1. Introduction
Picard–Vessiot theory is a Galois theory for linear ordinary differential equations. It studies the space of solutions of the operator in relation to an algebraic group, its differential Galois group. This theory has important connections with a long-standing conjecture, the Jacobian conjecture [1], which is related with the theory of commuting differential operators through the Dixmier conjecture [2] and Beret’s conjecture [3].
At the beginning of the 20th century, the solutions of differential equations as a function of a parameter caught the attention of some mathematicians. This interest seems clearly reflected in the work of J. Drach [4]. Recently, differential Galois theories began to be used to study spectral problems , for ordinary differential operators L of second order [5,6,7,8]. Following Drach’s ideology [4], the spectral parameter was considered an algebraic variable in [8], where a Picard–Vessiot theory for algebro-geometric operators with parameters began to develop. We can say that it is a Picard–Vessiot theory for commuting differential operators and, consequently, is governed by a spectral curve. Parametric Galoisian theories, for free algebraic parameters, have existed since the pioneer parametric Picard–Vessiot theory of Cassidy and Singer [9], and have been studied for second-order operators in [10], and for third-order operators in [11], but their connection with spectral problems has not been established thus far.
It has been well-known since the beginning of the 20th century [12,13], that the existence of an operator M that commutes with L forces the consideration of the spectral problem as a coupled spectral problem:
(1)
where and are tied by the constant coefficient polynomial equation that defines an algebraic curve . In the case of a Schrödinger operator L, the existence of a non-trivial M commuting with L is closely related to the Korteweg–de Vries hierarchy of integrable non-linear differential equations. The solutions of (1) depend on the geometry of the spectral curve , as it is exhaustively studied in the works of J. J. Morales-Ruiz and the authors of [7,8], where spectral Picard–Vessiot theory appears for the first time.In this work, we study the spectral problem for a third-order operator L, under the assumption that L has a nontrivial centralizer and that it contains M as in (1). The development of a spectral Picard–Vessiot theory for this type of operator is an open problem. Since a third-order L is related to a more complicated integrable hierarchy, the Boussinesq hierarchy, there is even a lack of explicit examples due to computational problems [14,15].
A famous theorem by I. Schur [16] implies that the centralizer of an ordinary differential operator L, with analytic coefficients, has as quotient field a function field in one variable. Therefore, such a centralizer can be seen as the affine ring of an algebraic curve [17]. In this work, we will call this curve the spectral curve of the operator L, and assume that L has coefficients in an arbitrary differential field of zero characteristic.
A fundamental goal driving this work is the development of a Picard–Vessiot theory for spectral problems , in the case of an algebro-geometric differential operator L and an algebraic parameter over . In order to construct a splitting field for , it is crucial to note that is not a free parameter, but rather it is governed by the spectral curve of L. For this purpose, to achieve the factorization of , a new coefficient field , containing and , must be considered.
We believe that the novel results that emerge from the study of the order three case will be important for the development of a spectral Picard–Vessiot theory for arbitrary order. Specifically, our main contribution is the identification of the spectral curve of L, as the curve defined by its centralizer, and the computation of its defining ideal in the case of third-order operators. The field of functions on the spectral curve of L is now controlled and allows effective computation of the first-order intrinsic right factor of . It is intrinsic because, via the specialization of the parameter to each constant value , it gives the rank one sheaf over , made famous by the works of Krichever [18] and Mumford [19], where the correspondence between commutative subrings of ODOs and affine spectral curves is established; see [20] for a recent review. In Section 2, an extended version of this contributions is included.
2. Contributions
We consider ordinary differential operators with coefficients in an arbitrary differential field , whose field of constants is algebraically closed and of zero characteristic. We use differential algebra, differential Galois theory and algebraic geometry to give a constructive proof of the following theorem.
The centralizer of a third-order ordinary differential operator L, with non-constant coefficients in Σ, is either isomorphic to a polynomial ring in one variable or to the coordinate ring of an irreducible (space) algebraic curve Γ.
A differential operator L whose centralizer is not a polynomial ring in one variable is often called algebro-geometric [21,22]. Our results establish a new framework to study spectral problems for algebro-geometric ODOs.
The authors defined spectral Picard–Vessiot fields for algebro-geometric second-order operators in [8] and provided algebraic tools to effectively compute a basis of solutions for the spectral problem in this case. The extension to a higher order of the techniques developed for second-order operators poses several important issues. The main problem treated in this paper is the definition and computation of spectral curves for algebro-geometric ODOs.
Computing algebro-geometric ODOs amounts to computing Gelfand–Dikii hierarchies and solutions of them [23], which are non-trivial problems. It is well-known that solutions of the KdV hierarchy provide algebro-geometric Schrödinger ODOs [7,8,24]. Solutions of the Boussinesq hierarchy [25] provide algebro-geometric operators of third order, and investigating those, we discovered that planar algebraic curves do not explain all nontrivial centralizers of third-order ODOs. We found an example of an algebro-geometric third-order ODO whose centralizer is the commutative -algebra generated by three ODOs [26]. Explaining the nature of this example was the seed for the results presented here. A general study of algebro-geometric differential operators of prime order is postponed until meaningful examples are found, for order five and higher.
Planar spectral curves have been commonly defined for pairs of commuting differential operators P and Q, since the seminal work of Burchnall and Chaundy [12]; see also [18,19] and the recent review [20]. Most results have been obtained in the case of commuting pairs with coprime orders, called the rank 1 case; see, for instance, [20]. Without this restriction, hence for any rank , we define an isomorphism between the coordinate ring of the planar spectral curve, an irreducible plane algebraic curve , and the commutative ring of differential operators . Namely, we prove that the ideal of the curve equals the ideal of all constant coefficient polynomials that are satisfied by the differential operators, commonly denoted , and known as Burchnall and Chaundy polynomials. We introduce in this paper the notion of Burchnall and Chaundy ideal of a pair and denote it by . Since P and Q are differential operators in a Euclidean domain, the ideal is naturally proved to be a prime ideal. These results are included in Theorem 3. From now on, we will call the spectral curve of the pair to distinguish it from the spectral curve associated with the centralizer of an ordinary differential operator.
It was known by E. Previato [27] and G. Wilson [14], in the case of differential operators with analytic coefficients, that a Burchnall and Chaundy polynomial for a commuting pair can be obtained by computing the differential resultant of and . In particular, they proved that such a resultant is a constant coefficient polynomial. We generalize this result in Theorem 5 to differential operators with coefficients in . We go further, proving that the ideal of all Burchnall and Chaundy polynomials for the pair is generated by an irreducible polynomial , the radical of . For this purpose, we use the differential elimination ideal determined by and ,
the ideal of all differential operators generated by and in , for which the derivation has been eliminated. Our main result about arbitrary commuting pairs of ODOs establishes the intrinsic nature of the differential resultant of and . We prove the following result in Theorem 8.Given commuting ordinary differential operators P and Q with coefficients in Σ, let be the radical of . The following statements hold:
The principal ideal generated by f in equals the prime ideal .
The differential ideal generated by f in equals the radical of the elimination ideal .
For an algebraically closed field of constants, the classical theory of Picard and Vessiot [28] allows us to formally manage solutions, and it will be important to give formal proofs of our results. Specializing the parameter to , the Picard–Vessiot extension of for is the splitting field for . If algebraic variables and over are considered, the differential operators and belong to the differential field and to use classical Picard–Vessiot theory, the algebraic closure of can be considered to prove Theorem 5.
Centralizers are maximal commutative rings. In some cases, the centralizer of a differential operator L is the affine ring of a planar curve, as in the case of algebro-geometric Schrödinger operators where . Let us fix now an algebro-geometric differential operator L of third order with coefficients in . One of the motivations to give a constructive proof of Theorem 1 is to show that in general, is not the coordinate ring of the planar curve. As a consequence of the results of K. Goodearl in [29], it has the structure of a free -module
where is of minimal order congruent with . Observe that contains the rank one algebras , but also , whose rank may be grater than one, being the rank of a set, the greatest common divisor of the orders of its elements; see [17]. We introduce the notion of Burchnall–Chaundy ideal of L to explain the ring multiplicative structure of . This is the ideal of all polynomials in such that . We prove that is a prime ideal that naturally contains the Burchnall–Chaundy ideals of pairs of differential operators in the centralizer of L. Three BC ideals are of crucial importance in this work: all of them contained in , but even more so we prove in Theorem 10 that which is one of the main contributions of this article, since it allows the computation of the spectral curve. The constructive proof of Theorem 1 is achieved in Corollary 3, where is shown to be isomorphic to the coordinate ring of an irreducible space algebraic curve in , which we define as the spectral curve of L.An important consequence of the constructive proof of Theorem 1 is to discover the appropriate coefficient field where a third-order algebro-geometric operator would have an intrinsic right factor. Recall that is not a free parameter; it is governed by the space spectral curve of L. Extending the ideal to the ring of polynomials in three variables with coefficients in , we can think of an extended curve whose ring of regular functions is
Over its quotient field , a factorization of is guaranteed, where can be computed by means of the greatest common right divisors or , which are proved to coincide over . The precise statement is contained in Theorem 12. We can think of as a global factor, since for almost every point , our methods produce a factorization of . The specialization of to each point of provides the rank one sheaf given by the solutions of . Therefore, starting with the centralizer of L, both the spectral curve and the previous rank one sheaf of the Mumford correspondence in [19] are now effectively computed.It remains as a future project to define and prove the existence of the spectral Picard–Vessiot field of , for an algebro-geometric differential operator L. It would be a differential field extension of , the minimal extension containing all the solutions, and it requires a full factorization of over . This requires further investigations where the parametric Picard–Vessiot theory, introduced by Cassidy and Singer in [9] and studied in [10,11], may be relevant.
The paper is organized as follows. In Section 3, we establish the appropriate framework to study algebro-geometric differential operators, in the case of a third-order operator L. The novel notions of Burchnall–Chaundy (BC) ideal, of a pair of ODOs and the BC ideal of a fixed differential operator L are introduced in Section 4. These BC ideals are proved to be prime ideals describing the algebra and the centralizer , as coordinate rings of algebraic curves, respectively, in Theorems 3 and 4. These allow the definitions of the spectral curves of a pair and of a third-order operator L.
Section 5 and Section 6 are dedicated to computing generators of these BC ideals. Section 5 contains the proof of Theorem 2, the first part in Theorem 6 and the second in Theorem 8. The effective description of is achieved by means of any basis of the centralizer and the computation of differential resultants. The choice of a basis of is discussed in Section 6.1. We use Gröbner bases in Section 6.2 to control the projection of the space spectral curve onto the plane and obtain a set of generators of in Theorem 10.
In Section 7, we show that the algebro-geometric hypothesis implies the existence of a right factor of , over a new differential field of coefficients , defined in Section 7.1. The computation of is delicate and it is achieved by means of differential subresultants in Section 7.2. Finally, Section 8 contains the first computed example of a non-planar spectral curve. It is used to illustrate how to globally factor the linear operator . All computations in this work are performed with Maple [30].
To make this work as self-contained as possible, two appendices are included. Appendix A contains the results needed to understand centralizers of differential operators, giving special importance to Goodearl’s construction of a basis of . In Appendix B, we review definitions and proofs of results on differential resultants and subresultants in relation to the factorization of ODOs.
For concepts in differential algebra, we refer to [28,31,32]. A differential ring is a ring R with a derivation ∂ on R. A differential ideal I is an ideal of R invariant under the derivation. We denote by
which is called the ring of constants of R. Assuming that R is a differential domain, its field of fractions is a differential field with extended derivation A differential field is a differential ring which is a field. Given , we denote by . Note that is a field whenever is. We assume that is algebraically closed and has characteristic 0.Let us consider algebraic variables and with respect to ∂. Thus, and , and we can extend the derivation ∂ of to the polynomial ring having ring of constants . Similarly, we consider algebraic variables and with respect to ∂ and extend the derivation to the polynomial ring having ring of constants .
We denote by the set of non-negative integers including 0. Notation regarding ODOs in and their centralizers is included in Appendix A.
3. Algebro-Geometric ODOs
Let us consider an ordinary differential operator L in , with non-constant coefficients, that is . We will assume throughout this work that L has a non-trivial centralizer, as in Definition A1; this means that the centralizer of L in
does not equal the polynomial ring . Therefore, we have a chain of strict inclusions In Appendix A, we review results of K. Goodearl in [29] regarding the structure of centralizers.We call a differential operator
By the generalization of Schur’s theorem in [16], (A6) in the Appendix A, we know that is commutative, it is a differential domain, and is an (abstract) algebraic curve. Spectral curves are defined for any commutative subring B of as ; see [19,20]. As observed in [17], Section 2.1, the centralizer is a maximal commutative subring of . We emphasize in this work the importance of spectral curves for maximal commutative subrings of , namely spectral curves for centralizers.
Given an algebro-geometric differential operator , we define the
Let us consider now a third-order ordinary differential operator L in . We will assume throughout this work that L has a non-trivial centralizer. In this work, we identify a defining ideal for the abstract curve . This is Corollary 3, obtained as a consequence of Theorem 10.
By Theorem A1, the centralizer is a free -module of rank 3. Let be a basis of as a -module. The construction of a basis is reviewed in Appendix A. Each is a monic operator in of minimal order . Therefore,
(2)
From the composition of the basis, we obtain the following result.Let L be a third-order operator in . The following statements are equivalent:
-
L has a nontrivial centralizer .
-
There exists an operator A in of order m, relatively prime with 3, such that .
The second statement of Corollary 1 is the traditional definition of algebro-geometric operator; see, for instance, [21,22]. This statement highlights that contains a pair of operators of rank one, the rank of the pair being the greatest common divisor of their orders. A discussion on the rank of a set of differential operators appears in [17] .
Since centralizers are maximal commutative subrings of , given a differential operator M that commutes with L, we have the sequence of inclusions
and all of them could be strict. In the case of a third-order operator, the following ring diagram of inclusions illustrates the ring structure of .(3)
Making use of the ring structure of , it may happen that equals a polynomial in or vice versa, , with q a univariate polynomial. In this situation, . For instance, this is the case whenever , then has order 4 and is the centralizer of a second-order operator.
4. Burchnall–Chaundy Ideals and Spectral Curves
A polynomial with constant coefficients satisfied by a commuting pair of differential operators P and Q is called a
4.1. Burchnall–Chaundy Ideal of a Pair
Given commuting differential operators P and Q in , to avoid meaningless situations, we will assume that and both have positive order. We define the ring homomorphism
(4)
It is a ring homomorphism since P and Q commute, and they commute with the elements of , the field of constants of . The image of is the -algebra Observe that is an ideal of . In this setting, given , we will denote(5)
Thus, a polynomial g belongs to the kernel of if and only if , that is,Given commuting differential operators P and Q in , both of positive order, we define the
An important consequence of working with differential operators in a Euclidean domain is the following lemma.
Given commuting differential operators P and Q in , both of positive order, the ideal is a prime ideal in .
To start, is a nonzero ideal by [29], Theorem 1.13. For details, let us consider the centralizer of P. Since is a finitely generated -module, there exists such that
Given and in , let us assume that . Observe that and are differential operators in the Euclidean domain . Since is a ring homomorphism
implies or . Thus, or , proving that is prime. □Given commuting differential operators P and Q in , both of positive order, we define the
(6)
We proved in Lemma 1 that is a prime ideal. Then the algebraic variety is an irreducible algebraic variety in . This variety cannot be a single point , because this would mean that , implying and , which contradicts . Therefore, is an irreducible algebraic curve in .
The following result summarizes the situation.
Given differential operators P and Q in , both of positive order, if P and Q commute, then the following statements hold:
-
is a commutative domain isomorphic to , the coordinate ring of the spectral curve .
-
There exists an irreducible polynomial such that .
As a consequence of the construction of and Lemma 1, we obtain 1. In addition, recall that is an algebraically closed field. By Remark 2, is an irreducible algebraic curve in , meaning that its defining ideal is generated by an irreducible polynomial . □
In Section 5, we treat the computation of the defining polynomial f of the spectral curve .
4.2. Burchnall–Chaundy Ideal of a Third-Order Operator
Let us consider a third-order operator L in with nontrivial centralizer . We will define next the Burchnall–Chaundy ideal of the operator L.
Given a basis of , we can define the ring homomorphism
(7)
We will denote monomials in by , , whose image is . Thus, the image of is the centralizer of L,Given , we will denote
(8)
Given a third-order operator L in , with nontrivial centralizer and given a basis of , we define a
(9)
We will call the elements of the BC idealGiven a third-order operator L in , with nontrivial centralizer , the ideal is a prime ideal in .
Given a basis , since L and commute, by Lemma 1, is a nonzero ideal. Observe that , which implies that is also a nonzero ideal.
As in Lemma 1, we can prove that is a prime ideal using the fact that is a ring homomorphism and that is a Euclidean domain. □
We obtain the following result as a consequence of the construction of and Lemma 2.
Given a third-order operator L in , with nontrivial centralizer , and given any basis of , then the next isomorphism holds:
Thus, is the affine ring of an algebraic curve whose defining ideal is .
Given a third-order operator L in , with nontrivial centralizer , we define the
We will compute in Section 6.2 a finite set of defining polynomials for . In general, the spectral curve of a third-order operator will be a space curve, and only in some special cases will it be a planar curve.
In some cases, described in Remark 1, if then . Therefore, the spectral curve of L would be the spectral curve of the pair a plane curve.
5. Elimination Ideals for Commuting Pairs of ODOs
Given a pair of commuting differential operators in , both of positive order, we know that the BC-ideal is a prime ideal by Lemma 1. This section is dedicated to the proof of Theorem 2, which is obtained as a consequence of Theorem 8 in this section. Moreover, we start proving in Theorem 5 a generalization of a known result of E. Previato [27] and G. Wilson [14] to the case of a pair of commuting operators with coefficients in .
5.1. Generalized Previato–Wilson Theorem
Let us assume that P and Q have positive orders n and m, and leading coefficients and , respectively. Observe that the operators and have coefficients in the differential domain ; see the notation in Section 1. Let us consider the differential resultant of and , as defined in Appendix B:
By (A8), it provides a nonzero polynomial in , where belong to and have degree less than or equal to m.The next theorem was proved by E. Previato in [27], for differential operators whose coefficients are analytic functions; see also [14]. For completion, we review the proof given in [7], on this occasion for differential operators with coefficients in an arbitrary differential field with algebraically closed field of constant of zero characteristic.
Let us consider differential operators P and Q in , both of positive order. If P and Q commute, then is a polynomial in .
Recall that and are differential operators with coefficients in , whose field of fractions is . We can extend the derivation of to its algebraic closure , whose field of constants is known to be algebraically closed; see [33], Corollary 3.3.1. Since the ring of constants of is , then it holds that .
Let us consider a fundamental system of solutions of in a Picard–Vessiot extension of for , whose field of constants is and whose derivation is defined by . The natural extension of to allows us to consider the action of on the -linear space of the solutions of .
Since and commute, then are solutions of . Therefore, there exists an matrix M with coefficients in such that
Taking determinants and using Poisson’s formula in Theorem A4, we obtain Thus, . □Let us consider . For a commuting pair P, Q, the next corollary characterizes the existence of common solutions of the eigenvalue problem
(10)
It follows from Theorem A5 and the fact thatGiven commuting differential operators P and Q in , both of positive order, the spectral problem (10) has a non-trivial (common) solution in a Picard–Vessiot extension of Σ for (or ) if and only if .
5.2. Computing the Burchnall–Chaundy Ideal of a Pair
It is ensured by Theorem 5 that the differential resultant is a polynomial in . Resultants are also called eliminants, and this is the feature of resultants we will emphasize next. Let us consider the left ideal generated by and in :
(11)
and the elimination ideal(12)
which is a two-sided ideal of , and(13)
which is a two-sided ideal of . Observe that by Lemma A3, with differential domain of coefficients , and by Theorem 5, it follows that Thus, both elimination ideals are nonzero.It was proved by Wilson in [14] that , that is, , in the case of differential operators P and Q whose coefficients are complex-valued smooth functions of x defined in some (real or complex) neighborhood of . We use here the argument of [14], Proposition 5.3 in a more general framework, where differential operators have coefficients in an arbitrary differential field , with field of constants algebraically closed of zero characteristic. For this purpose, we develop the following construction.
Considering as a -vector space with basis , we can define the -linear map
(14)
In this setting, given , we will also denote . Thus, Observe that the restriction of to the subring of constants of is the ring homomorphism defined in (4), and also that(15)
We will proceed next to calculate a generator for both kernels. We will begin by demonstrating that the differential resultant belongs to the ideal , that is, h is a Burchnall–Chaundy polynomial. This is Theorem 6, and to prove it, we will need some auxiliary results.
Given commuting differential operators P and Q in , both of positive order, with the previous notation, it holds that
Given a polynomial g in , then
Observe that the differential operator has finite order, .Let be a constant in . Since is algebraically closed, there exists in the nonempty set . By Corollary 2, there exists a common eigenfunction for the coupled spectral problem (10). Consequently, the following is an infinite set of linearly independent eigenfunctions
since eigenfunctions associated to different eigenvalues are linearly independent. Now, for every , it holds that where and are the result of evaluating the coefficients of C and D in . Moreover, is included in the -linear space of solutions of the equation , whose dimension is q. Then is the zero operator. □As a consequence of Lemma 3, Formula (15) and Theorem 5, the following result is proved.
Given commuting differential operators P and Q in , both of positive order, then
Moreover, for some non-zero natural number and f is the irreducible polynomial such that .
The radical of the ideal generated by h in equals ; see, for instance, [34], page 178, Proposition 9, where f is the radical of h, that is,
(16)
is the square-free part of h. If and , observe that, by Section 5.1, the degrees in and of h are, respectively,(17)
then divides , and , .Let us consider the pure lexicografic monomial ordering in with . Given , by the Division Theorem (see [34], Theorem 3, page 64), we can write , where ,
(18)
since the leading monomial of f does not divide . We call the normal form of gLet us consider a polynomial , and the differential ideal generated by g in the differential ring , whose field of constants is . Observe that and , since . Furthermore, is the ideal of generated by g, but we use the notation to distinguish it from the ideal generated in by g.
Given commuting differential operators P and Q in , both of positive order, with the previous notation, it holds that .
By Theorem 6, we know that . We observe that . Let us prove the other inclusion.
Given , let us consider its normal form w.r.t. f as in (18). If we denote , then , . Let us define
Observe that and , only because f is constant. Since and then is a zero operator. If we assume that at least one of the terms of H is nonzero, that is, , then by (A3) for distinct . Reorganizing and dividing by , it holds that(19)
where and , . By (17), we have . If , then by (19), and , contradicting that and are coprime. If , then , with . Hence, , implying , which is a contradiction. Therefore, we conclude that , in other words, that is the zero polynomial. This proves that , that is, . □Let us consider commuting differential operators P and Q in , both of positive order, and , with . Then is a prime differential ideal in .
As defined in (6), let be the spectral curve of the pair . For each point , by Corollary 2, there exists a common solution in a Picard–Vessiot extension of for .
Let us consider and assume that . Observe that
Thus, we can write the following partition of : Having an algebraically closed field of constants implies that at least one of the two components, say , has an infinite number of points.We then have an infinite set of linearly independent eigenfunctions . For every , we have
This contradicts the fact that is a finite-order differential operator. This proves that . By Lemma 4, . □We proceed next to emphasize the relationship between the elimination ideals defined in (12) and (13) and the ideal of Burchnall–Chaundy polynomials of a commuting pair of ODOs.
Let us consider commuting differential operators P and Q in , both of positive order, and , with . It holds that
-
The radical of the elimination ideal equals .
-
The radical of the elimination ideal equals .
By Lemma 3 and Lemma 4, we obtain
using also Equality (15) and Theorem 6. Now, taking radicals and applying Theorem 7, the following inclusions hold: where the last equality is satisfied since is an algebraically closed field; see [35].Moreover, the differential resultant is a polynomial in , by Theorem 6. Consequently, f belongs to the radical . Thus, we have . □
Summarizing, the spectral curve of a commuting pair is defined by the irreducible polynomial , with . The commutative -algebra is isomorphic to the coordinate ring of this curve:
In addition, the prime differential ideal determines a differential domain whose fraction field, denoted by , will play a key role in the study of the coupled spectral problem:In some cases, the centralizer of a differential operator L is the affine ring of a planar curve, as in the case of algebro-geometric Schrödinger operators where . The spectral Picard–Vessiot extension of for , introduced in [8], is a Liouvillian extension of determined by the solution of the first-order greatest common right divisor of and , as differential operators with coefficients in .
6. Centralizers as Coordinate Rings
Let be a third-order operator, whose centralizer is non-trivial. We proved in Lemma 2 that is a prime ideal, and in Theorem 4 that
Our next goal is to obtain a computational description of . We will use differential resultants for this purpose.6.1. Normalized Basis
By Theorem 6, each operator A in the centralizer , together with L, satisfies the algebraic equation defined by an irreducible polynomial defined by the differential resultant , that is,
The Tschirnhaus transformation of gives a new polynomial(20)
satisfied by L and the operator(21)
Observe that, by Theorems 3 and 8, In addition, and by (20) , is an irreducible polynomial because is irreducible over . Thus, Theorem 8 impliesRecall that any basis of is a minimal-order basis; see Proposition A2 in Appendix A.
Let be a third-order operator, whose centralizer is non-trivial. We will call a
The next result shows that any other basis of as a -module is determined by a normalized basis.
Let be a third-order operator, whose centralizer is non-trivial. Let us consider a normalized basis of . Then for any other basis of , it holds that
(22)
with , , . Moreover, if is also a normalized basis, then
We know that . W.l.o.g., let us assume that . Since is a minimal-order basis, let us assume that has minimal-order congruence with i (). Then and , together with (2), implies
with , and .Let us prove that is identically zero. We know that , where is a polynomial of degree 3 in . Thus,
(23)
Since is a normalized basis, it holds that(24)
Computing , for instance, with Maple and using (24) to replace , we obtain only the monomials and of degree three in and in (25). Comparing with the r.h.s. of (23), we obtain that the coefficients of the monomials of degree three must vanish:(25)
Thus, must be identically zero and we obtain If is also a normalized basis, the r.h.s. of (23) cannot have a term , then must be the zero polynomial and . We analogously prove that . □The previous result explains that a normalized basis is unique up to multiplication by constants. We include this result for completion but we do not need to choose the normalized basis in the remaining parts of this work.
6.2. Generators of the Burchnall–Chaundy Ideal of a Third-Order ODO
Let us consider a basis of . Let us denote
if is the normalized basis, then Recall that are irreducible and , ; see Section 5.2. In addition, let us denote by the irreducible polynomial in obtained as the radical of We have .The situation so far is described by the following chain of ideals in and we will determine first if the last two inclusions are identities:
(26)
Let us consider the following algebraic varieties defined by the previous ideals: By the inclusions in (26), we obtain . Observe that is the intersection of the irreducible surfaces defined by and ; therefore, is a space algebraic curve.The Zariski closure of the projection of onto the plane is a plane algebraic curve defined by the square free part of the algebraic resultant w.r.t. ; see [34], Chapter 3, §2 and [36], Section 2. This projected curve could be irreducible or not. Observe that , and therefore, . Thus, there are two possible situations:
Irreducible. If is irreducible, then and is an irreducible curve.
Non-irreducible. If is not irreducible, then properly divides r and . Including in the ideal allows us to select one irreducible component of . The example in Section 8 illustrates this situation.
We will prove next that is the defining ideal of an irreducible space curve, in both situations. By Remark 2, the irreducible component of determined by is a proper curve; it cannot be a point.
Let us assume that the projection over the plane of the algebraic curve β, and therefore, γ is birational. Note that this holds for almost all projections (see [35] [Fulton], p. 155), and hence, we can make this assumption w.l.o.g. More precisely, a valid projection direction can be achieved by an affine change of coordinates (see [36], Section 2, which establishes an isomorphism between coordinate rings; see, for instance, [37], Theorem 2.24).
A Gröbner basis of w.r.t. the pure lexicographic monomial ordering with verifies:
-
The polynomial is an implicit representation of the Zariski closure of the projection of γ onto the plane . Furthermore, because is irreducible over . Thus, this projection is an irreducible algebraic curve .
-
On the other hand, since this projection is assumed birational on γ, then contains a linear polynomial in λ, say .
Let us consider the pure lexicographic monomial ordering with in . Given , by the Division Theorem [34], Theorem 3, page 64, we can write
(27)
where . We callLet us consider a Gröbner basis of w.r.t. the pure lexicographic order with . The normal form of w.r.t is a polynomial in whose degree in is less than .
By the Division Theorem, no monomial of is divisible by the leading terms and ; thus, the result follows. □
We are ready to prove that .
Let L be a third-order operator in , whose centralizer is non-trivial. Given a basis of the centralizer, let , be the irreducible polynomials over such that and . Then
Given , let be its normal form w.r.t . Then . Furthermore, by Lemma 5, then and . We conclude that is identically zero, that is, . □
Let be a third-order operator, whose centralizer is non-trivial. Then
is an irreducible algebraic curve whose defining ideal is , with defined above. Furthermore,
7. Parametric Factorization of Algebro-Geometric ODOs
Let L be a third-order operator in , whose centralizer is non-trivial. We consider next the factorization of , for an algebraic parameter over . We know that is not a free parameter, but that it is governed by the spectral curve of L. Thus,
Recall that is a differential ring with the extended derivation ∂; see the notation in Section 1.Let us consider a basis of and the irreducible polynomials , in such that and . By Theorem 10,
(28)
As differential operators in , the pairs and are right coprime, since their differential resultants are nonzero, by Theorem A2. To consider the factorization of , we need an appropriate differential field of coefficients.7.1. Coefficient Field for Factorization
Associated to L we have defined in (9) the prime ideal ; see Lemma 2. We will prove in this section that the differential ideal is a prime ideal in . Thus, we can define the differential domain
(29)
and its fraction field , which therefore is a differential field.Considering as a -vector space with basis
we can define the -linear map(30)
where is the ring homomorphism defined in (7). Observe that given and , then(31)
Let us consider a Gröbner basis of w.r.t. the pure lexicographic order with . Given , the normal form of g w.r.t verifies
With notation as in (27), by (31), we have
Thus, and by Lemma 5, we know that . Therefore, by Lemma 4,□
With the previous notation, it holds that .
We will show next that the inclusion of in is natural. Since has constant coefficients, the next differential ideal coincides with the ideal generated by in :
In addition, by (31), , which proves that .Conversely, given , by Lemma 6, . In addition, by Lemma 5, . Therefore is identically zero, proving that . □
Let L be a third-order operator in , whose centralizer is non-trivial. Then the differential ideal is a prime ideal in .
Given , let us assume that , with . By Lemma 6, . By Lemma 5, and using (27), we can write
Thus, Therefore, . This implies that , which is a prime ideal in , by Theorem 7. We can conclude that , which shows that , for or , proving that is a prime ideal. □We are now ready to work over the field , the fraction field of the domain defined in (29). Regarding the differential structure of , we consider the standard differential structure of the quotient ring given by a derivation defined by:
(32)
Observe that is a derivation in because is a differential ideal. By abuse of notation, we will denote by ∂ the derivation and its extension to the fraction field .7.2. The Intrinsic Right Factor
We will study next the factorization of as a differential operator with coefficients in the differential field , as defined in Section 7.1.
As differential operators in , and , have a non-trivial common factor, namely their monic greatest common right divisor . For instance, this follows from the Resultant Theorem A5, since is zero in . Similarly, for some positive integer r, it holds that , which is zero in , implying that is a differential operator of order greater than or equal to one.
We will prove next that and are indeed differential operators of order 1 in . Moreover, the goal of this section is to show that and are equal as differential operators in , providing a right factor of that is intrinsic to the nature of L. This factor is linked to the hypothesis of having a nontrivial centralizer.
We chose differential subresultants to compute greatest common right divisors, because they have fairly explicit expressions for their computation that also allow us to obtain some important theoretical conclusions. As defined in Appendix B.1, let us consider the first differential subresultants:
where, for(33)
(34)
Observe that the coefficients are nonzero polynomials in , but we need to check if they are nonzero in . Recall that from Section 3, .With the notation established above, the class , , in is non-zero. In addition, if , then the class of is non-zero.
By the construction of the matrices as in (A11), the degree in of is less than 3, for . Thus, , does not belong to because otherwise it is included in
which is not possible. Similarly, if , then the degree in of is less than and does not belong to because otherwise, it is included in Observe that in the case then has degree in . □By Lemma 8, are nonzero differential operators in of order one. The First Subresultant Theorem A3 implies that equals the first subresultant. We can make them monic and write
(35)
The next lemma will be important to prove that the right factors coincide over the field of the spectral curve . For this purpose, let us define the elimination ideal
(36)
With the previous notation, it holds that .
Given a polynomial g in , then
(37)
With the notation in Remark 5 and by Corollary 3, every point of is determined by a point of the projection of the curve onto the plane , in the following way:
(38)
except for a finite number of points such that .As in Remark 5, let us consider a Gröbner basis of w.r.t. the pure lexicographic monomial ordering with . Recall that , with . By (38), since , we have
(39)
Given , there exists , the Zariski closure of the projection of onto the plane . Let be the Picard–Vessiot extension of for . By Corollary 2, since , there exists , a common eigenfunction of the spectral problem
(40)
Thus, is an infinite set of linearly independent eigenfunctions for (40). Observe that, by (39), Thus,(41)
To finish the argument, for every , by (40) and (41),
(42)
(43)
using (37) and obtaining , and in . Since is an infinite set, we can conclude that , which proves the result. □Let us denote by the class in . Looking at these functions as representatives of objects in , we can establish the following result.
With the previous notation. Let us consider as in (35), then
(44)
If , then .By the definition of subresultants in Appendix B.1, we know that
The next one is a differential polynomial in : Observe that it belongs to the elimination ideal as defined in (36). In addition, by Lemma 1, Therefore, in . This proves (44). □We are ready to state an important application of the results of this paper.
Let L be a third-order operator in , whose centralizer is non-trivial. Let Γ be the spectral curve of L. Then has an order one factor in . We call
(45)
By the First Subresultant Theorem A3,
and by Lemma 9, Thus, (45) follows. □As a consequence of (45) and (35) , the intrinsic right factor of as a differential operator in can be computed as , for any of minimal order for or .
Let us assume that is in normal form. By direct computation, we obtain the factorization
(46)
in , under the condition(47)
The intrinsic right factor is a global factor in the following sense. For every , where Z is a finite number of points in , we obtain a right factor in of
where is the greatest common right divisor of and in and is the result of replacing in by . The set of points Z to be removed can be described as where , i = 1, 2, are finite sets of points, which is a consequence of (33) together with Remark 5. Furthermore, is the greatest common right divisor of , and for every .The factorization of an ordinary differential operator with coefficient in a differential field can be achieved by means of the so-called eigenring of L, defined by M. Singer in [38] as
where denotes the equivalence class of in the module . The centralizer can be seen as a subring of the eigenring, but note that the representative of order smaller than 3 in the eigenring of an element of the centralizer may not belong itself to the centralizer. In the example of Section 8, since , then has a representative in that does not belong to the centralizer.
One could study the factorization of , for any by means of the algorithms based on the eigenring, as in [39]. What these algorithms would not do is to identify the spectral curve. Furthermore, these algorithms would not allow a formal treatment of λ as an algebraic parameter over the coefficient field Σ. These algorithms would partially identify the spectral curve. They would treat the parameter λ in the algebraic closure of and therefore consider a branch of the spectral curve. See the example in Section 8.
8. Example of Non-Planar Spectral Curve
There are many examples of rank 1 operators whose centralizers are the ring of a plane algebraic curve; see references in. In this section, we present the first explicit example of a centralizer isomorphic to the ring of a non-planar spectral curve. We explained the computational method used to obtain this example in [15,26], based on Wilson’s almost commuting bases [14]. We use this example to illustrate the new results of this paper.
Let us consider the differential operator in , where :
(48)
The centralizer of L equals with The Burchall–Chaundy ideal of L equals generated by the irreducible polynomials Since with , then is strictly contained in . A Gröbner basis of in w.r.t. the pure lexicographic order is with and .The curve defined by is a non-planar curve parameterized by
(49)
This is the first explicit example of a non-planar spectral curve.The first differential subresultants of , and pairwise are equal to
see (33), with We have and ; thus, and Let us call the result of replacing by the parameterization of in , which equals(50)
Thus, as in (46). At every point of the spectral curve of L, we recover a right factor , for .Observe that whenever the spectral curve is a rational curve with rational parameterization , as in the previous example, one can consider the factorization of as a differential operator in , since the algebraic variable with , would be a free parameter over . To achieve a full factorization, it is a future project to use the parameterized Picard–Vessiot theory introduced by Cassidy and Singer in [9], and studied in [11]. It would be interesting to explore how this theory explains the factorization over the field of the spectral curve, using the results on second-order operators in [10] combined with the intrinsic factorization of Section 7.2.
9. Discussion
In this work, the computation of the space algebraic curve that describes the centralizer of third-order ODO L was achieved. We proved that is the spectral curve of a commutative algebra with three generators and, in general, it is a non-planar curve. We provided an effective description of the defining ideal of that will allow a parametric factorization of , for an algebraic parameter . The computation of the intrinsic right factor, of the Mumford correspondence [19], of was only described before in the case of planar spectral curves [25], in connection with the Baker–Akheizer function.
It is an ongoing project to develop the spectral Picard–Vessiot theory of the third-order operator . We intend to follow our previously successful strategy, in the case of second-order Schrödinger operators [8]. In analogy with the Galois theory of univariate polynomials, a full factorization of will be studied, in combination with ideas coming from the parametric Picard–Vessiot theory [9,10,11], that must be adapted to a non-free algebraic parameter . Through this journey, spectral curves, planar or not, will govern the factorization of and the hidden free parameters will emerge.
Conceptualization; methodology; validation; formal analysis; investigation; resources; writing—original draft preparation; writing—review and editing; supervision; project administration; funding acquisition: S.L.R. and M.-A.Z. All authors have read and agreed to the published version of the manuscript.
Not applicable.
Not applicable.
Data are contained within the article.
A sincere acknowledgement to Rafael Hernandez Heredero for the enlightening conversations on this topic, which contributed to a better understanding of the mechanisms behind integrability. The authors really appreciate the exchange of ideas regarding the non-planar spectral curves with Emma Previato; we are very grateful for the conversations we had on spectral problems and for her constant encouragement to pursuit this project.
The first author is a member of the Research Group “Modelos matemáticos no lineales”, Polytechnic University of Madrid, Madrid, Spain. The authors declare no conflicts of interest.
Footnotes
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
References
1. Crespo, T.; Hajto, Z. Picard–Vessiot theory and the Jacobian problem. Israel J. Math.; 2011; 186, pp. 401-406. [DOI: https://dx.doi.org/10.1007/s11856-011-0145-y]
2. Mironov, A.E.; Zheglov, A.B. Commuting ordinary differential operators with polynomial coefficients and automorphisms of the first Weyl algebra. Int. Math. Res. Not.; 2016; 2016, pp. 2974-2993. [DOI: https://dx.doi.org/10.1093/imrn/rnv218]
3. Guo, J.; Zheglov, A. On some questions around Berest’s conjecture. arXiv; 2024; arXiv: 2203.13343
4. Drach, J. Détermination des cas de réduction de l’équation différentielle d2y/dx2 = [ϕ(x) + h]y. C. R. Math. Acad. Sci. Paris; 1919; 168, pp. 47-50.
5. Brezhnev, Y. Spectral/quadrature duality: Picard-Vessiot theory and finite-gap potentials. Contemp. Math.; 2012; 563, 1.
6. Brezhnev, Y. Elliptic solitons, Fuchsian equations, and algorithms. St. Petersburg Math. J.; 2013; 24, pp. 555-574. [DOI: https://dx.doi.org/10.1090/S1061-0022-2013-01253-2]
7. Morales-Ruiz, J.J.; Rueda, S.; Zurro, M. Factorization of KdV Schrödinger operators using differential subresultants. Adv. Appl. Math.; 2020; 120, 102065. [DOI: https://dx.doi.org/10.1016/j.aam.2020.102065]
8. Morales-Ruiz, J.J.; Rueda, S.; Zurro, M. Spectral Picard–Vessiot fields for Algebro-geometric Schrödinger operators. Ann. L’Institut Fourier; 2021; 71, pp. 1287-1324. [DOI: https://dx.doi.org/10.5802/aif.3425]
9. Cassidy, P.; Singer, M. Galois Theory of Parameterized Differential Equations and Linear Differential Algebraic Groups. Differ. Equ. Quantum Groups (Irma Lect. Math. Theor. Phys.); 1991; 9, pp. 113-157.
10. Arreche, C. On the computation of the parameterized differential Galois group for a second-order linear differential equation with differential parameters. J. Symbolic Comput.; 2016; 75, pp. 25-55. [DOI: https://dx.doi.org/10.1016/j.jsc.2015.11.006]
11. Minchenko, A.; Ovchinnikov, A. Calculating Galois groups of third-order linear differential equations with parameters. Commun. Contemp. Math.; 2018; 20, 1750038. [DOI: https://dx.doi.org/10.1142/S0219199717500389]
12. Burchnall, J.; Chaundy, T. Commutative ordinary differential operators. Proc. Lond. Math. Soc.; 1923; s2-21, pp. 420-440. [DOI: https://dx.doi.org/10.1112/plms/s2-21.1.420]
13. Burchnall, J.; Chaundy, T. Commutative ordinary differential operators II. The Identity Pn = Qm. Proc. R. Soc. Lond.; 1931; 134, pp. 471-485.
14. Wilson, G. Algebraic curves and soliton equations. Geometry Today; Arbarello, E.; Procesi, C.; Strickland, E. Birkhäuser: Boston, MA, USA, 1985; pp. 303-329.
15. Jiménez-Pastor, A.; Rueda, S.L.; Zurro, M.A. Computing Almost-Commuting Basis of Ordinary Differential Operators. ACM Commun. Comput. Algebra; 2023; 57, pp. 111-118. [DOI: https://dx.doi.org/10.1145/3637529.3637531]
16. Schur, I. Über vertauschbare lineare Differentialausdrücke. Berl. Math. Gesellschaft Sitzungsbericht. Arch. Der Math. Beil.; 1904; 3, pp. 2-8.
17. Previato, E.; Rueda, S.L.; Zurro, M.A. Commuting Ordinary Differential Operators and the Dixmier Test. Sigma Symmetry Integr. Geom. Methods Appl.; 2019; 15, 23. [DOI: https://dx.doi.org/10.3842/SIGMA.2019.101]
18. Krichever, I.M. Commutative rings of ordinary linear differential operators. Funct. Anal. Appl.; 1978; 12, pp. 175-185. [DOI: https://dx.doi.org/10.1007/BF01681429]
19. Mumford, D. An Algebro-Geometric Construction of Commuting Operators and of Solutions to the Toda Lattice Equation, Korteweg de Vries Equation and Related Non-Linear Equations. Proceedings of the International Symposium on Algebraic Geometry; Kyoto, Japan, 7–11 July 1977; pp. 115-153.
20. Zheglov, A. Algebra, Geometry and Analysis of Commuting Ordinary Differential Operators; Moscow State University: Moscow, Russia, 2020.
21. Weikard, R. On commuting differential operators. Electron. J. Differ. Equ.; 2000; 2000, pp. 1-11.
22. Grigorenko, N. Algebraic-geometric operators and Galois differential theory. Ukrainian Math. J.; 2009; 61, pp. 14-29. [DOI: https://dx.doi.org/10.1007/s11253-009-0200-6]
23. Gel’fand, I.M.; Dikey, L.A. Asymptotic behaviour of the resolvent of Sturm-Liouville equations and the algebra of the Korteweg-de Vries equations. Russian Math. Surveys; 1975; 30, pp. 77-113. [DOI: https://dx.doi.org/10.1070/RM1975v030n05ABEH001522]
24. Veselov, A.P. On Darboux-Treibich-Verdier Potentials. Lett. Math. Phys.; 2011; 96, pp. 209-216. [DOI: https://dx.doi.org/10.1007/s11005-010-0420-6]
25. Dickson, R.; Gesztesy, F.; Unterkofler, K. A New Approach to the Boussinesq Hierarchy. Math. Nachr.; 1999; 198, pp. 51-108. [DOI: https://dx.doi.org/10.1002/mana.19991980105]
26. Heredero, R.H.; Rueda, S.; Zurro, M. Computing centralizers of third order operators. EACA 2022. Xvii Meet. Comput. Algebra Appl.; 2023; pp. 103-106.
27. Previato, E. Another algebraic proof of Weil’s reciprocity. Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. Rend. Lincei (9) Mat. Appl.; 1991; 2, pp. 167-171.
28. Put, M.v.d.; Singer, M.F. Galois theory of linear differential equations. Grundlehren der Mathematischen Wissenschaften; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2012; Volume 328.
29. Goodearl, K. Centralizers in differential, pseudo-differential and fractional differential operator rings. Rocky Mt. J. Math.; 1983; 13, pp. 573-618. [DOI: https://dx.doi.org/10.1216/RMJ-1983-13-4-573]
30. Maplesoft. Maple (Version 2023); Maplesoft, a Division of Waterloo Maple Inc.: Waterloo, ON, Canada, 2023.
31. Crespo, T.; Hajto, Z. Algebraic Groups and Differential Galois Theory. Graduate Studies in Mathematics; America Mathematics Sociation: Providence, RI, USA, 2011; Volume 122.
32. Morales-Ruiz, J.J. Differential Galois Theory and Non-Integrability of Hamiltonian Systems; Birkhäuser: Berlin, Germany, 1999.
33. Bronstein, M. Symbolic Integration I: Transcendental Functions (Vol. 1); Springer: Berlin/Heidelberg, Germany, 2013.
34. Cox, D.; Little, J.; O’Shea, D. Ideals, Varieties, and Algorithms. An Introduction to Computational Algebraic Geometry and Commutative Algebra; Undergraduate Texts in Mathematics Springer International Publishing: New York, NY, USA, 2007.
35. Fulton, W. Algebraic Curves: An Introduction to Algebraic Geomentry; Benjamin: New York, NY, USA, 1969.
36. Abhyankar, S.; Bajaj, C. Automatic parameterization of rational curves and surfaces IV: Algebraic space curves. ACM Trans. Graph.; 1989; 8, pp. 325-334. [DOI: https://dx.doi.org/10.1145/77269.77273]
37. Sendra, J.; Winkler, J.; Pérez-Díaz, S. Rational Algebraic Curves: A Computer Algebra Approach; Number 22 in Algorithms and Computation in Mathematics Springer: Berlin/Heidelberg, Germany, 2007.
38. Singer, M.F. Testing reducibility of linear differential operators: A group theoretic perspective. Appl. Algebra Eng. Commun. Comput.; 1996; 7, pp. 77-104. [DOI: https://dx.doi.org/10.1007/BF01191378]
39. Acosta-Humánez, P.; Morales-Ruiz, J.; Weil, J. Galoisian approach to integrability of Schrödinger equation. Rep. Math. Phys.; 2011; 67, pp. 305-374. [DOI: https://dx.doi.org/10.1016/S0034-4877(11)60019-0]
40. Atiyah, M.; MacDonald, I. Introduction to Commutative Algebra; Addison–Wesley: Reading, MA, USA, 1969.
41. Bueso, J.; Gómez-Torrecillas, J.G.; Verschoren, A. Algorithmic Methods in Non-Commutative Algebra; Klumer Academic Publishers: Dordrecht, The Netherlands, 2003.
42. McCallum, S.; Winkler, F. Resultants: Algebraic and Differential. Techn. Rep. J. Kepler Univ.; 2018; RISC18-08, 21.
43. Chardin, M. Differential Resultants and Subresultants. Proceedings FCT’91. Lecture Notes in Computer Science; Springer: Berlin/Heidelberg, Germany, 1991; Volume 529, pp. 471-485.
44. Li, Z. A subresultant theory for Ore polynomials with applications. Proceedings of the 1998 International Symposium on Symbolic and Algebraic Computation, ISSAC ’98; Rostock, Germany, 13–15 August 1998; pp. 132-139.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2024 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Spectral curves are algebraic curves associated to commutative subalgebras of rings of ordinary differential operators (ODOs). Their origin is linked to the Korteweg–de Vries equation and to seminal works on commuting ODOs by I. Schur and Burchnall and Chaundy. They allow the solvability of the spectral problem
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details


1 Applied Mathematics Department, E.T.S. Arquitectura, Polytechnic University of Madrid, Avda. Juan de Herrera 4, 28040 Madrid, Spain;
2 Mathematics Department, Autonomous University of Madrid, Ctra. Colmenar Km. 15, 28049 Madrid, Spain