Skip to main content
Log in

Adiabatic Lindbladian Evolution with Small Dissipators

  • Original research
  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

We consider a time-dependent small quantum system weakly coupled to an environment, whose effective dynamics we address by means of a Lindblad equation. We assume the Hamiltonian part of the Lindbladian is slowly varying in time and the dissipator part has small amplitude. We study the properties of the evolved state of the small system as the adiabatic parameter and coupling constant both go to zero, in various asymptotic regimes. In particular, we analyse the deviations of the transition probabilities of the small system between the instantaneous eigenspaces of the Hamiltonian with respect to their values in the purely Hamiltonian adiabatic setup, as a function of both parameters.

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.

Similar content being viewed by others

References

  1. Abou Salem, W., Fröhlich, J.: Adiabatic theorems and reversible isothermal processes. Lett. Math. Phys. 72, 153–163 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  2. Albert, V.V., Bradlyn, B., Fraas, M., Jiang, L.: Geometry and response of Lindbladians. Phys. Rev. X 6, 041031 (2016)

    Google Scholar 

  3. Avron, J.E., Elgart, A.: Adiabatic theorem without a gap condition. Commun. Math. Phys. 203, 445–463 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Avron, J.E., Fraas, M., Graf, G.M., Grech, P.: Adiabatic theorems for generators of contracting evolutions. Commun. Math. Phys. 314, 163–191 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  5. Avron, J.E., Fraas, M., Graf, G.M., Grech, P.: Landau–Zener tunneling for dephasing lindblad evolutions. Commun. Math. Phys. 305(3), 633–639 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Avron, J.E., Seiler, R., Yaffe, L.G.: Adiabatic theorems and applications to the quantum Hall effect. Commun. Math. Phys. 110, 33–49 (1987)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Bachmann, S., De Roeck, W., Fraas, M.: The adiabatic theorem and linear response theory for extended quantum systems. Commun. Math. Phys. 361, 997–1027 (2018)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. Ballesteros, M., Crawford, N., Fraas, M., Fröhlich, J., Schubnel, B.: Perturbation theory for weak measurements in quantum mechanics, systems with finite-dimensional state space. Ann. H. Poincaré 20, 299–335 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  9. Benoist, T., Bernardin, C., Chétrite, R., Chhaibi, R., Najnudel, J., Pellegrini, C.: Emergence of jumps in quantum trajectories via homogeneization. Commun. Math. Phys. 387, 1821–1867 (2021)

    Article  ADS  MATH  Google Scholar 

  10. Benoist, T., Fraas, M., Jaksic, V., Pillet, C.-A.: Full statistics of erasure processes: Isothermal adiabatic theory and a statistical Landauer principle. Rev. Roumaine Math. Pures Appl. 62, 259–286 (2017)

    MathSciNet  MATH  Google Scholar 

  11. Born, M., Fock, V.: Beweis des Adiabatensatzes. Z. Phys. 51, 165–180 (1928)

    Article  ADS  MATH  Google Scholar 

  12. Carles, R., Fermanian-Kammerer, C.: A nonlinear adiabatic theorem for coherent states. Nonlinearity 24, 1–22 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  13. Carles, R., Fermanian-Kammerer, C.: A nonlinear Landau–Zener formula. J. Stat. Phys. 152, 619–656 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Davies, E.B.: Quantum Theory of Open Systems. Academic Press, London (1976)

    MATH  Google Scholar 

  15. Davies, E.B.: Linear operators and their spectra, Cambridge Studies in Advanced Mathematics 106, CUP (2007)

  16. Davies, E.B., Spohn, H.: Open quantum systems with time-dependent Hamiltonians and their linear response. J. Stat. Phys. 19, 511–523 (1978)

    Article  ADS  MathSciNet  Google Scholar 

  17. Dranov, A., Kellendonk, J., Seiler, R.: Discrete time adiabatic theorems for quantum mechanical systems. J. Math. Phys. 39, 1340–1349 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. Falconi, M., Faupin, J., Fröhlich, J., Schübnel, B.: Scattering Theory for Lindblad Master Equations. Commun. Math. Phys. 350, 1185–1218 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Fraas, M., Hänggli, L.: On Landau–Zener transitions for dephasing Lindbladians. Annales Henri Poincaré 18(7), 2447–2465 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Fermanian-Kammerer, C., Joye, A.: A nonlinear quantum adiabatic approximation. Nonlinearity 33, 4715–4751 (2020)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Gang, Z., Grech, P.: Adiabatic theorem for the Gross–Pitaevskii equation. Commun. PDE 42, 731–756 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  22. Haack, G., Joye, A.: Perturbation analysis of quantum reset models. J. Stat. Phys. 183, 17 (2021)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  23. Hänggli, L.: Aspects of system-environment evolutions, ETH-Zürich Doctoral thesis (2018). https://doi.org/10.3929/ethz-b-000299145

  24. Hanson, E.P., Joye, A., Pautrat, Y., Raquépas, R.: Landauer’s principle in repeated interaction systems. Commun. Math. Phys. 349, 285–327 (2017)

  25. Hanson, E.P., Joye, A., Pautrat, Y., Raquépas, R.: Landauer’s principle for trajectories of repeated interaction systems. Ann. H. Poincaré 19, 1939–1991 (2018)

  26. Joye, A.: Proof of the Landau–Zener formula. Asymp. Anal. 9, 209–258 (1994)

    MathSciNet  MATH  Google Scholar 

  27. Joye, A.: General adiabatic evolution with a gap condition. Commun. Math. Phys. 275, 139–162 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. Joye, A., Kunz, H., Pfister, C.-E.: Exponential decay and geometric aspect of transition probabilities in the adiabatic limit. Ann. Phys. 208, 299–332 (1991)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Joye, A., Merkli, M., Spehner, D.: Adiabatic transitions in a two-level system coupled to a free Boson reservoir. Ann. H. Poincaré 21, 3157–3199 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  30. Joye, A., Pfister, C.-E.: Exponentially small adiabatic invariant for the Schrödinger equation. Commun. Math. Phys. 140, 15–41 (1991)

    Article  ADS  MATH  Google Scholar 

  31. Joye, A., Pfister, C.-E.: Superadiabatic evolution and adiabatic transition probability between two non-degenerate levels isolated in the spectrum. J. Math. Phys. 34, 454–479 (1993)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Joye, A., Pfister, C.-E.: Quantum adiabatic evolution. In: Fannes, M., Meas, C., Verbeure, A. (eds.) Leuven Conference Proceedings; On the Three Levels Micro-, Meso- and Macro-Approaches in Physics, pp. 139–148. Plenum, New-York (1994)

    Google Scholar 

  33. Kato, T.: On the adiabatic theorem of quantum mechanics. J. Phys. Soc. Jpn. 5, 435–439 (1950)

    Article  ADS  Google Scholar 

  34. Kato, T.: Perturbation Theory for Linear Operators. Springer, Berlin, Heidelberg, New York (1980)

    MATH  Google Scholar 

  35. Krein, S.G.: Linear differential equations in Banach space. In: Translations of Mathematical Monographs, vol. 29. AMS (1971)

  36. Landau, L.: Zur Theorie der Energieübertragung. II. Phys. Z. Sowjet. 2, 46–51 (1932)

    MATH  Google Scholar 

  37. Lindblad, G.: On the generators of quantum dynamical semigroups. Commun. Math. Phys. 48, 119–130 (1976)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Liu, J., Li, S.-C., Fu, L.-B., Ye, D.-F.: Nonlinear Adiabatic Evolution of Quantum Systems. Springer, Singapore (2018)

    Book  Google Scholar 

  39. Macieszczak, K., Guta, M., Lesanovsky, I., Garrahan, J.P.: Towards a theory of metastability in open quantum dynamics. Phys. Rev. Lett. 116, 240404 (2016)

    Article  ADS  Google Scholar 

  40. Nenciu, G.: On the adiabatic theorem of quantum mechanics. J. Phys. A Math. Gen. 13, 15–18 (1980)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  41. Nenciu, G.: Linear adiabatic theory. Exponential estimates. Commun. Math. Phys. 152, 479–496 (1993)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  42. Norris, J.R.: Markov Chains. Cambridge University Press, Cambridge (1997)

    Book  MATH  Google Scholar 

  43. Nenciu, G., Rasche, G.: On the adiabatic theorem for nonself-adjoint Hamiltonians. J. Phys. A 25, 5741–5751 (1992)

    Article  ADS  MathSciNet  Google Scholar 

  44. Reed, M., Simon, B.: Methods of Modern Mathematical Physics. Academic Press, London (1972)

    MATH  Google Scholar 

  45. Schmid, J.: Adiabatic theorems with and without spectral gap condition for non-semisimple spectral values. In: Exner, P., König, W., Neidhardt, H. (eds.) Mathematical Results in Quantum Mechanics: Proceedings of the QMath12 Conference. World Scientific Publishing, Singapore (2014)

  46. Schrader, R.: Perron-Frobenius theory for positive maps on trace ideals. In: Mathematical Physics in Mathematics and Physics (Siena, 2000), 361–378, Fields Inst. Commun., 30, AMS Providence, RI (2001)

  47. Sparber, C.: Weakly nonlinear time–adiabatic theory. Ann. H. Poincaré 17, 913–936 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  48. Teufel, S.: A note on the adiabatic theorem without gap condition. Lett. Math. Phys. 58, 261–266 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  49. Teufel, S., Wachsmuth, J.: Spontaneous Decay of Resonant Energy Levels for Molecules with Moving Nuclei. Commun. Math. Phys. 315, 966–738 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  50. Yin, G., Zhang, Q.: Continuous-Time Markov Chains and Applications, Stochastic Modelling and Applied Probability, vol. 37, Springer

  51. Zener, C.: Non-adiabatic crossing of energy levels. Proc. R. Soc. Lond. Ser. A 137, 692–702 (1932)

    ADS  MATH  Google Scholar 

Download references

Acknowledgements

This work is partially supported by the ANR grant NONSTOPS (ANR-17-CE40-0006-01)

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alain Joye.

Additional information

Communicated by H.Spohn.

Publisher's Note

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

Appendix: Integration by Parts

Appendix: Integration by Parts

We present here a reformulation of the integration by parts argument used in [ASY] to prove the adiabatic theorem of quantum mechanics, suited to our setup.

Let \({{\mathcal {Z}}}\) be a Banach space and assume \({{\mathcal {G}}}:[0,1]\rightarrow {{\mathcal {B}}}({{\mathcal {Z}}})\), \({{\mathcal {K}}}:[0,1]\rightarrow {{\mathcal {B}}}({{\mathcal {Z}}})\) are bounded operator valued \(C^\infty \) functions on [0, 1], in the norm sense. Let \(\varepsilon >0\), and consider the two-parameter propagators \(({{\mathcal {X}}}(t,s))_{1\le s\le t\le 1}\) and \(({{\mathcal {Y}}}(t,s))_{1\le s\le t\le 1}\), solution to the equations

$$\begin{aligned} \left\{ \begin{array}{l} \varepsilon \partial _t{{\mathcal {X}}}(t,s)={{\mathcal {G}}}(t){{\mathcal {X}}}(t,s), \\ {{\mathcal {X}}}(s,s)={{\mathbb {I}}}, \ \ 0\le s \le t \le 1, \end{array} \right. \end{aligned}$$
(8.1)

and

$$\begin{aligned} \left\{ \begin{array}{l} \varepsilon \partial _t{{\mathcal {Y}}}(t,s)=({{\mathcal {G}}}(t)+\varepsilon {{\mathcal {K}}}(t)){{\mathcal {Y}}}(t,s), \\ {{\mathcal {Y}}}(s,s)={{\mathbb {I}}}, \ \ 0\le s \le t \le 1. \end{array} \right. \end{aligned}$$
(8.2)

The smooth propagators \({{\mathcal {X}}}(t,s)\) and \({{\mathcal {Y}}}(t,s)\) are determined by the corresponding Dyson series, both depend on \(\varepsilon >0\) with norms that diverge as \(\varepsilon \rightarrow 0\), a priori . Moreover, they satisfy the integral relation

$$\begin{aligned} {{\mathcal {X}}}(t,r)={{\mathcal {Y}}}(t,r)-\int _r^t {{\mathcal {Y}}}(t,s) {{\mathcal {K}}}(s) {{\mathcal {X}}}(s,r)ds, \ \ \forall 1\ge t\ge r \ge 0. \end{aligned}$$
(8.3)

Assume the existence of gaps in the spectrum of \({{\mathcal {G}}}(t)\), uniformly in \(t\in [0,1]\). For \(d\in {{\mathbb {N}}}^*\),

$$\begin{aligned} \sigma ({{\mathcal {G}}}(t))=\cup _{1\le j\le d}\, \sigma _j(t)\subset {{\mathbb {C}}}, \ \ \inf _{t\in [0,1], 1\le j\ne k\le d}\mathrm{dist} (\sigma _j(t),\sigma _k(t))\ge G >0. \end{aligned}$$
(8.4)

Consider the corresponding spectral projector

$$\begin{aligned} {{\mathcal {P}}}_j(t)=-\frac{1}{2 \pi \mathrm{i}}\int _{\gamma _j} ({{\mathcal {G}}}(s)-z)^{-1}dz, \end{aligned}$$
(8.5)

where \(\gamma _j\) is a simple loop in \(\rho ({{\mathcal {G}}}(t))\), the resolvent set of \({{\mathcal {G}}}(t)\), encircling \(\sigma _j(t)\) and such that for all \(k\ne j\), \(\text{ int }\gamma _j\cap \sigma _k(t)=\emptyset \). For \({{\mathcal {B}}}:[0,1]\rightarrow {{\mathcal {B}}}({{\mathcal {Z}}})\), a smooth bounded operator valued function, define for any \(t\in [0,1]\)

$$\begin{aligned} {{\mathcal {R}}}_j({{\mathcal {B}}})(t)=-\frac{1}{2\pi \mathrm{i}} \oint _{\gamma _j} ({{\mathcal {G}}}(t)-z)^{-1}{{\mathcal {B}}}(t)({{\mathcal {G}}}(t)-z)^{-1}dz, \end{aligned}$$
(8.6)

with the same loop \(\gamma _j\) as in (8.5). This operator is smooth as well, and satisfies the identity

$$\begin{aligned}{}[{{\mathcal {G}}}(t), {{\mathcal {R}}}_j({{\mathcal {B}}})(t)]=[{{\mathcal {B}}}(t),{{\mathcal {P}}}_j(t)]. \end{aligned}$$
(8.7)

Remark 8.1

If \(\sigma _k(t)=\{g_k(t)\}\) for all \(1\le k\le d\), \(({{\mathcal {G}}}(t)-z)^{-1}=\sum _{1\le k \le d} \frac{{{\mathcal {P}}}_k(t)}{(g_k(t)-z)}\), and

$$\begin{aligned} {{\mathcal {R}}}_j({{\mathcal {B}}})(t)=\sum _{1\le k\le d\atop k\ne j} \frac{{{\mathcal {P}}}_j(t){{\mathcal {B}}}(t){{\mathcal {P}}}_k(t)+{{\mathcal {P}}}_k(t){{\mathcal {B}}}(t){{\mathcal {P}}}_j(t)}{g_k(t)-g_j(t)}. \end{aligned}$$
(8.8)

Lemma 8.2

Suppose \({{\mathcal {K}}}(t)\) is off-diagonal for all \(t\in [0,1]\), i.e. s.t. \({{\mathcal {P}}}_j(t){{\mathcal {K}}}(t){{\mathcal {P}}}_j(t)\equiv 0\), \(\forall 1\le j\le d\). Then

$$\begin{aligned} {{\mathcal {X}}}(t,r)-{{\mathcal {Y}}}(t,r)&= \frac{1}{2}\sum _{1\le j\le d} \varepsilon \Big ({{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(t){{\mathcal {X}}}(t,r)- {{\mathcal {Y}}}(t,r) {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(r)\Big )\nonumber \\&\quad +\frac{1}{2}\sum _{1\le j\le d} \varepsilon \int _r^t \Big \{ {{\mathcal {Y}}}(t,s){{\mathcal {K}}}(s)[{{\mathcal {G}}}(s), {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)] {{\mathcal {X}}}(s,r)\nonumber \\&\quad -{{\mathcal {Y}}}(t,s) (\partial _s{{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)){{\mathcal {X}}}(s,r) \Big \} ds. \end{aligned}$$
(8.9)

Proof

The operator \({{\mathcal {K}}}\) being off-diagonal and (8.7) give

$$\begin{aligned} {{\mathcal {K}}}(t)=\frac{1}{2}\sum _{1\le j\le d}\big [[{{\mathcal {K}}}(t), {{\mathcal {P}}}_j(t)\big ],{{\mathcal {P}}}_j(t)]=\frac{1}{2}\sum _{1\le j\le d}\big [{{\mathcal {G}}}(t), {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(t)\big ].\qquad \end{aligned}$$
(8.10)

Hence, using (8.3), (8.1) and (8.2),

$$\begin{aligned} {{\mathcal {X}}}(t,r)-{{\mathcal {Y}}}(t,r) =&-\frac{1}{2}\sum _{1\le j\le d}\int _r^t {{\mathcal {Y}}}(t,s) \big [{{\mathcal {G}}}(s), {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)\big ] {{\mathcal {X}}}(s,r)ds \end{aligned}$$
(8.11)

where, for each integral in the summand

$$\begin{aligned} -\int _r^t {{\mathcal {Y}}}(t,s) [&{{\mathcal {G}}}(s), {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)] {{\mathcal {X}}}(s,r)ds\nonumber \\ =&\varepsilon \int _r^t \Big \{(\partial _s{{\mathcal {Y}}}(t,s)) {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s){{\mathcal {X}}}(s,r)\nonumber \\&\qquad \quad \quad + {{\mathcal {Y}}}(t,s){{\mathcal {K}}}(s)\big [{{\mathcal {G}}}(s), {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)\big ] {{\mathcal {X}}}(s,r) \nonumber \\&\qquad \qquad + {{\mathcal {Y}}}(t,s) {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)\partial _s{{\mathcal {X}}}(s,r) \Big \} ds. \end{aligned}$$
(8.12)

Thanks to the smoothness of all operators in the integrand, we have

$$\begin{aligned} (\partial _s{{\mathcal {Y}}}(t,s) )&{{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s){{\mathcal {X}}}(s,r)+{{\mathcal {Y}}}(t,s) {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)\partial _s{{\mathcal {X}}}(s,r)\nonumber \\&= \partial _s({{\mathcal {Y}}}(t,s) {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s){{\mathcal {X}}}(s,r)) -{{\mathcal {Y}}}(t,s) (\partial _s{{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)){{\mathcal {X}}}(s,r), \end{aligned}$$
(8.13)

which yields the sought for identity. \(\square \)

As a corollary of Lemma (8.2), if one of the propagators \(({{\mathcal {X}}}(t,s))_{0\le s\le t\le 1}\) or \(({{\mathcal {Y}}}(t, s))_{0\le s\le t\le 1}\) is uniformly bounded in \(\varepsilon \), so is the other, and their difference goes to zero with \(\varepsilon \):

Corollary 8.3

Assume \(\exists \ \varepsilon _1>0, C_1<\infty \) such that \(\sup _{0<\varepsilon \le \varepsilon _1\atop 0\le s\le t\le 1}\Vert {{\mathcal {X}}}(t,s)\Vert \le C_1\). Then, \(\exists \ \varepsilon _2>0, C_2<\infty \) such that \(\sup _{0<\varepsilon \le \varepsilon _2 \atop 0\le s\le t\le 1}\Vert {{\mathcal {Y}}}(t,s)\Vert \le C_2\). The same statement holds for \({{\mathcal {X}}}\) and \({{\mathcal {Y}}}\) exchanged. Moreover, \(\exists \ C_3<\infty \) such that for all \( \varepsilon <\varepsilon _2\),

$$\begin{aligned} \sup _{0\le s\le t\le 1}\Vert {{\mathcal {X}}}(t,s)-{{\mathcal {Y}}}(t,s)\Vert \le C_3 \varepsilon , \end{aligned}$$
(8.14)

and, whenever \({{\mathcal {K}}}(0)=0\), there exists \(C_4<\infty \) so that for all \(t\in [0,1]\)

$$\begin{aligned} \Vert {{\mathcal {X}}}(t,0)-{{\mathcal {Y}}}(t,0)\Vert \le C_4 t \varepsilon . \end{aligned}$$
(8.15)

Remark 8.4

If both \({{\mathcal {X}}}(t,s)\) and \({{\mathcal {Y}}}(t,s)\) are a priori uniformly bounded, estimate (8.14) holds for all \(\varepsilon \).

Proof

Set

$$\begin{aligned} 2C_0=\max&\Big ( \sum _{1\le j\le d}\sup _{0\le s\le 1}\Vert {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)\Vert , \sum _{1\le j\le d}\sup _{0\le s\le 1}\Vert \partial _s {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)\Vert , \nonumber \\&\sum _{1\le j\le d}\sup _{0\le s\le 1}\Vert {{\mathcal {K}}}(s)\big [{{\mathcal {G}}}(s), {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)\big ]\Vert \Big ) \end{aligned}$$
(8.16)

and consider \(\varepsilon <\varepsilon _1\). Lemma (8.2) yields the bound

$$\begin{aligned} \Vert {{\mathcal {Y}}}(t,r)\Vert&\le C_1+\varepsilon C_0 (\Vert {{\mathcal {Y}}}(t,r)\Vert +C_1)+\varepsilon 2C_0C_1\sup _{0\le s \le t\le 1} \Vert {{\mathcal {Y}}}(t,s)\Vert \end{aligned}$$
(8.17)

so that, taking the supremum over \(0\le r\le t\le 1\) and for \(\varepsilon < \min (\varepsilon _1, 1/(2C_0(1+2C_1))) := \varepsilon _2\), we get in turn

$$\begin{aligned} \sup _{0\le r \le t\le 1}\Vert {{\mathcal {Y}}}(t,r)\Vert&\le \frac{1}{(1-\varepsilon C_0(1+2C_1))}(C_1(1+\varepsilon C_0)) \le C_1\frac{3+4C_1}{1+2C_1}\equiv C_2. \end{aligned}$$
(8.18)

Then, inserting this estimate into (8.9), one gets, uniformly in \(0\le s\le t\le 1\),

$$\begin{aligned} \Vert {{\mathcal {X}}}(t,s)-{{\mathcal {Y}}}(t,s)\Vert \le C_3 \varepsilon , \end{aligned}$$
(8.19)

with \(C_3=C_0(C_1+C_2+2C_1C_2)\). Finally, for the initial time \(s=0\), the integrated contribution in (8.9) reduces to \(\frac{1}{2}\sum _{1\le j\le d} \varepsilon {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(t){{\mathcal {X}}}(t,0)\) when \({{\mathcal {K}}}(0)=0\), and since either \({{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(t)=0\) or

$$\begin{aligned} {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(t)=\int _0^t\partial _s {{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(s)ds, \end{aligned}$$
(8.20)

we have in any case

$$\begin{aligned} \Big \Vert \frac{1}{2}\sum _{1\le j\le d}{{\mathcal {R}}}_j([{{\mathcal {K}}}, {{\mathcal {P}}}_j])(t)\Big \Vert \le tC_0. \end{aligned}$$
(8.21)

The integral term in (8.9) is of order \(\varepsilon t\), so that the bound (8.15) holds with \(C_4=C_0C_1(1+2C_2)\).

The fact that \({{\mathcal {X}}}\) and \({{\mathcal {Y}}}\) can be exchanged in all arguments above follows from the structure of the RHS of (8.9). \(\square \)

Proof of Lemma 3.2:

We briefly prove estimate (3.7). Here the Banach space is \({{\mathcal {Z}}}={{\mathcal {H}}}\), the propagators are \({{\mathcal {X}}}=U\), \({{\mathcal {Y}}}=V\), and the generators are constructed with \({{\mathcal {G}}}=-iH\) and \({{\mathcal {K}}}=K=\sum _{1\le j\le d} P'_jP_j\).

Using \(P(t)P'(t)P(t)\equiv 0\) for any smooth projector P(t), one has the identities \(P_j(t)K(t) P_j(t)\equiv 0\), for all \(1\le j\le d\), and actually both U and V are bounded a priori, since they are unitary. Moreover, \(K(0)=0\), under Reg. Hence Lemma 3.2 derives from Corollary 8.3. \(\square \)

Proof of Lemma 3.12:

As a second application, we derive here estimate (3.37). We need to show that

$$\begin{aligned} P_n(t)\int _0^t{{\mathcal {V}}}^0(t,s)\circ {{\mathcal {L}}}^1_{s}\circ {{\mathcal {V}}}^0(s,0) (\rho _j)ds\, P_m(t)={{{\mathcal {O}}}}(\varepsilon ). \end{aligned}$$
(8.22)

We first note that by Lemma 3.7, \({{\mathcal {V}}}^0(s,0) (\rho _j)={\tilde{\rho }}_j(s)\), where \(\partial _s {\tilde{\rho }}_j(s)= [K(s), {\tilde{\rho }}_j(s)]\) is continuous in trace norm and \(\varepsilon \)-independent. Moreover,

$$\begin{aligned} P_n(t){{\mathcal {V}}}^0(t,s)(\cdot )P_m(t)=P_n(t)\big ({{\mathcal {V}}}^0(t,0)\circ {{\mathcal {V}}}^0(0,s)(P_n(s) \cdot P_m(s))\big )P_m(t), \end{aligned}$$
(8.23)

thanks to the intertwining property (7.19). Using the definition of \({{\mathcal {L}}}_s^1\) , we have

$$\begin{aligned}&P_n(s) {{\mathcal {L}}}^1_{s}\circ {{\mathcal {V}}}^0(s,0) (\rho _j) P_m(s)=\sum _{l\in I}P_n(s)(\Gamma _l(s) {\tilde{\rho }}_j(s)\Gamma _l^*(s)P_m(s) \nonumber \\&\quad -\delta _{mj}\frac{1}{2}P_n(s)\Gamma _l^*(s)\Gamma _l(s) {\tilde{\rho }}_j(s)P_m(s) - \frac{1}{2} \delta _{nj}P_n(s) {\tilde{\rho }}_j(s)\Gamma _l^*(s)\Gamma _l(s) P_m(s), \end{aligned}$$
(8.24)

where all terms are independent of \(\varepsilon \). Hence, to get the result, we are lead to show that for a smooth trace class operator \([0,1]\ni s\rightarrow F(s)\), such that \(\partial _sF(s)\in {{\mathcal {T}}}({{\mathcal {H}}})\), independent of \(\varepsilon \), and \(n\ne m\),

$$\begin{aligned} \int _0^tV^0(0,s) P_n(s)F(s)P_m(s)V^0(s,0)ds\, ={{{\mathcal {O}}}}(\varepsilon ). \end{aligned}$$
(8.25)

We have thanks to (8.7) with \({{\mathcal {G}}}=H\)

$$\begin{aligned} P_n(s)F(s)P_m(s)=[P_n(s)F(s)P_m(s), P_m(s)]=[H(s),{{\mathcal {R}}}_m(P_nFP_m)(s)]\qquad \end{aligned}$$
(8.26)

so that, by a slight variation of Lemma 8.2

$$\begin{aligned}&\int _0^tV^0 (0,s) [H(s),{{\mathcal {R}}}_m(P_nFP_m)(s)]V^0(s,0)ds\nonumber \\&\quad =-\mathrm{i}\varepsilon V^0(0,s) {{\mathcal {R}}}_m(P_nFP_m)(s)]V^0(s,0)|_0^t +\mathrm{i}\varepsilon \int _0^tV^0(0,s) \Big \{ \partial _s{{\mathcal {R}}}_m(P_nFP_m)(s)\nonumber \\&\qquad -K(s){{\mathcal {R}}}_m(P_nFP_m)(s)+{{\mathcal {R}}}_m(P_nFP_m)(s)K(s)\Big \}V^0(s,0)ds. \end{aligned}$$
(8.27)

As F(s) and its derivative are trace class, the expression above is \({{{\mathcal {O}}}}(\varepsilon )\) in trace norm. \(\square \)

Let us finally note that, making use of the projectors appearing (8.23), we can further integrate by parts the last integral term, provided \(\partial _sF(s)\) is continuously differentiable in trace norm, in which case

$$\begin{aligned}&\int _0^tV^0(0,s) P_n(s)F(s)P_m(s)V^0(s,0)ds\nonumber \\&\quad = \mathrm{i}\varepsilon \big ({{\mathcal {R}}}_m(P_nFP_m)(0) -V^0(0,t) {{\mathcal {R}}}_m(P_nFP_m)(t)V^0(t,0)\big )+{{{\mathcal {O}}}}(\varepsilon ^2). \end{aligned}$$
(8.28)

Proof of Lemma 3.13:

For any \(\rho \in {{\mathcal {T}}}({{\mathcal {H}}})\), we need to consider

$$\begin{aligned} \int _0^t\int _0^{s_1}\dots \int _0^{s_{n-1}} {{\mathcal {V}}}^0(t,s_1)\circ {{\mathcal {L}}}^1_{s_1} \circ {{\mathcal {V}}}^0(s_1,s_2) \circ {{\mathcal {L}}}^1_{s_2}\dots {{\mathcal {L}}}^1_{s_n} \circ {{\mathcal {V}}}^0(s_n,0) (\rho )ds_{n}\dots ds_{2}ds_{1}. \end{aligned}$$
(8.29)

Noting with (3.15) that for any \(0\le s,t\le 1\) \( {{\mathcal {V}}}^0(t,s)(\rho )={{\mathcal {V}}}^0(t,0)\circ {{\mathcal {V}}}^0(0,s)(\rho ), \) we can write (8.29) as \({{\mathcal {V}}}^0(t,0)(I_n(t))\), where \(I_n(t)\in {{\mathcal {T}}}({{\mathcal {H}}})\) is defined inductively by

$$\begin{aligned} I_n(t)&=\int _0^t {{\mathcal {V}}}^0(0,s_1)\circ {{\mathcal {L}}}^1_{s_1} \circ {{\mathcal {V}}}^0(s_1,0)(I_{n-1}(s_1))ds_1,\nonumber \\ I_1(t)&=\int _0^t {{\mathcal {V}}}^0(0,s_n) \circ {{\mathcal {L}}}^1_{s_n} \circ {{\mathcal {V}}}^0(s_n,0)(\rho )ds_n. \end{aligned}$$
(8.30)

Lemma 3.12 shows the existence of \(C_1<\infty \), such that for all \(0\le t\le 1\) and \(\varepsilon \) small enough,

$$\begin{aligned} \Vert I_1(t)-{{\mathcal {P}}}_0(0)(I_1(t))\Vert _1\le \varepsilon C_1 \Vert \rho \Vert _1. \end{aligned}$$
(8.31)

Let us show by induction that for for each n, there exists \(C_n<\infty \) so that for \(\varepsilon \) small enough

$$\begin{aligned} \Vert I_n(t)-{{\mathcal {P}}}_0(0)(I_n(t))\Vert _1\le \varepsilon C_n \Vert \rho \Vert _1. \end{aligned}$$
(8.32)

Assuming the result for \(n\ge 1\), we consider the step \(n+1\). We get

$$\begin{aligned}&I_{n+1}(t)-{{\mathcal {P}}}_0(0)(I_{n+1}(t))=\sum _{1\le j\ne k\le d}P_j(0)I_{n+1}(t)P_k(0)\nonumber \\&\quad = \sum _{1\le j\ne k\le d}\int _0^t P_j(0)\big \{{{\mathcal {V}}}^0(0,s) \circ {{\mathcal {L}}}^1_{s} \circ {{\mathcal {V}}}^0(s,0)(I_n(s))\big \}P_k(0)ds\nonumber \\&\quad = \sum _{1\le j\ne k\le d}\int _0^t\!\! P_j(0)\big \{{{\mathcal {V}}}^0(0,s) \circ {{\mathcal {L}}}^1_{s} \circ {{\mathcal {V}}}^0(s,0)\circ {{\mathcal {P}}}_0(0)(I_n(s)))\big \}P_k(0)ds +{{{\mathcal {O}}}}_n(\Vert \rho \Vert _1 \varepsilon ), \end{aligned}$$
(8.33)

by the induction hypothesis and recalling the operator \({{\mathcal {V}}}^0(t,s)\) is isometric and \({{\mathcal {L}}}_s^1\) is uniformly bounded. Then we observe that for any \(j\ne k\), and any \([0,1]\ni s\mapsto A(s)\in {{\mathcal {T}}}({{\mathcal {H}}})\), \(C^1\) in trace norm, see Lemmas 3.4, 3.5 and (3.23)

$$\begin{aligned}&\int _0^t P_j(0)\big \{{{\mathcal {V}}}^0(0,s) (A(s))\big \}P_k(0)ds\nonumber \\&\quad =\int _0^t e^{\frac{\mathrm{i}}{\varepsilon }\int _{0}^s(e_j-e_k)(u)du}P_j(0)\big \{{{\mathcal {W}}}^0(0,s) (A(s))\big \}P_k(0)ds\nonumber \\&\quad =\frac{-\mathrm{i}\varepsilon }{e_j(s)-e_k(s)}e^{\frac{\mathrm{i}}{\varepsilon }\int _{0}^s(e_j-e_k)(u)du}P_j(0)\big \{{{\mathcal {W}}}^0(0,s) (A(s))\big \}P_k(0)\Big |_0^t\nonumber \\&\qquad +\int _0^t\mathrm{i}\varepsilon e^{\frac{\mathrm{i}}{\varepsilon }\int _{0}^s(e_j-e_k)(u)du}P_j(0)\partial _s\Big (\frac{{{\mathcal {W}}}^0(0,s) (A(s))}{e_j(s)-e_k(s)}\Big )P_k(0)ds. \end{aligned}$$
(8.34)

The trace norm of the RHS is bounded above by \(\varepsilon c (\sup _{0\le s\le 1}\Vert A(s)\Vert _1+\sup _{0\le s\le 1}\Vert \partial _s A(s)\Vert _1)\), where c is a constant independent of \(\varepsilon \). The integral term of the RHS of (8.33) has the form (8.34) with

$$\begin{aligned} A(s)= {{\mathcal {L}}}^1_{s} \circ {{\mathcal {V}}}^0(s,0)\circ {{\mathcal {P}}}_0(0)(I_n(s))= {{\mathcal {L}}}^1_{s} \circ {{\mathcal {W}}}^0(s,0)\circ {{\mathcal {P}}}_0(0)(I_n(s)), \end{aligned}$$
(8.35)

where \({{\mathcal {W}}}^0(s,0)\) and \({{\mathcal {L}}}_s^1\) are smooth, independent of \(\varepsilon \) and bounded on \({{\mathcal {T}}}({{\mathcal {H}}})\), while \(I_n(s)\) and \(\partial _t I_n(s)\) are continuous and bounded in trace norm by a constant (uniform in \(\varepsilon \)) time \(\Vert \rho \Vert _1\), see (8.30), which ends the proof. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Joye, A. Adiabatic Lindbladian Evolution with Small Dissipators. Commun. Math. Phys. 391, 223–267 (2022). https://doi.org/10.1007/s00220-021-04306-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-021-04306-5

Navigation