Skip to main content
Log in

Absolutely continuous spectrum for Schrödinger operators with random decaying matrix potentials on the strip

  • Published:
Letters in Mathematical Physics Aims and scope Submit manuscript

Abstract

We consider a family of random Schrödinger operators on the discrete strip with decaying random \(\ell ^2\) matrix potential. We prove that the spectrum is almost surely pure absolutely continuous, apart from random, possibly embedded, eigenvalues which may accumulate at band edges.

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. Anderson, P.W.: Absence of diffusion in certain random lattices. Phys. Rev. 109, 1492–1505 (1958). https://doi.org/10.1103/PhysRev.109.1492

    Article  ADS  Google Scholar 

  2. Aizenman, M., Sims, R., Warzel, S.: Stability of the absolutely continuous spectrum of random Schrödinger operators on tree graphs. Prob. Theor. Rel. Fields 136, 363–394 (2006). https://doi.org/10.1007/s00440-005-0486-8

    Article  MATH  Google Scholar 

  3. Aizenman, M., Molchanov, S.: Localization at large disorder and extreme energies: an elementary derivation. Commun. Math. Phys. 157, 245–278 (1993). https://doi.org/10.1007/BF02099760

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Aizenman, M., Warzel, S.: Resonant delocalization for random Schrödinger operators on tree graphs. J. Eur. Math. Soc. 15(4), 1167–1222 (2013). https://doi.org/10.4171/JEMS/389

    Article  MathSciNet  MATH  Google Scholar 

  5. Bourgain, J., Kenig, C.: On localization in the continuous Anderson-Bernoulli model in higher dimension. Inv. Math. 161, 389–426 (2005). https://doi.org/10.1007/s00222-004-0435-7

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Carmona, R.: One-dimensional Schrödinger operators with random or deterministic potentials, new spectral types. J. Funct. Anal. 51, 229–258 (1983). https://doi.org/10.1016/0022-1236(83)90027-7

    Article  MathSciNet  MATH  Google Scholar 

  7. Carmona, R., Lacroix, J.: Spectral Theory of Random Schrödinger Operator. Birkhäuser, Boston (2013)

    MATH  Google Scholar 

  8. Carmona, R., Klein, A., Martinelli, F.: Anderson localization for Bernoulli and other singular potentials. Commun. Math. Phys. 108, 41–66 (1987). https://doi.org/10.1007/BF01210702

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Deift, P., Killip, R.: On the absolutely continuous spectrum of one-dimensional Schrödinger operators with square summable potentials. Commun. Math. Phys. 203, 341–347 (1999). https://doi.org/10.1007/s002200050615

    Article  ADS  MATH  Google Scholar 

  10. Froese, R., Hasler, D., Spitzer, W.: Transfer matrices, hyperbolic geometry and absolutely continuous spectrum for some discrete Schrödinger operators on graphs. J. Funct. Anal. 230, 184–221 (2006). https://doi.org/10.1016/j.jfa.2005.04.004

    Article  MathSciNet  MATH  Google Scholar 

  11. Froese, R., Hasler, D., Spitzer, W.: Absolutely continuous spectrum for the Anderson Model on a tree: A geometric proof of Klein’s Theorem. Commun. Math. Phys. 269, 239–257 (2007). https://doi.org/10.1007/s00220-006-0120-3

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Froese, R., Hasler, D., Spitzer, W.: On the ac spectrum of one-dimensional random Schrödinger operators with matrix-valued potentials. Math. Phys. Anal. Geom. 13, 219–233 (2010). https://doi.org/10.1007/s11040-010-9076-9

    Article  MathSciNet  MATH  Google Scholar 

  13. Fröhlich, J., Spencer, T.: Absence of difusion in the Anderson tight binding model for large disorder or low energy. Commun. Math. Phys. 88, 151–184 (1983). https://doi.org/10.1007/BF01209475

    Article  ADS  MATH  Google Scholar 

  14. Germinet, F., Klein, A.: Bootstrap multiscale analysis and localization in random media. Commun. Math. Phys. 222, 415–448 (2001). https://doi.org/10.1007/s002200100518

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Germinet, F., Klein, A.: A comprehensive proof of localization for continuous Anderson models with singular random potentials. J. Eur. Math. Soc. 15, 53–143 (2013). https://doi.org/10.4171/JEMS/356

    Article  MathSciNet  MATH  Google Scholar 

  16. Gol’dsheid, Y., Molchanov, S., Pastur, L.: Pure point spectrum of stochastic one dimensional Schrödinger operators. Funct. Anal. Appl. 11, 1–10 (1977). https://doi.org/10.1007/BF01135526

    Article  Google Scholar 

  17. Keller, M., Lenz, D., Warzel, S.: Absolutely continuous spectrum for random operators on trees of finite cone type. J. D’ Anal. Math. 118, 363–396 (2012). https://doi.org/10.1007/s11854-012-0040-4

    Article  MathSciNet  MATH  Google Scholar 

  18. Killip, R., Simon, B.: Sum rules for Jacobi matrices and their applications to spectral theory. Ann. Math. 158, 253–321 (2003). https://doi.org/10.4007/annals.2003.158.253

    Article  MathSciNet  MATH  Google Scholar 

  19. Kiselev, A., Last, Y., Simon, B.: Modified Prüfer and EFGP transforms and the spectral analysis of one-dimensional Schrödinger operators. Commun. Math. Phys. 194, 1–45 (1997). https://doi.org/10.1007/s002200050346

    Article  ADS  MATH  Google Scholar 

  20. Klein, A.: Extended states in the Anderson model on the Bethe lattice. Adv. Math. 133, 163–184 (1998). https://doi.org/10.1006/aima.1997.1688

    Article  MathSciNet  MATH  Google Scholar 

  21. Klein, A., Lacroix, J., Speis, A.: Localization for the Anderson model on a strip with singular potentials. J. Funct. Anal. 94, 135–155 (1990). https://doi.org/10.1016/0022-1236(90)90031-F

    Article  MathSciNet  MATH  Google Scholar 

  22. Klein, A., Sadel, C.: Absolutely continuous spectrum for random Schrödinger operators on the Bethe strip. Math. Nachr. 285, 5–26 (2012). https://doi.org/10.1002/mana.201100019

    Article  MathSciNet  MATH  Google Scholar 

  23. Kunz, H., Souillard, B.: Sur le spectre des operateurs aux differences finies aleatoires. Commun. Math. Phys. 78, 201–246 (1980). https://doi.org/10.1007/bf01942371

    Article  ADS  MATH  Google Scholar 

  24. Last, Y., Simon, B.: Eigenfunctions, transfer matrices, and absolutely continuous spectrum of one-dimensional Schrödinger operators. Invent. Math. 135(2), 329–367 (1999). https://doi.org/10.1007/s002220050288

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Li, L.: Anderson-Bernoulli localization with large disorder on the 2D lattice. Commun. Math. Phys. 393, 151–214 (2022). https://doi.org/10.1007/s00220-022-04366-1

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Li, L., Zhang, L.: Anderson-Bernoulli Localization on the 3D lattice and discrete unique continuation principle. Duke Math. J. 17, 327–415 (2022). https://doi.org/10.1215/00127094-2021-0038

    Article  MATH  Google Scholar 

  27. Sadel, C.: Relations between transfer and scattering matrices in the presence of hyperbolic channels. J. Math. Phys. 52, 1235111 (2011). https://doi.org/10.1063/1.3669483

    Article  MathSciNet  MATH  Google Scholar 

  28. Sadel, C.: Absolutely continuous spectrum for random Schrödinger operators on tree-strips of finite cone type. Ann. Henri Poincaré 14, 737–773 (2013). https://doi.org/10.1007/s00023-012-0203-y

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Sadel, C.: Absolutely continuous spectrum for random Schrödinger operators on the Fibbonacci and similar tree-strips. Math. Phys. Anal. Geom. 17, 409–440 (2014). https://doi.org/10.1007/s11040-014-9163-4

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Sadel, C.: Anderson transition at two-dimensional growth rate on antitrees and spectral theory for operators with one propagating channel. Ann. Henri Poincare 17, 1631–1675 (2016). https://doi.org/10.1007/s00023-015-0456-3

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Sadel, C.: Spectral theory of one-channel operators and application to absolutely continuous spectrum for Anderson type models. J. Funct. Anal. 274(8), 2205–2244 (2018). https://doi.org/10.1016/j.jfa.2018.01.017

    Article  MathSciNet  MATH  Google Scholar 

  32. Sadel, C.: Transfer matrices for discrete Hermitian operators and absolutely continuous spectrum. J. Funct. Anal. (2021). https://doi.org/10.1016/j.jfa.2021.109151

    Article  MathSciNet  MATH  Google Scholar 

  33. Sadel, C., Virág, B.: A central limit theorem for products of random matrices and GOE statistics for the Anderson model on long boxes. Commun. Math. Phys. 343(3), 881–919 (2016). https://doi.org/10.1007/s00220-016-2600-4

    Article  ADS  MathSciNet  MATH  Google Scholar 

  34. Tautenhahn, M.: Localization criteria for Anderson Models on locally finite graphs. J. Stat. Phys. 144, 60–75 (2011). https://doi.org/10.1007/s10955-011-0248-1

    Article  ADS  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This work has been supported by the Chilean Grants FONDECYT Nr. 1201836 and the Nucleo Mileneo MESCD.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian Sadel.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

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

Appendix A. Transfer matrices and spectral averaging formula on the strip

Appendix A. Transfer matrices and spectral averaging formula on the strip

As above, we consider operators of the form

$$\begin{aligned} (H\Psi )_n=-\Psi _{n-1}-\Psi _{n+1}+B _n \Psi _n \end{aligned}$$

on \(\ell ^2({{\mathbb {Z}}}_+)\otimes {{\mathbb {C}}}^l\). Solving the eigenvalue equation \(H\Psi =z\Psi \) leads to the transfer matrices

$$\begin{aligned} T^z_n=\begin{pmatrix} B_n-zI &{}\quad -I \\ I &{}\quad \textbf{0} \end{pmatrix} \quad \text {and the equation}\quad \begin{pmatrix} \Psi _{n+1} \\ \Psi _n \end{pmatrix}=T^z_n \begin{pmatrix} \Psi _n \\ \Psi _{n-1} \end{pmatrix} \end{aligned}$$

Then, for \(n>m\) we define the products

$$\begin{aligned} T^z_{m,n}=T^z_n T^z_{n-1} \cdots T^z_{m+1} T^z_m \quad \text {leading to}\quad \begin{pmatrix} \Psi _{n+1} \\ \Psi _n \end{pmatrix}=T^z_{m,n} \begin{pmatrix} \Psi _m \\ \Psi _{m-1} \end{pmatrix} \end{aligned}$$

for a formal solution of \(H\Psi =z\Psi \).

1.1 A.1. Transfer matrices and resolvent boundary data

Let \(n>m\), be non-negative integers. With \(H_{m,n}\), we denote the restriction of H to \(\ell ^2(\{m,m+1,\ldots ,n\}) \otimes {{\mathbb {C}}}^l\), that is

$$\begin{aligned} H_{m,n}=\begin{pmatrix} B_m &{}\quad -I &{}\quad \\ -I &{}\quad B_{m+1} &{}\quad -I \\ {} &{}\quad \ddots &{}\quad \ddots &{}\quad \ddots \\ {} &{} &{} \quad \ddots &{}\quad \ddots &{}\quad -I \\ {} &{} &{} &{}\quad -I &{}\quad B_n \end{pmatrix} \end{aligned}$$

Then, we define the m to n boundary resolvent data for \(z \not \in \sigma (H_{m,n})\) by

$$\begin{aligned} \begin{pmatrix} \alpha ^z_{m,n} &{}\quad \beta ^z_{m,n} \\ \gamma ^z_{m,n} &{}\quad \delta ^z_{m,n} \end{pmatrix}= \begin{pmatrix} P_m^* \\ P_n^* \end{pmatrix} (H_{m,n}-z)^{-1} \begin{pmatrix} P_m&\quad P_n \end{pmatrix} \end{aligned}$$
(A.1)

where \(P_k\) is the natural embedding of \(\ell ^2(\{k\})\otimes {{\mathbb {C}}}^l\) into \(\ell ^2(\{m,m+1,\ldots ,n\})\otimes {{\mathbb {C}}}^l\) for \(m\le k \le n\). This means, e.g., \(\alpha ^z_{m,n}=P_m^* (H_{m,n}-z)^{-1} P_m\), and in this setup

$$\begin{aligned} P_m=\begin{pmatrix} I\\ \textbf{0}\\ \vdots \\ \textbf{0} \end{pmatrix}\,,\quad P_n=\begin{pmatrix} \textbf{0}\\ \vdots \\ \textbf{0}\\ I \end{pmatrix}\quad \in \;\;{{\mathbb {C}}}^{(n-m+1)l \times l}. \end{aligned}$$

Note that \(\alpha ^z_{m,n}, \beta ^z_{m,n}, \gamma ^z_{m,n}, \delta ^z_{m,n}\) are all \(l\times l\) matrices.

Proposition A.1

Let be given \(n\ge m\in {{\mathbb {Z}}}_+\) and let \(z\not \in \sigma (H_{m,n})\) and let \(\beta ^z_{m,n}\) be invertible. Then,

$$\begin{aligned} T^z_{m,n}\,=\, \begin{pmatrix} (\beta ^z_{m,n})^{-1} &{} - (\beta ^z_{m,n})^{-1}\alpha ^z_{m,n}\\ \delta ^z_{m,n} (\beta ^z_{m,n})^{-1} &{} \gamma ^z_{m,n} - \delta ^z_{m,n} (\beta ^z_{m,n})^{-1} \alpha ^z_{m,n} \end{pmatrix} \end{aligned}$$

Proof

For \(\Psi =(\Psi _n)_n\) with \(\Psi _n \in {{\mathbb {C}}}^l\), we define the notations:

$$\begin{aligned} {\hat{\Psi }}_k:= \Psi _k, \quad k<m. \quad \quad \quad {\hat{\Psi }}_m:= \left( \begin{array}{cc} \Psi _m \\ \Psi _{m+1} \\ \vdots \\ \Psi _n \end{array} \right) , \qquad {\hat{\Psi }}_{m+1}: = \Psi _{n+1}, \end{aligned}$$

and we use \(P_m\) and \(P_n\) as in (A.1), then we have

$$\begin{aligned} \Psi _{m}=P_m^* {\hat{\Psi }}_m, \quad \Psi _n=P_n^* {\hat{\Psi }}_m . \end{aligned}$$

Moreover, let \(\Phi =H\Psi \) and introduce similar notations \({\hat{\Phi }}_k\), then we get

$$\begin{aligned} {\hat{\Phi }}_m=H_{m,n}{\hat{\Psi }}_m-P_m {\hat{\Psi }}_{m-1}-P_n {\hat{\Psi }}_{m+1}. \end{aligned}$$

With z being the spectral parameter, \({\hat{\Phi }}_m = z {\hat{\Psi }}_m\) leads to

$$\begin{aligned} P_n {\hat{\Psi }}_{m+1} = (H_{m,n}-z){\hat{\Psi }}_m-P_m {\hat{\Psi }}_{m-1} \end{aligned}$$

Multiplying with \(P_m^*(H_{m,n}-z)^{-1}\) from the left, noting that \({\hat{\Psi }}_{m+1}=\Psi _{n+1}\) and using (A.1) gives

$$\begin{aligned} \beta ^z_{m,n} \Psi _{n+1} =\Psi _m-\alpha ^z_{m,n} \Psi _{m-1} \implies \Psi _{n+1} =(\beta ^z_{m,n})^{-1}\Psi _m-(\beta ^z_{m,n})^{-1}\alpha ^z_{m,n} \Psi _{m-1} \end{aligned}$$

Multiplying from the left with \(P_n^*(H_{m,n}-z)^{-1}\) instead of \(P_m^*(H_{m,n}-z)^{-1})\) leads to

$$\begin{aligned} \delta ^z_{m,n} \Psi _{n+1}=\Psi _n-\gamma ^z_{m,n} \Psi _{m-1} \end{aligned}$$

Replacing \(\Psi _{n+1}\) with the formula above and resolving for \(\Psi _n\) leads to

$$\begin{aligned} \Psi _n=\delta ^z_{m,n} (\beta ^z_{m,n})^{-1}\Psi _m+(\gamma ^z_{m,n} -\delta ^z_{m,n} (\beta ^z_{m,n})^{-1} \alpha ^z_{m,n})\Psi _{m-1} \end{aligned}$$

Finally, we have:

$$\begin{aligned} \begin{pmatrix} \Psi _{n+1} \\ \Psi _{n} \end{pmatrix}=\begin{pmatrix} (\beta ^z_{m,n})^{-1} &{}\quad - (\beta ^z_{m,n})^{-1}\alpha ^z_{m,n}\\ \delta ^z_{m,n} (\beta ^z_{m,n})^{-1} &{}\quad \gamma ^z_{m,n} - \delta ^z_{m,n} (\beta ^z_{m,n})^{-1} \alpha ^z_{m,n} \end{pmatrix} \begin{pmatrix} \Psi _m \\ \Psi _{m-1} \end{pmatrix} \end{aligned}$$

As \(\Psi _m, \Psi _{m-1}\) determine the solution to \(H\Psi =z\Psi \) uniquely, the matrix must be \(T^z_{m,n}\)\(\square \)

1.2 A.2. Spectral averaging formula

Here, we state the strip-equivalent of the spectral average formula from Carmona-Lacroix [7, Theorem III.3.2 and III.3.6]. It is a special case of [32, Theorem 1]. First, we need to fix a vector in the root-slice. Thus, we choose some \({\vec {x}}\in {{\mathbb {C}}}^l\) which we identify with \(\delta _0 \otimes {\vec {x}}\in \ell ^2\{{{\textbf{Z}}}_+\} \otimes {{\mathbb {C}}}^l\). Let us assume that \(\Vert {\vec {x}}\Vert =1\), so that \({{\vec {x}}}^* {\vec {x}}= 1\). Furthermore, identifying \({\vec {x}}\,^*\) with a linear map from \({{\mathbb {C}}}^l\) to \({{\mathbb {C}}}\), we have a \(l-1\)-dimensional kernel consisting of the vectors orthogonal to \({\vec {x}}\),

$$\begin{aligned} {{\mathbb {K}}}:=\hbox {ker}({\vec {x}}\,^*) = \{{\vec v}\in {{\mathbb {C}}}^l:\,{\vec {x}}\,^* {\vec v}= 0 \} \,=\,\{{\vec v}\in {{\mathbb {C}}}^l:\, {\vec {x}}\cdot {\vec v}= 0\}. \end{aligned}$$

Then, in this special case, the work of [32] simply replaces \(T^z_0\) by the set of \(2l \times 2\) matrices

$$\begin{aligned} {{\mathbb {T}}}^z_0\,=\, \left\{ \begin{pmatrix} (B_n-zI) ({\vec {x}}+{\vec v}) &{}\quad - {\vec {x}}+(B_n-zI) {\vec w}\\ {\vec {x}}+{\vec v}&{}\quad {\vec w} \end{pmatrix}\,:\, {\vec v},{\vec w}\in {{\mathbb {K}}}\; \right\} \;\subset \; {{\mathbb {C}}}^{2l \times 2}\;. \end{aligned}$$
(A.2)

Note that

$$\begin{aligned} T^z_0 = \begin{pmatrix} B_0 -zI &{}\quad -I \\ I &{}\quad \textbf{0} \end{pmatrix} \quad \text {and}\quad {{\mathbb {T}}}^z_0 = T^z_0\,\left\{ \begin{pmatrix} {\vec {x}}+{\vec v}&{}\quad {\vec w}\\ \textbf{0}&{}\quad {\vec {x}} \end{pmatrix}\,:\, {\vec v},{\vec w}\in {{\mathbb {K}}}\right\} \end{aligned}$$
(A.3)

where we adopt the notation that \(T{{\mathbb {A}}}=\{TA:\, A \in {{\mathbb {A}}}\}\) for sets of matrices \({{\mathbb {A}}}\).

Moreover, we consider the spectral measure \(\mu _{{\vec {x}}}\) at the vector \({\vec {x}}\equiv \delta _0 \otimes {\vec {x}}\), that means

$$\begin{aligned} \int f d\mu _{{\vec {x}}}\,=\, \langle \delta _0 \otimes {\vec {x}}, \, f(H)\,(\delta _0 \otimes {\vec {x}})\, \rangle . \end{aligned}$$

Now, using that the operator H cannot have compactly (finitely) supported eigenfunctions, Theorem 1 in [32] implies the following:

Proposition A.2

[32] In the sense of a weak limit for finite measures, one finds that

$$\begin{aligned} \textrm{d}\mu _{{\vec {x}}}(E)\,&=\,\lim _{n\rightarrow \infty } \,\frac{1}{\pi }\, \frac{\textrm{d} E}{ \min \limits _{T^E \in {{\mathbb {T}}}^E_0}\left\| T^E_{1,n} \, T^E \left( \begin{array}{cc} 1 \\ 0 \end{array} \right) \right\| ^2}\,=\, \lim _{n\rightarrow \infty } \frac{1}{\pi }\, \frac{\textrm{d}E}{\min \limits _{\vec v \in {{\mathbb {K}}}} \left\| T^E_{0,n} \left( \begin{array}{cc} {\vec {x}}+\vec v \\ 0 \end{array} \right) \right\| ^2 } \end{aligned}$$

Using the symplectic structure of the transfer matrices and the Banach–Alaoglu theorem, one can obtain a criterion for absolute continuity (see [32]).

Proposition A.3

If one finds \(\vec {u}_{E,n} \in {{\mathbb {C}}}^m\) for \(E \in (a,b)\), \(n \in {{\mathbb {N}}}\), such that

$$\begin{aligned} \liminf _{n \rightarrow \infty } \;\int _a^b\, \left\| \,T^E_{0,n} \left( \begin{array}{cc} \vec {u}_{E,n} \\ {\vec {x}} \end{array} \right) \,\right\| ^4\;\textrm{d}E\,<\,\infty \end{aligned}$$

then, the measure \(\mu _{{\vec {x}}}\) is absolutely continuous in the interval (ab).

Proof

First, in [32] it was shown that the minimum \(\min \limits _{\vec v \in {{\mathbb {K}}}} \left\| T^E_{0,n} \left( \begin{array}{cc} {\vec {x}}+\vec v \\ 0 \end{array} \right) \right\| \) is achieved at a very specific vector which we call \(\vec {v}_{E,n}\in {{\mathbb {K}}}\). Defining

$$\begin{aligned} f_n(E):= \pi ^{-1} \left\| T^E_{0,n} \left( \begin{array}{cc} {\vec {x}}+\vec v_{E,n} \\ 0 \end{array} \right) \right\| ^{-2} \end{aligned}$$

we see from Proposition A.2 that \(\mu _{{\vec {x}}}\) is the weak limit of \(f_n(E) \textrm{d}E\) in the interval (ab). Note that

$$\begin{aligned}&\left( T^E_{0,n} \left( \begin{array}{cc} \vec {u}_{E,n} \\ {\vec {x}} \end{array} \right) \right) ^* \left( \begin{array}{cc} \textbf{0}&{} -I \\ I &{} \textbf{0} \end{array} \right) T^E_{0,n} \left( \begin{array}{cc} {\vec {x}}+ \vec v_{E,n} \\ 0 \end{array} \right) \\&\quad =\, \left( \begin{array}{cc} \vec {u}^*_{E,n}&{\vec {x}}\,^* \end{array} \right) \, \big (T^E_{0,n} \big )^* \left( \begin{array}{cc} \textbf{0}&{} -I \\ I &{} \textbf{0} \end{array} \right) T^E_{0,n} \left( \begin{array}{cc} {\vec {x}}+ \vec v_{E,n} \\ 0 \end{array} \right) \\&\quad =\, \left( \begin{array}{cc} \vec {u}^*_{E,n}&{\vec {x}}\,^* \end{array} \right) \left( \begin{array}{cc} \textbf{0}&{} -I \\ I &{} \textbf{0} \end{array} \right) \left( \begin{array}{cc} {\vec {x}}+ \vec v_{E,n} \\ 0 \end{array} \right) \, =\, \left( \begin{array}{cc} \vec {u}^*_{E,n}&{\vec {x}}\,^* \end{array} \right) \left( \begin{array}{cc} 0 \\ {\vec {x}}+ \vec v_{E,n} \end{array} \right) \,=\,1 \end{aligned}$$

where we use \(\Vert {\vec {x}}\Vert =1\) and \({\vec {x}}\,^* \vec v_{E,n}=0\) as \(\vec v_{E,n} \in {{\mathbb {K}}}\). Now, using the Cauchy–Schwartz inequality, this gives

$$\begin{aligned} 1\,\le \, \left\| \,T^E_{0,n} \left( \begin{array}{cc} \vec {u}_{E,n} \\ {\vec {x}} \end{array} \right) \,\right\| \; \cdot \; \left\| T^E_{0,n} \left( \begin{array}{cc} {\vec {x}}+\vec v_{E,n} \\ 0 \end{array} \right) \right\| \end{aligned}$$

and hence

$$\begin{aligned} \pi ^2 |f_n(E)|^2\,=\,\frac{1}{\left\| T^E_{0,n} \left( \begin{array}{cc} {\vec {x}}+\vec v_{E,n} \\ 0 \end{array} \right) \right\| ^4}\,\le \, \left\| \,T^E_{0,n} \left( \begin{array}{cc} \vec {u}_{E,n} \\ {\vec {x}} \end{array} \right) \,\right\| ^4. \end{aligned}$$

Thus, the estimate given implies that

$$\begin{aligned} \liminf _{n\rightarrow \infty } \int _a^b |f_n(E)|^2\; \textrm{d}E\,<\, \infty . \end{aligned}$$

This means, along a suitable sub-sequence, the norm of \(f_n\) in \(L^2(a,b)\) is bounded. By Banach–Alaoglu, there is a sub-sequence ( o better, a sub-sub-sequence of the suitable sub-sequence) \(f_{n_k}\) which converges weakly in \(L^2(a,b)\) to a limit \(f \in L^2(a,b)\). Noting that bounded continuous functions \(g \in C_b(a,b)\) are also in \(L^2(a,b)\), one has

$$\begin{aligned} \lim _{k \rightarrow \infty } \int _a^b g(E)\; f_{n_k}(E)\,\textrm{d}E\,=\, \int _a^b g(E) \,f(E)\,\textrm{d}E. \end{aligned}$$

for all \(g\in C_b(a,b)\). But since \(f_{n_k}(E)\textrm{d}E\) converges weakly to the measure \(\mu _{{\vec {x}}}\), this means that in the interval (ab) we have

$$\begin{aligned} \textrm{d}\mu _{{\vec {x}}}(E)\,=\,f(E)\,\textrm{d}E\; \end{aligned}$$

which is an absolutely continuous measure in (ab) with a density in \(L^2(a,b)\). \(\square \)

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

González, H., Sadel, C. Absolutely continuous spectrum for Schrödinger operators with random decaying matrix potentials on the strip. Lett Math Phys 113, 9 (2023). https://doi.org/10.1007/s11005-023-01632-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11005-023-01632-8

Keywords

Mathematics Subject Classification

Navigation