Irreducibility of the Bloch Variety for Finite-Range Schrödinger Operators
Abstract.
We study the Bloch variety of discrete Schrödinger operators associated with a complex periodic potential and a general finite-range interaction, showing that the Bloch variety is irreducible for a wide class of lattice geometries in arbitrary dimension. Examples include the triangular lattice and the extended Harper lattice.
1. Introduction
1.1. Setting and Main Theorem
We will study periodic finite-range Schrödinger operators of the form
(1.1) |
acting in , where is periodic and is a Toeplitz operator given by
Here, is finitely supported and will as usual denote both the potential and the corresponding multiplication operator . We say that is -periodic for if for all and each , where denotes the standard th basis vector.
In particular, let us note that the approach discussed herein does not rely on reality of the potential or self-adjointness of . The case in which
corresponds to , the discrete Laplacian.
Our main result is irreducibility of the Bloch variety for all operators of the form (1.1) subject to a suitable condition on . In particular, under mild assumptions on , the result holds universally for all periodic , including complex-valued potentials. We will define the Bloch variety precisely later in the manuscript (see Section 1.2); for now, the reader may think of it as a relation between the energy, , and the quasi-momentum, , which is given by the zero set of a function that is a polynomial in and a Laurent polynomial in .
Let us describe some of the main objects and assumptions used in this work. Starting with we generate the Laurent polynomial
(1.2) |
where we employ the standard multi-index notation .
Let us state our assumptions here in a moderately informal manner. For further definitions, details, and a more precise account, we refer the reader to Section 2, where we define the component of lowest degree and the fundamental domain . Let denote the lowest degree component of in the sense that . Given and , the vector is defined by
Given , we define . Our main assumptions are the following:
-
()
The degree of is negative.
- ()
Theorem 1.1.
Remark 1.2.
Let us make a few comments about Theorem 1.1.
-
(1)
The precise definitions of the Bloch variety and irreducibility modulo periodicity are given in Section 1.2. To prove Theorem 1.1, we use standard Floquet theory to obtain a Laurent polynomial with the property that belongs to the Bloch variety if and only if . We show that is irreducible as a function of and . Since the Bloch variety is defined as a set of pairs , it is only irreducible after one quotients out the relevant action. This is indeed necessary as shown by the free Laplacian, see Equation (1.22) and Figure 4 in [14]. The main idea of the proof is to reduce from the operator to by focusing on the lowest-degree component of (after a suitable change of variables). We show that reducibility combined with Assumptions () and () would imply mutually contradictory properties of the lowest-degree component.
- (2)
-
(3)
The strength of the result comes from the generality of the operators under consideration. For instance, this can handle operators in higher dimensions on rather general graphs; specifically, one can handle -periodic graphs for which acts transitively on vertices (see Remark 6.1 for further details). We emphasize that our result does not necessarily imply irreducibility of the Fermi varieties associated with .
Theorem 1.1 is the main motivation for this work. It will follow from a more general result formulated in Theorem 2.6 below.
The above assumptions are satisfied and straightforward to verify in many cases of interest. To illustrate scope of applications, we enumerate some corollaries.
We first note that Theorem 1.1 provides a direct proof of the irreduciblity of the Bloch variety for all discrete Schrödinger operators on .
Corollary 1.3.
If denotes the Laplacian on , then for any periodic , the Bloch variety of is irreducible modulo periodicity.
The corollary above was already known via results about the Fermi variety – see the discussion in Section 1.2 for additional details and references. Thus, we supply an alternative argument, working directly on the Bloch variety.
More significantly, Theorem 1.1 also enables one to prove irreducibility of the Bloch variety for other lattice geometries in arbitrary dimension. To remain concrete, we present a couple of two dimensional examples but the reader may readily recognize from the proofs that generalizations are possible (compare Remark 6.1).
Corollary 1.4.
If denotes the Laplacian on the extended Harper lattice, and are coprime, and is -periodic, then the Bloch variety of is irreducible modulo periodicity.
Corollary 1.5.
If denotes the Laplacian on the triangular lattice, then for any periodic , the Bloch variety of is irreducible modulo periodicity.
Note that irreducibility of the Bloch variety is potentially sensitive to modifications in the hopping terms (i.e., the matrix elements of ). To the best of our knowledge, even the results of Corollaries 1.4 and 1.5 are new. For further details, including definitions of the triangular and extended Harper lattices, see Section 6. To emphasize the distinction between the above models, we present the corresponding polynomials below, recalling that Equation (1.2) provides the dictionary between and .
-
(i)
For the discrete Laplacian on ,
-
(ii)
For the extended Harper lattice
-
(iii)
For the triangular lattice,
In particular, in dimension , adds to next nearest neighbor terms and is symmetric with respect to the map for The polynomial does not possess this symmetry, nonetheless the corresponding variety still falls into the scope of Theorem 1.1. The triangular lattice is depicted in Figure 2. Applying a simple shear transformation reduces the triangular lattice to the square lattice with additional edges, as shown in Figure 2, and hence places the Laplacian on the triangular lattice into the context of the paper after a suitable change of coordinates.
1.2. Definitions and Context
Let us now give relevant definitions and context. Given , , let . We say that a function is -periodic (-periodic, or just periodic) if for all and all .
Definition 1.6.
Let . For and , the space consists of those for which
(1.3) |
where we write and use the multi-index notation . Naturally, is a Hilbert space of finite dimension .
If is -periodic, the corresponding Bloch variety is given by
(1.4) |
where we write . We employ here a standard abuse of notation in which represents both the self-adjoint operator in and the difference operator acting in, say, .
Definition 1.7.
We will say the Bloch variety is irreducible modulo periodicity if for every two irreducible components and of , there exists such that .
Definition 1.8.
Given , the Fermi surface (variety) is defined as the level set of the Bloch variety:
We should mention that reducible Fermi and Bloch varieties are known to occur for periodic graph operators, e.g., [23, 8]. One challenging problem in the study of periodic operators is to prove the (ir)reducibility of the Bloch and Fermi varieties [10, 13, 5, 3, 4, 17, 23, 8, 9, 21]. For instance, irreducibility of the Bloch variety implies that in case for some open set , the knowledge of allows one to recover . Besides its own importance in algebraic geometry, the (ir)reducibility of these varieties is crucial in the study of spectral properties of periodic elliptic operators. In particular, this has implications for the structure of spectral band edges [19, 7], the isospectrality [18] and the existence of embedded eigenvalues for operators perturbed by a local defect [16, 15, 22, 12, 1, 6, 20]. Based on existing evidence, Kuchment conjectures that the Bloch variety of any periodic second-order elliptic operator is irreducible [14, Conjecture 5.17].
There have been many works that address irreducibility of the Bloch and Fermi varieties (see, e.g., [3, 4, 2, 5, 10, 13, 19]). In two dimensions, Bättig [2] showed that the Bloch variety is irreducible (modulo periodicity). In [10], Gieseker, Knörrer and Trubowitz proved that is irreducible except for finitely many values of . When , irreducibility of for any was proved by Bättig [4]. In a recent paper of the second author [19] it is showed that when is irreducible for any and when is irreducible for all , where is the average of over one periodicity cell. It follows from these works that when the Bloch variety is irreducible (modulo periodicity). For continuous periodic Schrödinger operators, when Knörrer and Trubowitz [13] proved that the Bloch variety is irreducible (modulo periodicity) and for , Bättig, Knörrer and Trubowitz proved that the Fermi variety at any level is irreducible (modulo periodicity) for separable periodic potentials [5]. In [17], irreducibility of the Fermi variety for all but finitely many energies is proved for a suitable class of planar periodic graphs.
In [10, 13, 5, 3, 4, 2], algebraic geometry techniques are employed to construct the toroidal and directional compactifications of Fermi and Bloch varieties and understand asymptotics of their defining (Laurent) polynomials. The perspective employed by us in the current manuscript is inspired by [19]. In general terms, the goal is to explicitly calculate asymptotics of the (Laurent) polynomials at and show that these asymptotics contain enough information about the original variety. Concretely, the proof is based on changing variables, studying of the lowest degree components of a family of (Laurent) polynomials in several variables and degree arguments. With regards to the Bloch variety, we expand the approach of [19] in different directions. As a consequence, for the main result of Theorem 2.6 below, the underlying lattice may be of very general nature and contain somewhat arbitrary finite-range connections (see () and () below for a more precise statement of the assumptions). In particular, we obtain irreducibility for the Bloch variety corresponding to periodic Schrödinger operators on the triangular lattice and the extended Harper lattice; see Section 6 for a precise description of these examples. While our approach is inspired by [19], we do not follow the same path. By working directly with the lowest degree components, we can eschew a discussion of asymptotic statements about the varieties themselves.
The structure of the paper is as follows. We precisely formulate Theorem 2.6, our main result, in Section 2. Section 3 contains preparatory technical results that are then employed in Section 4 to prove Theorem 2.6. We elucidate the connection between this result and periodic operators in Section 5, which also contains some relevant background on periodic long-range Schrödinger operators. We conclude in Section 6 with the proof of Theorem 1.1 and some relevant examples and applications.
Acknowledgements
We are grateful to the anonymous reviewer for carefully reading the manuscript and for helpful suggestions that improved the paper.
2. Main Result
To state the main result, we begin by recalling some crucial terminology.
Definition 2.1.
Suppose is a Laurent monomial in variables, that is, with for and . The degree of is defined as . Abusing notation slightly, we also denote for the multi-index .
Definition 2.2.
Given a Laurent polynomial
let . Then, the lowest degree component of is defined to be the Laurent polynomial
One of the crucial properties of this notion is the following: denoting the lowest-degree component of by , one has , which enables one to relate factorizations of a polynomial to factorizations of its lowest-degree component. Obviously, some care is needed to deduce nontrivial consequences from this observation in the context of our main result.
Let us write for the set of polynomials in . Similarly, we write for the set of Laurent polynomials111This involves a minor, albeit common abuse of notation, since one has the relation in . in .
Definition 2.3.
Recall that a polynomial is called reducible if there exist nonconstant polynomials such that and irreducible otherwise. Similarly, we say that a Laurent polynomial is irreducible if it can not be factorized non-trivially, that is, there are no non-monomial Laurent polynomials such that .
Notice that nonconstant monomials are units in the algebra of Laurent polynomials, which accounts for a small subtlety. That is, one must be somewhat careful here with zeros at and . The polynomial is reducible in but is a unit in . In practice, this should cause no confusion, and we will write that is irreducible in (respectively in ) if we wish to emphasize the sense in which irreducibility is meant in a specific context.
Remark 2.4.
If is an irreducible Laurent polynomial in variables, then the corresponding variety is irreducible as an analytic set.222The converse is clearly false, which may be considered by considering the variety associated with , where is irreducible. This issue can be elegantly resolved using the language of schemes [11]. Thus, the overall strategy of our work is to show that a suitable Laurent polynomial that describes the Bloch variety is irreducible. Concretely, we may consider the set which consists of those such that enjoys a nontrivial solution . By Floquet theory, one may determine a suitable Laurent polynomial such that is precisely the zero set of (see Section 5). Thus, since if and only if , to show that is irreducible modulo periodicity, it suffices to show that the corresponding Laurent polynomial is irreducible.
Let us begin by collecting some notation that we will use throughout the paper. Given , we define the lattice by
(2.1) |
and the fundamental cell, , by
(2.2) |
Given and , let
(2.3) |
and denote by the vector . We also let
(2.4) |
Let be a Laurent polynomial and define
(2.5) |
We shall work with Laurent polynomials in variables . Abusing notation somewhat, we write (respectively ) for the set of polynomials (respectively the set of Laurent polynomials) in and . The Laurent polynomials of interest in the present work are those of the form
(2.6) |
where the summation runs over in an arbitrary collection of proper subsets of and . In fact, is a Laurent polynomial in the variable and a polynomial in . Collecting terms, we see that
(2.7) |
where and .
Note that we do not exclude the case , our convention being that . These are exactly the types of polynomials that one produces by expanding the determinant of the Floquet operator associated with a suitable periodic operator, hence their interest in the current work.
For each , the constant is assumed to be independent of and . Assume further that is invariant under action of each , i.e.,
(2.8) |
Remark 2.5.
The assumptions (2.6) and (2.8) include the central example where
and the matrices and are defined by
(2.9) |
(2.10) |
Here denotes the discrete Fourier transform of , defined as in (5.1). For further discussion, see Section 5, especially Proposition 5.3.
Let us note the key properties are that is a diagonal matrix and the entries of are independent of . Consequently, neither self-adjointness of or real-valuedness of is a crucial ingredient.
Since is invariant under the action of each , it is elementary to check (cf. Lemma 3.1) that there exists such that
(2.11) |
Our goal is to show that is irreducible as a Laurent polynomial under the assumptions below.
-
()
, where denotes the lowest degree component of , (see Definition 2.2).
-
()
Letting , the polynomials are pairwise distinct.
The reader may readily check that (with addition of indices computed mod ). Thus, to check Assumption () in practice, it suffices to show that for every .
Theorem 2.6.
As mentioned in Remark 2.5, the connection to Schrödinger operators and Theorem 1.1 will be established in Section 5.
Remark 2.7.
Let us collect some notation from the previous paragraphs that will be repeatedly used throughout the proofs.
-
(1)
(resp. ) denotes the set of polynomials (resp. Laurent polynomials) in .
-
(2)
.
-
(3)
is the lowest degree component of .
-
(4)
, , is arbitrary.
-
(5)
, , .
-
(6)
For , .
-
(7)
-
(8)
is given by
-
(9)
-
(10)
for , .
-
(11)
is defined by
-
(12)
for ordered -tuples and .
3. Technical Lemmas
Lemma 3.1.
Suppose is a Laurent polynomial in and . With notation as in Remark 2.7, one has that for every if and only if there is a Laurent polynomial such that
(3.1) |
Proof.
If satisfying (3.1) exists, then it readily follows from the definition of that
To see that the converse implication holds, write
be another Laurent polynomial. Note that
Thus (3.1) holds if and only if for all
and hence satisfying (3.1) exists if and only if for all and . Thus, if (3.1) does not hold, we must have for some , say for some . Choose , and note that for this choice of one has
concluding the proof. ∎
Definition 3.2.
For each , define as follows. We let be the lowest exponent of in in case this exponent is negative and otherwise.
Lemma 3.3.
Let be a Laurent polynomial in and let be the lowest degree component of . Then, the polynomials
are irreducible in for each . Moreover, under Assumption (), we conclude that for any , and are relatively prime.
Proof.
Assume for the sake of contradiction that is reducible. Since the degree of in is one, we must have that
(3.2) |
for non-constant polynomials and . Since does not divide in , we see that there exist non-zero polynomials and such that
From (3.2) and the definition of we obtain . In particular, where are integers with for . Since is nonconstant, for at least one . In particular,
Consequently, (3.2) implies that the polynomial is divisible by for some However, the lowest degree of in is, by definition, equal to . Thus is not divisible by , contradicting (3.2). Consequently, is irreducible.
To prove the second statement of the lemma, assume and share a nontrivial common factor. By irreducibility, they must be constant multiples of one another. However, from the definition, this is only possible if , which contradicts Assumption (). ∎
Let us introduce the auxiliary polynomial
(3.3) |
with as in Lemma 3.3 for . By a direct calculation, is invariant under the replacement333Later on, we will call this the action of on a polynomial. , so, as a consequence of Lemma 3.1, there exists such that
(3.4) |
Remark 3.5.
It is important that we pass to the lift here, since is clearly reducible.
Proof of Lemma 3.4.
Suppose for the sake of establishing a contradiction that is reducible, and write
(3.5) |
for non-constant polynomials and . Let and . Combining (3.4) and (3.5) yields
Moreover, by definition and are both invariant under the action of each . Recall from Lemma 3.3 that each is irreducible. Therefore, each is a factor of either or . By invariance of (respectively ) under the action of each and since, by Lemma 3.3, and are relatively prime for , we conclude the following: if (respectively ) has a factor of then it must have a factor of
However, this, together with (3.5), implies that either or must be constant, which is a contradiction. Thus, we conclude that is irreducible. ∎
Lemma 3.6.
Let be given by (2.11) and let be any irreducible factor of . Then must depend on
4. Proof of Theorem 2.6
Before proceeding with the proof of the main result, Theorem 2.6, let us introduce some notation.
Definition 4.1.
For each denote by the lowest exponent of in in case this exponent is negative and otherwise. Clearly, with given in Definition 3.2.
Proof of Theorem 2.6.
Let . Then is a Laurent polynomial in the variables . Let , be as in Definition 4.1. In case for some , the lowest power of in is .
Claim 4.2.
For each , does not divide .
Proof of Claim. Indeed, if , this is clear from the definitions, since is the smallest power of in and hence is the smallest power of in . Otherwise, , and the claim can be seen from (2.7).
Since also does not divide , Claim 4.2 implies that reducibility of the Laurent polynomial is equivalent to reducibility of the polynomial .
Now, assume for the sake of contradiction that is reducible. There exist and non-constant polynomials , , in such that
(4.2) |
Let us recall the auxiliary polynomial given by
Let . Then, by (2.11) and (4.2), we have that
(4.3) |
By definition of in (2.6) one sees that replacing by for allows us to conclude that the lowest degree component of is given by , where
(4.4) |
We denote by the lowest degree components of , . From (4.3) it must be that
(4.5) |
and hence
(4.6) |
Given , is a polynomial in . Thus, there exists such that
(4.7) |
By (4.4), (4.6) and (4.7), we reach, recalling the definition of in (3.4),
(4.8) |
By Lemma 3.4, is irreducible, so there exists such that has a factor . We conclude that the highest power of in (hence in ) is at least . Since and, by Lemma 3.6 and Claim 4.2, , , must depend on we reach a contradiction since the highest power of on the left-hand side of (4.3) is equal to . ∎
5. Floquet Theory for Long-Range Operators
Let us summarize some of the important points about Floquet theory for operators with long-range interactions. This is well-known, especially in the continuum case; see the survey [14] and references therein. We are unaware of a precise reference in the discrete setting for long-range operators, so we included the details for the reader’s convenience.
Let us assume that is bounded. Writing for the matrix elements, we further assume that is translation-invariant in the sense that
and that satisfies the decay estimate
for constants . By translation-invariance, is fully encoded by via
We denote the Fourier transform on by , where
for and then extended to by Plancherel.
By the assumptions on , the symbol is analytic, real-analytic whenever , and a trigonometric polynomial whenever is finitely supported. For example, when denotes the Laplacian on ,
Recall that is -periodic and denotes the period lattice. We define the dual lattice and
The discrete Fourier transform of a -periodic function is defined by
(5.1) |
Of course, this also makes sense for and satisfies for any and any . One can check the inversion formula
(5.2) |
which holds for any -periodic .
Let denote the torus.
Proposition 5.1.
For any ,
and
Proof.
These follow from direct calculations using the definitions of and assumptions on and and the inversion formula (5.2). ∎
Let us now define ,
and by where
As usual, this is initially defined for (say) vectors, but has a unique extension to a unitary operator on via Plancherel.
Proposition 5.2.
The operator is unitary. If is -periodic, then
where denotes the restriction of to with boundary conditions
(5.3) |
Proof.
Unitarity of follows from Parseval’s formula. The form of follows from a direct calculation.∎
Given , let be the Floquet-Bloch transform defined on as follows: for any vector on , , we set
Therefore,
Let , and define the Laurent series by
(5.4) |
Using multi-index notation, we may rewrite this as
(5.5) |
Proposition 5.3.
Remark 5.4.
In particular, depends on and is independent of , while depends only on with no dependence on .
Proof of Proposition 5.3.
By a direct calculation, we see that is unitary, so it suffices to prove that . Let be with the potential set to zero. We are going to show and separately. To prove that , it suffices to show that for any ,
It is worth mentioning that satisfies (5.3) so that is well defined. With the given definitions, for any ,
(5.7) |
Putting together (2.4) and (5.7),
(5.8) |
Similarly,
(5.9) |
The proof of is similar. ∎
6. Proof of Theorem 1.1 and Examples
Proof of Theorem 1.1.
The Bloch variety precisely consists of those such that there is a nontrivial solution of satisfying the boundary conditions as in (1.4). Thus, with and as in Proposition 5.3, the Bloch variety is the zero set of the Laurent polynomial defined by (2.11) where
After using the standard permutation expansion for this determinant, we see that is of the form (2.6) (with given via (5.5)). By a brief calculation, one can check that satisfies (2.8). Namely, if denotes the shift with addition computed modulo , one can check that
Since , (2.8) follows. Thus, the result follows from Theorem 2.6. ∎
Let us conclude by discussing a few examples of how to obtain the generator for which Theorem 2.6 is applicable. In particular, the examples below show that the framework of the present paper allows one to consider different discrete geometries. We start with the most basic example of the Laplacian on , where
In this case, it readily follows from (5.5) that
(6.1) |
Proof of Corollary 1.3.
We then proceed to the description of a couple of two dimensional examples. The triangular lattice is given by specifying the vertex set
with edges given by . Applying the shear transformation , , one can view this graph as having vertices in and
In particular, the nearest-neighbor Laplacian on the triangular lattice is equivalent to the operator such that
Making use of (5.5) one finds that
(6.2) |
Proof of Corollary 1.5.
Finally, in the Extended Harper Model
Equation (5.5) now implies that
The lowest degree component is . The proof of Corollary 1.4 follows in just the same way as before; notice that we need the periods to be coprime to ensure that Assumption () is met.
Remark 6.1.
To conclude, let us say a few words about when our results can be applied in the general setting of periodic operators on periodic graphs. That is, given a periodic graph and a periodic operator thereupon, when can one apply a suitable change of coordinates as in the case of the triangular lattice to reduce an operator on ? In short, this can be done to periodic operators on any -periodic graph with transitive vertex action. Let us describe this in a little more detail.
Suppose is a locally finite graph with vertices on which acts freely by graph automorphisms. Denote the action of on the vertex by . This gives a unitary representation of viz. , . One may consider operators of the form acting in the natural Hilbert space , where
-
(1)
commutes with the -action (that is, for all )
-
(2)
is diagonal, that is, for a suitable function .
-
(3)
is invariant under the action of a full-rank subgroup of , that is, there is a subgroup of rank such that for all .
In item (3), notice that one can always take to be of the form for some , in which case we say is -periodic as before. If in addition, acts transitively on , then choosing arbitrarily some , one has a one-to-one correspondence via . Of course, this induces a unitary operator in an obvious manner. In this case, it is clear that is a periodic operator of the form studied in the present paper.
References
- [1] K. Ando, H. Isozaki, and H. Morioka. Spectral properties of Schrödinger operators on perturbed lattices. Ann. Henri Poincaré, 17(8):2103–2171, 2016.
- [2] D. Bättig. A toroidal compactification of the two dimensional Bloch-manifold. PhD thesis, ETH Zurich, 1988.
- [3] D. Bättig. A directional compactification of the complex Fermi surface and isospectrality. In Séminaire sur les Équations aux Dérivées Partielles, 1989–1990, pages Exp. No. IV, 11. École Polytech., Palaiseau, 1990.
- [4] D. Bättig. A toroidal compactification of the Fermi surface for the discrete Schrödinger operator. Comment. Math. Helv., 67(1):1–16, 1992.
- [5] D. Bättig, H. Knörrer, and E. Trubowitz. A directional compactification of the complex Fermi surface. Compositio Math., 79(2):205–229, 1991.
- [6] N. Do, P. Kuchment, and F. Sottile. Generic properties of dispersion relations for discrete periodic operators. J. Math. Phys., 61(10):103502, 19, 2020.
- [7] M. Faust and F. Sottile. Critical points of discrete periodic operators. arXiv preprint arXiv:2206.13649, 2022.
- [8] L. Fisher, W. Li, and S. P. Shipman. Reducible Fermi surface for multi-layer quantum graphs including stacked graphene. Comm. Math. Phys., 385(3):1499–1534, 2021.
- [9] D. Gieseker, H. Knörrer, and E. Trubowitz. An overview of the geometry of algebraic Fermi curves. In Algebraic geometry: Sundance 1988, volume 116 of Contemp. Math., pages 19–46. Amer. Math. Soc., Providence, RI, 1991.
- [10] D. Gieseker, H. Knörrer, and E. Trubowitz. The geometry of algebraic Fermi curves, volume 14 of Perspectives in Mathematics. Academic Press, Inc., Boston, MA, 1993.
- [11] R. Hartshorne. Algebraic geometry. Graduate Texts in Mathematics, No. 52. Springer-Verlag, New York-Heidelberg.
- [12] H. Isozaki and H. Morioka. A Rellich type theorem for discrete Schrödinger operators. Inverse Probl. Imaging, 8(2):475–489, 2014.
- [13] H. Knörrer and E. Trubowitz. A directional compactification of the complex Bloch variety. Comment. Math. Helv., 65(1):114–149, 1990.
- [14] P. Kuchment. An overview of periodic elliptic operators. Bull. Amer. Math. Soc. (N.S.), 53(3):343–414, 2016.
- [15] P. Kuchment and B. Vainberg. On absence of embedded eigenvalues for Schrödinger operators with perturbed periodic potentials. Comm. Partial Differential Equations, 25(9-10):1809–1826, 2000.
- [16] P. Kuchment and B. Vainberg. On the structure of eigenfunctions corresponding to embedded eigenvalues of locally perturbed periodic graph operators. Comm. Math. Phys., 268(3):673–686, 2006.
- [17] W. Li and S. P. Shipman. Irreducibility of the Fermi surface for planar periodic graph operators. Lett. Math. Phys., 110(9):2543–2572, 2020.
- [18] W. Liu. Fermi isospectrality for discrete periodic Schrödinger operators. Comm. Pure Appl. Math. to appear.
- [19] W. Liu. Irreducibility of the Fermi variety for discrete periodic Schrödinger operators and embedded eigenvalues. Geom. Funct. Anal., 32(1):1–30, 2022.
- [20] W. Liu. Topics on Fermi varieties of discrete periodic Schrödinger operators. J. Math. Phys., 63(2):Paper No. 023503, 13, 2022.
- [21] W. Shaban and B. Vainberg. Radiation conditions for the difference Schrödinger operators. Appl. Anal., 80(3-4):525–556, 2001.
- [22] S. P. Shipman. Eigenfunctions of unbounded support for embedded eigenvalues of locally perturbed periodic graph operators. Comm. Math. Phys., 332(2):605–626, 2014.
- [23] S. P. Shipman. Reducible Fermi surfaces for non-symmetric bilayer quantum-graph operators. J. Spectr. Theory, 10(1):33–72, 2020.