Skip to main content

One-dimensional Discrete Anderson Model in a Decaying Random Potential: from A.C. Spectrum to Dynamical Localization

  • Conference paper
  • First Online:
Spectral Theory and Mathematical Physics

Abstract

We consider a one-dimensional Anderson model where the potential decays in average like n α, α > 0. This simple model is known to display a rich phase diagram with different kinds of spectrum arising as the decay rate α varies.

We review an article of Kiselev, Last and Simon where the authors show a.c. spectrum in the super-critical case \(\alpha >\frac 12\), a transition from singular continuous to pure point spectrum in the critical case \(\alpha =\frac 12\), and dense pure point spectrum in the sub-critical case \(\alpha <\frac 12\). We present complete proofs of the cases \(\alpha \ge \frac 12\) and simplify some arguments along the way. We complement the above result by discussing the dynamical aspects of the model. We give a simple argument showing that, despite of the spectral transition, transport occurs for all energies for \(\alpha =\frac 12\). Finally, we discuss a theorem of Simon on dynamical localization in the sub-critical region \(\alpha <\frac 12\). This implies, in particular, that the spectrum is pure point in this regime.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

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

    Article  MathSciNet  Google Scholar 

  2. M. Aizenman, S. Warzel, Random Operators: Disordered effects on Quantum Spectra and Dynamics. Graduate Studies in Mathematics, vol 168 (American Mathematical Society, Providence, 2016)

    Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  4. K. Azuma, Weighted sums of certain dependent random variables. Tôhoku Math. J. 19(3), 357–367 (1967)

    Article  MathSciNet  Google Scholar 

  5. J. Bourgain, On random Schrödinger operators on . Discret Contin. Dyn. Syst. 8, 1–15 (2002)

    Google Scholar 

  6. J. Bourgain, Random lattice schrödinger operators with decaying potential: some higher dimensional phenomena, in Geometric Aspects of Functional Analysis. Lectures Notes in Mathematics, vol. 1807 (Springer, Berlin, 2003), pp. 70–98

    Google Scholar 

  7. O. Bourget, G. Moreno, A. Taarabt, Dynamical localization for the one-dimensional continuum Anderson model in a decaying random potential. Ann. Henri Poincaré, 21(8), (2020)

    Google Scholar 

  8. O. Bourget, G. Moreno, A. Taarabt, One-dimensional discrete dirac operators in a decaying random potential I: Spectrum and dynamics. Math. Phys. Anal. Geom. 23(20), (2020).

    Google Scholar 

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

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

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

    Google Scholar 

  12. D. Damanik, A. Gorodetski, An extension of the Kunz-Souillard approach to localization in one dimension and applications to almost-periodic Schrödinger operators. Adv. Math. 297, 149–173 (2016)

    Article  MathSciNet  Google Scholar 

  13. D. Damanik, G. Stolz, A continuum version of the Kunz–Souillard approach to localization in one dimension. J. Für die reine und Angewandte Math. (Crelles J.) 660, 99–130 (2011)

    MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  16. F. Delyon, Appearance of a purely singular continuous spectrum in a class of random Schrödinger operators. J. Statist. Phys. 40, 621–630 (1985)

    Article  MathSciNet  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  18. R. Durrett, Probability: Theory and Examples. Cambridge Series in Statistical and Probabilistic Mathematics, 4th edn. (Cambridge University Press, New York, 2010)

    Google Scholar 

  19. A. Figotin, F. Germinet, A. Klein, P. Müller, Persistence of Anderson localization in Schrödinger operators with decaying random potentials. Ark. Mat. 45. 15–30 (2007)

    Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  21. L. Ge, X. Zhao, Exponential dynamical localization in expectation foe one dimensional Anderson model. J. Spect. Theory

    Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  23. F. Germinet, A. Taarabt, Spectral properties of dynamical localization for Schrödinger operators. Rev. Math. Phys. 25(9), 1350016 (2013)

    Google Scholar 

  24. F. Germinet, A. Kiselev, S. Tcheremchantsev, Transfer matrices and transport for Schrödinger operators. Ann. Inst. Fourier 54, 787–830 (2004)

    Article  MathSciNet  Google Scholar 

  25. F. Germinet, A. Klein, J. Schenker, Dynamical delocalization in random Landau Hamiltonians. Ann. Math. 166, 215–244 (2007)

    Article  MathSciNet  Google Scholar 

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

    Article  Google Scholar 

  27. V. Jaks̆ić, Y. Last, Spectral structure of Anderson type Hamiltonians. Inven. Math. 141, 561–577 (2000)

    Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  29. A. Klein, Extended states in the anderson model on the bethe lattice. Adv. Math. 133, 163–184 (1998)

    Article  MathSciNet  Google Scholar 

  30. W. Kirsch, M. Krishna, J. Obermeit, Anderson model with decaying randomness: mobility edge. Math. Z. 235, 421–433 (2000)

    Article  MathSciNet  Google Scholar 

  31. A. Kisselev, Y. Last, B. Simon, Modified Prüfer and EFGP transforms and the spectral analysis of one-dimensional schrödinger operators. Comm. Math. Phys. 194, 1–45 (1998)

    Article  MathSciNet  Google Scholar 

  32. S. Kotani, N. Ushiroya, One-dimensional Schrödinger operators with random decaying potentials. Comm. Math. Phys. 115, 247–266 (1988)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  35. Y. Last, B. Simon, Eigenfunctions, transfer matrices, and absolutely continuous spectrum of one-dimensional Schrödinger operators. Invent. Math. 135, 329–367 (1999)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  37. B. Simon, Spectral Analysis of Rank One Perturbations and Applications. CRM Lectures Notes, vol. 8 (American Mathematical Society, Providence, 1995), pp. 109–149

    Google Scholar 

Download references

Acknowledgements

O. Bourget is partially supported by Fondecyt grant 1161732. G. Moreno Flores is 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’. A. Taarabt is partially supported by Fondecyt grant 11190084.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Amal Taarabt .

Editor information

Editors and Affiliations

Appendix A: Some Technical Lemmas

Appendix A: Some Technical Lemmas

1.1 A.1 Two Results on Unimodular Matrices

We say that a matrix is unimodular if it has determinant equal to 1. The following lemma is used to compare the asymptotics of the transfer matrix with those of the sequence (R n)n.

Lemma A.1

Let A be an unimodular matrix and let \(\hat \theta = (\cos \theta ,\, \sin \theta )\). Then, for all pair of angles \(|\theta _1-\theta _2|\leq \frac {\pi }{2}\),

$$\displaystyle \begin{aligned} \begin{array}{rcl} \| A \| \leq \sin \left( \tfrac{|\theta_1-\theta_2|}{2}\right)^{-1} \max \{ \| A \hat \theta_1\|,\, \| A\hat \theta_2\| \}. \end{array} \end{aligned} $$
(A.1)

Proof

First, there exists angles θ 0 and σ 0 such that \(A^* \hat \sigma _0 = \| A \| \hat \theta _0\). Then, for any angle θ,

$$\displaystyle \begin{aligned} \begin{array}{rcl} | \cos{}(\theta-\theta_0) | = |\langle \hat \theta, \hat \theta_0 \rangle | = \frac{1}{\| A \|} | \langle\hat \theta, A^*\hat \sigma_0 \rangle | = \frac{1}{\| A \|} | \langle A \hat \theta, \hat \sigma_0 \rangle | \leq \frac{\| A \hat \theta \|}{\| A \|}. \end{array} \end{aligned} $$

This way, for any pair of angles, we obtain

$$\displaystyle \begin{aligned} \begin{array}{rcl} \| A \| \max\{ | \cos{}(\theta_1-\theta_0) |,\, | \cos{}(\theta_2-\theta_0) | \} \leq \max\{ \| A\hat \theta_1\|,\, \| A\hat \theta_2\|\}. \end{array} \end{aligned} $$

To conclude, just note that for all |γ|≤ π∕2, the minimum of the function \(x\mapsto \max \{|\cos {}(x)|, \) \( |\cos {}(x+\gamma )|\}\) is attained at π∕2 − γ∕2 and is equal to \(|\cos {}(\pi /2-\gamma /2)|=|\sin {}(\gamma /2)|\). □

The following lemma is used to find eigenfunctions with the proper decay.

Lemma A.2

For a unimodular matrix withA∥ > 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) = ∥A(1, 0)T∥∕∥A(0, 1)T∥.

Let (A n)n be a sequence of unimodular matrices withA n∥ > 1 and write 𝜗 n = 𝜗(A ) and r n = r(A n). Assume that

  1. (i)

    limnA n∥ = ∞,

  2. (ii)

    \(\displaystyle \lim _{n\to \infty } \frac {\| A_{n+1}A_n^{-1}\|}{\| A_n \| \, \| A_{n+1}\|}=0\).

Then,

  1. 1.

    (𝜗 n)n has a limit 𝜗 ∈ (−π∕2, π∕2) if and only if (r n)n has a limit r ∈ [0, ). If 𝜗 n →±π∕2, then r n →∞ but, if r n →∞, we can only conclude that |𝜗 n|→ π∕2.

  2. 2.

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

    $$\displaystyle \begin{aligned} \begin{array}{rcl}{} \lim_{n\to\infty} \frac{\log \| A_n \hat\vartheta_{\infty}\|}{\log \| A_n \|} = -1 \quad \mathit{\text{if and only if}} \quad \limsup_n \frac{\log |r_n - r_{\infty}|}{\log \| A_n \|} \leq -2.\qquad \quad \end{array} \end{aligned} $$
    (A.2)

Proof

Recall 𝜗 n denotes the unique angle in (−π∕2, π∕2] such that \(\| A_n \hat \vartheta _n\| = \| A_n \|{ }^{-1}\). Let \(\vartheta _n^{\prime } \in (\vartheta _n-\pi /2, \vartheta _n+\pi /2]\) be the unique angle such that \(\| A_n \hat \vartheta _n^{\prime }\| = \| A_n \|\) and let \(\vartheta _n^\perp \) be either 𝜗 n − π∕2 or 𝜗 n + π∕2 and such that \(\vartheta _n^{\prime }\) lies between 𝜗 n and \(\vartheta _n^{\perp }\).

Let \(\displaystyle v_0 = \begin {pmatrix} 1 \\0 \end {pmatrix}\), \(\displaystyle w_0 = \begin {pmatrix} 0\\1 \end {pmatrix}\), v n = A nv 0 and w n = A nw 0. Then,

$$\displaystyle \begin{aligned} \begin{array}{rcl} \hat\vartheta_n = \cos\vartheta_n v_0 + \sin\vartheta_nw_0, \end{array} \end{aligned} $$
(A.3)

and applying A n on both sides,

$$\displaystyle \begin{aligned} \begin{array}{rcl} \pm\frac{1}{\|A_n\|} \hat\vartheta_n= A_n \hat\vartheta_n = \cos \vartheta_n v_n + \sin \vartheta_n w_n \simeq 0, \end{array} \end{aligned} $$
(A.4)

as ∥A n∥→. Hence,

$$\displaystyle \begin{aligned} \begin{array}{rcl} r_n = r(A_n) = \frac{\|v_n\|}{\|w_n\|} \simeq |\tan \vartheta_n|. \end{array} \end{aligned} $$
(A.5)

This shows that, if 𝜗 n → 𝜗 , then \(r_n\to |\tan \vartheta _\infty |\in [0,\infty ]\). On the other hand, if r n → r ∈ [0, ], then \(|\vartheta _n| \to \arctan (\rho _{\infty })\), with the convention \(\arctan (\infty )=\pi /2\). If ρ  = , we only get that |𝜗 n|→ π∕2. Otherwise, it is enough to prove that 𝜗 n − 𝜗 n−1 → 0 to show the convergence of 𝜗 n. For this, we use the decomposition

$$\displaystyle \begin{aligned} \begin{array}{rcl}{} \hat\vartheta = \cos{}(\vartheta-\vartheta_n) \hat\vartheta_n + \sin{}(\vartheta-\vartheta_n) \hat\vartheta_n^\perp. \end{array} \end{aligned} $$
(A.6)

Applied to 𝜗 n+1, this yields

$$\displaystyle \begin{aligned} \begin{array}{rcl} \hat\vartheta_{n+1} = \cos{}(\vartheta_{n+1}-\vartheta_n) \hat\vartheta_n + \sin{}(\vartheta_{n+1}-\vartheta_n) \hat\vartheta_n^\perp, \end{array} \end{aligned} $$
(A.7)

and

$$\displaystyle \begin{aligned} \begin{array}{rcl} A_n\hat\vartheta_{n+1} &\displaystyle =&\displaystyle \pm \| A_n \|{}^{-1}\cos{}(\vartheta_{n+1}-\vartheta_n) \hat\vartheta_n + \sin{}(\vartheta_{n+1}-\vartheta_n) A_n\hat\vartheta_n^\perp\\ &\displaystyle \simeq&\displaystyle \sin{}(\vartheta_{n+1}-\vartheta_n) A_n\hat\vartheta_n^\perp. \end{array} \end{aligned} $$
(A.8)

Hence,

$$\displaystyle \begin{aligned} \begin{array}{rcl} \|A_n\hat\vartheta_n^\perp\| |\sin{}(\vartheta_{n+1}-\vartheta_n)| &\displaystyle \simeq&\displaystyle \|A_n\hat\vartheta_{n+1}\|\\ &\displaystyle \leq&\displaystyle \| A_n A_{n+1}^{-1}\| \| A_{n+1} \hat\vartheta_{n+1}\| = \frac{\| A_{n+1}^{-1}A_n \|}{\| A_{n+1} \|}. \end{array} \end{aligned} $$
(A.9)

In the last step, we used that the A n’s are unimodular to switch from \(\| A_n A_{n+1}^{-1}\|\) to \(\| A_{n+1}^{-1} A_n \|\). By condition (ii), it is enough to show that \(\|A_n\hat \vartheta _n^\perp \| \gtrsim \|A_n\|\). For this, decompose

$$\displaystyle \begin{aligned} \begin{array}{rcl} \vartheta_n^\perp = \alpha_n \hat\vartheta_n^{\prime} + \beta_n \hat\vartheta_n, \end{array} \end{aligned} $$
(A.10)

for some coefficients such that α 2 + β 2 = 1. Note that, by construction, |α n| > |β n| and hence |α n|≥ 1∕2. Therefore,

$$\displaystyle \begin{aligned} \begin{array}{rcl} \| A_n\hat\vartheta_n^\perp \|{}^2 &\displaystyle =&\displaystyle \alpha^2 \| A_n \|{}^2 + \beta^2 \|A_n\|{}^{-2} \geq \frac 14 \|A_n\|{}^2. \end{array} \end{aligned} $$

This finishes the proof Part 1.

To prove Part 2, assume 𝜗 n → 𝜗 ∈ (0, π∕2) and apply (A.6) to 𝜗 = 𝜗 to obtain

$$\displaystyle \begin{aligned} \begin{array}{rcl} A_n\vartheta_{\infty} = \pm \| A_n \|{}^{-1} \cos{}(\vartheta_{\infty}-\vartheta_n) \hat\vartheta_n + \sin{}(\vartheta_{\infty}-\vartheta_n) A_n \hat\vartheta_n^{\perp}. \end{array} \end{aligned} $$
(A.11)

Recalling that ∥A n∥→, 𝜗 n − 𝜗 → 0 and \(\|A_n \hat \vartheta _n^\perp \| \asymp \|A_n\|\),

$$\displaystyle \begin{aligned} \begin{array}{rcl} \frac{\log \|A_n \vartheta_{\infty}\|}{\log \|A_n\|} \simeq \max \left\{ -1, \frac{\log |\sin{}(\vartheta_{\infty}-\vartheta_{n})|}{\log\|A_n\|}+1\right\}. \end{array} \end{aligned} $$
(A.12)

Hence, the left condition in (A.2) is satisfied if and only if

$$\displaystyle \begin{aligned} \begin{array}{rcl} \limsup_n \frac{\log |\sin{}(\vartheta_{\infty}-\vartheta_{n})|}{\log\|A_n\|} \leq -2. \end{array} \end{aligned} $$
(A.13)

But, disregarding multiplicative constants which will disappear in the limit,

$$\displaystyle \begin{aligned} \begin{array}{rcl} \sin{}(\vartheta_{\infty}-\vartheta_n) \asymp \tan{}(\vartheta_{\infty}-\vartheta_n) \asymp \tan\vartheta_{\infty}-\tan\vartheta_n = r_{\infty} - r_n. \end{array} \end{aligned} $$
(A.14)

We apply this with A n = T ω,n to prove Proposition 5.5. Define

$$\displaystyle \begin{aligned} \begin{array}{rcl} \begin{pmatrix} x^{(1)}_{n+1} \\ x^{(1)}_n \end{pmatrix} = {\mathbf{T}}_{\omega,n} \begin{pmatrix} 1 \\ 0 \end{pmatrix} \quad {\text and} \quad \begin{pmatrix} x^{(2)}_{n+1} \\ x^{(2)}_n1 \end{pmatrix} = {\mathbf{T}}_{\omega,n} \begin{pmatrix} 0 \\ 1 \end{pmatrix} \end{array} \end{aligned} $$
(A.15)

and let \(R^{(i)}_n,\, \theta ^{(i)}_n,\, n\geq 1\), i = 1, 2 be the corresponding Prüfer radii and phases. We let \(r_n=R^{(1)}_n/R^{(2)}_n\) and 𝜗 n be as in Lemma A.2. Then it follows from (3.5) and some elementary trigonometry that

$$\displaystyle \begin{aligned} \begin{array}{rcl} x^{(1)}_{n+1}x^{(2)}_n-x^{(1)}_nx^{(2)}_{n+1}=R^{(1)}_n R^{(2)}_n \sin k \sin{}(\theta^{(2)}_n-\theta^{(1)}_n), \end{array} \end{aligned} $$
(A.16)

where \(\bar {\theta }^{(i)}_n = nk -\theta ^{(i)}_n\). On the other hand,

$$\displaystyle \begin{aligned} \begin{array}{rcl} x^{(1)}_{n+1}x^{(2)}_n-x^{(1)}_nx^{(2)}_{n+1} &\displaystyle =&\displaystyle \det\left( {\mathbf{T}}_{\omega,n} \begin{pmatrix} 1 &\displaystyle 0 \\ 0 &\displaystyle 1 \end{pmatrix} \right) =1. \end{array} \end{aligned} $$
(A.17)

This, together with the convergence

$$\displaystyle \begin{aligned} \begin{array}{rcl} \lim_{n\to\infty} \frac{R^{(i)}_n}{\log n} = \beta,\quad i=1,2, \end{array} \end{aligned} $$
(A.18)

gives

$$\displaystyle \begin{aligned} \begin{array}{rcl}{} \lim_{n\to\infty}\frac{\log |\sin{}(\theta^{(2)}_n-\theta^{(1)}_n)|}{\log n} =\lim_{n\to\infty}\frac{\log |\sin{}(\bar\theta^{(2)}_n-\bar\theta^{(1)}_n)|}{\log n} = -2\beta. \end{array} \end{aligned} $$
(A.19)

Remember the decomposition (5.10). We have to estimate the difference of the expansions for \(\log R^{(1)}_n\) and \(\log R^{(2)}_n\). By (A.19), \(|\sin {}(\bar \theta ^{(2)}_n-\bar \theta ^{(1)}_n)| \lesssim n^{-\beta +\epsilon }\), for any 𝜖 > 0. Hence, there exists random sequences and such that \(\bar \theta ^{(1)}_n-\bar \theta ^{(2)}_n=m_n\pi + \Delta _n\) and \(|\Delta _n|\lesssim n^{-\beta +\epsilon }\). Therefore,

$$\displaystyle \begin{aligned} \sin{}(2\bar\theta^{(2)}_n) = \sin{}(2\bar\theta^{(1)}_n+2\Delta_n) \simeq \sin{}(2\bar\theta^{(1)}_n)+2\cos{}(2\bar\theta^{(1)}_n)\Delta_n. \end{aligned} $$
(A.20)

This shows that

$$\displaystyle \begin{aligned} \begin{array}{rcl} \left|V_j \left(\sin{}(2\bar{\theta}^{(1)}_j) - \sin{}(2\bar{\theta}^{(2)}_j)\right)\right| &\displaystyle \lesssim&\displaystyle j^{-\frac{1}{2}-2\beta + \epsilon},\\ \text{and} \quad \left|V_j^2 \left(\cos^2(\bar{\theta}^{(1)}_j) - \cos^2(\bar{\theta}^{(2)}_j)\right)\right| &\displaystyle \lesssim&\displaystyle j^{-1-2\beta + \epsilon}, \end{array} \end{aligned} $$

by a similar argument. Hence,

$$\displaystyle \begin{aligned} \begin{array}{rcl}{} \log r_n &\displaystyle =&\displaystyle -\sum^n_{j=1} \frac{V_{\omega,j}}{\sin k} \left(\sin{}(2\bar{\theta}^{(1)}_j) - \sin{}(2\bar{\theta}^{(2)}_j) \right) + \sum^n_{j=1} A_j \end{array} \end{aligned} $$
(A.21)

where the first sum is a convergent martingale by Lemma 5.3 with \(\gamma =\frac 12 + 2\beta -\epsilon \) and the second one is absolutely convergent as A j = O(j −1−2β+𝜖). This shows that r n → r ∈ (0, ) almost surely which implies that 𝜗 n has a limit 𝜗 ≠ 0, π∕2 by the first part of Lemma A.2.

The equivalence (A.2) in our context corresponds to

$$\displaystyle \begin{aligned} \begin{array}{rcl} \lim_{n\to\infty}\frac{\log R_n( \vartheta_{\infty})}{\log n}=-\beta \quad \text{if and only if} \quad \limsup_n \frac{\log |r_n-r_{\infty}|}{\log n} \leq -2\beta.\qquad \end{array} \end{aligned} $$
(A.22)

Let us denote by \(\log r_n = M_n + S_n\) the decomposition (A.21) and \(\log r_{\infty }=M_{\infty }+S_{\infty }\) where M and S are the almost sure limits of M n and S n respectively. Then,

$$\displaystyle \begin{aligned} \begin{array}{rcl} |r_{\infty}-r_n| &\displaystyle =&\displaystyle e^{M_{\infty}+S_{\infty}}\left| 1-\mathrm{e}^{M_n-M_{\infty}+S_n-S_{\infty}}\right| \simeq \mathrm{e}^{M_{\infty}+S_{\infty}} \end{array} \end{aligned} $$
(A.23)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle \left| M_n-M_{\infty}+S_n-S_{\infty}\right| \lesssim \mathrm{e}^{M_{\infty}+S_{\infty}} n^{-2\beta+2\epsilon}, \end{array} \end{aligned} $$
(A.24)

by the last statement of Lemma 5.3 with \(\gamma = \frac 12 + 2\beta - \epsilon \). This finishes the proof of Proposition 5.5.

1.2 A.2 Analysis of the Prüfer Phases

We begin with a simple observation: from (3.6),

$$\displaystyle \begin{aligned} \begin{array}{rcl} \left| e^{i(\theta_{n+1}-\theta_n)}-1\right| = \left| \frac{\rho_{n+1}}{\rho_n}-1\right|\leq \frac{| V_{\omega,a}|}{|\sin k|}\leq 1, \end{array} \end{aligned} $$
(A.25)

for n large enough. Hence, for n large enough, |θ n+1 − θ n| < π∕2 and

$$\displaystyle \begin{aligned} \begin{array}{rcl} |\theta_{n+1}-\theta_n| \leq \frac{\pi}{2} |\sin{}(\theta_{n+1}-\theta_n)| \leq \frac{\pi}{2} \left|e^{i(\theta_{n+1}-\theta_n)}-1\right| \lesssim n^{-\alpha}. \end{array} \end{aligned} $$
(A.26)

This can be written in the equivalent form

$$\displaystyle \begin{aligned} \begin{array}{rcl}{} |\bar{\theta}_{n+1}-\bar{\theta}_n-k|\leq c_0 n^{-\alpha}, \end{array} \end{aligned} $$
(A.27)

for some c 0 > 0, which will be more suitable for our purposes.

The next lemma provides the control of the Prüfer phases needed to complete the proof of Proposition 5.2.

Lemma A.3

Let \(0<\alpha \leq \frac 12\). Let E ∈ (−2, 2) corresponding to values of k different from \(\frac {\pi }{4}\), \(\frac {\pi }{2}\) and \(\frac {3\pi }{4}\). Then,

(A.28)

Proof

The key to the proof is [31, Lemma 8.5] which states: suppose that is not in . Then, there exists a sequence of integers q l → such that

$$\displaystyle \begin{aligned} \begin{array}{rcl}{} \left|\sum^{q_l}_{j=1} \cos \theta_j \right| \leq 1 + \sum^{q_l}_{j=1} \left| \theta_j-\theta_0 - jy \right|, \end{array} \end{aligned} $$
(A.29)

for all .

We will treat the term with \(\cos {}(4\bar {\theta }_j)\) as the other one is similar. We will take y = 4k above. Let n be large enough so that it can be written as n = n 0 + Kq l with \(n_0\ge q_l^2\) and \(4c_0 n_0^{-\alpha }\le q_l^{-2} \) (where c 0 is the constant from (A.27)). Then,

$$\displaystyle \begin{aligned} \begin{array}{rcl} \left|\sum_{j=n_0+1}^n j^{-2\alpha}\cos{}(4\bar\theta_j)\right| &\displaystyle =&\displaystyle \left|\sum_{m=0}^K \sum_{r=1}^{q_l}(n_0+mq_l+r)^{-2\alpha}\cos{}(4\bar\theta(n_0+mq_l+r))\right| \end{array} \end{aligned} $$
(A.30)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle \leq&\displaystyle \sum_{m=0}^K(n_0+mq_l)^{-2\alpha} \left|\sum_{r=1}^{q_l}\cos{}(4\bar\theta(n_0+mq_l+r))\right| \end{array} \end{aligned} $$
(A.31)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle +&\displaystyle \sum_{m=0}^K \sum_{r=1}^{q_l} \left|(n_0+mq_l+r)^{-2\alpha}-(n_0+mq_l)^{-2\alpha}\right| \end{array} \end{aligned} $$
(A.32)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle =:&\displaystyle A + B. \end{array} \end{aligned} $$
(A.33)

By (A.29),

$$\displaystyle \begin{aligned} \begin{array}{rcl} A &\displaystyle \leq&\displaystyle \sum_{m=0}^{K}(n_0+mq_l)^{-2\alpha}\left(1+4\sum_{r=1}^{q_l}|\bar\theta(n_0+mq_l+r)-\bar\theta(n_0+mq_l)-kr|\right). \end{array} \end{aligned} $$
(A.34)

Now, by (A.27),

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle 4\sum_{r=1}^{q_l}|\bar\theta(n_0+mq_l+r)-\bar\theta(n_0+mq_l)-kr| \leq c_0 \sum_{r=1}^{q_l}\sum_{s=1}^r(n_0+mq_l+r)^{-\alpha} \end{array} \end{aligned} $$
(A.35)
(A.36)

Thus,

$$\displaystyle \begin{aligned} \begin{array}{rcl} A \leq 2 \sum_{m=0}^{K}(n_0+mq_l)^{-2\alpha} \leq 2 q_l^{-2\alpha} \sum_{m=0}^{K}(n_0q_l^{-1}+m)^{-2\alpha} \leq c_1 q_l^{-2\alpha} K^{1-2\alpha} \leq c_1 q_l^{-1} n^{1-2\alpha}, \end{array} \end{aligned} $$

for some finite c 1 > 0. To estimate B, we use that

$$\displaystyle \begin{aligned} \begin{array}{rcl} \left| (n_0+mq_l+r)^{-2\alpha}-(n_0+mq_l)^{-2\alpha}\right| \leq c_2 (n_0+mq_l)^{-2\alpha-1} r, \end{array} \end{aligned} $$
(A.37)

for some finite c 2 > 0 which shows that

$$\displaystyle \begin{aligned} \begin{array}{rcl} B &\displaystyle \leq&\displaystyle c_2 \sum_{m=0}^K \sum_{r=1}^{q_l} (n_0+mq_l)^{-2\alpha-1} r \leq c_2 q_l^2 n_0^{-1} \sum_{m=0}^K (1+n_0^{-1}mq_l)^{-1}(n_0+mq_l)^{-2\alpha}\qquad \end{array} \end{aligned} $$
(A.38)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle \leq&\displaystyle c_2 \sum_{m=0}^K (n_0+mq_l)^{-2\alpha},\qquad \end{array} \end{aligned} $$
(A.39)

where we used \(q_l^2 n_0^{-1}\leq 1\). This last sum can be estimated as above. Summarizing,

$$\displaystyle \begin{aligned} \begin{array}{rcl} \left|\sum^{q_l}_{j=1} \cos \theta_j \right| \leq c_3 q_l^{-1} n^{1-2\alpha}, \end{array} \end{aligned} $$
(A.40)

for some finite c 3 > 0. This finishes the proof. □

Rights and permissions

Reprints and permissions

Copyright information

© 2020 The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG

About this paper

Check for updates. Verify currency and authenticity via CrossMark

Cite this paper

Bourget, O., Flores, G.R.M., Taarabt, A. (2020). One-dimensional Discrete Anderson Model in a Decaying Random Potential: from A.C. Spectrum to Dynamical Localization. In: Miranda, P., Popoff, N., Raikov, G. (eds) Spectral Theory and Mathematical Physics. Latin American Mathematics Series(). Springer, Cham. https://doi.org/10.1007/978-3-030-55556-6_3

Download citation

Publish with us

Policies and ethics