Skip to main content
Log in

Nature Abhors a Vacuum: A Simple Rigorous Example of Thermalization in an Isolated Macroscopic Quantum System

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

Abstract

We show, without relying on any unproven assumptions, that a low-density free fermion chain exhibits thermalization in the following (restricted) sense. We choose the initial state as a pure state drawn randomly from the Hilbert space in which all particles are in half of the chain. This represents a nonequilibrium state such that the half chain containing all particles is in equilibrium at infinite temperature, and the other half chain is a vacuum. We let the system evolve according to the unitary time evolution determined by the Hamiltonian and, at a sufficiently large typical time, measure the particle number in an arbitrary macroscopic region in the chain. In this setup, it is proved that the measured number is close to the equilibrium value with probability very close to one. Our result establishes the presence of thermalization in a concrete model in a mathematically rigorous manner. The key for the proof is a new strategy to show that a randomly generated nonequilibrium initial state typically has a large enough effective dimension by using only mild verifiable assumptions. In the present work, we first give general proof of thermalization based on two assumptions, namely, the absence of degeneracy in energy eigenvalues and a property about the particle distribution in energy eigenstates. We then justify these assumptions in a concrete free-fermion model, where the absence of degeneracy is established by using number-theoretic results. This means that our general result also applies to any lattice gas models in which the above two assumptions are justified. To confirm the potential wide applicability of our theory, we discuss some other models for which the essential assumption about the particle distribution is easily verified, and some non-random initial states whose effective dimensions are sufficiently large.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

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

Instant access to the full article PDF.

Similar content being viewed by others

Data Availibility Statement

All data and information relevant to this study are presented in the paper.

Notes

  1. It is a common misconception that the prediction of equilibrium statistical mechanics should always be compared with an averaged quantity in the corresponding physical system. In fact, the law of large numbers guarantees that the statistical mechanical expectation value accurately predicts the outcome of a single measurement in the equilibrium state, provided that both the system and the quantity to be measured are macroscopic.

  2. Researchers who emphasize microscopic mechanisms may not call the process thermalization since everything is governed by free particle dynamics. Our point is to focus only on macroscopically observable phenomena, assuming that the observer has no access to microscopic mechanisms.

  3. All the results in Sect. 2 are also valid for a system of hardcore bosons.

  4. To be precise this is true only when the final state represents the equilibrium state at infinite temperature (as in our case). In general, if the initial state \(|\Phi (0)\rangle \) has energy close to E then the effective dimension is believed to be close to the dimension of the corresponding energy shell, i.e., the Hilbert space spanned by energy eigenstates whose eigenvalues are close to E. One can argue, although very heuristically, that a large effective dimension is a consequence of (a strong form of) ETH. Consider a system described by a short-ranged translation-invariant Hamiltonian \(\hat{H}\) and assume that ETH is valid. For simplicity, we take the initial state \(|\Phi (0)\rangle \) to be a translation invariant product state. (We assume \(|\Phi (0)\rangle \) is not an eigenstate of \(\hat{H}\).) Then \(|\Phi (0)\rangle \) has energy distribution peaked around some value E. Let \(|\Psi _j\rangle \) be the eigenstate of \(\hat{H}\) with eigenvalue \(E_j\). Since ETH asserts that energy eigenstates with close eigenvalues are similar to each other, it is reasonable to assume that the overlap \(|\langle \Phi (0)|\Psi _j\rangle |^2\) is almost independent of j as long as \(E_j\simeq E\). This implies that \(D_\textrm{eff}\) is almost identical to the dimension of the energy shell around E.

  5. We note that the diagonal entropy \(S_\textrm{d}\) studied in these works is believed to be related to the effective dimension as \(D_\textrm{eff}\sim \exp [S_\textrm{d}]\).

  6. \({\varDelta }\hat{H}\) may not be small in the class of models considered in Appendix B.

  7. The theorem was proved by one of us in [53]. See also Proposition 10.1 in [17] for a similar statement for a slightly complicated model.

  8. See [55] for elementary proofs of the two lemmas.

  9. We learned the lemma and its proof from Wataru Kai and Kazuaki Miyatani.

  10. As is suggested by this conclusion, one can prove, by using essentially the same argument, the absence of degeneracy in certain open fermion chains with a suitable boundary potential.

  11. In this trivial model, the energy eigenvalues for a pair of sites (x, 1) and (x, 2) are either zero (when there is no particles), \(\pm s_x+w_x\) (when there is one particle), or \(2w_x+u_x\) (when there are two particles). The total energy eigenvalues are the sums of these eigenvalues and are nondegenerate if we choose \(s_x\), \(w_x\), and \(u_x\) properly.

  12. A non-optimal but simple example of a Golomb ruler is obtained by taking N such that \(L=2^N-1\) is a (Mersenne) prime, and setting \(x_j=2^{j-1}\) for \(j=1,\ldots ,N\). In this case, the particle density \(\rho \simeq N/2^N\) is exponentially small in N.

  13. In a Golomb ruler, \(x_{j}-x_{Q(j)}-x_{P'(j)}+x_{Q'(j)}=0\mod L\) holds only if (1) \(j=P'(j)\) and \(Q(j)=Q'(j)\), or (2) \(j=Q(j)\) and \(P'(j)=Q'(j)\). Now we decompose a set \(\{1,2,\ldots , N\}\) into two subsets, A and B, where (1) holds in A and (2) holds in B. Then, \(P'\), Q, \(Q'\) can be expressed in the form of \(P'=\textrm{id}^A\oplus \pi ^A\), \(Q=\pi ^B\oplus \textrm{id}^B\), and \(Q'=\pi ^A\oplus \pi ^B\), where \(\pi ^A\) and \(\pi ^B\) represent permutations on A and B, respectively. With the above form of permutations, we easily see \((-1)^{P'QQ'}=1\) if \(x_{j}-x_{Q(j)}-x_{P'(j)}+x_{Q'(j)}=0\mod L\) holds for any j.

  14. A derangement is a permutation in which no entry stays at the original position.

  15. The floor function \(\lfloor x\rfloor \) is the largest integers less than or equal to x.

References

  1. There is a 27-minute video that explains the main results of the present work: https://youtu.be/eUIX8WdJftc

  2. 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, The European Phys. J. H 35 201–237 (2010). arxiv:1003.2133

  3. Kinoshita, T., Wenger, T., Weiss, D.S.: A quantum Newton’s cradle. Nature 440, 900 (2006)

    Article  ADS  Google Scholar 

  4. Trotzky, S., Chen, Y.-A., Flesch, A., McCulloch, I.P., Schollwöck, U., Eisert, J., Bloch, I.: Probing the relaxation towards equilibrium in an isolated strongly correlated one-dimensional Bose gas. Nat. Phys. 8, 325 (2012). arxiv:1101.2659

    Article  Google Scholar 

  5. Gring, M., Kuhnert, M., Langen, T., Kitagawa, T., Rauer, B., Schreitl, M., Mazets, I., Adu Smith, D., Demler, E., Schmiedmayer, J.: Relaxation and Prethermalization in an Isolated Quantum System. Science 337, 1318 (2012). arxiv:1112.0013

    Article  ADS  Google Scholar 

  6. Langen, T., Erne, S., Geiger, R., Rauer, B., Schweigler, T., Kuhnert, M., Rohringer, W., Mazets, I.E., Gasenzer, T., Schmiedmayer, J.: Experimental observation of a generalized Gibbs ensemble. Science 348, 207 (2015). arxiv:1411.7185

    Article  ADS  MathSciNet  Google Scholar 

  7. Kaufman, A.M., Tai, M.E., Lukin, A., Rispoli, M., Schittko, R., Preiss, P.M., Greiner, M.: Quantum thermalization through entanglement in an isolated many-body system. Science 353, 794 (2016). arxiv:1603.04409

  8. D’Alessio, L., Kafri, Y., Polkovnikov, A., Rigol, M.: From quantum chaos and eigenstate thermalization to statistical mechanics and thermodynamics, Adv. Phys. 65, 239–362 (2016). arxiv:1509.06411

  9. Gogolin, C., Eisert, J.: Equilibration, thermalisation, and the emergence of statistical mechanics in closed quantum systems. Rep. Prog. Phys. 79, 056001 (2016). arxiv:1503.07538

  10. 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. European Phys. J. H 35, 173–200 (2010). arxiv:1003.2129

    Article  ADS  Google Scholar 

  11. Rigol, M., Srednicki, M.: Alternatives to Eigenstate Thermalization. Phys. Rev. Lett. 108, 110601 (2012). arxiv:1108.0928

    Article  ADS  Google Scholar 

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

    Article  ADS  Google Scholar 

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

    Article  ADS  Google Scholar 

  14. 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

  15. 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 

  16. Reimann, P.: Generalization of von Neumann’s Approach to Thermalization, Phys. Rev. Lett. 115, 010403 (2015). arxiv:1507.00262

  17. Tasaki, H.: Typicality of thermal equilibrium and thermalization in isolated macroscopic quantum systems, J. Stat. Phys. 163, 937–997 (2016). arxiv:1507.06479

  18. Reimann, P.: Foundation of Statistical Mechanics under Experimentally Realistic Conditions. Phys. Rev. Lett. 101, 190403 (2008). arxiv:0810.3092

    Article  ADS  Google Scholar 

  19. 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 

  20. 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

  21. Reimann, P.: Equilibration of Isolated Macroscopic Quantum Systems under Experimentally Realistic Conditions. Phys. Scr. 86, 058512 (2012). arxiv:1210.5821

    Article  Google Scholar 

  22. 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

  23. Shiraishi, N.: Analytic model of thermalization: Quantum emulation of classical cellular automata, Phys. Rev. E 97, 062144 (2018). arxiv:1709.06315

  24. Grad, H.: Statistical mechanics, thermodynamics, and fluid dynamics of systems with an arbitrary number of integrals. Pure and Applied Mathematics 5, 455 (1952)

    Article  MathSciNet  Google Scholar 

  25. Jancel, R.: Foundations of Classical and Quantum Statistical Mechanics. Pergamon Press, UK (1969)

    Google Scholar 

  26. Cazalilla, M.A.: Effect of Suddenly Turning on Interactions in the Luttinger Model. Phys. Rev. Lett. 97, 156403 (2006)

    Article  ADS  Google Scholar 

  27. 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

  28. Basko, D.M., Aleiner, I.L., Altshuler, B.L.: Metal-insulator transition in a weakly interacting many-electron system with localized single-particle states. Ann. Phys. 321, 1126 (2006). arXiv:cond-mat/0506617

  29. Pal, A., Huse, D.A.: Many-body localization phase transition. Phys. Rev. B 82, 174411 (2010). arxiv:1010.1992

    Article  ADS  Google Scholar 

  30. Serbyn, M., Papić, Z., Abanin, D.A.: Local Conservation Laws and the Structure of the Many-Body Localized States. Phys. Rev. Lett. 111, 127201 (2013). arxiv:1305.5554

    Article  ADS  Google Scholar 

  31. Friesdorf, M., Werner, A.H., Brown, W., Scholz, V.B., Eisert, J.: Many-Body Localization Implies that Eigenvectors are Matrix-Product States. Phys. Rev. Lett. 114, 170505 (2015). arxiv:1409.1252

    Article  ADS  Google Scholar 

  32. Nandkishore, R., Huse, D.A.: Many body localization and thermalization in quantum statistical mechanics. Ann. Rev. Cond. Matt. Phys. 6, 15 (2015). arxiv:1404.0686

    Article  ADS  Google Scholar 

  33. Imbrie, J.Z.: On Many-Body Localization for Quantum Spin Chains. J. Stat. Phys. 163, 998 (2016). arxiv:1403.7837

    Article  ADS  MathSciNet  Google Scholar 

  34. Bernien, H., Schwartz, S., Keesling, A., Levine, H.y, Omran, A., Pichler, H., Choi, S., Zibrov, A.S., Endres, M., Greiner, M., Vuletic, V., Lukin, M.D.: Probing many-body dynamics on a 51-atom quantum simulator. Nature 551, 579 (2017). arxiv:1707.04344

  35. Shiraishi, N., Mori, T.: Systematic Construction of Counterexamples to Eigenstate Thermalization Hypothesis. Phys. Rev. Lett. 119, 030601 (2017). arxiv:1702.08227

  36. Mori, T., Shiraishi, N.: Thermalization without eigenstate thermalization hypothesis after a quantum quench. Phys. Rev. E 96, 022153 (2017). arxiv:1707.05921

  37. Turner, C.J., Michailidis, A.A., Abanin, D.A., Serbyn, M., Papić, Z.: Weak ergodicity breaking from quantum many-body scars. Nature Physics 14, 745 (2018). arxiv:1711.03528

  38. Moudgalya, S., Rachel, S., Bernevig, B.A., Regnault, N.: Exact Excited States of Non-Integrable Models. Phys. Rev. B 98, 235155 (2018). arxiv:1708.05021

  39. Moudgalya, S., Regnault, N., Bernevig, B.A.: Entanglement of exact excited states of Affleck-Kennedy-Lieb-Tasaki models: Exact results, many-body scars, and violation of the strong eigenstate thermalization hypothesis. Phys. Rev. B 98, 235156 (2018). arxiv:1806.09624

  40. Lin, C.-J., Motrunich, O.I.: Exact Quantum Many-Body Scar States in the Rydberg-Blockaded Atom Chain. Phys. Rev. Lett. 122, 173401 (2019). arxiv:1810.00888

  41. Ho, W.W., Choi, S., Pichler, H., Lukin, M.D.: Periodic Orbits, Entanglement, and Quantum Many-Body Scars in Constrained Models: Matrix Product State Approach. Phys. Rev. Lett. 122, 040603 (2019). arxiv:1807.01815

  42. Shiraishi, N.: Connection between quantum-many-body scars and the AKLT model from the viewpoint of embedded Hamiltonians. J. Stat. Mech. 083103 (2019). arxiv:1904.05182

  43. Serbyn, M., Abanin, D.A., Papić, Z.: Quantum Many-Body Scars and Weak Breaking of Ergodicity. Nat. Phys. 17, 675 (2021). arxiv:2011.09486

  44. Shiraishi, N., Matsumoto, K.: Undecidability in quantum thermalization, Nature Comm. 12, 5084 (2021). https://www.nature.com/articles/s41467-021-25053-0

  45. Rigol, M., Muramatsu, A., Olshanii, M.: Hard-core bosons on optical superlattices: Dynamics and relaxation in the superfluid and insulating regimes, Phys. Rev. A 74, 053616 (2006). arxiv:cond-mat/0612415

  46. Rigol, M., Fitzpatrick, M.: Initial state dependence of the quench dynamics in integrable quantum systems. Phys. Rev. A 84, 033640 (2011). arxiv:1107.5811

    Article  ADS  Google Scholar 

  47. Pandey, S., Bhat, J.M., Dhar, A., Goldstein, S., Huse, D.A., Kulkarni, M., Kundu, A., Lebowitz, J.L.: Boltzmann entropy of a freely expanding quantum ideal gas, J. Stat. Phys. 190, article number 142, (2023). arxiv:2303.12330

  48. Santos, L.F., Polkovnikov, A., Rigol, M.: Entropy of Isolated Quantum Systems after a Quench. Phys. Rev. Lett. 107, 040601 (2011). arxiv:1103.0557

    Article  ADS  Google Scholar 

  49. Rigol, M.: Quantum Quenches in the Thermodynamic Limit. Phys. Rev. Lett. 112, 170601 (2014). arxiv:1401.2160

    Article  ADS  Google Scholar 

  50. Rigol, M.: Fundamental Asymmetry in Quenches Between Integrable and Nonintegrable Systems, Phys. Rev. Lett. 116, 100601 (2016). arxiv:1511.04447

  51. Farrelly, T., Brandao, F.G.S.L., Cramer, M.: Thermalization and Return to Equilibrium on Finite Quantum Lattice Systems, Phys. Rev. Lett. 118, 140601 (2017). arxiv:1610.01337

  52. Ullah, N.: Invariance hypothesis and higher correlations of Hamiltonian matrix elements, Nuc. Phys. 58, 65 (1964).https://www.sciencedirect.com/science/article/pii/002955826490522X

  53. Tasaki, H.: The approach to thermal equilibrium and “thermodynamic normality” — An observation based on the works by Goldstein, Lebowitz, Mastrodonato, Tumulka, and Zanghi in 2009, and by von Neumann in 1929, (unpublished note 2010). arxiv:1003.5424

  54. Abdul-Rahman, H., Stolz, G.: A uniform area law for the entanglement of eigenstates in the disordered XY chain, J. Math. Phys. 56, 121901 (2015). arxiv:1505.02117

  55. Tasaki, H.: Two number-theoretic theorems (that we found useful for quantum physics) and their elementary proofs, YouTube video (2023). https://youtu.be/YrCoBv0acgs

  56. Tignol, J.-P.: Galois’ Theory of Algebraic Equations (second edition), (World Scientific, 2015)

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

    Google Scholar 

  58. Stewart, I., Tall, D.: Algebraic number theory and Fermat’s last theorem. Chapman and Hall, USA (2020)

    Google Scholar 

  59. Mathworld, Golomb ruler. https://mathworld.wolfram.com/GolombRuler.html

  60. Singer, J.: A Theorem in Finite Projective Geometry and Some Applications to Number Theory. Trans. Am. Math. Soc. 43, 377–385 (1938). https://www.jstor.org/stable/1990067

  61. Erdös, P., Turán, P.: On a problem of Sidon in additive number theory and some related problems, J. London Math. Soc. 16, 212–215 (1941). https://londmathsoc.onlinelibrary.wiley.com/doi/abs/10.1112/jlms/s1-16.4.212

  62. Mathworld, derangement. https://mathworld.wolfram.com/Derangement.html

  63. Teufel, S., Tumulka, R., Vogel, C.: Time Evolution of Typical Pure States from a Macroscopic Hilbert Subspace, J. Stat. Phys. 190, 69 (2023). arxiv:2210.10018

Download references

Acknowledgements

It is a pleasure to thank Shelly Goldstein, Takashi Hara, Shu Nakamura, and Marcos Rigol for useful discussions. We also thank Shin Nakano for his patient guidance in number theory, and Wataru Kai and Kazuaki Miyatani for letting us know of Lemma 3.4 and its proof. N.S. was supported by JSPS Grants-in-Aid for Early-Career Scientists No. JP19K14615, and H.T. by JSPS Grants-in-Aid for Scientific Research No. 22K03474.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hal Tasaki.

Ethics declarations

Conflict of interest

The authors declare no Conflict of interest.

Additional information

Communicated by Anatoli Polkovnikov.

Publisher's Note

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

Appendices

Appendices

A: Models With Degenerate Energy Eigenvalues

Our general discussion in Sect. 2 is based on the crucial assumption, Assumption 2.1, that all the energy eigenvalues are nondegenerate. Here we shall see how one can treat models in which the degree of degeneracy is at most \(d_\textrm{max}\). We find that our thermalization results remain valid as long as \(d_\textrm{max}\) is not too large. Unfortunately, we do not know of any examples where a nontrivial upper bound for the degree of degeneracy is known.

Let \(E_j\) with \(j=1,\ldots ,N_\textrm{el}\) be the distinct energy eigenvalues. We denote by \(|\Psi _{j,\ell }\rangle \) with \(\ell =1,\ldots ,d_j\) the energy eigenstates corresponding to \(E_j\), where \(d_j\) is the degree of degeneracy of \(E_j\). We assume that the collection of \(|\Psi _{j,\ell }\rangle \) with all j, \(\ell \) forms an orthonormal basis of \(\mathcal{H}_\textrm{tot}\).

We first examine the discussion in Sect. 2.2 about the effective dimension. A straightforward generalization of the definition (2.6) of the effective dimension is

$$\begin{aligned} D_\textrm{eff}=\biggl (\sum _{j=1}^{N_\textrm{el}}\sum _{\ell =1}^{d_j}\bigl |\left\langle \Phi (0)|\Psi _{j,\ell }\right\rangle \bigr |^4\biggr )^{-1}. \end{aligned}$$
(A.1)

When energy eigenvalues are degenerate, however, it is convenient to employ the definition

$$\begin{aligned} \widetilde{D}_\textrm{eff}=\biggl (\sum _{j=1}^{N_\textrm{el}}\langle \Phi (0)|\hat{P}_j|\Phi (0)\rangle ^2\biggr )^{-1}, \end{aligned}$$
(A.2)

where \(\hat{P}_j=\sum _{\ell =1}^{d_j}|\Psi _{j,\ell }\rangle \langle \Psi _{j,\ell }|\) is the projection onto the space corresponding to the energy eigenvalue \(E_j\). Clearly, (A.2) reduces to the original (A.1) when there is no degeneracy. To evaluate (A.2), we note that

$$\begin{aligned} \langle \Phi (0)|\hat{P}_j|\Phi (0)\rangle ^2&=\sum _{\ell ,\ell '=1}^{d_j}\bigl |\left\langle \Phi (0)|\Psi _{j,\ell }\right\rangle \bigr |^2\,\bigl |\left\langle \Phi (0)|\Psi _{j,\ell '}\right\rangle \bigr |^2 \nonumber \\&\le \frac{1}{2}\sum _{\ell ,\ell '=1}^{d_j}\Bigl (\bigl |\left\langle \Phi (0)|\Psi _{j,\ell }\right\rangle \bigr |^4+\bigl |\left\langle \Phi (0)|\Psi _{j,\ell '}\right\rangle \bigr |^4\Bigr ) \nonumber \\&=d_j\sum _{\ell =1}^{d_j}\bigl |\left\langle \Phi (0)|\Psi _{j,\ell }\right\rangle \bigr |^4, \end{aligned}$$
(A.3)

where we noted \(ab\le (a^2+b^2)/2\) to get the second line. We thus find

$$\begin{aligned} \widetilde{D}_\textrm{eff}^{-1}\le \sum _{j=1}^{N_\textrm{el}}d_j\sum _{\ell =1}^{d_j}\bigl |\left\langle \Phi (0)|\Psi _{j,\ell }\right\rangle \bigr |^4 \le d_\textrm{max}\sum _{j=1}^{N_\textrm{el}}\sum _{\ell =1}^{d_j}\bigl |\left\langle \Phi (0)|\Psi _{j,\ell }\right\rangle \bigr |^4 =\frac{d_\textrm{max}}{D_\textrm{eff}} \end{aligned}$$
(A.4)

where \(d_\textrm{max}=\max _jd_j\).

Suppose that (2.5) in Assumption 2.2 is valid for the energy eigenstates \(|\Psi _{j,\ell }\rangle \). Then Theorem 2.3 guarantees the crucial lower bound (2.7) for \(D_\textrm{eff}\) defined as (A.1). We thus find from (A.4) that

$$\begin{aligned} \frac{D_\textrm{tot}}{\widetilde{D}_\textrm{eff}}\le d_\textrm{max}\,e^{\rho N}. \end{aligned}$$
(A.5)

We see that \(\widetilde{D}_\textrm{eff}\) is large provided that \(d_\textrm{max}\) is not too large. Note that the degeneracy does not essentially change the behavior of the effective dimension if \(d_\textrm{max}\) grows subexponentially in N.

We move onto the discussion in Sect. 2.3 about the time evolution. Taking into account the degeneracy, the expression (2.14) for the time evolution reads

$$\begin{aligned} |\Phi (t)\rangle =e^{-i\hat{H}t}|\Phi (0)\rangle =\sum _{j=1}^{N_\textrm{el}}e^{-iE_jt}\hat{P}_j|\Phi (0)\rangle =\sum _{j=1}^{N_\textrm{el}}e^{-iE_jt}\,|\widetilde{\Psi }_j\rangle \, \sqrt{\langle \Phi (0)|\hat{P}_j|\Phi (0)\rangle }, \nonumber \\ \end{aligned}$$
(A.6)

where we defined

$$\begin{aligned} |\widetilde{\Psi }_j\rangle =\frac{\hat{P}_j|\Phi (0)\rangle }{\Vert \hat{P}_j|\Phi (0)\rangle \Vert }. \end{aligned}$$
(A.7)

Correspondingly, (2.20) is modified as

$$\begin{aligned} \lim _{T\uparrow \infty }\frac{1}{T}\int _0^Tdt\,\langle \Phi (t)|\hat{P}^{\Gamma ,\varepsilon }_\textrm{neq}|\Phi (t)\rangle&=\sum _{j=1}^{N_\textrm{el}}\langle \Phi (0)|\hat{P}_j|\Phi (0)\rangle \langle \widetilde{\Psi }_j|\hat{P}^{\Gamma ,\varepsilon }_\textrm{neq}|\widetilde{\Psi }_j\rangle \nonumber \\&\le \sqrt{ \biggl (\sum _{j=1}^{N_\textrm{el}}\langle \Phi (0)|\hat{P}_j|\Phi (0)\rangle ^2\biggr ) \biggl (\sum _{j=1}^{N_\textrm{el}}\langle \widetilde{\Psi }_j|\hat{P}^{\Gamma ,\varepsilon }_\textrm{neq}|\widetilde{\Psi }_j\rangle ^2\biggr ) } \nonumber \\&\le \sqrt{ \biggl (\sum _{j=1}^{N_\textrm{el}}\langle \Phi (0)|\hat{P}_j|\Phi (0)\rangle ^2\biggr ) \biggl (\sum _{j=1}^{N_\textrm{el}}\sum _{\ell =1}^{d_j}\langle \Psi _{j,\ell }|\hat{P}^{\Gamma ,\varepsilon }_\textrm{neq}|\Psi _{j,\ell }\rangle ^2\biggr ) } \nonumber \\&\le \sqrt{ D_\textrm{tot}\widetilde{D}_\textrm{eff}^{-1}\,\langle \hat{P}^{\Gamma ,\varepsilon }_\textrm{neq}\rangle _\infty . } \end{aligned}$$
(A.8)

Therefore the rest of the discussion remains valid if we replace \(D_\textrm{eff}\) with \(\widetilde{D}_\textrm{eff}\).

B: Models Satisfying Assumption 2.2

In this Appendix, we present two classes of models in which we can prove Assumption 2.2 about the particle distribution (under suitable assumptions about nondegeneracy). If we could also justify Assumption 2.1 about the nondegeneracy of energy eigenvalues, we would have further rigorous examples of thermalization. Unfortunately, we still do not know how nondegeneracy can be proved, although we believe it to be highly plausible.

The first class of models is that of interacting fermions on a specific class of lattices, while the second class is that of free fermions on arbitrary lattices with \(\mathbb {Z}_2\) symmetry.

1.1 B.1. Interacting Fermions on Double-Lattice

First, we discuss a class of lattice gas models on a double lattice with special symmetry. In these models, we can easily verify the bound (2.5) for any energy eigenstate corresponding to a nondegenerate energy eigenvalue. See Lemma B.1 below. This means that Assumption 2.2 about the particle distribution is automatically valid if Assumption 2.1 about nondegeneracy of energy eigenvalues is valid. The class of models in fact contains many non-trivial interacting models for which we generally expect that energy eigenvalues are nondegenerate. We thus expect that the present class contains many examples in which our thermalization theorem, Theorem 2.4, is valid. Unfortunately, we are not able to prove nondegeneracy in concrete models, except for trivial decoupled models. See the discussion at the end of the present subsection.

We shall describe the class of models and state the basic observation, i.e., Lemma B.1. Although we here describe models of fermions for notational simplicity, extensions to hardcore bosons or quantum spin systems are trivial.

Let \(\Lambda _0\) be a lattice with L/2 sites, and \(\Lambda _1\) and \(\Lambda _2\) be copies of \(\Lambda _0\). Sites in \(\Lambda _1\) and \(\Lambda _2\) are denoted as (x, 1) and (x, 2), respectively, with \(x\in \Lambda _0\). We consider a model of fermions on the whole lattice \(\Lambda =\Lambda _1\cup \Lambda _2\).

We assume that the Hamiltonian \(\hat{H}\) conserves the total particle number and is invariant under the exchange of two sites (x, 1) and (x, 2) for each \(x\in \Lambda _0\). The latter is a highly nontrivial (and artificial) assumption, which enables us to prove the desired bound (2.5) easily. To be more precise we define for each \(x\in \Lambda _0\) the unitary operator \(\hat{U}_x\) that swaps (x, 1) and (x, 2). It is defined by \(\hat{U}_x|\Phi _\textrm{vac}\rangle =|\Phi _\textrm{vac}\rangle \), \(\hat{U}_x\,\hat{c}_{(x,1)}\hat{U}_x^\dagger =\hat{c}_{(x,2)}\), \(\hat{U}_x\,\hat{c}_{(x,2)}\hat{U}_x^\dagger =\hat{c}_{(x,1)}\), and \(\hat{U}_x\,\hat{c}_{(y,\nu )}\hat{U}_x^\dagger =\hat{c}_{(y,\nu )}\) for \(y\ne x\) and \(\nu =1,2\). Note that \((\hat{U}_x)^2=\hat{1}\). Our symmetry assumption is that \([\hat{U}_x,\hat{H}]=0\) for any \(x\in \Lambda _0\).

If we restrict ourselves to models with standard particle-hopping and two-body interactions, the most general Hamiltonian takes the form

$$\begin{aligned} \hat{H}&=\mathop {\sum _{x,y\in \Lambda _0}}_{(x\ne y)}\bigl \{t_{x,y}(\hat{c}^\dagger _{(x,1)}+\hat{c}^\dagger _{(x,2)})(\hat{c}_{(y,1)}+\hat{c}_{(y,2)}) +\frac{v_{x,y}}{2}(\hat{n}_{(x,1)}+\hat{n}_{(x,2)})(\hat{n}_{(y,1)}+\hat{n}_{(y,2)})\bigr \} \nonumber \\&\qquad +\sum _{x\in \Lambda _0}\bigl \{s_x(\hat{c}^\dagger _{(x,1)}\hat{c}_{(x,2)}+\hat{c}^\dagger _{(x,2)}\hat{c}_{(x,1)})+w_x(\hat{n}_{(x,1)}+\hat{n}_{(x,2)})+u_x\,\hat{n}_{(x,1)}\hat{n}_{(x,2)}\bigr \}, \end{aligned}$$
(B.1)

where \(t_{x,y}=(t_{y,x})^*\in \mathbb {C}\), \(v_{x,y}=v_{y,x}\in \mathbb {R}\), and \(s_x,w_x,u_x\in \mathbb {R}\). We defined the number operator by \(\hat{n}_{(x,\sigma )}=\hat{c}^\dagger _{(x,\sigma )}\hat{c}_{(x,\sigma )}\).

Here is the basic observation in the present appendix.

Lemma B.1

Let \(|\Psi \rangle \) be the normalized eigenstate corresponding to a nondgenerate energy eigenvalue of \(\hat{H}\). Then we have

$$\begin{aligned} \langle \Psi |\hat{P}_1|\Psi \rangle \le 2^{-N}, \end{aligned}$$
(B.2)

which is the same as (2.5).

Proof

For a fixed particle number N, we define the basis states of the model by

$$\begin{aligned} |\Xi _{S_1,S_2}\rangle =\Bigl (\prod _{x\in S_1}\hat{c}^\dagger _{(x,1)}\Bigr )\Bigl (\prod _{x\in S_2}\hat{c}^\dagger _{(x,2)}\Bigr )|\Phi _\textrm{vac}\rangle , \end{aligned}$$
(B.3)

where \(S_1\) and \(S_2\) are arbitrary subsets of \(\Lambda _0\) such that \(|S_1|+|S_2|=N\). Take any normalized eigenstate \(|\Psi \rangle \) of \(\hat{H}\) and expand it in the above basis as

$$\begin{aligned} |\Psi \rangle =\mathop {\sum _{S_1,S_2\subset \Lambda _0}}_{(|S_1|+|S_2|=N)}\psi _{S_1,S_2}|\Xi _{S_1,S_2}\rangle , \end{aligned}$$
(B.4)

where \(\psi _{S_1,S_2}\in \mathbb {C}\) are coefficients which satisfy \(\sum |\psi _{S_1,S_2}|^2=1\). The symmetry of the Hamiltonian and the nondgeneracy imply \(\hat{U}_x|\Psi \rangle =\pm |\Psi \rangle \) for any \(x\in \Lambda _0\). This means that the expansion coefficients satisfy

$$\begin{aligned} |\psi _{S,\emptyset }|=|\psi _{S\backslash S',S'}|, \end{aligned}$$
(B.5)

for any \(S\subset \Lambda _0\) with \(|S|=N\) and any \(S'\subset S\). We thus have

$$\begin{aligned} |\psi _{S,\emptyset }|^2=\frac{1}{2^N}\sum _{S'\subset S}|\psi _{S\backslash S',S'}|^2, \end{aligned}$$
(B.6)

which, when summed over S, yields

$$\begin{aligned} \mathop {\sum _{S\subset \Lambda _0}}_{(|S|=N)}|\psi _{S,\emptyset }|^2=\frac{1}{2^N}\mathop {\sum _{S\subset \Lambda _0}}_{(|S|=N)}\sum _{S'\subset S}|\psi _{S\backslash S',S'}|^2\le \frac{1}{2^N}\mathop {\sum _{S_1,S_2\subset \Lambda _0}}_{(|S_1|+|S_2|=N)}|\psi _{S_1,S_2}|^2=\frac{1}{2^N}. \end{aligned}$$
(B.7)

Noting that the left-hand side is \(\langle \Psi |\hat{P}_1|\Psi \rangle \), we get (B.2). \(\square \)

This model considered here is generally non-integrable, and we expect that its energy eigenvalues are nondegenerate. It is desirable to find models in which the absence of degeneracy can be proved rigorously.

Unfortunately, the only case we can prove nondegeneracy is a trivial decoupled model with \(t_{x,y}=v_{x,y}=0\) for all \(x,y\in \Lambda _0\). We readily see that the energy eigenvalues are nondegenerate if \(s_x\), \(w_x\), and \(u_x\) with \(x\in \Lambda _0\) are chosen to be different from each other.Footnote 11 Therefore we can fully justify our main theorem, Theorem 2.4, for the model, but we should note that the result is trivial. In the initial state \(|\Phi (0)\rangle \), each pair of sites (x, 1) and (x, 2) is either empty or occupied by one partilce at (x, 1). The time evolution then takes place independently for each pair of sites. If there is a particle in a pair, then a superposition of two states with a particle at (x, 1) and at (x, 2) is generated. This, when viewed macroscopically, results in thermalization. We must say that there is nothing interesting in this observation.

1.2 B.2. Free Fermions With \(\mathbb {Z}_2\) Symmetry

Next, we discuss a class of free fermion models in which the bound (2.5) for the particle distribution, and hence Assumption  2.2 can be justified. We here follow the strategy outlined at the end of Sect. 3.3 and justify the inequality (3.34) for the fermion operators corresponding to single-particle energy eigenstates.

Let \(\Lambda \) be an arbitrary lattice, and consider the most general free fermion Hamiltonian

$$\begin{aligned} \hat{H}=\sum _{x,y\in \Lambda }t_{x,y}\,\hat{c}^\dagger _x\hat{c}_y, \end{aligned}$$
(B.8)

where the hopping amplitude satisfies \(t_{x,y}=(t_{y,x})^*\in \mathbb {C}\). Note that the diagonal element \(t_{x,x}\in \mathbb {R}\) represents the single-body potential.

We assume that the model has \(\mathbb {Z}_2\) symmetry in the sense that there is a one-to-one map \(p:\Lambda \rightarrow \Lambda \) such that \(p^2=\textrm{id}\), and that the Hamiltonian is invariant under the transformation p, i.e., \(t_{p(x),p(y)}=t_{x,y}\) for any \(x,y\in \Lambda \). We also assume that \(\Lambda \) is disjointly decomposed as \(\Lambda =\Lambda _1\cup \Lambda _2\) and that \(p(\Lambda _1)\subset \Lambda _2\).

As an example, consider the chain \(\Lambda =\{1,\ldots ,L\}\) with odd L, and let p be the inversion \(p(x)=L+1-x\). Then the decomposition with \(\Lambda _1=\{1,\ldots ,(L-1)/2\}\) and \(\Lambda _2=\{(L+1)/2,\ldots ,L\}\) satisfies the above property.

Let \(\varvec{\psi }=(\psi _x)_{x\in \Lambda }\) be a normalized single-particle energy eigenstate. To be precise, it satisfies the Schrödinger equation \(\epsilon \,\psi _x=\sum _{y\in \Lambda }t_{x,y}\psi _x\) for all \(x\in \Lambda \) with the (single-particle) energy eigenvalue \(\epsilon \). Let us further assume that the energy eigenvalue \(\epsilon \) is nondegenerate. Then, with respect to the symmetry transformation p, the corresponding wave function \(\varvec{\psi }\) is either symmetric, i.e., \(\psi _{p(x)}=\psi _x\) for all \(x\in \Lambda \), or antisymmetric, i.e., \(\psi _{p(x)}=-\psi _x\) for all \(x\in \Lambda \). We then see that

$$\begin{aligned} \sum _{x\in \Lambda _1}|\psi _x|^2=\frac{1}{2}\sum _{x\in \Lambda _1}\bigl (|\psi _x|^2+|\psi _{p(x)}|^2\bigr ) \le \frac{1}{2}\sum _{x\in \Lambda }|\psi _x|^2=\frac{1}{2}, \end{aligned}$$
(B.9)

where we noted that \(p(\Lambda _1)\subset \Lambda \backslash \Lambda _1\).

Let \(\hat{a}^\dagger _{\varvec{\psi }}=\sum _{x\in \Lambda }\psi _x\,\hat{c}^\dagger _x\) be the creation operator of the state \(\varvec{\psi }\). It can be decomposed as \(\hat{a}^\dagger _{\varvec{\psi }}=\hat{b}^\dagger _{1,\varvec{\psi }}+\hat{b}^\dagger _{2,\varvec{\psi }}\) with \(\hat{b}^\dagger _{1,\varvec{\psi }}=\sum _{x\in \Lambda _1}\psi _x\,\hat{c}^\dagger _x\) and \(\hat{b}^\dagger _{2,\varvec{\psi }}=\sum _{x\in \Lambda _2}\psi _x\,\hat{c}^\dagger _x\). This corresponds to the decomposition (3.30). We also see from (B.9) that \(\hat{b}^\dagger _{1,\varvec{\psi }}\) satisfies \(\Vert \hat{b}_{1,\varvec{\psi }}\,\hat{b}^\dagger _{1,\varvec{\psi }}\Vert =\sum _{x\in \Lambda _1}|\psi _x|^2\le 1/2\), which corresponds to the desired (3.34).

We now assume that single-particle energy eigenvalues \(\epsilon _1,\ldots ,\epsilon _{|\Lambda |}\) are all nondegenerate, and denote by \(\hat{a}^\dagger _j\) the creation operator of the single-particle energy eigenstate corresponding to \(\epsilon _j\). Then the foregoing discussion shows that each \(\hat{a}^\dagger _j\) is decomposed as (3.30), and the operator for the sublattice \(\Lambda _1\) satisfies the bound (3.34). Repeating the derivation in section 3.3, we see an N-body energy eigenstate of the form

$$\begin{aligned} |\Psi \rangle =\hat{a}^\dagger _{j_1}\cdots \hat{a}^\dagger _{j_N}|\Phi _\textrm{vac}\rangle , \end{aligned}$$
(B.10)

satisfies the desired bound (2.5).

Interestingly, it was only necessary to assume the nondegeneracy of single-particle energy eigenvalues to prove the desired bound (2.5) in this model. To ensure the presence of thermalization, we have to assume further that N-body energy eigenvalues are nondegenerate. It is rather likely that degeneracy is absent in a sufficiently complex free fermion model, but we do not know how to justify the claim. We also note that the p-symmetry may not be exact. It can be violated by a small perturbation as long as the bound (2.5) remains valid.

C: Effective Dimensions of Some Initial Particle Configurations in the Free Fermion Chain

In the main text, the initial state \(|\Phi (0)\rangle \) is drawn randomly from the small Hilbert space \(\mathcal{H}_1\). Conceptually speaking, it may be desirable to consider the time evolution starting from a non-random simple initial state. Here we again treat free fermion chains and examine the effective dimensions of some initial states in which particles have definite positions.

In Sect. C.2, we observe that the initial state where particles are arranged in a periodic manner has an effective dimension that is large but not large enough to guarantee thermalization. This observation suggests that a random initial state is mandatory in a free fermion model if we demand the effective dimension to be extremely large. This is very likely to be a common property for integrable models. In a non-integrable model, on the other hand, it is believed that even a regular initial state generally has an effective dimension almost as large as the total dimension.

In Sect. C.3, we consider an artificial class of initial configurations (Golomb ruler configurations) and show that the corresponding effective dimensions are almost as large as the total dimension. This leads to a statement about thermalization with a non-random initial state. In this class of models, however, the particle density inevitably tends to zero according to \(\rho \sim N^{-1}\) as the particle number grows.

1.1 C.1. General formula for \(D_\textrm{eff}\)

We consider the free fermion chain as defined in section 3. Let the initial particle configuration be \(\varvec{x}=(x_1,x_2,\ldots ,x_N)\) with \(x_j\in \{1,\ldots ,L\}\) such that \(x_j<x_{j+1}\) for \(j=1,\ldots ,N-1\), and define the corresponding N fermion state as

$$\begin{aligned} |\Phi _{\varvec{x}}\rangle =\hat{c}^\dagger _{x_1}\hat{c}^\dagger _{x_2}\cdots \hat{c}^\dagger _{x_N}|\Phi _\textrm{vac}\rangle . \end{aligned}$$
(C.1)

We set \(|\Phi _{\varvec{x}}\rangle \) as the initial state \(|\Phi (0)\rangle \). Then we see from (2.6) that the effective dimension is given by

$$\begin{aligned} D_\textrm{eff}^{-1}=\sum _{\varvec{k}\in \tilde{\mathcal{K}}_N}\bigl |\langle \Phi _{\varvec{x}}|\Psi _{\varvec{k}}\rangle \bigr |^4, \end{aligned}$$
(C.2)

where \(\tilde{\mathcal{K}}_N=\{(k_1,\ldots ,k_N)\,|\,k_j<k_{j+1}\}\subset \mathcal{K}^N\). (The k-space \(\mathcal{K}\) is defined in (3.2).) Noting that (3.3) implies \(\{\hat{c}_x,\hat{a}^\dagger _k\}=e^{ikx}/\sqrt{L}\), we see

$$\begin{aligned} \langle \Phi _{\varvec{x}}|\Psi _{\varvec{k}}\rangle =\langle \Phi _\textrm{vac}|\hat{c}_{x_N}\cdots \hat{c}_{x_1}\hat{a}^\dagger _{k_1}\cdots \hat{a}^\dagger _{k_N}|\Phi _\textrm{vac}\rangle =L^{-N/2}\sum _P(-1)^P\prod _{j=1}^Ne^{ik_jx_{P(j)}}, \end{aligned}$$
(C.3)

where the summation is over all possible N! permutations P of \(\{1,\ldots ,N\}\) and \((-1)^P\) is the signature of P. It is useful to regard \(\varvec{k}\) in the above expression as an element in \(\mathcal{K}^N\) rather than its physical subspace \(\tilde{\mathcal{K}}_N\). This replacement is justified since \(\bigl |\langle \Phi _{\varvec{x}}|\Psi _{\varvec{k}}\rangle \bigr |\) is invariant under any permutations of \(k_1,\ldots ,k_N\) and equals zero if \(k_j=k_{j'}\) for some \(j\ne j'\). We can thus rewrite (C.2) as

$$\begin{aligned} D_\textrm{eff}^{-1}=\frac{1}{N!}\sum _{\varvec{k}\in \mathcal{K}^N}\bigl |\langle \Phi _{\varvec{x}}|\Psi _{\varvec{k}}\rangle \bigr |^4. \end{aligned}$$
(C.4)

This rewriting is useful since one can now sum independently over \(k_1,\ldots ,k_N\in \mathcal{K}\).

From (C.3), we see that

$$\begin{aligned} \bigl |\langle \Phi _{\varvec{x}}|\Psi _{\varvec{k}}\rangle \bigr |^2&= \frac{1}{L^N}\sum _{P,Q}(-1)^{PQ}\prod _{j=1}^Ne^{ik_j(x_{P(j)}-x_{Q(j)})} \nonumber \\&=\frac{N!}{L^N} +\frac{1}{L^N}\mathop {\sum _{P,Q}}_{(P\ne Q)}(-1)^{PQ}\prod _{j=1}^Ne^{ik_j(x_{P(j)}-x_{Q(j)})}, \end{aligned}$$
(C.5)

and

$$\begin{aligned} \bigl |\langle \Phi _{\varvec{x}}|\Psi _{\varvec{k}}\rangle \bigr |^4=C_1+C_2(\varvec{k})+C_3(\varvec{k}), \end{aligned}$$
(C.6)

with

$$\begin{aligned} C_1=\Bigl (\frac{N!}{L^N}\Bigr )^2,\quad C_2(\varvec{k})=\frac{2N!}{L^{2N}}\mathop {\sum _{P,Q}}_{(P\ne Q)}(-1)^{PQ}\prod _{j=1}^Ne^{ik_j(x_{P(j)}-x_{Q(j)})},\end{aligned}$$
(C.7)
$$\begin{aligned} C_3(\varvec{k})=\frac{1}{L^{2N}}\mathop {\sum _{P,Q,P',Q'}}_{(P\ne Q,\,P'\ne Q')}(-1)^{PQP'Q'} \prod _{j=1}^Ne^{ik_j\{(x_{P(j)}-x_{Q(j)})-(x_{P'(j)}-x_{Q'(j)})\}}. \end{aligned}$$
(C.8)

We shall evaluate the sum (C.4) by using the decomposition (C.6). Clearly

$$\begin{aligned} \frac{1}{N!}\sum _{\varvec{k}\in \mathcal{K}^N}C_1=\frac{N!}{L^N}, \end{aligned}$$
(C.9)

The remaining sums are evaluated by using the standard formula

$$\begin{aligned} \sum _{k\in \mathcal{K}}e^{ikx}={\left\{ \begin{array}{ll} L,&{}x=0\mod L;\\ 0,&{}\text {otherwise}, \end{array}\right. } \end{aligned}$$
(C.10)

where \(x\in \mathbb {Z}\). Note that in the expression for \(C_2(\varvec{k})\) in (C.7), one has \(x_{P(j)}-x_{Q(j)}\ne 0\) for at least one j because \(P\ne Q\). We thus see

$$\begin{aligned} \frac{1}{N!}\sum _{\varvec{k}\in \mathcal{K}^N}C_2(\varvec{k})=0. \end{aligned}$$
(C.11)

The sum of \(C_3(\varvec{k})\) is generally nonzero and can be evaluated as

$$\begin{aligned} \frac{1}{N!}\sum _{\varvec{k}\in \mathcal{K}^N}C_3(\varvec{k})&= \frac{1}{L^NN!}\mathop {\sum _{P,Q,P',Q'}}_{(P\ne Q,\,P'\ne Q')}(-1)^{PQP'Q'}\nonumber \\&\quad \cdot \prod _{j=1}^N\chi [(x_{P(j)}-x_{Q(j)})-(x_{P'(j)}-x_{Q'(j)})=0\mod L], \end{aligned}$$
(C.12)

where the characteristic function is defined as \(\chi [\text {true}]=1\) and \(\chi [\text {false}]=0\). Let us write the right-hand side of (C.12) as \(\mathcal{S}_{\varvec{x}}/L^N\). From (C.4), (C.6), (C.9), (C.11), and (C.12), we see that the effective dimension of the initial state \(|\Phi (0)\rangle =|\Phi _{\varvec{x}}\rangle \) is given by

$$\begin{aligned} D_\textrm{eff}=\frac{L^N}{N!+\mathcal{S}_{\varvec{x}}}. \end{aligned}$$
(C.13)

Our main task is to evaluate the sum \(\mathcal{S}_{\varvec{x}}\) defined in (C.12) for a given particle configuration \(\varvec{x}\). For later convenience we sum over P in (C.12) (and write \(P^{-1}Q\), \(P^{-1}P'\), and \(P^{-1}Q'\) as Q, \(P'\), and \(Q'\), respectively) to rewrite the expression as

$$\begin{aligned} \mathcal{S}_{\varvec{x}}=\mathop {\sum _{Q,P',Q'}}_{(Q\ne \textrm{id},\,P'\ne Q')}(-1)^{QP'Q'} \prod _{j=1}^N\chi [x_{j}-x_{Q(j)}-x_{P'(j)}+x_{Q'(j)}=0\mod L]. \nonumber \\ \end{aligned}$$
(C.14)

1.2 C.2. \(D_\textrm{eff}\) in Periodic Configurations

First, we consider regular particle configurations with a period \(p=1,2,\ldots \). Fix p, and choose the chain length L and the particle number N such that \(L=pN\). We consider the initial particle distribution given by

$$\begin{aligned} x_j=pj, \end{aligned}$$
(C.15)

for \(j=1,\ldots ,N\).

Then (C.14) is computed as

$$\begin{aligned} \mathcal{S}_{\varvec{x}}&=\mathop {\sum _{Q,P',Q'}}_{(Q\ne \textrm{id},\,P'\ne Q')}(-1)^{QP'Q'} \prod _{j=1}^N\chi [pj-{pQ(j)}-{pP'(j)}+{pQ'(j)}=0\mod L] \nonumber \\&=\mathop {\sum _{Q,P',Q'}}_{(Q\ne \textrm{id},\,P'\ne Q')}(-1)^{QP'Q'} \prod _{j=1}^N\chi [j-{Q(j)}-{P'(j)}+{Q'(j)}=0\mod N], \end{aligned}$$
(C.16)

which depends only on N and is independent of L and p. Thus, we can evaluate the above sum by employing a useful choice of p. Fortunately, this sum becomes trivial for \(p=1\), and therefore we compute the sum in the case of \(p=1\). A fermion system with \(L=N=1\), which is fully filled, has one-dimensional Hilbert space and hence \(D_\textrm{eff}=1\). We see from (C.13) that \(\mathcal{S}_{\varvec{x}}=N^N-N!\). Recalling the L independence of \(\mathcal{S}_{\varvec{x}}\), we get a remarkably simple result

$$\begin{aligned} D_\textrm{eff}=\Bigl (\frac{L}{N}\Bigr )^N=e^{-(\rho \log \rho )L}, \end{aligned}$$
(C.17)

for any L and N (such that \(L=pN\)), where \(\rho =1/p\) is the particle density. We thus see that the effective dimension grows exponentially with the system size L, as expected. But it turns out that it is not large enough. Combining (C.17) with (2.1), we see

$$\begin{aligned} \frac{D_\textrm{tot}}{D_\textrm{eff}}\sim e^{\{-(1-\rho )\log (1-\rho )\}L}=e^{\{\rho +O(\rho ^2)\}L}, \end{aligned}$$
(C.18)

and hence \(D_\textrm{eff}\) is considerably smaller compared with the total dimension \(D_\textrm{tot}\). This conclusion is consistent with the numerical result in [46]. We conclude that our strategy of the proof of Theorem 2.4 is ineffective for this initial state. Interestingly, it was found numerically in [45, 46] that the free fermion chain with similar initial states exhibits thermalization in some sense.

1.3 C.3. \(D_\textrm{eff}\) in Golomb-Ruler Configurations

We next discuss a class of particle configurations for which the effective dimension \(D_\textrm{eff}\) is easily evaluated and turns out to be almost as large as the total dimension \(D_\textrm{tot}\). In these settings, however, the particle density inevitablyapproaches zero as N gets larger.

A sequence of natural numbers \(\varvec{x}=(x_1,\ldots ,x_N)\) is called a Golomb ruler [59] if for any \(j\ne k\), one has \(x_j-x_k=x_\ell -x_m\) only when \(j=\ell \) and \(k=m\). The periodic boundary counterpart (in which one replaces the condition \(x_j-x_k=x_\ell -x_m\) by \(x_j-x_k=x_\ell -x_m \mod L\)) is called a modular Golomb ruler. The optimal (minimum) system size L of a modular Golomb ruler for given N is \(L=N(N-1)\), since \(x_j-x_k\) takes \(N(N-1)\) distinct positive integers. The optimal configuration, if exists, is called a perfect difference set. Interestingly, perfect difference sets are proven to exist if \(N-1\) is a prime power \(p^n\) [60].

We set the configuration of N particles as a modular Golomb ruler \(\varvec{x}=(x_1,\ldots ,x_N)\) (\(x_1<x_2\cdots <x_N\)). By taking \(x_1=1\) and choosing the system size L as a prime such that \(L\ge 2x_N-1\), we see that for any \(j\ne k\), one has \(x_j-x_k=x_\ell -x_m\mod L\) only when \(j=\ell \) and \(k=m\) (i.e., a modular Golomb ruler). A nontrivial and asymptotically optimal exampleFootnote 12 of a Golomb ruler can be found in [61], where the following sequence

$$\begin{aligned} x_j=1+2N(j-1)+\{(j-1)^2\mod N\}, \end{aligned}$$
(C.19)

for \(j=1,\ldots ,N\) with a prime \(N>2\) is shown to be a Golomb ruler. Since \(1+2N(N-1)\le x_N\le 1+2N(N-1)+N-1\), the aforementioned construction leads to the chain length as \(L\simeq 4N^2\) with the particle density \(\rho \simeq (4N)^{-1}\). Note that the optimal (densest) Golomb ruler has density \(\rho =O(N^{-1})=O(L^{-1/2})\), and thus the above construction is asymptotically optimal.

We shall fix an arbitrary initial particle configuration \(\varvec{x}\) that forms a modular Golomb ruler and evaluate its effective dimension. We first bound the sign factor in (C.14) as \((-1)^{QP'Q'}\le 1\) to get

$$\begin{aligned} \mathcal{S}_{\varvec{x}}\le \mathop {\sum _{Q,P',Q'}}_{(Q\ne \textrm{id},\,P'\ne Q')} \prod _{j=1}^N\chi [x_{j}-x_{Q(j)}-x_{P'(j)}+x_{Q'(j)}=0\mod L]. \end{aligned}$$
(C.20)

In fact, it can be shown that this is an equalityFootnote 13, but the upper bound is enough for our purpose.

Let us fix a permutation \(Q\ne \textrm{id}\), and examine the conditions for \(\prod _{j=1}^N\chi [\cdots ]=1\), i.e., \(x_{j}-x_{Q(j)}-x_{P'(j)}+x_{Q'(j)}=0\mod L\) for all \(j=1,\ldots ,N\). If j is such that \(Q(j)\ne j\), the condition for \(\varvec{x}\) implies \(P'(j)=j\) and \(Q'(j)=Q(j)\). We see there is no choice for \(P'(j)\) and \(Q'(j)\). If \(j=Q(j)\), on the other hand, the only requirement is \(P'(j)=Q'(j)\). There is some freedom for choosing \(P'(j)\) and \(Q'(j)\).

Let \(n_Q\) be the number of j such that \(Q(j)=j\). Since \(Q\ne \textrm{id}\), we see \(n_Q=0,1,\ldots ,N-2\). From the above consideration, we see that there are \(n_Q!\) choices for \(P'\) (and thus \(Q'\)) for fixed Q. We thus find

$$\begin{aligned} (\text {RHS of } (C.20))=\mathop {\sum _{Q}}_{(Q\ne \textrm{id})}n_Q!=\sum _{n=0}^{N-2}n!\,\mathcal{N}(n), \end{aligned}$$
(C.21)

where \(\mathcal{N}(n)\) is the number of \(Q\ne \textrm{id}\) such that \(n_Q=n\). The value of \(\mathcal{N}(n)\) is computed explicitly as

$$\begin{aligned} \mathcal{N}(n)=\left( {\begin{array}{c}N\\ n\end{array}}\right) d_{N-n}, \end{aligned}$$
(C.22)

where \(d_m\) is the m-th de Montmort number (also known as the m-th derangement number or the subfactorial of m). The de Montmort number counts the number of derangementFootnote 14 on n elements. Fortunately, the de Montmort number \(d_m\) is explicitly computed as

$$\begin{aligned} d_m=\left\lfloor \frac{m!+1}{e}\right\rfloor \end{aligned}$$
(C.23)

with the floor functionFootnote 15\(\lfloor \cdot \rfloor \) [62]. This expression, with (C.20) and (C.21), leads to a simple upper bound

$$\begin{aligned} \mathcal{S}_x\le \sum _{n=0}^{N-2} \frac{N!}{e}\biggl (1+\frac{1}{(N-n)!}\biggr ) =\frac{N!}{e}\biggl (N-1+\sum _{m=2}^N\frac{1}{m!}\biggr ) \le \frac{N!}{e}(N-3+e). \end{aligned}$$
(C.24)

Substituting this into (C.13), we can bound the effective dimension from below as

$$\begin{aligned} D_\textrm{eff}\ge \frac{eL^N}{(N+2e-3)N!}. \end{aligned}$$
(C.25)

Thus the ratio between the total dimension and the effective dimension is bounded as

$$\begin{aligned} \frac{D_\textrm{tot}}{D_\textrm{eff}}\le \frac{(N+2e-3)\,L!}{e(L-N)!\,L^N}\le \frac{N+2e-3}{e}. \end{aligned}$$
(C.26)

Note that \(D_\textrm{tot}=\left( {\begin{array}{c}L\\ N\end{array}}\right) \) is approximated by \((L/N)^N\) when \(N\ll L\), and hence grows super-exponentially in N. (If we take the initial configuration (C.19) then \(D_\textrm{tot}\sim (4N)^N\).) This means that \(D_\textrm{eff}\) satisfying (C.26) is extremely close to \(D_\textrm{tot}\).

As we have stressed, such a large effective dimension is expected in a non-integrable quantum many-body system, but not in an integrable system. Here we have a large \(D_\textrm{eff}\) in a free fermion model because of the artificial Golomb-ruler configuration. But recall that this choice is possible only in the extremely low density \(\rho =O(N^{-1})\).

The above observation about the large effective dimension leads to a statement about thermalization. Take a sufficiently large and arbitrary prime N and a prime L such that \(L\ge 2x_N-1\) with \(x_N\) given by (C.19). We consider the system of N fermions on the chain \(\{1,\ldots ,L\}\) with the Hamiltonian (3.1). We take the phase factor \(\theta \) for which the energy eigenvalues (3.7) are nondegenerate. (See Lemma 3.1.)

For simplicity we restrict our observable only to the particle number in the left half of the chain, i.e.,

$$\begin{aligned} \hat{N}_\textrm{left}=\sum _{j=1}^{(L+1)/2}\hat{n}_j. \end{aligned}$$
(C.27)

The equilibrium value is of course \(\langle \hat{N}_\textrm{left}\rangle _\infty =N/2\). Let the initial state be \(|\Phi (0)\rangle =|\Phi _{\varvec{x}}\rangle =\hat{c}^\dagger _{x_1}\cdots \hat{c}^\dagger _{x_N}|\Phi _\textrm{vac}\rangle \), where the configuration \(x_1,\ldots ,x_N\) is given by (C.19). Since

$$\begin{aligned} \frac{\hat{N}_\textrm{left}}{N}|\Phi (0)\rangle =|\Phi (0)\rangle , \end{aligned}$$
(C.28)

the initial state is highly nonequilibrium with respect to the quantity \(\hat{N}_\textrm{left}/N\).

Then by using the large deviation type estimate

$$\begin{aligned} \biggl \langle \hat{P}\biggl [ \Bigl |\frac{\hat{N}_\textrm{left}}{N}-\frac{1}{2}\Bigr |\ge \varepsilon \biggr ]\biggr \rangle _\infty \le e^{-(4\varepsilon ^2/3)N}, \end{aligned}$$
(C.29)

which follows from (2.22), and the standard argument (as in the proof of Theorem 2.4), we can prove the following.

Theorem C.1

For any \(\varepsilon >0\), there exists a constant \(T>0\) and a subset (a collection of intervals) \(G\subset [0,T]\) with

$$\begin{aligned} \frac{\mu (G)}{T}\ge 1-e^{-(\varepsilon ^2/4)N}, \end{aligned}$$
(C.30)

where \(\mu (G)\) is the total length of the intervals in G. Suppose that one performs a measurement of the number operator \(\hat{N}_\textrm{left}\) in the state \(|\Phi (t)\rangle \) with arbitrary \(t\in G\). Then, with probability larger than \(1-e^{-(\varepsilon ^2/4)N}\), the measurement result \(N_\textrm{left}\) satisfies

$$\begin{aligned} \Bigl |\frac{N_\textrm{left}}{N}-\frac{1}{2}\Bigr |\le \varepsilon , \end{aligned}$$
(C.31)

i.e., the state is found in thermal equilibrium.

Thus thermalization starting from a deterministic initial state has been established without any unproven assumptions. Here one can choose the precision \(\varepsilon >0\) arbitrarily. But in order to make the factor \(e^{-(\varepsilon ^2/4)N}\) negligibly small, one must take N large enough, which means that the density becomes lower.

D: Possible Extension to the Finite Temperature Situation

Throughout the present paper, we only focused on situations where the initial and the final states correspond to infinite temperature thermal states. See, in particular, Sect. 2.4. We believe that our results can be extended to finite temperature settings with extra technical efforts. Although we do not elaborate on the extension, we here briefly discuss the setting and essential steps in the proof.

We consider the free fermion Hamiltonian (1.1), (3.1). Decompose the Hamiltonian as in (2.30), where we choose \(\Lambda _1\) as the half-chain \(\{1,\ldots ,(L-1)/2\}\). It follows that \(\Vert {\varDelta }\hat{H}\Vert =h_0\) is independent of the system size. Denote by \(|\tilde{\Psi }_j\rangle \in \mathcal{H}_1\) the normalized eigenstate of \(\hat{H}_1\) with eigenvalue \(\tilde{E}_j\). For the energy density \(u\in (-2,0)\) and the energy width \({\varDelta }u>0\), we define the nonequilibrium microcanonical energy shell by

$$\begin{aligned} \mathcal{H}_1^u=\textrm{span}\bigl \{|\tilde{\Psi }_j\rangle \,\Bigl |\, \bigl |\tfrac{\tilde{E}_j}{N}-u\bigr |\le {\varDelta }u\bigr \}\subset \mathcal{H}_1. \end{aligned}$$
(D.1)

Noting that \(\hat{H}_2|\tilde{\Psi }_j\rangle =0\), we observe

$$\begin{aligned} \langle \tilde{\Psi }_j|(\hat{H}-\tilde{E}_j)^2|\tilde{\Psi }_j\rangle =\langle \tilde{\Psi }_j|\{(\hat{H}_1-\tilde{E}_j)+{\varDelta }\hat{H}\}^2|\tilde{\Psi }_j\rangle =\langle \tilde{\Psi }_j|\{({\varDelta }\hat{H})^2|\tilde{\Psi }_j\rangle \le (h_0)^2,\nonumber \\ \end{aligned}$$
(D.2)

which implies that \(|\tilde{\Psi }_j\rangle \) is (with a minor error when N is large) a superposition of \(|\Psi _k\rangle \) such that \(|E_k-\tilde{E}_j|\lesssim h_0\). We thus find that any state \(|\Phi (0)\rangle \in \mathcal{H}_1^u\) and its time-evolution \(|\Phi (t)\rangle =e^{-i\hat{H}t}|\Phi (0)\rangle \) belongs (again, with minor errors when N is large) to the standard microcanonical energy shell

$$\begin{aligned} \mathcal{H}_\textrm{tot}^u=\textrm{span}\bigl \{|\Psi _j\rangle \,\Bigl |\, \bigl |\tfrac{E_j}{N}-u\bigr |\le {\varDelta }u'\bigr \}\subset \mathcal{H}_\textrm{tot}, \end{aligned}$$
(D.3)

with \({\varDelta }u'>{\varDelta }u\).

In the finite temperature setting, we choose initial state \(|\Phi (0)\rangle \) randomly and uniformly from the nonequilibrium energy shell \(\mathcal{H}^u_1\). The goal is to show that Theorem 2.4 (with suitable modifications of constants) is valid for the time-evolved state \(|\Phi (0)\rangle \).

Recalling that \(|\Phi (0)\rangle \) (essentially) belongs to \(\mathcal{H}^u_\textrm{tot}\), our strategy for the proof will be to properly replace \(\mathcal{H}_1\) and \(\mathcal{H}_\textrm{tot}\) in the original proof with \(\mathcal{H}^u_1\) and \(\mathcal{H}^u_\textrm{tot}\), respectively. Let us see how the proof of the most important estimate of the effective dimension, Theorem 2.3, is modified. Interestingly, a small modification is sufficient. Denoting by \(\hat{P}^u_1\) the projection onto \(\mathcal{H}^u_1\), and by \(D_1^u\) the dimension of \(\mathcal{H}^u_1\), we find

$$\begin{aligned} \overline{D_\textrm{eff}^{-1}}&=\sum _{j=1}^{D_\textrm{tot}}\overline{\bigl |\langle \Phi (0)|\hat{P}^u_1|\Psi _j\rangle \bigr |^4} =\frac{2}{D^u_1(D^u_1+1)}\sum _{j=1}^{D_\textrm{tot}}\Vert \hat{P}^u_1|\Psi _j\rangle \Vert ^4 \nonumber \\&\le \frac{2}{D^u_1(D^u_1+1)}\sum _{j=1}^{D_\textrm{tot}}\Vert \hat{P}_1|\Psi _j\rangle \Vert ^2\,\Vert \hat{P}^u_1|\Psi _j\rangle \Vert ^2 \nonumber \\&\le \frac{2}{D^u_1(D^u_1+1)2^N}{\text {Tr}}[\hat{P}^u_1]=\frac{2}{(D^u_1+1)2^N}, \end{aligned}$$
(D.4)

which is a faithful extension of the key inequality (2.10). The analog of Theorem 2.3 is proved if we properly estimate the ratio \(D^u_\textrm{tot}/D^u_1\). Another nontrivial (but technical) step for the proof of the desired extension of Theorem 2.4 is the derivation of the large-deviation upper bound (2.22) for the microcanonical average.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shiraishi, N., Tasaki, H. Nature Abhors a Vacuum: A Simple Rigorous Example of Thermalization in an Isolated Macroscopic Quantum System. J Stat Phys 191, 82 (2024). https://doi.org/10.1007/s10955-024-03289-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-024-03289-6

Keywords

Navigation