Skip to main content

Advertisement

Log in

Typicality of Thermal Equilibrium and Thermalization in Isolated Macroscopic Quantum Systems

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

Abstract

Based on the view that thermal equilibrium should be characterized through macroscopic observations, we develop a general theory about typicality of thermal equilibrium and the approach to thermal equilibrium in macroscopic quantum systems. We first formulate the notion that a pure state in an isolated quantum system represents thermal equilibrium. Then by assuming, or proving in certain classes of nontrivial models (including that of two bodies in thermal contact), large-deviation type bounds (which we call thermodynamic bounds) for the microcanonical ensemble, we prove that to represent thermal equilibrium is a typical property for pure states in the microcanonical energy shell. We believe that the typicality, along with the empirical success of statistical mechanics, provides a sound justification of equilibrium statistical mechanics. We also establish the approach to thermal equilibrium under two different assumptions; one is that the initial state has a moderate energy distribution, and the other is the energy eigenstate thermalization hypothesis.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. Throughout the present paper, “pure state” implies a quantum mechanical pure state, which is a vector in the Hilbert space (or the many-body wave function). We treat mixed states in Appendix 3.

  2. To relate our theory to cold atom experiments may be an important future issue.

  3. The corresponding energy shell \(\mathcal {H}_{V,u}\) for for quantum case is defined in Sect. 2.1. See also Sect. 9.2 for a comparison between classical and quantum cases.

  4. We believe that the ergodic property of (classical) dynamical systems, although being quite interesting and deep, has little to do with the justification of equilibrium statistical mechanics.

  5. We must note, however, that one does not choose a state randomly in reality. See Sect. 4.2 for further discussion about the preparation of states.

  6. We note in passing that numerical simulations (based, e.g., on the molecular dynamics or the Markov process Monte Carlo methods) for equilibrium states of a macroscopic system work effectively because they are designed to generate states which belong to the majority (and hence represent thermal equilibrium). In this sense the Monte Carlo simulation in statistical mechanics is essentially different from Monte Carlo calculations of low-dimensional integrals.

  7. In a system (such as a quantum spin system) where the energy is bounded from above, there is a range of energy (corresponding to “negative temperatures”) where the density of states \(d\Omega _V(U)/dU\) is strictly decreasing in U. In such a region the entropy density \(\sigma (u)\) defined from the number of states (rather than the density of states) as in (2.1) becomes constant, and the assumption does not hold. But this is not an essential problem. The region with “negative temperatures” becomes a normal one by simply switching the sign of the Hamiltonian; we can apply our theory to this reversed Hamiltonian.

  8. In concrete models like quantum spin systems, we shall prove these properties.

  9. By \(A:=B\) (or, equivalently, \(B=:A\)) we mean that A is defined in terms of B.

  10. Small in the sense that \(\sigma (u)-\sigma (u- \Delta u)\ll \sigma (u)\).

  11. We have set the upper bound in (2.3) to u in order to make some formula simple. One can change it to, say, \(u+ \Delta u\).

  12. But we can also treat quantities which are not macroscopic. See Appendix 2.

  13. Although it is physically natural that one uses a fixed precision for the density, the critical reader might argue that the precision can be made higher for larger V. In fact our convention to use a fixed precision is also motivated by the (theoretical) fact that it leads us to an upper bound for the fluctuation, the thermodynamic bound (2.16), which is exponentially small in V.

  14. This projection, which acts on the whole Hilbert space \(\mathcal{H}_{V,\mathrm {tot}}\), should not be confused with the similar projection operator (also denoted as \(1-P_\mathrm {eq}\) or \(\hat{P}_\mathrm {neq}\)) which appears in [6, 2830]. The latter is the projection onto a subspace of the energy shell \(\mathcal {H}_{V,u}\). See Sect. 2.4.

  15. The value of \(\alpha \) must be set to be smaller than the constant \(\gamma \), which appears in the thermodynamic bound (2.16). The optimal value of \(\gamma \) is model specific, and there is no room for our choice.

  16. The quick reader can skip the treatment of multiple quantities.

  17. This characterization of thermal equilibrium is referred to as TMATE in [19].

  18. The problem whether n Hermitian matrices which “almost commute” can be approximated by n mutually commuting Hermitian matrices has a long history. The case \(n=2\) was solved affirmatively by Lin [32], while it is known [33] that the statement does not hold in general for \(n\ge 3\). In this sense Ogata’s result [31] for quantum spin systems is nontrivial and important.

  19. We are not arguing that one should literally measure the operators \(\tilde{M}_V^{(1)},\ldots ,\tilde{M}_V^{(n)}\). We have discussed this mathematical construction because it ensures that simultaneous approximate measurement of the n quantities is possible.

  20. The bound (2.16) may be called the “global large deviation upper bound for the microcanonical ensemble”.

  21. Throughout the present paper \(\left\| \cdot \right\| \) denotes the operator norm, i.e., \(\left\| \mathsf {O}\right\| :=\sup _{|\varphi \rangle \ne 0}\left\| \mathsf {O}|\varphi \rangle \right\| /\left\| |\varphi \rangle \right\| \).

  22. Consider, for example, the canonical distribution for the classical three dimensional ferromagnetic Ising model at very low temperature without external magnetic field. If the model is defined on a finite but large lattice, the canonical distribution represents the mixture of the two phases where the spins align upward or downward.

  23. This is similar to MITE formulated in [19]. But we have here fixed the decomposition \(\mathcal{H}_\mathrm{res}\otimes \mathcal{H}_\mathrm{sys}\), while various decompositions are considered in [19].

  24. It is indeed a deep problem to determine to what extent the prediction of equilibrium statistical mechanics should be regarded as realistic. See Sect. 9.3 and [19] for related issues.

  25. In fact we can prove (2.19) for some models by modifying our proof of the thermodynamic bound. See (8.23).

  26. An essential difference appears when we consider the time evolution. (Note that time evolution is not considered in [19].) In the prescription of [19], the energy shell \(\mathcal {H}_{V,u}\) is redefined according to the modified Hamiltonian \(\hat{\mathsf {H}}_V\). Since the time evolution must be determined by the original Hamiltonian \(\mathsf {H}_V\), the (redefined) energy shell \(\mathcal {H}_{V,u}\) is not invariant under the time evolution. Note that in our formalism neither the Hamiltonian nor the energy shell is redefined.

  27. It can be shown quite generally that there are plenty of u which satisfies the condition (3.1). Note that, in terms of the inverse function \(u(\beta )\), the condition (3.1) is equivalent to the continuity of \(u(\beta )\). But since \(u(\beta )\) is monotonically nonincreasing, there are at most countably many values of \(\beta \) at which \(u(\beta )\) is discontinuous.

  28. Physically speaking the condition should hold in general except at the triple point. The triple point must be excluded since energy is not distributed equally between the two subsystems at the point.

  29. In general we write \(a\simeq b\) when a and b are almost equal, and \(a\sim b\) when they are roughly equal or of the same order. For V dependent quantities, we write \(f(V)\simeq g(V)\) when \(\lim _{V\uparrow \infty }f(V)/g(V)=1\), and \(f(V)\sim g(V)\) when \(\lim _{V\uparrow \infty }V^{-1}\log [f(V)/g(V)]=0\).

  30. We believe that \(\sigma (u)\) is differentiable for all u, but cannot prove it in general.

  31. The probability is with respect to the random choice of \(|\varphi \rangle \).

  32. One can indeed get a stronger bound by using the Chebyshev estimate (see, e.g., [14, 15]) \(\mathrm{Prob}[|\mathsf {P}\!_\mathrm {neq}-\langle \mathsf {P}\!_\mathrm {neq}\rangle ^\mathrm {mc}_{V,u}|\ge s]\le n^2/(s^2\,D_{V,u})\).

  33. When the notion of pure states representing (thermal) equilibrium is defined through the expectation values as in (2.17), the typicality of (thermal) equilibrium follows only from the fact that the dimension of the relevant Hilbert space is large. See, e.g., [1, 1416].

  34. Take a spacial region A in the system, and consider the density matrix \(\rho _\mathrm{A}:=\mathrm{Tr}_{\mathcal{H}_{\bar{\mathrm{A}}}}[|\varphi \rangle \langle \varphi |]\), where the trace is taken over the subspace corresponding to the region out side A. Then the canonical typicality [1214] implies that the entropy \(S_\mathrm{A}=-\mathrm{Tr}[\rho _\mathrm{A}\log \rho _\mathrm{A}]\) is typically close to that of the canonical distribution, and is hence proportional to the volume of A.

  35. A method for preparing a nonequilibrium state (in a numerical or a cold-atom experiment) is to start from the ground state of a certain Hamiltonian, and then quickly change the Hamiltonian (see the next part). Since a ground state generally has small entanglement, the state cannot have too strong entanglement after a finite time.

  36. Physically speaking, the change of Hamiltonian is caused by an external agent, who must be in a nonequilibrium state to perform operations. On Earth, such nonequilibrium states are prepared by using energy from the sun. In the larger time scale, the origin of nonequilibrium goes back to the Big Bang.

  37. A quench from the ground state may be discussed in a similar manner.

  38. In terms of the Boltzmann entropy \(S=k_\mathrm {B}\log D\), the inequality (4.8) means \(S'(u')\ge S(u)+k_\mathrm {B} \Delta \sigma \,V\). It is quite normal that the entropy increases after a sudden change of Hamiltonian.

  39. We can extend our results about thermalization to the case where the system is initially in a mixed state. See Appendix 3.

  40. \(|\mathcal{G}|\) stands for the total length (i.e., the Lebesgue measure) of the intervals in \(\mathcal{G}\).

  41. Note that we do not have such freedom (or ambiguity) for geometric quantities such as the volume or mechanical quantity such as the energy. Entropy is a delicate quantity.

  42. For the Gibbs state \(\rho _{\beta }:=e^{-\beta \mathsf {H}_V}/\mathrm{Tr}[e^{-\beta \mathsf {H}_V}]\), the von Neumann entropy \(-\mathrm{Tr}[\rho _\beta \log \rho _\beta ]\) coincides with the (most macroscopic) thermodynamic entropy.

  43. Note that the standard thermodynamic entropy S(UVN) must be a function of controllable parameters. Since the value of \(\mathsf {M}_V\) cannot be controlled, S(UVNM) should be regarded as a nonequilibrium entropy. Such an entropy may be defined microscopically, for example, as \(S(U,V,N,M):=k_\mathrm {B}\log \mathrm{Tr}_{\mathcal {H}_{V,u}}[\mathsf {P}[|\mathsf {M}_V-M|\le V\delta ]]\) in the spirit of Boltzmann.

  44. An explicit calculation shows \(\overline{\sum _{j\in J_{V,u}}|c_j|^4}=2/(D_{V,u}+1)\), where the bar indicates the average over all normalized states as in Sect. 4. If we again choose \(|\varphi (0)\rangle \in \tilde{\mathcal {H}}_{V,u}\) randomly, we see that \(\mathrm {Prob}[\sum _j|c_j|^4\ge e^{\eta V}/D_{V,u}] =\overline{\chi [\sum _j|c_j|^4\ge e^{\eta V}/D_{V,u}]} \le \overline{\sum _j|c_j|^4\,e^{-\eta V}\,D_{V,u}} \le 2\,e^{-\eta V}\).

  45. It was stated first in the footnote of the unpublished work [48] and then as Theorem A.2 of [30].

  46. Given \(\gamma \) and \(\eta \) with \(\gamma >\eta \), one may choose \(\alpha \) and \(\nu \) such that \(\alpha +\nu <(\gamma -\eta )/2\).

  47. Note that we are here talking about the thermal equilibrium characterized by the particular \(\mathsf {P}\!_\mathrm {neq}\). Even in the system of two bodies not in contact, there can be nontrivial thermalization within each body.

  48. The energy eigenvalues are then degenerate since the two subsystems are identical. But this is not essential since the degeneracy may be lifted by making very small difference between the two Hamiltonians \(\mathsf {H}^{(1)}_{V/2}\) and \(\mathsf {H}^{(2)}_{V/2}\).

  49. It holds for the highly artificial toy model for heat conduction discussed in Sect. 1.3. See [3] for another artificial example.

  50. This statement is a trivial variation of Theorem 4.1. When the thermodynamic bound (2.16) is valid with \(\gamma >\kappa \), the bound \(\langle \varphi |\mathsf {P}\!_\mathrm {neq}|\varphi \rangle \le e^{-\kappa V}\) is valid with probability larger than \(1-e^{-(\gamma -\kappa )V}\).

  51. If this is the case, there is little (or no) question about our conclusion in Sect. 4.2 that typical properties correspond to thermal equilibrium.

  52. This simple strategy may not be the optimal way for the proof, especially when n is large.

  53. This is the proper rate function of the large deviation theory when (8.9) holds as an equality.

  54. The problem of two subsystems exchanging particles as well as energy can be treated in almost the same manner.

  55. The quantity \(\phi _\beta (\lambda )=\lim _{V\uparrow \infty }V^{-1}\log \omega _\beta (e^{\lambda \mathsf {B}_V})\), where \(\omega _\beta (\cdot )\) denotes the equilibrium state (or, more precisely, the KMS state) for the infinite lattice, is treated in [34, 37]. But by using the property called asymptotically decoupledness [38], which is satisfied in the present models, it can be shown that this defines the same quantity as (8.40). See [35, 36, 38].

  56. In [34, 35], the large deviation principle (for small enough \(\beta (u)\)) was proved when the operator \(\mathsf {m}_x\) (see (3.15)) acts only on a single site x. But by combining the derivations in [35] with the cluster expansion technique developed in Sect. 3.1 of [57], one can prove the desired results for a general local operator \(\mathsf {m}_x\) (Rey-Bellet, private communication).

  57. The function \(\phi _\beta (\lambda )\) of (8.10) is obtained by \(\phi _\beta (\lambda )=\beta \{f(\beta ,0)-f(\beta ,\lambda /\beta )\}-\lambda m\), where m is the equilibrium value of \(M_V/V\).

  58. It should be noted however that the observed time scales for thermalization differs considerably depending on the system.

  59. The idea to look for a typical property in the space of systems is due to von Neumann [1, 2].

  60. But see [62] for the discussion about the quick decay in physically realistic situations.

  61. It is known for \(p\in (1/2,1)\) that \(\sum _{N_+\ge pV}2^{-V}{V\atopwithdelims ()N_+}\le \exp [-V\{\log 2-S_2(p)\}]\).

  62. The thermodynamic bound (2.16) of course follows by summing up (10.5).

  63. We write \(\{\mathsf {A},\mathsf {B}\}=\mathsf {A}\mathsf {B}+\mathsf {B}\mathsf {A}\).

  64. One can introduce a different phase for the second term.

  65. Here the exceptional values of \(\theta \) depends on \(\lambda \).

  66. Here we do not make use of Theorem 6.2, but directly construct examples.

  67. Note that one inevitably has \(z_0(\varvec{n})\ne z_0(\varvec{n}')\) and \(z_\mathrm{coup}(\varvec{n})\ne z_\mathrm{coup}(\varvec{n}')\) when the equality holds.

  68. Proof: (10.44) implies \(s_n-s_{n-1}>s_{n+1}-s_n\). By repeatedly using this, one finds \(s_p-s_{p-1}>s_{q+1}-s_q\) for any \(p\le q\). This means \(s_p+s_q>s_{p-1}+s_{q+1}\), which justifies the claim.

  69. The numbering of \(\ell =1,2,\ldots ,\Omega _m\) can be done in an arbitrary manner.

  70. This, again, implies the thermodynamic bound (2.16).

  71. It is possible to let \(\Lambda _0\) depend on V, even to let \(\Lambda _0=\Lambda \). We do not go into such extensions, since a finite \(\Lambda _0\) is usually sufficient.

  72. This is a consequence of the equivalence of the microcanonical and the canonical ensembles. The proof in this case is elementary since the system is a union of noninteracting small parts.

  73. Proof: For an an arbitrary density matrix \(\rho \), one has \(\sum _j\langle \psi _j|\rho |\psi _j\rangle ^2\le \sum _{j,j'}\langle \psi _j|\rho |\psi _{j'}\rangle \langle \psi _{j'}|\rho |\psi _j\rangle =\sum _j\langle \psi _j|\rho ^2|\psi _j\rangle \).

  74. We do not use the assumption \(\lambda \gg 1\) in what follows. The main result is valid even when \(\lambda =0\). See the remark after the theorem.

References

  1. von Neumann, J.: Beweis des Ergodensatzes und des H-Theorems in der neuen Mechanik. Z. Phys. 57, 30 (1929); English translation (by R. Tumulka), Proof of the Ergodic Theorem and the H-Theorem in Quantum Mechanics. Eur. Phys. J. H 35, 201–237. arXiv:1003.2133

  2. Goldstein, S., Lebowitz, J.L., Tumulka, R., Zanghì, N.: Long-time behavior of macroscopic quantum systems: commentary accompanying the English translation of John von Neumann’s 1929 article on the quantum ergodic theorem. Eur. Phys. J. H 35, 173–200 (2010). arXiv:1003.2129

    Article  Google Scholar 

  3. Tasaki, H.: From quantum dynamics to the canonical distribution: general picture and a rigorous example. Phys. Rev. Lett. 80, 1373–1376 (1998). arXiv:cond-mat/9707253

  4. Reimann, P.: Foundation of statistical mechanics under experimentally realistic conditions. Phys. Rev. Lett. 101, 190403 (2008). arXiv:0810.3092

    Article  ADS  Google Scholar 

  5. Linden, N., Popescu, S., Short, A.J., Winter, A.: Quantum mechanical evolution towards thermal equilibrium. Phys. Rev. E 79, 061103 (2009). arXiv:0812.2385

    Article  ADS  MathSciNet  Google Scholar 

  6. Goldstein, S., Lebowitz, J.L., Mastrodonato, C., Tumulka, R., Zanghì, N.: On the approach to thermal equilibrium of macroscopic quantum systems. Phys. Rev. E 81, 011109 (2010). arXiv:0911.1724

    Article  ADS  Google Scholar 

  7. Tasaki, H.: The approach to thermal equilibrium and thermodynamic normality—an observation based on the works by Goldstein, Lebowitz, Mastrodonato, Tumulka, and Zanghiin 2009, and by von Neumann in 1929, unpublished note (2013). arXiv:1003.5424

  8. Reimann, P., Kastner, M.: Equilibration of isolated macroscopic quantum systems. New J. Phys. 14, 043020 (2012). http://iopscience.iop.org/1367-2630/14/4/043020

  9. Reimann, P.: Equilibration of isolated macroscopic quantum systems under experimentally realistic conditions. Phys. Scr. 86, 058512 (2012). arXiv:1210.5821

    Article  MATH  Google Scholar 

  10. Sato, J., Kanamoto, R., Kaminishi, E., Deguchi, T.: Exact relaxation dynamics of a localized many-body state in the 1D bose gas. Phys. Rev. Lett. 108, 110401 (2012). arXiv:1112.4244

    Article  ADS  Google Scholar 

  11. Reimann, P.: Generalization of von Neumann’s approach to thermalization. Phys. Rev. Lett. 115, 010403 (2015). arXiv:1507.00262

  12. Popescu, S., Short, A.J., Winter, A.: Entanglement and the foundation of statistical mechanics. Nat. Phys. 2(11), 754–758 (2006)

    Article  Google Scholar 

  13. Goldstein, S., Lebowitz, J.L., Tumulka, R., Zanghì, N. (2006). Canonical typicality. Phys. Rev. Lett. 96, 050403 (2006). arXiv:cond-mat/0511091

  14. Sugita, A.: On the foundation of quantum statistical mechanics (in Japanese), RIMS (Res. Inst. Math. Sc., Kyoto) Kokyuroku 1507, 147–159 (2006). http://www.kurims.kyoto-u.ac.jp/kyodo/kokyuroku/contents/pdf/1507-15+

  15. Sugita, A.: On the basis of quantum statistical mechanics. Nonlinear Phenom. Complex Syst. 10, 192–195 (2007). arXiv:cond-mat/0602625

  16. Reimann, P.: Typicality for generalized microcanonical ensembles. Phys. Rev. Lett 99, 160404 (2007). arXiv:0710.4214

    Article  ADS  Google Scholar 

  17. Sugiura, S., Shimizu, A.: Thermal pure quantum states at finite temperature. Phys. Rev. Lett 108, 240401 (2012). arXiv:1112.0740

    Article  ADS  Google Scholar 

  18. Sugiura, S., Shimizu, A.: Canonical thermal pure quantum state. Phys. Rev. Lett. 111, 010401 (2013). arXiv:1302.3138

  19. Goldstein, S., Huse, D.A, Lebowitz, J.L., Tumulka, R.: Thermal equilibrium of a macroscopic quantum system in a pure state. Phys. Rev. Lett. 115, 100402 (2015). arXiv:1506.07494

  20. Deutsch, J.M.: Quantum statistical mechanics in a closed system. Phys. Rev. A 43, 2046 (1991)

    Article  ADS  MathSciNet  Google Scholar 

  21. Srednicki, M.: Chaos and quantum thermalization. Phys. Rev. E 50, 888 (1994)

    Article  ADS  Google Scholar 

  22. Horoi, M., Zelevinsky, V., Alex, B.: Brown. Chaos vs thermalization in the nuclear shell model. Phys. Rev. Lett. 74, 5194 (1995)

    Article  ADS  Google Scholar 

  23. Zelevinsky, V., Brown, B.A., Frazier, N., Horoi, M.: The nuclear shell model as a testing ground for many-body quantum chaos. Phys. Rep. 276, 85–176 (1996)

  24. Lebowitz, J.L.: Boltzmann’s entropy and time’s arrow. Phys. Today 46(9), 32–38 (1993)

    Article  MathSciNet  Google Scholar 

  25. Lebowitz, J.L.: From time-symmetric microscopic dynamics to time-asymmetric macroscopic behavior: an overview, pp. 63–88. In: Gallavotti, G., Reiter, W.L., Yngvason, J. (eds.) Boltzmann’s Legacy. European Mathematical Society (2008). arXiv:0709.0724

  26. Tasaki, H.: Statistical Mechanics I (in Japanese, Baifukan 2008)

  27. Ruelle, D.: Statistical Mechanics: Rigorous Results. World Scientific (1999)

  28. Goldstein, S., Hara, T., Tasaki, H.: Time scales in the approach to equilibrium of macroscopic quantum systems, Phys. Rev. Lett. 111, 140401 (2013). arXiv:1307.0572

  29. Goldstein, S., Hara, H., Tasaki, T.: Extremely quick thermalization in a macro- scopic quantum system for a typical nonequilibrium subspace. New. J. Phys. 17, 045002 (2015). http://iopscience.iop.org/1367-2630/17/4/045002

  30. Goldstein, S., Hara, T., Tasaki, H.: The approach to equilibrium in a macroscopic quantum system for a typical nonequilibrium subspace, preprint (2014). arXiv:1402.3380

  31. Ogata, Y.: Approximating macroscopic observables in quantum spin systems with commuting matrices. J. Funct. Anal. 264, 2005–2033 (2013). arXiv:1111.5933

    Article  MathSciNet  MATH  Google Scholar 

  32. Lin, H.: Almost commuting selfadjoint matrices and applications. In: Fillmore, P.A., Mingo, J.A. (eds.) Operator Algebras and their Applications (Fields Institute Communications 13), pp. 193–233. American Mathematical Society (1997)

  33. Davidson, K.R.: Almost commuting Hermitian matrices. Math. Scand. 56, 222–240 (1985). http://www.mscand.dk/article/view/12098

  34. Netočnỳ, K., Redig, F.: Large deviations for quantum spin systems. J. Stat. Phys. 117, 521–547 (2004). arXiv:math-ph/0404018

  35. Lenci, M., Rey-Bellet, L.: Large deviations in quantum lattice systems: one-phase region. J. Stat. Phys. 119, 715–746 (2005). arXiv:math-ph/0406065

  36. Hiai, F., Mosonyi, M., Ogawa, T.: Large deviations and Chernoff bound for certain correlated states on a spin chain. J. Math. Phys. 48(123301), 1–19 (2007). arXiv:0706.2141

    MathSciNet  MATH  Google Scholar 

  37. Ogata, Y.: Large deviations in quantum spin chains. Commun. Math. Phys. 296, 35–68 (2010). arXiv:0803.0113

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Ogata, Y., Rey-Bellet, L.: Ruelle-Lanford functions and large deviations for asymptotically decoupled quantum systems. Rev. Math. Phys. 23, 211–232 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  39. Eisert, J., Cramer, M., Plenio, M.B.: Area laws for the entanglement entropy. Rev. Mod. Phys. 82, 277 (2010). arXiv:0808.3773

    Article  ADS  MathSciNet  MATH  Google Scholar 

  40. Goldstein, S., Lebowitz, J.L.: On the (Boltzmann) entropy of nonequilibrium systems. Phys. D 193, 53–66 (2004). arXiv:cond-mat/0304251

  41. Goldstein, S., Garrido, P., Lebowitz, J.L.: The Boltzmann entropy for dense fluids not in local equilibrium. Phys. Rev. Lett. 92, 050602 (2004). arXiv:cond-mat/0310575

  42. Pal, A., Huse, D.A.: The many-body localization transition. Phys. Rev. B 82, 174411 (2010). arXiv:1003.2613

    Article  ADS  Google Scholar 

  43. Imbrie, J.Z.: On Many-Body Localization for Quantum Spin Chains, preprint (2014). arXiv:1403.7837

  44. Rigol, M., Dunjko, V., Yurovsky, V., Olshanii, M.: Relaxation in a completely integrable many-body quantum system: an ab initio study of the dynamics of the highly excited states of 1D lattice hard-core bosons. Phys. Rev. Lett. 98, 050405 (2007). arXiv:cond-mat/0604476

  45. Ilievski, E., De Nardis, J., Wouters, B., Caux, J.-S., Essler, F.H.L., Prosen, T. (2015). Complete generalized Gibbs ensembles in interacting theories. Phys. Rev. Lett. 115, 157201 (2015). arXiv:1507.02993

  46. Rigol, M.: Breakdown of thermalization in finite one-dimensional systems. Phys. Rev. Lett. 103, 100403 (2009). arXiv:0904.3746

    Article  ADS  Google Scholar 

  47. Rigol, M.: Quantum quenches and thermalization in one-dimensional fermionic systems. Phys. Rev. A 80, 053607 (2009). arXiv:0908.3188

    Article  ADS  Google Scholar 

  48. Goldstein, S., Hara, T., Tasaki, H.: The second law of thermodynamics for pure quantum states (version 3), unpublished note (2013). arXiv:1303.6393v3

  49. Rigol, M., Srednicki, M.: Alternatives to eigenstate thermalization. Phys. Rev. Lett. 108, 110601 (2012). arXiv:1108.0928

    Article  ADS  Google Scholar 

  50. Rigol, M., Dunjko, V., Olshanii, M.: Thermalization and its mechanism for generic isolated quantum systems. Nature 452, 854–858 (2008). arXiv:0708.1324

  51. Santos, L.F., Rigol, M.: Onset of quantum chaos in one-dimensional bosonic and fermionic systems and its relation to thermalization. Phys. Rev. E 81, 036206 (2010). arXiv:0910.2985

    Article  ADS  MathSciNet  Google Scholar 

  52. Polkovnikov, A., Sengupta, K., Silva, A., Vengalattore, M.: Nonequilibrium dynamics of closed interacting quantum systems. Rev. Mod. Phys. 83, 863 (2011). arXiv:1007.5331

    Article  ADS  Google Scholar 

  53. Beugeling, W., Moessner, R., Haque, M.: Finite-size scaling of eigenstate thermalization. Phys. Rev. E 89, 042112 (2014). arXiv:1308.2862

    Article  ADS  Google Scholar 

  54. Ellis, R.: Entropy, Large Deviations, and Statistical Mechanics. Taylor & Francis (2005)

  55. Dembo, A., Zeitouni, O.: Large Deviations Techniques and Applications. Springer (2009)

  56. Araki, H.: Gibbs states of a one dimensional quantum lattice. Commun. Math. Phys. 14, 120–157 (1969). http://projecteuclid.org/euclid.cmp/1103841726

  57. Frölich, J., Ueltschi, D.: Some Properties of Correlations of Quantum Lattice Systems in Thermal Equilibrium, preprint (2014). arXiv:1412.2534

  58. Robinson, D.: A proof of the existence of phase transitions in the anisotropic Heisenberg model, Commun. Math. Phys. 14, 195–204 (1969). https://projecteuclid.org/euclid.cmp/1103841775

  59. Ginibre, J.: Existence of phase transitions for quantum lattice systems. Commun. Math. Phys. 14, 205–234 (1969). https://projecteuclid.org/euclid.cmp/1103841776

  60. Newman, C.M.: Gaussian correlation inequalities for ferromagnets. Z. Wahrscheilichkeitstheorie Verw. Gebiete 33, 75–93 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  61. Lebowitz, J.L.: GHS and other inequalities. Commun. Math. Phys. 35, 87–92 (1974). http://projecteuclid.org/euclid.cmp/1103859553

  62. Reimann, P.: Typical fast thermalization processes in closed many-body systems. Nat. Commun. 7, 10821 (2016). arXiv:1603.00669

  63. Lieb, E.H., Robinson, D.: The finite group velocity of quantum spin systems, Commun. Math. Phys. 28, 251–257 (1972). http://projecteuclid.org/euclid.cmp/1103858407

  64. Hastings, M.B., Koma, T.: Spectral gap and exponential decay of correlations. Commun. Math. Phys. 256, 781–804 (2006). arXiv:math-ph/0507008

  65. Vershynina, A., Lieb, E.H.: Lieb-Robinson bounds. Scholarpedia 8(9), 31267 (2013). http://www.scholarpedia.org/article/Lieb-Robinson_bounds

  66. Tignol, J.-P.: Galois’ Theory of Algebraic Equations. World Scientific (2001)

  67. Ireland, K., Rosen, M.: A Classical Introduction to Modern Number Theory. Graduate Texts in Mathematics. Springer, Berlin (1990)

    Book  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hal Tasaki.

Appendices

Appendix 1: Three Toy Models

We shall discuss three simple solvable examples in which the assumptions made in Sects. 6 and 7 can be easily verified. Although the results are in a sense trivial, we hope that these elementary examples shed light on the general scenario, and also provide insight to truly nontrivial many body systems. The material in this section is partly based on our unpublished work [7].

1.1 Independent Spins Under Random Magnetic Field

Let us start with a trivial example of independent spins under random magnetic field. In this model, independent precession of each spin causes the “approach to equilibrium” for certain observables. Although everything is trivial, it may be a good idea to look at the simple (but genuinely quantum) mechanism that realizes the relaxation-like behavior. Interestingly, the same model also offers a counterexample to the assumptions.

We use the same notation as in Sect. 3.3, but here regard \(\Lambda \) simply as a set of V sites. We consider a system of \(S=1/2\) spins on \(\Lambda \), and take the Hamiltonian

$$\begin{aligned} \mathsf {H}_V=\sum _{x\in \Lambda }^Nh_x\,\mathsf {S}^{(1)}_x, \end{aligned}$$
(10.1)

where independent spins are under nonuniform magnetic field. The local magnetic field \(h_x\) is independently drawn from the interval \([-h_0,h_0]\) according to the uniform probability measure, where \(h_0>0\) is a fixed constant.

As in (3.17), we denote by \(|\varphi ^\pm _x\rangle \) the basis states of the local Hilbert space \(\mathcal{H}_x\). Then the eigenstates of \(\mathsf {H}_V\) is written as

$$\begin{aligned} |\Psi _{\varvec{\tau }}\rangle =\bigotimes _{x\in \Lambda }\frac{1}{\sqrt{2}} \Bigl \{|\varphi ^+_x\rangle +\tau _x|\varphi ^-_x\rangle \Bigr \}, \end{aligned}$$
(10.2)

where we used the multi-index (or the spin configuration) \(\varvec{\tau }:=(\tau _x)_{x\in \Lambda }\) with \(\tau _x=\pm 1\). The corresponding energy eigenvalue is

$$\begin{aligned} E_{\varvec{\tau }}=\frac{1}{2}\sum _xh_x\,\tau _x. \end{aligned}$$
(10.3)

Since \(h_x\) are drawn randomly, the energy eigenvalues are nondegenerate with probability one.

As in Sect. 3.3, we take the total magnetization (in the direction orthogonal to the magnetic filed) \(\mathsf {M}_V=\sum _{x\in \Lambda }\mathsf {S}^{(3)}_x\) as the thermodynamic quantity of interest. From the symmetry, one has \(\langle \mathsf {M}_V\rangle ^\mathrm {mc}_{V,u}=0\) for any V and u.

Since each spin independently points upward or downward (in the 3-direction) with probability 1 / 2 in the energy eigenstate (10.2), the probability distribution of \(\mathsf {M}_V\) is given by the binomial distribution

$$\begin{aligned} \langle \Psi _{\varvec{\tau }}|\,\mathsf {P}[\mathsf {M}_V=M]\,|\Psi _{\varvec{\tau }}\rangle =\frac{1}{2^V}\frac{V!}{N_+!\,N_-!}. \end{aligned}$$
(10.4)

Here \(N_+\) and \(N_-\) are the numbers of up and down spins, respectively, which are determined by \(N_++N_-=V\) and \((N_+-N_-)/2=M\). By recalling the standard large deviation property of the coin tossFootnote 61 [54, 55], (10.4) implies

$$\begin{aligned} \langle \Psi _{\varvec{\tau }}|\,\mathsf {P}\bigl [|\mathsf {M}_V|\ge M\bigr ]\,|\Psi _{\varvec{\tau }}\rangle \le 2\,e^{-\kappa (\delta )\,V}, \end{aligned}$$
(10.5)

with \(\kappa (\delta )=\log 2-S_2[(1/2)+\delta ]=2\delta ^2+O(\delta ^4)\), where \(S_2(p)=-p\log p-(1-p)\log (1-p)\) is the binary entropy. We thus find that the energy eigenstate thermalization hypothesis (7.1) is valid in any range of energy.Footnote 62

We thus conclude that, as far as one looks at the total magnetization \(\mathsf {M}_V=\sum _{x\in \Lambda }\mathsf {S}^{(3)}_x\), the model approaches (thermal) equilibrium from any initial state. As we have noted in the beginning, this “approach to equilibrium” is nothing but a trivial consequence of independent precession of each spin.

It is interesting to see what happens if we take \(\mathsf {M}_V'=\sum _{x\in \Lambda }\mathsf {S}^{(1)}_x\), which is the total spin in the 1-direction, as the thermodynamic quantity of interest. In this case we have \(\mathsf {M}_V'|\Psi _{\varvec{\tau }}\rangle =(\sum _{x\in \Lambda }\tau _x/2)|\Psi _{\varvec{\tau }}\rangle \), i.e., the energy eigenstate (10.2) is also the eigenstate of \(\mathsf {M}_V'\). Since the energy eigenvalue (10.3) is the sum of the continuously distributed random quantities, it happens in general that two energy eigenvalues \(E_{\varvec{\tau }}\) and \(E_{\varvec{\tau }'}\) which are extremely close to each other have radically different configurations \(\varvec{\tau }\) and \(\varvec{\tau }'\). As a consequence, the eigenvalue \(\sum _{x\in \Lambda }\tau _x/2\) shows an erratic behavior when viewed as a function of the energy eigenvalue \(E_{\varvec{\tau }}\). See the discussion in [42]. This observation implies that, when one is interested in \(\mathsf {M}_V'\), the model does not satisfy the energy eigenstate thermalization hypothesis.

This trivial example illustrates how the tendency of the approach to equilibrium can be lost in a system with a quenched disorder. Pal and Huse [42] numerically studied the \(S=1/2\) model with the Hamiltonian

$$\begin{aligned} \mathsf {H}=J\sum _{j=1}^N{\varvec{\mathsf {S}}}_j\cdot {\varvec{\mathsf {S}}}_{j+1}+\sum _{j=1}^Nh_j\,\mathsf {S}^{(3)}_j, \end{aligned}$$
(10.6)

where J is a constant and \(h_j\) is uniformly distributed in \([-h,h]\). A systematic analysis suggests that the “localization” observed above for \(J=0\) persists in a model with sufficiently small |J|. In this case, the model lacks the ability to relax to equilibrium by itself. For large enough |J|, the system enters the delocalized phase where it can relax to equilibrium. See [42] and references therein for further discussions about the localization in many body quantum systems and its relation to the problem of the approach to equilibrium. See also [43] for a recent rigorous result.

1.2 Free Fermions on a Double Chain

We next discuss a slightly less trivial (but still easily solvable) model of free fermions on a double chain. We shall confirm that all the assumptions made in Theorems 6.1 and 7.1 are valid in this model.

Fig. 5
figure 5

The free fermion model defined on a pair of chains is a simple solvable model where we can prove all the assumptions in Theorems 6.1 and 7.1. Solid lines represent intra-chain hopping, and dotted lines represent (weak) inter-chain hopping (or coupling). The model is equivalent to a free fermion on a single chain with nearest and next-nearest neighbor hopping

The Model: We consider a free fermion model defined on a double chain as depicted in Fig. 5. The two chains are identified with sets of odd and even integers, respectively, as

$$\begin{aligned} \Lambda _1=\{1,3,\ldots ,2L-1\},\quad \Lambda _2=\{2,4,\ldots ,2L\}, \end{aligned}$$
(10.7)

where L is a fixed integer. We also denote the whole lattice as

$$\begin{aligned} \Lambda =\Lambda _1\cup \Lambda _2=\{1,2,3,\ldots ,2L\}, \end{aligned}$$
(10.8)

and identify the volume with \(V=2L\).

For each \(x \in \Lambda \), let \(\mathsf {c}_{x}\) and \(\mathsf {c}^\dagger _{x}\) be the annihilation and the creation operators, respectively, of a fermion at x. They satisfy the standard canonical anticommutation relationsFootnote 63

$$\begin{aligned} \{\mathsf {c}^\dagger _{x},\mathsf {c}_{y}\}=\delta _{x,y},\quad \{\mathsf {c}_{x},\mathsf {c}_{y}\}=\{\mathsf {c}^\dagger _{x},\mathsf {c}^\dagger _{y}\}=0, \end{aligned}$$
(10.9)

for any \(x,y\in \Lambda \). We consider states with N fermions on the lattice. (We fix \(\rho =N/V\) when we make V large.) The whole Hilbert space is spanned by the states of the form \(\mathsf {c}^\dagger _{x_1}\ldots \mathsf {c}^\dagger _{x_N}|\Phi _\mathrm {vac}\rangle \), where \(x_j\in \Lambda \) with \(x_j<x_{j+1}\), and \(|\Phi _\mathrm {vac}\rangle \) is the normalized state with no fermions in the system. It satisfies \(\mathsf {c}_x|\Phi _\mathrm {vac}\rangle =0\) for any x.

We consider the Hamiltonian

$$\begin{aligned} \mathsf {H}_V=\frac{1}{2}\sum _{x\in \Lambda }(e^{i\theta }\mathsf {c}^\dagger _{x}\mathsf {c}_{x+2}+e^{-i\theta }\mathsf {c}^\dagger _{x+2}\mathsf {c}_{x}) +\frac{\lambda }{2}\sum _{x\in \Lambda }(e^{i\theta }\mathsf {c}^\dagger _{x}\mathsf {c}_{x+1}+e^{-i\theta }\mathsf {c}^\dagger _{x+1}\mathsf {c}_{x}), \end{aligned}$$
(10.10)

where \(\lambda \in (0,1]\) and \(\theta \in [0,2\pi )\) are parameters. The phases \(\theta \) is introduced (rather artificially) to avoid degeneracy.Footnote 64 See Proposition 10.1. We impose periodic boundary conditions, and make identifications \(\mathsf {c}_{2L+1}=\mathsf {c}_1\) and \(\mathsf {c}_{2L+2}=\mathsf {c}_1\).

The first term in (10.10) represents hopping within each chain, while the second term represents hopping between the two chains. The model is also interpreted as that on a single chain \(\Lambda \) with nearest neighbor and next-nearest neighbor hopping.

Energy Eigenstates and Eigenvalues: Define the set of wave numbers as

$$\begin{aligned} \mathcal{K}:=\left\{ \frac{2\pi }{2L}j\,\Bigl |\,j=1,2,\ldots ,2L\right\} . \end{aligned}$$
(10.11)

We introduce fermion operators \(\mathsf {a}^\dagger _k\) for \(k\in \mathcal{K}\), which are related with \(\mathsf {c}^\dagger _x\) by

$$\begin{aligned} \mathsf {a}^\dagger _k=\frac{1}{\sqrt{2L}}\sum _{x\in \Lambda }e^{ikx}\mathsf {c}^\dagger _x,\quad \mathsf {c}^\dagger _x=\frac{1}{\sqrt{2L}}\sum _{k\in \mathcal{K}}e^{-ikx}\mathsf {a}^\dagger _k, \end{aligned}$$
(10.12)

and satisfy the anticommutation relations

$$\begin{aligned} \{\mathsf {a}^\dagger _k,\mathsf {a}_{k'}\}=\delta _{k,k'},\quad \{\mathsf {a}^\dagger _k,\mathsf {a}^\dagger _{k'}\}=\{\mathsf {a}_k,\mathsf {a}_{k'}\}=0, \end{aligned}$$
(10.13)

for any \(k,k'\in \mathcal{K}\).

A standard calculation shows that the Hamiltonian (10.10) is diagonalized by using the \(\mathsf {a}\) operators as

$$\begin{aligned} \mathsf {H}_V=\sum _{k\in \mathcal{K}}\bigl \{\epsilon _0(k)+\lambda \epsilon _\mathrm{coup}(k)\bigr \}\mathsf {a}^\dagger _k\mathsf {a}_k, \end{aligned}$$
(10.14)

where

$$\begin{aligned} \epsilon _0(k)=\cos (2k+\theta ),\quad \epsilon _\mathrm{coup}(k)=\cos (k+\theta ). \end{aligned}$$
(10.15)

Take an arbitrary subset \(K\subset \mathcal{K}\) such that \(|K|=N\), and define

$$\begin{aligned} |\Psi _K\rangle :=\Bigl (\prod _{k\in K}\mathsf {a}^\dagger _{k}\Bigr )|\Phi _\mathrm {vac}\rangle . \end{aligned}$$
(10.16)

From (10.14) one finds that \(|\Psi _K\rangle \) is an eigenstate of \(\mathsf {H}_V\), i.e.,

$$\begin{aligned} \mathsf {H}_V|\Psi _K\rangle =E_K\,|\Psi _K\rangle , \end{aligned}$$
(10.17)

with the energy eigenvalue

$$\begin{aligned} E_K=\sum _{k\in K}\bigl \{\epsilon _0(k)+\lambda \epsilon _\mathrm{coup}(k)\bigr \}. \end{aligned}$$
(10.18)

It can be also shown that the corresponding number of states exhibits the standard behavior (2.1) when U / V is sufficiently small. The energy shell \(\mathcal {H}_{V,u}\) is spanned by \(|\Psi _K\rangle \) with

$$\begin{aligned} u- \Delta u<\frac{E_K}{V}\le u. \end{aligned}$$
(10.19)

In Sects. 6 and 7, we have assumed that the energy eigenvalues are nondegenerate. Although nondegeneracy is always achieved by adding a small (random) perturbation to any given Hamiltonian, it is nice to know that nondegeneracy is guaranteed under certain conditions. By making use of standard results in number theory [66, 67], we can prove that the present model generically has no degeneracy. See the end of the section for the proof.

Proposition 10.1

Let \(L>2\) be a prime number with \(N<L/2\). Fix an arbitrary constant \(\lambda _0>0\). For any \(\lambda \in (0,\lambda _0]\) except for (at most) a finite number of points, and for any \(\theta \in [0,2\pi )\) except for a finite number of pointsFootnote 65, the energy eigenvalues are nondegenerate, i.e., \(E_K=E_{K'}\) implies \(K=K'\).

From a physical point of view, it is absurd to assume that the chain length is equal to a prime number. We of course do not believe that this is really crucial. When some of the conditions of the proposition are not satisfied, we still expect the system to exhibit essentially the same behavior although there may be some accidental (and irrelevant) degeneracies in the energy eigenvalues.

Other Fermion Operators: Let \(\nu =1,2\) specify one of the two chains. We denote by \(\mathsf {N}_\nu :=\sum _{x\in \Lambda _\nu }\mathsf {c}^\dagger _{x}\mathsf {c}_{x}\) the number of fermions on the chain \(\nu \). We also define

$$\begin{aligned} \mathsf {a}^\dagger _{k,\nu }:=\frac{1}{\sqrt{L}}\sum _{x\in \Lambda _\nu }e^{ikx}\mathsf {c}^\dagger _x, \end{aligned}$$
(10.20)

which creates the state with wave number k only on the chain \(\nu \). We obviously have

$$\begin{aligned} \mathsf {a}^\dagger _k=\frac{1}{\sqrt{2}}(\mathsf {a}^\dagger _{k,1}+\mathsf {a}^\dagger _{k,2}), \end{aligned}$$
(10.21)

and

$$\begin{aligned}{}[\mathsf {N}_\nu ,\mathsf {a}^\dagger _{k,\nu '}]=\delta _{\nu ,\nu '}\mathsf {a}^\dagger _{k,\nu }. \end{aligned}$$
(10.22)

For any \(k\in \mathcal{K}\), we denote by \(\bar{k}\) the unique element in \(\mathcal{K}\) such that \(|k-\bar{k}|=\pi \). One readily finds \(\mathsf {a}^\dagger _{\bar{k},1}=-\mathsf {a}^\dagger _{k,1}\) and \(\mathsf {a}^\dagger _{\bar{k},2}=\mathsf {a}^\dagger _{k,2}\). Recalling (10.21), this implies

$$\begin{aligned} \mathsf {a}^\dagger _k\mathsf {a}^\dagger _{\bar{k}}=\frac{1}{2}(\mathsf {a}^\dagger _{k,1}+\mathsf {a}^\dagger _{k,2})(\mathsf {a}^\dagger _{\bar{k},1}+\mathsf {a}^\dagger _{\bar{k},2}) =\frac{1}{2}(\mathsf {a}^\dagger _{k,1}\mathsf {a}^\dagger _{\bar{k},2}+\mathsf {a}^\dagger _{k,2}\mathsf {a}^\dagger _{\bar{k},1}). \end{aligned}$$
(10.23)

From (10.20), we also get the anticommutation relations

$$\begin{aligned} \{\mathsf {a}^\dagger _{k,1},\mathsf {a}_{k',2}\}=0, \quad \{\mathsf {a}^\dagger _{k,1},\mathsf {a}_{k',1}\}= {\left\{ \begin{array}{ll} 1&{}k=k',\\ -1&{}\bar{k}=k',\\ 0&{}\text {otherwise},\\ \end{array}\right. } \quad \{\mathsf {a}^\dagger _{k,2},\mathsf {a}_{k',2}\}= {\left\{ \begin{array}{ll} 1&{}k=k',\\ 1&{}\bar{k}=k',\\ 0&{}\text {otherwise},\\ \end{array}\right. }\nonumber \\ \end{aligned}$$
(10.24)

for any \(k,k'\in \mathcal{K}\).

Energy Eigenstate Thermalization: As for the thermodynamic quantity of interest let us take

$$\begin{aligned} \mathsf {M}_V:=\mathsf {N}_1-\mathsf {N}_2, \end{aligned}$$
(10.25)

which is the difference of the particle numbers in the two chains. The equilibrium value of \(\mathsf {M}_V\) is obviously zero by the symmetry.

Let us examine the validity of the energy eigenstate thermalization with respect to \(\mathsf {M}_V\). For \(K\subset \mathcal{K}\) with \(|K|=N\), define

$$\begin{aligned} K_0:=\left\{ k\in K\,\Bigl |\,k\le \pi ,\bar{k}=k+\pi \in K\right\} . \end{aligned}$$
(10.26)

We can then write the energy eigenstate (10.16) as

$$\begin{aligned} |\Psi _K\rangle&=\pm \Bigl (\prod _{k\in K_0}\mathsf {a}^\dagger _{k}\mathsf {a}^\dagger _{\bar{k}}\Bigr ) \Bigl (\prod _{k\in K\backslash K_0}\mathsf {a}^\dagger _{k}\Bigr )|\Phi _\mathrm {vac}\rangle \nonumber \\&= \pm \left( \prod _{k\in K_0}\frac{\mathsf {a}^\dagger _{k,1}\mathsf {a}^\dagger _{\bar{k},2}+\mathsf {a}^\dagger _{k,2}\mathsf {a}^\dagger _{\bar{k},1}}{2}\right) \left( \prod _{k\in K\backslash K_0}\frac{\mathsf {a}^\dagger _{k,1}+\mathsf {a}^\dagger _{k,2}}{\sqrt{2}}\right) |\Phi _\mathrm {vac}\rangle , \end{aligned}$$
(10.27)

where we used (10.23) and (10.21).

Let \(N_0=2|K_0|\). Noting the commutation relation (10.22), we find that, in the state (10.27), \(N_0\) fermions are evenly distributed to the two chains, and each of the remaining \(N-N_0\) fermions belongs to one of the two chains with independent probability 1 / 2. The probability distribution for the numbers of fermions in the two chains is then

$$\begin{aligned}&p(N_1,N_2):=\langle \Psi _K|\,\mathsf {P}[\mathsf {N}_1=N_1,\mathsf {N}_2=N_2]\,|\Psi _K\rangle \nonumber \\&\quad ={\left\{ \begin{array}{ll} \dfrac{1}{2^{N-N_0}}\,\dfrac{(N-N_0)!}{\{N_1-\frac{N_0}{2}\}!\,\{N_2-\frac{N_0}{2}\}!}&{} \text {when}\, N_1+N_2=N, N_1\ge \dfrac{N_0}{2},\, \mathrm{and}\, N_2\ge \dfrac{N_0}{2}\\ 0&{}\text {otherwise}. \end{array}\right. } \end{aligned}$$
(10.28)

Given the binomial distribution (10.28) we again see from the standard result in large deviation (see footnote 61) that

$$\begin{aligned}&\langle \Psi _K|\,\mathsf {P}\bigl [\mathsf {M}_V\ge V\delta \bigr ]\,|\Psi _K\rangle =\mathop {\sum _{N_1,N_2}}_{(N_1-N_2\ge V\delta )}p(N_1,N_2) \nonumber \\&{\left\{ \begin{array}{ll} \le \exp \biggl [ -V\,(\rho -\rho _0)\biggl \{\log 2-S_2\Bigl (\dfrac{1}{2}+\dfrac{\delta }{2(\rho -\rho _0)}\Bigr )\biggr \}\biggr ] &{}\text {for}\, \delta \in (0,\rho -\rho _0]\\ =0&{}\text {for}\, \delta \in (\rho -\rho _0,\rho ] \end{array}\right. } \end{aligned}$$
(10.29)

where \(\rho _0=N_0/V\). Noting the symmetry \(\mathsf {M}_V\rightarrow -\mathsf {M}_V\) and the bound

$$\begin{aligned} (\rho -\rho _0)\biggl \{\log 2-S_2\Bigl (\dfrac{1}{2}+\dfrac{\delta }{2(\rho -\rho _0)}\Bigr )\biggr \} \ge \rho \,\biggl \{\log 2-S_2\Bigl (\dfrac{1}{2}+\dfrac{\delta }{2\rho }\Bigr )\biggr \}, \end{aligned}$$
(10.30)

we find for any \(\delta \in (0,\rho ]\) that

$$\begin{aligned} \langle \Psi _K|\,\mathsf {P}\bigl [|\mathsf {M}_V|\ge V\delta \bigr ]\,|\Psi _K\rangle \le 2 e^{-\kappa (\delta ) V}, \end{aligned}$$
(10.31)

where \(\kappa (\delta )\) is defined as the right-hand side of (10.30). So the energy eigenstate thermalization hypothesis (7.1) is valid for any energy eigenstate \(|\Psi _K\rangle \).

Note that (10.31) readily implies the thermodynamic bound

$$\begin{aligned} \Bigl \langle \mathsf {P}\bigl [|\mathsf {M}_V|\ge V\delta \bigr ]\Bigr \rangle ^\mathrm {mc}_{V,u} \le 2e^{-\kappa (\delta )V}, \end{aligned}$$
(10.32)

for any u.

Thermalization: Suppose that the conditions for the nondegeneracy in Proposition 10.1 are satisfied. Since the energy eigenstate thermalization (10.31) is valid, we see that the conditions for Theorem 7.1 are satisfied. Therefore we have a concrete example in which the approach to thermal equilibrium from an arbitrary initial state \(|\varphi (0)\rangle \in \mathcal {H}_{V,u}\) can be proved without any unjustified assumptions.

Nonequilibrium Initial States with a Moderate Energy Distribution: Note that Theorem 6.1, which states thermalization for certain initial states, does not provide any additional information when the stronger Theorem 7.1 is known to be valid. It is nevertheless useful to see explicitly that nonequilibrium initial states with a moderate energy distribution (as required in Theorem 6.1) are possible in this model.Footnote 66

Fix an energy interval \((u- \Delta u,u]\). To avoid technical complexity, we assume that the coupling \(\lambda \) satisfies \(0<\lambda \ll \Delta u\), and can be neglected when we estimate the energy. We further simplify the discussion by focusing only on the extreme nonequilibrium where \(\mathsf {M}_V=\mathsf {N}_1-\mathsf {N}_2\) takes its maximum value N.

Take a subset \(\tilde{K}\subset \mathcal{K}\) such that \(|\tilde{K}|=N\) and \(k\le \pi \) for any \(k\in \tilde{K}\). We also assume that the total energy satisfies

$$\begin{aligned} u- \Delta u<\frac{1}{V}\sum _{k\in \tilde{K}}\epsilon _0(k)\le u, \end{aligned}$$
(10.33)

where we ignored the terms including \(\lambda \) in (10.18). We denote by \(D_{V,u}^\mathrm {neq}\) the total number of \(\tilde{K}\) satisfying all these conditions.

We define

$$\begin{aligned} |\Gamma _{\tilde{K}}\rangle :=\Bigl (\prod _{k\in \tilde{K}}\mathsf {a}^\dagger _{k,1}\Bigr )|\Phi _\mathrm {vac}\rangle , \end{aligned}$$
(10.34)

which is in the energy shell \(\mathcal {H}_{V,u}\) (under the assumption \(\lambda \ll \Delta u\)), and is an extreme nonequilibrium state with \(\mathsf {M}_V|\Gamma _{\tilde{K}}\rangle =N|\Gamma _{\tilde{K}}\rangle \). We thus find that \(D_{V,u}^\mathrm {neq}\) is the dimension of the nonequilibrium subspace in this case. Note that, corresponding to the thermodynamic bound (2.16), we have \(D_{V,u}^\mathrm {neq}\sim e^{-\gamma V}D_{V,u}\) (which essentially is the definition of \(\gamma >0\)).

Since \(\mathsf {a}^\dagger _{k,1}=(\mathsf {a}^\dagger _k-\mathsf {a}^\dagger _{\bar{k}})/\sqrt{2}\), we can rewrite (10.34) as

$$\begin{aligned} |\Gamma _{\tilde{K}}\rangle =\Biggl (\prod _{k\in \tilde{K}}\frac{\mathsf {a}^\dagger _k-\mathsf {a}^\dagger _{\bar{k}}}{\sqrt{2}}\Biggr )|\Phi _\mathrm {vac}\rangle . \end{aligned}$$
(10.35)

Expanding the product, we see that this state is a linear combination of \(2^N\) energy eigenstates (10.16). This means that \(|\Gamma _{\tilde{K}}\rangle \) has the effective dimension \(D_\mathrm {eff}=2^N\).

The rest is easy. Consider an initial state

$$\begin{aligned} |\Phi (0)\rangle =\sum _{\tilde{K}}\alpha _{\tilde{K}}|\Gamma _{\tilde{K}}\rangle , \end{aligned}$$
(10.36)

where \(\tilde{K}\) is summed over the \(D_{V,u}^\mathrm {neq}\) subsets \(\tilde{K}\) satisfying the above conditions, and all \(|\alpha _{\tilde{K}}|\) are nearly equal. Clearly \(|\Phi (0)\rangle \) is in \(\mathcal {H}_{V,u}\), is an extreme nonequilibrium state, and has the effective dimension

$$\begin{aligned} D_\mathrm {eff}\sim 2^{N}D_{V,u}^\mathrm {neq}\sim 2^{N}e^{-\gamma V}D_{V,u}= e^{-(\gamma -\rho \log 2)V}D_{V,u}. \end{aligned}$$
(10.37)

By comparing this with (6.2), we can choose \(\eta =\gamma -\rho \log 2\). The condition \(\gamma >\eta \) required in Theorem 6.1 is thus satisfied.

Proof of Proposition 10.1: To prove the absence of degeneracy, it is convenient to introduce the standard occupation number description. With each \(K\subset \mathcal{K}\), we associate a 2L-tuple \(\varvec{n}=(n_j)_{j=1,2,\ldots ,2L}\) by

$$\begin{aligned} n_j={\left\{ \begin{array}{ll} 1&{}\text {if}\, \frac{2\pi }{2L}j\in K,\\ 0&{}\text {if}\, \frac{2\pi }{2L}j\not \in K.\\ \end{array}\right. } \end{aligned}$$
(10.38)

Then the energy eigenvalue (10.18) is written as

$$\begin{aligned} E_K=\sum _{k\in K}\bigl \{\cos (2k+\theta )+\lambda \cos (k+\theta )\bigr \} =\mathfrak {R}\Bigl [e^{i\theta }\bigl \{z_0(\varvec{n})+\lambda \,z_\mathrm{coup}(\varvec{n})\bigr \}\Bigr ], \end{aligned}$$
(10.39)

with

$$\begin{aligned} z_0(\varvec{n}):=\sum _{j=1}^{2L}n_j\,\exp \left[ i\frac{2\pi }{L}j\right] ,\quad z_\mathrm{coup}(\varvec{n}):=\sum _{j=1}^{2L}n_j\,\exp \left[ i\frac{2\pi }{2L}j\right] . \end{aligned}$$
(10.40)

In the following lemma, \(\varvec{n}=(n_j)_{j=1,2,\ldots ,2L}\) and \(\varvec{n}'=(n'_j)_{j=1,2,\ldots ,2L}\) denote general 2L-tuples whose elements are 0 and 1. We also write \(|\varvec{n}|=\sum _{j=1}^{2L}n_j\).

Lemma 10.1

Let \(L>2\) be prime. Take any \(\varvec{n}\) and \(\varvec{n}'\) such that \(|\varvec{n}|=|\varvec{n}'|<L/2\). Then one has \(z_0(\varvec{n})=z_0(\varvec{n}')\) and \(z_\mathrm{coup}(\varvec{n})=z_\mathrm{coup}(\varvec{n}')\) simultaneously if and only if \(\varvec{n}=\varvec{n}'\).

Proof of Proposition 10.1 given Lemma 10.1

The lemma implies that, when \(\varvec{n}\ne \varvec{n}'\), the equality \(z_0(\varvec{n})+\lambda \,z_\mathrm{coup}(\varvec{n})=z_0(\varvec{n}')+\lambda \,z_\mathrm{coup}(\varvec{n}')\) may hold only accidentally,Footnote 67 and becomes invalid by an infinitesimal change of \(\lambda \). We thus find that, for any \(\lambda \in (0,1]\) except for a finite number of points, the complex quantities \(z_0(\varvec{n})+\lambda \,z_\mathrm{coup}(\varvec{n})\) (with all possible \(\varvec{n}\) such that \(|\varvec{n}|=N\)) are all distinct.

Suppose that \(\lambda \) is fixed to a non-exceptional value. The energy eigenvalue \(E_K\) may still degenerate if \(e^{i\theta }\{z_0(\varvec{n})+\lambda \,z_\mathrm{coup}(\varvec{n})\}\) and \(e^{i\theta }\{z_0(\varvec{n}')+\lambda \,z_\mathrm{coup}(\varvec{n}')\}\) (with \(\varvec{n}\ne \varvec{n}'\)) happen to have the same real part. Such a degeneracy is lifted by infinitesimally changing \(\theta \). Thus degeneracy in \(E_K\) can take place only for a finite number of values of \(\theta \).

We shall state a mathematical lemma which is the essence of Lemma 10.1. Let \(\zeta :=e^{i(2\pi /L)}\). Take an L-tuple \(\varvec{m}=(m_j)_{j=1,\ldots ,L}\) with \(m_j\in \mathbb {Z}\). Let \(|\varvec{m}|=\sum _{j=1}^L|m_j|\) and \(\tilde{z}(\varvec{m}):=\sum _{j=1}^Lm_j\zeta ^j\). It is crucial to note that here j runs from 1 to L.

Lemma 10.2

One has \(\tilde{z}(\varvec{m})\ne 0\) for any \(\varvec{m}\) such that \(0<|\varvec{m}|<L\). One also has \(\tilde{z}(\varvec{m})\ne \tilde{z}(\varvec{m}')\) for any \(\varvec{m},\varvec{m}'\) such that \(\varvec{m}\ne \varvec{m}'\) and \(|\varvec{m}|+|\varvec{m}'|<L\).

Proof

This lemma is a straightforward consequence of the classical result by Gauss known as “the irreducibility of the cyclotomic polynomials of prime index” (see, for example, Chapter 12, Sect. 3 of [66] or Chapter 13, Sect. 2 of [67]). It implies that the \(L-1\) complex numbers \(\zeta \), \(\zeta ^2\), \(\ldots \), \(\zeta ^{L-1}\) are rationally independent, i.e., if \(\sum _{n=1}^{L-1}m_n\,\zeta ^n=0\) with integers \(m_1,\ldots ,m_{L-1}\), one inevitably has \(m_1=m_2=\cdots =m_{L-1}=0\).

To show the first claim, we note that \(|\varvec{m}|<L\) implies that there is \(j_1\in \{1,2,\ldots ,L\}\) such that \(m_{j_1}=0\). Then one finds that \(\zeta ^{-j_1}\,\tilde{z}(\varvec{m})=\sum _{n=1}^{L-1}\tilde{m}_n\,\zeta ^n\) with \(\tilde{m}_n\in \mathbb {Z}\). Since not all of \(\tilde{m}_n\) are vanishing, we have \(\sum _{n=1}^{L-1}\tilde{m}_n\,\zeta ^n\ne 0\), and hence \(\tilde{z}(\varvec{m})\ne 0\).

To show the second claim, we observe that \(\tilde{z}(\varvec{m})-\tilde{z}(\varvec{m}')=\tilde{z}(\varvec{m}'')\) with \(m''_j=m_j-m'_j\). Since \(0<|\varvec{m}''|\le |\varvec{m}|+|\varvec{m}'|<L\), the first claim shows that \(\tilde{z}(\varvec{m})-\tilde{z}(\varvec{m}')\ne 0\). \(\square \)

Proof of Lemma 10.1 given Lemma 10.2

We shall rewrite \(z_0(\varvec{n})\) and \(z_\mathrm{coup}(\varvec{n})\) in the form of \(\tilde{z}(\varvec{m})\). As for \(z_0(\varvec{n})\), we readily find

$$\begin{aligned} z_0(\varvec{n})=\sum _{j=1}^{2L}n_j\,\exp \left[ i\frac{2\pi }{L}j\right] =\sum _{j=1}^L(n_j+n_{j+L})\zeta ^j. \end{aligned}$$
(10.41)

To deal with \(z_\mathrm{coup}(\varvec{n})\) we note that \(e^{i(\pi /L)j}=\zeta ^{j/2}\) for even j, and \(e^{i(\pi /L)j}=-\zeta ^{(j\pm L)/2}\) for odd j. Then we get

$$\begin{aligned} z_\mathrm{coup}(\varvec{n})=\sum _{j=1}^{2L}n_j\,\exp \Bigl [i\frac{\pi }{L}j\Bigr ] =\mathop {\sum _{j=1}^L}_{(j\,\text {even})}(n_j-n_{j+L})\zeta ^{j/2} -\mathop {\sum _{j=1}^L}_{(j\,\text {odd})}(n_j-n_{j+L})\zeta ^{(j+L)/2}.\nonumber \\ \end{aligned}$$
(10.42)

Take \(\varvec{n}\) and \(\varvec{n}'\) such that \(|\varvec{n}|=|\varvec{n}'|<L/2\). Then, from the expression (10.41) and Lemma 10.2, we see that \(z_0(\varvec{n})=z_0(\varvec{n}')\) if and only if \(n_j+n_{j+L}=n'_j+n'_{j+L}\) for all \(j=1,2,\ldots ,L\). Similarly from (10.42) and Lemma 10.2, we see that \(z_\mathrm{coup}(\varvec{n})=z_\mathrm{coup}(\varvec{n}')\) if and only if \(n_j-n_{j+L}=n'_j-n'_{j+L}\) for all \(j=1,2,\ldots ,L\). Therefore \(z_0(\varvec{n})=z_0(\varvec{n}')\) and \(z_\mathrm{coup}(\varvec{n})=z_\mathrm{coup}(\varvec{n}')\) are simultaneously valid if and only if \(\varvec{n}=\varvec{n}'\). \(\square \)

1.3 Toy Model for Two Identical Bodies in Contact

Finally we shall see an artificial model of two identical bodies exchanging energy, the situation treated in Sect. 3.1. We should warn the reader that, unlike the two simple models that we have discussed in Sects. 1.1 and 1.2, the present example is “made up” so that our scenario of typicality and thermalization works perfectly. Nevertheless we hope that this concrete example will be of help in developing intuitions about nontrivial models.

Let us specify the first subsystem. We assume that the eigenvalues of the Hamiltonian \(\mathsf {H}^{(1)}\) is \(\epsilon _0V n\) with \(n=1,2,\ldots \), where \(\epsilon _0>0\) is a fixed constant. To mimic the behavior of a macroscopic system, we require that each level with n is \(\Omega ^{(1)}_n\) fold degenerate, where

$$\begin{aligned} \Omega ^{(1)}_n=\exp [Vs_n]. \end{aligned}$$
(10.43)

We assume that the “entropy density” \(s_n\) is strictly increasing in n, and strictly concave, i.e.,

$$\begin{aligned} 2s_n>s_{n-1}+s_{n+1}, \end{aligned}$$
(10.44)

for any \(n=2,3,\ldots \). We denote the eigenstate of \(\mathsf {H}^{(1)}\) as \(|\psi ^{(1)}_{n,j}\rangle \) where \(n=1,2,\ldots \), and \(j=1,2,\ldots ,\Omega ^{(1)}_n\). It satisfies \(\mathsf {H}^{(1)}|\psi ^{(1)}_{n,j}\rangle =\epsilon _0V n|\psi ^{(1)}_{n,j}\rangle \).

The second subsystem is an exact copy of the first, and we denote by \(|\psi ^{(2)}_{n',j'}\rangle \) the energy eigenstate of the Hamiltonian \(\mathsf {H}^{(2)}\).

For \(m=2,3,\ldots \), let \(\mathcal{N}_m:=\{(n,n')\,|\,n,n'\in \{1,2,\ldots \},\ n+n'=m\}\). Then for any \((n,n')\in \mathcal{N}_m\), the tensor product \(|\psi ^{(1)}_{n,j}\rangle \otimes |\psi ^{(2)}_{n',j'}\rangle \) is an eigenstate of the noninteracting Hamiltonian \(\mathsf {H}^{(1)}\otimes \mathsf {1}+\mathsf {1}\otimes \mathsf {H}^{(2)}\) with the eigenvalue \(\epsilon _0V m\). The degeneracy of this eigenvalue is given by

$$\begin{aligned} \Omega _m=\sum _{(n,n')\in \mathcal{N}_m}\Omega ^{(1)}_n\,\Omega ^{(2)}_{n'} =\sum _{(n,n')\in \mathcal{N}_m}\exp [V(s_n+s_{n'})]. \end{aligned}$$
(10.45)

We also denote by \(\mathcal{H}_m\) the corresponding \(\Omega _m\) dimensional eigenspace.

Suppose that m is even. Take \((n,n')\in \mathcal{N}_m\), and write \(n=(m/2)+r\) and \(n'=(m/2)-r\). From concavity (10.44), it follows that the quantity \(s_n+s_{n'}=s_{(m/2)+r}+s_{(m/2)-r}\) attains its maximum at \(r=0\) and decreases strictlyFootnote 68 as r deviates from 0. We thus find that

$$\begin{aligned} \frac{\Omega ^{(1)}_{(m/2)+r}\,\Omega ^{(2)}_{(m/2)-r}}{\Omega _{m}} \le \frac{\Omega ^{(1)}_{(m/2)+r}\,\Omega ^{(2)}_{(m/2)-r}}{\Omega ^{(1)}_{m/2}\,\Omega ^{(2)}_{m/2}} =e^{-\tilde{\kappa }(m,r)\,V}, \end{aligned}$$
(10.46)

for any \(|r|\ge 1\), where

$$\begin{aligned} \tilde{\kappa }(m,r):=2s_{m/2}-\{s_{(m/2)+r}+s_{(m/2)-r}\}>0, \end{aligned}$$
(10.47)

is strictly increasing in |r| (for a fixed m). This bound will be useful below.

We shall design the interaction Hamiltonian \(\mathsf {H}_\mathrm {int}\) so that to leave each subspace \(\mathcal{H}_m\) invariant, and mix up all the basis states in it. To make the model trivially solvable, we shall go through the following highly artificial construction. For each \(m=2,3,\ldots \), list up all the basis states \(|\psi ^{(1)}_{n,j}\rangle \otimes |\psi ^{(2)}_{n',j'}\rangle \) with \((n,n')\in \mathcal{N}_m\), and renumber themFootnote 69 as \(|\Phi _{m,\ell }\rangle \), where \(\ell =1,2,\ldots ,\Omega _m\). We then define \(\mathsf {H}_\mathrm {int}\) by

$$\begin{aligned} \langle \Phi _{m,\ell }|\mathsf {H}_\mathrm {int}|\Phi _{m',\ell '}\rangle = {\left\{ \begin{array}{ll} (V\epsilon _1/2)\,e^{i\theta }&{}\text {if}\, m=m' \,\mathrm{and}\, \ell '=\ell +1\\ (V\epsilon _1/2)\,e^{-i\theta }&{}\text {if}\, m=m' \,\mathrm{and}\, \ell '=\ell -1\\ 0&{}\text {otherwise}, \end{array}\right. } \end{aligned}$$
(10.48)

where we take the “periodic boundary condition” and identify \(\ell =\Omega _m+1\) with \(\ell =1\). With this artificial choice, the interaction Hamiltonian \(\mathsf {H}_\mathrm {int}\), restricted on \(\mathcal{H}_m\), is exactly the Hamiltonian of the tight-binding model on a chain of length \(\Omega _m\). The phase \(\theta \) is introduced to avoid degeneracy.

Then the energy eigenstate of the full Hamiltonian \(\mathsf {H}^{(1)}\otimes \mathsf {1}+\mathsf {1}\otimes \mathsf {H}^{(2)}+\mathsf {H}_\mathrm {int}\) is readily obtained as

$$\begin{aligned} |\Psi _{m,q}\rangle =\frac{1}{\sqrt{\Omega _m}} \sum _{\ell =1}^{\Omega _m}\exp \Bigl [i\frac{2\pi q\ell }{\Omega _m}\Bigr ]|\Phi _{m,\ell }\rangle , \end{aligned}$$
(10.49)

where \(q=1,\ldots ,\Omega _m\) for each \(m=2,3,\ldots \). The corresponding energy eigenvalue is

$$\begin{aligned} E_{m,q}=V\biggl \{\epsilon _0m+\epsilon _1\cos \Bigl [\frac{2\pi q}{\Omega _m}+\theta \Bigr ]\biggr \}. \end{aligned}$$
(10.50)

By taking \(\epsilon _0>\epsilon _1>0\) and assuming that \(\theta /\pi \) is irrational, we find that the energy eigenvalues are nondegenerate.

Take, for simplicity, an even m, and choose u and \( \Delta u\) so that \(u- \Delta u=\epsilon _0m-\epsilon _1\) and \(u=\epsilon _0m+\epsilon _1\). Then the energy shell \(\mathcal {H}_{V,u}\) coincides with the subspace \(\mathcal{H}_m\). We shall again look at the energy difference \(\mathsf {M}_V=\mathsf {H}^{(1)}\otimes \mathsf {1}-\mathsf {1}\otimes \mathsf {H}^{(2)}\). By construction each \(|\Phi _{m,\ell }\rangle \), which is indeed \(|\psi ^{(1)}_{n,j}\rangle \otimes |\psi ^{(2)}_{n',j'}\rangle \), is an eigenstate of \(\mathsf {M}_V\), i.e., \(\mathsf {M}_V|\psi ^{(1)}_{n,j}\rangle \otimes |\psi ^{(2)}_{n',j'}\rangle =V\epsilon _0(n-n')|\psi ^{(1)}_{n,j}\rangle \otimes |\psi ^{(2)}_{n',j'}\rangle \). Since the energy eigenstate \(|\Psi _{m,q}\rangle \) is a linear combination of all \(|\psi ^{(1)}_{n,j}\rangle \otimes |\psi ^{(2)}_{n',j'}\rangle \) with \((n,n')\in \mathcal{N}_m\) as in (10.49), we readily see that

$$\begin{aligned} \langle \Psi _{m,q}|\,\mathsf {P}[\mathsf {M}_V=V\epsilon _0s]\,|\Psi _{m,q}\rangle = \frac{\Omega ^{(1)}_{(m+s)/2}\,\Omega ^{(2)}_{(m-s)/2}}{\Omega _{m}} \le e^{-\tilde{\kappa }(m,s/2)\,V}, \end{aligned}$$
(10.51)

where the inequality follows from (10.46). Then we immediately see that the energy eigenstate thermalization hypothesis (7.1) holds asFootnote 70

$$\begin{aligned} \langle \Psi _{m,q}|\,\mathsf {P}[|\mathsf {M}_V|\ge V\delta ]\,|\Psi _{m,q}\rangle \le e^{-\kappa (m,\delta )\,V}, \end{aligned}$$
(10.52)

where \(\kappa (m,\delta )\simeq \tilde{\kappa }(m,\delta /(2\epsilon _0))\).

We can thus conclude from Theorem 7.1 that the present model exhibits thermalization (in the sense that the energy is almost evenly distributed into the two subsystems) from any initial state.

It is also instructive to see that one can take (initial) states with moderate energy distribution in this model. From (10.49), one finds that any \(|\Phi _{m,\ell }\rangle \) can be written in terms of the energy eigenstates as

$$\begin{aligned} |\Phi _{m,\ell }\rangle =\frac{1}{\sqrt{\Omega _m}} \sum _{q=1}^{\Omega _m}\exp \Bigl [-i\frac{2\pi q\ell }{\Omega _m}\Bigr ]|\Psi _{m,q}\rangle . \end{aligned}$$
(10.53)

Since all the expansion coefficients have the same amplitude, the effective dimension of any \(|\Phi _{m,\ell }\rangle \), which is indeed \(|\psi ^{(1)}_{n,j}\rangle \otimes |\psi ^{(2)}_{n',j'}\rangle \), takes the maximum possible value \(D_\mathrm {eff}=D_{V,u}\). This means that there are many nonequilibrium states which satisfies the condition (6.2) with \(\eta =0\). We also see that the condition for Theorem 6.2 is satisfied in this case, and hence an overwhelming majority of nonequilibrium states satisfy (6.2).

Appendix 2: Treatment of Non-extensive Quantities

In the main body of the present paper we have treated only macroscopic (extensive) quantities \(\mathsf {M}_V^{(1)},\ldots ,\mathsf {M}_V^{(n)}\) of a macroscopic system. Although this is sufficient for our motivation to reproduce equilibrium thermodynamics, one can, if necessary, treat quantities which are not extensive by a slight modification.

Correlation Functions: We concentrate on a model defined on the d-dimensional \(L\times \cdots \times L\) hypercubic lattice \(\Lambda \) whose sites are denoted as \(x,y,\ldots \in \Lambda \), and write \(V=L^d\). We assume that the Hamiltonian \(\mathsf {H}_V\) is translationally invariant.

Let \(\mathsf {f}_o\) and \(\mathsf {g}_o\) be operators which act only on a finite number of sites, and denote by \(\mathsf {f}_x\) and \(\mathsf {g}_x\) their translations. Suppose that one is interested in the correlation function

$$\begin{aligned} c_r(u):=\lim _{V\uparrow \infty }\Bigl \langle \mathsf {f}_x\,\mathsf {g}_{x+r}+(\mathsf {f}_x\,\mathsf {g}_{x+r})^\dagger \Bigr \rangle ^\mathrm {mc}_{V,u}, \end{aligned}$$
(11.1)

for \(r\in \mathbb {Z}^d\). We shall argue that the value \(c_r(u)\) (for limited r) can be treated as the equilibrium value of a macroscopic quantity.

For a fixed r, we define

$$\begin{aligned} \mathsf {C}_{V,r}:=\sum _{x\in \Lambda }\bigl \{\mathsf {f}_x\,\mathsf {g}_{x+r}+(\mathsf {f}_x\,\mathsf {g}_{x+r})^\dagger \bigr \}, \end{aligned}$$
(11.2)

which is regarded as an extensive quantity. Because of the translation invariance, we have

$$\begin{aligned} c_r(u)=\lim _{V\uparrow \infty }\frac{1}{V}\langle \mathsf {C}_{V,r}\rangle ^\mathrm {mc}_{V,u}. \end{aligned}$$
(11.3)

We then take a sufficiently large subset \(\Lambda _0\subset \mathbb {Z}^d\) (which is independent of V), and include all \(\mathsf {C}_{V,r}\) with \(r\in \Lambda _0\) into the list of macroscopic quantities to consider.Footnote 71 We expect that the thermodynamic bound (2.16) is still valid in general after including \(\mathsf {C}_{V,r}\). In fact Proposition 3.2 for quantum spin systems extends to this case as it is (with a worse constant).

In this manner we can treat, within our scheme, the correlation function \(c_r(u)\) as the equilibrium value of the macroscopic quantity \(\mathsf {C}_{V,r}\). All the results about typicality (Sect. 4) and thermalization (Sects. 5, 6, and 7) can be applied as they are when the assumptions are verified. In most cases it is enough to take \(\Lambda _0\) sufficiently large (to exceed the correlation length) in order to recover essential physics described by the correlation function.

Probability Distribution in a Small System: Consider a quantum mechanical system defined on a region (which can be a lattice) with a small volume \(V_0\). Physical quantities in this system should exhibit relatively large fluctuation in thermal equilibrium. With a little trick (see, e.g., [26]), one can also treat the probability distribution of such a fluctuating quantity within our scheme.

Let a self-adjoint operator \(\mathsf {f}\) be the quantity of interest of the small system. We prepare N identical copies of the small system, and consider a combined system of all the copies. There are no interactions between small subsystems. Denoting by \(\mathsf {f}^{(j)}\) the quantity \(\mathsf {f}\) in the j-th copy, we define, for \(a<b\), the operator

$$\begin{aligned} \mathsf {N}_{[a,b]}:=\sum _{j=1}^N\mathsf {P}\bigl [\mathsf {f}^{(j)}\in [a,b]\bigr ], \end{aligned}$$
(11.4)

which counts the number of copies in which the value of \(\mathsf {f}\) falls in to the interval [ab].

Let \(\langle \cdots \rangle ^\mathrm{mc}_{N,u}\) be the microcanonical expectation corresponding to the energy range \([(u- \Delta u)N,uN]\) for the whole system. Then it is easily found thatFootnote 72

$$\begin{aligned} p_{\beta (u)}(a,b):=\lim _{N\uparrow \infty }\frac{1}{N}\langle \mathsf {N}_{[a,b]}\rangle ^\mathrm{mc}_{N,u} \end{aligned}$$
(11.5)

is the probability that the value of \(\mathsf {f}\) falls into [ab] in a single small system described by the canonical distribution with \(\beta (u)\).

Let us divide the spectrum of \(\mathsf {f}\) in to n intervals as \([f_\mathrm{min},f_\mathrm{max}]=\bigcup _{j=1}^n[a_{j-1},a_j]\). We then regard \(\mathsf {N}_{[a_{j-1},a_j]}\) with \(j=1,\ldots ,n\) as our macroscopic quantities. It is again not hard to prove that the thermodynamic bound (2.16) is valid for any u.

From the results in Sect. 4, we can thus essentially recover the probability distribution \(p_\beta (a,b)\) from a single pure state of a large system (constructed by combining copies of the original small system). Unfortunately the large system, as it is, never exhibit thermalization since small parts do not interact with each other. It is likely that the system thermalizes if we add weak interactions between the small systems, but this is not easy to prove.

Appendix 3: Mixed Initial State

Our results about thermalization in Sects. 5, 6, and 7 readily extends to the case where the initial state is a mixed state. In what follows we only consider density matrices whose supports are the energy shell \(\mathcal {H}_{V,u}\). Corresponding to Definitions 2.1 or 2.2, we say that a state \(\rho \) represents thermal equilibrium if

$$\begin{aligned} \mathrm{Tr}[\rho \,\mathsf {P}\!_\mathrm {neq}]\le e^{-\alpha V}, \end{aligned}$$
(12.1)

where \(\mathsf {P}\!_\mathrm {neq}\) is defined by (2.8), (2.12) or (2.15) depending on the situation and the treatment.

Take an initial state \(\rho (0)\), and let \(\rho (t)=e^{-i\mathsf {H}_Vt}\rho (0)\,e^{i\mathsf {H}_Vt}\) be its time evolution. Then the statement corresponding to Lemma 5.1 reads

Lemma 12.1

Suppose that there is \(\tau >0\) and it holds that

$$\begin{aligned} \frac{1}{\tau }\int _0^\tau dt\,\mathrm{Tr}[\rho (t)\,\mathsf {P}\!_\mathrm {neq}] \le e^{-(\alpha +\nu ) V}. \end{aligned}$$
(12.2)

Then there exists a collection of intervals \(\mathcal{G}\subset [0,\tau ]\) such that \(|\mathcal{G}|/\tau \ge 1-e^{-\nu V}\), and we have for any \(t\in \mathcal{G}\) that

$$\begin{aligned} \mathrm{Tr}[\rho (t)\,\mathsf {P}\!_\mathrm {neq}]\le e^{-\alpha V}, \end{aligned}$$
(12.3)

which means that \(\rho (t)\) represents thermal equilibrium in the sense of (12.1).

To discuss the extension of Theorem 6.1, it is useful to define two effective dimensions. The first is

$$\begin{aligned} D_\mathrm {eff}^\mathrm{full}:=\frac{1}{\mathrm{Tr}[\{\rho (t)\}^2]}, \end{aligned}$$
(12.4)

which is the inverse of the purity \(\mathrm{Tr}[\{\rho (t)\}^2]\), and is clearly independent of t. Note that one has \(D_\mathrm {eff}^\mathrm{full}=1\) when \(\rho (t)\) represents a pure state.

The effective dimension which corresponds to \(D_\mathrm {eff}\) of (6.2) is defined using the energy eigenstate basis \(\{|\psi _j\rangle \}_{j=1,\ldots ,D_{V,u}}\) as

$$\begin{aligned} D_\mathrm {eff}^\mathrm{diag}:=\Bigl \{\sum _j\bigl (\langle \psi _j|\rho (t)|\psi _j\rangle \bigr )^2\Bigr \}^{-1}=\frac{1}{\mathrm{Tr}[(\rho _\mathrm{diag})^2]}, \end{aligned}$$
(12.5)

where

$$\begin{aligned} \rho _\mathrm{diag}:=\sum _k|\psi _j\rangle \langle \psi _j|\rho (t)|\psi _j\rangle \langle \psi _j| \end{aligned}$$
(12.6)

is the density matrix for the “diagonal ensemble” obtained by deleting the off-diagonal elements of \(\rho (t)\). Note that \(\rho _\mathrm{diag}\) and \(D_\mathrm {eff}^\mathrm{diag}\) are independent of t.

These effective dimensions satisfy the bound \(D_\mathrm {eff}^\mathrm{full}\le D_\mathrm {eff}^\mathrm{diag}\) because of the monotonicity of the purity.Footnote 73

We first note that the state is in thermal equilibrium to begin with if \(D_\mathrm {eff}^\mathrm{full}\) is sufficiently large.

Theorem 12.1

If the thermodynamic bound (2.16) is valid, and one has

$$\begin{aligned} D_\mathrm {eff}^\mathrm{full}\ge e^{-\eta V}D_{V,u}, \end{aligned}$$
(12.7)

with \(\eta \) satisfying \(\gamma -\eta >2(\alpha +\nu )\), then \(\rho (t)\) represents thermal equilibrium for any t (including \(t=0\)).

Proof

From the Schwarz inequality, we get

$$\begin{aligned} \mathrm{Tr}[\rho (t)\mathsf {P}\!_\mathrm {neq}]&=\mathrm{Tr}_{\mathcal {H}_{V,u}}[\rho (t)\mathsf {P}\!_\mathrm {neq}] \le \sqrt{ \mathrm{Tr}_{\mathcal {H}_{V,u}}[\{\rho (t)\}^2]\,\mathrm{Tr}_{\mathcal {H}_{V,u}}[\mathsf {P}\!_\mathrm {neq}] } \nonumber \\&=\sqrt{\frac{D_{V,u}\langle \mathsf {P}\!_\mathrm {neq}\rangle ^\mathrm {mc}_{V,u}}{D_\mathrm {eff}^\mathrm{full}}}. \end{aligned}$$
(12.8)

Then the statement follows exactly as in the proof of Theorem 6.1. \(\square \)

Of course it is much more interesting if (12.7) is not valid. The following is a straightforward extension of Theorem 6.1.

Theorem 12.2

If the thermodynamic bound (2.16) is valid, and one has \(D_\mathrm {eff}^\mathrm{diag}\ge e^{-\eta V}D_{V,u}\) with \(\eta \) satisfying \(\gamma -\eta >2(\alpha +\nu )\), then \(\rho (t)\) approaches thermal equilibrium (in the sense of Lemma 5.1).

Proof

Note that the nondegeneracy of \(E_j\) implies

$$\begin{aligned} \lim _{\tau \uparrow \infty }\frac{1}{\tau }\int _0^\tau dt\,\mathrm{Tr}[\rho (t)\,\mathsf {P}\!_\mathrm {neq}]&= \lim _{\tau \uparrow \infty }\frac{1}{\tau }\int _0^\tau dt\sum _{j,j'\in J_{V,u}}e^{i(E_j-E_{j'})t}\langle \psi _{j'}|\rho (0)|\psi _j\rangle \langle \psi _j|\mathsf {P}\!_\mathrm {neq}|\psi _{j'}\rangle \nonumber \\&=\sum _{j\in J_{V,u}}\langle \psi _{j}|\rho (0)|\psi _j\rangle \langle \psi _j|\mathsf {P}\!_\mathrm {neq}|\psi _{j}\rangle =\mathrm{Tr}[\rho _\mathrm{diag}\mathsf {P}\!_\mathrm {neq}]. \end{aligned}$$
(12.9)

Then the rest of the proof is the same as that of Theorem 12.1. \(\square \)

Theorem 7.1, which makes use of the energy eigenstate thermalization hypothesis, also extends to the case with mixed initial state. We omit the details since the extension is trivial.

Appendix 4: Typicality of the Canonical Expectation Values

We have exclusively discussed the microcanonical setting in the present paper. Here we apply some of the techniques in the present paper to the setting where the system of interest is coupled to a heat bath, and show the typicality of the canonical expectation values.

Setting: We consider a macroscopic quantum system with volume V, and denote by \(\mathcal{H}_\mathrm {S}\) and \(\mathsf {H}_\mathrm {S}\) the Hilbert space and the Hamiltonian, respectively. The system is coupled to a heat bath (reservoir) which itself is a macroscopic quantum system with volume \(V_\mathrm {B}=\lambda V\), Hilbert space \(\mathcal{H}_\mathrm {B}\), and Hamiltonian \(\mathsf {H}_\mathrm {B}\). Here \(\lambda \) is a fixed constant which is usually taken to beFootnote 74 \(\lambda \gg 1\).

The whole system, i.e., the system plus the heat bath, has the volume \({V_\mathrm {tot}}=(1+\lambda )V\), the Hilbert space \(\mathcal{H}_\mathrm {tot}=\mathcal{H}_\mathrm {S}\otimes \mathcal{H}_\mathrm {B}\), and the Hamiltonian

$$\begin{aligned} \mathsf {H}_\mathrm {tot}:=\mathsf {H}_\mathrm {S}\otimes \mathsf {1}+\mathsf {1}\otimes \mathsf {H}_\mathrm {B}+\mathsf {H}_\mathrm {int}, \end{aligned}$$
(13.1)

where \(\mathsf {H}_\mathrm {int}\) describes the interaction between the system and the bath. We assume \(\left\| \mathsf {H}_\mathrm {int}\right\| \le h_0V^\zeta \) with constants \(h_0>0\) and \(0<\zeta <1\). The energy shell \(\mathcal {H}_{{V_\mathrm {tot}},u}\) for the whole system is defined as in Sect. 2.1 by replacing V with \({V_\mathrm {tot}}\).

Canonical Typicality: Popescu, Short, and Winter [12] and Goldstein, Lebowitz, Tumulka, and Zanghì [13] independently stated the important result known as the canonical typicality (see also Sugita [14]).

Let \({\rho }^\mathrm {mc}_{{V_\mathrm {tot}},u}\) be the microcanonical density matrix for the whole system, defined as in (2.5) by replacing V with \({V_\mathrm {tot}}\). Let \(\rho _\mathrm {S}:=\mathrm{Tr}_\mathrm {B}[\,{\rho }^\mathrm {mc}_{{V_\mathrm {tot}},u}\,]\) be the reduced density matrix for the system, where \(\mathrm {Tr}_\mathrm {B}\) denotes the trace in \(\mathcal{H}_\mathrm {B}\). It is usually expected (and can be proved with suitable assumptions) that \(\rho _\mathrm {S}\) is close to the canonical density matrix for the system with the inverse temperature \(\beta (u)\) defined in (2.2).

Take an arbitrary normalized state \(|\varphi \rangle \in \tilde{\mathcal {H}}_{{V_\mathrm {tot}},u}\) from the energy shell, and write the corresponding density matrix for the system as \(\rho _\varphi :=\mathrm{Tr}_\mathrm{B}[\,|\varphi \rangle \langle \varphi |\,]\). It has been shown that, when the state \(|\varphi \rangle \) is sampled from \(\tilde{\mathcal {H}}_{{V_\mathrm {tot}},u}\) randomly according to the uniform measure as in Sect. 4.1, the density matrices \(\rho _\mathrm {S}\) and \(\rho _\varphi \) are very close to each other with probability close to one. This is the canonical typicality.

Therefore, when one is only interested in physical quantities of the system, it is typical that the state of the system is described by the canonical distribution. We believe that this provides a satisfactory characterization and justification of the canonical distribution.

Main Result: To complete the justification of the canonical distribution from the operational point of view, we still need to show that (i) the result of a single measurement of a macroscopic quantity of the system almost coincides with the expectation value with respect to \(\rho _\mathrm {S}\), and (ii) \(\rho _\mathrm {S}\) is indeed close to the canonical density matrix.

Both (i) and (ii) can be done starting from the canonical typicality. But let us here present a result (which does not make an explicit use of the canonical typicality) which attains the goal directly.

Define the canonical expectation for the system as

$$\begin{aligned} \langle \cdots \rangle ^\mathrm {can}_{V,\beta }:= \frac{\mathrm{Tr}_\mathrm{S}[\,(\cdots )\,e^{-\beta \mathsf {H}_\mathrm{S}}\,]}{\mathrm{Tr}_\mathrm{S}[\,e^{-\beta \mathsf {H}_\mathrm{S}}\,]}, \end{aligned}$$
(13.2)

where \(\mathrm{Tr}_\mathrm{S}[\cdots ]\) denotes the trace over the space \(\mathcal{H}_\mathrm{S}\). Note that neither \(\mathsf {H}_\mathrm{int}\) nor \(\mathsf {H}_\mathrm{B}\) appears in the definition (13.2).

We assume that the whole system satisfies the condition for the number of states (2.1), with V replaced by \({V_\mathrm {tot}}\).

As in Sect. 2.2, we let \(\mathsf {M}_V^{(1)},\ldots ,\mathsf {M}_V^{(n)}\) be extensive quantities and define the nonequilibrium projection \(\mathsf {P}\!_\mathrm {neq}\) as (2.8), (2.12) or (2.15) , but by replacing \(m^{(i)}(u)\simeq \langle \mathsf {M}_V^{(i)}\rangle ^\mathrm{mc}_{V,u}/V\) by the canonical expectation value \(\langle \mathsf {M}_V^{(i)}\rangle ^\mathrm{can}_{V,\beta }/V\). We then make a crucial assumption that the canonical version of the thermodynamic bound

$$\begin{aligned} \langle \mathsf {P}\!_\mathrm {neq}\rangle ^\mathrm{can}_{V,\beta }\le e^{-\gamma '(\beta )\,V} \end{aligned}$$
(13.3)

is valid for sufficiently large V with \(\gamma '(\beta )>0\). This is almost the standard large deviation upper bound, which is expected to be valid in general, and can be proved for models treated in Sects. 3.2 and 3.3.

Theorem 13.1

Take the energy density u such that (3.1) is valid, and let \(\beta =\beta (u)\). Assume that the bound (13.3) holds. We choose a normalized state \(|\varphi \rangle \in \tilde{\mathcal {H}}_{{V_\mathrm {tot}},u}\) randomly according to the uniform measure as in Theorem 4.1. Then with probability larger than \(1-e^{-\nu 'V}\), we have

$$\begin{aligned} \langle \varphi |\mathsf {P}\!_\mathrm {neq}\otimes \mathsf {1}|\varphi \rangle \le e^{-\alpha 'V}, \end{aligned}$$
(13.4)

for sufficiently large V, where the constants satisfy \(\nu '\simeq \gamma '(\beta )-\alpha '\).

The theorem says that, for an overwhelming majority of the states in the energy shell, the result of a single measurement of \(\mathsf {M}_V^{(i)}\) almost coincides with the canonical expectation value \(\langle \mathsf {M}_V^{(i)}\rangle ^\mathrm{can}_{V,\beta }\) with probability close to one. This provides a rather satisfactory justification of the canonical distribution.

As we noted before the ration \(\lambda \) need not be large. This is because the proof makes use of the equivalence of the microcanonical and the canonical ensembles. When \(\lambda \) is large, however, we see that the inverse temperature \(\beta \) is essentially determined by the properties and the energy of the heat bath.

Proof of Theorem 13.1

We shall prove that

$$\begin{aligned} \overline{\langle \varphi |\mathsf {P}\!_\mathrm {neq}\otimes \mathsf {1}|\varphi \rangle }\le e^{-\gamma ''V}, \end{aligned}$$
(13.5)

where the left-hand side is the average over \(|\varphi \rangle \in \tilde{\mathcal {H}}_{{V_\mathrm {tot}},u}\) as defined in (4.1) and \(\gamma ''\simeq \gamma '(\beta )\). Then by using the Markov inequality as in the proof of Theorem 4.1, we get the desired (13.4).

To bound the average, we note that

$$\begin{aligned} \overline{\langle \varphi |\mathsf {P}\!_\mathrm {neq}\otimes \mathsf {1}|\varphi \rangle }= \langle \mathsf {P}\!_\mathrm {neq}\otimes \mathsf {1}\rangle ^\mathrm{mc}_{{V_\mathrm {tot}},u} \le \eta (u){V_\mathrm {tot}}\langle \mathsf {P}\!_\mathrm {neq}\otimes \mathsf {1}\rangle ^\mathrm{can}_{{V_\mathrm {tot}},\beta (u)}, \end{aligned}$$
(13.6)

where \(\langle \cdots \rangle ^\mathrm{mc}_{{V_\mathrm {tot}},u}\) and \(\langle \cdots \rangle ^\mathrm{can}_{{V_\mathrm {tot}},\beta }\) are the microcanonical and the canonical expectations, respectively, of the whole system with Hamiltonian (13.1). The equality and the inequality in (13.6) follow from (4.3) and (8.8), respectively. Then exactly as in (8.31), the final expectation is bounded as

$$\begin{aligned} \langle \mathsf {P}\!_\mathrm {neq}\otimes \mathsf {1}\rangle ^\mathrm{can}_{{V_\mathrm {tot}},\beta }\le e^{2\beta h_0V^\zeta }\langle \mathsf {P}\!_\mathrm {neq}\rangle ^\mathrm{can}_{V,\beta }, \end{aligned}$$
(13.7)

where the expectation in the right-hand side is the one defined in (13.2). By combining (13.6), (13.7), and the assumption (13.3), we get (13.5). \(\square \)

It is a pleasure to thank Shelly Goldstein and Takashi Hara, my collaborators in closely related topics, for valuable discussions and inspirations, Shin Nakano, Yoshiko Ogata, and Luc Rey-Bellet for their indispensable help in mathematical issues, and Marcus Cramer, Fabian Essler, Tatsuhiko Ikeda, Joel Lebowitz, Elliott Lieb, Takashi Mori, Hidetoshi Nishimori, Peter Reimann, Marcos Rigol, Takahiro Sagawa, Udo Seifert, Akira Shimizu, and Yu Watanabe for useful discussions and comments. The present work was supported in part by JSPS Grants-in-Aid for Scientific Research no. 25400407.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tasaki, H. Typicality of Thermal Equilibrium and Thermalization in Isolated Macroscopic Quantum Systems. J Stat Phys 163, 937–997 (2016). https://doi.org/10.1007/s10955-016-1511-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-016-1511-2

Keywords

Navigation