Skip to main content
Log in

One-dimensional Discrete Dirac Operators in a Decaying Random Potential I: Spectrum and Dynamics

  • Published:
Mathematical Physics, Analysis and Geometry Aims and scope Submit manuscript

Abstract

We study the spectrum and dynamics of a one-dimensional discrete Dirac operator in a random potential obtained by damping an i.i.d. environment with an envelope of type nα for α > 0. We recover all the spectral regimes previously obtained for the analogue Anderson model in a random decaying potential, namely: absolutely continuous spectrum in the super-critical region \(\alpha >\frac 12\); a transition from pure point to singular continuous spectrum in the critical region \(\alpha =\frac 12\); and pure point spectrum in the sub-critical region \(\alpha <\frac 12\). From the dynamical point of view, delocalization in the super-critical region follows from the RAGE theorem. In the critical region, we exhibit a simple argument based on lower bounds on eigenfunctions showing that no dynamical localization can occur even in the presence of point spectrum. Finally, we show dynamical localization in the sub-critical region by means of the fractional moments method and provide control on the eigenfunctions.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Andrei, E.Y., Du, X., Duerr, F., Lucian, A., Skachko, I.: Fractional quantum Hall effect and insulating phase of Dirac electrons in graphene. Nature 462, 192–5 (2009)

    ADS  Google Scholar 

  2. Aizenman, M., Molchanov, S.: Localization at large disorder and at extreme energies: an elementary derivation. Comm. Math. Phy. 157, 245–278 (1993)

    ADS  MathSciNet  MATH  Google Scholar 

  3. Aizenman, M., Sims, R., Warzel, S.: stability of the absolutely continuous spectrum of random schrödinger operators on tree graphs, Probab. Theory Rel. Fields 136, 363–394 (2006)

    MATH  Google Scholar 

  4. Aizenman, M., Warzel, S.: Random Operators: Disordered effects on Quantum spectra and dynamics, Graduate Studies in Mathematics 168 AMS (2016)

  5. Azuma, K.: Weighted sums of certain dependent random variables. Tô,hoku Math. J. 19(3), 357–367 (1967)

    MathSciNet  MATH  Google Scholar 

  6. Barbaroux, J.-M., Cornean, H., Zalczer, S.: Localization for gapped Dirac Hamiltonians with random pertubations: Application to graphene antidot lattices, arXiv:1812.01868

  7. Bissbort, U., Esslinger, T., Greif, D., Hofstetter, W., Jotzu, G., Messer, N., Uehlinger, T.: Artificial graphene with tunable interactions. Phys. Rev. Lett. 111, 185307 (2013)

    ADS  Google Scholar 

  8. Bourgain, J.: On random schrödinger operators on \(\mathbb {Z}^{2}\). Discret Contin. Dyn. Syst. 8, 1–15 (2002)

    Google Scholar 

  9. Bourgain, J.: Random Lattice Schrödinger Operators with Decaying Potential: Some Higher Dimensional Phenomena, Geometric Aspects of Functional Analysis, Lectures Notes in Math., vol. 1807, pp 70–98. Springer, Berlin-Heidelberg (2003)

    Google Scholar 

  10. Bourget, O., Moreno Flores, G.R., Taarabt, A.: Dynamical localization for the one-dimensional continuum Anderson model in a decaying random potential, preprint

  11. Bucaj, V.: On the Kunz-Souillard approach to localization for the discrete one dimensional generalized Anderson model, preprint.

  12. Bucaj, V: The Kunz-Souillard approach to localization for jacobi operators. Oper. Matrices. 12(4), 1099–1127 (2018)

    MathSciNet  MATH  Google Scholar 

  13. Bucaj, V., Damanik, D., Fillman, J., Gerbuz, V., VandenBoom, T., Wang, F., Zhang, Z.: Localization for the one-dimensional Anderson model via positivity and large deviations for the Lyapunov exponent. Trans. Amer. Math. Soc. 372, 3619–3667 (2019)

    MathSciNet  MATH  Google Scholar 

  14. Bolotin, K.I., Ghahari, F., Kim, P., Shulman, M.D., Stormer, H.L.: Observation of the fractional quantum Hall effect in graphene. Nature 462, 196–9 (2009)

    ADS  Google Scholar 

  15. Basu, C., Macía, E., Domínguez-Adame, F., Roy, C.L., Sánchez, A.: Localization of relativistic electrons in a One-Dimensional disordered system. J. Phys. A 27, 3285–3291 (1994)

    ADS  Google Scholar 

  16. Bolotin, K.I., Jiang, Z., Sikes, K.J., et al.: Ultrahigh electron mobility in suspended graphene. Solid State Commun. 146, 351–5 (2008)

    ADS  Google Scholar 

  17. Carmona, R.: Exponential localization in one dimensional disordered systems, Duke. Math. J. 49, 191–213 (1982)

    MathSciNet  MATH  Google Scholar 

  18. Carmona, R., Klein, A., Martinelli, F.: Anderson localization for bernoulli and other singular potentials, commun. Math. Phys. 108, 41–66 (1987)

    ADS  MathSciNet  MATH  Google Scholar 

  19. Carvalho, S., de Oliveira, C., Prado, R.: Sparse one-dimensional discrete Dirac operators II: Spectral properties. J. Math. Phys 073501, 52 (2011)

    MathSciNet  MATH  Google Scholar 

  20. Carvalho, S., de Oliveira, C., Prado, R.: Dynamical localization for discrete anderson dirac operators. J. Stat. Phys. 167(2), 260–296 (2017)

    ADS  MathSciNet  MATH  Google Scholar 

  21. Comets, F., Yoshida, N.: Branching random walks in Space–Time random environment: Survival probability, global and local growth rates. J. Theor. Prob. 24, 657–687 (2011)

    MathSciNet  MATH  Google Scholar 

  22. Cycon, H.L., Froese, R.G., Kirsch, W., Simon, B.: Schrödinger Operators with Applications to Quantum Mechanics and Global Geometry, Texts and Monographs in Physics, Springer Study Edition. Springer-Verlag, Berlin (1987)

    MATH  Google Scholar 

  23. Castro Neto, A.H., Guinea, F., Peres, N.M.R., Novoselov, K.S., Geim, A.K.: The electronic properties of graphene. Rev. Modern Phys. 81, 109–162 (2009)

    ADS  Google Scholar 

  24. Damanik, D., Gorodetski, A: An extension of the Kunz-Souillard approach to localization in one dimension and applications to almost-periodic Schrödinger operators, Adv. Math (2016)

  25. De Bièvre, S., Germinet, F.: dynamical Localization for the Random Dimer schrödinger Operator. J. Stat. Phys. 98(5-6), 1134–1148 (2000)

    MATH  Google Scholar 

  26. Delyon, F.: appearance of a purely singular continuous spectrum in a class of random schrödinger operators. J. Statist. Phys. 40, 621–630 (1985)

    ADS  MathSciNet  MATH  Google Scholar 

  27. Delyon, F., Simon, B., Souillard, B.: From power pure point to continuous spectrum in disordered systems. Ann. Henri Poincaré 42(6), 283–309 (1985)

    MathSciNet  MATH  Google Scholar 

  28. Del Rio, R., Jitomirskaya, S., Last, Y., Simon, B.: What is localization?. Phys. Rev. Lett. 75, 117–119 (1995)

    ADS  Google Scholar 

  29. Del Rio, R., Jitomirskaya, S., Last, Y., Simon, B.: Operators with singular continuous spectrum IV: Hausdorff dimensions, rank one pertubations and localization. J. Anal. Math. 69, 153–200 (1996)

    MathSciNet  MATH  Google Scholar 

  30. de Oliveira, C., Prado, R.: Dynamical delocalization for the 1D Bernoulli discrete Dirac operator. J. Phys. A 38, 115–119 (2005)

    MathSciNet  MATH  Google Scholar 

  31. de Oliveira, C., Prado, R.: Spectral and localization properties for the one-dimensional Bernoulli discrete Dirac operator. J. Math. Phys. 072105, 46 (2005)

    MathSciNet  MATH  Google Scholar 

  32. de Oliveira, C., Prado, R.: Dynamical lower bounds for 1D Dirac operators. Math. Z. 259(1), 45–60 (2008)

    MathSciNet  MATH  Google Scholar 

  33. de Oliveira, C., Prado, R.: Sparse 1D discrete Dirac operators I: Quantum transport. J. Math. Anal. Appl. 385, 947–960 (2012)

    MathSciNet  MATH  Google Scholar 

  34. Durrett, R.: Probability: Theory and Examples, Cambridge Series in Statistical and Probabilistic Mathematics, Fourth Edition. Cambridge University Press, New York (2010)

    Google Scholar 

  35. Dean, C.R., Wang, L., Maher, P., et al.: Hofstadter’s butterfly and the fractal quantum Hall effect in moire superlattices. Nature 497, 598–602 (2013)

    ADS  Google Scholar 

  36. Figotin, A., Germinet, F., Klein, A., Müller, P.: persistence of Anderson localization in schrödinger operators with decaying random potentials. Ark. Mat. 45, 15–30 (2007)

    MathSciNet  MATH  Google Scholar 

  37. Froese, R., Hasler, D., Spitzer, W.: Absolutely continuous spectrum for the Anderson model on a tree: a geometric proof of Klein’s theorem. Comm. Math. Phys. 269, 239–257 (2007)

    ADS  MathSciNet  MATH  Google Scholar 

  38. Germinet, F., Klein, A.: Bootstrap multiscale analysis localization in random media. Commun. Math Phys. 222, 415–448 (2001)

    ADS  MathSciNet  MATH  Google Scholar 

  39. Germinet, F., Kiselev, A., Tcheremchantsev, S.: transfer matrices and transport for schrödinger operators. Ann. Inst. Fourier 54, 787–830 (2004)

    MathSciNet  MATH  Google Scholar 

  40. Germinet, F., Taarabt, A.: spectral properties of dynamical localization for schrödinger operators. Rev. Math. Phys. 25, 9 (2013)

    MATH  Google Scholar 

  41. Novoselov, KS, Geim, AK, Morozov, SV, et al.: Two-dimensional gas of massless Dirac fermions in graphene. Nature 438, 197–200 (2005)

    ADS  Google Scholar 

  42. Goldsheid, I., Molchanov, S., Pastur, L.: a pure point spectrum of the stochastic one-dimensional schrödinger equation. Funct. Anal. Appl. 11, 1–10 (1977)

    Google Scholar 

  43. Golénia, S., Haugomat, T.: On the a.c. spectrum of 1D discrete Dirac operator. Methods Funct. Anal. Topology 20(3), 252–273 (2014)

    MathSciNet  MATH  Google Scholar 

  44. Hamza, E., Stolz, G.: Lyapunov exponents for unitary Anderson models. J. Math. Phys. 043301, 48 (2008)

    MATH  Google Scholar 

  45. Hunt, B., Sanchez-Yamagishi, J.D., Young, A.F., et al.: Massive Dirac Fermions and Hofstadter Butterfly in a van der Waals Heterostructure. Science 340, 1427–30 (2013)

    ADS  Google Scholar 

  46. Jitomirskaya, S., Zhu, X.: Large deviations of the Lyapunov exponent and localization for the 1D Anderson model. Comm. Math. Phys. 370(1), 311–324 (2019)

    ADS  MathSciNet  MATH  Google Scholar 

  47. Klein, A.: Extended states in the Anderson model on the Bethe lattice. Adv. Math. 133, 163–184 (1998)

    MathSciNet  MATH  Google Scholar 

  48. Kiselev, A., Last, Y., Simon, B.: modified prüfer and EFGP transforms and the spectral analysis of one-dimensional schrödinger operators. Comm. Math. Phys. 194, 1–45 (1998)

    ADS  MathSciNet  MATH  Google Scholar 

  49. Katsnelson, M.I., Novoselov, K.S., Geim, A.K.: Chiral tunnelling and the Klein paradox in graphene. Nat. Phys. 2(9), 620–625 (2006)

    Google Scholar 

  50. Kiselev, A., Remling, C., Simon, B.: effective perturbation methods for one-dimensional schrödinger operators. J. Diff. Equ. 151, 290–312 (1999)

    ADS  MATH  Google Scholar 

  51. Krishna, M.: Anderson model with decaying randomness: existence of extended states. Proc. Indian Acad. Sci. (Math. Sci.) 100, 285–294 (1990)

    MathSciNet  MATH  Google Scholar 

  52. Kunz, H., Souillard, B.: sur le spectre des opérateurs aux différences finies aléatoires. Comm. Math. Phys. 78, 201–246 (1980)

    ADS  MathSciNet  MATH  Google Scholar 

  53. Last, Y., Simon, B.: eigenfunctions, transfer matrices, and absolutely continuous spectrum of one-dimensional schrödinger operators. Invent. Math. 135, 329 (1999)

    ADS  MathSciNet  MATH  Google Scholar 

  54. Novoselov, K.S., Geim, A.K., Morozov, S.V., Jiang, D., Zhang, Y., Dubonos, S.V., Grigorieva, I.V., Firsov, A.A.: Electric field effect in atomically thin carbon films. Science 306(5696), 666–669 (2004)

    ADS  Google Scholar 

  55. Rahu, S., Haldane, F.D.M.: Analogs of quantum-Hall-effect edge states in photonic crystals. Phys. Rev. A 78, 033834 (2008)

    ADS  Google Scholar 

  56. Rahu, S., Haldane, F.D.M.: Possible realization of directional optical waveguides in photonic crystals with broken Time-Reversal symmetry. Phys. Rev. Lett. 100, 013904 (2008)

    ADS  Google Scholar 

  57. Roy, C.L., Basu, C.: Relativistic study of electrical conduction in disordered systems. Phys. Rev. B 45, 14293–14301 (1992)

    ADS  Google Scholar 

  58. Sarma, S.D., Adam, S., Hwang, E.H., Rossi, E.: Electronic transport in two-dimensional graphene. Rev. Mod. Phys. 83, 407 (2011)

    ADS  Google Scholar 

  59. Simon, B.: Some Jacobi matrices with decaying potential and dense point spectrum. Comm. Math. Phys. 87, 253–258 (1982)

    ADS  MathSciNet  MATH  Google Scholar 

  60. Simon, B.: Spectral Analysis of rank one perturbations and applications, CRM Lectures Notes Vol. 8, Amer. Math. Soc, Providence, RI (1995)

  61. Zhang, Y., Tan, Y.W., Stormer, H.L., Kim, P.: Experimental observation of the quantum Hall effect and Berry’s phase in graphene. Nature 438, 201–4 (2005)

    ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gregorio R. Moreno Flores.

Additional information

Publisher’s Note

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

Olivier Bourget Partially supported by Fondecyt grant 1161732

Gregorio R. Moreno Flores Partially supported by Fondecyt grant 1171257, Núcleo Milenio ‘Modelos Estocásticos de Sistemas Complejos y Desordenados’ and MATH Amsud ‘Random Structures and Processes in Statistical Mechanics’

Amal Taarabt Partially supported by Fondecyt grant 11190084

Appendix A: Some Technical Estimates

Appendix A: Some Technical Estimates

1.1 A.1. Unimodular Matrices

The following lemmas correspond to [48, Lemma 2.2 and 8.7]. The first one allows us to establish the upper bound in Lemma 3.2. The proof of Proposition 6.2 is given after the second one. At the end of the section, we state [53, Theorem 8.3] which is used to prove pure point spectrum in the sub-critical regime.

Lemma A.1

Let A be an unimodular matrix and let \(\hat \theta = ({\cos \limits } \theta , {\sin \limits } \theta )\). Then, for all pair of angles \(|\theta _{1}-\theta _{2}|\leqslant \frac {\pi }{2}\),

$$ \begin{array}{@{}rcl@{}} \| A \| \leqslant \sin \left( \tfrac{|\theta_{1}-\theta_{2}|}{2}\right)^{-1} \max \{ \| A \hat \theta_{1}\|, \| A\hat \theta_{2}\| \}. \end{array} $$

Proof

See [48, Lemma 2.2]. □

The following lemma is used to find eigenfunctions with the proper decay and is the key to Proposition 6.2.

Lemma A.2

For a unimodular matrix with ∥A∥ > 1, define 𝜗 = 𝜗(A) as the unique angle \(\vartheta \in (-\frac {\pi }{2},\frac {\pi }{2}]\) such that \(\| A \hat \vartheta \| = \| A \|^{-1}\). We also define \(r(A)=\left \lVert A\left (\begin {array}{l}1\\0 \end {array}\right ) \right \rVert .\left \lVert A\left (\begin {array}{l}0\\1 \end {array}\right ) \right \rVert ^{-1} \).

Let (An)n be a sequence of unimodular matrices with ∥An∥ > 1 and write 𝜗n = 𝜗(An) and rn = r(An). Assume that

  1. (i)

    \(\underset {n\to \infty }{\lim } \| A_{n}\| = \infty \),

  2. (ii)

    \( \underset {n\to \infty }{\lim } \frac {\| A_{n+1}A_{n}^{-1}\|}{\| A_{n} \| \| A_{n+1}\|}=0\).

Then,

  1. (1)

    (𝜗n)n has a limit \(\vartheta _{\infty }\in (-\pi /2,\pi /2)\) if and only if (rn)n has a limit \(r_{\infty }\in [0,\infty )\). If 𝜗n →±π/2, then \(r_{n}\to \infty \) but, if \(r_{n}\to \infty \), we can only conclude that |𝜗n|→ π/2.

  2. (2)

    Suppose (𝜗n)n has a limit \(\vartheta _{\infty }\neq 0, \frac {\pi }{2}\). Then,

    $$ \underset{n\to\infty}{\lim} \frac{\log \| A_{n} \hat\vartheta_{\infty}\|}{\log \| A_{n} \|} = -1 \quad \text{if and only if} \quad \limsup_{n} \frac{\log |r_{n} - r_{\infty}|}{\log \| A_{n} \|} \leqslant -2. $$
    (A.1)

Proof

See [48, Lemma 8.7]. □

We apply this with An = Tω,n in order to prove Proposition 6.2. For positive sequences (bn)n and (cn)n, we write \(b_{n} \simeq c_{n}\) if \(\underset {n\to \infty }{\lim } \frac {b_{n}}{c_{n}}=1\) and denote \(b_{n} \lesssim c_{n}\) if there exists a constant K > 0 such that \(b_{n} \leqslant K c_{n}\) for n large enough, and bncn if \(b_{n} \lesssim c_{n} \lesssim b_{n}\).

Proof of Proposition 6.2

Define

$$ {\Phi}^{(1)}_{n} = \left( \begin{array}{l} \phi^{(1)}_{+,n} \\ \phi^{(1)}_{-,n} \end{array}\right) = \textbf{T}_{\omega,n-1} \left( \begin{array}{l} 1 \\ 0 \end{array}\right), \quad \quad {\Phi}^{(2)}_{n} = \left( \begin{array}{l} \phi^{(2)}_{+,n} \\ \phi^{(2)}_{-,n} \end{array}\right) = \textbf{T}_{\omega,n-1} \left( \begin{array}{l} 0 \\ 1 \end{array}\right), $$

and let \(R^{(i)}_{n}, \theta ^{(i)}_{n}, n\geq 1\), i = 1, 2 be the corresponding Prüfer radii and phases. We let \(r_{n}=\frac {R^{(1)}_{n}}{R^{(2)}_{n}}\) and 𝜗n be as in Lemma A.2. Recall the relation \({\Phi }_{n}={\mathcal {P}}_{n} {\Psi }_{n}\). In particular,

$$ {\Phi}_{n}^{(i)} = \left( \begin{array}{l} \phi^{(i)}_{+,n} \\ \phi^{(i)}_{-,n} \end{array}\right) = (-1)^{n-1} R_{n} \left( \begin{array}{l} -\sqrt{p_{2}}\cos(\bar{\theta}_{n}^{(i)}) \\ \sqrt{-p_{1}}\cos(\bar{\theta}_{n}^{(i)}+k) \end{array}\right). $$

Thus it follows from some elementary trigonometry that

$$ \phi^{(1)}_{+,n}\phi^{(2)}_{-,n}-\phi^{(1)}_{-,n}\phi^{(2)}_{+,n} = R^{(1)}_{n} R^{(2)}_{n} \sin(2k) \sin(\theta^{(1)}_{n}-\theta^{(2)}_{n}), $$

where \(\bar {\theta }^{(i)}_{n} = \theta ^{(i)}_{n}-(2n-1)k\). On the other hand,

$$ \phi^{(1)}_{+,n}\phi^{(2)}_{-,n}-\phi^{(1)}_{-,n}\phi^{(2)}_{+,n} = \det\left( \textbf{T}_{\omega,n} \left( \begin{array}{ll} 1 0 \\ 0 1 \end{array}\right) \right) =1. $$

This, together with the convergence

$$ \lim_{n\to\infty} \frac{R^{(i)}_{n}}{\log n} = \beta,\quad i=1,2, $$

gives

$$ \lim_{n\to\infty}\frac{\log |\sin(\theta^{(2)}_{n}-\theta^{(1)}_{n})|}{\log n} =\lim_{n\to\infty}\frac{\log |\sin(\bar\theta^{(2)}_{n}-\bar\theta^{(1)}_{n})|}{\log n} = -2\beta. $$
(A.2)

Remember the decomposition (4.2) that we summarize as

$$ \begin{array}{@{}rcl@{}} (R^{(i)}_{n+1})^{2} &=& \left( 1 -\frac{p_{2}}{\sin(2k)}\sin(2\bar\theta_{n}^{(i)})V_{\omega,1}(n)\right.\\ && \left.+\frac{p_{1}}{\sin(2k)}\sin(2(\bar\theta^{(i)}_{n}-k))V_{\omega,2}(n+1) +E_{j}^{(i)} \right) (R^{(i)}_{n})^{2}. \end{array} $$

We have to estimate the difference of the expansions for \(\log R^{(1)}_{n}\) and \(\log R^{(2)}_{n}\). By (A.2), one has \(|\sin \limits (\bar \theta ^{(2)}_{n}-\bar \theta ^{(1)}_{n})| \lesssim n^{-\beta +\epsilon }\), for any 𝜖 > 0. Hence, there exist random sequences \((m_{n})_{n}\subset \mathbb {N}^{*}\) and \(({\Delta }_{n})_{n}\subset \mathbb {R}\) such that \(\bar \theta ^{(1)}_{n}-\bar \theta ^{(2)}_{n}=m_{n}\pi + {\Delta }_{n}\) and \(|{\Delta }_{n}|\lesssim n^{-\beta +\epsilon }\). Therefore,

$$ \sin(2\bar\theta^{(2)}_{n}) = \sin(2\bar\theta^{(1)}_{n}+2{\Delta}_{n}) \simeq \sin(2\bar\theta^{(1)}_{n})+2\cos(2\bar\theta^{(1)}_{n}){\Delta}_{n}. $$

This shows that

$$ \left|V_{\omega,1}(j) \left( \sin(2\bar{\theta}^{(1)}_{j}) - \sin(2\bar{\theta}^{(2)}_{j})\right)\right| \lesssim j^{-\frac{1}{2}-2\beta + \epsilon}. $$

By means of similar arguments, one can show that

$$ \left|V_{\omega,2}(j+1) \left( \sin(2(\bar{\theta}^{(1)}_{j}-k)) - \sin(2(\bar{\theta}^{(2)}_{j}-k))\right)\right| \lesssim j^{-\frac{1}{2}-2\beta + \epsilon}. $$

and

$$ |E_{j}^{(1)}-E_{j}^{(2)}| \lesssim j^{-1-2\beta + \epsilon}. $$

Hence,

$$ \begin{array}{@{}rcl@{}} \log r_{n} &=& -\frac{p_{2}}{\sin(2k)}\sum\limits^{n}_{j=1} \frac{V_{\omega,1}(j)}{\sin k} \left( \sin(2\bar{\theta}^{(1)}_{j}) - \sin(2\bar{\theta}^{(2)}_{j}) \right) \end{array} $$
(A.3)
$$ \begin{array}{@{}rcl@{}} && + \frac{p_{1}}{\sin(2k)}\sum\limits^{n}_{j=1} \frac{V_{\omega,2}(j+1)}{\sin k} \left( \sin(2(\bar{\theta}^{(1)}_{j}-k)) - \sin(2(\bar{\theta}^{(2)}_{j}-k)) \right) + \sum\limits^{n}_{j=1} A_{j} \end{array} $$
(A.4)

where the first two sums are convergent martingales by Lemma A.4 with \(\gamma =\frac 12 + 2\beta -\epsilon \) and the last one is absolutely convergent as Aj = O(j− 1 − 2β+𝜖). This shows that \(r_{n}\to r_{\infty }\in (0,\infty )\) almost surely which implies that 𝜗n has a limit \(\vartheta _{\infty }\neq 0, \frac {\pi }{2}\) by the first part of Lemma A.2.

The equivalence (A.1) in our context corresponds to

$$ \underset{n\to\infty}{\lim}\frac{\log R_{n}(\vartheta_{\infty})}{\log n}=-\beta \quad \text{if and only if} \quad \underset{n}{\limsup} \frac{\log |r_{n}-r_{\infty}|}{\log n} \leqslant -2\beta. $$

Let us denote by \(\log r_{n} = M_{n} + S_{n}\) the decomposition (A.4) and \(\log r_{\infty }=M_{\infty }+S_{\infty }\), where Mn is the martingale part and \(M_{\infty }\) and \(S_{\infty }\) are the almost sure limits of Mn and Sn respectively. Then,

$$ \begin{array}{@{}rcl@{}} |r_{\infty}-r_{n}| &=& \mathrm{e}^{M_{\infty}+S_{\infty}}\left| 1-\mathrm{e}^{M_{n}-M_{\infty}+S_{n}-S_{\infty}}\right| \simeq \mathrm{e}^{M_{\infty}+S_{\infty}}\left| M_{n}-M_{\infty}+S_{n}-S_{\infty}\right|\\ &\lesssim& \mathrm{e}^{M_{\infty}+S_{\infty}} n^{-2\beta+2\epsilon}, \end{array} $$

by the last statement of Lemma A.4 with \(\gamma = \frac 12 + 2\beta - \epsilon \) and Aj = O(j− 1 − 2β+𝜖). This finishes the proof of Proposition 6.2. □

We state [53, Theorem 8.3] which allowed us to prove pure point spectrum in the sub-critical regime:

Theorem A.3

Let (An)n≥ 1 be 2 × 2 real unimodular matrices and let An = AnA1 such that

$$ \underset{n\geq 1}{\sum} \frac{\| A_{n+1}\|}{\| \textbf{A}_{n} \|}<\infty. $$

Suppose there exists a monotone increasing function \(g:\mathbb {N}^{*} \to (0,\infty )\) such that

$$ \underset{n\to\infty}{\lim} \frac{\log \| A_{n} \|}{g(n)}=0 \quad \text{and} \quad \underset{n\to\infty}{\lim} \frac{\log \| \textbf{A}_{n} \|}{g(n)}=1, $$

and such that

$$ \underset{n\geq 1}{\sum} \mathrm{e}^{-\epsilon g(n)}<\infty, $$

for all 𝜖 > 0. Then, there exists an angle 𝜗0 such that

$$ \underset{n\to\infty}{\lim} \frac{\log \| \textbf{A}_{n} \widehat{\vartheta}_{0} \|}{g(n)}=-1. $$

1.2 A.2. A Martingale Inequality

The following corresponds to [48, Lemma 8.4]. We formulate it in full generality but provide a short proof under the assumption that Vω,i(n) = λnαωn,i for uniformly bounded random variables ωn,i.

Lemma A.4

Let (Zj)j be i.i.d. random variables with \(\mathbb {E}[Z_{n}]=0\) and \(\mathbb {E}[|Z_{n}|^{2}]\leqslant n^{-2\gamma }\) for some γ > 0. Let \(\mathcal {G}_{n}=\sigma (Z_{1},\cdots , Z_{n})\) and let \(Y_{n} \in \mathcal {G}_{n-1}\) for n ≥ 1 such that \(|Y_{n}|\leqslant 1\). Define

$$ M_{n} = \sum\limits^{n}_{j=1} Y_{j}Z_{j} \quad \text{and} \quad s_{n} = \sum\limits^{n}_{j=1} \frac{1}{j^{2\gamma}}. $$

Then, (Mn)n is a \(\mathcal {G}_{n}\)-martingale and

  1. (i)

    For \(\gamma \leqslant \frac 12\) and all ε > 0,

    $$ \underset{n\to\infty}{\lim} s_{n}^{-\frac{1+\varepsilon}{2}}M_{n}= 0,\quad \quad \mathbb{P}-a.s. $$
  2. (ii)

    For \(\gamma >\frac 12\), (Mn)n converges \(\mathbb {P}\)-almost surely to a finite (random) limit \(M_{\infty }\) and, for all \(\kappa <\gamma - \frac 12\), we have

    $$ \underset{n\to\infty}{\lim} n^{\kappa} \left( M_{\infty}-M_{n}\right) = 0,\quad \mathbb{P}-a.s. $$

Proof

The reader can consult the book [34] for the general properties of martingales used below. The sequence (Mn)n is a martingale thanks to our hypothesis on (Yn)n and (Zn)n: indeed, since \(M_{n}, Y_{n+1} \in \mathcal {G}_{n}\), Yn+ 1 is bounded and Zn+ 1 is independent of \(\mathcal {G}_{n}\) and centered, we have, \(\mathbb {P}\)-almost surely,

$$ \begin{array}{@{}rcl@{}} \mathbb{E}[M_{n+1} | \mathcal{G}_{n}] &=& \mathbb{E}\left[Y_{n+1}Z_{n+1} + M_{n} \Big{|}\mathcal{G}_{n}\right]\\ &=& Y_{n+1}\mathbb{E}[Z_{n+1}] + M_{n} = M_{n}. \end{array} $$

As stated above, we assume Zn = nγXn with \(|X_{n}|\leqslant 1\) and \(\mathbb {E}[X_{n}]=0\) to simplify the argument. Let \(\gamma \leqslant \frac 12\). We use Azuma’s inequality [5]: let (Mn)n be a martingale such that \(|M_{n}-M_{n-1}|\leqslant c_{n}\) for all n ≥ 1. Then,

$$ \mathbb{P}\left[ |M_{n}-M_{0}| \geq t\right] \leqslant 2 \exp\left\{ - \frac{t^{2}}{2 {\sum}^{n}_{j=1} {c_{j}^{2}}}\right\}. $$

In our case, M0 = 0, cj = 2jγ, and taking \(t=s_{n}^{\frac {1+\varepsilon ^{\prime }}{2}}\) for \(0<\varepsilon ^{\prime }<\varepsilon \), we obtain

$$ \mathbb{P}\left[ |M_{n}| \geq s_{n}^{\frac{1+\varepsilon^{\prime}}{2}}\right] \leqslant 2\ \mathrm{e}^{- C n^{\varepsilon^{\prime}}}, $$

for some C > 0. The claim (i) then follows from Borel-Cantelli’s lemma. Now, let \(\gamma >\frac 12\). Noticing that, for i < l,

$$ \mathbb{E}[X_{i}Y_{i} X_{l} Y_{l}] = \mathbb{E}[X_{l}]\mathbb{E}[X_{i}Y_{i} Y_{l}] =0, $$

we have

$$ \sup_{n}{\mathbb{E}[M_{n}^{2}}] = \sup_{n} \sum\limits^{n}_{j=1} \frac{{\mathbb{E}[X_{j}^{2}} {Y_{j}^{2}}]}{j^{2\gamma}} \leqslant \underset{j\geq 1}{\sum} \frac{1}{j^{2\gamma}}<\infty. $$

Hence, (Mn)n is bounded in L2 and, as a consequence, converges almost surely, i.e., there exists a random variable \(M_{\infty }\) such that \(\lim _{n\to \infty }M_{n}=M_{\infty }\), \(\mathbb {P}\)-a.s.. Finally, applying Azuma’s inequality to the martingale (Mn+iMn)i≥ 0, we obtain

$$ \mathbb{P}\left[ n^{\kappa}\left| M_{n+i}-M_{n}\right| \geq 1\right]\leqslant 2 \exp\left\{ - C n^{2(\gamma-\frac12 -\kappa)}\right\}, $$

for all i ≥ 0. Choosing \(\kappa <\gamma -\frac 12\), the last claim follows from Fatou’s lemma, the convergence of (Mn)n and Borel-Cantelli. □

Remark 1

The result above is proved in [48, Lemma 8.4] under the second moment assumption replacing our use of Azuma’s inequality by Doob’s inequality. For a short proof assuming bounded exponential moments, see [21, Lemma A.1].

1.3 A.3. Control of the Phases

The next lemma provides the control of the Prüfer phases needed to complete the proof of Proposition 4.1. The strategy is taken from [48]. Recalling the definitions of Qn,1 and Qn,2 from (4.5),

Lemma A.5

Assume (A3a) and (A4). Let \(0<\alpha \leqslant \frac 12\). For each fixed energy corresponding to a value of \(k \in (-\pi , -\tfrac {\pi }{2})\) different from \(-\frac {5\pi }{8},-\frac {3\pi }{4}\) and \(-\frac {7\pi }{8}\),

$$ \underset{n\to\infty}{\lim} \frac{Q_{n,i}}{{\sum}^{n}_{j=1} j^{-2\alpha}} = 0, $$

for i = 1, 2. Moreover, for each compact energy interval I ⊂Σ̈, the convergence is uniform over all initial values 𝜃0 ∈ [0, 2π) and EI corresponding to values of k different from \(-\frac {5\pi }{8},-\frac {3\pi }{4}\) and \(-\frac {7\pi }{8}\).

Proof

We will show that

$$ \limsup_{n\to\infty} \frac{{\sum}^{n}_{j=1}\mathbb{E}[ V_{\omega,j}^{2}]\cos 4 \bar \theta_{j}}{{\sum}^{n}_{j=1} j^{-2\alpha}} = 0, \qquad \mathbb{P}-\text{a.s.,} $$

the other terms being handled similarly. Note that the prefactor accompanying this term in the definition of Qn,1 is uniformly bounded over compact energy intervals. The computations below are uniform in the initial condition 𝜃0 and only assume \(k\notin \tfrac {\pi }{8}\mathbb {Z}\).

We begin with a simple observation: from (3.17), for any compact interval I ⊂Σ̈, there exists a constant \(C=C(I)\in (0,\infty )\) such that

$$ \begin{array}{@{}rcl@{}} && \left| \mathrm{e}^{i(\theta_{n+1}-\theta_{n})}-1\right| = \left| \frac{\zeta_{n+1}}{\zeta_{n}}-1\right| \\ && \leqslant C \left( |V_{\omega,1}(n)|+|V_{\omega,2}(n+1)|+|V_{\omega,1}(n)V_{\omega,2}(n+1)| \right) \leqslant 1, \end{array} $$

for nn(ω) for some \(n^{*}(\omega )=n^{*}(\omega ,I)<\infty \) thanks to (A4). Hence, for nn(ω), \(|\theta _{n+1}-\theta _{n}|< \frac {\pi }{2}\) and, recalling (A4) once more, we have

$$ |\theta_{n+1}-\theta_{n}| \leqslant \frac{\pi}{2} |\sin(\theta_{n+1}-\theta_{n})| \leqslant \frac{\pi}{2} \left|\mathrm{e}^{i(\theta_{n+1}-\theta_{n})}-1\right| \leqslant c_{0}(\omega)\ n^{-\frac{2\alpha}{3}}, $$

for some \(c_{0}(\omega )=c_{0}(\omega ,I)\in (0,\infty )\). This can be written in the equivalent form

$$ |\bar{\theta}_{n+1}-\bar{\theta}_{n}+2k|\leqslant c_{0}(\omega) \ n^{-\frac{2\alpha}{3}}, $$
(A.5)

which will be more suitable for our purposes. By possibly increasing the value of c0(ω), we can assume that (A.5) holds for all nn(ω) with \(c_{0}(\omega )<\infty \)\(\mathbb {P}\)-amost surely. For p ≥ 1, define

$$ \mathcal{E}_{p} = \left\{ \omega: c_{0}(\omega) \leqslant p\right\}, $$

and observe that \({\Omega }=\underset {p\geq 1}{\bigcup } \mathcal {E}_{p}\). The key proof is [48, Lemma 8.5] which states the following: suppose that \(y\in \mathbb {R}\) is not in \(\pi \mathbb {Z}\). Then, there exists a sequence of integers \(q_{l} \to \infty \) such that

$$ \left|\sum\limits^{q_{l}}_{j=1} \cos \theta_{j} \right| \leqslant 1 + \sum\limits^{q_{l}}_{j=1} \left| \theta_{j}-\theta_{0} - jy \right|, $$
(A.6)

for all \((\theta _{j})_{j\geq 0}\subset \mathbb {R}\). We take y = − 8k. Let p ≥ 1 and \(\omega \in \mathcal {E}_{p}\). Let n be large enough so that it can be written as n = n0 + Kql with \(n_{0}\ge {q_{l}^{2}}\) and \(4c_{0}(\omega ) n_{0}^{-\alpha }\le q_{l}^{-2}\). Then,

$$ \begin{array}{@{}rcl@{}} \left|\sum\limits_{j=n_{0}+1}^{n} j^{-2\alpha}\cos(4\bar\theta_{j})\right| &=& \left|\sum\limits_{m=0}^{K} \sum\limits_{r=1}^{q_{l}}(n_{0}+mq_{l}+r)^{-2\alpha}\cos(4\bar\theta(n_{0}+mq_{l}+r))\right| \\ &\leqslant& \sum\limits_{m=0}^{K}(n_{0}+mq_{l})^{-2\alpha} \left|\sum\limits_{r=1}^{q_{l}}\cos(4\bar\theta(n_{0}+mq_{l}+r))\right| \\ && + \sum\limits_{m=0}^{K} \sum\limits_{r=1}^{q_{l}} \left|(n_{0}+mq_{l}+r)^{-2\alpha}-(n_{0}+mq_{l})^{-2\alpha}\right| \\ &=:& A + B. \end{array} $$

We first estimae the term A using (A.6) to get

$$ A \leqslant \sum\limits_{m=0}^{K}(n_{0}+mq_{l})^{-2\alpha}\left( 1+4{\sum}_{r=1}^{q_{l}}|\bar\theta(n_{0}+mq_{l}+r)-\bar\theta(n_{0}+mq_{l})+2kr|\right). $$

Now, by (A.5) it follows that

$$ \begin{array}{@{}rcl@{}} && 4\sum\limits_{r=1}^{q_{l}}|\bar\theta(n_{0}+mq_{l}+r)-\bar\theta(n_{0}+mq_{l})+2kr| \leqslant c_{0} \sum\limits_{r=1}^{q_{l}}\sum\limits_{s=1}^{r}(n_{0}+mq_{l}+r)^{-\frac{2\alpha}{3}} \\ && \leqslant c_{0}(\omega)(n_{0}+mq_{l})^{-\frac{2\alpha}{3}}\sum\limits_{r=1}^{q_{l}} r \leqslant c_{0}(\omega){q_{l}^{2}} n_{0}^{-\frac{2\alpha}{3}} \leqslant 1. \end{array} $$

Thus,

$$ A \leqslant 2 \sum\limits_{m=0}^{K}(n_{0}+mq_{l})^{-2\alpha} \leqslant 2 q_{l}^{-2\alpha} \sum\limits_{m=0}^{K}(n_{0}q_{l}^{-1}+m)^{-2\alpha} \leqslant c_{1} q_{l}^{-2\alpha} \sum\limits_{j=1}^{K}j^{-2\alpha}, $$

for some finite c1 > 0. To estimate B, we use that

$$ \left| (n_{0}+mq_{l}+r)^{-2\alpha}-(n_{0}+mq_{l})^{-2\alpha}\right| \leqslant c_{2} (n_{0}+mq_{l})^{-2\alpha-1} r, $$

for some finite c2 > 0 which allows us to write

$$ \begin{array}{@{}rcl@{}} B &\leqslant& c_{2} \sum\limits_{m=0}^{K} \sum\limits_{r=1}^{q_{l}} (n_{0}+mq_{l})^{-2\alpha-1} r \leqslant c_{2} {q_{l}^{2}} n_{0}^{-1} \sum\limits_{m=0}^{K} (1+n_{0}^{-1}mq_{l})^{-1}(n_{0}+mq_{l})^{-2\alpha} \\ &\leqslant& c_{2} \sum\limits_{m=0}^{K} (n_{0}+mq_{l})^{-2\alpha}, \end{array} $$

where we used \({q_{l}^{2}} n_{0}^{-1}\leqslant 1\). This last sum can be estimated as above. Combining, we obtain

$$ \left|\sum\limits^{n}_{j=1} j^{-2\alpha} \cos 4 \bar{\theta}_{j} \right| \leqslant \sum\limits^{n_{0}}_{j=1}j^{-2\alpha} + c_{3} q_{l}^{-2\alpha} \sum\limits_{j=1}^{K}j^{-2\alpha}, $$

for some finite c3 > 0 and all \(\omega \in \mathcal {E}_{p}\). Hence,

$$ \underset{n\to\infty}{\limsup}\frac{\left|{\sum}^{n}_{j=1} j^{-2\alpha} \cos 4 \bar{\theta}_{j} \right|}{{\sum}^{n}_{j=1}j^{-2\alpha}} \leqslant c_{3} q_{l}^{-2\alpha}, $$

for all \(\omega \in \mathcal {E}_{p}\). We can then let \(l\to \infty \). As the events \(\mathcal {E}_{p}\) exhaust Ω, this finishes the proof. □

The next lemma provides the control of the phases needed to complete the proof of Proposition 4.2.

Lemma A.6

Let \(0<\alpha \leqslant \frac 12\). Assume (A1)-(A3a) and (A5). For each fixed energy corresponding to a value of \(k \in (-\pi , -\tfrac {\pi }{2})\) different from \(-\frac {5\pi }{8},-\frac {3\pi }{4}\) and \(-\frac {7\pi }{8}\),

$$ \underset{n\to\infty}{\lim} \frac{\mathbb{E}[Q_{n,i}]}{{\sum}^{n}_{j=1} j^{-2\alpha}} = 0, $$

for i = 1, 2. Moreover, for each compact energy interval I ⊂Σ̈, the convergence is uniform over all initial values 𝜃0 ∈ [0, 2π) and EI corresponding to values of k different from \(-\frac {5\pi }{8},-\frac {3\pi }{4}\) and \(-\frac {7\pi }{8}\).

Proof

By Borel-Cantelli ,

$$ |V_{\omega,i}(n)| \leqslant n^{-\frac{2\alpha}{3}-\frac{\varepsilon}{2}}, $$

for all nτ, i = 1, 2 for some \(\mathbb {P}\)-almost surely finite τ = τ(ω). Thanks to the uniform control of the previous lemma, we have

$$ \underset{n\to\infty}{\lim} \frac{\mathbb{E}\left[\left|{\sum}^{n}_{j=\tau(\omega)} j^{-2\alpha} \cos 4 \bar{\theta}_{j} \right|\right]}{{\sum}^{n}_{j=1}j^{-2\alpha}} = 0. $$

It is then enough to show that

$$ \mathbb{E}\left[ \sum\limits^{\tau(\omega)}_{j=1} j^{-2\alpha}\right] < \infty. $$

If \(\alpha \neq \frac 12\), we have

$$ \begin{array}{@{}rcl@{}} \mathbb{E}\left[ \sum\limits^{\tau(\omega)}_{j=1} j^{-2\alpha}\right] &=& \underset{k\geq 1}{\sum} \left( \sum\limits^{k}_{j=1} j^{-2\alpha} \right) \mathbb{P}[\tau(\omega)=k] \\ &\leqslant& C_{1} \underset{k\geq 1}{\sum} k^{1-2\alpha} \mathbb{P}[\tau(\omega)=k], \end{array} $$

for some finite C1 > 0. We can estimate the probability inside the sum:

$$ \begin{array}{@{}rcl@{}} \mathbb{P}[\tau(\omega)=k] &\leqslant& \mathbb{P}\left[|V_{j}| > j^{-\frac{2\alpha}{3}-\frac{\varepsilon}{2}}, \forall j<k\right] \\ &=& \prod\limits^{k-1}_{j=1} \mathbb{P}\left[|V_{j}| > j^{-\frac{2\alpha}{3}-\frac{\varepsilon}{2}}\right] \leqslant \prod\limits^{k-1}_{j=1} j^{(\frac{2\alpha}{3}+\frac{\varepsilon}{2})p} \mathbb{E}\left[|V_{j}|^{p}\right] \\ &\leqslant& {C_{2}^{k}} \prod\limits^{k-1}_{j=1} j^{-\varepsilon p} \leqslant {C_{2}^{k}} \mathrm{e}^{-c \varepsilon p k \log k}, \end{array} $$

for some finite C1 > 0 and c > 0. Hence,

$$ \mathbb{E}\left[ \sum\limits^{\tau(\omega)}_{j=1} j^{-2\alpha}\right] \leqslant C_{1} \underset{k\geq 1}{\sum} k^{1-2\alpha} {C_{2}^{k}} \mathrm{e}^{-c \varepsilon p k \log k} < \infty. $$

The case \(\alpha =\frac 12\) is similar. All the above estimates hold uniformly in EI corresponding to values of k different from \(-\frac {5\pi }{8},-\frac {3\pi }{4}\) and \(-\frac {7\pi }{8}\). □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bourget, O., Moreno Flores, G.R. & Taarabt, A. One-dimensional Discrete Dirac Operators in a Decaying Random Potential I: Spectrum and Dynamics. Math Phys Anal Geom 23, 20 (2020). https://doi.org/10.1007/s11040-020-09341-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11040-020-09341-7

Keywords

Mathematics Subject Classification (2010)

Navigation