Absolutely continuous spectrum for Schrödinger operators with random decaying matrix potentials on the strip
Abstract.
We consider a family of random Schrödinger operators on the discrete strip with decaying random 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.
Key words and phrases:
random decyaing potential, absolutely continuous spectrum, extended states2010 Mathematics Subject Classification:
Primary 82B44, Secondary 60H25, 47B361. Model and main result
We consider a random family of block-Jacobi operators on given by
(1.1) |
where means that , with , In the case one sets in (1.1). is a fixed Hermitian matrix () and finally we have a random Hermitian-matrix potential This means, we have some probability space and valued random variables . Moreover, we assume that the family is independent and that
(1.2) |
where denotes the expectation value.
We also define the ’unperturbed’ operator by eliminating the ,
(1.3) |
and can be seen as quasi-one dimensional discrete Schrödinger operators on a semi-infinite strip of width . The matrix maybe the adjacency matrix of a finite graph , in which case would be like a discrete Laplace operator on the product graph . is then a random perturbation of adding the matrix potentials at each level . This way, falls into the class of operators describing randomly perturbed quantum systems. The study of such systems was initiated by Anderson [1] with the today called Anderson model where one studies operators on with independent identically distributed potentials on each lattice site. In general for such models one finds Anderson localization at large disorder (large variance of the potential) and at the edges of the spectrum. Anderson localization means one has pure point spectrum and exponentially decaying eigenfunctions. There are two general methods to prove this, the fractional moment method [3] and multi-scale analysis [12, 13, 14]. The fractional moment method at high disorder works fine in any graphs with a finite upper bound on the connectivity of one point [32]. In dimension (line or strip) one finds localization for the Anderson model at any disorder [15, 21, 7, 19].
Except in one dimension, for a long time the high disorder Bernoulli Anderson model could not be handled, this means the i.i.d. potential has a Bernoulli distribution. A first breakthrough was done for the continuous model in [5], and recently, the high-disorder localization has also been shown for the discrete Bernoulli Anderson model in for and dimensions [23, 24].
From dimensions on, one expects some absolutely continuous spectrum at small enough disorder. However, this is still conjectural. Existence of absolutely continuous spectrum for Anderson models at low disorder has first been proved for infinite dimensional hyperbolic type graphs like regular trees and tree-like structures [18, 2, 4, 9, 10, 20, 16, 26, 27]. It has also been shown for the Anderson model on special graphs with a finite-dimensional growth, so called anti-trees and partial antitrees [28, 29].
As a mean to study critical transitions from absolutely continuous to pure point spectrum, random decaying potentials in one dimension were also investigated [17, 22, 11]. Here, we extend and improve on the result by Froese, Hasler and Spitzer [11] using methods similar to Last and Simon [22]. The key point for the absolutely continuous spectrum result in [22] has been the spectral average formula by Carmona-Lacroix [6, Theorem II.3.2]. Here, we use its generalization to strips, Proposition A.2 which is a special case of the broader generalization recently done in [30].
1.1. Spectrum and spectral bands
Without loss of generality, we may assume that in (1.1) is a diagonal matrix: If this is not the case, then, as , there is a unitary matrix such that is diagonal. Then, define the unitary operator by and one finds:
Now is diagonal and are random Hermitian matrices satisfying an inequality as (1.2). Thus, using this unitary conjugation, we may assume that is diagonal, hence
(1.4) |
with , being the eigenvalues of . As a consequence,
and the spectrum is purely absolutely continuous.
We call the j-th band of the spectrum of , are the band-edges of this band. Each band-edge can be internal, meaning inside of another band, or external, meaning an edge (boundary point) of the spectrum of . We consider the spectrum of without all the (external and internal) band-edges and define
(1.5) |
Note is open and . We also define the intersection of all open bands,
(1.6) |
which might be empty. For the essential spectrum we note that where is almost surely a compact operator, hence,
1.2. The main result
The main theorem of the whole thesis is the following:
Theorem 1.
Apart from discrete spectrum, (embedded isolated eigenvalues) the spectrum of is almost surely purely absolutely continuous in . Moreover, there are no embedded eigenvalues in the intersection of the bands, .
That means, there may be random embedded eigenvalues in which may only accumulate at the boundary , that is,
the internal and external band-edges.
In technical terms, this means,
there is a set of probability one, , such that for all and all compact subsets
, there is a finite (random) subset of eigenvalues , such that the spectrum of is purely absolutely continuous in .
Under the slightly stronger assumption that and it was aready shown in [11] that the spectrum is purely absolutely continuous in , and that there is absolutely continuous spectrum in all of . However, the proof method used there for the set does not exclude any other type of singular spectrum.
Note that in the line case, we have and purely absolutely continuous spectrum in this case has already been shown in [22]. On the line case it is also known that for any, also non-random -potential, one has absolutely continuous spectrum in [8], but again, any other type of embedded singular spectrum is possible (not excluded in the proof).
The general operator investigated here allows the case, were the operator (almost surely) splits into the direct sum of two strip operators (two separated strips). Then, adjusting one of the one may create an eigenvalue for , lying outside of its essential spectrum, but lying inside the essential spectrum of . In fact, one may have the part of belonging to non-random and create some fixed embedded eigenvalue (for all of ). Thus, without further ’channel-mixing’ assumptions, one can not expect to obtain pure absolutely continuous spectrum within . But under sufficient ’mixing’ created by the , this should be true.
Finitely, let us mention that Theorem 1 does also cover the ’full’ strip case going from to . That means, Theorem 1 also applies to random operators on (rather than ) of the form
where the are independent Hermitian matrices satisfying
To see this, we can transform the operator unitarily to an operator on of the above form (1.1) by defining
It is obvious to see that the random matrices do indeed satisfy the conditions as above.
To picture the transformation, just think of the case , the doubly infinite discrete line, and then flip over the negative line to make a half-infinity strip of width two. The off-diagonal blocks in then correspond to the connection in the shell of the strip, which corresponds to the points and on the line:
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/de61b06f-3190-4ad2-8a79-c2f4c23abc5d/line-folded.png)
2. Transfer matrices, elliptic and hyperbolic channels
The eigenvalue equation is a recursion that can be written in the matrix form as follows:
Iteration leads to the products of transfer matrices for ,
We may write
is basically the transfer matrix of the unperturbed operator . We will now write the transfer matrix in some basis which diagonalises . Recall, is assumed diagonal and its eigenvalues are . Adopting the notions of [25] we define:
Definition 1.
Let . We call the -th channel
-
(1)
Elliptic at if
-
(2)
Hyperbolic at if
-
(3)
Parabolic at if
Now fix some . Note that by the definition of , there are no parabolic channels and there is at least one elliptic channel at We assume the channels to be ordered such that:
Note that the set of all satisfying these inequalities is some open interval . We later vary slightly within this interval. For , and we define by
For with we define , , by
We define the diagonal matrices
such that
Thus, for the transfer matrices of we find
Also we note that
so that , and are the eigenvalues of . In order to diagonlize we introduce
where is the unit matrix of size , then
(2.1) |
with being diagonal, more precisely,
(2.2) |
We note that is indeed invertible for as and in this case. Defining
(2.3) |
we find
(2.4) |
Now, in order to work with uniform estimates we will restrict our consideration to a compact interval . Chosen such a compact interval and allowing complex values for and , we can extend the definitions of analytically to spectral parameters in the complex plane, , for small enough. This means , and the equations (2.1) and (2.2) still hold with replaced by . We will need this extension in some part to use analyticity arguments.
Choosing small enough, one can guarantee by compactness and analyticity arguments, that there is some such that
(2.5) |
Note for we have .
3. The key estimates
The estimates in this section are somewhat independent of the rest of the paper. But in many respects, it is the key part of the proof.
We consider the following general situation: Let be given independent random matrices of the form
(3.1) |
where for some fixed we have
(3.2) |
(Later we use and .)
Note that the second condition implies for any vector . Moreover, are independent random matrices satisfying, with some fixed constant ,
(3.3) |
For certain parts we will also assume the stricter bound (cf. Proposition 3.3). In fact, at real spectral parameters we will have this bound, however, for some arguments we need to allow some small imaginary part, which is why in general we only assume in this section.
Now let us consider the Markov process of matrices given by
Using the splitting into blocks of sizes and like above we write
(3.4) |
From the process we will define the process of pairs of matrices given by
(3.5) |
is a so called Schur-complement. Some standard calculations , see for instance [31], show that can be seen as the process of equivalence classes of defining
with being the identity matrix, and being any matrix. Note that the set of matrices of the form is a group.
In that sense if is invertible we get
and we find
which leads to the identities
(3.6) |
(3.7) |
provided that and exist.
Proof.
First we note that by straight forward calculations one finds for and some square matrix that
(3.8) |
From there we find:
To show this, we use the equivalence (3.8) noting
Now we proof by induction that is invertible and . First, we notice is invertible and . Now assume and being invertible. We find
and by the lower right block
Lemma 3.2 now shows invertibility of and hence of . Using (3.6) we find
This finishes the induction. ∎
Proof.
We will prove by induction that there exists such that for all we have for all . The induction start for is given by Proposition 3.1. Assume the statement is true for . Then we find such that for all
Using (3.6), Lemma 3.8 we find for that
For the last line we used the estimates . This finishes the induction and the first statement.
For the second statement, note
which gives
(3.9) |
where we use that as .
Moreover,to get the last assertion, note
Now, using
we obtain
Therefore, using ,
(3.10) |
Concerning probabilistic estimates, the main point of this section is the following proposition.
Proposition 3.4.
Proof.
Given a starting vector we define inductively by Using (3.7) we find
Now:
As and , we first note
The problematic terms, were we can not use the expectation outside the norm are and . For the other terms, we remark
(3.11) |
For the last step we use the bound (3.3) and the fact that is independent of and thus .
For the frst problematic term, note that
using the fact that is measurable and is independent of . Thus
(3.12) |
Now for the term we want to use a similar estiamte. However, one of hte problem is now that actually depends on and . However, is independent of and are measurable. Thus, we want to condition on . Furthermore, before that, in order to handle some the inverse, we use we use a resolvent identity together with (3.6) to find
giving
Thus,
where
Splitting up this way gives
Using the bounds from above, we see
and
Thus, we have in total the bound
(3.13) |
In summary we find
where and are some positive conitnuous functions in . Taking we find
Use for being some orthogonal basis to get the result. ∎
4. Applying the key estimates to the transfer matrices
The main point of this section will be to apply the estimates from Section 3 to the conjugated transfer matrices as developed in Section 2. Like indicated at the end of Section 2 we choose some compact interval such that for the first channels are elliptic and the other channels are hyperbolic. In the notations of the previous sections, we have , and the matrices and as defined in (3.1) are given by
where and depend analytically on . Using continuity and compactness arguments we have uniform estimates like
for all with and . This leads to
(4.1) |
for all .
In order to apply the results of Section 3 we need . We therefore will replace by
(4.2) |
where the latter expression is the indicator function on the event that on the probability space . This means essentially to replace the potential by
Note that by the estimates above
(4.3) |
We modify the transfer matrices accordingly and let
(4.4) |
With these definitions we note that
Similarly to the products we define
and
(4.5) |
Using the splitting into blocks of sizes and we write
(4.6) |
and we define the Schur complements
(4.7) |
The reason that we will work with the products from some on is that for large and some random , we will have that . More precise probabilistic arguments will be given later.
First, we need to check that the matrices do indeed satisfy the bounds we need:
Proposition 4.1.
There exists (depending on the chosen compact interval and the chosen ) such that for all we have
Proof.
First we note
uniformly for , which bounds the second term as needed.
Using the Cauchy-Schwartz Inequality in we find
For the first term we use (4.1), for the second term, we use Chebyshev’s inequality
in order to get
(4.8) |
for all . Thus,
which leads to
Now does the job. ∎
Thus, we can apply the results from Section 3.
Proposition 4.2.
Let which satisfies .
-
(i)
For all , all , and for all we have,
-
(ii)
For all , all
and, uniformly in we find
-
(iii)
We have for all , and that
-
(iv)
We find such that (uniformly) for all and all
Proof.
For part (i) note that with probability 1, for . We let be the set of probability one where . Then we have the same for and the limits follow from Proposition 3.3.
For part (ii) note first that is analytic in . Now, for fixed, one sees from the estimates in Proposition 3.3 that the convergence of the series is uniform for . Hence, the limiting function is analytic in . Moreover, if for all we have then one sees that with a uniform constant , we have . As for , we find arbitrarily small as and hence uniformly in .
Proposition 4.3.
There is a set of probability one, , , such that for any and any we find
Proof.
5. Absolutely continuous spectrum
In this section we finally prove Theorem 1. Recall in Proposition 4.2 we defined the set of probability one, where . For we find such that for and all we have . However, the is random and not uniform in . Therefore, we define the events
For , and we find and, hence,
Moreover,
The main work left to do now is to use Proposition 4.3 to obtain an estimate of the form as needed in Theorem A.3 which is a special case of [30, Theorem 4]. Thus, given a vector associated to some vector in the 0-th shell, we need to find vectors such that
Note that for one has
Lemma 5.1.
For assume
(5.1) |
then, for any one finds , such that
(5.2) |
If the condition (5.1) is fulfilled for specific and , then it is fulfilled in a neighborhood of . Moreover, given a fixed vector , one may get solutions and that depend continuously on in a neighborhood of .
Now let and use and denote , by and . Hence,
(5.3) |
Proof.
Using (5.2) in the decomposition of above, with and , the statement (5.3) follows directly. Thus, we need to check that we find and such that (5.2) is satisfied. Dividing the matrix horizontally in blocks of sizes and , and vertically into blocks of sizes and we may write
Note . Then, (5.2) is satisfied for if and only if
This is equivalent to
Thus, we find a solution for any , if is surjective (as a linear map from to ), which is exactly the rank condition given in the assumption.
Note, if this is fulfilled for some specific , and some specific spectral parameter , then we find a matrix such that ,. So in a neighborhood of and , this determinant is still not zero and we may use
and as above, . Thus, both depend continuously on . ∎
In the sequel need to use the form of the transfer matrices as in [30] using the resolvent boundary data of restrictions to finite graphs. Thus, let be the restriction of to , that is where is the natural embedding. Note that is a Hermitian matrix. One may define the resolvent boundary data from shell 0 to as in [30] by
(5.4) |
where is the natural embedding of into and can be regarded as an matrix. In this sense,
and and are all matrices. Then, one of the main points following from the work in [30] is the following formula, which we also prove in Appendix A.
Proposition 5.2.
(cf. Proposition A.1) If is not an eigenvalue of and is invertible, then
Now, we can continue with the following. Note, that and implies .
Lemma 5.3.
Given , and fixed, there exists such that and we have , where
Proof.
For notation we let . From Proposition 4.2 part (ii)we find that is uniformly small for sufficiently big. This means, for any , there exists such that for any and any we have
The needed for the statement will be chosen later.
Using the definitions (5.4) and Proposition 5.2 we find
By the other ways of writing the transfer matrix, we see that exists for any , at least after analytic continuation. We also note that exists for any value except for the eigenvalues of . Thus, it exists for any with .
In order to prove that is of full rank , it is sufficient to prove that is invertible, where . In particular, we consider
First, take the ’limit case’ and with (2.4) we find
were we note that by their definition, and commute. Thus, we find
where
Using , where , and compactness, we get with some uniform constant that
for all and all . This gives
for any and any . Note, is a diagonal matrix, such that
Moreover, as set above, all diagonal entries of are bigger than . We note, that the imaginary parts of have opposite sign and an absolute value smaller than for the corresponding values of . Thus, we find for that
In general, we will define the ”imaginary” part in algebra sense, that is , then
for and Hence, we finally obtain
Thus, if
then, the matrix is invertible and we have (for any ). By the dimensions of we also have . ∎
Proposition 5.4.
Let , where is the set as in Proposition 4.3.
Then, there is a finite set such that the spectrum of is purely absolutely continuous in
.
If there is no hyperbolic channel, that is , then the spectrum is purely absolutely continuous in .
Proof.
For some we find . Choose with , take as in Lemma 5.3 and consider some fixed . We note that we also have . Now, using the notation as above, has full rank for . By analyticity, the rank of is full for all but finitely many values of . We may now restrict to the real line again and let be the finite set of energies, where .
We consider now a compact interval . For all we find that has full rank . By compactness, the set has some positive distance, say , to the set of matrices of non full rank.
In order to get to the point of Lemma 5.3, let us introduce the notations
Again by compactness we note that for all . (Note, that is fixed now!). Thus we see that
for all . Therefore, if
then is of full rank .
Now, consider the compact set
By Lemma 5.1, for any , we find some neighborhood and solutions , to (5.2), that depend continuously on . Possibly shrinking the neighborhood a bit, we may assume it is compact, and thus, attains a maximum in . By compactness, can be covered by finitely many such compact neighborhoods . Making a specific choice in the overlaps of these finitely many neighborhood, we find piece-wise continuous functions
satisfying equation (5.2) such that for some constant and all we have
As mentioned in the proof of Proposition 4.2 part (ii), the convergence of for is uniform in , as such we find such that and all we have
Thus, for all , and all we may choose
By (5.3) we obtain that
(5.5) |
for and all .
Hence, using that we get by Proposition 4.3 that
Hence, Theorem A.3 gives that the spectral measure at is purely absolutely continuous in .
As was arbitrary, and the closures of is the whole Hilbert space,
we find that the spectrum of is purely absolutely continuous in .
Now, the set can be written as countable union of intervals such that .
Therefore, the spectrum of is purely absolutely continuous in . Note that may be eigenvalues of , but they do not have to be. If is not an eigenvalue,
then the spectrum of is also purely absolutely continuous in a neighborhood of .
Thus, we only need to subtract the eigenvalues from the set .
Note, in the intersection of all the bands, that is, if , one has , and one can choose any family of uniformly bounded vectors to get pure absolutely continuous spectrum in . There is no need to subtract a finite set of values.
∎
Theorem 1 now essentially follows directly from this proposition:
Proof.
First note that the set does not depend on the interval analyzed above, but does. Using compact intervals inside with rational boundary points we may write as countable union of open intervals, whose closure is inside ,
As does not contain any band-edges, for each the type of the -th channel does not change in . Therefore, one can make the whole analysis as done for the compact interval above for the interval . In particular, there is a corresponding set of probability one for the set . We then let
and note . Let and let be compact. Using compactness, there is a finite sub-collection of these intervals, , , such that
Theorem 5.4 gives that there is a finite set of eigenvalues, such that the spectrum of in is purely absolutely continuous.
Letting , which is finite, we see that the spectrum in is purely absolutely continuous.
Due to the last comment in Theorem 5.4, the spectrum of is purely absolutely continuous in the intersection of all bands (which might be an empty set).
∎
6. Acknowledgement
This work has been supported by the Chilean grants FONDECYT Nr. 1161651, FONDECYT Nr. 1201836 and the Nucleo Mileneo MESCD.
Appendix A Transfer matrices and spectral averaging formula on the strip
As above we consider operators of the form
on . Solving the eigenvalue equation leads to the transfer matrices
Then, for we define the products
for a formal solution of .
A.1. Transfer matrices and resolvent boundary data
Let , be non-negative integers. With we denote the restriction of to , that is
Then we define the to boundary resolvent data for by
(A.1) |
where is the natural embedding of into for . This means, e.g. , and in this setup
Note that are all matrices.
Proposition A.1.
Let be given and let and let be invertible. Then,
Proof.
With being the spectral parameter, leads to
Multiplying with from the left, noting that and using (A.1) gives
Multiplying from the left with instead of leads to
Replacing with the formula above and resolving for leads to
Finally, we have:
As determine the solution to uniquely, the matrix must be . ∎
A.2. Spectral averaging formula
Here we state the strip-equivalent of the spectral average formula from Carmona-Lacroix [6, Theorem III.3.2 and III.3.6]. It is a special case of [30, Theorem 1]. First, we need to fix a vector in the root-slice. Thus, we choose some which we identify with . Let us assume that , so that . Furthermore, identifying with a linear map from to , we have a dimensional kernel consisting of the vectors orthogonal to ,
Then, in this special case, the work of [30] simply replaces by the set of matrices
(A.2) |
Note that
(A.3) |
where we adopt the notation that for sets of matrices .
Moreover we consider the spectral measure at the vector , that means
Now, using that the operator can not have compactly (finitely) supported eigenfunctions, Theorem 1 in [30] implies the following:
Proposition A.2.
[30] In the sense of a weak limit for finite measures one finds that
Using the symplectic structure of the transfer matrices and the Banach-Alaoglu theorem one can obtain a criterion for absolute continuity (see [30]).
Proposition A.3.
If one finds for , , such that
then, the measure is absolutely continuous in the interval .
Proof.
First, in [30] it was shown that the minimum is achieved at a very specific vector which we call . Defining
we see from Theorem A.2 that is the weak limit of in the interval . Note that
where we use and as . Now, using the Cauchy Schwartz inequality, this gives
and hence
Thus, the estimate given implies that
This means, along a suitable sub-sequence, the norm of in is bounded. By Banach-Alaaoglu, there is a sub-sequence ( o better, a sub-sub-sequence of the suitable sub-sequence) which converges weakly in to a limit . Noting that bounded continuous functions are also in one has
for all . But since converges weakly to the measure this means that in the interval we have
which is an absolutely continuous measure in with a density in . ∎
References
- [1] Anderson, P.W. (1958), Absence of diffusion in certain random lattices, Physical Review 109, 1492–1505, doi: 10.1103/PhysRev.109.1492
- [2] Aizenman, M. & Sims, R. & Warzel, S. (2006), Stability of the absolutely continuous spectrum of random Schrödinger operators on tree graphs, Prob. Theor. Rel. Fields, 136, 363-394, doi: 10.1007/s00440-005-0486-8
- [3] Aizenman, M. & Molchanov, S. (1993), Localization at large disorder and extreme energies: an elementary derivation, Commun. Math. Phys. 157, 245-278, doi: 10.1007/BF02099760
- [4] Aizenman, M. & Warzel, S. (2013), Resonant delocalization for random Schrödinger operators on tree graphs, J. Eur. Math. Soc., 15 (4), 1167-1222, doi: 10.4171/JEMS/389
- [5] Bourgain, J. & . Kenig, C. (2005) On localization in the continuous Anderson-Bernoulli model in higher dimension, Inv. Math., 161, 389-426, doi: 10.1007/s00222-004-0435-7
- [6] Carmona, R. & Lacroix, J. (2013), Spectral theory of random Schrödinger operator, Boston: Birkhäuser.
- [7] Carmona, R. & Klein, A. & Martinelli, F. (1987) Anderson localization for Bernoulli and other singular potentials, Commun. Math. Phys. 108, 41-66, doi: 10.1007/BF01210702
- [8] Deift, P. & Killip, R. (1999) On the absolutely continuous spectrum of one-dimensional Schrödinger operators with square summable potentials, Communications in Mathematical Physics 203, 341-347, doi: 10.1007/s002200050615
- [9] Froese R., & Hasler, D. & Spitzer, W. (2006), Transfer matrices, hyperbolic geometry and absolutely continuous spectrum for some discrete Schrödinger operators on graphs, Journal of Functional Analysis 230, 184-221, doi:10.1016/j.jfa.2005.04.004
- [10] Froese, R. & D. Hasler, D. & Spitzer, W. (2007), Absolutely continuous spectrum for the Anderson Model on a tree: A geometric proof of Klein’s Theorem, Communications in Mathematical Physics 269, 239-257, doi: 10.1007/s00220-006-0120-3
- [11] Froese, R. & Hasler, D. & Spitzer, W. (2010) On the ac spectrum of one-dimensional random Schrödinger operators with matrix-valued potentials, Mathematical Physics Analysis and Geometry 13, 219-233, doi: 10.1007/s11040-010-9076-9
- [12] Fröhlich, J. & Spencer, T. (1983) Absence of difusion in the Anderson tight binding model for large disorder or low energy, Communications in Mathematical Physics 88, 151-184, doi: 10.1007/BF01209475
- [13] Germinet, F. & A. Klein (2001), Bootstrap multiscale analysis and localization in random media, Communications in Mathematical Physics 222, 415–448, doi: 10.1007/s002200100518
- [14] Germinet, F. & Klein, A. (2013), A comprehensive proof of localization for continuous Anderson models with singular random potentials, Journal of the European Mathematical Society 15, 53–143, doi: 10.4171/JEMS/356
- [15] Gol’dsheid, Y & Molchanov, S. & Pastur, L. (1977) Pure point spectrum of stochastic one dimensional Schrödinger operators, Funct. Anal. Appl. 11, 1-10, doi:10.1007/BF01135526
- [16] Keller, M. & Lenz, D. & Warzel, S. (2012) Absolutely continuous spectrum for random operators on trees of finite cone type, J. D’ Analyse Math. 118, 363-396, doi: 10.1007/s11854-012-0040-4
- [17] Kiselev, A. & Last, Y. & Simon, B. (1997), Modified Prüfer and EFGP transforms and the spectral analysis of one-dimensional Schrödinger operators, Communications in Mathematical Physics 194, 1-45, doi: 10.1007/s002200050346
- [18] Klein, A. (1998), Extended states in the Anderson model on the Bethe lattice, Advances in Mathematics 133, 163-184, doi:10.1006/aima.1997.1688
- [19] Klein, A. & Lacroix, J. & Speis A. (1990), Localization for the Anderson model on a strip with singular potentials, J. Funct. Anal. 94, 135-155, doi: 10.1016/0022-1236(90)90031-F
- [20] Klein, A. & Sadel, C. (2012) Absolutely Continuous Spectrum for Random Schrödinger Operators on the Bethe Strip, Mathematische Nachrichten 285, 5-26, doi: 10.1002/mana.201100019
- [21] Kunz, H. & Souillard, B. (1980) Sur le spectre des operateurs aux differences finies aleatoires, Commun. Math. Phys. 78, 201-246, doi:10.1007/bf01942371
- [22] Last, Y. & Simon, B. (1999), Eigenfunctions, transfer matrices, and absolutely continuous spectrum of one-dimensional Schrödinger operators. Inventiones Mathematicae, 135(2), 329-367. doi:10.1007/s002220050288
- [23] Li, L. (2020) Anderson-Bernoulli localization with large disorder on the 2D lattice, preprint, arXiv:2002.11580
- [24] Li L. & Zhang, L. (2019) Anderson-Bernoulli Localization on the 3D lattice and discrete unique continuation principle, preprint, arXiv:1906.04350
- [25] Sadel, C. (2011), Relations between transfer and scattering matrices in the presence of hyperbolic channels. Journal of Mathematical Physics, 52, 1235111, doi:10.1063/1.3669483
- [26] Sadel, C. (2013), Absolutely continuous spectrum for random Schrödinger operators on tree-strips of finite cone type, Annales Henri Poincaré, 14, 737-773, doi: 10.1007/s00023-012-0203-y
- [27] Sadel, C. (2014), Absolutely continuous spectrum for random Schrödinger operators on the Fibbonacci and similar tree-strips, Mathematical Physics Analysis and Geometry 17, 409-440, doi: 10.1007/s11040-014-9163-4
- [28] Sadel, C. (2016) Anderson transition at two-dimensional growth rate on antitrees and spectral theory for operators with one propagating channel, Annales Henri Poincare 17, 1631-1675, doi: 10.1007/s00023-015-0456-3
- [29] Sadel, C. (2018). Spectral theory of one-channel operators and application to absolutely continuous spectrum for Anderson type models. Journal of Functional Analysis, 274(8), 2205-2244. doi:10.1016/j.jfa.2018.01.017
- [30] Sadel, C. (2021), Transfer matrices for discrete Hermitian operators and absolutely continuous spectrum, Journal of Functional Analysis, 281(8), published online doi:10.1016/j.jfa.2021.109151
- [31] Sadel, C. & Virág, B. (2016). A Central Limit Theorem for Products of Random Matrices and GOE Statistics for the Anderson Model on Long Boxes. Communications in Mathematical Physics, 343(3), 881-919. doi:10.1007/s00220-016-2600-4
- [32] Tautenhahn, M. (2011), Localization criteria for Anderson Models on locally finite graphs, Journal of Statistical Physics 144, 60-75, doi: 10.1007/s10955-011-0248-1