Skip to main content
Log in

Quantifying the irreversibility of channels

  • Research Articles
  • Published:
Theoretical and Mathematical Physics Aims and scope Submit manuscript

Abstract

In contrast to unitary evolutions, which are reversible, generic quantum processes (operations and quantum channels) are often irreversible. However, the degree of irreversibility is different for different channels, and it is desirable to have a quantitative characterization of irreversibility. In this paper, by exploiting the channel–state duality implemented by the Jamiołkowski–Choi isomorphism, we quantify the irreversibility of channels via entropy of the Jamiołkowski–Choi states of the corresponding channels and compare it with the notions of entanglement fidelity and entropy exchange. General properties of a reasonable measure of irreversibility are discussed from an intuitive perspective, and entropic measures of irreversibility are introduced. Several relations between irreversibility, entanglement fidelity, the degree of nonunitality, and decorrelating power are established. Some measures of irreversibility for a variety of prototypical channels are evaluated explicitly, revealing some information-theoretic aspects of the structure of channels from the perspective of irreversibility.

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

References

  1. E. Fermi, Thermodynamics, Dover, Mineola, NY (2012).

    Google Scholar 

  2. E. Schrödinger, Statistical Thermodynamics, Dover, Mineola, NY (1989).

    Google Scholar 

  3. L. D. Landau and E. M. Lifshitz, Course of Theoretical Physics, Vol. 5: Statistical Physics, Butterworth-Heinemann, Oxford (1980).

    Google Scholar 

  4. H. B. Callen, Thermodynamics and an Introduction to Thermostatistics, Wiley, New York (1991).

    Google Scholar 

  5. W. H. Zurek, “Pointer basis of quantum apparatus: Into what mixture does the wave packet collapse?,” Phys. Rev. D, 24, 1516–1525 (1981).

    Article  ADS  MathSciNet  Google Scholar 

  6. W. H. Zurek, “Decoherence and the transition from quantum to classical,” Physics Today, 44, 36–44 (1991).

    Article  Google Scholar 

  7. E. Joos, H. D. Zeh, C. Kiefer, D. Giulini, J. Kupsch, I.-O. Stamatescu, Decoherence and the Appearance of a Classical World in Quantum Theory, Springer, Berlin (2003).

    Book  Google Scholar 

  8. M. Schlosshauer, “Decoherence, the measurement problem, and interpretations of quantum mechanics,” Rev. Mod. Phys., 76, 1267–1305 (2004).

    Article  ADS  Google Scholar 

  9. M. Schlosshauer, Decoherence and the Quantum-To-Classical Transition, Springer, Berlin (2007).

    Google Scholar 

  10. H.-P. Breuer and F. Petruccione, The Theory of Open Quantum Systems, Oxford Univ. Press, Oxford (2007).

    Book  Google Scholar 

  11. M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information, Cambridge Univ. Press, Cambridge (2010).

    Google Scholar 

  12. J. Watrous, The Theory of Quantum Information, Cambridge Univ. Press, Cambridge (2018).

    Book  Google Scholar 

  13. A. S. Holevo, Quantum Systems, Channels, Information: A Mathematical Introduction (De Gruyter Studies in Mathematical Physics, Vol. 16), De Gruyter, Berlin, Boston (2019).

    Book  Google Scholar 

  14. B. W. Schumacher, “Sending entanglement through noisy quantum channels,” Phys. Rev. A, 54, 2614–2628 (1996).

    Article  ADS  Google Scholar 

  15. B. Schumacher and M. A. Nielsen, “Quantum data processing and error correction,” Phys. Rev. A, 54, 2629–2635 (1996).

    Article  ADS  MathSciNet  Google Scholar 

  16. M. A. Nielsen, “The entanglement fidelity and quantum error correction,” arXiv: quant-ph/9606012.

  17. H. Barnum, M. A. Nielsen, and B. W. Schumacher, “Information transmission through a noisy quantum channel,” Phys. Rev. A, 57, 4153–4175 (1998).

    Article  ADS  Google Scholar 

  18. M. A. Nielsen, C. M. Caves, B. Schumacher, and H. Barnum, “Information-theoretic approach to quantum error correction and reversible measurement,” Proc. R. Soc. London Ser. A, 454, 277–304 (1998).

    Article  ADS  MathSciNet  Google Scholar 

  19. M. Horodecki, P. Horodecki, and R. Horodecki, “General teleportation channel, singlet fraction, and quasidistillation,” Phys. Rev. A, 60, 1888–1898 (1999).

    Article  ADS  MathSciNet  Google Scholar 

  20. M. A. Nielsen, “A simple formula for the average gate fidelity of a quantum dynamical operation,” Phys. Lett. A, 303, 249–252 (2002).

    Article  ADS  MathSciNet  Google Scholar 

  21. S. Luo, “Information conservation and entropy change in quantum measurements,” Phys. Rev. A, 82, 052103, 5 pp. (2010).

    Article  ADS  MathSciNet  Google Scholar 

  22. S. Luo and N. Li, “Decoherence and measurement-induced correlations,” Phys. Rev. A, 84, 052309, 8 pp. (2011).

    Article  ADS  Google Scholar 

  23. H.-P. Breuer, E.-M. Laine, and J. Piilo, “Measure for the degree of non-Markovian behavior of quantum processes in open systems,” Phys. Rev. Lett., 103, 210401, 4 pp. (2009).

    Article  ADS  MathSciNet  Google Scholar 

  24. Á. Rivas, S. F. Huelga, and M. B. Plenio, “Entanglement and non-Markovianity of quantum evolutions,” Phys. Rev. Lett., 105, 050403, 4 pp. (2010).

    Article  ADS  MathSciNet  Google Scholar 

  25. S. Luo, S. Fu, and H. Song, “Quantifying non-Markovianity via correlations,” Phys. Rev. A, 86, 044101, 4 pp. (2012).

    Article  ADS  Google Scholar 

  26. D. Chruściński and S. Maniscalco, “Degree of non-Markovianity of quantum evolution,” Phys. Rev. Lett., 112, 120404, 5 pp. (2014).

    Article  ADS  Google Scholar 

  27. H.-P. Breuer, E.-M. Laine, J. Piilo, and B. Vacchini, “Non-Markovian dynamics in open quantum systems,” Rev. Mod. Phys., 88, 021002, 24 pp. (2016).

    Article  ADS  Google Scholar 

  28. H. Spohn, “Entropy production for quantum dynamical semigroups,” J. Math. Phys., 19, 1227–1230 (1978).

    Article  ADS  MathSciNet  Google Scholar 

  29. G. E. Crooks, “Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences,” Phys. Rev. E, 60, 2721–2726 (1999).

    Article  ADS  Google Scholar 

  30. L. M. Martyushev and V. D. Seleznev, “Maximum entropy production principle in physics, chemistry and biology,” Phys. Rep., 426, 1–45 (2006).

    Article  ADS  MathSciNet  Google Scholar 

  31. J. M. R. Parrondo, C. Van den Broeck, and R. Kawai, “Entropy production and the arrow of time,” New J. Phys., 11, 073008, 14 pp. (2009).

    Article  ADS  Google Scholar 

  32. S. Deffner and E. Lutz, “Nonequilibrium entropy production for open quantum systems,” Phys. Rev. Lett., 107, 140404, 5 pp. (2011).

    Article  ADS  Google Scholar 

  33. T. Sagawa and M. Ueda, “Role of mutual information in entropy production under information exchanges,” New J. Phys., 15, 125012, 23 pp. (2013).

    Article  ADS  MathSciNet  Google Scholar 

  34. M. Lostaglio, K. Korzekwa, D. Jennings, and T. Rudolph, “Quantum coherence, time- translation symmetry, and thermodynamics,” Phys. Rev. X, 5, 021001, 11 pp. (2015); arXiv: 1410.4572.

    Google Scholar 

  35. H. Hossein-Nejad, E. J. O’Reilly, and A. Olaya-Castro, “Work, heat and entropy production in bipartite quantum systems,” New J. Phys., 17, 075014, 11 pp. (2015).

    Article  ADS  Google Scholar 

  36. M. Popovic, B. Vacchini, and S. Campbell, “Entropy production and correlations in a controlled non-Markovian setting,” Phys. Rev. A, 98, 012130, 8 pp. (2018).

    Article  ADS  Google Scholar 

  37. J. P. Santos, L. C. Céleri, F. Brito, G. T. Landi, and M. Paternostro, “Spin-phase-space-entropy production,” Phys. Rev. A, 97, 052123, 10 pp. (2018).

    Article  ADS  Google Scholar 

  38. K. Ptaszyǹski and M. Esposito, “Entropy production in open systems: The predominant role of intraenvironment correlations,” Phys. Rev. Lett., 123, 200603, 6 pp. (2019).

    Article  ADS  Google Scholar 

  39. P. Strasberg and A. Winter, “First and second law of quantum thermodynamics: A consistent derivation based on a microscopic definition of entropy,” PRX Quantum, 2, 030202, 26 pp. (2021).

    Article  ADS  Google Scholar 

  40. G. T. Landi and M. Paternostro, “Irreversible entropy production: From classical to quantum,” Rev. Mod. Phys., 93, 035008, 58 pp. (2021).

    Article  ADS  MathSciNet  Google Scholar 

  41. A. Belenchia, M. Paternostro, and G. T. Landi, “Informational steady states and conditional entropy production in continuously monitored systems: The case of Gaussian systems,” Phys. Rev. A, 105, 022213, 14 pp. (2022).

    Article  ADS  MathSciNet  Google Scholar 

  42. A. Jamiołkowski, “Linear transformations which preserve trace and positive semidefinitness of operators,” Rep. Math. Phys., 3, 275–278 (1972).

    Article  ADS  MathSciNet  Google Scholar 

  43. M.-D. Choi, “Completely positive maps on complex matrices,” Linear Algebra Appl., 10, 285–290 (1975).

    Article  MathSciNet  Google Scholar 

  44. M. Jiang, S. Luo, and S. Fu, “Channel-state duality,” Phys. Rev. A, 87, 022310, 8 pp. (2013).

    Article  ADS  Google Scholar 

  45. K. Życzkowski and I. Bengtsson, “On duality between quantum maps and quantum states,” Open Syst. Inf. Dyn., 11, 3-42 (2004).

    Article  MathSciNet  Google Scholar 

  46. Y. Sun, N. Li, and S. Luo, “Fidelity-disturbance-entropy tradeoff in quantum channels,” Phys. Rev. A, 105, 062458, 9 pp. (2022).

    Article  ADS  MathSciNet  Google Scholar 

  47. K. Banaszek, “Fidelity balance in quantum operations,” Phys. Rev. Lett., 86, 1366–1369 (2001).

    Article  ADS  Google Scholar 

  48. S. Luo, S. Fu, and N. Li, “Decorrelating capabilities of operations with application to decoherence,” Phys. Rev. A, 82, 052122, 12 pp. (2010).

    Article  ADS  Google Scholar 

  49. G. Lüders, “Über die Zustandsänderung durch den Meßprozeß,” Ann. Physik, 443, 322–328 (1950).

    Article  ADS  Google Scholar 

  50. J. M. Renes, R. Blume-Kohout, A. J. Scott, and C. M. Caves, “Symmetric informationally complete quantum measurements,” J. Math. Phys, 45, 2171–2180 (2004).

    Article  MathSciNet  Google Scholar 

  51. R. F. Werner and A. S. Holevo, “Counterexample to an additivity conjecture for output purity of quantum channels,” J. Math. Phys., 43, 4353–4357 (2002).

    Article  ADS  MathSciNet  Google Scholar 

  52. F. Buscemi, G. Chiribella, and G. M. D’Ariano, “Inverting quantum decoherence by classical feedback from the environment,” Phys. Rev. Lett., 95, 090501, 4 pp. (2005).

    Article  ADS  Google Scholar 

  53. K. Brádler, P. Hayden, D. Touchette, and M. M. Wilde, “Trade-off capacities of the quantum Hadamard channels,” Phys. Rev. A, 81, 062312, 23 pp. (2010).

    Article  ADS  Google Scholar 

  54. C. H. Bennett, G. Brassard, C. Crépeau, R. Jozsa, A. Peres, and W. K. Wootters, “Teleporting an unknown quantum state via dual classical and Einstein–Podolsky–Rosen channels,” Phys. Rev. Lett., 70, 1895–1899 (1993).

    Article  ADS  MathSciNet  Google Scholar 

  55. S. Popescu, “Bell’s inequalities versus teleportation: What is nonlocality?,” Phys. Rev. Lett., 72, 797–799 (1994).

    Article  ADS  MathSciNet  Google Scholar 

  56. M. Horodecki and P. Horodecki, “Reduction criterion of separability and limits for a class of distillation protocols,” Phys. Rev. A, 59, 4206–4216 (1999).

    Article  ADS  Google Scholar 

  57. K. Banaszek, “Optimal quantum teleportation with an arbitrary pure state,” Phys. Rev. A, 62, 024301, 4 pp. (2000).

    Article  ADS  Google Scholar 

  58. G. Bowen and S. Bose, “Teleportation as a depolarizing quantum channel, relative entropy, and classical capacity,” Phys. Rev. Lett., 87, 267901, 4 pp. (2001).

    Article  ADS  Google Scholar 

  59. M. S. Pinsker, Information and Information Stability of Random Variables and Processes, Holden-Day, San Francisco, CA (1964).

    Google Scholar 

  60. C. Tsallis, “Possible generalization of Boltzmann–Gibbs statistics,” J. Stat. Phys., 52, 479–487 (1988).

    Article  ADS  MathSciNet  Google Scholar 

Download references

Funding

This work was supported by the National Key R&D Program of China (grant No. 2020YFA0712700) and the National Natural Science Foundation of China (grant No. 12005104).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shunlong Luo.

Ethics declarations

The authors of this work declare that they have no conflicts of interest.

Additional information

Prepared from an English manuscript submitted by the author; for the Russian version, see Teoreticheskaya i Matematicheskaya Fizika, 2024, Vol. 218, pp. 492–521 https://doi.org/10.4213/tmf10607.

Publisher’s note. Pleiades Publishing remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: Proofs of Propositions 1–5

Below, we give the detailed proofs of Propositions 15. We further discuss an alternative measure of irreversibility in terms of the Tsallis entropy, which is easier to compute than the irreversibility based on the von Neumann entropy. We make a comparative study of this quantity and that introduced in the main text.

Appendix A: Proof of Proposition 1

1. By the properties of von Neumann entropy, we have \(0\le S(J_{\mathcal E})\le\ln d^2\), \(S(J_{\mathcal E})=0\) if and only if \(J_{\mathcal E}\) is a pure state, and \(S(J_{\mathcal E})=\ln d^2\) if and only if \(J_{\mathcal E}=\frac{1}{d^2}\mathbf 1\otimes\mathbf 1\) is a maximally mixed state on \(H\otimes H\). Consequently,

$$0\le S(\mathcal E)=\frac{1}{2}S(J_\mathcal E)\le\ln d.$$
It is clear that \(S(\mathcal E)=0\) if and only if \(J_{\mathcal E}\) is a pure state, and \(S(\mathcal E)=\ln d\) if and only if \(J_{\mathcal E}=\mathbf 1\otimes\mathbf 1/d^2\) is a maximally mixed state on \(H\otimes H\). Next, we show that \(J_{\mathcal E}\) is a pure state if and only if \(\mathcal E\) is a unitary channel, and \(J_{\mathcal E}=\frac{1}{d^2}\mathbf 1\otimes\mathbf 1\) is equivalent to \(\mathcal E\) being the completely depolarizing channel. For the first equivalence, we suppose that \(\mathcal E(\rho)=U\rho U^\dagger\) is a unitary channel; then
$$J_{\mathcal E}=|\Phi^{+}_U\rangle\langle\Phi^{+}_U|,$$
where \(|\Phi^{+}_U\rangle=\frac{1}{\sqrt d}\sum_{i}|i\rangle\otimes U|i\rangle\) is a pure state, with \(S(\mathcal E)=0\). Conversely, we suppose that
$$J_{\mathcal E}=\frac{1}d\sum_{i,j}|i\rangle\langle j|\otimes\mathcal E(|i\rangle\langle j|)$$
is a pure state; then \(J_{\mathcal E}\) is a rank-1 operator, which makes \(\mathcal E(|i\rangle\langle j|)\) an operator with rank not greater than 1 for any \(i\) and \(j\). Because \(\mathcal E(|i\rangle\langle i|)\) are quantum states, \(\operatorname{rank}(\mathcal E(|i\rangle\langle i|))=1\) for any \(i=1,2,\ldots,d\). If there is \(\mathcal E(|i\rangle\langle j|)=0\) for some \(i\neq j\), we have \(\operatorname{rank}(J_\mathcal E)\ge\operatorname{rank}(\mathcal E(|i\rangle\langle i|))+\operatorname{rank}(\mathcal E(|j\rangle\langle j|))=2\), which contradicts the rank condition \(\operatorname{rank}(J_{\mathcal E})=1\). Thus, \(\mathcal E(|i\rangle\langle j|)\) is a rank-1 operator for any \(i\) and \(j\). Combining \(\operatorname{rank}(J_{\mathcal E})=1\) and \(\operatorname{rank}(\mathcal E(|i\rangle\langle i|))=1\) for any \(i\), we see that \(\mathcal E(|i\rangle\langle j|)\) can be written as
$$\mathcal E(|i\rangle\langle j|)=|\phi_i\rangle\langle\phi_j|,\qquad i,j=1,2,\ldots,d,$$
where \(\{|\phi_i\rangle\colon i=1,2,\ldots,d\}\) is a set of pure states. In what follows, we show that \(\{|\phi_i\rangle:i=1,2,\ldots,d\}\) constitutes an orthonormal basis. By
$$\mathcal E(|i\rangle\langle i|)=\sum_k E_k|i\rangle\langle i|E_k^\dagger=|\phi_i\rangle\langle\phi_i|,$$
we have \(E_k|i\rangle=c_{k i}|\phi_i\rangle\) for some complex numbers \(c_{ki}\). Thus,
$$\mathcal E(|i\rangle\langle j|)=\biggl(\sum_k c_{ki}\bar c_{kj}\biggr)|\phi_i\rangle\langle\phi_j|=|\phi_i\rangle\langle\phi_j|,$$
which implies that \(\sum_k c_{k i}\bar c_{k j}=1\) for any \(i\), \(j\). The orthogonality of \(\{|\phi_i\rangle\colon i=1,2,\ldots,d\}\) follows from
$$\langle i|j\rangle=\Big\langle i\bigg|\sum_k E_k^\dagger E_k\bigg|j\Big\rangle= \biggl(\sum_k\bar c_{ki}c_{kj}\biggr)\langle\phi_i|\phi_j\rangle= \langle\phi_i|\phi_j\rangle,\qquad i,j=1,2,\ldots,d.$$
Thus, \(\mathcal E\) maps the orthonormal basis \(\{|i\rangle\colon i=1,2,\ldots,d\}\) to the orthonormal basis \(\{|\phi_i\rangle\colon i=1,2,\ldots,d\}\). Because \(S(\mathcal E)\) is independent of the choice of the orthonormal basis \(\{|i\rangle\colon i=1,2,\ldots,d\}\), \(\mathcal E\) maps any orthonormal basis to an orthonormal basis, which implies that \(\mathcal E\) is a unitary channel.

To prove the second equivalence, we suppose that \(\mathcal E\) is the completely depolarizing channel. Then

$$J_{\mathcal E}=\frac{1}{d}\sum_{i,j}|i\rangle\langle j|\otimes\mathcal E(|i\rangle\langle j|)=\frac{1}{d^2}\mathbf 1\otimes\mathbf 1$$
is the maximally mixed state on \(H\otimes H\). Conversely, if \(J_{\mathcal E}=\mathbf 1\otimes\mathbf 1/d^2\) is the maximally mixed state, then we conclude from
$$J_{\mathcal E}=\frac{1}d\sum_{i,j}|i\rangle\langle j|\otimes\mathcal E(|i\rangle\langle j|)= \frac{1}d\sum_{i}|i\rangle\langle i|\otimes\mathcal E(|i\rangle\langle i|)+ \frac{1}d\sum_{i\neq j}|i\rangle\langle j|\otimes\mathcal E(|i\rangle\langle j|)$$
that \(\mathcal E(|i\rangle\langle j|)=0\) for any \(i\neq j\) and \(\mathcal E(|i\rangle\langle i|)=\mathbf 1/d\) for any \(i\). Therefore,
$$\mathcal E(\rho)=\mathcal E\biggl(\sum_{i,j}\langle i|\rho|j\rangle|i\rangle\langle j|\biggr)= \sum_{i,j}\langle i|\rho|j\rangle\mathcal E(|i\rangle\langle j|)=\frac{1}d \mathbf 1$$
for any state \(\rho\), i.e., \(\mathcal E\) is the completely depolarizing channel.

2. Direct calculation shows that

$$J_{p_1\mathcal E_1+p_2\mathcal E_2}=p_1J_{\mathcal E_1}+p_2J_{\mathcal E_2}.$$
By the concavity of von Neumann entropy, we have
$$\begin{aligned} \, S(p_1\mathcal E_1+p_2\mathcal E_2)&=\frac{1}{2}S(J_{p_1\mathcal E_1+p_2\mathcal E_2})=\frac{1}{2}S(p_1J_{\mathcal E_1}+p_2J_{\mathcal E_2})\ge \\ &\ge\frac{1}{2}\bigl(p_1S(J_{\mathcal E_1})+p_2S(J_{\mathcal E_2})\bigr)=p_1S(\mathcal E_1)+p_2S(\mathcal E_2). \end{aligned}$$

3. Let \(U\) and \(V\) be any unitary operators. Then

$$\begin{aligned} \, J_{\mathcal E_U \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E_V}&= \frac{1}{d}\sum_{i,j}|i\rangle\langle j|\otimes U\mathcal E( V|i\rangle\langle j|V^\dagger)U^\dagger= \\ &=(V^\dagger\otimes U)\biggl(\frac{1}{d}\sum_{i,j}V|i\rangle\langle j|V^\dagger\otimes\mathcal E(V|i\rangle\langle j|V^\dagger)\biggr) (V^\dagger\otimes U)^\dagger. \end{aligned}$$
Thus
$$S(\mathcal E_U \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E_V)= \frac{1}{2}S(J_{\mathcal E_U \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E_V})= \frac{1}{2}S \biggl(\frac{1}{d}\sum_{i,j}V|i\rangle\langle j|V^\dagger\otimes\mathcal E(V|i\rangle\langle j|V^\dagger)\biggr)=\frac{1}{2}S(J_{\mathcal E})=S(\mathcal E)$$
whence item 3 follows.

4. Let \(H^a\) and \(H\) have the respective dimensions \(d_a\) and \(d\), and orthonormal bases \(\{|\mu\rangle\}\) and \(\{|i\rangle\}\). For the channel \(\mathcal I^a\otimes\mathcal E\) on the composite system Hilbert space \(H^a\otimes H\), we have

$$\begin{aligned} \, d_adJ_{\mathcal I^a\otimes\mathcal E}&= \sum_{\mu,\nu,i,j}|\mu\rangle\langle\nu|\otimes|i\rangle\langle j|\otimes|\mu\rangle \langle\nu|\otimes\mathcal E(|i\rangle\langle j|)= \\ &=\sum_{\mu,\nu,i,j}(\mathbf 1^a\otimes F_{}\otimes\mathbf 1)(|\mu\rangle\langle\nu|\otimes|\mu\rangle \langle\nu|\otimes|i\rangle\langle j|\otimes\mathcal E(|i\rangle\langle j|)) (\mathbf 1^a\otimes F_{}^\dagger\otimes\mathbf 1)= \\ &=d_ad(\mathbf 1^a\otimes F_{}\otimes\mathbf 1)(|\Phi^{+}_a\rangle\langle\Phi^{+}_a|\otimes J_{\mathcal E})(\mathbf 1^a\otimes F^\dagger_{}\otimes\mathbf 1), \end{aligned}$$
where \(F=\sum_{\mu i}|i\rangle\langle\mu|\otimes|\mu\rangle\langle i|\) is the swap operator on the composite system Hilbert space \(H^a\otimes H\) and \(|\Phi^{+}_a\rangle=\frac{1}{\sqrt{d_a}}\sum_\mu|\mu\rangle\otimes|\mu\rangle\). Consequently,
$$\begin{aligned} \, S(\mathcal I^a\otimes\mathcal E)=\frac{1}{2}S(J_{\mathcal I^a\otimes\mathcal E})&= \frac{1}{2}S\bigl((\mathbf 1^a\otimes F\otimes\mathbf 1)(|\Phi_a^{+}\rangle\langle\Phi_a^{+}|\otimes J_{\mathcal E})(\mathbf 1^a\otimes F^\dagger_{}\otimes\mathbf 1)\bigr)= \\ &=\frac{1}{2}S(J_{\mathcal E}\otimes|\Phi_a^{+}\rangle\langle\Phi_a^{+}|)=\frac{1}{2} S(J_{\mathcal E})=S(\mathcal E), \end{aligned}$$
whence item 4 follows.

5. Let \(\mathcal F\) be a unital channel on a \(d\)-dimensional system satisfying \(\mathcal F(\mathbf 1)=\mathbf 1\). By the definition of Jamiołkowski–Choi states, we have \(J_{\mathcal F \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E}=\mathcal I\otimes\mathcal F(J_{\mathcal E})\). It is obvious that \(\mathcal I\otimes{\mathcal F}\) is also a unital channel. In view of the monotonicity of von Neumann entropy for a unital channel, we obtain

$$S(\mathcal F \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E)= \frac{1}{2}S(J_{\mathcal F \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E})=\frac{1}{2}S(\mathcal I\otimes\mathcal F(J_{\mathcal E}))\ge \frac{1}{2}S(J_{\mathcal E})=S(\mathcal E).$$

6. Let \(\mathcal E^a\) and \(\mathcal E^b\) be channels on systems \(a\) and \(b\), with orthonormal bases \(\{|\mu\rangle\}\) and \(\{|i\rangle\}\). Because

$$J_{\mathcal E^a\otimes\mathcal E^b}= \frac{1}{d_ad_b}\sum_{\mu\nu ij}|\mu\rangle\langle\nu|\otimes|i\rangle\langle j|\otimes\mathcal E^a(|\mu\rangle\langle\nu|)\otimes\mathcal E^b(|i\rangle\langle j|),$$
we have
$$\begin{aligned} \, (\mathbf 1^a\otimes F_{ab}^\dagger\otimes\mathbf 1^b)&J_{\mathcal E^a\otimes\mathcal E^b}(\mathbf 1^a\otimes F_{ab}\otimes\mathbf 1^b)= \\ &=\frac{1}{d_ad_b}\sum_{\mu,\nu,i,j}|\mu\rangle\langle\nu|\otimes\mathcal E^a(|\mu\rangle\langle\nu|)\otimes |i\rangle\langle j|\otimes\mathcal E^b(| i\rangle\langle j|)=J_{\mathcal E^a}\otimes J_{\mathcal E^b}, \end{aligned}$$
where \(F_{ab}=\sum_{\mu i}|\mu\rangle\langle i|\otimes|i\rangle\langle\mu|\) is the swap operator on the system Hilbert space \(H^a\otimes H^b\). Consequently,
$$\begin{aligned} \, S(\mathcal E^a\otimes\mathcal E^b)=\frac{1}{2}S(J_{\mathcal E^a\otimes\mathcal E^b})&= \frac{1}{2}S \bigl((\mathbf 1^a\otimes F_{ab}^\dagger\otimes\mathbf 1^b) J_{\mathcal E^a\otimes\mathcal E^b}(\mathbf 1^a\otimes F_{ab}^\dagger\otimes\mathbf 1^b)\bigr)= \\ &=\frac{1}{2}S(J_{\mathcal E^a}\otimes J_{\mathcal E^b})= \frac{1}{2}S(J_{\mathcal E^a})+\frac{1}{2}S(J_{\mathcal E^b})= S(\mathcal E^a)+S(\mathcal E^b), \end{aligned}$$
whence item 6 follows.

Appendix B: Proof of Proposition 2

For any random unitary channel \(\mathcal E_{\mathrm{ru}}\), we have

$$J_{\mathcal E_{\mathrm{ru}}}= \sum_k p_k\biggl(\frac{1}{d}\sum_{i,j}|i\rangle\langle j|\otimes U_k|i\rangle\langle j|U_k^\dagger\biggr)= \sum_k p_k J_{\mathcal E_{U_k}},$$
whence
$$0\le S(J_{\mathcal E_{\mathrm{ru}}})= S\biggl(\sum_k p_k J_{\mathcal E_{U_k}}\biggr)\le H(\{p_k\})+\sum_k p_k S(J_{\mathcal E_{U_k}})=H(\{p_k\}),$$
which implies that
$$0\le S(\mathcal E_{\mathrm{ru}})\le\frac{1}{2}H(\{p_k\}).$$
By the properties of von Neumann entropy, \(S(J_{\mathcal E_{\mathrm{ru}}})=0\) if and only if \(\mathcal E_{\mathrm{ru}}\) is actually a unitary channel and \(S(J_{\mathcal E_{\mathrm{ru}}})=H(\{p_k\})\) if and only if the states \(J_{\mathcal E_{U_k}}\) have support on orthogonal subspaces, i.e., \( \operatorname{tr} (J_{\mathcal E_{U_k}}J_{\mathcal E_{U_l}})=0\) for any \(k\neq l\). Direct manipulation of the definition of a Jamiołkowski–Choi state shows that
$$\operatorname{tr} (J_{\mathcal E_{U_k}}J_{\mathcal E_{U_l}})=\frac{1}{d^2}| \operatorname{tr} U_k^\dagger U_l|^2.$$
Therefore, \(S(J_{\mathcal E_{\mathrm{ru}}})=H(\{p_k\})\) if and only if \( \operatorname{tr} (U_k^\dagger U_l)=0\) for any \(k\neq l\).

Appendix C: Proof of Proposition 3

To prove Proposition 3, we first recall Pinsker’s inequality, which states that [59]

$$ S(\rho|\sigma)\ge 2T^2(\rho,\sigma)$$
(15)
for any states \(\rho\) and \(\sigma\), where \(S(\rho|\sigma)= \operatorname{tr} \rho (\ln\rho-\ln\sigma)\) is the quantum relative entropy, and
$$T(\rho,\sigma)=\frac{1}{2} \operatorname{tr} |\rho-\sigma|$$
is the trace distance between states \(\rho\) and \(\sigma\) with \(|A|=\sqrt{A^\dagger A}\) for any operator \(A\).

By Pinsker’s inequality (15) and the triangle inequality for the trace distance, we have

$$S(J_\mathcal E|J_{\mathrm{cde}})\ge 2T^2(J_\mathcal E,J_{\mathrm{cde}})\ge 2\bigl(T(J_\mathcal E,J_{\mathcal I})-T(J_{\mathcal I},J_{\mathrm{cde}})\bigr)^2= 2\biggl(1-\frac{1}{d^2}-T(J_\mathcal E,J_{\mathcal I})\biggr)^{\!2},$$
where the equality holds by virtue of \(T(J_{\mathcal I},J_{\mathrm{cde}})=1-1/d^2\). By the well-known inequality relating fidelity and trace distance [11]
$$T(\rho,\sigma)\le\sqrt{1-F(\rho,\sigma)},$$
we further have
$$\begin{aligned} \, \sqrt{S(J_\mathcal E|J_{\mathrm{cde}})}&\ge\sqrt{2}\biggl(1-\frac1{d^2}-T(J_\mathcal E,J_{\mathcal I})\biggr)\ge \\ &\ge\sqrt{2}\biggl(1-\frac{1}{d^2}-\sqrt{1-F(J_\mathcal E,J_{\mathcal I})}\biggr)= \sqrt{2}\biggl(1-\frac{1}{d^2}-\sqrt{1-F(\mathcal E)}\biggr), \end{aligned}$$
whence the desired inequality follows.

Appendix D: Proof of Proposition 4

By Eqs. (9) and (12), we have

$$\begin{aligned} \, D(\mathcal E)&=I(J_{\mathcal I})-I(J_{\mathcal E})= 2S(\mathbf 1/d)-\bigl( S(\mathbf 1/d) +S(\mathcal E(\mathbf 1/d))- S(J_\mathcal E)\bigr)= \\ &=\ln d-S(\mathcal E(\mathbf 1/d))+S(J_\mathcal E)=N(\mathcal E)+2S(\mathcal E). \end{aligned}$$

Appendix E: Proof of Proposition 5

1. By Eq. (13), we have \(D(\mathcal E)\ge0\) and the equality holds if and only if \(S(J_\mathcal E)=0\) and \({S(\mathcal E(\mathbf 1/d)|\mathbf 1/d)\!=\!0}\), which implies that \(\mathcal E\) is a unitary channel. For the upper bound \(D(\mathcal E)\le 2\ln d\), noting that

$$\begin{aligned} \, D(\mathcal E)&=S(J_\mathcal E)+S(\mathcal E(\mathbf 1/d)|\mathbf 1/d)= S(J_\mathcal E)+S(\mathbf 1/d)-S(\mathcal E(\mathbf 1/d))= \\ &=2S(\mathbf 1/d)-S(J_\mathcal E|\mathbf 1/d\otimes\mathcal E(\mathbf 1/d))\le 2S(\mathbf 1/d)=2\ln d, \end{aligned}$$
we have \(D(\mathcal E)\le 2\ln d\) and \(D(\mathcal E)=2\ln d\) if and only if \(J_\mathcal E=\mathbf 1/d\otimes\mathcal E(\mathbf 1/d)\), which implies that \(\mathcal E(\rho)=\mathcal E(\mathbf 1/d)\) for any state \(\rho\).

2. Direct calculations show that

$$J_{p_1\mathcal E_1+p_2\mathcal E_2}=p_1J_{\mathcal E_1}+p_2J_{\mathcal E_2}.$$
By the joint convexity of relative entropy, we have
$$\begin{aligned} \, I(J_{p_1\mathcal E_1+p_2\mathcal E_2})&=S(J_{p_1\mathcal E_1+p_2\mathcal E_2}|\mathbf 1/d\otimes(p_1\mathcal E_1+p_2\mathcal E_2)(\mathbf 1/d))= \\ &=S(p_1J_{\mathcal E_1}+p_2J_{\mathcal E_2}|p_1\mathbf 1/d\otimes\mathcal E_1(\mathbf 1/d)+p_2\mathbf 1/d\otimes\mathcal E_2(\mathbf 1/d))\le \\ &\le p_1S(J_{\mathcal E_1}|\mathbf 1/d\otimes\mathcal E_1(\mathbf 1/d))+p_2S(J_{\mathcal E_2}|\mathbf 1/d\otimes\mathcal E_2(\mathbf 1/d))= \\ &=p_1I(J_{\mathcal E_1}) +p_2I(J_{\mathcal E_2}), \end{aligned}$$
whence item 2 follows.

3. By the unitary invariance of von Neumann entropy, we have

$$\begin{aligned} \, I(J_{\mathcal E_U \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E_V})&= S(\mathbf 1/d)+S(\mathcal E_U \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E_V(\mathbf 1/d))- S(\mathcal I\otimes\mathcal E_U \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E_V(|\Phi^{+}\rangle\langle\Phi^{+}|))= \\ &=S(\mathbf 1/d)+S(\mathcal E(\mathbf 1/d))-S(\mathcal I\otimes\mathcal E(|\Phi^{+}\rangle\langle\Phi^{+}|))=I(J_{\mathcal E}) \end{aligned}$$
for any unitary operators \(U\) and \(V\), which naturally implies the desired property.

4. Let \(H^a\) and \(H\) have respective dimensions \(d_a\) and \(d\), and orthonormal bases \(\{|\mu\rangle\}\) and \(\{|i\rangle\}\). For the channel \(\mathcal I^a\otimes\mathcal E\) on the composite system Hilbert space \(H^a\otimes H\), we have

$$\begin{aligned} \, d_a dJ_{\mathcal I^a\otimes\mathcal E}&= \sum_{\mu\nu ij}|\mu\rangle\langle\nu|\otimes|i\rangle\langle j|\otimes|\mu\rangle\langle\nu|\otimes\mathcal E (|i\rangle\langle j|)= \\ &=\sum_{\mu\nu ij}(\mathbf 1^a\otimes F_{ab}\otimes\mathbf 1) (|\mu\rangle\langle\nu|\otimes|\mu\rangle\langle\nu|\otimes|i\rangle\langle j|\otimes\mathcal E(|i\rangle\langle j|)) (\mathbf 1^a\otimes F^\dagger_{ab}\otimes\mathbf 1)= \\ &=d_a d(\mathbf 1^a\otimes F_{ab}\otimes\mathbf 1) (|\Phi^{+}_a\rangle\langle\Phi^{+}_a|\otimes J_{\mathcal E}) (\mathbf 1^a\otimes F^\dagger_{ab}\otimes\mathbf 1) \end{aligned}$$
where \(F_{ab}=\sum_{\mu i}|i\rangle\langle\mu|\otimes|\mu\rangle\langle i|\) is the swap operator on the composite system \(H^a\otimes H\) and \(|\Phi^{+}_a\rangle=\frac{1}{\sqrt{d_a}}\sum_\mu|\mu\rangle\otimes|\mu\rangle\). Consequently,
$$\begin{aligned} \, S(J_{\mathcal I^a\otimes\mathcal E})&= S\bigl((\mathbf 1^a\otimes F_{ab}\otimes\mathbf 1) (|\Phi_a^{+}\rangle\langle\Phi_a^{+}|\otimes J_{\mathcal E}) (\mathbf 1^a\otimes F^\dagger_{ab}\otimes\mathbf 1)\bigr)= \\ &=S(J_{\mathcal E}\otimes|\Phi_a^{+}\rangle\langle\Phi_a^{+}|)=S(J_{\mathcal E}) \end{aligned}$$
and
$$\begin{aligned} \, S(\mathbf 1^a/d_a\otimes\mathcal E(\mathbf 1/d)|\mathbf 1^a/{d_a}\otimes\mathbf 1/{d})&= S(\mathbf 1^a/d_a)+S(\mathbf 1/d)-S(\mathbf 1^a/d_a\otimes\mathcal E(\mathbf 1/d))= \\ &=S(\mathbf 1^a/d_a)+S(\mathbf 1/d)-S(\mathbf 1^a/d_a)-S(\mathcal E(\mathbf 1/d))= \\ &=S(\mathbf 1/d)-S(\mathcal E(\mathbf 1/d))=S(\mathcal E(\mathbf 1/d)|\mathbf 1/d), \end{aligned}$$
whence item 4 follows.

5. Let \(\mathcal F\) be any channel on a system Hilbert space \(H\). By the definition of Jamiołkowski–Choi states, we have \(J_{\mathcal F \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E}=\mathcal I\otimes\mathcal F(J_{\mathcal E})\). In view of the monotonicity of relative entropy, item 5 follows from

$$\begin{aligned} \, I(J_{\mathcal F \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E})=I(\mathcal I\otimes\mathcal F(J_{\mathcal E}))&= S(\mathcal I\otimes\mathcal F(J_{\mathcal E})|\mathcal I\otimes\mathcal{F}(\mathbf 1/d\otimes\mathcal E(\mathbf 1/d)))\le \\ &\le S(J_{\mathcal E}|\mathbf 1/d\otimes\mathcal E(\mathbf 1/d))=I(J_{\mathcal E}). \end{aligned}$$

6. Let \(\mathcal E^a\) and \(\mathcal E^b\) be channels on the systems \(a\) and \(b\), with respective Hilbert spaces \(H^a\) and \(H^b\), and orthonormal bases \(\{|\mu\rangle\}\) and \(\{|i\rangle\}\). Because

$$J_{\mathcal E^a\otimes\mathcal E^b}=\frac{1}{d_ad_b}\sum_{\mu\nu ij} |\mu\rangle\langle\nu|\otimes|i\rangle\langle j|\otimes\mathcal E^a(|\mu\rangle\langle\nu|)\otimes\mathcal E^b(|i\rangle\langle j|),$$
we have
$$(\mathbf 1^a\otimes F^\dagger\otimes\mathbf 1^b) J_{\mathcal E^a\otimes\mathcal E^b}(\mathbf 1^a\otimes F\otimes\mathbf 1^b)=\frac{1}{d_ad_b}\sum_{\mu\nu ij} |\mu\rangle\langle\nu|\otimes\mathcal E^a(|\mu\rangle\langle\nu|)\otimes|i\rangle\langle j|\otimes\mathcal E^b(|i\rangle\langle j|)= J_{\mathcal E^a}\otimes J_{\mathcal E^b},$$
where \(F_{ab}=\sum_{\mu i}|\mu\rangle\langle i|\otimes|i\rangle\langle\mu|\) is the swap operator on the system Hilbert space \(H^a\otimes H^b\). Consequently,
$$S(J_{\mathcal E^a\otimes\mathcal E^b})= S\bigl((\mathbf 1^a\otimes F_{ab}\otimes\mathbf 1^b)J_{\mathcal E^a\otimes\mathcal E^b} (\mathbf 1^a\otimes F^\dagger_{ab}\otimes\mathbf 1^b)\bigr)=S(J_{\mathcal E^a}\otimes J_{\mathcal E^b})=S(J_{\mathcal E^a})+S(J_{\mathcal E^b}),$$
and
$$\begin{aligned} \, S(\mathcal E^a(\mathbf 1^a/d_a)\otimes\mathcal E^b&(\mathbf 1^b/d_b)|\mathbf 1^a/d_a\otimes\mathbf 1^b/d_b)=S(\mathbf 1^a/d_a\otimes\mathbf 1^b/d_b)-S(\mathcal E^a(\mathbf 1^a/d_a)\otimes\mathcal E^b(\mathbf 1^b/d_b))= \\ &=S(\mathbf 1^a/d_a)+S(\mathbf 1^b/d_b)-S(\mathcal E^a(\mathbf 1^a/d_a))-S(\mathcal E^b(\mathbf 1^b/d_b))= \\ &=S(\mathcal E^a(\mathbf 1^a/d_a)|\mathbf 1^a/d_a)+S(\mathcal E^b(\mathbf 1^b/d_b|\mathbf 1^b/d_b), \end{aligned}$$
whence item 6 follows.

Appendix F: An alternative measure of irreversibility

We recall that the Tsallis \(r\)-entropy

$$S_r(\rho)=\frac {1- \operatorname{tr} \rho^r}{r-1},\qquad r\in\mathbb{R},$$

is a simple and significant quantity characterizing the mixedness of a state \(\rho\) [60]. The case \(r=1\) is understood as the limit \(r\to 1\) and actually corresponds to the von Neumann entropy. If we take \(\rho\) to be \(J_\mathcal E\) of a channel \(\mathcal E\), then we can regard

$$S_r(\mathcal E)=\frac{1}{2} S_r(J_\mathcal E)$$

as a measure of irreversibility of the channel \(\mathcal E(\rho)=\sum_kE_k\rho E_k^\dagger\). For \(r\in\mathbb{N}\), \(S_r(\mathcal E)\) has the explicit form

$$S_r^{}(\mathcal E)=\frac{1}{2(r-1)} \biggl(1-\frac{1}{d^r}\sum_{k_1,\ldots,k_r} \operatorname{tr} (E_{k_1}^\dagger E_{k_2}^{}) \operatorname{tr} (E_{k_2}^\dagger E_{k_3}^{})\ldots \operatorname{tr} (E_{k_r}^\dagger E_{k_1}^{})\biggr).$$

In particular, the Tsallis 2-entropy \(S_2(\rho)=1- \operatorname{tr} \rho^2\) is the linear entropy, and if we take \(\rho\) to be \(J_\mathcal E\) of a channel \(\mathcal E\), then we can regard

$$S_2(\mathcal E)=\frac{1}{2} S_2(J_\mathcal E)$$

as a measure of irreversibility of the channel \(\mathcal E\). It is interesting to note that this quantity can also be expressed as

$$S_2(\mathcal E)=\frac{1}{2}-\frac{1}{2d^2}\sum_{k}\|\mathcal E(X_k)\|^2$$

where \(\{X_k\colon k=1,\ldots, d^2\}\) is any orthonormal basis of the operator space \(L(H)\) of all observables (Hermitian operators) on \(H\) with the Hilbert–Schmidt inner product \(\langle A|B\rangle= \operatorname{tr} AB\). The quantity \(S_2(\mathcal E)\) satisfies the following properties, which parallel those of \(S(\mathcal E)\).

  1. 1.

    We have

    $$0\le S_2(\mathcal E)\le\frac{1}{2}-\frac{1}{2d^2},$$

    and \(S_2(\mathcal E)=0\) if and only if \(\mathcal E\) is a unitary channel, while \(S_2(\mathcal E)\) attains the maximum value \((d^2-1)/2d^2\) if and only if \(\mathcal E\) is the completely depolarizing channel \(\mathcal E_{\mathrm{cde}}(\rho)=\mathbf 1/d\) for any state \(\rho\).

  2. 2.

    \(S_2(\mathcal E)\) is concave in \(\mathcal E\), i.e.,

    $$S_2(p_1\mathcal E_1+p_2\mathcal E_2)\ge p_1 S_2(\mathcal E_1)+p_2 S_2(\mathcal E_2)$$

    for \(p_1,p_2\ge 0\), \(p_1+p_2=1\), and any channels \(\mathcal E_1\) and \(\mathcal E_2\).

  3. 3.

    \(S_2(\,{\cdot}\,)\) is invariant under composition with unitary dynamics in the sense that

    $$S_2(\mathcal E_U \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E)=S_2(\mathcal E \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E_U)=S_2(\mathcal E)$$

    for any unitary channel \(\mathcal E_U(\rho)=U\rho U^\dagger\) with \(U\) being any unitary operator on the system Hilbert space.

  4. 4.

    \(S_2(\,{\cdot}\,)\) is ancilla-independent in the sense that \(S_2(\mathcal I^a\otimes\mathcal E)=S_2(\mathcal E)\), where \(\mathcal I^a\) is the identity channel on any ancilla system \(a\).

  5. 5.

    We have

    $$S_2(\mathcal E^a\otimes\mathcal E^b)=S_2(\mathcal E^a)+S_2(\mathcal E^b)-S_2(\mathcal E^a)S_2(\mathcal E^b),$$

    where \(\mathcal E^a\) and \(\mathcal E^b\) are channels on systems \(a\) and \(b\). This is a kind of nonextensitivity of the Tsallis entropy.

  6. 6.

    \(S_2(\,{\cdot}\,)\) is monotonic in the sense that

    $$S_2(\mathcal F \mathbin{\stackrel{\scriptscriptstyle{\circ}}{{}_{\vphantom{.}}}} \mathcal E)\ge S_2(\mathcal E)$$

    for any unital channel \(\mathcal F\).

The measure of irreversibility \(S_2(\mathcal E)\) can be explicitly evaluated for various channels studied in Sec. 7. We list the results, together with those for \(S(\mathcal E)\), in Table 1.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Luo, S., Sun, Y. Quantifying the irreversibility of channels. Theor Math Phys 218, 426–451 (2024). https://doi.org/10.1134/S004057792403005X

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S004057792403005X

Keywords

Navigation