. 154, 188–203 (2014)

    Article  MATH  MathSciNet  Google Scholar 

  19. Sagawa, T., Hayakawa, H.: Geometrical expression of excess entropy production. Phys. Rev. E 84, 051110 (2011)

    Article  ADS  Google Scholar 

  20. Yuge, T., Sagawa, T., Sugita, A., Hayakawa, H.: Geometrical excess entropy production in nonequilibrium quantum systems. J. Stat. Phys. 153, 412–441 (2013)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  21. Komatsu, T.S., Nakagawa, N., Sasa, S., Tasaki, H.: Exact equalities and thermodynamic relations for nonequilibrium steady states. J. Stat. Phys. 159, 1237–1285 (2015)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  22. Sinitsyn, N.A., Nemenman, I.: The Berry phase and the pump flux in stochastic chemical kinetics. Europhys. Lett. 77, 58001 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  23. Yuge, T., Sagawa, T., Sugita, A., Hayakawa, H.: Geometrical pumping in quantum transport: quantum master equation approach. Phys. Rev. B 86, 235308 (2012)

    Article  ADS  Google Scholar 

  24. Watanabe, K., Hayakawa, H.: Non-adiabatic effect in quantum pumping for a spin-boson system. Prog. Theor. Exp. Phys. 2014, 113A01 (2014)

    Article  MATH  Google Scholar 

  25. Nakajima, S., Taguchi, M., Kubo, T., Tokura, Y.: Interaction effect on adiabatic pump of charge and spin in quantum dot. Phys. Rev. B 92, 195420 (2015)

    Article  ADS  Google Scholar 

  26. Esposito, M., Harbola, U., Mukamel, S.: Nonequilibrium fluctuations, fluctuation theorems, and counting statistics in quantum systems. Rev. Mod. Phys. 81, 1665–1702 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  27. Berry, M.V.: Quantal phase factors accompanying adiabatic change. Proc. R Soc. A 392, 45–57 (1984)

    Article  ADS  MATH  MathSciNet  Google Scholar 

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

    MATH  Google Scholar 

  29. Schaller, G., Brandes, T.: Preservation of positivity by dynamical coarse graining. Phys. Rev. A 78, 022106 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  30. Majenz, C., Albash, T., Breuer, H.P., Lidar, D.A.: Coarse graining can beat the rotating-wave approximation in quantum Markovian master equations. Phys. Rev. A 88, 012103 (2013)

    Article  ADS  Google Scholar 

  31. Brouwer, P.W.: Scattering approach to parametric pumping. Phys. Rev. B 58, 10135(R) (1998)

    Article  ADS  Google Scholar 

  32. Splettstoesser, J., Governale, M., König, J., Fazio, R.: Adiabatic pumping through interacting quantum dots. Phys. Rev. Lett. 95, 246803 (2005)

    Article  ADS  Google Scholar 

  33. Pluecker, T., Wegewijs, M.R., Splettstoesser, J.: Gauge freedom in observables and Landsberg’s nonadiabatic geometric phase: pumping spectroscopy of interacting open quantum systems. Phys. Rev. B 95, 155431 (2017)

    Article  ADS  Google Scholar 

  34. Cavaliere, F., Governale, M., König, J.: Nonadiabatic pumping through interacting quantum dots. Phys. Rev. L 103, 136801 (2009)

    Article  ADS  Google Scholar 

  35. Calvo, H.L., Classen, L., Splettstoesser, J., Wegewijs, M.R.: Interaction-induced charge and spin pumping through a quantum dot at finite bias. Phys. Rev. B 86, 245308 (2012)

    Article  ADS  Google Scholar 

  36. Utsumi, Y., Entin-Wohlman, O., Aharony, A., Kubo, T., Tokura, Y.: Fluctuation theorem for heat transport probed by a thermal probe electrode. Phys. Rev. B 89, 205314 (2014)

    Article  ADS  Google Scholar 

  37. Mandal, D., Jarzynski, C.: Analysis of slow transitions between nonequilibrium steady states. J. Stat. Mech. 2016, 063204 (2016)

    Article  MathSciNet  Google Scholar 

  38. Gorini, V., Frigerio, A., Verri, M., Kossakowski, A., Sudarshan, E.C.G.: Properties of quantum Markovian master equations. Rep. Math. Phys. 13, 149 (1978)

    Article  ADS  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgements

We acknowledge helpful discussions with S. Okada. Part of this work is supported by JSPS KAKENHI (26247051).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yasuhiro Tokura.

Appendices

Appendix A: Derivative of the von Neumann Entropy

We show that

$$\begin{aligned} \frac{\partial S_\mathrm{{vN}}(\rho _0(\alpha ))}{\partial \alpha ^n}=-\text{ Tr }_S \Big [\ln \rho _0(\alpha ) \frac{\partial \rho _0(\alpha )}{\partial \alpha ^n} \Big ] . \end{aligned}$$
(154)

From the definition of the von Neumann entropy, the LHS of the above equation is given by

$$\begin{aligned} \frac{\partial S_\mathrm{{vN}}(\rho _0(\alpha ))}{\partial \alpha ^n}=-\text{ Tr }_S \Big [\ln \rho _0(\alpha ) \frac{\partial \rho _0(\alpha )}{\partial \alpha ^n} \Big ]-\text{ Tr }_S \Big [ \frac{\partial \ln \rho _0(\alpha )}{\partial \alpha ^n} \rho _0(\alpha ) \Big ]. \end{aligned}$$
(155)

Using (151), the second term of the RHS of the above equation becomes

$$\begin{aligned} -\text{ Tr }_S \Big [ \frac{\partial \ln \rho _0(\alpha )}{\partial \alpha ^n} \rho _0(\alpha ) \Big ]= & {} - \text{ Tr }_S \Big [ \int _0^\infty ds \ \frac{1}{\rho _0(\alpha )+s}\frac{\partial \rho _0(\alpha )}{\partial \alpha ^n}\frac{1}{\rho _0(\alpha )+s} \rho _0(\alpha ) \Big ] \nonumber \\= & {} - \text{ Tr }_S \Big [ \int _0^\infty ds \ \frac{\rho _0(\alpha )}{(\rho _0(\alpha )+s)^2}\frac{\partial \rho _0(\alpha )}{\partial \alpha ^n} \Big ] \nonumber \\= & {} - \text{ Tr }_S \Big [ \frac{\partial \rho _0(\alpha )}{\partial \alpha ^n} \Big ]=0. \end{aligned}$$
(156)

Then, we obtain (154).

Appendix B: Derivation of the Relation Between \(k_{i,b}\) and \(\rho _{i,b}^{\kappa }\)

In this section, we examin the relation of the coefficients of the expansion of \(\rho _0^{\kappa }(\alpha )\) and \(l_0^\prime (\alpha )\) in (105) of 3.2.

First, we investigate \(k_{i,b}\) in (105). (92) can be rewritten as

$$\begin{aligned} \hat{K}^\dagger (\alpha )l_0^\prime (\alpha )+[\mathcal {K}^\prime (\alpha )]^\dagger 1 = J^\sigma _\mathrm{{ss}}(\alpha ). \end{aligned}$$
(157)

Here, \(J^\sigma _\mathrm{{ss}}(\alpha ) = \mathcal {O}(\varepsilon ^2)\) holds because \(i_\mathrm{{ss}}^{H_b}(\alpha ),i_\mathrm{{ss}}^{N_b}(\alpha )=\mathcal {O}(\varepsilon )\) and

$$\begin{aligned} J_\mathrm{{ss}}^\sigma (\alpha )=\sum _b \left( -i_\mathrm{{ss}}^{H_b}(\alpha )\varepsilon _{1,b}+i_\mathrm{{ss}}^{N_b}(\alpha )\varepsilon _{2,b}\right) \end{aligned}$$
(158)

since

$$\begin{aligned} \sum _b i_\mathrm{{ss}}^{X_b}(\alpha )= -\text{ Tr }_S\Big [X_S \sum _b \mathcal {L}_b(\alpha ) \rho _0(\alpha )\Big ]=0,\ \ \ (X=N,H). \end{aligned}$$
(159)

Then we obtain

$$\begin{aligned} \overline{\partial _{i,b}\mathcal {K}^\prime } ^\dagger 1+\overline{K}^\dagger k_{i,b}+\overline{\partial _{i,b}\mathcal {L}_{b}} ^\dagger \overline{l_0^\prime }= & {} 0, \end{aligned}$$
(160)

in \(\mathcal {O}(\varepsilon _{i,b})\). Here, \(\partial _{i,b}X {\mathop {=}\limits ^{\mathrm {def}}}\partial X/\partial \alpha _{i,b}\) and \(\overline{K}{\mathop {=}\limits ^{\mathrm {def}}}\overline{\hat{K}}\). The first term of the LHS is

$$\begin{aligned} \overline{\partial _{i,b}\mathcal {K}^\prime } ^\dagger 1= & {} \frac{\partial [\mathcal {K}^\prime ]^\dagger 1}{\partial \alpha _{i,b}}\Big \vert _{\alpha _{i,b}=\overline{\alpha _i}} \nonumber \\= & {} \frac{\partial \mathcal {L}_b^\dagger [\alpha _{1,b}H_S-\alpha _{2,b}N_S]}{\partial \alpha _{i,b}}\Big \vert _{\alpha _{i,b}=\overline{\alpha _i}} \nonumber \\= & {} \overline{\partial _{i,b} \mathcal {L}_b} ^\dagger [\overline{\beta }H_S-\overline{\beta \mu }N_S]+\overline{\Pi _b}^\dagger \frac{\partial [\alpha _{1,b}H_S-\alpha _{2,b}N_S]}{\partial \alpha _{i,b}}. \end{aligned}$$
(161)

The third term of the LHS becomes

$$\begin{aligned} \overline{\partial _{i,b}\mathcal {L}_{b}} ^\dagger \overline{l_0^\prime }= & {} \overline{\partial _{i,b}\mathcal {L}_{b}} ^\dagger ( -\overline{\beta }H_S+\overline{\beta \mu }N_S+\overline{c^\prime }^*1) \nonumber \\= & {} -\overline{\partial _{i,b}\mathcal {L}_{b}} ^\dagger (\overline{\beta }H_S-\overline{\beta \mu }N_S). \end{aligned}$$
(162)

Here, we used \(\overline{\partial _{i,b}\mathcal {L}_{b}}^\dagger 1=0\) derived from \(\hat{K}^\dagger 1=0\). Then, (160) becomes

$$\begin{aligned} \overline{K}^\dagger k_{1,b}+\overline{\Pi _b}^\dagger H_S= & {} 0, \end{aligned}$$
(163)
$$\begin{aligned} \overline{K}^\dagger k_{2,b}-\overline{\Pi _b}^\dagger N_S= & {} 0 . \end{aligned}$$
(164)

Next, we show the relation between \(k_{i,b}\) and \(\rho _{i,b}^{(-1)}\). (102) leads

$$\begin{aligned} \overline{K_\kappa }\rho _{i,b}^{(\kappa )}+\overline{\partial _{i,b}\mathcal {L}_b}\rho _\mathrm{{gc}} = 0 , \end{aligned}$$
(165)

in \(\mathcal {O}(\varepsilon _{i,b})\). Here, \(\overline{K_\kappa }{\mathop {=}\limits ^{\mathrm {def}}}\overline{\hat{K}}_\kappa \). By the way,

$$\begin{aligned} \mathcal {L}_b \rho _\mathrm{{gc}}(\alpha _S;\beta _b,\beta _b \mu _b)=0, \end{aligned}$$
(166)

holds. Differentiating this equation by \(\alpha _{i,b}\), we obtain

$$\begin{aligned} \overline{\partial _{i,b}\mathcal {L}_b}\rho _\mathrm{{gc}}= & {} -\overline{\mathcal {L}_b}\overline{\frac{\rho _\mathrm{{gc}}(\alpha _S;\beta _b,\beta _b \mu _b)}{\partial \alpha _{i,b}}} = \overline{\mathcal {L}_b}\frac{\partial [\alpha _{1,b} H_S-\alpha _{2,b}N_S]}{\partial \alpha _{i,b}}\rho _\mathrm{{gc}}(\alpha _S;\overline{\beta },\overline{\beta \mu }).\quad \quad \end{aligned}$$
(167)

Substituting these equations into (165), we obtain

$$\begin{aligned} \overline{K}_\kappa \rho _{1,b}^{(\kappa )}+\overline{\Pi _b}(H_S \rho _\mathrm{{gc}})= & {} 0, \end{aligned}$$
(168)
$$\begin{aligned} \overline{K}_\kappa \rho _{2,b}^{(\kappa )}-\overline{\Pi _b}(N_S \rho _\mathrm{{gc}})= & {} 0. \end{aligned}$$
(169)

Now, we use

$$\begin{aligned} \overline{\Pi _b}(\bullet \rho _\mathrm{{gc}})=(\overline{\Pi _b}^\dagger \bullet )\rho _\mathrm{{gc}} , \end{aligned}$$
(170)

which is derived from KMS condition (83). Using this relation, we rewire (168) and (169) as

$$\begin{aligned} \overline{K}_\kappa \rho _{1,b}^{(\kappa )}+(\overline{\Pi _b}^\dagger H_S) \rho _\mathrm{{gc}}= & {} 0, \end{aligned}$$
(171)
$$\begin{aligned} \overline{K}_\kappa \rho _{2,b}^{(\kappa )}-(\overline{\Pi _b}^\dagger N_S) \rho _\mathrm{{gc}}= & {} 0 . \end{aligned}$$
(172)

Multiplying \(\rho _\mathrm{{gc}}^{-1}\) from the right, we obtain

$$\begin{aligned} (\overline{K}_\kappa \rho _{1,b}^{(\kappa )})\rho _\mathrm{{gc}}^{-1}+\overline{\Pi _b}^\dagger H_S= & {} 0, \end{aligned}$$
(173)
$$\begin{aligned} (\overline{K}_\kappa \rho _{2,b}^{(\kappa )})\rho _\mathrm{{gc}}^{-1}-\overline{\Pi _b}^\dagger N_S= & {} 0 . \end{aligned}$$
(174)

(170) can be rewritten as

$$\begin{aligned} (\overline{\Pi _b}Y)\rho _\mathrm{{gc}}^{-1}=\overline{\Pi _b}^\dagger (Y\rho _\mathrm{{gc}}^{-1}), \end{aligned}$$
(175)

for any \(Y=\bullet \rho _\mathrm{{gc}} \in \mathrm{\varvec{B}}\) by multiplying \(\rho _\mathrm{{gc}}^{-1}\) from the right. (175) leads

$$\begin{aligned} (\overline{\Pi }\rho _{i,b}^{(\kappa )})\rho _\mathrm{{gc}}^{-1} =\overline{\Pi }^\dagger (\rho _{i,b}^{(\kappa )}\rho _\mathrm{{gc}}^{-1}) , \end{aligned}$$
(176)

where \(\overline{\Pi }{\mathop {=}\limits ^{\mathrm {def}}}\sum _b \overline{\Pi _b}\). By the way, \([H_S(\alpha _S),\rho _0^{(\kappa )}(\alpha )]= 0\) holds similarly to (86). Differentiating this equation by \(\alpha _{i,b}\), we obtain

$$\begin{aligned}{}[H_S(\alpha _S),\rho _{i,b}^{(\kappa )}] = 0. \end{aligned}$$
(177)

This relation leads

$$\begin{aligned} (\overline{H_\kappa ^{\times }}\rho _{i,b}^{(\kappa )})\rho _\mathrm{{gc}}^{-1}=\overline{H_\kappa ^{\times }}(\rho _{i,b}^{(\kappa )}\rho _\mathrm{{gc}}^{-1}) = \overline{H_{-\kappa }^{\times }}^\dagger (\rho _{i,b}^{(\kappa )}\rho _\mathrm{{gc}}^{-1}), \end{aligned}$$
(178)

where \(H_\kappa ^{\times }\bullet {\mathop {=}\limits ^{\mathrm {def}}}-i[H_S(\alpha _S)+\kappa H_\mathrm{{L}}(\alpha ),\bullet ]\). We used \((H_\kappa ^{\times })^\dagger =-H_\kappa ^{\times }\). In the first equality, we used that \(\rho _\mathrm{{gc}}\) commutes with \(H_S\) and \(H_\mathrm{{L}}\). (176) and (178) lead

$$\begin{aligned} (\overline{K}_\kappa \rho _{i,b}^{(\kappa )})\rho _\mathrm{{gc}}^{-1} = \overline{K}_{-\kappa }^\dagger (\rho _{i,b}^{(\kappa )}\rho _\mathrm{{gc}}^{-1}). \end{aligned}$$
(179)

Substituting this into (173) and (174), we obtain

$$\begin{aligned} \overline{K}_{-\kappa }^\dagger (\rho _{1,b}^{(\kappa )}\rho _\mathrm{{gc}}^{-1})+\overline{\Pi _b}^\dagger H_S= & {} 0 , \end{aligned}$$
(180)
$$\begin{aligned} \overline{K}_{-\kappa }^\dagger (\rho _{2,b}^{(\kappa )}\rho _\mathrm{{gc}}^{-1})-\overline{\Pi _b}^\dagger N_S= & {} 0 . \end{aligned}$$
(181)

Subtracting (180) ((181)) for \(\kappa =-1\) from (163) ((164)), we obtain

$$\begin{aligned} \overline{K}^\dagger (k_{i,b}-\rho _{i,b}^{(-1)}\rho _\mathrm{{gc}}^{-1}) = 0. \end{aligned}$$
(182)

This means

$$\begin{aligned} k_{i,b} = \rho _{i,b}^{(-1)}\rho _\mathrm{{gc}}^{-1}+\overline{c_{i,b}}1, \end{aligned}$$
(183)

where \(\overline{c_{i,b}}\) is an arbitrary complex number.

Appendix C: Definition of Entropy Production of the Markov Jump Process

Except (192), this section is based on Ref. [21]. We consider the Markov jump process on the states \(n=1,2,\ldots ,\mathcal {N}\):

$$\begin{aligned} n(t)= n_k (t_k \le t<t_{k+1}), \ t_0= 0<t_1<t_2 \cdots<t_n<t_{N+1}=\tau . \end{aligned}$$
(184)

where \(N=0,1,2,\ldots \) is the total number of jumps. We denote the above path by

$$\begin{aligned} \hat{n}= (N,(n_0,n_1,\ldots ,n_N),(t_1,t_2,\ldots ,t_N)) . \end{aligned}$$
(185)

The probability to find the system in a state n is \(p_n(t)\) and it obeys the master equation (119). We suppose the trajectory of the control \(\hat{\alpha }=\big ( \alpha (t) \big )_{t=0}^{\tau }\) is smooth. Now we introduce

$$\begin{aligned} \theta _{nm}(\alpha ) {\mathop {=}\limits ^{\mathrm {def}}}\left\{ \begin{array}{ll} - \ln \frac{K_{nm}(\alpha )}{K_{mn}(\alpha )} &{}\quad K_{nm}(\alpha ) \ne 0 \\ 0 &{}\quad K_{nm}(\alpha ) = 0 \end{array} \right. . \end{aligned}$$
(186)

If \(n\ne m\), this is entropy production of process \(m \rightarrow n\). The entropy production of process (185) is defined by

$$\begin{aligned} \Theta ^{\hat{\alpha }}[\hat{n}] = \sum _{k=1}^N \theta _{n_kn_{k-1}}(\alpha _{t_k}). \end{aligned}$$
(187)

Then the weight (the transition probability density) associated with a path \(\hat{n}\) is

$$\begin{aligned} \mathcal {T}^{\hat{\alpha }}[\hat{n}] = \prod _{k=1}^N K_{n_kn_{k-1}}(\alpha _{t_k}) \exp \Big [\sum _{k=0}^N \int _{t_k}^{t_{k+1}}dt \ K_{n_kn_k}(\alpha _t) \Big ] . \end{aligned}$$
(188)

The integral over all the paths is defined by

$$\begin{aligned} \int \mathcal {D}{\hat{n}}\ Y[{\hat{n}}] {\mathop {=}\limits ^{\mathrm {def}}}\sum _{N=0}^\infty \sum _{n_0,n_1,\ldots ,n_N}^{n_{k-1}\ne n_k}\int _0^\tau dt_1 \int _{t_1}^\tau dt_2 \int _{t_3}^\tau dt_3 \cdots \int _{t_{N-1}}^\tau dt_N \ Y[{\hat{n}}], \end{aligned}$$
(189)

and the expectation value of \(X[{\hat{n}}]\) is defined by

$$\begin{aligned} \langle X \rangle ^{\hat{\alpha }}{\mathop {=}\limits ^{\mathrm {def}}}\int \mathcal {D}{\hat{n}}\ X[{\hat{n}}]p_{n_0}^\mathrm{{ss}}(\alpha _0) \mathcal {T}^{\hat{\alpha }}[{\hat{n}}] . \end{aligned}$$
(190)

Here, \(p_{n}^\mathrm{{ss}}(\alpha )\) is the instantaneous stationary probability distribution characterized by \(\sum _m K_{nm}(\alpha )p_m^\mathrm{{ss}}(\alpha ) = 0\). We introduce a matrix \(K^\lambda (\alpha )\) by

$$\begin{aligned}{}[K^\lambda (\alpha )]_{nm} {\mathop {=}\limits ^{\mathrm {def}}}K_{nm}(\alpha )e^{i\lambda \theta _{nm}(\alpha )} . \end{aligned}$$
(191)

Then, the k-th order moment of the entropy production is given by

$$\begin{aligned} \langle (\Theta ^{\hat{\alpha }}[{\hat{n}}])^k \rangle ^{\hat{\alpha }}= \frac{\partial ^k}{\partial (i\lambda )^k}\Big \vert _{\lambda =0} \sum _{n,m} \Big [ \mathrm{{T}} \exp \Big [ \int _0^\tau dt \ K^\lambda (\alpha _t) \Big ] \Big ]_{nm} p_m^\mathrm{{ss}}(\alpha _0) . \end{aligned}$$
(192)

In particular, the average is given by

$$\begin{aligned} \sigma ^\mathrm{{C}} {\mathop {=}\limits ^{\mathrm {def}}}\langle \Theta ^{\hat{\alpha }}[{\hat{n}}] \rangle ^{\hat{\alpha }}= \int _0^\tau dt \ \sum _{n,m} \sigma _{nm}^\mathrm{{C}}(\alpha _t)p_m(t), \end{aligned}$$
(193)

where

$$\begin{aligned} \sigma _{nm}^\mathrm{{C}}(\alpha ) {\mathop {=}\limits ^{\mathrm {def}}}K_{nm}(\alpha ) \theta _{nm}(\alpha ) = -K_{nm}(\alpha ) \ln \frac{K_{nm}(\alpha )}{K_{mn}(\alpha )}. \end{aligned}$$
(194)

According to Ref. [21], for a quasi-static operation,

$$\begin{aligned} \sigma _\mathrm{{ex}}^\mathrm{{C}}= S_\mathrm{{Sh}}[p^\mathrm{{ss}}(\alpha _\tau )]-S_\mathrm{{Sh}}[p^\mathrm{{ss}}(\alpha _0)]+\mathcal {O}(\varepsilon ^2\delta ), \end{aligned}$$
(195)

holds where

$$\begin{aligned} \sigma _\mathrm{{ex}}^\mathrm{{C}}{\mathop {=}\limits ^{\mathrm {def}}}\sigma ^\mathrm{{C}}-\int _0^\tau dt \ \sum _{n,m} \sigma _{nm}^\mathrm{{C}}(\alpha _t)p_m^\mathrm{{ss}}(\alpha _t) , \end{aligned}$$
(196)

and \( S_\mathrm{{Sh}}[p]{\mathop {=}\limits ^{\mathrm {def}}}-\sum _n p_n\ln p_n\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nakajima, S., Tokura, Y. Excess Entropy Production in Quantum System: Quantum Master Equation Approach. J Stat Phys 169, 902–928 (2017). https://doi.org/10.1007/s10955-017-1895-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1895-7

Keywords

Navigation