arxiv: v3 [math.fa] 23 Sep 2016

Similar documents
Linear Systems and Matrices

Determinants Chapter 3 of Lay

Linear Algebra. Matrices Operations. Consider, for example, a system of equations such as x + 2y z + 4w = 0, 3x 4y + 2z 6w = 0, x 3y 2z + w = 0.

MORE NOTES FOR MATH 823, FALL 2007

EIGENVALUES AND EIGENVECTORS 3

MATH2210 Notebook 2 Spring 2018

a 11 a 12 a 11 a 12 a 13 a 21 a 22 a 23 . a 31 a 32 a 33 a 12 a 21 a 23 a 31 a = = = = 12

Determinants. Beifang Chen

MATH 2030: EIGENVALUES AND EIGENVECTORS

Chapter 2:Determinants. Section 2.1: Determinants by cofactor expansion

Chapter SSM: Linear Algebra Section Fails to be invertible; since det = 6 6 = Invertible; since det = = 2.

Fundamentals of Engineering Analysis (650163)

Linear Algebra M1 - FIB. Contents: 5. Matrices, systems of linear equations and determinants 6. Vector space 7. Linear maps 8.

Equality: Two matrices A and B are equal, i.e., A = B if A and B have the same order and the entries of A and B are the same.

Linear Algebra: Lecture notes from Kolman and Hill 9th edition.

1 Matrices and Systems of Linear Equations. a 1n a 2n

1 Matrices and Systems of Linear Equations

ELEMENTARY LINEAR ALGEBRA WITH APPLICATIONS. 1. Linear Equations and Matrices

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS. + + x 1 x 2. x n 8 (4) 3 4 2

Foundations of Matrix Analysis

Representations of Sp(6,R) and SU(3) carried by homogeneous polynomials

0.1 Rational Canonical Forms

Basic Concepts of Group Theory

SPRING 2006 PRELIMINARY EXAMINATION SOLUTIONS

A Brief Outline of Math 355

LINEAR ALGEBRA WITH APPLICATIONS

Determinant lines and determinant line bundles

The Determinant: a Means to Calculate Volume

Problem 1A. Suppose that f is a continuous real function on [0, 1]. Prove that

5.3 Determinants and Cramer s Rule

Math113: Linear Algebra. Beifang Chen

Chapter 2 The Group U(1) and its Representations

Quantum Computing Lecture 2. Review of Linear Algebra

1. General Vector Spaces

Lecture 7. Econ August 18

Finite-dimensional spaces. C n is the space of n-tuples x = (x 1,..., x n ) of complex numbers. It is a Hilbert space with the inner product

Chapter 1: Systems of linear equations and matrices. Section 1.1: Introduction to systems of linear equations

Vector Spaces. Chapter 1

Problem 1A. Find the volume of the solid given by x 2 + z 2 1, y 2 + z 2 1. (Hint: 1. Solution: The volume is 1. Problem 2A.

NOTES FOR LINEAR ALGEBRA 133

Eigenvalues and Eigenvectors A =

On rational approximation of algebraic functions. Julius Borcea. Rikard Bøgvad & Boris Shapiro

Determinants - Uniqueness and Properties

Math 215B: Solutions 1

Elementary linear algebra

SUMS OF ENTIRE FUNCTIONS HAVING ONLY REAL ZEROS

Chapter 5. Linear Algebra. A linear (algebraic) equation in. unknowns, x 1, x 2,..., x n, is. an equation of the form

Upper triangular matrices and Billiard Arrays

NOTES ON LINEAR ALGEBRA. 1. Determinants

Linear Algebra. Min Yan

Chapter 4 - MATRIX ALGEBRA. ... a 2j... a 2n. a i1 a i2... a ij... a in

The Riemann Hypothesis Project summary

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

LECTURE-13 : GENERALIZED CAUCHY S THEOREM

Matrix Lie groups. and their Lie algebras. Mahmood Alaghmandan. A project in fulfillment of the requirement for the Lie algebra course

0 Sets and Induction. Sets

Math Camp Lecture 4: Linear Algebra. Xiao Yu Wang. Aug 2010 MIT. Xiao Yu Wang (MIT) Math Camp /10 1 / 88

Vector Spaces. Vector space, ν, over the field of complex numbers, C, is a set of elements a, b,..., satisfying the following axioms.

LINEAR ALGEBRA BOOT CAMP WEEK 4: THE SPECTRAL THEOREM

Introduction to Groups

Lecture Notes in Linear Algebra

MAT Linear Algebra Collection of sample exams

Remark 1 By definition, an eigenvector must be a nonzero vector, but eigenvalue could be zero.

Hilbert Spaces. Hilbert space is a vector space with some extra structure. We start with formal (axiomatic) definition of a vector space.

Product distance matrix of a tree with matrix weights

Chapter 5 Eigenvalues and Eigenvectors

12 CHAPTER 1. PRELIMINARIES Lemma 1.3 (Cauchy-Schwarz inequality) Let (; ) be an inner product in < n. Then for all x; y 2 < n we have j(x; y)j (x; x)

Week 9-10: Recurrence Relations and Generating Functions

e j = Ad(f i ) 1 2a ij/a ii

EXERCISES IN MODULAR FORMS I (MATH 726) (2) Prove that a lattice L is integral if and only if its Gram matrix has integer coefficients.

Linear Algebra review Powers of a diagonalizable matrix Spectral decomposition

Commutative Banach algebras 79

ELEMENTARY SUBALGEBRAS OF RESTRICTED LIE ALGEBRAS

MATH 240 Spring, Chapter 1: Linear Equations and Matrices

Remark By definition, an eigenvector must be a nonzero vector, but eigenvalue could be zero.

Q N id β. 2. Let I and J be ideals in a commutative ring A. Give a simple description of

Theorem A.1. If A is any nonzero m x n matrix, then A is equivalent to a partitioned matrix of the form. k k n-k. m-k k m-k n-k

Linear Algebra March 16, 2019

LINEAR ALGEBRA REVIEW

Recursion Systems and Recursion Operators for the Soliton Equations Related to Rational Linear Problem with Reductions

II. Determinant Functions

Question: Given an n x n matrix A, how do we find its eigenvalues? Idea: Suppose c is an eigenvalue of A, then what is the determinant of A-cI?

Introduction to Linear Algebra. Tyrone L. Vincent

DETERMINANTS. , x 2 = a 11b 2 a 21 b 1

implies that if we fix a basis v of V and let M and M be the associated invertible symmetric matrices computing, and, then M = (L L)M and the

Lecture Summaries for Linear Algebra M51A

October 25, 2013 INNER PRODUCT SPACES

and let s calculate the image of some vectors under the transformation T.

ELEMENTARY LINEAR ALGEBRA

Linear Algebra Primer

A matrix is a rectangular array of. objects arranged in rows and columns. The objects are called the entries. is called the size of the matrix, and

2 Two-Point Boundary Value Problems

Continued fractions for complex numbers and values of binary quadratic forms

Letting be a field, e.g., of the real numbers, the complex numbers, the rational numbers, the rational functions W(s) of a complex variable s, etc.

A matrix A is invertible i det(a) 6= 0.

Vector Spaces ปร ภ ม เวกเตอร

HASSE-MINKOWSKI THEOREM

ELEMENTARY LINEAR ALGEBRA

A matrix over a field F is a rectangular array of elements from F. The symbol

REPRESENTATION THEORY OF S n

Transcription:

AN INVERSE PROBLEM FOR A CLASS OF CANONICAL SYSTEMS AND ITS APPLICATIONS TO SELF-RECIPROCAL POLYNOMIALS MASATOSHI SUZUKI arxiv:13080228v3 [mathfa] 23 Sep 2016 Abstract A canonical system is a kind of first-order system of ordinary differential equations on an interval of the real line parametrized by complex numbers It is known that any solution of a canonical system generates an entire function of the Hermite-Biehler class In this paper, we deal with the inverse problem to recover a canonical system from a given entire function of the Hermite-Biehler class satisfying appropriate conditions This inverse problem was solved by de Branges in 1960s However his results are often not enough to investigate a Hamiltonian of recovered canonical system In this paper, we present an explicit way to recover a Hamiltonian from a given exponential polynomial belonging to the Hermite-Biehler class After that, we apply it to study distributions of roots of self-reciprocal polynomials 1 Introduction Let H(a) be a 2 2 symmetric matrix-valued function defined almost everywhere on a finite interval I = [a 1,a 0 ) (0 < a 1 < a 0 < ) with respect to the Lebesgue measure da We refer to a first-order system of differential equations a d [ A(a,z) 0 1 A(a,z) A(a,z) 1 = z H(a), lim = (11) da B(a, z) 1 0 B(a, z) aրa 0 B(a, z) 0] on I parametrized by z C as a quasi-canonical system (on I) A column vector-valued function (A(,z),B(,z)) : I C 2 1 is called a solution if it consists of absolutely continuous functions and satisfies (11) almost everywhere on I for every fixed z C A quasi-canonical system (11) is called a canonical system if (H1) H(a) is a real positive semidefinite symmetric matrix for almost every a I, (H2) H(a) 0 on any subset of I with positive Lebesgue measure, (H3) H(a) is locally integrable on I with respect to da/a, and H(a) is called its Hamiltonian Even if a quasi-canonical system (11) is not a canonical system, a matrix valued function H(a) is often called a Hamiltonian in this paper by abuse of language if there is no confusion A number of different second-order differential equations, like Schrödinger equations and Sturm Liouville equations of appropriate form, and systems of first-order differential equations like Dirac type systems of appropriate form are reduced to a canonical system Fundamental results on the spectral theory of canonical systems was established in works of Gohberg Kreĭn [5], de Branges [3] and many other authors (see the survey articles Winkler [27], Woracek [28] and references there in for historical details on canonical systems) Note that the variable a in (11) is the variable on the multiplicative group R >0 By the change of variable a = e x, (11) is transformed into the equation for the variable x on the additive group R, and the right endpoint of a for the initial value is transformed into the left endpoint of x for the initial value The transformed equation is the one treated by de Branges, Kreĭn, Kac and others (See the final paragraph of the introduction for the reason why we use a multiplicative variable) The subject of the present paper is an inverse spectral problem for canonical systems In order to state it, we review the theory of the Hermite Biehler class 2000 Mathematics Subject Classification 30C15, 34A55, 34L40 1

2 M SUZUKI We use the notation F (z) = F( z) for functions of the complex variable z and denote by C + the (open) upper half-plane {z = x +iy C y > 0} An entire function E(z) satisfying E (z) < E(z) for every z C + (12) and having no real zeros is said to be a function of the Hermite Biehler class, or the class HB for short (This definition of the class HB is equivalent to the definition of Levin [11, 1 of Chap VII] if we replace the word the upper half-plane by the lower half-plane, because (12) implies that E(z) has no zeros in C + We adopt the above definition for the convenience to use the theory of canonical systems via the theory of de Branges spaces) Suppose that (11) is a canonical system endowed with a solution (A(a,z),B(a,z)) Then E(a,z) := A(a,z) ib(a,z) is a function of the class HB for every fixed regular point a I Inparticular, lim aրa0 E(a,z) = 1 and E(a 1,z) is a function of theclass HB Therefore, an inverse problem for canonical systems is to recover their Hamiltonians from given entire functions of the class HB satisfying appropriate conditions Usually such inverse problem is difficult to solve in general, because it is an inverse spectral problem ([8, Section 2], [17, Section 7]) However it is already solved by the theory of de Branges in 1960s ([3]) For example, for fixed I and an entire function E(z) of exponential type belonging to the class HB and satisfying (1+x2 ) 1 E(x) 2 dx < and E(0) = 1, there exists a canonical system (that is, there exists a Hamiltonian H(a) on I) such that A(a 1,z) = 1 2 (E(z)+E (z)), B(a 1,z) = i 2 (E(z) E (z)) for its solution (A(a,z),B(a,z)) (see [17, Theorem 73] with [4, Lemma 33]) Moreover, such canonical system is uniquely determined by E(z) and I under appropriate normalizations More general situation is treated in Kac [7] (see also [27] and [28]) As above, de Bragnes s theory ensures the existence of canonical systems or Hamiltonians for given functions of the class HB, but it does not provide explicit or useful expressions of Hamiltonians In fact, an explicit form of H(a) is not known except for a few examples of E(z) as in [3, Chapter 3], [4, Section 8] and some additional examples constructed from such known examples using transformation rules for Hamiltonians and Weyl functions ([26]) In this paper, we deal with the above inverse problem for a special class of exponential polynomials together with the problem of explicit constructions of H(a), and apply the results to the studying of the distribution of roots of self-reciprocal polynomials Let R := R\{0}, g Z >0 and q > 1 We denote by C a vector of length 2g +1 of the form C = (C g,c g 1,,C g ) R R 2g 1 R and consider exponential polynomials along with two associated functions E(z) := E q (z;c) := C m q imz (13) A(z) := A q (z;c) := 1 2 (E q(z;c)+e q(z;c)), (14) B(z) := B q (z;c) := i 2 (E q(z;c) Eq (z;c)) (15) AbasicfactisthatanexponentialpolynomialE(z) of(13)belongstotheclasshbifand only if it has no zeros on the closed upper half-plane C + R = {z = x+iy C y 0} (see [11, Chapter VII, Theorem 6], for example) Therefore, there exists a Hamiltonian H(a) of a canonical system corresponding to E(z) if E(z) has no zeros on C + R

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 3 Beyond such results standing on a general theory, we show that there exists a symmetric matrix valued function H(a) of a quasi-canonical system corresponding to an exponential polynomial E(z) if E(z) has no zeros on the real line at least It is constructed explicitly as follows We define two lower triangular matrices E + and E of size 2g +1 by E ± = E ± (C) := C g C (g 1) C g C 0 C 1 C g and define square matrices J n of size 2g +1 by J n = J n (2g+1) := C ±(g 1) C ±(g 2) C g C ±g C ±(g 1) C 0 C (g 1) C g J (n) 0 0 0, J(n) := for 1 n 2g, where J (n) is the antidiagonal matrix of size n with 1 s on the antidiagonal line By using the above matrices, we define the numbers n by n = n (C) := det(e+ (C)+E (C)J n ) det(e + (C) E (C)J n ) 1 1 (16) for 1 n 2g and 0 = 0 (C) := 1 We write n = if det(e + E J n ) = 0 In addition, we define the real-valued locally constant function γ : [1,q g ) R by γ(a) = γ(a;c) := (C) n (C) for q ()/2 a < q n/2 (17) Then, we obtain the following results for the inverse problem associated with the exponential polynomial (13) Theorem 11 Let q > 1 and C R R 2g 1 R Let E(z) be the exponential polynomial defined by (13) Define the square matrix D n (C) := C g C g+1 C g C g 1 C g C g+ C g+n 2 C g C g n+1 C g n+2 C g C g C g 1 C g n+1 C g C g+1 C g+ C g C g n+2 C g C g+n 2 C g C g C g of size 2n Suppose that E(z) has no zeros on the real line and that detd n (C) 0 for every 1 n 2g Then, (1) det(e + ±E J n ) 0 for every 1 n 2g, (2) the pair of functions (A(a,z),B(a,z)) defined in (213) below satisfies a d A(a,z) 0 1 A(a,z) = z H(a) (a [1,q da B(a, z) 1 0 B(a, z) g ), z C) (18)

4 M SUZUKI together with the boundary condition A(1,z) A(z) = B(1, z) B(z), lim aրq g A(a,z) = B(a, z) E(0), (19) 0 where A(z) and B(z) are functions of (14) and (15), respectively, and γ(a;c) 1 0 H(a) = H(a;C) := (110) 0 γ(a; C) We mention another way of constructing (γ(a),a(a,z),b(a,z)) in Section 6 The positive definiteness of the Hamiltonian H(a) in (110) is characterized by the following to-be-expected way Theorem 12 Let q > 1 and C R R 2g 1 R Let E(z) be the exponential polynomial defined by (13) and let H(a) be the matrix valued function defined by (110) (1) Suppose that E(z) belongs to the class HB Then H(a) is well-defined and is positive definite for every 1 a < q g Hence the quasi-canonical system attached to (18) and (19) is a canonical system and (A(a,z),B(a,z))/E(0) is its solution (2) Suppose that H(a) is well-defined and is positive definite for every 1 a < q g Then E(z) belongs to the class HB As mentioned before, E(z) of (13) belongs to the class HB if and only if it has no zeros on C + R On the other hand, the following conditions are equivalent to each other by 0 = 1 and definitions (16), (17) and (110): (i) H(a) is positive definite for every 1 a < q g, (ii) γ(a) > 0 for every 1 a < q g, (iii) 0 < n < for every 1 n 2g Therefore, we obtain the following corollary: Corollary 13 An exponential polynomial E(z) of (13) has no zeros on C + R if and only if 0 < n < for every 1 n 2g The converse of Theorem 11 is the direct problem for quasi-canonical systems (11) with the Hamiltonians of the form (110) It is easier than the inverse problem, because the Hamiltonians have a simple form: H(a) = diag(γ(a) 1,γ(a)), and γ(a) is a (special) locally constant function Theorem 14 Let q > 0, g Z >0, and let γ(a) be a locally constant function on [1,q g ) such that γ(a) = γ n R for q ()/2 a < q n/2 (1 n 2g) Then the quasicanonical system (11) with H(a) = diag(γ(a) 1,γ(a)) on [1,q g ) has the unique solution (A(a, z), B(a, z)) whose components have the forms [ (q m) iz q m) ] iz A(a,z) = α m (n) +(, a a +n [ (q ) m iz ) ] (111) q m iz ib(a,z) = β m (n) ( a a +n for q ()/2 a < q n/2 and 1 n 2g, where α ± m(n) and β m(n) ± are real constants depending only on the set {γ n } of values of γ(a) Therefore, for q ()/2 a < q n/2, E(a,z) := A(a,z) ib(a,z) is the exponential polynomial [ ( ) q m iz ) ] q m iz E(a,z) = (α m (n)+β m (n)) +(α m (n) β m (n))( a a +n Moreover, E(a,0) = 1 and E(a,z) has no real zeros for any fixed 1 a q g In particular, each Hamiltonian of the form (110) with (17) yields an exponential polynomial E(1,z) = A(1,z) ib(1,z) having no real zeros

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 5 Theorem 14 does not guarantee that E(1,z) has the form (13) In fact, H(a) on [0,q) with γ(a) = 1 for 1 a < q 1/2 and γ(a) = 1 for q 1/2 a < q yields the constant function E(1,z) = 1, and H(a) on [0,q) with γ(a) = 1 for 1 a < q yields the function E(1,z) = q iz In these examples, E(1,z) does not have the form (13) A sufficient condition for which E(1,z) has the form (13) is the following Theorem 15 Under the notation of Theorem 14, E(1, z) is an exponential polynomial of the form (13) if γ n > 0 for every 1 n 2g and γ 1 1 If the assumption of Theorem 15 is satisfied, it is proved that E(1,z) belongs to the class HB in a way similar to the proof of Theorem 12 (2) in Section 52 On the other hand, γ n > 0 for every 1 n 2g and γ 1 1 if we start from the exponential polynomial of the form (13) by Theorem 12 and (58) below Hence, as a consequence of the above theorems, the exponential polynomials (13) belonging to the class HB are characterized in terms of positive definiteness of Hamiltonians Now we turn to an application of Corollary 13 A nonzero polynomial P(x) = c 0 x n +c 1 x + +c x+c n with real coefficients is called a self-reciprocal polynomial of degree n if c 0 0 and it satisfies the self-reciprocal condition P(x) = x n P(1/x), or equivalently, c 0 0 and c k = c n k for every 0 k n The roots of a selfreciprocal polynomial either lie on the unit circle T = {z C z = 1} or are distributed symmetrically with respect to T Therefore, a basic problem is to finda nice condition on coefficients of a self-reciprocal polynomial for which all its roots lie on T There are quite many results for this problem in literatures For example, see books of Marden[14], Milovanović Mitrinović Rassias [15], Takagi [24, Section 10] and the survey paper of Milovanović Rassias [16] for several systematic treatments on roots of polynomials As an application of Corollary 13, we study roots self-reciprocal polynomials of even degree The restriction on the degree is not essential, because if P(x) is a self-reciprocal polynomial of odd degree, then there exists a self-reciprocal polynomial P(x) of even degree and an integer r 1 such that P(x) = (x+1) r P(x) In contrast, the realness of coefficients is essential We denote by P g (x) a self-reciprocal polynomial of degree 2g of the form g 1 P g (x) = c k (x (2g k) +x k )+c g x g, c 0 0 (112) k=0 and identify the polynomial with the vector c = (c 0,c 1,,c g ) R R g consisting of its coefficients For a vector c R R g and a real number q > 1, we define the numbers δ n (c) (1 n 2g) by δ n (c) := det(e+ (C)+E (C)J n ) det(e + (C) E (C)J n ) where C is the vector defined by C m T m := c g m (1 logq m )T m + m=0 { 1 if n is even, glogq if n is odd, (113) c g m (1+logq m )T m (114) Theorem 16 Let g Z >0, q > 1 Let c R R g be coefficients of a self-reciprocal polynomial P g (x) of the form (112) Define C R R 2g 1 R by (114) Then, for every 1 n 2g, it is independent of q whether det(e + ±E J n ) is zero, and the numbers δ n (c) of (113) are independent of q if det(e + ±E J n ) is not zero Moreover, a necessary and sufficient condition for all roots of P g (x) to be simple roots on T is that 0 < δ n (c) < for every 1 n 2g Remark As a function of indeterminate elements (c 0,,c g ), δ n (c) is a rational function of (c 0,,c g ) over Q m=1

6 M SUZUKI Thecriterion of Theorem 16 itself may not benew, becauseitseems that thequantity δ n (c)intheorem16probablyessentially coincidewithquantities inaclassical theoryon the studying of the roots of polynomials in Section 75 The author has not yet thought rigorously whether δ n (c) and the classical quantities indeed have the same meaning However, at least, the theory of the presented paper provides a new bridge between the studying on the roots of polynomials and the theory of canonical systems The author believe that this new connection deserve to be recorded in a literature In order to deal with the case that all roots of P g (x) lie on T but P g (x) may have a multiple zero, we modify the above definition of δ n (c) as follow For a vector c R R g and real numbers q > 1, ω > 0, we define the numbers δ n (c;q ω ) (1 n 2g) by δ n (c;q ω ) = det(e+ (C)+E (C)J n ) 1 if n is even, det(e + (C) E (C)J n ) q gω q gω (115) q gω +q gω if n is odd, where C is the vector defined by C m T m := c g m q mω T m + m=0 c g m q mω T m (116) Theorem 17 Let g Z >0, q > 1 Let c R R g be coefficients of a self-reciprocal polynomial P g (x) of the form (112) Then a necessary and sufficient condition for all roots of P g (x) to lie on T is that 0 < δ n (c;q ω ) < for every 1 n 2g and ω > 0 m=1 Furthermore, the quantities δ n (c) and δ n (c;q ω ) are related as follows: Theorem 18 Let δ n (c) and δ n (c;q ω ) be as above Then lim δ n(c;q ω ) = δ n (c) q ω ց1 as a rational function of c = (c 0,,c g ) over Q Suppose that all roots of a self-reciprocal polynomial (112) lie on T and are simple Then we have where implied constants depend only on c δ n (c;q ω ) = δ n (c)+o(logq ω ) as q ω ց 1, Finally, we comment on the reason why we use a multiplicative variable on R >0 and the right endpoint for the initial value in (11) It comes from the author s personal motivation for the work of this paper: it is the inverse problem for entire functions of the class HB obtained by Mellin transforms of functions on [1, ) This problem was stimulated by Burnol [1], in particular Sections 5 8, and was partially treated in [23] under the number theoretic setting In [1], Burnol often use a multiplicative variable on R >0 and the left endpoint (= + ) corresponds to the initial value In order to apply the method of [1] to the above inverse problem, the author noted on the formulas and X 1 1 f(x)x izdx x = lim X X 1 f(x)x izdx x f(x)x izdx logx/logq x = lim 1 f(q n )q inz q 1+ logq Here exponential polynomials appear in the left hand side of the second formula If we fix X > 1, q > 1 and put 2g = logx/logq, we obtain an exponential polynomial of degree 2g Recall that the results of the paper corresponds the exponential polynomial to the canonical system on [1,q g ) Then, if we imagine the limiting situation q 1 and n=0

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 7 X +, it is expected (but not rigorous) that a canonical system corresponds to the Mellin transform 1 f(x)x iz dx x should be a system on [1, ) To realize the above heuristic discussion to the original motivation, the right endpoint may be more useful and convenient than the left endpoint for the initial value, although the left endpoint is useful for the initial value in the usual theory of canonical systems Anyway, the author believe that the above number theoretic aspect of de Branges theory is an important one The paper is organized as follows In Section 2, we describe the outline of the proof of Theorem 11 In Section 3, we prepare several lemmas and notations in order to prove statements of Section 2 and Theorem 11 in Section 4 In Section 5, we prove Theorems 12, 14 and 15 In Section 6, we mention a way of constructing (γ(a),a(a,z),b(a,z)) which is different from the way of Section 2 and 4 In Section 7, we prove Theorems 16, 17 and 18 and compare these with classical results on the roots of self-reciprocal polynomials Acknowledgments The author thanks Shigeki Akiyama for suggesting the book of Takagi ([24, Section 10]) The author also thanks the referee for a number of helpful suggestions for improvement in the article and for careful reading In particular, Theorems 14 and 15 were added by the suggestion of the referee This work was supported by KAKENHI (Grant-in-Aid for Young Scientists (B)) No 21740004 and No 25800007 2 Outline of the proof of Theorem 11 21 Hilbert spaces and operators Let q > 1 and let L 2 (T q ) be the completion of the space of (2π/logq)-periodic continuous functions on R with respect to the L 2 -norm f 2 L 2 (T := f,f q) L 2 (T q),where f,g L 2 (T q) := logq 2π I q f(z)g(z)dz andi q = [0,2π/(logq)) Every f L 2 (T q ) has the Fourier expansion f(z) = k Z u(k)qikz with {u(k)} k Z l 2 (Z) and f 2 L 2 (T q) = k Z u(k) 2, where l 2 (Z) is the Hilbert space of sequences {u(k) C k Z} satisfying k Z u(k) 2 < ([18, Chapter 4]) For 0 < a q Z/2 = {q n/2 n Z}, we define the vector space V a := { φ(z) = a iz f(z)+a iz g(z) f, g L 2 (T q ) } of functions of z R As a vector space, V a is isomorphic to the direct sum L 2 (T q ) L 2 (T q ), since a iz f(z) + a iz g(z) = 0 if and only if (f,g) = (0,0) In fact, if one of f and g is zero and a iz f(z)+a iz g(z) = 0, the another one should be zero On the other hand, if f 0, g 0 and a iz f(z)+a iz g(z) = 0, we have a 2iz = f(z)/g(z), and hence a q Z/2 The maps p 1 : (a iz f(z)+a iz g(z)) a iz f(z) and p 2 : (a iz f(z)+a iz g(z)) a iz g(z) are projections from V a to the firstand the second components of the direct sum, respectively In addition, we define the inner product in V a by φ 1,φ 2 = f 1,g 1 L 2 (T q) + g 1,g 2 L 2 (T q), where φ j (z) = a iz f j (z) + a iz g j (z) (j = 1,2) Then, V a forms a Hilbert space with respect to the inner product and is isomorphic to the (orthogonal) direct sum L 2 (T q ) L 2 (T q ) of Hilbert spaces ([9, Chapter 5]) We put X(k) := (q k /a) iz, Y(l) := X(l) 1 = (q l /a) iz (21) for k,l Z and a > 0 We regard X(k) and Y(l) as functions of z, functions of (a,z) or symbols depending on the situation For a fixed 0 < a q Z/2, the countable set consisting of all X(k) and Y(l) is linearly independent over C as a set of functions of z, since the linear dependence of {X(k), Y(l)} k,l Z implies the existence of a nontrivial

8 M SUZUKI pair of f,g L 2 (T q ) satisfying a iz f(z)+a iz g(z) = 0 Using these vectors, we have { V a = φ = u(k)x(k)+ } v(l)y(l) {u(k)} k Z, {v(l)} l Z l 2 (Z) k Z l Z and L 2 (T q ) L 2 (T q ), { } p 1 V a = u(k)x(k) {u(k)} k Z l 2 (Z) L 2 (T q ), k Z { } p 2 V a = v(l)y(l) {v(l)} l Z l 2 (Z) L 2 (T q ) l Z φ,φ = u(k)u (k)+ v(l)v (l) (φ,φ V a ), k Z l Z φ 2 = φ,φ = u(k) 2 + v(l) 2 (φ V a ), (22) k Z l Z where φ = k Z u(k)x(k)+ l Z v(l)y(l) and φ = k Z u (k)x(k)+ l Z v (l)y(l) On the other hand, we have φ 2 = logq p 1 φ(z)p 1 φ(z)dz + logq p 2 φ(z)p 2 φ(z)dz (23) 2π I q 2π I q for φ V a, since logq (a ±iz f 1 (z))(a 2π ±iz f 2 (z))dz = u 1 (k)u 2 (k) = f 1,f 2 L 2 (T q) I q k Z for f j (z) = k Z u j(k)q ikz L 2 (T q ) (j = 1,2) Note that, for φ V a, p 1 φ and p 2 φ are not periodic functions of z, but the integrals I p jφ(z)p j φ (z)dz (j = 1,2, φ,φ V a ) are independent of the intervals I = [α,α+2π/logq] (α R) We write φ V a as φ(z) (resp φ(a,z)) to emphasize that φ is a function of z (resp (a,z)) If we regard X(k) and Y(l) as symbols, V a is abstract Hilbert space endowed with the norm defined by (22) and is isomorphic to l 2 (Z) l 2 (Z) For each nonnegative integer n, we define the closed subspace V a,n of V a by { } V a,n = φ n = u n (k)x(k) + v n (l)y(l) {u n(k)} k=0, {v n(l)} l= l2 (Z) k=0 k= and denote by P n the projection operator V a V a,n Then, for the conjugate operator P n := JP nj : V a V a of P n by the involution J : φ(z) φ( z), we have P n φ = JP njφ = u(k)x(k) + v(l)y(l) (φ V a ), k= because J acts on V a as exchanging of coefficients between X(k) and Y(l): Jφ = v(k)x(k) + u(l)y(l) (φ V a ) k= l= Therefore, P n maps V a,n into V a,n for every nonnegative integer n In addition, JP n also maps V a,n into V a,n for every nonnegative n, because l=0 l=0 JP n φ n = v n (l)x(l)+ u n (k)y(k) (24) for φ n V a,n Note that P 0 Va,0 = JP 0 Va,0, since P 0 J V a,0 = 0 by definition k=0

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 9 For an exponential polynomial E(z) of (13), we have E (z) = C m q imz = E q (z;c ) for C := (C g,c (g 1),,C g ) R R 2g 1 R by a simple calculation By using E(z) and E (z), we define two multiplication operators E : φ(z) E(z)φ(z), E : φ(z) E (z)φ(z) (25) These operators map V a into V a, because E and E are expressed as E = C m T m, E = C m T m by using shift operators T m : V a V a (m Z) defined by T m v = u(k)x(k +m)+ v(l)y(l m) k= l= Both E and E are bounded on V a, since E g C m T m M and E g C m T m M for M = max{c m g m g} Suppose that E(z) has no zeros on the real line Then the operator E is invertible on V a (Lemma 41(1)) Thus the operator is well-defined We have Θ := E 1 E (26) (Θφ)(z) = Θ(z)φ(z) for Θ(z) := E (z) E(z), φ V a (27) and Θ(V a,n ) V a,n for each nonnegative integer n (Lemma 41(2)) 22 Quasi-canonical systems associated with exponential polynomials Under the above settings, a quasi-canonical system associated with an exponential polynomial E(z) of (13) is constructed starting from solutions of the set of linear equations (I±ΘJP n )φ ± n = ΘX(0) (φ ± n V a,n, 0 n 2g), (28) where I is the identity operator The set of 4g+2 equations (28) is a discrete analogue of the (right) Mellin transform of differential equations (117a) and (117b) in [1, Section 6] (see also [23, Section 4]) Under the assumption that both I±ΘJP n are invertible on V a,n for every 0 n 2g, that is, (I±ΘJP n ) 1 exists as a bounded operator on V a,n, we define A n (a,z) := 1 2 (I+J)Eφ+ n (a,z), B n (a,z) := 1 2i (I J)Eφ n (a,z) (29) by using unique solutions of (28) Then they are entire functions of z and are extended to functions of a on (0, ) (by formula (45)) Here we define functions A (a,z) and B (a,z) of (a,z) [1,q g ) C by A (a,z) : = A n (a,z) B (a,z) : = B n(a,z) for q ()/2 a < q n/2 (210) The function A (a,z) is discontinuous at a [1,q g ) q Z/2 in general, because A n(q n/2,z) = A n+1(q n/2,z)

10 M SUZUKI may not be hold It is similar about B (a,z) However, we will see that α n = α n (C) := A (q()/2,z) A n(q ()/2,z), β n = β n (C) := B (q()/2,z) B n(q ()/2,z) (211) are independent of z for every 1 n 2g (Proposition 48) Therefore, we obtain functions A(a,z) and B(a,z) of (a,z) [1,q g ) C which are continuous for a and entire for z by the modification for 1 n 2g and A n (a,z) : = α 1 α n A n(a,z), B n (a,z) : = β 1 β n B n(a,z) A(a,z) : = A n (a,z) B(a,z) : = B n (a,z) (212) for q ()/2 a < q n/2 (213) Moreover, (A(a, z), B(a, z)) satisfies the system (18) endowed with the boundary condition (19) for the locally constant function γ(a) defined by and γ n = γ n (C) := α 1(C) α n (C) β 1 (C) β n (C) (214) γ(a) := γ(a;c) := γ n if q ()/2 a < q n/2 (215) (Proposition 45 and Theorem 418) This γ(a) is equal to the function of (17) (Proposition 48) Therefore, (A(a, z), B(a, z))/e(0) is a solution of a quasi-canonical system on [1,q g ) if E(0) 0 and α n,β n 0 for every 1 n 2g As a summary of the above argument, we obtain Theorem 11 See Section 4 for details In order to prove Theorems 12, 14 and 15, we use the theory of de Branges spaces which is a kind of reproducing kernel Hilbert spaces consisting of entire functions Roughly, the positivity of H(a) corresponds to the positivity of the reproducing kernels of de Branges spaces See Section 5 for details 3 Preliminaries 31 Identities for matrices The following basic facts of linear algebra are used often in later sections Lemma 31 Let A and B be square matrices of the same size Then, A B det = det(a+b)det(a B) B A Lemma 32 Let A, B, C, D be square matrices of the same size If deta 0, detd 0 and det(d CA 1 B) 0, 1 [ A B (A BD = 1 C) 1 A 1 B(D CA 1 B) 1 ] C D D 1 C(A BD 1 C) 1 (D CA 1 B) 1 Lemma 33 (Laplace s expansion formula) Let A be a square matrix of size n Let i = (i 1,i 2,,i k ) (resp j = (j 1,j 2,,j k )) be a list of indices of k rows (resp columns), where 1 k < n and 0 i 1 < i 2 < i k < n (resp 1 j 1 < j 2 < < j k < n) Denote by A(i,j) the submatrix of A obtained by keeping the entries in the intersection of any row and column that are in the lists Denote by A c (i,j) the submatrix of A obtained

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 11 by removing the entries in the rows and columns that are in the lists By Laplace s formula for determinants, we get deta = j ( 1) i + j deta(i,j)deta c (i,j), where i = i 1 +i 2 + +i k, j = j 1 +j 2 + +j k, and the summation is taken over all k-tuples j = (j 1,j 2,,j k ) for which 1 j 1 < j 2 < < j k < n Proof See [6, Chap IV], for example 32 Definition of special matrices, I We define several special matrices for a convenience of later arguments The matrix E 0 ± is the square matrix of size 8g defined by E 0 ± = E± 0 (C) := e ± 0 (C) 0 0 e± 0 (C), where e ± 0 (C) are lower triangular matrices of size 4g defined by C g C (g 1) C g e ± 0 = C e± 0 (C) := ±(g 1) C (g 1) C g C ±g C (g 1) C g 0 C±(g 1) C±g C ±(g 1) C (g 1) C g 0 0 C ±g C ±(g 1) C (g 1) C g Replacing the column at the left edge of e 0 by the zero column vector, 0 0 0 C g C (g 1) Cg e 1 = e 1 (C) := C (g 1) C(g 1) C g C g C(g 1) C g 0 C (g 1) 0 C g C (g 1) C(g 1) C g 0 0 0 C g C (g 1) C (g 1) C g We define and E nj = E nj(c) := E 0 J = E 0 J(C) := 0, E 0,1 J = E 0,1J(C) := 0 0 e 1,n (C) e 2,n (C) 0, E n,1 J = E n,1j(c) := 0 e 3,n (C) e 2,n (C) 0

12 M SUZUKI by setting e 1,n = e 1,n (C) := e 2,n = e 2,n (C) := C g C (g 1) C g 0 C (g 1) C (g 1) 0 C g C (g 1) C g 0 0 0 0 0 C g C (g 1) C g C (g 1) C (g 1) 0 C g C (g 1) C g = = 2g n n 2g n 4g n 2g n 2g +n 2g +n, 2g n and e 3,n = e 3,n (C) := 0 0 0 C g 0 0 C (g 1) 0 C g 0 C (g 1) C (g 1) 0 0 C g 0 C (g 1) C g 0 0 0 0 0 = 2g n n 2g 2g +n 2g n for 1 n 2g, where the right-hand sides mean the size of each block of matrices in middle terms In addition, we define square matrices e 0 of size 4g by e 0 = e 0 (C) := C g C g 1 C (g 1) C g C g 1 C g C g C (g 1) C g C g 1 C (g 1) C g C g 1 C g C (g 1) C (g 1) C g C g

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 13 so that J (8g) E + 0 = 0 e 0 e 0 0 Replacing n columns from the left edge of e 0 by zero column vectors, C g C g 1 C (g 1) C g C g 1 C g C g C (g 1) e 0,n = e 0,n (C) : = 0 Cg 1 C (g 1) C g C g C (g 1) C g = n 4g n 4g We denote by I (m) the identity matrix of size m and by J n (m) of size m: J n (m) = We also use the notation J (n) 0 0 0, J (n) = 1 the following square matrix 1 χ n = t[ 1 0 0 ] = the unit column vector of length n For a square matrix M of size N, we denote by [M] տn (resp [M] րn ) the square matrix of size n( N) in the top left (resp the top right) corner, and denote by [M] ցn (resp [M] ւn ) the square matrix of size n( N) in the lower right (resp the lower left) corner: M = [ [M]տn = [M] ցn n [M] ւn n [M] րn 33 Definition of special matrices, II We define the square matrix P k (m k ) of size 2k+2 endowed with the parameter m k and the (2k+2) (2k+4) matrix Q k for every nonnegative integer k as follows For k = 0,1, P 1 (m 1 ) := P 0 := [ 1 0 0 1 1 0 0 0 0 1 0 0 0 0 1 0 0 1 0 m 1 ], Q 0 :=, Q 1 := [ 1 1 0 0 0 0 1 1 ], ] 1 0 1 0 0 0 0 1 0 0 0 0 0 0 0 1 0 1 0 0 0 0 0 0 For k 2, we define P k (m k ) and Q k blockwisely as follows V + k 0 W + P k (m k ) := 0 V k 0 k, Q k := 0 W k 0I k m k 0I k 0 k,k+2 0 k,k+2,

14 M SUZUKI where 0I k := 0 k,1 I k, mk 0I k = 0 k,1 m k I k, 0k,l is the k l zero matrix, I k is the identity matrix of size k, 1 0 V + k := 1 1 ( ) k+3 (k+1) if k is odd, 2 1 V k := := := 1 0 1 1 1 1 1 0 1 1 1 0 1 1 0 1 1 1 1 ( k +2 2 ( k +1 ) (k +1) if k is even, 2 ) (k +1) if k is odd, ( ) k+2 (k+1) if k is even, 2 and matrices W ± k are defined by adding column vectors t (1 0 0) to the right-side end of matrices V ± k : 1 0 1 W + k := 1 1 ( ) k+3 (k+2) if k is odd, 2 1 1 0 1 1 1 ( ) := k +2 (k +2) if k is even, 2 1 1 W k := := 1 0 1 1 1 1 0 1 1 0 1 1 1 1 1 Lemma 34 For k 1, we have where ε 2j+1 = detp k (m k ) = { { ( k +1 ( k+2 ε 2j+1 2 j m j+1 2j+1 if k = 2j +1, ε 2j 2 j m j 2j if k = 2j, 2 2 ) (k +2) if k is odd, ) (k+2) if k is even { +1 j 2,3 mod4 1 j 0,1 mod4, ε +1 j 0,1 mod4 2j = 1 j 2,3 mod4 In particular, P k (m k ) is invertible if and only if m k 0

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 15 Proof This is trivial for k = 1 Suppose that k = 2j + 1 3 and write P k (m k ) as (v 1 v 2k+2 ) by its column vectors v l At first, we make the identity matrix I k+2 at the left-upper corner by exchanging the columns v (k+5)/2,,v k+1 and v k+2,,v (3k+3)/2 so that detp k (m k ) = det(v 1 v (k+3)/2 v k+2 v (3k+3)/2 v (k+5)/2 v k+1 v (3k+5)/2 v 2k+2 ) Then, by eliminating every 1 and m k under I k+2 of the left-upper corner, we have Ik+2 detp k (m k ) = det 0 k,k+2 Z k Here Z k is the k k matrix Z k = Zk,1 Z k,2 0 j+1,j Z k,3 for which Z k,1 is the j j antidiagonal matrix with 1 on the antidiagonal line and m k j (j +1) Zk,2 m k = Z k,3 m k 2m k (j +1) (j +1) 2mk The above formula of detp k (m k ) implies the desired result The case of even k is proved in a way similar to the case of odd k Lemma 35 We have P k (m k ) 1 Q k = Mk,1 M k,2 M k,3 M k,4 (2k +2) (2k +4) (31) for every 0 k 2g, where M k,1, M k,2, M k,3, M k,4 are (k+1) (k+2) matrices defined by 1 0 1 0 0 0 M0,1 M 0,2 1 1 0 0 M1,1 M =, 1,2 = 0 1 0 0 0 0 M 0,3 M 0,4 0 0 1 1 M 1,3 M 1,4 0 0 0 1 0 1 0 1/m 1 0 0 0 0 and 2 2 1 1 M k,1 = 1 1 1 2 2 if k 3 is odd, 1 1 0 1 1 0 2 2 1 1 = 1 2 1 1 if k 2 is even, 1 1 0 1 1 0

16 M SUZUKI 2 2 1 1 M k,4 = 1 1 1 2 0 if k 3 is odd, 1 1 0 1 1 0 2 2 1 1 = 1 2 1 1 if k 2 is even, 1 1 0 1 1 0 0 0 1 1 M k,2 = m k 1 1 2 0 if k 3 is odd, 1 1 0 1 1 0 0 0 1 1 = m k 2 1 1 if k 2 is even, 1 1 0 1 1 0 0 0 1 1 M k,3 = 1 1 1 2m k 2 if k 3 is odd, 1 1 0 1 1 0 0 0 1 1 = 1 2m k 1 1 if k 2 is even 1 1 0 1 1 0 Proof According to definitions of P k (m k ), Q k and M k,l (1 l 4), the identity Mk,1 M P k (m k ) k,2 = Q M k,3 M k k,4 is checked by an elementary way

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 17 4 Proof of Theorem 11 In this section, we complete the proof of Theorem 11 by showing each statement in Section 2 We fix g Z >0, q > 1 and C R R 2g 1 R throughout this section Lemma 41 Let E be the multiplication operator defined by (25) for E(z) of (13) Suppose that E(z) has no zeros on the real line Then (1) E is invertible on V a Moreover, for the operator Θ of (26), we have (2) Θ(V a,n ) V a,n for each nonnegative integer n Proof (1) The multiplication by 1/E(z) maps a sum of X(k) (resp Y(l)) to another sum of X(k) (resp Y(l)) as well as multiplication by E(z), since 1 E(z) = q igz 2g m=0 C m gq = C m q imz imz for some sequence { C m } of real numbers Therefore, it is sufficient to prove that E is invertible on p 1 V a and p 2 V a We have 1/E(z) L (T q ) by the assumption Therefore, multiplication by1/e(z) definesaboundedoperatore 1 onp i V a withthenorm E 1 = 1/E L (T q) for i = 1,2 (2) We have Θ(z) = 1 on the real line by definition (27) Thus, multiplication by Θ(z) defines a bounded operator on V a with the norm Θ = 1 We can write 2g m=0 Θ(z) = C g mq imz 2g m=0 C m gq = R m q imz (R m R) imz m=0 Therefore, Θ(z)φ n (z) = R m u n (k)x(k +m)+ R m v n (l)y(l m) = m=0 k=0 ( m=0 k=m This shows Θ(V a,n ) V a,n Let us consider the equations ) R m u n (k m) X(m)+ m=g m=0l= 0 m= (+m l= ) R m v n (l m) Y(m) (E±E JP n )φ ± n = E X(0) (φ ± n V a,n, 1 n 2g) (41) instead of (28) They are equivalent to (41) if E is invertible Lemma 42 Let 0 < a q Z/2 Let D n (C) be matrices define in Theorem 11 Suppose that E(z) of (13) has no zeros on the real line and that detd n (C) 0 for every 1 n 2g Then ΘJP n defines a compact operator on V a,n for each 0 n 2g and the resolvent set of ΘJP n Va,n contains both ±1 In particular, I ±ΘJP n are invertible on V a,n and (41) has unique solutions in V a,n for each 0 n 2g Remark If a q Z/2 (0, ), I±ΘJP n may not be invertible Proof It is trivial for n = 0, since P 0 Va,0 = 0 On V a,n for n 1, the Hilbert-Schmidt norm of ΘJP n is finite In fact, ΘJP n Va,n 2 2 = ΘJP n X(k) 2 + ΘJP n Y(l) 2 k=0 JP n X(k) 2 + k=0 l= l= JP n Y(l) 2 k=0 1 < 1+ l=0

18 M SUZUKI by(22)and(24) Therefore, ΘJP n Va,n isahilbert-schmidtoperatoronv a,n, andhence it is compact Thus, either λ C\{0} is an eigenvalue of ΘJP n Va,n or an element of the resolvent set of ΘJP n Va,n Assume that ΘJP n φ n = ±φ n Then ΘJP n φ n = φ n Because Θ and p i (i = 1,2) commute (see the proof of Lemma 41(2)), ΘJP n φ n 2 = logq 2π by (23) While, φ n 2 = logq 2π = logq 2π k=0 by (23) Thus, it must be In this case, and Θ(z)p 1 JP n φ n (z) 2 dz + logq Θ(z)p 2 JP n φ n (z) 2 dz I q 2π I q p 1 JP n φ n (z) 2 dz + logq p 2 JP n φ n (z) 2 dz I q 2π I q = u(k) 2 + v(l) 2 E JP n φ n = ±Eφ n = ± l=0 p 1 φ n (z) 2 dz + logq I q 2π k=0 k=0 I q p 2 φ n (z) 2 dz = φ n = u(k)x(k) + v(l)y(l) l=0 C m v(k)x(k +m)+ k=0 C m u(k)x(k +m)± l=0 l=0 u(k) 2 + k=0 l= C m u(l)y(l m) C m v(l)y(l m) v(l) 2 Comparing the coefficients of {X(k)} g k g+ and {X(k)} g k g+ in the equality E JP n φ n ±Eφ n = 0, yields where A ± = A ± t[ u n (0) u n () v n (0) v n () ] = 0, (42) ±C g C g ±C g+1 ±C g C g 1 C g ±C g+ ±C g+n 2 ±C g C g n+1 C g n+2 C g ±C g ±C g 1 ±C g n+1 C g C g+1 C g+ ±C g ±C g n+2 C g C g+n 2 ±C g C g Here deta ± 0 by detd n (C) 0 Therefore (42) has no nontrivial solutions, thus φ n = 0 Consequently, both ±1 are not eigenvalue, and hence they belong to the resolvent set ofθjp n Va,n whichmeansthati±θjp n Va,n existsandboundedonv a,n In the remaining part of the section, we assume that E(z) 0 for z R and detd n (C) 0 for 1 n 2g so that both E±E JP n are invertible on V a,n for every 0 n 2g (cf Lemma 41 and Lemma 42)

Let φ ± n = k=0 u± n (k)x(k) + Using coefficient of φ ± n and putting INVERSE PROBLEMS FOR CANONICAL SYSTEMS 19 l= v± n (l)y(l) be solutions of (41) for 0 n 2g v ± n(n) = v ± n(n+1) = v ± n(2g 1) = 0 if 0 n 2g 1, we define the column vectors Φ ± n of length 8g by Φ ± n = t[ u ± n (0) u± n (1) u± n (4g 1) v± n (2g 1) v± n (2g 2) v± n ( 2g)] Substituting φ ± n in (41) and then comparing coefficients of X(k) and Y(l), we obtain and (E + 0 ±E n J)Φ± n = E 0 χ 8g, (Φ ± n V a,n, 0 n 2g) (43) 2 j=0 C g j u ± n(k +j) = 0, for every K 4g and L 2g 1 2 j=0 C g j v ± n(l j) = 0 (44) Lemma 43 Let 0 n 2g We have det(e + 0 ±E nj) 0 if E±E JP n is invertible Proof For fixed n, equations (41) have unique solutions φ ± n, if E ± E JP n are invertible On the other hand, the solutions φ ± n correspond to solutions of the pair of linear equations (43) and (44) Therefore, det(e 0 + ±E nj) 0 On the other hand, Eφ ± n = E X(0) E JP n φ ± n ( = C m (X(m) v ± n (k)x(k +m)+u ± n (k)y(k m))) by (41) Therefore, we can write k=0 g+ Eφ ± n = ) (p ± n (k)x(k)+q± n (k)y(k) k= g (45) for some real numbers p ± n(k) and q n(k) ± Hence Eφ ± n(a,z) are extended to smooth functions of a on (0, ) by the right-hand side of (45) We use the same notation for such extended functions We put p ± n(k) = q n(k) ± = 0 for every g+n k 3g 1 if 0 n 2g 1 and define the column vectors Ψ ± n of length 8g by Ψ ± n = t[ p ± n ( g) p± n ( g +1) p± n (3g 1) q± n (3g 1) q± n (3g 2) q± n ( g)] Then, we have by g+ Eφ ± n = ) (p ± n (k)x(k)+q± n (k)y(k) = k= g k=0 C m u n (k)x(k +m)+ Ψ ± n = E+ 0 Φ± n (46) l= C m v n (l)y(l m) Lemma 44 Let 0 n 2g We have p ± n(k)±q n(k) ± = 0 if g +1 k g + or g k g+, and p ± n(g)±q n(g) ± = C g Therefore, (I±J)Eφ ± n = (p ± n(k)±q n(k))(x(k) ± ±Y(k)) k= g+n

20 M SUZUKI Proof By definition of matrices, the equality (E 0 + ±E nj)φ ± n = E0 χ 8g is written as e + 0 ±e 1,n ±e 2,n e + Φ± n = E 0 χ 8g (47) 0 On the other hand, e + 0 ±e 0 ±e 0 e + 0 Φ± n = p ± n( g)±q n( g) ± p ± n( g+1)±q n( g ± +1) p ± n(3g 1)±q n(3g ± 1) ±(p ± n(3g 1)±q n(3g ± 1)), ±(p ± n(3g 2)±q n(3g ± 2)) ±(p ± n( g)±q ± n( g)) since E + 0 ±J(8g) E + 0 = e + 0 ±e 0 ±e 0 e + 0 by definition of e 0 and p ± n ( g)±q± n ( g) p ± n ( g +1)±q± n ( g +1) (I (8g) ±J (8g) )E 0 + Φ± n = (I (8g) ±J (8g) )Ψ ± n = p ± n (3g 1)±q± n (3g 1) ±(p ± n (3g 1)±q± n (3g 1)) ±(p ± n (3g 2)±q± n (3g 2)) ±(p ± n ( g)±q± n ( g)) by (46) In addition, e + 0 ±e 0 ±e 0 e + 0 Φ± n = e + 0 ±e 0,n ±e 0 e + 0 Φ± n, since v n(n) ± = v n(n ± +1) = v n(2g ± 1) = 0 for 1 n 2g 1 by definition of Φ ± n Therefore, p ± n ( g)±q± n ( g) p ± n ( g +1)±q± n ( g +1) e + 0 ±e 0,n ±e 0 e + Φ± n = p ± n(3g 1)±q n(3g ± 1) ±(p ± n(3g 1)±q n(3g ± 1)) (48) 0 ±(p ± n (3g 2)±q± n (3g 2)) ±(p ± n( g)±q n( g)) ±

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 21 Here we find that 2g rows of both E 0 + e + ±E nj and 0 e 0,n e 0 e + with indices (2g+1,2g+ 0 2,,4g) and n rows of both E 0 + e + ±E nj and 0 e 0,n e 0 e + with indices (8g n+1,8g 0 n+2,,8g) have the same entries Therefore, by comparing 2g rows of (47) and (48) with indices (2g +1,2g +2,,4g), we have p ± n(g)±q ± n(g) = C g and p ± n(k)±q n(k) ± = 0 (g +1 k 3g 1) Similarly, by comparing n rows of (47) and (48) with indices (8g n + 1,8g n + 2,,8g), we have Hence we complete the proof p ± n(k)±q ± n(k) = 0 ( g k g+) We define the column vectors A n and B n of length 8g by A n = A n(c) := (I +J)Ψ + n = (I +J)E 0 + Φ+ n, B n = B n(c) := (I J)Ψ n = (I J)E + 0 Φ n, where I = I (8g) and J = J (8g), and define the row vector F(a,z) of length 8g by F(a,z) := [ X( g) X( g +1) X(g) 0 0 Y(g) Y(g 1) Y( g) ] Then, we obtain by (210) and Lemma 44 Proposition 45 We have a d (49) A n(a,z) = 1 2 F(a,z) A n, B n(a,z) = 1 2i F(a,z) B n (410) da A n (a,z) = zb n for every 0 n 2g (a,z), a d da B n (a,z) = za n (a,z) We provide two lemmas in order to prove Proposition 45 Lemma 46 We have for every 0 n 2g u + n(k) = u n(k) (0 k 4g 1), v n + (k) = v n (k) ( 2g k 2g 1) Proof We have Φ ± n = (E+ 0 ±E nj) 1 E0 χ 8g with E 0 + ±E n J = e + 0 ±e 1,n ±e 2,n e + 0 by (43) and definition of EnJ Applying Lemma 32 to A = D = P := e + 0, B = ±Q := ±e 1,n and C = ±R := ±e 2,n, we have [ Φ ± n = (P QP 1 R) 1 P 1 Q(P RP 1 Q) 1 ] P 1 R(P QP 1 R) 1 (P RP 1 Q) 1 E0 χ 8g This shows Lemma 46, since all 4g entries of E0 χ 8g with indices (4g+1,4g+2,,8g) are zero

22 M SUZUKI Lemma 47 We have p + n (k)+q+ n (k) = p n (k) q n (k) ( g k g +) for every 0 n 2g, where p ± n (k)±q± n (k) = 0 if g k g + or g +1 k g + by Lemma 44 Proof By (41), Eφ ± n = E X(0) E JP n φ ± n = C m X(m) Therefore, (I+J)Eφ + n = + On the other hand, (I J)Eφ n = = C m X(m) C m Y(m) C m X(m) C m Y(m)+ C m X(m) C m Y(m)+ C m v n ± (k)x(k +m) k=0 ( ( ( ( C m C m u ± n (l)y(l m) l=0 ) u + n(k)x(k +m)+c m v n(k)x(k + +m) k=0 C m k=0 ) u + n (l)y(l+m)+c m v n + (l)y(l+m) l=0 C m ( l=0 ) u n(k)x(k +m) C m vn(k)x(k +m) k=0 C m C m ( k=0 ) u n(l)y(l+m) C m vn(l)y(l+m) l=0 l=0 ) u + n(k)x(k +m)+c m v n(k)x(k + +m) k=0 C m k=0 ) u + n(l)y(l+m)+c m v n(l)y(l+m) + by Lemma 46 Comparing the right-hand sides of the above formulas of (I ± J)Eφ ± n with formulas of (I±J)Eφ ± n in Lemma 44, we obtain Lemma 47 Proof of Proposition 45 By definition of X(k) and Y(l), (45), and Lemma 44, (I±J)Eφ ± n (a,z) = (p ± n (k)±q± n (k))((qk /a) iz ±(q k /a) iz ) k= g+n Therefore, the differentiability of A n(a,z) and Bn(a,z) with respect to a is trivial, and a d da (I+J)Eφ+ n(a,z) = iz (p + n(k)+q n(k))((q + k /a) iz (q k /a) iz ), a d 1 da i (I J)Eφ n(a,z) = z k= g+n k= g+n l=0 l=0 (p n(k) q n(k))((q k /a) iz +(q k /a) iz ) Applying Lemma 47 to the right-hand sides, we obtain Proposition 45 by definition (29) Asmentionedbefore, A n (a,z)andb n (a,z)of(29)aresmoothfunctionsofaon(0, ) for every 0 n 2g However A (a,z) and B (a,z) of (210) may be discontinuous at a (0, ) q Z/2 Therefore, as mentioned in Section 2, the next aim is to make modifications so that A (a,z) and B (a,z) become continuous functions of a on [1,q g ) The essential part is the following proposition

INVERSE PROBLEMS FOR CANONICAL SYSTEMS 23 Proposition 48 Let 1 n 2g Suppose that E±E JP is invertible on V a, for every q (n 2)/2 < a < q n/2 and E±E JP n is invertible on V a,n for every q ()/2 < a < q n/2 Define α n and β n by (211) Then, we have det(e + +E J 1 ) if n = 1, α n = det(e + ) det(e + +E J 2 ) det(e + ) det(e + 0 ) det(e + 0 +E 1 J) if n = 2, det(e + +E J n ) det(e 0 + +E n 2 J) det(e + +E J n 2 ) det(e 0 + +E J) if 3 n 2g, det(e + E J 1 ) det(e + if n = 1, ) det(e + E J 2 ) det(e 0 + ) β n = det(e + ) det(e 0 + +E 1 J) if n = 2, det(e + E J n ) det(e 0 + E n 2 J) det(e + E J n 2 ) det(e 0 + E J) if 3 n 2g In particular, α n and β n depend only on C We prove Proposition 48 after preparing several lemmas Lemma 49 For every 0 n 2g, we have (411) (412) det(e + 0 +E nj) = det(e + 0 E nj) (413) Proof Suppose that n 1, since EnJ = 0 by definition We have E 0 + ±E n J = e + 0 ±e 1,n ±e 2,n e + 0 A 0 ±B C D 0 = ±B E A = 2g +n 4g 2n 2g +n 2g +n 4g 2n 2g +n by definition of e 1,n and e 2,n, where [ A = [e + 0 ] տ2g+n, B = [e [e + 2,n ] ւ2g+n, D = 0 ] ց2g n [e + 0 ] տ2g n and the right-hand side means the size of each block of the matrix of the second line Applying Lemma 33 to the (4g 2g n) columns of E 0 + ± E nj with indices (2g + n+1,,4g), and then applying Lemma 33 to the (4g 2g n) rows of the resulting matrix with indices (2g +n+1,,4g), we obtain det(e 0 + A ±B ±E nj) = det(d)det ±B A By Lemma 31, det This equality shows (413) A ±B = det(d)det(a+b)det(a B) ±B A ]

24 M SUZUKI Lemma 410 We have adj(e + 0 ±e 0 J(4g) n )e 0 χ 4g = adj(e + 0 ±e 1 J(4g) n )e 0 χ 4g for every 1 n 2g, where adj(a) = (deta)a 1 is the adjugate matrix of A Proof Let w = t[ C g C g 0 0 ] be the column vector of length 4g Denote by u n,j and v n,j the column vectors of e + 0 ±e 0 J(4g) n e + 0 ±e 0 J(4g) n = [ u n,1 u n,4g ], e + 0 ±e 1 J(4g) n = [ v n,1 v n,4g ] By Cramer s formula, it is sufficient to prove that det[u n,1 u n,j 1 j w un,j+1 u n,4g ] = det[v n,1 v n,j 1 j w vn,j+1 v n,4g ] and e + 0 ±e 1 J(4g) n, respectively: (414) for every 1 j 2g By definition of e 0 and e 1, u n,j = v n,j for every j n, 1 j 4g and u n,n = v n,n +w Therefore, (414) is trivial if j = n If 1 j < n, det[u n,1 u n,j 1 j w un,j+1 u n,4g ] = det[v n,1 v n,j 1 j w vn,j+1 v n,4g ]+det[v n,1 j w n w vn,4g ] = det[v n,1 v n,j 1 j w vn,j+1 v n,4g ]+0 Hence we obtain (414) Lemma 411 We have and for 3 n 2g det(e + 0 ±e 1 J(4g) 1 ) = det(e + 0 ±e 1 J(4g) 2 ) = det(e + 0 ) det(e + 0 ±e 1 J(4g) n ) = det(e + 0 ±e 0 J(4g) n 2 ) Proof We have det(e + 0 ± e 1 J(4g) 1 ) = det(e + 0 ), since e 1 J(4g) 1 = 0 On the other hand, det(e + 0 ±e 1 J(4g) 2 ) = C 4g g = det(e+ 0 ), since e+ 0 is a lower triangular matrix of size 4g and For 3 n 2g, we have e 1 J(4g) 2 = e + 0 ±e 1 J(4g) n = 0 0 0 0 C g 0 0 [e + 0 ±e 0 J(4g) n 2 ] տ4g 1 The 4g-th row and the 4g-th column of e + 0 ± e 0 J(4g) n 2 are [0 0 C g C g ] and t [0 0 C g ], respectively Therefore, det([e + 0 ±e 0 J(4g) n 2 ] տ4g 1) = C 1 g det(e+ 0 ±e 0 J(4g) n 2 ) Hence, we obtain the equality of Lemma 411

Lemma 412 For 1 n 2g, we have det(e + 0 ±E n INVERSE PROBLEMS FOR CANONICAL SYSTEMS 25 J) = C8g 2n g det([e 0 + +E 0 J(8g) n ] տn )det([e 0 + E 0 J(8g) n ] տn ) (415) Proof By the proof of Lemma 49, det(e + 0 ±E n Here J) = C4g 2n g det([e + 0 ] տ2g+n +[e 2,n ] ւ2g+n)det([e + 0 ] տ2g+n [e 2,n ] ւ2g+n) [e + 0 ] տ2g+n ±[e 2,n ] ւ2g+n = [E + 0 ±E 0 J(8g) n ] տ2g+n, since [e + 0 ] տ2g+n = [E + 0 ] տ2g+n and [e 2,n ] ւ2g+n = [E 0 J(8g) n ] տ2g+n by definition of matrices Applying Lemma 33 to the 2g columns of [E + 0 ± E 0 J(8g) n ] տ2g+n with indices (n+1,n+2,,2g +n), we have det[e + 0 ±E 0 J(8g) n ] տ2g+n = C 2g g det([e+ 0 ±E 0 J(8g) n ] տn ), since all entries above [E 0 + ±E 0 J(8g) n ] ց2g in [E 0 + ±E 0 J(8g) n ] տ2g+n are zero and [E 0 + ± E0 J(8g) n ] ց2g is lower triangular with diagonal (C g,,c g ) Therefore, we obtain (415) Lemma 413 We have det(e + 0 ±E 0 J(8g) n ) ( 1 (u ± n ()+v± n (0))) = det(e + 0 ±E 0 J(8g) n 2 ) Proof Let w = t[ C g C g 0 0 ] be the column vector of length 8g Denote by u ± n,j the columns of E+ 0 ±E nj By Cramer s formula, det(e + 0 ±E nj)(1 (u ± n()+v ± n(0))) = det(e + 0 ±E nj) det[u ± n,1 n w u ± n,8g ] det[u ± n,1 6g w u ± n,8g ] = det(e + 0 ±E n J) det[u± n,1 n w u ± n,8g ] det[u ± n,1 6g ±w u ± n,8g ] = det(e + 0 ±E nj) det[u ± n,1 n w u ± n,8g ] det[u ± n,1 6g u ± n,6g u± n,8g ]+det[u± n,1 6g u ± n,6g w u± n,8g ] = det[u ± n,1 n w u ± n,8g ]+det[u ± n,1 6g u ± n,6g w u± n,8g ] n 6g 6g n w = det[u ± n,1 w ±w u ± n,8g ] det[u± n,1 u ± n,6g w u± n,8g ] +det[u ± n,1 6g u ± n,6g w u± n,8g ] n 6g = det[u ± n,1 u ± n,6g w Therefore, it is sufficient to prove det(e + 0 ±E 0 J(8g) n 2 )det(e+ 0 ±E n J) = det(e + 0 ±E 0 J(8g) n )det[u ± n,1 u ± n,6g w u± n,8g ] u ± n,6g n w 6g Applying Lemma 33 to the columns (n+1,n+2,,8g), (416) u ± n,6g w u± n,8g ] det(e + 0 ±E 0 J(8g) n 2 ) = det([e+ 0 ±E 0 J(8g) n 2 ] տn 2)det([E + 0 ±E 0 J(8g) n 2 ] ց8g n+2) = C 8g n+2 g det([e 0 + ±E 0 J(8g) n 2 ] տn 2),

26 M SUZUKI since all entries above [E 0 + ±E 0 J(8g) n 2 ] ց8g n+2 are zero and [E 0 + ±E 0 J(8g) n 2 ] ց8g n+2 is lower triangular with diagonal (C g,,c g ) By the proof of Lemma 49, det(e + 0 ±E n J) = C4g n g det([e 0 + ] տ2g+n +[e 1,n ] ր2g+n)det([e 0 + ] տ2g+n [e 1,n ] ր2g+n), since [e 1,n ] ր2g+n = [e 2,n ] ւ2g+n We observe that By Lemma 412, [E + 0 ] տ2g+n ±[e 1,n ] ր2g+n = [E + 0 ±E 0 J(8g) n ] տ2g+n (LHS of (416)) = C 16g 2n+2 g det([e 0 + ±E 0 J(8g) n 2 ] տn 2) On the right hand side of (416), we have det([e + 0 +E 0 J(8g) n ] տn )det([e + 0 E 0 J(8g) n ] տn ) det(e 0 + ±E 0 J(8g) n ) = C 8g n g det([e + 0 ±E 0 J(8g) n ] տn ) by applying Lemma 33 to the columns (n+1,n+2,,8g) of E + 0 ±E 0 J(8g) n Therefore, it is sufficient to prove D ± n := det[u ± n,1 u ± n,6g n w 6g u ± n,6g w u± n,8g ] = C 8g n+2 g det([e 0 + E 0 J(8g) n ] տn )det([e 0 + ±E 0 J(8g) n 2 ] տn 2) Subtracting the (6g j)-th column from the (n j)-th column for 1 j, D ± n = det[v± n,1 v± n, u ± n,6g n w 6g u ± n,6g w u± n,8g ], (417) where v ± n,n j = u± n,n j u n,6g j for 1 j Successively, adding the (n j+1)-th column to the (6g j 1)-th column for 1 j n 2, D ± n = det[v ± n,1 v± n, n u ± n,6g w v n,6g n+1 v n,6g 2 6g u ± n,6g w u± n,8g ], where v n,6g j 1 ± = u± n,6g j 1 v± n,n j+1 for 1 j n 2 Note that the columns (n+1,,6g n) are not changed Put M ± n := [ v ± n,1 v± n, n u ± n,6g w v n,6g n+1 v n,6g 2 By the deformation of the matrix, we observe that [M ± n ] տ4g = [E + 0 E 0 J(8g) n ] տ4g Therefore, by applying Lemma 33 to the columns (n+1,n+2,,4g), D n ± = C4g n g det([e 0 + E 0 J(8g) n ] տn )det([m n ± ] ց4g) 6g u ± n,6g w u± n,8g] Here we note that [M n ± ] ց2g+n = [E 0 + E 0 J(8g) n 2 ] տ2g+n and that the columns (4g+1,,6g n) are not changed by the deformation of matrix Then, by applying Lemma 33 to the rows (1,,2g n) of [M n ± ] ց2g+n, D n ± = C 6g 2n g det([e 0 + E 0 J(8g) n ] տn )det([e 0 + E 0 J(8g) n 2 ] տ2g+n) Now (417) is proved = C 8g 2n+2 g det([e 0 + E 0 J(8g) n ] տn )det([e 0 + E 0 J(8g) n 2 ] տn 2)

Lemma 414 We have for every 1 n 2g Proof If 1 n 2g, we have Therefore, INVERSE PROBLEMS FOR CANONICAL SYSTEMS 27 det(e 0 + ±E 0 J(8g) n ) = C 4g g det(e+ 0 ±e 0 J(4g) n ) = C 6g 1 g E 0 + ±E 0 J(8g) n = det(e + ±E J (2g+1) n ) e + 0 ±e 0 J(4g) n 0 0 e + 0 det(e 0 + ±E 0 J(8g) n ) = C 4g g det(e+ 0 ±e 0 J(4g) n ), since det(e + 0 ) = C4g g Further, if 1 n 2g, we have e + 0 ±e 0 J(4g) n = E + ±E J (2g+1) n 0 [e + 0 ] ց(2g 1) Hence det(e 0 + ±E 0 J(8g) n ) = C 6g 1 g det(e + ±E J n (2g+1) ), and we complete the proof Lemma 415 We have det(e 0 + ±E 1,1 J) = det(e+ 0 ) and det(e 0 + ±E n,1 J) = det(e+ 0 ±E J) for 2 n 2g Proof By the proof of Lemma 49, for det(e 0 + ±E J) = C4g 2n+2 g det(a +B )det(a B ) (418) A = [e + 0 ] տ2g+, B = [e 2, ] ւ2g+ Ontheotherhand,applyingLemma33tothecolumns(2g+n+1,,4g) ofe 0 + ±E n,1 J, and then applying Lemma 33 to the rows (2g +n+1,4g) of the resulting matrix, [ det(e 0 + An ±B ±E n,1j) = C4g 2n g det ], ±B n A n [ where B = [e 2, ] An ±B ւ2g+n Applying Lemma 33 to the first row of ] ±B n A n and then applying Lemma 33 to the (4g +2)-th column of the resulting matrix, An ±B det[ ] = C ±B n A g 2 det A ±B n ±B A Hence = C 2 g det(a +B )det(a B ) det(e 0 + ±E n,1j) = C4g 2n+2 g det(a +B )det(a B ) (419) We complete the proof by comparing (418) and (419)