This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.


Generalized Hurwitz polynomials

Mikhail Tyaglov
Institut für Mathematik, Technische Universität Berlin
The work was supported by the Sofja Kovalevskaja Research Prize of Alexander von Humboldt Foundation. Email: tyaglov@math.tu-berlin.de
(August 19, 2025)
Abstract

We describe a wide class of polynomials, which is a natural generalization of Hurwitz stable polynomials. We also give a detailed account of so-called self-interlacing polynomials, which are dual to Hurwitz stable polynomials but have only real and simple zeroes. All proofs are given using properties of rational functions mapping the upper half-plane of the complex plane to the lower half-plane. Matrices with self-interlacing spectra and other applications of generalized Hurwitz polynomials are discussed.

Introduction

The polynomials with zeroes in the open left half-plane of the complex plane are called Hurwitz stable, or just stable. They play an important role in several areas of mathematics and engineering such as stability and bifurcation theory, control theory etc. The Hurwitz stable polynomials are well-studied and closely related to the moment problem, spectral theory, operator theory and orthogonal polynomials. The Hurwitz stable polynomials were studied by C. Hermite, E. Routh, A. Stodola, A. Hurwitz [13, 35, 17, 22, 12]. The present paper was motivated by some problems of the bifurcation theory and concerns the most natural (from our point of view) generalization of Hurwitz stable polynomials in the following way.

Let

p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2})

be a real polynomial, where p0p_{0} and p1p_{1} are its even and odd parts, respectively. Consider its associated rational function

Φ(u)=p1(u)p0(u).\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}~.

In the present paper we study polynomials whose associated function Φ\Phi is a so-called R-function, that is, the function mapping the upper half-plane of the complex plane to the lower half-plane. These functions have real, simple, and interlacing zeroes and poles [19, 16]. Such functions play a very important role in spectral theory of self-adjoint operators [37], in the moment problem [2, 1, 23], in the theory of the Stieltjes string [19], in the theory of oscillation operators [10], in the theory of Jacobi and Stieltjes continued fractions [41] etc. In particular, it is very well known [22, 7, 12, 6] that the polynomial pp is Hurwitz stable if and only if its associated function Φ\Phi is an R-function with negative zeroes and poles, and degp0degp1\deg p_{0}\geqslant\deg p_{1}. In Section 3 of the present paper, we prove this fact using methods of complex analysis, and then from this fact we obtain all main properties of Hurwitz stable polynomials and the main criteria of the polynomial stability including connections with total positivity and Stieltjes continued fractions. Section 3.2 is devoted to polynomials with zeroes in the closed half-plane of the complex plane. Note that in [3, 20] there was proved that the infinite Hurwitz matrix of a quasi-stable polynomial is totally nonnegative. In Section 3.2, we proved the converse statement, which is new.

In Section 4, we investigate polynomials whose associated function Φ\Phi maps the upper half-plane to the lower half-plane (with degp0degp1\deg p_{0}\geqslant\deg p_{1}) but has only positive zeroes and poles in contrast with the Hurwitz stable case . We call those polynomials self-interlacing. If p(z)p(z) is a self-interlacing polynomial, then it has only simple real zeroes which interlace with zeroes of the polynomial p(z)p(-z). More exactly, if λi\lambda_{i}, i=1,,ni=1,\ldots,n, are the zeroes of the self-interlacing polynomial pp (of type I), then

0<λ1<λ2<λ3<<(1)n1λn.0<\lambda_{1}<-\lambda_{2}<\lambda_{3}<\ldots<(-1)^{n-1}\lambda_{n}.

This fact is very curious, because the deep structure of self-interlacing polynomials is very close to the structure of the Hurwitz polynomials but the former ones have only real and simple zeroes. The main result that reveals this connection is Theorem 4.8, which says that a polynomial p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) is self-interlacing if and only if the polynomial q(z)=p(z2)zp1(z2)q(z)=p(-z^{2})-zp_{1}(z^{2}) is Hurwitz stable. Section 4 is devoted to the comprehensive description of self-interlacing polynomials and includes analogues of the Hurwitz criterion of stability and of the Liénard and Chipart criterion of stability, and some other criteria of self-interlacing. We also give some examples of matrices whose characteristic polynomials are self-interlacing and describe some numerical observations related to the self-interlacing polynomials. Note that there are three R-functions connected to the self-interlacing polynomials. Namely, let pp be self-interlacing. Then it has only real and simple zeroes, so its logarithmic derivative pp\dfrac{p^{\prime}}{p} is an R-function (see e.g. [16]). Moreover, pp is self-interlacing if and only if its associated function Φ=p1p0\Phi=\dfrac{p_{1}}{p_{0}} is an R-function with positive zeroes and poles (Theorem 4.3). Finally, pp is self-interlacing if and only if the function p(z)p(z)-\dfrac{p(z)}{p(-z)} is an R-function (see the proof of Theorem 4.3).

The full generalization of the Hurwitz stable polynomials (in terms of the function Φ\Phi) is given in Section 5. We describe the polynomials whose associated function Φ\Phi is an R-function without any restriction on the signs of its zeroes and poles. We call such polynomials generalized Hurwitz polynomials. More exactly, the polynomial pp of degree nn is called generalized Hurwitz of order kk (of type I), 0k[n+12]0\leqslant k\leqslant\left[\dfrac{n+1}{2}\right], if it has exactly kk zeroes in the closed right half-plane, all of which are nonnegative and simple:

0μ1<μ2<<μk,0\leqslant\mu_{1}<\mu_{2}<\cdots<\mu_{k},

such that p(μi)0p(-\mu_{i})\neq 0, i=1,,ki=1,\ldots,k, and pp has an odd number of zeroes, counting multiplicities, on each interval (μk,μk1),,(μ3,μ2),(μ2,μ1)(-\mu_{k},-\mu_{k-1}),\ldots,(-\mu_{3},-\mu_{2}),(-\mu_{2},-\mu_{1}). Moreover, the number of zeroes of pp on the interval (μ1,0)(-\mu_{1},0) (if any) is even, counting multiplicities. The other real zeroes lie on the interval (,μk)(-\infty,-\mu_{k}): an odd number of zeroes, counting multiplicities, when n=2ln=2l, and an even number of zeroes, counting multiplicities, when n=2l+1n=2l+1. All nonreal zeroes of pp (if any) are located in the open left half-plane of the complex plane.

We prove a generalization of Hurwitz theorem, Theorem 5.6, which, in fact, is an analogue of Hurwitz criterion of stability for the generalized Hurwitz polynomials. We also establish a generalization of the Liénard and Chipart criterion of stability, Theorem 5.8. Note that in [21] G. F. Korsakov made an attempt to prove this generalization of the Liénard and Chipart criterion. However, his methods did not allow him to prove the whole theorem (for details, see Section 5.1) .

We also should note that V. Pivovarchik in [31] describes some aspects of the generalized Hurwitz polynomials, however, in different form, replacing the lower half-plane by the upper half-plane. In particular, he indirectly establishes that the polynomial pp is generalized Hurwitz (shifted Hermite-Biehler with symmetry, in Pivovarchik’s terminology) if and only if its associated function Φ\Phi is an R-function. Precisely, the cases considered in [31] are degp0>degp1\deg p_{0}>\deg p_{1} and degp0=degp1\deg p_{0}=\deg p_{1} with the leading coefficients of p0p_{0} and p1p_{1} of the same sign. Neither generalizations of Hurwitz theorem nor generalizations of the Liénard and Chipart criterion, nor the theory of self-interlacing polynomials, nor connections of the generalized Hurwitz polynomials with continued fractions were established in [31]. However, V. Pivovarchik with H. Woracek described [33, 34] a generalization of his shifted Hermite-Biehler polynomials with symmetry (which are generalized Hurwitz polynomials up to rotation of the upper half-plane to the left half-plane) to entire functions. The main result of [33, 34] is an analogue of our Theorem 5.3 for entire functions, provided by two different methods. The method in [34] coincides with that of [31], where as [34] uses the theory of de Branges and Pontryagin spaces. Our proof of Theorem 5.3 differs from the methods used by V. Pivovarchik and H. Woracek and is based on using the properties of R-functions.

It is curious that V. Pivovarchik came to his shifted Hermite-Biehler polynomials and entire functions with symmetry due to his investigation of spectra of quadratic operator pencils [30] and Sturm-Liouville operators whose boundary conditions contain spectral parameter [32, 40]. The last two articles deal with a generalization of the Regge problem, which has a direct relation to the scattering theory. So the distribution of zeroes of the generalized Hurwitz polynomials has a physical interpretation.

Finally, in Section 5.2 we explain how the generalized Hurwitz polynomials can be used in bifurcation theory. Also, we describe how specific zeroes of a real polynomial p(z,α)p(z,\alpha) move with the movement of the parameter α\alpha along \mathbb{R} if p(z,α)p(z,\alpha) is generalized Hurwitz for all real α\alpha.

1 Auxiliary definitions and theorems

Consider a rational function

R(z)=q(z)p(z),R(z)=\dfrac{q(z)}{p(z)}, (1.1)

where pp and qq are polynomials with complex coefficients

p(z)=a0zn+a1zn1++an,\displaystyle p(z)=a_{0}z^{n}+a_{1}z^{n-1}+\dots+a_{n},\qquad a0,a1,,an,a0>0,\displaystyle a_{0},a_{1},\dots,a_{n}\in\mathbb{R},\ a_{0}>0, (1.2)
q(z)=b0zn+b1zn1++bn,\displaystyle q(z)=b_{0}z^{n}+b_{1}z^{n-1}+\dots+b_{n},\qquad b0,b1,,bn,\displaystyle b_{0},b_{1},\dots,b_{n}\in\mathbb{R}, (1.3)

so degp=n\deg p=n and degqn\deg q\leqslant n. If the greatest common divisor of pp and qq has degree ll, 0ln0\leq l\leq n, then the rational function RR has exactly r=nlr=n-l poles.

1.1 Hankel matrices and Hankel minors

Expand the function (1.1) into its Laurent series at \infty:

R(z)=s1+s0z+s1z2+s2z3+.R(z)=s_{-1}+\dfrac{s_{0}}{z}+\dfrac{s_{1}}{z^{2}}+\dfrac{s_{2}}{z^{3}}+\cdots. (1.4)

Here sj=0s_{j}=0 for j<n1mj<n-1-m and sn1m=b0a0s_{n-1-m}=\dfrac{b_{0}}{a_{0}}, where m=degqm=\deg q.

The sequence of coefficients of negative powers of zz

s0,s1,s2,s_{0},s_{1},s_{2},\ldots

defines the infinite Hankel matrix S:=S(R):=si+ji,j=0S\mathop{{:}{=}}S(R)\mathop{{:}{=}}\|s_{i+j}\|_{i,j=0}^{\infty}.

Definition 1.1.

For a given infinite sequence (sj)j=0(s_{j})_{j=0}^{\infty}, consider the determinants

Dj(S)=|s0s1s2sj1s1s2s3sjsj1sjsj+1s2j2|,j=1,2,3,,D_{j}(S)=\begin{vmatrix}s_{0}&s_{1}&s_{2}&\dots&s_{j-1}\\ s_{1}&s_{2}&s_{3}&\dots&s_{j}\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ s_{j-1}&s_{j}&s_{j+1}&\dots&s_{2j-2}\end{vmatrix},\quad j=1,2,3,\dots, (1.5)

i.e., the leading principal minors of the infinite Hankel matrix SS. These determinants are referred to as the Hankel minors or Hankel determinants.

An infinite matrix is said to have finite rank rr if all its minors of order greater than rr equal zero whereas there exists at least one nonzero minor of order rr. Kronecker [24] proved that, for any infinite Hankel matrix, any minor of order rr where rr is the rank of the matrix, is a multiple of its leading principal minor of order rr. This implies the following result.

Theorem 1.2 (Kronecker [24]).

An infinite Hankel matrix S=si+j0S=\|s_{i+j}\|_{0}^{\infty} has finite rank rr if and only if

Dr(S)\displaystyle D_{r}(S) \displaystyle\neq 0,\displaystyle 0, (1.6)
Dj(S)\displaystyle D_{j}(S) =\displaystyle= 0,for all j>r.\displaystyle 0,\quad\hbox{\rm for all }j>r. (1.7)

The following theorem was established by Gantmacher in [12].

Theorem 1.3.

An infinite Hankel matrix S=si+j0S=\|s_{i+j}\|_{0}^{\infty} has finite rank if and only if the sum of the series

R(z)=s0z+s1z2+s2z3+R(z)=\dfrac{s_{0}}{z}+\dfrac{s_{1}}{z^{2}}+\dfrac{s_{2}}{z^{3}}+\cdots

is a rational function of zz. In this case the rank of the matrix SS is equal to the number of poles of the function RR.

Theorems 1.2 and 1.3 imply the following: if the greatest common divisor of the polynomials pp and qq defined in (1.2)–(1.3) has degree ll, 0lm0\leqslant l\leqslant m, then the formulæ (1.6)–(1.7) hold for r=nlr=n-l for the rational function (1.1).

Let D^j(S)\widehat{D}_{j}(S) denote the following determinants

D^j(S)=|s1s2s3sjs2s3s4sj+1sjsj+1sj+2s2j1|,j=1,2,3,\widehat{D}_{j}(S)=\begin{vmatrix}s_{1}&s_{2}&s_{3}&\dots&s_{j}\\ s_{2}&s_{3}&s_{4}&\dots&s_{j+1}\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ s_{j}&s_{j+1}&s_{j+2}&\dots&s_{2j-1}\end{vmatrix},\quad j=1,2,3,\dots (1.8)

With a slight abuse of notation, we will also write Dj(R):=Dj(S(R))D_{j}(R)\mathop{{:}{=}}D_{j}(S(R)) and D^j(R):=D^j(S(R))\widehat{D}_{j}(R)\mathop{{:}{=}}\widehat{D}_{j}(S(R)) if the matrix S=S(R)S=S(R) is made of the coefficients (1.4) of the function RR.

Theorems 1.2 and 1.3 have the following simple corollaries, which will be useful later.

Corollary 1.4 ([16]).

A rational function RR represented by the series (1.4) has at most rr poles if and only if

D^j(R)=0forj=r+1,r+2,.\widehat{D}_{j}(R)=0\quad{\rm for}\;\;j=r+1,r+2,\ldots.
Corollary 1.5 ([16]).

A rational function RR with exactly rr poles represented by the series (1.4) has a pole at the point 0 if and only if

D^r1(R)0andD^r(R)=0.\widehat{D}_{r-1}(R)\neq 0\quad\text{and}\quad\widehat{D}_{r}(R)=0.

Otherwise, D^r(R)0\widehat{D}_{r}(R)\neq 0.

In the sequel, we are also interested in the number of sign changes in the sequences of Hankel minors Dj(R)D_{j}(R) and D^j(R)\widehat{D}_{j}(R), which we denote by111Here we set D0(R)=D^0(R)1D_{0}(R)=\widehat{D}_{0}(R)\equiv 1. V(D0(R),D1(R),,Dr(R))\operatorname{{\rm V}}(D_{0}(R),D_{1}(R),\ldots,D_{r}(R)) and V(D^0(R),D^1(R),,D^r(R))\operatorname{{\rm V}}(\widehat{D}_{0}(R),\widehat{D}_{1}(R),\ldots,\widehat{D}_{r}(R)), respectively. These numbers exist, since RR is a real function.

In his remarkable work [9], G. Frobenius proved the following fact that allow us to calculate the number of sign changes in the sequence D0(R),D1(R),,Dr(R)D_{0}(R),D_{1}(R),\ldots,D_{r}(R) in the case when some minors Dj(R)D_{j}(R) are zero.

Theorem 1.6 (Frobenius [9, 11]).

If, for some integers ii and jj (0i<j)(0\leq i<j),

Di(R)0,Di+1(R)=Di+2(R)==Di+j(R)=0,Di+j+1(R)0,D_{i}(R)\neq 0,\quad D_{i+1}(R)=D_{i+2}(R)=\cdots=D_{i+j}(R)=0,\quad D_{i+j+1}(R)\neq 0, (1.9)

then the number V(D0(R),D1(R),D2(R),,Dr(R))\operatorname{{\rm V}}(D_{0}(R),D_{1}(R),D_{2}(R),\ldots,D_{r}(R)) of Frobenius sign changes should be calculated by assigning signs as follows:

signDi+ν(R)=(1)ν(ν1)2signDi(R),ν=1,2,,j.\mathop{\rm sign}\nolimits D_{i+\nu}(R)=(-1)^{\tfrac{\nu(\nu-1)}{2}}\mathop{\rm sign}\nolimits D_{i}(R),\quad\nu=1,2,\ldots,j. (1.10)
Definition 1.7.

A matrix (finite or infinite) is called totally nonnegative (strictly totally positive) if all its minors are nonnegative (positive).

Definition 1.8.

An infinite matrix of finite rank rr is called mm-totally nonnegative (mm-strictly totally positive) of rank rr, mrm\leqslant r, if all its minors are nonnegative (positive) up to order mm inclusively.

An rr-totally nonnegative (rr-strictly totally positive) matrix of rank rr is called totally nonnegative (strictly totally positive) of rank rr.

Definition 1.9.

An infinite matrix is called positive (nonnegative) definite of rank rr if all its leading principal minors are positive (nonnegative) up to order rr inclusive.

For the matrix AA (finite or infinite), we denote its minor of order jj constructed with rows i1,i2,,iji_{1},i_{2},\ldots,i_{j} and columns l1,l2,,ljl_{1},l_{2},\ldots,l_{j} by

A(i1i2ijl1l2lj).A\begin{pmatrix}i_{1}&i_{2}&\dots&i_{j}\\ l_{1}&l_{2}&\dots&l_{j}\\ \end{pmatrix}.
Definition 1.10 ([10]).

An infinite matrix AA of finite rank rr is called mm-sign regular, mrm\leqslant r, if all its minors up to order mm (inclusively) satisfy the following inequalities:

(1)k=1jik+k=1jlkA(i1i2ijl1l2lj)>0,j=1,,m.(-1)^{\sum\limits_{k=1}^{j}i_{k}+\sum\limits_{k=1}^{j}l_{k}}A\begin{pmatrix}i_{1}&i_{2}&\dots&i_{j}\\ l_{1}&l_{2}&\dots&l_{j}\end{pmatrix}>0,\qquad j=1,\ldots,m. (1.11)

An rr-sign regular matrix of rank rr is called sign regular of rank rr.

In [12] there was proved the following criterion of total positivity of infinite Hankel matrices of finite order.

Theorem 1.11.

An infinite Hankel matrix S=si+j0S=\|s_{i+j}\|_{0}^{\infty} of finite rank rr is strictly totally positive if and only if the following inequalities hold:

Dj(S)>0,D^j(S)>0,j=1,,r.\begin{split}&D_{j}(S)>0,\\ &\widehat{D}_{j}(S)>0,\end{split}\qquad j=1,\ldots,r. (1.12)

In [16], the following fact was proved.

Theorem 1.12.

An infinite Hankel matrix S=si+j0S=\|s_{i+j}\|_{0}^{\infty} of finite rank rr is sign regular if and only if the following inequalities hold:

Dj(S)>0,(1)jD^j(S)>0,j=1,,r.\begin{split}&D_{j}(S)>0,\\ &(-1)^{j}\widehat{D}_{j}(S)>0,\end{split}\qquad j=1,\ldots,r. (1.13)

1.2 Matrices and minors of Hurwitz type

In this section we introduce infinite matrices (matrices of Hurwitz type) associated with the polynomials pp and qq defined in (1.2)–(1.3) and discuss the Hurwitz formulaæthat connect Hankel matrices with matrices of Hurwitz type.

Definition 1.13.

Given polynomials pp and qq from (1.2)–(1.3), define the infinite matrix H(p,q)H(p,q) as follows: if degq<degp\deg q<\deg p, that is, if b0=0b_{0}=0, then

H(p,q)=(a0a1a2a3a4a50b1b2b3b4b50a0a1a2a3a400b1b2b3b4),H(p,q)=\begin{pmatrix}a_{0}&a_{1}&a_{2}&a_{3}&a_{4}&a_{5}&\dots\\ 0&b_{1}&b_{2}&b_{3}&b_{4}&b_{5}&\dots\\ 0&a_{0}&a_{1}&a_{2}&a_{3}&a_{4}&\dots\\ 0&0&b_{1}&b_{2}&b_{3}&b_{4}&\dots\\ \vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\ddots\end{pmatrix}, (1.14)

if degq=degp\deg q=\deg p, that is, b00b_{0}\neq 0, then

H(p,q)=(b0b1b2b3b4b50a0a1a2a3a40b0b1b2b3b400a0a1a2a3).H(p,q)=\begin{pmatrix}b_{0}&b_{1}&b_{2}&b_{3}&b_{4}&b_{5}&\dots\\ 0&a_{0}&a_{1}&a_{2}&a_{3}&a_{4}&\dots\\ 0&b_{0}&b_{1}&b_{2}&b_{3}&b_{4}&\dots\\ 0&0&a_{0}&a_{1}&a_{2}&a_{3}&\dots\\ \vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\ddots\end{pmatrix}. (1.15)

The matrix H(p,q)H(p,q) is called an infinite matrix of Hurwitz type. We denote the leading principal minor of H(p,q)H(p,q) of order jj, j=1,2,j=1,2,\ldots, by ηj(p,q)\eta_{j}(p,q).

Remark 1.14.

The matrix H(p,q)H(p,q) is of infinite rank since its submatrix obtained by deleting the even or odd rows of the original matrix is a triangular infinite matrix with a00a_{0}\neq 0 on the main diagonal.

Together with the infinite matrix H(p,q)H(p,q), we consider its specific finite submatrices:

Definition 1.15.

Let the polynomials pp and qq are given by (1.2)–(1.3).

If degq<degp=n\deg q<\deg p=n, then we construct the 2n×2n{2n}\times{2n} matrix

2n(p,q)=(b1b2b3bn0000a0a1a2an1an0000b1b2bn1bn0000a0a1an2an1an00000a0a1a2an00000b1b2bn00000a0a1an1an),\mathcal{H}_{2n}(p,q)=\begin{pmatrix}b_{1}&b_{2}&b_{3}&\dots&b_{n}&0&0&\dots&0&0\\ a_{0}&a_{1}&a_{2}&\dots&a_{n-1}&a_{n}&0&\dots&0&0\\ 0&b_{1}&b_{2}&\dots&b_{n-1}&b_{n}&0&\dots&0&0\\ 0&a_{0}&a_{1}&\dots&a_{n-2}&a_{n-1}&a_{n}&\dots&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&\dots&a_{0}&a_{1}&a_{2}&\dots&a_{n}&0\\ 0&0&0&\dots&0&b_{1}&b_{2}&\dots&b_{n}&0\\ 0&0&0&\dots&0&a_{0}&a_{1}&\dots&a_{n-1}&a_{n}\\ \end{pmatrix}, (1.16)

if degq=degp=n\deg q=\deg p=n, then we construct the (2n+1)×(2n+1)(2n{+}1)\times(2n{+}1) matrix

2n+1(p,q)=(a0a1a2an1an000b0b1b2bn1bn0000a0a1an2an1an000b0b1bn2bn1bn00000a0a1a2an0000b0b1b2bn00000a0a1an1an).\mathcal{H}_{2n+1}(p,q)=\begin{pmatrix}a_{0}&a_{1}&a_{2}&\dots&a_{n-1}&a_{n}&0&\dots&0&0\\ b_{0}&b_{1}&b_{2}&\dots&b_{n-1}&b_{n}&0&\dots&0&0\\ 0&a_{0}&a_{1}&\dots&a_{n-2}&a_{n-1}&a_{n}&\dots&0&0\\ 0&b_{0}&b_{1}&\dots&b_{n-2}&b_{n-1}&b_{n}&\dots&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&\dots&a_{0}&a_{1}&a_{2}&\dots&a_{n}&0\\ 0&0&0&\dots&b_{0}&b_{1}&b_{2}&\dots&b_{n}&0\\ 0&0&0&\dots&0&a_{0}&a_{1}&\dots&a_{n-1}&a_{n}\\ \end{pmatrix}. (1.17)

Both kinds of matrices 2n(p,q)\mathcal{H}_{2n}(p,q) and 2n+1(p,q)\mathcal{H}_{2n+1}(p,q) are called finite matrices of Hurwitz type. The leading principal minors of these matrices will be denoted by222That is, j(p,q)\nabla_{j}(p,q) is the leading principal minor of the matrix 2n(p,q)\mathcal{H}_{2n}(p,q) of order jj if degq<degp\deg q<\deg p. Otherwise (when degq=degp\deg q=\deg p), j(p,q)\nabla_{j}(p,q) denotes the leading principal minor of the matrix 2n+1(p,q)\mathcal{H}_{2n+1}(p,q) of order jj. j(p,q)\nabla_{j}(p,q).

In his celebrated work [17], A. Hurwitz found relationships between the minors Dj(R)D_{j}(R) and D^j(R)\widehat{D}_{j}(R) defined in (1.5) and (1.8) and the leading principal minors ηi(p,q)\eta_{i}(p,q) of the matrix H(p,g)H(p,g) (see Definition 1.13).

Lemma 1.16 ([17, 22, 12, 6, 16]).

Let the polynomials pp and qq be defined in (1.2)–(1.3) and let

R(z)=q(z)p(z)=s1+s0z+s1z2+R(z)=\dfrac{q(z)}{p(z)}=s_{-1}+\dfrac{s_{0}}{z}+\dfrac{s_{1}}{z^{2}}+\cdots

The following relations hold between the determinants ηj(p,q)\eta_{j}(p,q) and Dj(R),D^j(R)D_{j}(R),\widehat{D}_{j}(R) defined in (1.5) and (1.8), respectively.

  • If degq<degp\deg q<\deg p, then

    η2j(p,q)=a02jDj(R),j=1,2,;\eta_{2j}(p,q)=a_{0}^{2j}D_{j}(R),\quad j=1,2,\ldots; (1.18)
    η2j+1(p,q)=(1)ja02j+1D^j(R),j=0,1,2,\eta_{2j+1}(p,q)=(-1)^{j}a_{0}^{2j+1}\widehat{D}_{j}(R),\quad j=0,1,2,\ldots (1.19)
  • If degq=degp\deg q=\deg p, then

    η2j+1(p,q)=b0a02jDj(R),j=0,1,2,;\eta_{2j+1}(p,q)=b_{0}a_{0}^{2j}D_{j}(R),\quad j=0,1,2,\ldots; (1.20)
    η2j(p,q)=(1)j1b0a02j1D^j1(R),j=1,2,\eta_{2j}(p,q)=(-1)^{j-1}b_{0}a_{0}^{2j-1}\widehat{D}_{j-1}(R),\quad j=1,2,\ldots (1.21)

From (1.14)–(1.17) we have

  • If degq<degp\deg q<\deg p, then

    i(p,q)=a01ηi+1(p,q),i=1,2,,2n;\nabla_{i}(p,q)=a_{0}^{-1}\eta_{i+1}(p,q),\quad i=1,2,\ldots,2n; (1.22)
  • If degq=degp\deg q=\deg p, then

    i(p,q)=b01ηi+1(p,q),i=1,2,,2n+1.\nabla_{i}(p,q)=b_{0}^{-1}\eta_{i+1}(p,q),\quad i=1,2,\ldots,2n+1. (1.23)

1.3 Rational functions mapping the upper half-plane to the lower half-plane

All main theorems related to the generalized Hurwitz polynomials are based on properties of the following very important class of rational functions:

Definition 1.17.

A rational function FF is called R-function if it maps the upper half-plane of the complex plane to the lower half-plane333In [16] these functions are called R-functions of negative type.:

Imz>0ImF(z)<0.\operatorname{Im}z>0\Rightarrow\operatorname{Im}F(z)<0.

By now, these functions, as well as their meromorphic analogues, have been considered by many authors and have acquired various names. For instance, these functions are called strongly real functions in the monograph [36] due to their property to take real values only for real values of the argument (more general and detailed consideration can be found in [7], see also [16]).

The following theorem provides the most important (for the present study) properties of R-functions. Some parts of this theorem was considered in [29, 22, 7, 12, 4, 6]. An extended version of this theorem can be found (with proof) in [16, Theorem 3.4].

Theorem 1.18.

Let hh and gg be real polynomials such that degh1degg=n\deg h-1\leqslant\deg g=n. For the real rational function

F=hgF=\dfrac{h}{g}

with exactly r(n)r\,(\leqslant n) poles, the following conditions are equivalent:

  • 1)1)

    FF is an R-function:

    Imz>0ImF(z)<0;\operatorname{Im}z>0\Rightarrow\operatorname{Im}F(z)<0; (1.24)
  • 2)2)

    The function FF can be represented in the form

    F(z)=αz+β+j=1rγjz+ωj,α0,β,ωj,\displaystyle F(z)=-\alpha z+\beta+\sum^{r}_{j=1}\frac{\gamma_{j}}{z+\omega_{j}},\quad\alpha\geqslant 0,\,\,\,\beta,\omega_{j}\in\mathbb{R}, (1.25)

    where

    γj=h(ωj)g(ωj)>0,j=1,,r;\gamma_{j}=\dfrac{h(\omega_{j})}{g^{\prime}(\omega_{j})}>0,\quad j=1,\ldots,r; (1.26)
  • 3)3)

    The zeroes of the polynomials g~(z)\widetilde{g}(z) and h~(z)\widetilde{h}(z), where g~=g/f\widetilde{g}=g/f, h~=h/f\widetilde{h}=h/f, f=gcd(g,h)f=\gcd(g,h), are real, simple and interlacing, that is, between any two consecutive zeroes of one of polynomials there is exactly one zero of the other polynomial, counting multiplicity, and

    ω:g~(ω)h~(ω)g~(ω)h~(ω)<0;\exists\;\omega\in\mathbb{R}\,:\,\,\,\widetilde{g}(\omega)\widetilde{h}^{\prime}(\omega)-\widetilde{g}^{\prime}(\omega)\widetilde{h}(\omega)<0; (1.27)
  • 4)4)

    The polynomial

    g(z)=λp(z)+μq(z)g(z)=\lambda p(z)+\mu q(z) (1.28)

    has only real zeroes for any real λ\lambda and μ\mu, λ2+μ20\lambda^{2}+\mu^{2}\neq 0, and the condition (1.27) is satisfied;

  • 5)5)

    Let the function FF be represented by the series

    F(z)=s2z+s1+s0z+s1z2+s2z3+,F(z)=s_{-2}z+s_{-1}+\frac{s_{0}}{z}+\frac{s_{1}}{z^{2}}+\frac{s_{2}}{z^{3}}+\dots, (1.29)

    then s20s_{-2}\leqslant 0 and s1s_{-1}\in\mathbb{R}. The following inequalities hold

    Dj(F)>0,j=1,2,,r,D_{j}(F)>0,\quad j=1,2,\dots,r, (1.30)

    where determinants Dj(F)D_{j}(F) are defined in (1.5).

Remark 1.19.

If a function FF maps the upper half-plane to the upper half-plane444In [16] such functions are called R-functions of positive type., then the function F-F is an R-function.

Note that from (1.25) it follows that if FF is an R-function, then the function FF is decreasing between its poles, so F-F is obviously increasing. Therefore, from Theorem 1.18 we obtain the following well-known fact.

Corollary 1.20.

Let gg and hh be real polynomials such that |deghdegg|1|\deg h-\deg g|\leqslant 1 and let the zeroes of the polynomials gg and hh be real and simple. If zeroes of gg and hh interlace, then the function hg\dfrac{h}{g} is decreasing or increasing between its poles.

Remark 1.21.

If RR defined in (1.1) is an R-function, then by Theorem 1.18, we have degqdegp1\deg q\geqslant\deg p-1, that is, the equality b0=0b_{0}=0 implies b10b_{1}\neq 0 or, in other words, b02+b120b_{0}^{2}+b_{1}^{2}\neq 0.

In the sequel, we also use the following fact [16], which is a simple consequence of a theorem of V. Markov [7].

Theorem 1.22.

Let pp and qq be real coprime polynomials such that |degpdegq|1|\deg p-\deg q|\leqslant 1. If the function R=q/pR=q/p is an RR-function, then the functions Rj=q(j)/p(j)R_{j}=q^{(j)}/p^{(j)}, j=1,,degp1j=1,\ldots,\deg p-1, are also RR-functions.

The following simple properties of R-functions can be established using Theorem 1.18.

Theorem 1.23.

Let FF and GG be R-functions. Then

  • 1)

    the function R=F+GR=F+G is an R-function;

  • 2)

    the function F(z)-F(-z) is an R-function;

  • 3)

    the function 1F(z)-\dfrac{1}{F(z)} is an R-function;

  • 4)

    the function zF(z)zF(z) is an R-function if all poles of FF are positive and F(z)0F(z)\to 0 as zz\to\infty.

Now consider again the function RR defined in (1.1). By (1.2)–(1.3), the degree of its numerator is not greater than the degree of its denominator. If RR is an R-function, then one can find the numbers of its negative and positive poles (see e.g. [16]).

Theorem 1.24.

Let a rational function RR with exactly rr poles be an R-function of negative type and let RR have a series expansion (1.29). Then the number rr_{-} of negative poles of RR equals555Recall that the number V(1,D^1(R),D^2(R),,D^k(R))\operatorname{{\rm V}}(1,\widehat{D}_{1}(R),\widehat{D}_{2}(R),\ldots,\widehat{D}_{k}(R)) of Frobenius sign changes must be calculated according to Frobenius rule provided by Theorem 1.6.

r=V(1,D^1(R),D^2(R),,D^k(R)),r_{-}=\operatorname{{\rm V}}(1,\widehat{D}_{1}(R),\widehat{D}_{2}(R),\ldots,\widehat{D}_{k}(R)), (1.31)

where k=r1k=r-1, if R(0)=R(0)=\infty, and k=rk=r, if |R(0)|<|R(0)|<\infty. The determinants D^j(R)\widehat{D}_{j}(R) are defined in (1.8).

As was shown in [12] (see also [16]), if RR is an R-function has exactly rr poles, all of which are of the same sign, then all minors D^j(R)\widehat{D}_{j}(R) are non-zero up to order rr.

Corollary 1.25 ([12]).

Let a rational function RR with exactly rr poles (counting multiplicities) be an R-function of negative type. All poles of RR are positive if and only if

D^j(R)>0,j=1,,r,\widehat{D}_{j}(R)>0,\qquad j=1,\ldots,r,

where the determinants D^j(R)\widehat{D}_{j}(R) are defined in (1.8).

Corollary 1.26 ([16]).

Let a rational function RR with exactly rr poles be an R-function of negative type. All poles of RR are negative if and only if

(1)jD^j(R)>0,j=1,,r,(-1)^{j}\widehat{D}_{j}(R)>0,\qquad j=1,\ldots,r,

where the determinants D^j(R)\widehat{D}_{j}(R) are defined in (1.8).

One can use (1.18)–(1.23) to obtain criteria for the function RR defined in (1.1) to be an R-function in terms of Hurwitz minors.

Definition 1.27.

For a sequence of real numbers a0,a1,,ana_{0},a_{1},\ldots,a_{n}, we denote the number of sign changes in the sequence of its nonzero entries by v(a0,a1,,an)v(a_{0},a_{1},\ldots,a_{n}). The number v(a0,a1,,an)v(a_{0},a_{1},\ldots,a_{n}) is usually called the number of strong sign changes of the sequence a0,a1,,ana_{0},a_{1},\ldots,a_{n}.

In [16], the following theorem was established.

Theorem 1.28.

Let RR be a real rational function as in (1.1). If RR is an R-function with exactly nn poles, that is, gcd(p,q)1\gcd(p,q)\equiv 1, then the number of its positive poles equals v(a0,a1,,an)v(a_{0},a_{1},\ldots,a_{n}). In particular, RR has only negative poles if and only if 666In fact, the coefficients must be of the same signs, but we assume that a0>0a_{0}>0 (see (1.2)). aj>0a_{j}>0 for j=1,2,,nj=1,2,\ldots,n, and it has only positive poles if and only if aj1aj<0a_{j-1}a_{j}<0 for j=1,2,,nj=1,2,\ldots,n.

Remark 1.29.

It is easy to see that Theorem 1.28 is also true for an R-function with degree of numerator greater than degree of denominator.

For R-functions with only negative poles, there are a few more criteria [16].

Theorem 1.30.

The function (1.1), where degq<degp\deg q<\deg p, is an R-function and has exactly nn negative poles if and only if one of the following equivalent conditions holds

  • 1)1)

    an>0,an1>0,,a0>0,1(p,q)>0,3(p,q)>0,,2n1(p,q)>0a_{n}>0,a_{n-1}>0,\ldots,a_{0}>0,\quad\nabla_{1}(p,q)>0,\nabla_{3}(p,q)>0,\ldots,\nabla_{2n-1}(p,q)>0;

  • 2)2)

    an>0,bn>0,bn1>0,,b1>0,1(p,q)>0,3(p,q)>0,,2n1(p,q)>0a_{n}>0,b_{n}>0,b_{n-1}>0,\ldots,b_{1}>0,\quad\nabla_{1}(p,q)>0,\nabla_{3}(p,q)>0,\ldots,\nabla_{2n-1}(p,q)>0;

  • 3)3)

    an>0,an1>0,,a0>0,2(p,q)>0,4(p,q)>0,,2n(p,q)>0a_{n}>0,a_{n-1}>0,\ldots,a_{0}>0,\quad\nabla_{2}(p,q)>0,\nabla_{4}(p,q)>0,\ldots,\nabla_{2n}(p,q)>0;

  • 4)4)

    an>0,bn>0,bn1>0,,b1>0,2(p,q)>0,4(p,q)>0,,2n(p,q)>0a_{n}>0,b_{n}>0,b_{n-1}>0,\ldots,b_{1}>0,\quad\nabla_{2}(p,q)>0,\nabla_{4}(p,q)>0,\ldots,\nabla_{2n}(p,q)>0;

where i(p,q)\nabla_{i}(p,q) are defined in Definition 1.15.

In the case degq=degp\deg q=\deg p we have similar criteria.

Theorem 1.31.

The function (1.1), where degq(z)=degp(z)\deg q(z)=\deg p(z), is an R-function and has exactly nn negative poles if and only if one of the following equivalent conditions holds

  • 1)1)

    an>0,an1>0,,a0>0,2(p,q)>0,4(p,q)>0,,2n(p,q)>0a_{n}>0,a_{n-1}>0,\ldots,a_{0}>0,\quad\nabla_{2}(p,q)>0,\nabla_{4}(p,q)>0,\ldots,\nabla_{2n}(p,q)>0;

  • 2)2)

    an>0,bn>0,bn1>0,,b0>0,2(p,q)>0,4(p,q)>0,,2n(p,q)>0a_{n}>0,b_{n}>0,b_{n-1}>0,\ldots,b_{0}>0,\quad\nabla_{2}(p,q)>0,\nabla_{4}(p,q)>0,\ldots,\nabla_{2n}(p,q)>0;

  • 3)3)

    an>0,an1>0,,a0>0,1(p,q)>0,3(p,q)>0,,2n+1(p,q)>0a_{n}>0,a_{n-1}>0,\ldots,a_{0}>0,\quad\nabla_{1}(p,q)>0,\nabla_{3}(p,q)>0,\ldots,\nabla_{2n+1}(p,q)>0;

  • 4)4)

    an>0,bn>0,bn1>0,,b0>0,1(p,q)>0,3(p,q)>0,,2n+1(p,q)>0a_{n}>0,b_{n}>0,b_{n-1}>0,\ldots,b_{0}>0,\quad\nabla_{1}(p,q)>0,\nabla_{3}(p,q)>0,\ldots,\nabla_{2n+1}(p,q)>0;

where i(p,q)\nabla_{i}(p,q) are defined in Definition 1.15.

At last, we recall a relation between R-functions with nonpositive poles and some properties of Hurwitz matrices [16].

Theorem 1.32 (Total Nonnegativity of the Hurwitz Matrix).

The following are equivalent:

  • 1)1)

    The polynomials pp and qq defined in (1.2)–(1.3) have only nonpositive zeroes (or q(z)0q(z)\equiv 0), and the function RR defined in (1.1) is either an R-function (or R(z)0R(z)\equiv 0).

  • 2)2)

    The infinite matrix of Hurwitz type H(p,q)H(p,q) defined in (1.14)–(1.15) is totally nonnegative.

Theorem 1.32 implies the following corollary [16].

Corollary 1.33.

Let the polynomials pp and qq be defined in (1.2)–(1.3). The following are equivalent:

  • 1)1)

    The function R=pqR=\dfrac{p}{q} is an R-function with exactly nn poles, all of which are negative.

  • 2)2)

    The infinite matrix of Hurwitz type H(p,q)H(p,q) defined in (1.14)–(1.15) is totally nonnegative and ηk+1(p,q)0\eta_{k+1}(p,q)\neq 0.

  • 3)3)

    The finite matrix of Hurwitz type k(p,q)\mathcal{H}_{k}(p,q) defined in (1.16)–(1.17) is nonsingular and totally nonnegative. Here k=2nk=2n if degq<degp\deg q<\deg p, and k=2n+1k=2n+1 whenever degq=degp\deg q=\deg p.

The total nonnegativity of the matrices of Hurwitz matrix also implies the following curious result [16].

Theorem 1.34.

Let the polynomials pp and qq defined in (1.2)–(1.3) have only nonpositive zeroes, and let the function R=q/pR=q/p be an R-function of negative type. Given any two positive integers777The number rr is choosen such that b02+br20b_{0}^{2}+b_{r}^{2}\neq 0. rr and ll such that rln<(l+1)rrl\leqslant n<(l+1)r, the polynomials

pr,l(z)\displaystyle p_{r,l}(z) =\displaystyle= a0zl+arzl1+a2rzl2++arl,\displaystyle a_{0}z^{l}+a_{r}z^{l-1}+a_{2r}z^{l-2}+\ldots+a_{rl},
qr,l(z)\displaystyle q_{r,l}(z) =\displaystyle= b0zl+brzl1+b2rzl2++brl\displaystyle b_{0}z^{l}+b_{r}z^{l-1}+b_{2r}z^{l-2}+\ldots+b_{rl}

have only negative zeroes, and the function Rr,l=qr,l/pr,lR_{r,l}=q_{r,l}/p_{r,l} is an R-function.

1.4 Stieltjes continued fractions

Let a real rational function FF finite at infinity be expanded into its Laurant series at infinity:

F(z)=s1+s0z+s1z2+s2z3+F(z)=s_{-1}+\dfrac{s_{0}}{z}+\dfrac{s_{1}}{z^{2}}+\dfrac{s_{2}}{z^{3}}+\cdots (1.32)
Definition 1.35.

The function FF is said to have Stieltjes continued fraction expansion if FF can be represented in the form

F(z)=c0+1c1z+1c2+1c3z+1+1T,cj0,whereT={c2rif|F(0)|<,c2r1zifF(0)=.F(z)=c_{0}+\dfrac{1}{c_{1}z+\cfrac{1}{c_{2}+\cfrac{1}{c_{3}z+\cfrac{1}{\ddots+\cfrac{1}{T}}}}},\quad c_{j}\neq 0,\quad\text{where}\quad T=\begin{cases}&c_{2r}\qquad\ \,\text{if}\ |F(0)|<\infty,\\ &c_{2r-1}z\quad\text{if}\ F(0)=\infty.\end{cases} (1.33)

Here rr is the number of poles of FF, counting multiplicity.

There exists the following criterion for a rational function to have a Stieltjes continued fraction expansion [41] (see also [16]).

Theorem 1.36.

Suppose that a rational function FF defined in (1.32) has exactly rr poles. The function FF has a Stieltjes continued fraction expansion (1.33) if and only if it satisfies the conditions

Dj(F)0,j=1,2,,r,D_{j}(F)\neq 0,\quad j=1,2,\ldots,r, (1.34)
D^j(F)0,j=1,2,,r1,\widehat{D}_{j}(F)\neq 0,\quad j=1,2,\ldots,r-1, (1.35)
Dj(F)=D^j(F)=0,j=r+1,r+2,,D_{j}(F)=\widehat{D}_{j}(F)=0,\quad j=r+1,r+2,\ldots, (1.36)

where Dj(F)D_{j}(F) and D^j(F)\widehat{D}_{j}(F) are defined in (1.5) and (1.8), respectively.

The coefficients of the continued fraction (1.33) can be found by the formulæ

c2j=Dj2(F)D^j1(F)D^j(F),j=1,2,,r,c_{2j}=-\dfrac{D_{j}^{2}(F)}{\widehat{D}_{j-1}(F)\cdot\widehat{D}_{j}(F)},\quad j=1,2,\ldots,r, (1.37)
c2j1=D^j12(F)Dj1(F)Dj(F),j=1,2,,r,c_{2j-1}=\dfrac{\widehat{D}_{j-1}^{2}(F)}{D_{j-1}(F)\cdot D_{j}(F)},\quad j=1,2,\ldots,r, (1.38)

where D0(F)=D^0(F)1D_{0}(F)=\widehat{D}_{0}(F)\equiv 1.

We note that the determinant D^r(F)\widehat{D}_{r}(F) in Theorem 1.36 can be equal zero. But in this case (and only in this case), the function FF has a pole at 0 according to Corollary 1.5, and therefore T=c2r1zT=c_{2r-1}z in (1.33).

Let again the polynomials pp and qq be defined in (1.2)–(1.3). Consider the function R=qpR=\dfrac{q}{p} and suppose that RR has a Stieltjes continued fraction expansion (1.33), where rn=degpr\leqslant n=\deg p. Then from (1.18)–(1.21) and (1.37)–(1.38) we obtain:

  • if degq(z)<degp(z)\deg q(z)<\deg p(z), then

    c2j1\displaystyle c_{2j-1} =\displaystyle= η2j12(p,q)η2j2(p,q)η2j(p,q),j=1,2,,r;\displaystyle\dfrac{\eta_{2j-1}^{2}(p,q)}{\eta_{2j-2}(p,q)\cdot\eta_{2j}(p,q)},\qquad j=1,2,\ldots,r; (1.39)
    c2j\displaystyle c_{2j} =\displaystyle= η2j2(p,q)η2j1(p,q)η2j(p,q),j=1,2,,[k2];\displaystyle\dfrac{\eta_{2j}^{2}(p,q)}{\eta_{2j-1}(p,q)\cdot\eta_{2j}(p,q)},\qquad j=1,2,\ldots,\left[\dfrac{k}{2}\right]; (1.40)
  • if degq(z)=degp(z)\deg q(z)=\deg p(z), then

    c2j1\displaystyle c_{2j-1} =\displaystyle= η2j2(p,q)η2j1(p,q)η2j+1(p,q),j=1,2,,r;\displaystyle\dfrac{\eta_{2j}^{2}(p,q)}{\eta_{2j-1}(p,q)\cdot\eta_{2j+1}(p,q)},\qquad j=1,2,\ldots,r; (1.41)
    c2j\displaystyle c_{2j} =\displaystyle= η2j+12(p,q)η2j(p,q)η2j+2(p,q),j=0,1,2,,[k2];\displaystyle\dfrac{\eta_{2j+1}^{2}(p,q)}{\eta_{2j}(p,q)\cdot\eta_{2j+2}(p,q)},\qquad\quad j=0,1,2,\ldots,\left[\dfrac{k}{2}\right]; (1.42)

where k=2r1k=2r-1 if R(0)=R(0)=\infty, k=2rk=2r if |R(0)|<|R(0)|<\infty, and ηj(p,q)\eta_{j}(p,q) are the leading principal minors of the infinite Hurwitz matrix H(p,q)H(p,q) (see Definition 1.13). Here we set η0(p,q)1\eta_{0}(p,q)\equiv 1, and [ρ][\rho] denotes the largest integer not exceeding ρ\rho.

From Theorems 1.181.24 and 1.36 and from the formulæ (1.37)–(1.38) one obtain the following theorem [16].

Theorem 1.37.

Let the function RR with exactly rr poles, counting multiplicities, be defined (1.1). If RR has a Stieltjes continued fraction expansion (1.33), then RR is an R-function if and only if

c2j1>0,j=1,2,,r.c_{2j-1}>0,\quad j=1,2,\ldots,r. (1.43)

Moreover, the number of negative poles of the function RR equals the number of positive coefficients c2jc_{2j}, j=1,2,,kj=1,2,\ldots,k, where k=rk=r, if |R(0)|<|R(0)|<\infty, and k=r1k=r-1, if R(0)=R(0)=\infty.

Note that every R-function with poles of the same sign always has a Stieltjes continued fraction expansion.

Corollary 1.38 (A. Markov, Stieltjes, [27, 38, 39]).

A real rational function RR with exactly rr poles, counting multiplicities, is an R-function with all nonpositive poles if and only if RR has a Stieltjes continued fraction expansion (1.33), where

ci>0,i=1,2,,2r1.c_{i}>0,\quad i=1,2,\ldots,2r-1.

Moreover, if |R(0)|<|R(0)|<\infty, then c2r>0c_{2r}>0.

Corollary 1.39 ([16]).

A real rational function RR with exactly rr poles, counting multiplicities, is an R-function with all nonnegative poles if and only if RR has a Stieltjes continued fraction expansion (1.33), where

(1)i1ci>0,i=1,2,,2r1,(-1)^{i-1}c_{i}>0,\quad i=1,2,\ldots,2r-1,

and if |R(0)|<|R(0)|<\infty, then c2r<0c_{2r}<0.

2 Even and odd parts of polynomials. Associated function

Consider a real polynomial

p(z)=a0zn+a1zn1++an,a1,,an,a0>0.p(z)=a_{0}z^{n}+a_{1}z^{n-1}+\dots+a_{n},\qquad a_{1},\dots,a_{n}\in\mathbb{R},\ a_{0}>0. (2.1)

In the rest of the paper we use the following notation

l=[n2],l=\left[\dfrac{n}{2}\right], (2.2)

where n=degpn=\deg p, and [ρ][\rho] denotes the largest integer not exceeding ρ\rho.

The polynomial pp can always be represented as follows

p(z)=p0(z2)+zp1(z2),p(z)=p_{0}(z^{2})+zp_{1}(z^{2}), (2.3)

where

for n=2ln=2l,

p0(u)=a0ul+a2ul1++an,p1(u)=a1ul1+a3ul2++an1,\begin{split}&p_{0}(u)=a_{0}u^{l}+a_{2}u^{l-1}+\ldots+a_{n},\\ &p_{1}(u)=a_{1}u^{l-1}+a_{3}u^{l-2}+\ldots+a_{n-1},\end{split} (2.4)

and for n=2l+1n=2l+1,

p0(u)=a1ul+a3ul1++an,p1(u)=a0ul+a2ul1++an1.\begin{split}&p_{0}(u)=a_{1}u^{l}+a_{3}u^{l-1}+\ldots+a_{n},\\ &p_{1}(u)=a_{0}u^{l}+a_{2}u^{l-1}+\ldots+a_{n-1}.\end{split} (2.5)

The polynomials p0(z2)p_{0}(z^{2}) and p1(z2)p_{1}(z^{2}) satisfy the following equalities:

p0(z2)=p(z)+p(z)2,p1(z2)=p(z)p(z)2z.\begin{split}&\displaystyle p_{0}(z^{2})=\frac{p(z)+p(-z)}{2},\\ &\displaystyle p_{1}(z^{2})=\frac{p(z)-p(-z)}{2z}.\end{split} (2.6)

Introduce the following function888In the book [12, Chapter XV], F. Gantmacher used the function p1(u)p0(u)-\dfrac{p_{1}(-u)}{p_{0}(-u)}.:

Φ(u)=p1(u)p0(u).\Phi(u)=\displaystyle\frac{p_{1}(u)}{p_{0}(u)}. (2.7)
Definition 2.1.

We call Φ\Phi the function associated with the polynomial pp.

From (2.6) and (2.7) one can derive the following relations:

zΦ(z2)=p(z)p(z)p(z)+p(z)=1p(z)p(z)1+p(z)p(z),z\Phi(z^{2})=\displaystyle\frac{p(z)-p(-z)}{p(z)+p(-z)}=\displaystyle\frac{1-\displaystyle\frac{p(-z)}{p(z)}}{1+\displaystyle\frac{p(-z)}{p(z)}}, (2.8)
p(z)p(z)=1zΦ(z2)1+zΦ(z2).\frac{p(-z)}{p(z)}=\displaystyle\frac{1-z\Phi(z^{2})}{1+z\Phi(z^{2})}\,.

Now let us introduce two Hurwitz matrices associated with the polynomial pp.

Definition 2.2.

The matrix

H(p)=(a0a2a4a6a8a100a1a3a5a7a90a0a2a4a6a800a1a3a5a7)H_{\infty}(p)=\begin{pmatrix}a_{0}&a_{2}&a_{4}&a_{6}&a_{8}&a_{10}&\dots\\ 0&a_{1}&a_{3}&a_{5}&a_{7}&a_{9}&\dots\\ 0&a_{0}&a_{2}&a_{4}&a_{6}&a_{8}&\dots\\ 0&0&a_{1}&a_{3}&a_{5}&a_{7}&\dots\\ \vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\ddots\end{pmatrix} (2.9)

is called the infinite Hurwitz matrix of the polynomial pp. The leading principal minors of the matrix H(p)H_{\infty}(p) will be denoted by ηj(p)\eta_{j}(p), j=1,2,j=1,2,\ldots

Remark 2.3.

According to Definitions 1.13 and 2.2, we have H(p)=H(p0,p1)H_{\infty}(p)=H(p_{0},p_{1}), where p0p_{0} and p1p_{1} are the even and odd parts of the polynomial pp, respectively.

Together with the infinite matrix H(p)H_{\infty}(p), we consider its specific finite submatrix.

Definition 2.4.

The n×nn\times n matrix

n(p)=(a1a3a5a700a0a2a4a6000a1a3a5000a0a2a4000000an100000an2an)\mathcal{H}_{n}(p)=\begin{pmatrix}a_{1}&a_{3}&a_{5}&a_{7}&\dots&0&0\\ a_{0}&a_{2}&a_{4}&a_{6}&\dots&0&0\\ 0&a_{1}&a_{3}&a_{5}&\dots&0&0\\ 0&a_{0}&a_{2}&a_{4}&\dots&0&0\\ \vdots&\vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&0&\dots&a_{n-1}&0\\ 0&0&0&0&\dots&a_{n-2}&a_{n}\end{pmatrix} (2.10)

is called the finite Hurwitz matrix or the Hurwitz matrix.

The leading principal minors of this matrix we denote by Δj(p)\Delta_{j}(p):

Δj(p)=|a1a3a5a7a2j1a0a2a4a6a2j20a1a3a5a2j30a0a2a4a2j40000aj|,j=1,,n,\Delta_{j}(p)=\begin{vmatrix}a_{1}&a_{3}&a_{5}&a_{7}&\dots&a_{2j-1}\\ a_{0}&a_{2}&a_{4}&a_{6}&\dots&a_{2j-2}\\ 0&a_{1}&a_{3}&a_{5}&\dots&a_{2j-3}\\ 0&a_{0}&a_{2}&a_{4}&\dots&a_{2j-4}\\ \vdots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&0&\dots&a_{j}\end{vmatrix},\quad j=1,\ldots,n, (2.11)

where we set ai0a_{i}\equiv 0 for i>ni>n.

Definition 2.5.

The determinants Δj(p)\Delta_{j}(p), j=1,,nj=1,\ldots,n, are called the Hurwitz determinants or the Hurwitz minors of the polynomial pp.

Remark 2.6.

From Definitions 1.15 and 2.4 it follows that Hn(p)=H2l(p0,p1)H_{n}(p)=H_{2l}(p_{0},p_{1}) if n=2ln=2l, and Hn(p)=H2l+1(p0,p1)H_{n}(p)=H_{2l+1}(p_{0},p_{1}) if n=2l+1n=2l+1, where p0p_{0} and p1p_{1} are defined in (2.3)–(2.4).

Obviously,

ηj(p)=a0Δj1(p),j=1,,n,\eta_{j}(p)=a_{0}\Delta_{j-1}(p),\quad j=1,\dots,n, (2.12)

where Δ0(p)1\Delta_{0}(p)\equiv 1.

Suppose that degp0degp1\deg p_{0}\geqslant\deg p_{1} and expand the function Φ\Phi into its Laurent series at infinity:

Φ(u)=p1(u)p0(u)=s1+s0u+s1u2+s2u3+s3u4+,\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}=s_{-1}+\frac{s_{0}}{u}+\frac{s_{1}}{u^{2}}+\frac{s_{2}}{u^{3}}+\frac{s_{3}}{u^{4}}+\dots, (2.13)

where s10s_{-1}\neq 0 if degp0=degp1\deg p_{0}=\deg p_{1}, and s1=0s_{-1}=0 if degp0>degp1\deg p_{0}>\deg p_{1}.

From (1.18)–(1.23) we obtain the following relations [17, 12] between the determinants Dj(Φ)D_{j}(\Phi), D^j(Φ)\widehat{D}_{j}(\Phi) defined in (1.5) and (1.8), the Hurwitz minors Δj(p)\Delta_{j}(p), and the determinants ηj(p)\eta_{j}(p) (j=1,2,,n)(j=1,2,\dots,n).

  • 1)

    If n=2ln=2l, then

    Δ2j1(p)=a01η2j(p)=a02j1Dj(Φ),Δ2j(p)=a01η2j+1(p)=(1)ja02jD^j(Φ),j=1,2,,l;\begin{split}&\Delta_{2j-1}(p)=a_{0}^{-1}\eta_{2j}(p)=a_{0}^{2j-1}D_{j}(\Phi),\\ &\Delta_{2j}(p)=a_{0}^{-1}\eta_{2j+1}(p)=(-1)^{j}a_{0}^{2j}\widehat{D}_{j}(\Phi),\end{split}\qquad j=1,2,\dots,l; (2.14)
  • 2)

    If n=2l+1n=2l+1, then

    Δ2j(p)=a01η2j+1(p)=(a0s1)2jDj(Φ),j=1,2,,l;Δ2j+1(p)=a01η2j+2(p)=(1)j(a0s1)2j+1D^j(Φ),j=0,1,,l,\begin{split}&\Delta_{2j}(p)=a_{0}^{-1}\eta_{2j+1}(p)={\left(\frac{a_{0}}{s_{-1}}\right)}^{2j}D_{j}(\Phi),\qquad j=1,2,\dots,l;\\ &\Delta_{2j+1}(p)=a_{0}^{-1}\eta_{2j+2}(p)=(-1)^{j}{\left(\frac{a_{0}}{s_{-1}}\right)}^{2j+1}\widehat{D}_{j}(\Phi),\qquad j=0,1,\dots,l,\end{split} (2.15)

    where D^0(Φ)1\widehat{D}_{0}(\Phi)\equiv 1.

By (2.3)–(2.5) and (2.7) and by Theorem 1.2 and Corollary 1.4, we have

Dj(Φ)=D^j(Φ)=0,j>l,D_{j}(\Phi)=\widehat{D}_{j}(\Phi)=0,\quad j>l, (2.16)

where ll is defined in (2.2). So in the sequel, we deal only with the determinants Dj(Φ)D_{j}(\Phi), D^j(Φ)\widehat{D}_{j}(\Phi) of order at most ll.

Also in the sequel, we deal only with the determinants ηj(p)\eta_{j}(p) of order at most n+1n+1, since by the formulæ (1.18)–(1.21) and (2.16) and by Remark 2.3, we have

ηj(p)=0,forj>n+1.\eta_{j}(p)=0,\quad\text{for}\quad j>n+1. (2.17)

In Section 5 we also consider the case when n=2l+1n=2l+1 with a1=0a_{1}=0 and a30a_{3}\neq 0. In this case, we have degp0=degp11=l1\deg p_{0}=\deg p_{1}-1=l-1, so the function Φ\Phi has the form

Φ(u)=p1(u)p0(u)=s2u+s1+s0u+s1u2+s2u3+s3u4+=a0ul+a2ul1++a2la3ul1+a5ul2++a2l+1.\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}=s_{-2}u+s_{-1}+\frac{s_{0}}{u}+\frac{s_{1}}{u^{2}}+\frac{s_{2}}{u^{3}}+\frac{s_{3}}{u^{4}}+\dots=\dfrac{a_{0}u^{l}+a_{2}u^{l-1}+\ldots+a_{2l}}{a_{3}u^{l-1}+a_{5}u^{l-2}+\ldots+a_{2l+1}}. (2.18)

It is easy to see that t2=a0a3t_{-2}=\dfrac{a_{0}}{a_{3}}, so we can represent Φ\Phi as a sum Φ(u)=a0a3u+p2(u)p0(u)\Phi(u)=\dfrac{a_{0}}{a_{3}}\,u+\dfrac{p_{2}(u)}{p_{0}(u)}, where

p2(u)=(a2a0a3a5)ul1+(a4a0a3a7)ul2++(a2l2a0a3a2l+1)u+a2l.p_{2}(u)=\left(a_{2}-\dfrac{a_{0}}{a_{3}}\,a_{5}\right)u^{l-1}+\left(a_{4}-\dfrac{a_{0}}{a_{3}}\,a_{7}\right)u^{l-2}+\ldots+\left(a_{2l-2}-\dfrac{a_{0}}{a_{3}}\,a_{2l+1}\right)u+a_{2l}.

Note that for the function

Θ(u):=p2(u)p0(u)=Φ(u)a0a3u=s1+s0u+s1u2+s2u3+s3u4+,\Theta(u):=\dfrac{p_{2}(u)}{p_{0}(u)}=\Phi(u)-\dfrac{a_{0}}{a_{3}}\,u=s_{-1}+\frac{s_{0}}{u}+\frac{s_{1}}{u^{2}}+\frac{s_{2}}{u^{3}}+\frac{s_{3}}{u^{4}}+\dots,

the following equalities hold

Dj(Θ)=Dj(Φ),andD^j(Θ)=D^j(Φ)forj=1,2,D_{j}(\Theta)=D_{j}(\Phi),\quad\text{and}\quad\widehat{D}_{j}(\Theta)=\widehat{D}_{j}(\Phi)\quad\text{for}\quad j=1,2,\ldots (2.19)

By Definition 1.15 and by the formulæ (1.20), and (1.23) applied to the polynomials p0p_{0} and p2p_{2}, we have, for j=1,2,j=1,2,\ldots,

a0a32j+1Dj(Θ)=a0a32j(p0,p2)=a0a3|a3a5a4j+1a2a0a3a5a4a0a3a7a4ja0a3a4j+30a3a4j10a2a0a3a5a4j2a0a3a4j+100a2j+2a0a3a2j+5|==|0a3a5a7a4j+3a0a2a4a6a4j+200a3a5a4j+100a2a0a3a5a4a0a3a7a4ja0a3a4j+3000a3a4j1000a2a0a3a5a4j2a0a3a4j+10000a2j+2a0a3a2j+5|==|0a3a5a7a4j+3a0a2a4a6a4j+200a3a5a4j+10a0a2a4a4j000a3a4j100a0a2a4j20000a2j+2|=Δ2j+2(p)=a0a32j+1Dj(Φ),\begin{array}[]{c}-a_{0}a_{3}^{2j+1}D_{j}(\Theta)=-a_{0}a_{3}\nabla_{2j}(p_{0},p_{2})=-a_{0}a_{3}\begin{vmatrix}a_{3}&a_{5}&\dots&a_{4j+1}\\ a_{2}-\dfrac{a_{0}}{a_{3}}\,a_{5}&a_{4}-\dfrac{a_{0}}{a_{3}}\,a_{7}&\dots&a_{4j}-\dfrac{a_{0}}{a_{3}}\,a_{4j+3}\\ 0&a_{3}&\dots&a_{4j-1}\\ 0&a_{2}-\dfrac{a_{0}}{a_{3}}\,a_{5}&\dots&a_{4j-2}-\dfrac{a_{0}}{a_{3}}\,a_{4j+1}\\ \vdots&\vdots&\ddots&\vdots\\ 0&0&\dots&a_{2j+2}-\dfrac{a_{0}}{a_{3}}\,a_{2j+5}\end{vmatrix}=\\ =\begin{vmatrix}0&a_{3}&a_{5}&a_{7}&\dots&a_{4j+3}\\ a_{0}&a_{2}&a_{4}&a_{6}&\dots&a_{4j+2}\\ 0&0&a_{3}&a_{5}&\dots&a_{4j+1}\\ 0&0&a_{2}-\dfrac{a_{0}}{a_{3}}\,a_{5}&a_{4}-\dfrac{a_{0}}{a_{3}}\,a_{7}&\dots&a_{4j}-\dfrac{a_{0}}{a_{3}}\,a_{4j+3}\\ 0&0&0&a_{3}&\dots&a_{4j-1}\\ 0&0&0&a_{2}-\dfrac{a_{0}}{a_{3}}\,a_{5}&\dots&a_{4j-2}-\dfrac{a_{0}}{a_{3}}\,a_{4j+1}\\ \vdots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&0&\dots&a_{2j+2}-\dfrac{a_{0}}{a_{3}}\,a_{2j+5}\end{vmatrix}=\\ =\begin{vmatrix}0&a_{3}&a_{5}&a_{7}&\dots&a_{4j+3}\\ a_{0}&a_{2}&a_{4}&a_{6}&\dots&a_{4j+2}\\ 0&0&a_{3}&a_{5}&\dots&a_{4j+1}\\ 0&a_{0}&a_{2}&a_{4}&\dots&a_{4j}\\ 0&0&0&a_{3}&\dots&a_{4j-1}\\ 0&0&a_{0}&a_{2}&\dots&a_{4j-2}\\ \vdots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&0&\dots&a_{2j+2}\end{vmatrix}=\Delta_{2j+2}(p)=-a_{0}a_{3}^{2j+1}D_{j}(\Phi),\end{array}

where we set ai0a_{i}\equiv 0 for i>2l+1i>2l+1.

Denote the coefficients of the polynomial p2p_{2} by bjb_{j}:

bj=a2j+2a0a3a2j+5,j=0,1,,l1,b_{j}=a_{2j+2}-\dfrac{a_{0}}{a_{3}}a_{2j+5},\qquad j=0,1,\ldots,l-1, (2.20)

where a2l+30a_{2l+3}\equiv 0. Then the function uΦ(u)u\Phi(u) has the form

uΦ(u)=a0a3u2+b0a3u+p3(u)p0(u),u\Phi(u)=\dfrac{a_{0}}{a_{3}}\,u^{2}+\dfrac{b_{0}}{a_{3}}\,u+\dfrac{p_{3}(u)}{p_{0}(u)},

where

p3(u):=up2(u)b0a3up0(u)=(b1b0a3a5)ul1+(b2b0a3a7)ul2++(bl1b0a3a2l+1)u.p_{3}(u):=up_{2}(u)-\dfrac{b_{0}}{a_{3}}\,up_{0}(u)=\left(b_{1}-\dfrac{b_{0}}{a_{3}}\,a_{5}\right)u^{l-1}+\left(b_{2}-\dfrac{b_{0}}{a_{3}}\,a_{7}\right)u^{l-2}+\ldots+\left(b_{l-1}-\dfrac{b_{0}}{a_{3}}\,a_{2l+1}\right)u.

By Definition 1.15 and by the formulæ (2.19), (1.20) and (1.23) applied to the polynomials p0p_{0} and p3p_{3} we get (putting bi0b_{i}\equiv 0 for i>l1i>l-1):

a0a32j+2D^j(Θ)=a0a32(p0,p3)=a0a32|a3a5a4j+1b1b0a3a5b2b0a3a7b2jb0a3a4j+30a3a4j10b1b0a3a5b2j1b0a3a4j+100bj+1b0a3a2j+5|==a0a3|a3a5a7a4j+30a3a5a4j+10b1b0a3a5b2b0a3a7b2jb0a3a4j+300a3a4j100b1b0a3a5b2j1b0a3a4j+1000bj+1b0a3a2j+5|=(1)ja0a3|a3a5a7a4j+3b0b1b2b2j0a3a5a4j+10b0b1b2j100a3a4j100b0b2j2000a2j+3|==(1)j+1|0a3a5a7a9a4j+5a0a2a4a6a8a4j+400a3a5a7a4j+300a2a0a3a5a4a0a3a7a6a0a3a9a4j+2a0a3a4j+5000a3a5a4j+1000a2a0a3a5a4a0a3a7a4ja0a3a4j+30000a3a4j10000a2a0a3a5a4j2a0a3a4j+100000a2j+3|==(1)j+1Δ2j+3(p)=a0a32j+2D^j(Φ),\begin{array}[]{c}a_{0}a_{3}^{2j+2}\widehat{D}_{j}(\Theta)=a_{0}a_{3}^{2}\nabla(p_{0},p_{3})=a_{0}a_{3}^{2}\begin{vmatrix}a_{3}&a_{5}&\dots&a_{4j+1}\\ b_{1}-\dfrac{b_{0}}{a_{3}}\,a_{5}&b_{2}-\dfrac{b_{0}}{a_{3}}\,a_{7}&\dots&b_{2j}-\dfrac{b_{0}}{a_{3}}\,a_{4j+3}\\ 0&a_{3}&\dots&a_{4j-1}\\ 0&b_{1}-\dfrac{b_{0}}{a_{3}}\,a_{5}&\dots&b_{2j-1}-\dfrac{b_{0}}{a_{3}}\,a_{4j+1}\\ \vdots&\vdots&\ddots&\vdots\\ 0&0&\dots&b_{j+1}-\dfrac{b_{0}}{a_{3}}\,a_{2j+5}\end{vmatrix}=\\ =a_{0}a_{3}\begin{vmatrix}a_{3}&a_{5}&a_{7}&\dots&a_{4j+3}\\ 0&a_{3}&a_{5}&\dots&a_{4j+1}\\ 0&b_{1}-\dfrac{b_{0}}{a_{3}}\,a_{5}&b_{2}-\dfrac{b_{0}}{a_{3}}\,a_{7}&\dots&b_{2j}-\dfrac{b_{0}}{a_{3}}\,a_{4j+3}\\ 0&0&a_{3}&\dots&a_{4j-1}\\ 0&0&b_{1}-\dfrac{b_{0}}{a_{3}}\,a_{5}&\dots&b_{2j-1}-\dfrac{b_{0}}{a_{3}}\,a_{4j+1}\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\dots&b_{j+1}-\dfrac{b_{0}}{a_{3}}\,a_{2j+5}\end{vmatrix}=(-1)^{j}a_{0}a_{3}\begin{vmatrix}a_{3}&a_{5}&a_{7}&\dots&a_{4j+3}\\ b_{0}&b_{1}&b_{2}&\dots&b_{2j}\\ 0&a_{3}&a_{5}&\dots&a_{4j+1}\\ 0&b_{0}&b_{1}&\dots&b_{2j-1}\\ 0&0&a_{3}&\dots&a_{4j-1}\\ 0&0&b_{0}&\dots&b_{2j-2}\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\dots&a_{2j+3}\end{vmatrix}=\\ =(-1)^{j+1}\begin{vmatrix}0&a_{3}&a_{5}&a_{7}&a_{9}&\dots&a_{4j+5}\\ a_{0}&a_{2}&a_{4}&a_{6}&a_{8}&\dots&a_{4j+4}\\ 0&0&a_{3}&a_{5}&a_{7}&\dots&a_{4j+3}\\ 0&0&a_{2}-\dfrac{a_{0}}{a_{3}}\,a_{5}&a_{4}-\dfrac{a_{0}}{a_{3}}\,a_{7}&a_{6}-\dfrac{a_{0}}{a_{3}}\,a_{9}&\dots&a_{4j+2}-\dfrac{a_{0}}{a_{3}}\,a_{4j+5}\\ 0&0&0&a_{3}&a_{5}&\dots&a_{4j+1}\\ 0&0&0&a_{2}-\dfrac{a_{0}}{a_{3}}\,a_{5}&a_{4}-\dfrac{a_{0}}{a_{3}}\,a_{7}&\dots&a_{4j}-\dfrac{a_{0}}{a_{3}}\,a_{4j+3}\\ 0&0&0&0&a_{3}&\dots&a_{4j-1}\\ 0&0&0&0&a_{2}-\dfrac{a_{0}}{a_{3}}\,a_{5}&\dots&a_{4j-2}-\dfrac{a_{0}}{a_{3}}\,a_{4j+1}\\ \vdots&\vdots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&0&0&\dots&a_{2j+3}\end{vmatrix}=\\ \\ =(-1)^{j+1}\Delta_{2j+3}(p)=a_{0}a_{3}^{2j+2}\widehat{D}_{j}(\Phi),\end{array}

where we set ai0a_{i}\equiv 0 for i>2l+1i>2l+1. Here we used the equalities D^j(Θ)=Dj(uΘ)\widehat{D}_{j}(\Theta)=D_{j}(u\Theta), j=1,2,j=1,2,\ldots

Thus, we obtain the following formulæ

Δ2j(p)=a01η2j+1(p)=a0a32j1Dj1(Φ)=a32j2Δ2(p)Dj1(Φ),j=2,3,,l,Δ2j+1(p)=a01η2j+2(p)=(1)ja0a32jD^j1(Φ)=(1)j1a32j2Δ3(p)D^j1(Φ),j=2,3,,l,\begin{split}&\Delta_{2j}(p)=a_{0}^{-1}\eta_{2j+1}(p)=-a_{0}a_{3}^{2j-1}D_{j-1}(\Phi)=a_{3}^{2j-2}\Delta_{2}(p)D_{j-1}(\Phi),\quad\qquad\qquad\qquad j=2,3,\dots,l,\\ &\Delta_{2j+1}(p)=a_{0}^{-1}\eta_{2j+2}(p)=(-1)^{j}a_{0}a_{3}^{2j}\widehat{D}_{j-1}(\Phi)=(-1)^{j-1}a_{3}^{2j-2}\Delta_{3}(p)\widehat{D}_{j-1}(\Phi),\qquad j=2,3,\dots,l,\end{split} (2.21)

where ηi(p)\eta_{i}(p) are the leading principal minors of the matrix H(p)H_{\infty}(p) defined in (2.9). Here we used the formulæ

Δ2(p)=a01η3(p)=a0a3,Δ3(p)=a01η4(p)=a0a32,\Delta_{2}(p)=a_{0}^{-1}\eta_{3}(p)=-a_{0}a_{3},\quad\Delta_{3}(p)=a_{0}^{-1}\eta_{4}(p)=-a_{0}a_{3}^{2}, (2.22)

that follow from the equality a1=0a_{1}=0.

Note that, as above, the following holds:

Dj(Φ)=D^j(Φ)=0,j>l1.D_{j}(\Phi)=\widehat{D}_{j}(\Phi)=0,\quad j>l-1.

3 Stable polynomials

This section is devoted to some basic facts of the theory of Hurwitz stable and quasi-stable polynomials. We expound those facts from the viewpoint of the theory of R-functions.

3.1 Hurwitz polynomials

Definition 3.1 ([17, 12]).

The polynomial pp defined in (2.1) is called Hurwitz or Hurwitz stable if all its zeroes lie in the open left half-plane of the complex plane.

At first, we recall the following simple necessary condition for polynomials to be Hurwitz stable. This condition is called usually Stodola theorem [12, 6].

Theorem 3.2 (Stodola).

If the polynomial pp is Hurwitz stable, then all its coefficients are positive999More exactly, the coefficients must be of the same sign, but a0>0a_{0}>0 by (2.1)..

Now we show a connection between Hurwitz polynomials and R-functions. This connection will allow us to apply all statements from the Section 1 to Hurwitz polynomials to obtain such basic criteria of Hurwitz stability as Hurwitz criterion and Liénard and Chipart criterion (Theorems (3.5) and 3.7, respectively). We prove the following theorem using a method suggested by Yu.S. Barkovsky [5].

Theorem 3.3 ([12, 6]).

A real polynomial pp defined in (2.1) is Hurwitz stable if and only if its associated function Φ\Phi defined in (2.7) is an R-function with exactly ll poles, all of which are negative, and the limit limu±Φ(u)\displaystyle\lim_{u\to\pm\infty}\Phi(u) is positive whenever n=2l+1n=2l+1. The number ll is defined in (2.2).

Proof.

Let the polynomial pp be Hurwitz stable, that is,

p(λ)=0Reλ<0.p(\lambda)=0\ \ \Longrightarrow\ \ \operatorname{Re}\lambda<0. (3.1)

First, we show that

|p(z)p(z)|<1,z:Rez>0.\displaystyle\left|\frac{p(-z)}{p(z)}\right|<1,\quad\forall z:\,\operatorname{Re}{z}>0. (3.2)

Note that the polynomials p(z)p(z) and p(z)p(-z) have no common zeroes if pp is Hurwitz stable, so the function p(z)p(z)\dfrac{p(-z)}{p(z)} has exactly nn poles. The Hurwitz stable polynomial pp can be represented in the form

p(z)=a0k(zλk)j(zξj)(zξ¯j),p(z)=a_{0}\prod_{k}(z-\lambda_{k})\prod_{j}\left(z-\xi_{j}\right)\left(z-\overline{\xi}_{j}\right),

where λk<0,Reξj<0\lambda_{k}<0,\operatorname{Re}\xi_{j}<0 and Imξj0\operatorname{Im}\xi_{j}\neq 0. Then we have

|p(z)p(z)|=k|z+λk||zλk|j|z+ξj||z+ξ¯j||zξ¯j||zξj|.\displaystyle\left|\frac{p(-z)}{p(z)}\right|=\prod_{k}\frac{\left|z+\lambda_{k}\right|}{\left|z-\lambda_{k}\right|}\prod_{j}\frac{\left|z+\xi_{j}\right|\left|z+\overline{\xi}_{j}\right|}{\left|z-\overline{\xi}_{j}\right|\left|z-\xi_{j}\right|}. (3.3)

It is easy to see that the function of type z+aza¯\dfrac{z+a}{z-\overline{a}} , where Rea<0\operatorname{Re}a<0, maps the right half-plane of the complex plane to the unit disk. In fact,

|z+aza¯|2=(Rez+Rea)2+(Imz+Ima)2(RezRea)2+(Imz+Ima)2<1wheneverRez>0andRea<0.\left|\dfrac{z+a}{z-\overline{a}}\right|^{2}=\dfrac{(\operatorname{Re}z+\operatorname{Re}a)^{2}+(\operatorname{Im}z+\operatorname{Im}a)^{2}}{(\operatorname{Re}z-\operatorname{Re}a)^{2}+(\operatorname{Im}z+\operatorname{Im}a)^{2}}<1\qquad\text{whenever}\quad\operatorname{Re}z>0\quad\text{and}\quad\operatorname{Re}a<0.

Now from (3.3) it follows that the function p(z)p(z)\displaystyle\frac{p(-z)}{p(z)} also maps the right half-plane to the unit disk as a product of functions of such a type. Thus, the inequality (3.2) is valid.

At the same time, the fractional linear transformation z1z1+z\displaystyle z\mapsto\frac{1-z}{1+z} conformally maps the unit disk to the right half-plane:

|z|<1Re(1z1+z)=1|z|2|1+z|2>0.|z|<1\implies\operatorname{Re}\left(\frac{1-z}{1+z}\right)=\dfrac{1-|z|^{2}}{|1+z|^{2}}>0. (3.4)

Consequently, from the relations (2.8), (3.2) and (3.4) we obtain that the function zΦ(z2)z\Phi(z^{2}) maps the right half-plane to itself, so the function zΦ(z2)-z\Phi(-z^{2}) maps the upper half-plane of the complex plane to the lower half-plane:

Imz>0Re(iz)>0Re[izΦ(z2)]=Im[zΦ(z2)]>0Im[zΦ(z2)]<0.\operatorname{Im}z>0\implies\operatorname{Re}(-iz)>0\implies\operatorname{Re}\left[-iz\Phi(-z^{2})\right]=\operatorname{Im}\left[z\Phi(-z^{2})\right]>0\implies\operatorname{Im}\left[-z\Phi(-z^{2})\right]<0.

Since pp is Hurwitz stable by assumption, the polynomials p(z)p(z) and p(z)p(-z) have no common zeroes, therefore, p0p_{0} and p1p_{1} also have no common zeroes, and p0(0)0p_{0}(0)\neq 0 by (2.3). Moreover, by Theorem 3.2 we have a0>0a_{0}>0 and a1>0a_{1}>0, so degp0=l\deg p_{0}=l (see (2.4)–(2.5)). Thus, the number of poles of the function zΦ(z2)-z\Phi(-z^{2}) equals the number of zeroes of the polynomial p0(z2)p_{0}(-z^{2}), i.e. exactly 2l2l.

So according to Theorem 1.18, the function zΦ(z2)-z\Phi(-z^{2}) can be represented in the form (1.25), where all poles are located symmetrically with respect to 0 and β=0\beta=0, since zΦ(z2)-z\Phi(-z^{2}) is an odd function. Denote the poles of zΦ(z2)-z\Phi(-z^{2}) by ±ν1,,±νl\pm\nu_{1},\ldots,\pm\nu_{l} such that

0<ν1<ν2<<νl.0<\nu_{1}<\nu_{2}<\ldots<\nu_{l}.

Note that ν10\nu_{1}\neq 0, since p0(0)0p_{0}(0)\neq 0 as we mentioned above.

Thus, the function zΦ(z2)-z\Phi(-z^{2}) can be represented in the following form

zΦ(z2)=αz+j=1lγjzνj+j=1lγjz+νj=αz+j=1l2γjzz2νj2,α0,γj,νj>0.-z\Phi(-z^{2})=-\alpha z+\sum_{j=1}^{l}\frac{\gamma_{j}}{z-\nu_{j}}+\sum_{j=1}^{l}\frac{\gamma_{j}}{z+\nu_{j}}=-\alpha z+\sum_{j=1}^{l}\frac{2\gamma_{j}z}{z^{2}-\nu_{j}^{2}},\quad\alpha\geqslant 0,\,\gamma_{j},\nu_{j}>0.

Dividing this equality by z-z and changing variables as follows z2u-z^{2}\to u, 2γjβj2\gamma_{j}\to\beta_{j}, νj2ωj\nu_{j}^{2}\to\omega_{j}, we obtain the following representation of the function Φ\Phi:

Φ(u)=p1(u)p0(u)=α+j=1lβju+ωj,\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}=\alpha+\sum_{j=1}^{l}\frac{\beta_{j}}{u+\omega_{j}}, (3.5)

where α0,βj>0\alpha\geqslant 0,\,\beta_{j}>0 and

0<ω1<ω2<<ωl.0<\omega_{1}<\omega_{2}<\ldots<\omega_{l}.

Here α=0\alpha=0 whenever n=2ln=2l, and α=a0a1>0\alpha=\displaystyle\frac{a_{0}}{a_{1}}>0 whenever n=2l+1n=2l+1. Since Φ\Phi can be represented in the form (3.5), we have that by Theorem 1.18, Φ\Phi is an R-function with exactly ll poles, all which are negative, and limu±Φ(u)=α>0\displaystyle\lim_{u\to\pm\infty}\Phi(u)=\alpha>0 as n=2l+1n=2l+1.

Conversely, let the polynomial pp be defined in (2.1) and let its associated function Φ\Phi be an R-function with exactly ll poles, all of which are negative, and limu±Φ(u)>0\displaystyle\lim_{u\to\pm\infty}\Phi(u)>0 as n=2l+1n=2l+1. We will show that pp is Hurwitz stable.

By Theorem 1.18, Φ\Phi can be represented in the form (3.5), where α=limu±Φ(u)0\alpha=\displaystyle\lim_{u\to\pm\infty}\Phi(u)\geqslant 0 such that α>0\alpha>0 if n=2l+1n=2l+1, and α=0\alpha=0 if n=2ln=2l. Thus, the polynomial p0p_{0} has only negative zeroes, and the polynomials p0p_{0} and p1p_{1} have no common zeroes. Together with (2.3) and (2.7), this implies that the set of zeroes of the polynomial pp coincides with the set of roots of the equation

zΦ(z2)=1.z\Phi(z^{2})=-1. (3.6)

Let λ\lambda be a zero of the polynomial pp and therefore, of the equation (3.6). Then from (3.5) and (3.6) we obtain

1=Re[λΦ(λ2)]=[α+j=1lβj|λ|2+ωj|λ2+ωj|2]Reλ,-1=\operatorname{Re}\left[\lambda\Phi(\lambda^{2})\right]=\left[\alpha+\sum_{j=1}^{l}\beta_{j}\frac{|\lambda|^{2}+\omega_{j}}{|\lambda^{2}+\omega_{j}|^{2}}\right]\operatorname{Re}\lambda,

where α0\alpha\geqslant 0, and βj,ωj>0\beta_{j},\,\omega_{j}>0 for j=1,,lj=1,\ldots,l. Thus, if λ\lambda is a zero of pp, then Reλ<0\operatorname{Re}\lambda<0, so pp is Hurwitz stable. ∎

Consider the Hankel matrix S(Φ)=si+ji,j=0S(\Phi)=\|s_{i+j}\|^{\infty}_{i,j=0} constructed with the coefficients of the series (2.13) and its determinants Dj(Φ)D_{j}(\Phi) and D^j(Φ)\widehat{D}_{j}(\Phi) defined in (1.5) and (1.8), respectively. From Theorems 1.183.3 and 1.12 and from Corollary 1.26 we obtain the following Hurwitz stability criteria:

Theorem 3.4.

Let a real polynomial pp be defined by (2.1). The following conditions are equivalent:

  • 1)

    the polynomial pp is Hurwitz stable;

  • 2)

    the following hold

    s1>0forn=2l+1,Dj(Φ)>0,j=1,,l,(1)jD^j(Φ)>0,j=1,,l,\begin{split}&s_{-1}>0\quad\text{for}\quad n=2l+1,\\ &D_{j}(\Phi)>0,\qquad j=1,\ldots,l,\\ &(-1)^{j}\widehat{D}_{j}(\Phi)>0,\qquad j=1,\ldots,l,\end{split} (3.7)

    where l=[n2]l=\left[\dfrac{n}{2}\right];

  • 3)

    the matrix S(Φ)S(\Phi) is sign regular of rank ll, and s1>0s_{-1}>0 for n=2l+1n=2l+1.

Proof.

In fact, by Theorem 3.3, the polynomial pp is Hurwitz stable if and only if its associated function Φ\Phi is an R-function with exactly ll poles, all of which are negative and limu±Φ(u)=s10\lim\limits_{u\to\pm\infty}\Phi(u)=s_{-1}\geqslant 0. According to Theorem 1.18 and Corollary 1.26, this is equivalent to the inequalities (3.7). But these inequalities also are equivalent to the sign regularity of the matrix S(Φ)S(\Phi) by Theorem 1.12. ∎

Our next theorem provides stability criteria in terms of coefficients of the polynomial pp.

Theorem 3.5.

Given a real polynomial pp of degree nn as in (2.1), the following conditions are equivalent:

  • 1)

    the polynomial pp is Hurwitz stable;

  • 2)

    all Hurwitz determinants Δj(p)\Delta_{j}(p) are positive:

    Δ1(p)>0,Δ2(p)>0,,Δn(p)>0;\Delta_{1}(p)>0,\ \Delta_{2}(p)>0,\dots,\ \Delta_{n}(p)>0; (3.8)
  • 3)

    the determinants ηj(p)\eta_{j}(p) are positive up to order n+1n+1:

    η1(p)>0,η2(p)>0,,ηn+1(p)>0;\eta_{1}(p)>0,\ \eta_{2}(p)>0,\dots,\ \eta_{n+1}(p)>0; (3.9)
  • 4)

    the matrix n(p)\mathcal{H}_{n}(p) defined in (2.10) is nonsingular and totally nonnegative;

  • 5)

    the matrix H(p)H_{\infty}(p) defined in (2.9) is totally nonnegative with nonzero minor ηn+1(p)\eta_{n+1}(p).

Note that the equivalence of 1)1) and 2)2) is the famous Hurwitz criterion of stability. The implications 1)4)1)\Longrightarrow 4) and 1)5)1)\Longrightarrow 5) were proved in [3, 20]. The implication 4)1)4)\Longrightarrow 1) was, in fact, proved in [3]. However, the implication 5)1)5)\Longrightarrow 1) is probably new.

Proof.

Indeed, by Theorem 3.4, the polynomial pp is Hurwitz stable if and only if the inequalities (3.7) hold. According to the formulæ (2.14)–(2.15), these inequalities are equivalent to (3.8). By (2.12), the inequalities (3.8) are equivalent to (3.9), since η1(p)=a0>0\eta_{1}(p)=a_{0}>0 (see (1.14)–(1.15) and Remark 2.3).

Furthermore, by Theorem 3.3, the polynomial pp is Hurwitz stable if and only if its associated function Φ\Phi is an R-function with exactly l=[n2]l=\left[\dfrac{n}{2}\right], all of which are negative, and limu±Φ(u)=a0a1=s1>0\lim\limits_{u\to\pm\infty}\Phi(u)=\dfrac{a_{0}}{a_{1}}=s_{-1}>0 if n=2l+1n=2l+1. According to Corollary 1.33 and Remark 2.3, this is equivalent to the total nonnegativity of the matrix H(p)H_{\infty}(p). Moreover, ηn+1(p)0\eta_{n+1}(p)\neq 0 by Corollary 1.33. Thus, the condition 1)1) is equivalent to the condition 5)5).

The conditions 1)1) and 4)4) are equivalent also by Corollary 1.33 and Remark 2.6. ∎

Remark 3.6.

Note that total nonnegativity of the Hurwitz matrix of a Hurwitz polynomial was established in [3, 20, 14].

Now we are in a position to prove a few another famous criteria of Hurwitz stability, which are known as the Liénard and Chipart criterion and its modifications (see [25, 12]). These criteria are simple consequences of Theorems 1.301.31 and 3.3.

Theorem 3.7.

The polynomial pp given by (2.1) is Hurwitz stable if and only if one of the following conditions holds

  • 1)1)

    an>0,an2>0,an4>0,,Δn1(p)>0,Δn3(p)>0,Δn5(p)>0,a_{n}>0,a_{n-2}>0,a_{n-4}>0,\ldots,\qquad\Delta_{n-1}(p)>0,\Delta_{n-3}(p)>0,\Delta_{n-5}(p)>0,\ldots;

  • 2)2)

    an>0,an1>0,an3>0,,Δn1(p)>0,Δn3(p)>0,Δn5(p)>0,a_{n}>0,a_{n-1}>0,a_{n-3}>0,\ldots,\qquad\Delta_{n-1}(p)>0,\Delta_{n-3}(p)>0,\Delta_{n-5}(p)>0,\ldots;

  • 3)3)

    an>0,an2>0,an4>0,,Δn(p)>0,Δn2(p)>0,Δn4(p)>0,a_{n}>0,a_{n-2}>0,a_{n-4}>0,\ldots,\qquad\Delta_{n}(p)>0,\Delta_{n-2}(p)>0,\Delta_{n-4}(p)>0,\ldots;

  • 4)4)

    an>0,an1>0,an3>0,,Δn(p)>0,Δn2(p)>0,Δn4(p)>0,a_{n}>0,a_{n-1}>0,a_{n-3}>0,\ldots,\qquad\Delta_{n}(p)>0,\Delta_{n-2}(p)>0,\Delta_{n-4}(p)>0,\ldots

Proof.

Let n=2ln=2l. By Definitions 1.15 and 2.4 and by Remark 2.6, we have Hn(p)=H2l(p0,p1)H_{n}(p)=H_{2l}(p_{0},p_{1}), where the polynomials p0p_{0} and p1p_{1} are defined in (2.3)–(2.4). Obviously, Δj(p)=Δj(p0,p1)\Delta_{j}(p)=\Delta_{j}(p_{0},p_{1}), j=1,,nj=1,\ldots,n. The assertion of the theorem follows now from Theorems 3.3 and 1.30.

In the case n=2l+1n=2l+1, the assertion of the theorem can be proved in the same way using Theorem 1.31 instead of Theorem 1.30. ∎

We now recall a connection between Hurwitz polynomials and Stieltjes continued fractions (see [12]).

Theorem 3.8.

The polynomial pp of degree n1n\geqslant 1 defined in (2.1) is Hurwitz stable if and only if its associated function Φ\Phi has the following Stieltjes continued fraction expansion:

Φ(u)=c0+1c1u+1c2+1c3u+1+1c2l1u+1c2l,withci>0,i=1,,2l,\Phi(u)=c_{0}+\dfrac{1}{c_{1}u+\cfrac{1}{c_{2}+\cfrac{1}{c_{3}u+\cfrac{1}{\ddots+\cfrac{1}{c_{2l-1}u+\cfrac{1}{c_{2l}}}}}}},\quad\text{with}\quad c_{i}>0,\quad i=1,\ldots,2l, (3.10)

where c0=0c_{0}=0 if nn is even, c0>0c_{0}>0 if nn is odd, and ll as in (2.2).

Proof.

In fact, by Theorem 3.3, the polynomial pp is Hurwitz stable if and only if the function Φ\Phi can be represented in the form (3.5), where α=c0=0\alpha=c_{0}=0 if n=2ln=2l, and α=c0=a0a1>0\alpha=c_{0}=\dfrac{a_{0}}{a_{1}}>0 if n=2l+1n=2l+1. Now the assertion of the theorem follows from Theorem 1.18 and Corollary 1.38. ∎

From the formulæ (1.37)–(1.38) it follows that the coefficients cic_{i} of (3.10) can be found by the following formulæ

c2j=Dj2(Φ)D^j1(Φ)D^j(Φ),j=1,2,,l,c_{2j}=-\dfrac{D_{j}^{2}(\Phi)}{\widehat{D}_{j-1}(\Phi)\cdot\widehat{D}_{j}(\Phi)},\quad j=1,2,\ldots,l, (3.11)
c2j1=D^j12(Φ)Dj1(Φ)Dj(Φ),j=1,2,,l,c_{2j-1}=\dfrac{\widehat{D}_{j-1}^{2}(\Phi)}{D_{j-1}(\Phi)\cdot D_{j}(\Phi)},\quad j=1,2,\ldots,l, (3.12)

where D0(Φ)=D^0(Φ)1D_{0}(\Phi)=\widehat{D}_{0}(\Phi)\equiv 1. If nn is odd, then c0=a0a1>0c_{0}=\dfrac{a_{0}}{a_{1}}>0.

Using the formulæ (3.11)–(3.12) and (2.14)–(2.15), we can represent the coefficients cic_{i} in terms of Hurwitz determinants Δi(p)\Delta_{i}(p):

  • 1)

    If n=2ln=2l, then

    ci=Δi12(p)Δi2(p)Δi(p),i=1,2,,n,c_{i}=\dfrac{\Delta^{2}_{i-1}(p)}{\Delta_{i-2}(p)\cdot\Delta_{i}(p)},\quad i=1,2,\dots,n, (3.13)
  • 2)

    If n=2l+1n=2l+1, then

    ci=Δi2(p)Δi1(p)Δi+1(p),i=0,1,2,,n,c_{i}=\dfrac{\Delta^{2}_{i}(p)}{\Delta_{i-1}(p)\cdot\Delta_{i+1}(p)},\quad i=0,1,2,\dots,n, (3.14)

where we set Δ1(p)1a0\Delta_{-1}(p)\equiv\dfrac{1}{a_{0}} and Δ0(p)1\Delta_{0}(p)\equiv 1.

At last, from Theorems 1.22 and 3.3 one can obtain the following simple result.

Theorem 3.9.

Let pp be a Hurwitz polynomial of degree n2n\geqslant 2 as in (2.1). Then all the polynomials

pj(z)=i=0n2j[ni2]([ni2]1)([ni2]+j1)aizn2ji,j=1,,[n2]1,p_{j}(z)=\sum\limits_{i=0}^{n-2j}\left[\dfrac{n-i}{2}\right]\left(\left[\dfrac{n-i}{2}\right]-1\right)\cdots\left(\left[\dfrac{n-i}{2}\right]+j-1\right)a_{i}z^{n-2j-i},\quad j=1,\ldots,\left[\dfrac{n}{2}\right]-1,

also are Hurwitz stable.

Proof.

By Theorem 3.3, if pp is Hurwitz stable, then the function Φ=p1/p0\Phi=p_{1}/p_{0} is an R-function. According to Theorem 1.22, all functions Φj=p1(j)/p0(j)\Phi_{j}=p_{1}^{(j)}/p_{0}^{(j)}, j=1,,[n2]1j=1,\ldots,\left[\dfrac{n}{2}\right]-1, are R-functions with negative zeroes and poles. Now Theorem 3.3 implies that all polynomials

pj(z)=p0(j)(z2)+zp1(j)(z2),j=1,,[n2]1,p_{j}(z)=p_{0}^{(j)}(z^{2})+zp_{1}^{(j)}(z^{2}),\quad j=1,\ldots,\left[\dfrac{n}{2}\right]-1,

are Hurwitz stable, as required. ∎

3.2 Quasi-stable polynomials

In this section we deal with polynomials whose zeroes lie in the closed left half-plane of the complex plane.

Definition 3.10.

A polynomial pp of degree nn is called quasi-stable with degeneracy index mm, 0mn0\leqslant m\leqslant n, if all its zeroes lie in the closed left half-plane of the complex plane and the number of zeroes of pp, counting multiplicities, on the imaginary axis equals mm. We call the number nmn-m the stability index of the polynomial pp.

Throughout this section we use the following notation

r=[n2][m2],r=\left[\dfrac{n}{2}\right]-\left[\dfrac{m}{2}\right], (3.15)

where nn and mm are degree and degeneracy index of the polynomial pp, respectively.

Obviously, any Hurwitz polynomial is quasi-stable with zero degeneracy index, that is, it has the smallest degeneracy index and the largest stability index (which equals the degree of the polynomial). Note that if the degeneracy index mm is even, then p(0)0p(0)\neq 0, and if mm is odd, then pp must have a zero at 0.

Moreover, if pp is a quasi-stable polynomial, then

p(z)=p0(z2)+zp1(z2)=f(z2)q(z)=f(z2)[q0(z2)+zq1(z2)],p(z)=p_{0}(z^{2})+zp_{1}(z^{2})=f(z^{2})q(z)=f(z^{2})\left[q_{0}(z^{2})+zq_{1}(z^{2})\right],

where f(u)f(u) is a real polynomial of degree [m2]\left[\dfrac{m}{2}\right] with nonpositive zeroes, and qq is a Hurwitz stable polynomial if mm is even, and it is a quasi-stable polynomial with degeneracy index 11 if mm is odd. Using this representation of quasi-stable polynomials, one can extend almost all results of Section 3.1 to quasi-stable polynomials in the same way, so we state them here without proofs. We only should take into account that the function

Φ(u)=p1(u)p0(u)=f(u)q1(u)f(u)q0(u)=q1(u)q0(u)\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}=\dfrac{f(u)q_{1}(u)}{f(u)q_{0}(u)}=\dfrac{q_{1}(u)}{q_{0}(u)} (3.16)

has a pole at 0 whenever p(0)=0p(0)=0.

The analogue of Stodola necessary condition for quasi-stable polynomials is the following.

Theorem 3.11.

If the polynomial pp defined in (2.1) is quasi-stable, then all its coefficients are nonnegative.

The next theorem is an extended version of Theorem 3.3.

Theorem 3.12.

The polynomial pp of degree nn defined in (2.1) is quasi-stable with degeneracy index mm if and only if the function Φ\Phi defined in (3.16) is an R-function of negative type with exactly rr poles all of which are nonpositive, and limu±Φ(u)\displaystyle\lim_{u\to\pm\infty}\Phi(u) is positive if n=2l+1n=2l+1. The number rr is defined in (3.15).

If we expand the function Φ\Phi into its Laurent series (2.13), then the Hankel matrix S(Φ)=si+j0S(\Phi)=\|s_{i+j}\|^{\infty}_{0} has rank rr, where rr as in (3.15).

Theorem 3.13.

Let the polynomial pp be defined in (2.1). The following conditions are equivalent:

  • 1)

    the polynomial pp is quasi-stable with even degeneracy index mm;

  • 2)

    the following hold

    s1>0forn=2l+1,s_{-1}>0\quad\text{for}\quad n=2l+1,
    Dj(Φ)>0,j=1,,r,D_{j}(\Phi)>0,\qquad j=1,\ldots,r,
    (1)jD^j(Φ)>0,j=1,,r1,(-1)^{j}\widehat{D}_{j}(\Phi)>0,\qquad j=1,\ldots,r-1,

    and (1)rD^r(Φ)>0(-1)^{r}\widehat{D}_{r}(\Phi)>0 for even mm, but D^r(Φ)=0\widehat{D}_{r}(\Phi)=0 for odd mm. Here rr is defined in (3.15);

  • 3)

    the matrix S(Φ)S(\Phi) is sign regular of rank rr for even mm, S(Φ)S(\Phi) is r1r-1-sign regular of rank rr for odd mm, and s1s_{-1} is positive if n=2l+1n=2l+1.

It is also easy to extend Theorem 3.5 to quasi-stable polynomials.

Theorem 3.14.

Let the polynomial pp of degree nn be defined in (2.1). The following conditions are equivalent:

  • 1)

    the polynomial pp is quasi-stable with degeneracy index mm;

  • 2)

    determinants Δj(p)\Delta_{j}(p) are positive up to order nmn-m:

    Δ1(p)>0,Δ2(p)>0,,Δnm(p)>0,Δnm+1(p)==Δn(p)=0;\Delta_{1}(p)>0,\ \Delta_{2}(p)>0,\dots,\ \Delta_{n-m}(p)>0,\ \Delta_{n-m+1}(p)=\ldots=\Delta_{n}(p)=0;
  • 3)

    determinants ηj(p)\eta_{j}(p) are positive up to order nm+1n-m+1:

    η1(p)>0,η2(p)>0,,ηnm+1(p)>0,ηnm+i(p)=0,i=2,3,;\eta_{1}(p)>0,\ \eta_{2}(p)>0,\dots,\ \eta_{n-m+1}(p)>0,\ \eta_{n-m+i}(p)=0,\quad i=2,3,\ldots;
  • 4)

    the matrix H(p)H_{\infty}(p) is totally nonnegative and

    ηnm+1(p)0,ηnm+i(p)=0,i=2,3,\eta_{n-m+1}(p)\neq 0,\,\eta_{n-m+i}(p)=0,\quad i=2,3,\ldots

The implication 1)4)1)\Longrightarrow 4) was proved, indeed, in [3, 20]. The implication 4)1)4)\Longrightarrow 1) seems to be new. Thus, we established that a polynomial is quasi-stable if and only if its infinite Hurwitz matrix is totally nonnegative.

The following interesting property of quasi-stable polynomials is a simple consequence of Theorems 1.34 and 3.3.

Theorem 3.15.

Let the polynomials pp of degree n=2ln=2l defined in (2.1) be quasi-stable. Given any positive integer r(n)r(\leqslant n), the polynomial

pr(z)=a0z2k+a2r1z2k1+a2rz2k2+a4r1z2k3+a4rz2k4++a2rk1z+a2rk,p_{r}(z)=a_{0}z^{2k}+a_{2r-1}z^{2k-1}+a_{2r}z^{2k-2}+a_{4r-1}z^{2k-3}+a_{4r}z^{2k-4}+\ldots+a_{2rk-1}z+a_{2rk},

where k=[nr]k=\left[\dfrac{n}{r}\right], also is quasi-stable.

Theorem 3.16.

Let the polynomials pp of degree n=2l+1n=2l+1 defined in (2.1) be quasi-stable. Given any positive integer r(n)r(\leqslant n), the polynomial

pr(z)=a0z2k+1+a1z2k+a2rz2k1+a2r+1z2k2+a4rz2k3+a4r+1z2k4++a2rkz+a2rk+1,p_{r}(z)=a_{0}z^{2k+1}+a_{1}z^{2k}+a_{2r}z^{2k-1}+a_{2r+1}z^{2k-2}+a_{4r}z^{2k-3}+a_{4r+1}z^{2k-4}+\ldots+a_{2rk}z+a_{2rk+1},

where k=[nr]k=\left[\dfrac{n}{r}\right], also is quasi-stable.

Although total nonnegativity of the infinite Hurwitz matrix is equivalent to quasi-stability of polynomials, total nonnegativity of the finite singular Hurwitz matrix is not equivalent to quasi-stability as was noticed by Asner [3]. It is clear from Theorem 3.14 that the finite Hurwitz matrix of a quasi-stable polynomial is totally nonnegative. However, given a real polynomial pp of degree nn, if n(p)\mathcal{H}_{n}(p) is totally nonnegative, pp is not undertaken to be quasi-stable. In fact [16], pp can be represented in the form p(z)=q(z)g(z2)p(z)=q(z)g(z^{2}), where the polynomial qq is Hurwitz stable, and the polynomial gg is chosen such that the matrix n(p)\mathcal{H}_{n}(p) is totally nonnegative but g(z2)g(z^{2}) has zeroes in the open right half-plane. Such choice is possible as was mentioned in [16].

The connection between quasi-stable polynomials and Stieltjes continued fractions is similar to Hurwitz stable polynomials’ one. Namely, one can prove the following extended version of Theorem 3.8

Theorem 3.17.

The polynomial pp of degree nn defined in (2.1) is quasi-stable with degeneracy index mm if and only if the function Φ\Phi has a Stieltjes continued fraction expansion:

Φ(u)=c0+1c1u+1c2+1c3u+1+1T,withci>0,T={c2r,ifmis even,c2r1u,ifmis odd.\Phi(u)=c_{0}+\dfrac{1}{c_{1}u+\cfrac{1}{c_{2}+\cfrac{1}{c_{3}u+\cfrac{1}{\ddots+\cfrac{1}{T}}}}},\quad\text{with}\quad c_{i}>0,\quad T=\begin{cases}&c_{2r},\ \text{if}\ m\ \text{is even},\\ &c_{2r-1}u,\ \text{if}\ m\ \text{is odd}.\end{cases} (3.17)

Here c0=0c_{0}=0 if nn is even, c0>0c_{0}>0 if nn is odd, and rr is defined in (3.15).

The coefficients cic_{i} in (3.17) also can be found by the formulæ (3.13)–(3.14).

Finally, we note that a quasi-stable polynomial pp with degeneracy index 11 can be represented as follows: p(z)=zq(z)p(z)=zq(z), where qq is a Hurwitz stable polynomial. Therefore, for quasi-stable polynomials with degeneracy index 11, one can establish an analogue of Liénard and Chipart criterion.

Theorem 3.18.

The polynomial pp given by (2.1) is quasi-stable with degeneracy index 11 if and only if one of the following conditions holds

  • 1)1)

    an=0,an1>0,an2>0,an4>0,,Δ1(p)>0,Δ3(p)>0,Δ5(p)>0,a_{n}=0,a_{n-1}>0,a_{n-2}>0,a_{n-4}>0,\ldots,\quad\Delta_{1}(p)>0,\Delta_{3}(p)>0,\Delta_{5}(p)>0,\ldots;

  • 2)2)

    an=0,an1>0,an3>0,an5>0,,Δ1(p)>0,Δ3(p)>0,Δ5(p)>0,a_{n}=0,a_{n-1}>0,a_{n-3}>0,a_{n-5}>0,\ldots,\quad\Delta_{1}(p)>0,\Delta_{3}(p)>0,\Delta_{5}(p)>0,\ldots;

  • 3)3)

    an=0,an1>0,an2>0,an4>0,,Δ2(p)>0,Δ4(p)>0,Δ6(p)>0,a_{n}=0,a_{n-1}>0,a_{n-2}>0,a_{n-4}>0,\ldots,\quad\Delta_{2}(p)>0,\Delta_{4}(p)>0,\Delta_{6}(p)>0,\ldots;

  • 4)4)

    an=0,an1>0,an3>0,an5>0,,Δ2(p)>0,Δ4(p)>0,Δ6(p)>0,a_{n}=0,a_{n-1}>0,a_{n-3}>0,a_{n-5}>0,\ldots,\quad\Delta_{2}(p)>0,\Delta_{4}(p)>0,\Delta_{6}(p)>0,\ldots

4 Self-interlacing polynomials

In Section 3 we established that the function Φ\Phi (see (2.7)) associated with a Hurwitz stable polynomial (quasi-stable polynomial) maps the upper half-plane of the complex plane to the lower half-plane and possesses only negative (nonpositive) poles. Now we are in a position to describe polynomials whose associated function Φ\Phi also maps the upper half-plane to the lower half-plane but has only positive (nonnegative) poles.

4.1 General theory

Definition 4.1.

A real polynomial p(z)p(z) is called self-interlacing if all its zeroes are real and simple and interlace zeroes of the polynomial p(z)p(-z).

In other words, if λi\lambda_{i} are the zeroes of a self-interlacing polynomial pp, then one of the following holds:

0<λ1<λ2<λ3<<(1)n1λn,0<\lambda_{1}<-\lambda_{2}<\lambda_{3}<\ldots<(-1)^{n-1}\lambda_{n}, (4.1)
0<λ1<λ2<λ3<<(1)nλn,0<-\lambda_{1}<\lambda_{2}<-\lambda_{3}<\ldots<(-1)^{n}\lambda_{n}, (4.2)

where n=degpn=\deg p.

Definition 4.2.

A real polynomial pp of degree nn is called self-interlacing of type I (of type II) if all its zeroes are real and simple and satisfy the inequalities (4.1) (respectively, (4.2)).

If a polynomial pp of degree nn is self-interlacing of type I, then its minimal absolute value zero is positive. Moreover, if n=2ln=2l, then pp has exactly ll negative zeroes and exactly ll positive zeroes. Its zero λn\lambda_{n}, which has the maximal absolute value is negative. If n=2l+1n=2l+1, then the polynomial pp has exactly ll negative zeroes and l+1l+1 positive zeroes. In this case, λn\lambda_{n} is positive.

Note that a polynomial p(z)p(z) is self-interlacing of type I if and only if the polynomial p(z)p(-z) is self-interlacing of type II. So in the sequel, we deal only with self-interlacing polynomials of type I.

Theorem 4.3.

Let pp be a real polynomial of degree n1n\geqslant 1 as in (2.1). The polynomial pp is self-interlacing of type I if and only if its associated function Φ\Phi defined in (2.7) is an R-function with exactly ll poles, all of which are positive, and limu±Φ(u)\displaystyle\lim_{u\to\pm\infty}\Phi(u) is negative whenever n=2l+1n=2l+1.

Proof.

Let pp be self-interlacing of type I. First, we show that the function

G(z)=p(z)p(z)G(z)=-\dfrac{p(-z)}{p(z)}

is an R-function. In fact, by Definition 4.1, the zeroes of the polynomials p(z)p(z) and p(z)p(-z) are real, simple and interlacing, that is, between any two consecutive zeroes of one polynomial there lies exactly one zero, counting multiplicity, of the other polynomial, so G(z)G(z) or G(z)-G(z) is an R-function according to Theorem 1.18. By Corollary 1.20, GG is monotone between its poles. So it remains to prove that the function GG is decreasing between its poles.

Let n=2ln=2l, then by (4.1), the maximal pole of GG is λn1>0\lambda_{n-1}>0, but its maximal zero is λn>0-\lambda_{n}>0, which is greater than λn1\lambda_{n-1} according to Definition 4.2 (see (4.1)). Therefore, in the interval (λn,+)(-\lambda_{n},+\infty), the function GG has no poles and zeroes. At the same time, limz±G(z)=1\displaystyle\lim_{z\to\pm\infty}G(z)=-1. Consequently, in the interval (λn,+)(-\lambda_{n},+\infty), G(z)G(z) decreases from 0 to 1-1. So, GG is an R-function. In the same way, one can prove that if n=2l+1n=2l+1, then GG is also an R-function.

Thus, if pp is self-interlacing of type I, then GG maps the upper half-plane to the lower half-plane. From (2.8) we obtain that

zΦ(z2)=1+G(z)1G(z).z\Phi(z^{2})=\dfrac{1+G(z)}{1-G(z)}. (4.3)

Since the fractional linear transformation z1+z1z\displaystyle z\mapsto\frac{1+z}{1-z} conformally maps the lower half-plane to the lower half-plane:

Imz<0Im(1z1+z)=2Imz|1+z|2<0,\operatorname{Im}z<0\implies\operatorname{Im}\left(\frac{1-z}{1+z}\right)=\dfrac{2\operatorname{Im}z}{|1+z|^{2}}<0, (4.4)

from (4.3)–(4.4) it follows that the function zΦ(z2)z\Phi(z^{2}) maps the upper half-plane to the lower half-plane, that is, zΦ(z2)z\Phi(z^{2}) is an R-function.

Since pp is self-interlacing of type I by assumption, the polynomials p(z)p(z) and p(z)p(-z) have no common zeroes, therefore p0p_{0} and p1p_{1} also have no common zeroes and p0(0)0p_{0}(0)\neq 0 by (2.3). Therefore, the number of poles of the function zΦ(z2)z\Phi(z^{2}) equals the number of zeroes of the polynomial p0(z2)p_{0}(z^{2}). If n=2ln=2l, then by (2.4), degp0=l\deg p_{0}=l, so zΦ(z2)z\Phi(z^{2}) has exactly 2l2l poles. If n=2l+1n=2l+1, then by (2.5), degp1=l\deg p_{1}=l and degp0l\deg p_{0}\leqslant l, so zΦ(z2)z\Phi(z^{2}) has at most 2l2l poles, and it has exactly 2l+12l+1 zeroes, since p0(0)0p_{0}(0)\neq 0 as we mentioned above. But zΦ(z2)z\Phi(z^{2}) is an R-function, therefore, it has exactly 2l2l poles by Theorem 1.18.

Thus, Theorem 1.18 implies that the function zΦ(z2)z\Phi(z^{2}) can be represented in the form (1.25), where all poles are located symmetrically with respect to 0 and β=0\beta=0, since zΦ(z2)z\Phi(z^{2}) is an odd function. Denote the poles of zΦ(z2)z\Phi(z^{2}) by ±ν1,,±νl\pm\nu_{1},\ldots,\pm\nu_{l} such that

0<ν1<ν2<<νl.0<\nu_{1}<\nu_{2}<\ldots<\nu_{l}.

Note that ν10\nu_{1}\neq 0 since p0(0)0p_{0}(0)\neq 0.

So the function zΦ(z2)z\Phi(z^{2}) can be represented in the following form

zΦ(z2)=αz+j=1lγjzνj+j=1lγjz+νj=αz+j=1l2γjzz2νj2,α0,γj,νj>0.z\Phi(z^{2})=-\alpha z+\sum_{j=1}^{l}\frac{\gamma_{j}}{z-\nu_{j}}+\sum_{j=1}^{l}\frac{\gamma_{j}}{z+\nu_{j}}=-\alpha z+\sum_{j=1}^{l}\frac{2\gamma_{j}z}{z^{2}-\nu_{j}^{2}},\quad\alpha\geqslant 0,\,\gamma_{j},\nu_{j}>0.

Divide this equality by zz and changing variables as follows z2uz^{2}\to u, 2γjβj2\gamma_{j}\to\beta_{j}, νj2ωj\nu_{j}^{2}\to\omega_{j}, we obtain the following representation of the function Φ\Phi:

Φ(u)=α+j=1lβjuωj,\Phi(u)=-\alpha+\sum_{j=1}^{l}\frac{\beta_{j}}{u-\omega_{j}}, (4.5)

where α0,βj>0\alpha\geqslant 0,\,\beta_{j}>0 and

0<ω1<ω2<<ωl.0<\omega_{1}<\omega_{2}<\ldots<\omega_{l}.

Here α=0\alpha=0 whenever n=2ln=2l, and α=a0a1<0-\alpha=\displaystyle\frac{a_{0}}{a_{1}}<0 whenever n=2l+1n=2l+1. Since Φ\Phi can be represented in the form (4.5), by Theorem 1.18, Φ\Phi is an R-function with exactly ll poles. Moreover, from (4.5) it also follows that all poles Φ\Phi are positive and limu±Φ(u)<0\displaystyle\lim_{u\to\pm\infty}\Phi(u)<0 if n=2l+1n=2l+1.

Conversely, let the polynomial pp be defined in (2.1), and let its associated function Φ\Phi be an R-function with exactly ll poles, all of which are positive, and let limu±Φ(u)<0\displaystyle\lim_{u\to\pm\infty}\Phi(u)<0 when n=2l+1n=2l+1. We will show that pp is self-interlacing of type I.

By Theorem 1.18, Φ\Phi can be represented in the form (4.5), where α=limu±Φ(u)0-\alpha=\displaystyle\lim_{u\to\pm\infty}\Phi(u)\leqslant 0 such that α>0\alpha>0 if n=2l+1n=2l+1, and α=0\alpha=0 if n=2ln=2l. Thus, the polynomial p0p_{0} has only positive zeroes, and the polynomials p0p_{0} and p1p_{1} have no common zeroes. Together with (2.3) and (2.7), this means that the set of zeroes of the polynomial pp coincides with the set of roots of the equation

zΦ(z2)=1.z\Phi(z^{2})=-1. (4.6)

At first, we study real zeroes of this equation, so we consider only real zz. Since p0p_{0} has only positive zeroes, we have p(0)=p0(0)0p(0)=p_{0}(0)\neq 0. Thus, we put z\{0}z\in\mathbb{R}\backslash\{0\}. Changing variables as follows z2u>0z^{2}\to u>0, we rewrite the equation (4.6) in the following form, which is equivalent to (4.6) for real nonzero zz:

{Φ(u)=1u,Φ(u)=1u,u>0.\begin{cases}\ \Phi(u)=-\dfrac{1}{\sqrt{u}},\\ \ \Phi(u)=\dfrac{1}{\sqrt{u}},\end{cases}\quad u>0. (4.7)

Note that all positive roots (if any) of the first equation (4.7) are squares of the positive roots of the equation (4.6), and all positive roots (if any) of the second equation (4.7) are squares of the negative roots of the equation (4.6).

Let n=2ln=2l. Then limu±Φ(u)=α=0\lim\limits_{u\to\pm\infty}\Phi(u)=-\alpha=0. Consider the function F1(u)=Φ(u)+1uF_{1}(u)=\Phi(u)+\dfrac{1}{\sqrt{u}} whose set of zeroes coincides with the set of roots of the first equation (4.7). The function F1(u)F_{1}(u) has the same positive poles as Φ(u)\Phi(u) does, that is, ω1,ω2,,ωl\omega_{1},\omega_{2},\ldots,\omega_{l}. Moreover, F1(u)F_{1}(u) is decreasing on the intervals (0,ω1),(ω1,ω2),,(ωl1,ωl)(0,\omega_{1}),(\omega_{1},\omega_{2}),\ldots,(\omega_{l-1},\omega_{l}), (ωl,+)(\omega_{l},+\infty) as a sum of functions that are decreasing on those intervals. Since F1F_{1} is decreasing on (ωl,+)(\omega_{l},+\infty) and F1(u)+0F_{1}(u)\to+0 as u+u\to+\infty, we have F1(u)>0F_{1}(u)>0 for u(ωl,+)u\in(\omega_{l},+\infty). Further, F1(u)F_{1}(u)\to-\infty as uωiu\nearrow\omega_{i} and F1(u)+F_{1}(u)\to+\infty as uωiu\searrow\omega_{i}, i=1,,li=1,\ldots,l. Also F1(u)+F_{1}(u)\to+\infty whenever u+0u\to+0, since |Φ(0)|=|p1(0)p0(0)|<|\Phi(0)|=\left|\dfrac{p_{1}(0)}{p_{0}(0)}\right|<\infty and 1u+\dfrac{1}{\sqrt{u}}\to+\infty as u+0u\to+0. Thus, on each of the intervals (0,ω1),(ω1,ω2),,(ωl1,ωl)(0,\omega_{1}),(\omega_{1},\omega_{2}),\ldots,(\omega_{l-1},\omega_{l}), the function F1F_{1} decreases from ++\infty to -\infty, and therefore, it has exactly one zero, counting multiplicity, on each of those intervals. Denote the zero of F1F_{1} on the interval (ωi1,ωi)(\omega_{i-1},\omega_{i}) by μi2\mu_{i}^{2}, i=2,,li=2,\ldots,l (μi>0\mu_{i}>0). Also we denote by μ12\mu_{1}^{2} (μ1>0\mu_{1}>0) the zero of F1F_{1} on the interval (0,ω1)(0,\omega_{1}). Since Φ\Phi is an R-function by assumption, it has exactly one zero, counting multiplicity, say ξi\xi_{i}, on each of the intervals (ωi,ωi+1)(\omega_{i},\omega_{i+1}), i=1,,l1i=1,\ldots,l-1. But F1(ξi)=1ξi>0F_{1}(\xi_{i})=\dfrac{1}{\sqrt{\xi_{i}}}>0, so we have

μ12<ω1<ξ1<μ22<ω2<ξ2<<ξl1<μl2<ωl.\mu^{2}_{1}<\omega_{1}<\xi_{1}<\mu^{2}_{2}<\omega_{2}<\xi_{2}<\ldots<\xi_{l-1}<\mu_{l}^{2}<\omega_{l}. (4.8)

Thus, the first equation (4.7) has exactly ll positive roots μi\mu_{i}, all of which are simple and satisfy the inequalities (4.8).

Consider now the function F2(u)=Φ(u)1uF_{2}(u)=\Phi(u)-\dfrac{1}{\sqrt{u}} whose set of zeroes coincides with the set of roots of the second equation (4.7). Since Φ(0)=i=1lβiωi<0\Phi(0)=-\displaystyle\sum_{i=1}^{l}\dfrac{\beta_{i}}{\omega_{i}}<0 by (4.5), and Φ(u)\Phi(u)\to-\infty as uω1u\nearrow\omega_{1}, the function Φ\Phi decreases from Φ(0)<0\Phi(0)<0 to -\infty on the interval (0,ω1)(0,\omega_{1}). Consequently, F2(u)<0F_{2}(u)<0 for all u(0,ω1)u\in(0,\omega_{1}). It is clear that F2(u)+F_{2}(u)\to+\infty as uωiu\searrow\omega_{i} and F2(u)F_{2}(u)\to-\infty as uωiu\nearrow\omega_{i}, i=1,,li=1,\ldots,l. Since F2(μi2)=1μi<0F_{2}(\mu^{2}_{i})=-\dfrac{1}{\mu_{i}}<0, the function F2F_{2} has an odd number of zeroes, counting multiplicities, on each of the intervals (ωi,μi2)(\omega_{i},\mu^{2}_{i}), i=1,,l1i=1,\ldots,l-1. Besides, F2(u)=F1(u)2u<0F_{2}(u)=F_{1}(u)-\dfrac{2}{\sqrt{u}}<0 whenever u[μi2,ωi+1)u\in[\mu^{2}_{i},\omega_{i+1}), i=1,,l1i=1,\ldots,l-1, because F1(u)<0F_{1}(u)<0 on those intervals. More exactly, since F2(ξi)=1ξi<0F_{2}(\xi_{i})=\dfrac{1}{\sqrt{\xi_{i}}}<0, i=1,,r1i=1,\ldots,r-1, the function F2(u)F_{2}(u) has an odd number of zeroes on each interval (ωi,ξi)(\omega_{i},\xi_{i}) and an even number of zeroes on each interval [ξi,μi+12)[\xi_{i},\mu^{2}_{i+1}), i=1,,r1i=1,\ldots,r-1. So F2F_{2} has at least l1l-1 zeroes in the interval (ω1,ωl)(\omega_{1},\omega_{l}). Consider now the interval (ωl,+)(\omega_{l},+\infty) and show that F2(u)0F_{2}(u)\to-0 as u+u\to+\infty. In fact, since the function Φ\Phi is an R-function by assumption, we have s0=limu±uΦ(u)=D1(Φ)>0s_{0}=\lim\limits_{u\to\pm\infty}u\Phi(u)=D_{1}(\Phi)>0 according Theorem 1.18, where s0s_{0} is the first coefficient101010Recall that s1=0s_{-1}=0 whenever n=2ln=2l. in the series (2.13). Therefore, Φ(u)u1\Phi(u)\sim u^{-1} as u+u\to+\infty. This implies the following: F2(u)u12F_{2}(u)\sim-u^{-\tfrac{1}{2}} as u+u\to+\infty. Thus, F2(ωl+ε)>0F_{2}(\omega_{l}+\varepsilon)>0 for all sufficiently small ε>0\varepsilon>0 and F2(u)<0F_{2}(u)<0 for all sufficiently large positive uu. Consequently, F2F_{2} has an odd number of zeroes, counting multiplicities, in the interval (ωl,+)(\omega_{l},+\infty). So F2F_{2} has at least ll zeroes in the interval (ω1,+)(\omega_{1},+\infty), and it has no zeroes in (0,ω1)(0,\omega_{1}).

Since the first equation (4.7) has exactly ll positive roots and the second equation (4.7) has at least ll positive roots, the equation (4.6) has at least 2l2l real roots, all of which are the zeroes of the polynomial pp. But degp=2l\deg p=2l by assumption, therefore, the second equation (4.7) also has exactly ll positive roots. Moreover, it has exactly one simple root in each of the intervals (ωi,ξi)(\omega_{i},\xi_{i}), i=1,,l1i=1,\dots,l-1, and exactly one simple root in the interval (ωl,+)(\omega_{l},+\infty). Thus, denoting the positive roots of the second equation (4.7) by ζi2\zeta_{i}^{2}, i=1,,li=1,\ldots,l (ζi<0\zeta_{i}<0), we have

ω1<ζ12<ξ1<ω2<ζ22<ξ2<<ξl1<ωl<ζl2.\omega_{1}<\zeta_{1}^{2}<\xi_{1}<\omega_{2}<\zeta_{2}^{2}<\xi_{2}<\ldots<\xi_{l-1}<\omega_{l}<\zeta_{l}^{2}. (4.9)

Now from (4.8)–(4.9) we obtain

0<μ1<ζ1<μ2<ζ2<μl<ζl.0<\mu_{1}<-\zeta_{1}<\mu_{2}<-\zeta_{2}\ldots<\mu_{l}<-\zeta_{l}. (4.10)

Recall that μi>0\mu_{i}>0 are the positive roots of the equation (4.6), and ζi<0\zeta_{i}<0 are the negative roots of the equation (4.6). Thus, all roots of the equation (4.6) (and therefore, all zeroes of the polynomial pp) are real and simple. Denote them by λi\lambda_{i} and enumerate such that

0<|λ1|<|λ2|<|λn|.0<|\lambda_{1}|<|\lambda_{2}|\ldots<|\lambda_{n}|.

Then we have λ2i1=μi>0\lambda_{2i-1}=\mu_{i}>0 and λ2i=ζi<0\lambda_{2i}=\zeta_{i}<0, i=1,,li=1,\ldots,l, and the inequalities (4.10) imply (4.1). Thus, pp is a self-interlacing polynomial of type I.

In the same way, one can show that if n=2l+1n=2l+1, the function Φ\Phi is an R-function with positive poles, and limu±Φ(u)<0\lim\limits_{u\to\pm\infty}\Phi(u)<0, then the polynomial pp is self-interlacing of type I. ∎

Remark 4.4.

Let us note that if the even and odd parts, p0p_{0} and p1p_{1}, of a given polynomial pp have positive interlacing zeroes, then the polynomials p0(z2)p_{0}(z^{2}) and zp1(z2)zp_{1}(z^{2}) have real interlacing zeroes, and therefore the polynomial p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) has real zeroes by Theorem 1.18 and Corollary 1.19 as a linear combination of polynomials with real interlacing zeroes. However, this notice does not help to investigate the self-interlacing property of polynomials.

Remark 4.5.

In the proof of Theorem 4.3, we also established that the squares of both positive and negative zeroes of the self-interlacing polynomial pp interlace both zeroes of p0p_{0} and zeroes of p1p_{1}.

Theorem 4.3 allow us to use properties of R-functions in order to obtain additional criteria of self-interlacing. At first, consider the Hankel matrix S(Φ)=si+ji,j=0S(\Phi)=\|s_{i+j}\|^{\infty}_{i,j=0} constructed with the coefficients of the series (2.13). From Theorems 1.184.3 and 1.11 and from Corollary 1.25 we obtain the following self-interlacing criteria:

Theorem 4.6.

Let a real polynomial pp be defined by (2.1). The following conditions are equivalent:

  • 1)

    the polynomial pp is self-interlacing of type I;

  • 2)

    the following hold

    s1<0forn=2l+1,Dj(Φ)>0,j=1,,l,D^j(Φ)>0,j=1,,l,\begin{split}&s_{-1}<0\quad\text{for}\quad n=2l+1,\\ &D_{j}(\Phi)>0,\qquad j=1,\ldots,l,\\ &\widehat{D}_{j}(\Phi)>0,\qquad j=1,\ldots,l,\end{split} (4.11)

    where l=[n2]l=\left[\dfrac{n}{2}\right];

  • 3)

    the matrix S(Φ)S(\Phi) is strictly totally positive of rank ll, and s1<0s_{-1}<0 whenever n=2l+1n=2l+1.

The determinants Dj(Φ)D_{j}(\Phi) and D^j(Φ)\widehat{D}_{j}(\Phi) are defined in (1.5) and (1.8), respectively.

Proof.

According to Theorem 4.3, the polynomial pp is self-interlacing of type I if and only if its associated function Φ\Phi is an R-function with exactly l=[n2]l=\left[\dfrac{n}{2}\right] poles, all of which are positive, and limu±Φ(u)=s10\lim\limits_{u\to\pm\infty}\Phi(u)=s_{-1}\leqslant 0. By Theorem 1.18 and Corollary 1.25, this is equivalent to the inequalities (4.11). But these inequalities are equivalent to the strictly total positivity of the matrix S(Φ)S(\Phi) by Theorem 1.11. ∎

From this theorem we obtain the following criterion of self-interlacing, which is an analogue of the Hurwitz stability criterion.

Theorem 4.7.

Given a real polynomial pp of degree nn as in (2.1), the following conditions are equivalent:

  • 1)

    pp is self-interlacing of type I;

  • 2)

    the Hurwitz determinants Δj(p)\Delta_{j}(p) satisfy the inequalities:

    Δn1(p)>0,Δn3(p)>0,,\Delta_{n-1}(p)>0,\ \Delta_{n-3}(p)>0,\ldots, (4.12)
    (1)[n+12]Δn(p)>0,(1)[n+12]1Δn2(p)>0,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]}\Delta_{n}(p)>0,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]-1}\Delta_{n-2}(p)>0,\ldots (4.13)
  • 3)

    the determinants ηj(p)\eta_{j}(p) up to order n+1n+1 satisfy the inequalities:

    ηn(p)>0,ηn2(p)>0,,η1(p)=a0>0,\eta_{n}(p)>0,\ \eta_{n-2}(p)>0,\ldots,\ \eta_{1}(p)=a_{0}>0, (4.14)
    (1)[n+12]ηn+1(p)>0,(1)[n+12]1ηn1(p)>0,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]}\eta_{n+1}(p)>0,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]-1}\eta_{n-1}(p)>0,\ldots (4.15)
Proof.

We establish the equivalence 1)2)1)\Longleftrightarrow 2) while the equivalence 2)3)2)\Longleftrightarrow 3) follows from (2.12).

By Theorem 4.6, pp is self-interlacing of type I if and only if the inequalities (4.11) hold. Now from (4.11) and (2.14)–(2.15) we obtain that pp is self-interlacing of type I if and only if the inequalities (4.12) hold and

forn=2l(1)jΔ2j(p)>0,j=1,,l;forn=2l+1(1)j+1Δ2j+1(p)>0,j=0,1,,l,\begin{array}[]{l}\text{for}\qquad n=2l\qquad\qquad\qquad(-1)^{j}\Delta_{2j}(p)>0,\qquad j=1,\ldots,l;\\ \\ \text{for}\qquad n=2l+1\qquad\qquad\;(-1)^{j+1}\Delta_{2j+1}(p)>0,\qquad j=0,1,\ldots,l,\end{array}

that is equivalent to (4.13). ∎

Note that the inequalities (4.13) and (4.15) are equivalent to the following ones:

Δn2i(p)Δn2i2(p)<0,andηn2i+1(p)ηn2i1(p)<0,i=0,1,,[n12],\Delta_{n-2i}(p)\Delta_{n-2i-2}(p)<0,\quad\text{and}\quad\eta_{n-2i+1}(p)\eta_{n-2i-1}(p)<0,\quad i=0,1,\ldots,\left[\dfrac{n-1}{2}\right],

where Δ0(p)=η0(p)=a0Δ1(p)1\Delta_{0}(p)=\eta_{0}(p)=a_{0}\Delta_{-1}(p)\equiv 1.

Analogues of Theorems 3.73.9 will be established in Section 4.3 using a connection between Hurwitz stable polynomials and self-interlacing polynomials of type I.

4.2 Interrelation between Hurwitz stable and self-interlacing polynomials

Comparing Theorems 3.3 and 4.3, one obtains the following fact, which has deep consequences and allows us to describe a lot of properties of self-interlacing polynomials.

Theorem 4.8.

A polynomial p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) is self-interlacing of type I if and only if the polynomial q(z)=p(z2)zp1(z2)q(z)=p(-z^{2})-zp_{1}(z^{2}) is Hurwitz stable.

Proof.

By Theorem 4.3, pp is self-interlacing if and only if the function Φ(u)=p1(u)p0(u)\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)} is an R-function with only positive poles, and limu±Φ(u)0\lim\limits_{u\to\pm\infty}\Phi(u)\leqslant 0 that is equivalent to the fact that the function Ψ(u)=p1(u)p0(u)\Psi(u)=-\dfrac{p_{1}(-u)}{p_{0}(-u)} is an R-function with only negative poles, and limu±Ψ(u)0\lim\limits_{u\to\pm\infty}\Psi(u)\geqslant 0. According to Theorem 3.3, this is equivalent to the polynomial qq being Hurwitz stable, as required. ∎

Remark 4.9.

Thus, we have that there is one-to-one correspondence between self-interlacing and Hurwitz stable polynomials. Given a real Hurwitz stable polynomials, we should change sings of some its coefficients to obtain a self-interlacing polynomial, and vise versa.

As an immediate consequence of Theorem 4.8, we obtain the following analogue of Stodola’s theorem, Theorem 3.2.

Theorem 4.10.

If the polynomial

p(z)=a0zn+a1zn1++an,a1,,an,a0>0,p(z)=a_{0}z^{n}+a_{1}z^{n-1}+\dots+a_{n},\qquad a_{1},\dots,a_{n}\in\mathbb{R},\ a_{0}>0, (4.16)

is self-interlacing of type I, then

  • for n=2ln=2l,

    (1)j(j1)2aj>0,j=0,1,,n;(-1)^{\tfrac{j(j-1)}{2}}a_{j}>0,\qquad j=0,1,\ldots,n;
  • for n=2l+1n=2l+1,

    (1)j(j+1)2aj>0,j=0,1,,n.(-1)^{\tfrac{j(j+1)}{2}}a_{j}>0,\qquad j=0,1,\ldots,n.
Proof.

Since pp is self-interlacing of type I, by Theorem 4.8, the polynomial111111We put the factor (1)n(n+1)2(-1)^{\tfrac{n(n+1)}{2}} in order to make the leading coefficient of the polynomial qq equal to a0>0a_{0}>0.

q(z)=(1)n(n+1)2[p0(z2)zp1(z2)]=b0zn+b1zn1+b2zn2++bn,b0=a0>0,q(z)=(-1)^{\tfrac{n(n+1)}{2}}[p_{0}(-z^{2})-zp_{1}(-z^{2})]=b_{0}z^{n}+b_{1}z^{n-1}+b_{2}z^{n-2}+\dots+b_{n},\quad b_{0}=a_{0}>0, (4.17)

is Hurwitz stable. It is easy to see that

bj=(1)j(j1)2aj,j=0,1,,nforn=2l;b_{j}=(-1)^{\tfrac{j(j-1)}{2}}a_{j},\qquad j=0,1,\ldots,n\qquad\qquad\qquad\quad\quad\text{for}\qquad n=2l; (4.18)

and

bj=(1)j(j+1)2aj,j=0,1,,nforn=2l+1.\quad\,\,\,\,\qquad\qquad\qquad b_{j}=(-1)^{\tfrac{j(j+1)}{2}}a_{j},\qquad j=0,1,\ldots,n\qquad\qquad\qquad\qquad\text{for}\qquad n=2l+1. (4.19)

By Stodola’s theorem, Theorem 3.2, all bjb_{j} are positive. ∎

Thus, a necessary form of a self-interlacing polynomial with positive leading coefficient is as follows:
for n=2ln=2l

p(z)=b0zn+b1zn1b2zn2b3zn3+b4zn4+b5zn5b6zn6,b0,,bn>0,p(z)=b_{0}z^{n}+b_{1}z^{n-1}-b_{2}z^{n-2}-b_{3}z^{n-3}+b_{4}z^{n-4}+b_{5}z^{n-5}-b_{6}z^{n-6}-\dots,\qquad b_{0},\dots,b_{n}>0,

for n=2l+1n=2l+1

p(z)=b0znb1zn1b2zn2+b3zn3+b4zn4b5zn5b6zn6+,b0,,bn>0.p(z)=b_{0}z^{n}-b_{1}z^{n-1}-b_{2}z^{n-2}+b_{3}z^{n-3}+b_{4}z^{n-4}-b_{5}z^{n-5}-b_{6}z^{n-6}+\dots,\qquad b_{0},\dots,b_{n}>0.

We also notice that if the polynomial pp defined in (4.16) is self-interlacing, then

an1an<0,a_{n-1}a_{n}<0, (4.20)

since Ψ(0)=Φ(0)=p1(0)/p0(0)=an1/an>0\Psi(0)=-\Phi(0)=-{p_{1}(0)}/{p_{0}(0)}=-{a_{n-1}}/{a_{n}}>0 (see the proof of Theorem 4.10).

Remark 4.11.

In fact, Theorem 4.10 was proved in [8] by another methods.

Let us point out at one more interesting connection between Hurwitz stable and self-interlacing polynomials. Let the polynomial pp be self-interlacing. If p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}), then p(iz)=p0(z2)+izp1(z2)p(iz)=p_{0}(-z^{2})+izp_{1}(-z^{2}), where i=1i=\sqrt{-1}. Consequently, the Hurwitz stable polynomials q(z)=p0(z2)zp1(z2)q(z)=p_{0}(-z^{2})-zp_{1}(-z^{2}) can be represented as follows

q(z)=p(iz)+p(iz)2p(iz)p(iz)2i=p(iz)1+i2+p(iz)1i2.q(z)=\dfrac{p(iz)+p(-iz)}{2}-\dfrac{p(iz)-p(-iz)}{2i}=p(iz)\dfrac{1+i}{2}+p(-iz)\dfrac{1-i}{2}.

So one can establish the following theorem.

Theorem 4.12.

Let the polynomial p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) be self-interlacing of type I and let its dual polynomial q(z)=p0(z2)zp1(z2)q(z)=p_{0}(-z^{2})-zp_{1}(-z^{2}) be the Hurwitz stable polynomial associated with pp. Then

p(λ)=0argq(iλ)=π4or5π4;p(\lambda)=0\Longleftrightarrow\arg q(i\lambda)=\dfrac{\pi}{4}\,\,\,\text{or}\,\,\,\dfrac{5\pi}{4};

and, respectively,

q(μ)=0argp(iμ)=π4or5π4.q(\mu)=0\Longleftrightarrow\arg p(i\mu)=\dfrac{\pi}{4}\,\,\,\text{or}\,\,\,\dfrac{5\pi}{4}.
Proof.

In fact, p(λ)=0p(\lambda)=0 if and only if λp1(λ2)p0(λ2)=1-\dfrac{\lambda p_{1}(\lambda^{2})}{p_{0}(\lambda^{2})}=1. At the same time, q(iλ)=p0(λ2)iλp1(λ2)q(i\lambda)=p_{0}(\lambda^{2})-i\lambda p_{1}(\lambda^{2}). Consequently, argq(iλ)=arctan(λp1(λ2)p0(λ2))=arctan1=π4or5π4\arg q(i\lambda)=\arctan\left(-\dfrac{\lambda p_{1}(\lambda^{2})}{p_{0}(\lambda^{2})}\right)=\arctan 1=\dfrac{\pi}{4}\,\text{or}\,\dfrac{5\pi}{4}.

The second assertion of the theorem can be proved analogously. ∎

4.3 Liénard and Chipart criterion, Stieltjes continued fractions and the signs of Hurwitz minors

Now we consider the polynomials p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) and q(z)=p0(z2)zp1(z2)q(z)=p_{0}(-z^{2})-zp_{1}(-z^{2}) and their associated functions Φ(u)=p1(u)p0(u)\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)} and Ψ(u)=p1(u)p0(u)\Psi(u)=-\dfrac{p_{1}(-u)}{p_{0}(-u)}. Let

Φ(u)=s1+s0u+s1u2+s2u3+s3u4+\Phi(u)=s_{-1}+\frac{s_{0}}{u}+\frac{s_{1}}{u^{2}}+\frac{s_{2}}{u^{3}}+\frac{s_{3}}{u^{4}}+\dots

Then we have

Ψ(u)=p1(u)p0(u)=t1+t0w+t1w2+t2w3+t3w4+t3w5+=s1+s0ws1w2+s2w3s3w4+s3w5\Psi(u)=-\dfrac{p_{1}(-u)}{p_{0}(-u)}=t_{-1}+\frac{t_{0}}{w}+\frac{t_{1}}{w^{2}}+\frac{t_{2}}{w^{3}}+\frac{t_{3}}{w^{4}}+\frac{t_{3}}{w^{5}}+\dots=-s_{-1}+\frac{s_{0}}{w}-\frac{s_{1}}{w^{2}}+\frac{s_{2}}{w^{3}}-\frac{s_{3}}{w^{4}}+\frac{s_{3}}{w^{5}}-\dots

Thus, the connection between sjs_{j} and tjt_{j} is as follows

tj=(1)jsj,j=1,0,1,2,t_{j}=(-1)^{j}s_{j},\qquad j=-1,0,1,2,\ldots

Consider the two infinite Hankel matrices S=si+j0S=\|s_{i+j}\|_{0}^{\infty} and T=ti+j0T=\|t_{i+j}\|_{0}^{\infty}.

In [16] there was proved the following formula.

Theorem 4.13.

Minors of the matrices SS and TT are connected as follows

T(i1i2imj1j2jm)=(1)k=1mik+k=1mjkS(i1i2imj1j2jm),m=1,2,T\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ j_{1}&j_{2}&\dots&j_{m}\\ \end{pmatrix}=(-1)^{\sum\limits_{k=1}^{m}i_{k}+\sum\limits_{k=1}^{m}j_{k}}S\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ j_{1}&j_{2}&\dots&j_{m}\\ \end{pmatrix},\quad m=1,2,\ldots

From this theorem one can easily obtain the following consequence.

Corollary 4.14.

Let rational functions Φ(u)\Phi(u) and Ψ(u)\Psi(u) be such that Ψ(u)=Φ(u)\Psi(u)=-\Phi(-u). Then the following equalities hold

Dj(Φ)=Dj(Ψ),j=1,2,D_{j}(\Phi)=D_{j}(\Psi),\quad j=1,2,\ldots
D^j(Φ)=(1)jD^j(Ψ),j=1,2,\widehat{D}_{j}(\Phi)=(-1)^{j}\widehat{D}_{j}(\Psi),\quad j=1,2,\ldots

From this corollary and from (2.14)–(2.15) one obtains

Δn+12j(q)=Δn+12j(p),j=1,,[n2],\Delta_{n+1-2j}(q)=\Delta_{n+1-2j}(p),\qquad j=1,\ldots,\left[\dfrac{n}{2}\right], (4.21)
Δn2j(q)=(1)[n+12]jΔn2j(p),j=0,1,,[n2],\Delta_{n-2j}(q)=\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]-j}\Delta_{n-2j}(p),\qquad j=0,1,\ldots,\left[\dfrac{n}{2}\right], (4.22)

where the polynomials pp and qq are defined in (4.16) and (4.17), respectively.

Now we are in position to prove analogues of Theorems 3.73.9 for self-interlacing polynomials.

Theorem 4.15.

The polynomial pp defined in (4.16) is self-interlacing of type I if and only if

(1)[n+12]an>0,(1)[n+12]1an2>0,(1)[n+12]2an4>0,(-1)^{\left[\tfrac{n+1}{2}\right]}a_{n}>0,\,(-1)^{\left[\tfrac{n+1}{2}\right]-1}a_{n-2}>0,\,(-1)^{\left[\tfrac{n+1}{2}\right]-2}a_{n-4}>0,\,\dots (4.23)

or

(1)[n+12]an>0,(1)[n+12]1an1>0,(1)[n+12]2an3>0,(-1)^{\left[\tfrac{n+1}{2}\right]}a_{n}>0,\,(-1)^{\left[\tfrac{n+1}{2}\right]-1}a_{n-1}>0,\,(-1)^{\left[\tfrac{n+1}{2}\right]-2}a_{n-3}>0,\,\dots (4.24)

and one of the following two conditions holds

  • 1)
    Δn1(p)>0,Δn3(p)>0,Δn5(p)>0,;\Delta_{n-1}(p)>0,\,\Delta_{n-3}(p)>0,\,\Delta_{n-5}(p)>0,\,\dots; (4.25)
  • 2)
    (1)[n+12]Δn(p)>0,(1)[n+12]1Δn2(p)>0,(1)[n+12]2Δn4(p)>0,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]}\Delta_{n}(p)>0,\,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]-1}\Delta_{n-2}(p)>0,\,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]-2}\Delta_{n-4}(p)>0,\,\ldots (4.26)
Proof.

If the polynomial pp is self-interlacing of type I, then by Theorems 4.7 and 4.10, all the conditions (4.23)–(4.26) hold.

Let the conditions (4.23) and (4.25) hold. Consider the polynomial qq defined in (4.17). From (4.23) and (4.18)–(4.19) it follows that bn>0,bn2>0,bn4>0,b_{n}>0,b_{n-2}>0,b_{n-4}>0,\ldots, and from (4.25), (2.14)–(2.15) and (4.21) we obtain that Δn1(q)>0,Δn3(q)>0,\Delta_{n-1}(q)>0,\Delta_{n-3}(q)>0,\ldots Thus, the polynomial qq satisfies the condition 1)1) of Theorem 3.7. Therefore, qq is Hurwitz stable, so pp is self-interlacing of type I according to Theorem 4.8.

Analogously, using (4.18)–(4.19), (2.14)–(2.15), (4.21)–(4.22) and Theorem 3.7 one can show that the conditions (4.24) and (4.25), or (4.23) and (4.26), or (4.24) and (4.26) imply Hurwitz stability of the polynomial qq, so by Theorem 4.8, pp is self-interlacing of type I. ∎

The following theorem presents a relation between self-interlacing polynomials and continued fractions of Stieltjes type.

Theorem 4.16.

The polynomial pp of degree nn defined in (4.16) is self-interlacing of type I if and only if its associated function Φ\Phi has the following Stieltjes continued fraction expansion:

Φ(u)=c0+1c1u+1c2+1c3u+1+1c2l1u+1c2l,with(1)i1ci>0,i=1,,2l,\Phi(u)=c_{0}+\dfrac{1}{c_{1}u+\cfrac{1}{c_{2}+\cfrac{1}{c_{3}u+\cfrac{1}{\ddots+\cfrac{1}{c_{2l-1}u+\cfrac{1}{c_{2l}}}}}}},\quad\text{with}\quad(-1)^{i-1}c_{i}>0,\quad i=1,\ldots,2l, (4.27)

where c0=0c_{0}=0 if nn is even, and c0>0c_{0}>0 if nn is odd, and ll as in (2.2).

Proof.

In fact, by Theorem 4.3, the polynomial pp is self-interlacing of type I if and only if the function Φ\Phi can be represented in the form (4.5), where α=c0=0-\alpha=c_{0}=0 if n=2ln=2l, and α=c0=a0a1<0-\alpha=c_{0}=\dfrac{a_{0}}{a_{1}}<0 if n=2l+1n=2l+1. Now the assertion of the theorem follows from Theorem 1.18 and Corollary 1.39. ∎

Theorems 3.9 and 4.8 imply the following theorem.

Theorem 4.17.

Let pp be a self-interlacing polynomial of type I of degree n2n\geqslant 2 as in (4.16). Then all the polynomials

pj(z)=i=0n2j[ni2]([ni2]1)([ni2]+j1)aizn2ji,j=1,,[n2]1,p_{j}(z)=\sum\limits_{i=0}^{n-2j}\left[\dfrac{n-i}{2}\right]\left(\left[\dfrac{n-i}{2}\right]-1\right)\cdots\left(\left[\dfrac{n-i}{2}\right]+j-1\right)a_{i}z^{n-2j-i},\quad j=1,\ldots,\left[\dfrac{n}{2}\right]-1,

also are self-interlacing of type I (if jj is even) or of type II (if jj is odd).

Let again pp be a self-interlacing polynomial of type Iof degree nn and let qq be its associated Hurwitz stable polynomial, that is, p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) and q(z)=(1)n(n+1)2[p0(z2)zp1(z2)]q(z)=(-1)^{\tfrac{n(n+1)}{2}}[p_{0}(-z^{2})-zp_{1}(-z^{2})]. We are in a position to find an interrelation between Hurwitz minors of these polynomials.

Let n=2ln=2l. By Theorem 4.8, the Hurwitz matrix of the polynomial qq has the form

n(q)=(a1a3a5a70a0a2a4a600a1a3a500a0a2a400000(1)[n2]an).\mathcal{H}_{n}(q)=\begin{pmatrix}a_{1}&-a_{3}&a_{5}&-a_{7}&\dots&0\\ a_{0}&-a_{2}&a_{4}&-a_{6}&\dots&0\\ 0&a_{1}&-a_{3}&a_{5}&\dots&0\\ 0&a_{0}&-a_{2}&a_{4}&\dots&0\\ \dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&(-1)^{\left[\frac{n}{2}\right]}a_{n}\\ \end{pmatrix}.

It is easy to see that the matrix n(q)\mathcal{H}_{n}(q) can be factorized as follows

n(q)=C~nn(p)E~n,\mathcal{H}_{n}(q)=\widetilde{C}_{n}\mathcal{H}_{n}(p)\widetilde{E}_{n}, (4.28)

where the n×nn\times n matrices E~n\widetilde{E}_{n} and C~n\widetilde{C}_{n} have the forms

E~n=(1000001000001000001000001),C~n=(1000001000001000001000001).\widetilde{E}_{n}=\begin{pmatrix}1&0&0&0&0&\dots\\ 0&-1&0&0&0&\dots\\ 0&0&1&0&0&\dots\\ 0&0&0&-1&0&\dots\\ 0&0&0&0&1&\dots\\ \vdots&\vdots&\vdots&\vdots&\vdots&\ddots\end{pmatrix},\qquad\widetilde{C}_{n}=\begin{pmatrix}1&0&0&0&0&\dots\\ 0&1&0&0&0&\dots\\ 0&0&-1&0&0&\dots\\ 0&0&0&-1&0&\dots\\ 0&0&0&0&1&\dots\\ \vdots&\vdots&\vdots&\vdots&\vdots&\ddots\end{pmatrix}.

All non-principal minors of these matrices are equal to zero. So since e~jj=(1)j1\widetilde{e}_{jj}=(-1)^{j-1} and c~jj=(1)(j1)(j2)2\widetilde{c}_{jj}=(-1)^{\tfrac{(j-1)(j-2)}{2}}, we have

C~n(i1i2imi1i2im)=(1)k=1m(ik1)(ik2)2,E~n(i1i2imi1i2im)=(1)k=1mikm,\widetilde{C}_{n}\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ i_{1}&i_{2}&\dots&i_{m}\end{pmatrix}=(-1)^{\sum\limits_{k=1}^{m}\tfrac{(i_{k}-1)(i_{k}-2)}{2}},\qquad\widetilde{E}_{n}\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ i_{1}&i_{2}&\dots&i_{m}\end{pmatrix}=(-1)^{\sum\limits_{k=1}^{m}i_{k}-m}, (4.29)

where 1i1<i2<<imn1\leqslant i_{1}<i_{2}<\ldots<i_{m}\leqslant n. Thus, the Cauchy–Binet formula together with (4.29) and (4.28) implies, for n=2ln=2l,

n(q)(i1i2imj1j2jm)=(1)k=1m(ik1)(ik2)2+k=1mjkmn(p)(i1i2imj1j2jm).\mathcal{H}_{n}(q)\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ j_{1}&j_{2}&\dots&j_{m}\end{pmatrix}=(-1)^{\sum\limits_{k=1}^{m}\tfrac{(i_{k}-1)(i_{k}-2)}{2}+\sum\limits_{k=1}^{m}j_{k}-m}\mathcal{H}_{n}(p)\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ j_{1}&j_{2}&\dots&j_{m}\end{pmatrix}. (4.30)

where 1i1<i2<<imj1<j2<<jmn1\leqslant\begin{array}[]{c}i_{1}<i_{2}<\ldots<i_{m}\\ j_{1}<j_{2}<\ldots<j_{m}\end{array}\leqslant n,

Let now n=2l+1n=2l+1. Then by Theorem 4.8, the Hurwitz matrix of the polynomial qq has the form

n(q)=(a1a3a5a70a0a2a4a600a1a3a500a0a2a400000(1)[n+12]an).\mathcal{H}_{n}(q)=\begin{pmatrix}-a_{1}&a_{3}&-a_{5}&a_{7}&\dots&0\\ a_{0}&-a_{2}&a_{4}&-a_{6}&\dots&0\\ 0&-a_{1}&a_{3}&-a_{5}&\dots&0\\ 0&a_{0}&-a_{2}&a_{4}&\dots&0\\ \dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&(-1)^{\left[\frac{n+1}{2}\right]}a_{n}\\ \end{pmatrix}.

In this case, n(q)\mathcal{H}_{n}(q) also can be factorized:

n(q)=C^nn(p)(E~n),\mathcal{H}_{n}(q)=\widehat{C}_{n}\mathcal{H}_{n}(p)(-\widetilde{E}_{n}), (4.31)

where the n×nn\times n matrix C^n\widehat{C}_{n} is as follows

C^n=(100000010000001000000100000010000001).\widehat{C}_{n}=\begin{pmatrix}1&0&0&0&0&0&\dots\\ 0&-1&0&0&0&0&\dots\\ 0&0&-1&0&0&0&\dots\\ 0&0&0&1&0&0&\dots\\ 0&0&0&0&1&0&\dots\\ 0&0&0&0&0&-1&\dots\\ \vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\ddots\end{pmatrix}.

All non-principal minors of the matrices C^n\widehat{C}_{n} and E~n-\widetilde{E}_{n} equal zero. The principal minors of these matrices can be easily calculated:

C^n(i1i2imi1i2im)=(1)k=1mik(ik1)2,E^n(i1i2imi1i2im)=(1)k=1mik,\widehat{C}_{n}\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ i_{1}&i_{2}&\dots&i_{m}\end{pmatrix}=(-1)^{\sum\limits_{k=1}^{m}\tfrac{i_{k}(i_{k}-1)}{2}},\qquad\widehat{E}_{n}\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ i_{1}&i_{2}&\dots&i_{m}\\ \end{pmatrix}=(-1)^{\sum\limits_{k=1}^{m}i_{k}}, (4.32)

where 1i1<i2<<imn1\leqslant i_{1}<i_{2}<\ldots<i_{m}\leqslant n. Thus, the Cauchy–Binet formula together with (4.32) and (4.31) implies, for n=2l+1n=2l+1,

n(q)(i1i2imj1j2jm)=(1)k=1mik(ik1)2+k=1mjkn(p)(i1i2imj1j2jm),\mathcal{H}_{n}(q)\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ j_{1}&j_{2}&\dots&j_{m}\end{pmatrix}=(-1)^{\sum\limits_{k=1}^{m}\tfrac{i_{k}(i_{k}-1)}{2}+\sum\limits_{k=1}^{m}j_{k}}\mathcal{H}_{n}(p)\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ j_{1}&j_{2}&\dots&j_{m}\end{pmatrix}, (4.33)

where 1i1<i2<<imj1<j2<<jmn1\leqslant\begin{array}[]{c}i_{1}<i_{2}<\ldots<i_{m}\\ j_{1}<j_{2}<\ldots<j_{m}\end{array}\leqslant n.

Since qq is Hurwitz stable, from Theorem 3.5 and from the formulæ (4.30) and (4.33) we obtain the following theorem.

Theorem 4.18.

Let pp be a self-interlacing polynomial of type I of degree nn and let n(p)\mathcal{H}_{n}(p) be its Hurwitz matrix defined in (2.10). Then

  • for n=2ln=2l,

    (1)k=1m(ik1)(ik2)2+k=1mjkmn(p)(i1i2imj1j2jm)0;(-1)^{\sum\limits_{k=1}^{m}\tfrac{(i_{k}-1)(i_{k}-2)}{2}+\sum\limits_{k=1}^{m}j_{k}-m}\mathcal{H}_{n}(p)\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ j_{1}&j_{2}&\dots&j_{m}\end{pmatrix}\geqslant 0;
  • for n=2l+1n=2l+1,

    (1)k=1mik(ik1)2+k=1mjkn(p)(i1i2imj1j2jm)0,(-1)^{\sum\limits_{k=1}^{m}\tfrac{i_{k}(i_{k}-1)}{2}+\sum\limits_{k=1}^{m}j_{k}}\mathcal{H}_{n}(p)\begin{pmatrix}i_{1}&i_{2}&\dots&i_{m}\\ j_{1}&j_{2}&\dots&j_{m}\end{pmatrix}\geqslant 0,

where 1i1<i2<<imj1<j2<<jmn1\leqslant\begin{array}[]{c}i_{1}<i_{2}<\ldots<i_{m}\\ j_{1}<j_{2}<\ldots<j_{m}\end{array}\leqslant n.


4.4 The second proof of the Hurwitz self-interlacing criterion

Given a polynomial

p(z)=a0zn+a1zn1++an,a1,,an,a0>0,p(z)=a_{0}z^{n}+a_{1}z^{n-1}+\dots+a_{n},\qquad a_{1},\dots,a_{n}\in\mathbb{R},\ a_{0}>0, (4.34)

we consider the following rational functions

R(z)=(1)np(z)p(z)andF(z)=1R(z)=p(z)(1)np(z)R(z)=\dfrac{(-1)^{n}p(-z)}{p(z)}\quad\text{and}\quad F(z)=\dfrac{1}{R(z)}=\dfrac{p(z)}{(-1)^{n}p(-z)} (4.35)

and expand them into their Laurent series at \infty:

R(z)=1+s0z+s1z2+s2z3+,F(z)=1+t0z+t1z2+t2z3+R(z)=1+\frac{s_{0}}{z}+\frac{s_{1}}{z^{2}}+\frac{s_{2}}{z^{3}}+\dots,\qquad F(z)=1+\frac{t_{0}}{z}+\frac{t_{1}}{z^{2}}+\frac{t_{2}}{z^{3}}+\dots

As we mentioned in Section 1.1, ranks of the matrices S=sj+k0S=\|s_{j+k}\|_{0}^{\infty} and T=tj+k0T=\|t_{j+k}\|_{0}^{\infty} are equal to the number of poles of the functions RR and FF, respectively. It is easy to see that rank of each matrix equals nn if the polynomials p(z)p(z) and p(z)p(-z) have no common zeroes. In the rest of this section we consider only such polynomials, so in this section, ranks of the matrices SS and TT always equal nn. Denoting

q(z)=(1)np(z),q(z)=(-1)^{n}p(-z),

we obtain R=qpR=\dfrac{q}{p} and F=pqF=\dfrac{p}{q}.

Lemma 4.19.

For the functions RR and FF defined in (4.35), the following formulæ are valid:

2j(p,q)=a02jDj(R)=(1)j(j+1)22ja0Δj1(p)Δj(p),j=1,2,\nabla_{2j}(p,q)=a_{0}^{2j}D_{j}(R)=(-1)^{\frac{j(j+1)}{2}}2^{j}a_{0}\Delta_{j-1}(p)\Delta_{j}(p),\quad j=1,2,\ldots (4.36)

and

2j(q,p)=a02jDj(F)=(1)j(j1)22ja0Δj1(p)Δj(p),j=1,2,,\nabla_{2j}(q,p)=a_{0}^{2j}D_{j}(F)=(-1)^{\frac{j(j-1)}{2}}2^{j}a_{0}\Delta_{j-1}(p)\Delta_{j}(p),\quad j=1,2,\ldots, (4.37)

where the determinants 2j(p,q)\nabla_{2j}(p,q) are the leading principal minors of the matrix 2n+1(p,q)\mathcal{H}_{2n+1}(p,q) of order 2j2j, and Δj(p)\Delta_{j}(p) are defined in (2.11), Δ0(p)1\Delta_{0}(p)\equiv 1.

Proof.

We prove the formulæ (4.36). The formulæ (4.37) follow from (4.36) since 2j(q,p)=(1)j2j(p,q)\nabla_{2j}(q,p)=(-1)^{j}\nabla_{2j}(p,q).

Assuming aj=0a_{j}=0 for j>nj>n, we have

2j(p,q)=|a0a1a2a3aj1aja2j2a2j1a0a1a2a3aj1aja2j2a2j10a0a1a2aj2aj1a2j3a2j20a0a1a2aj2aj1a2j3a2j200a0a1aj3aj2a2j4a2j300a0a1aj3aj2a2j4a2j30000a0a1aj1aj0000a0a1aj1aj|=\nabla_{2j}(p,q)=\begin{vmatrix}a_{0}&a_{1}&a_{2}&a_{3}&\dots&a_{j-1}&a_{j}&\dots&a_{2j-2}&a_{2j-1}\\ a_{0}&-a_{1}&a_{2}&-a_{3}&\dots&-a_{j-1}&a_{j}&\dots&a_{2j-2}&-a_{2j-1}\\ 0&a_{0}&a_{1}&a_{2}&\dots&a_{j-2}&a_{j-1}&\dots&a_{2j-3}&a_{2j-2}\\ 0&a_{0}&-a_{1}&a_{2}&\dots&a_{j-2}&-a_{j-1}&\dots&-a_{2j-3}&a_{2j-2}\\ 0&0&a_{0}&a_{1}&\dots&a_{j-3}&a_{j-2}&\dots&a_{2j-4}&a_{2j-3}\\ 0&0&a_{0}&-a_{1}&\dots&-a_{j-3}&a_{j-2}&\dots&a_{2j-4}&-a_{2j-3}\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&a_{0}&a_{1}&\dots&a_{j-1}&a_{j}\\ 0&0&0&0&\dots&a_{0}&-a_{1}&\dots&-a_{j-1}&a_{j}\\ \end{vmatrix}=
=2j|a0a1a2a3aj1aja2j2a2j1a00a200aja2j200a0a1a2aj2aj1a2j3a2j20a000aj200a2j200a0a1aj3aj2a2j4a2j300a000aj2a2j400000a0a1aj1aj0000a000aj|==2^{j}\begin{vmatrix}a_{0}&a_{1}&a_{2}&a_{3}&\dots&a_{j-1}&a_{j}&\dots&a_{2j-2}&a_{2j-1}\\ a_{0}&0&a_{2}&0&\dots&0&a_{j}&\dots&a_{2j-2}&0\\ 0&a_{0}&a_{1}&a_{2}&\dots&a_{j-2}&a_{j-1}&\dots&a_{2j-3}&a_{2j-2}\\ 0&a_{0}&0&0&\dots&a_{j-2}&0&\dots&0&a_{2j-2}\\ 0&0&a_{0}&a_{1}&\dots&a_{j-3}&a_{j-2}&\dots&a_{2j-4}&a_{2j-3}\\ 0&0&a_{0}&0&\dots&0&a_{j-2}&\dots&a_{2j-4}&0\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&a_{0}&a_{1}&\dots&a_{j-1}&a_{j}\\ 0&0&0&0&\dots&a_{0}&0&\dots&0&a_{j}\\ \end{vmatrix}=
=(2)j|a00a200aja2j200a10a3aj100a2j10a00a2aj200a2j200a100aj1a2j3000a000aj2a2j40000a1aj300a2j30000a000aj00000a1aj10|==(-2)^{j}\begin{vmatrix}a_{0}&0&a_{2}&0&\dots&0&a_{j}&\dots&a_{2j-2}&0\\ 0&a_{1}&0&a_{3}&\dots&a_{j-1}&0&\dots&0&a_{2j-1}\\ 0&a_{0}&0&a_{2}&\dots&a_{j-2}&0&\dots&0&a_{2j-2}\\ 0&0&a_{1}&0&\dots&0&a_{j-1}&\dots&a_{2j-3}&0\\ 0&0&a_{0}&0&\dots&0&a_{j-2}&\dots&a_{2j-4}&0\\ 0&0&0&a_{1}&\dots&a_{j-3}&0&\dots&0&a_{2j-3}\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&a_{0}&0&\dots&0&a_{j}\\ 0&0&0&0&\dots&0&a_{1}&\dots&a_{j-1}&0\\ \end{vmatrix}=
=(2)ja0|a10a30aj100a2j1a00a20aj200a2j200a10aj300a2j300a00aj400a2j40000a100aj+10000a000aj0a10a30aj1a2j300a00a20aj2a2j4000000a2aj000000a1aj10|==(-2)^{j}a_{0}\begin{vmatrix}a_{1}&0&a_{3}&0&\dots&a_{j-1}&0&\dots&0&a_{2j-1}\\ a_{0}&0&a_{2}&0&\dots&a_{j-2}&0&\dots&0&a_{2j-2}\\ 0&0&a_{1}&0&\dots&a_{j-3}&0&\dots&0&a_{2j-3}\\ 0&0&a_{0}&0&\dots&a_{j-4}&0&\dots&0&a_{2j-4}\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&a_{1}&0&\dots&0&a_{j+1}\\ 0&0&0&0&\dots&a_{0}&0&\dots&0&a_{j}\\ 0&a_{1}&0&a_{3}&\dots&0&a_{j-1}&\dots&a_{2j-3}&0\\ 0&a_{0}&0&a_{2}&\dots&0&a_{j-2}&\dots&a_{2j-4}&0\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&0&a_{2}&\dots&a_{j}&0\\ 0&0&0&0&\dots&0&a_{1}&\dots&a_{j-1}&0\\ \end{vmatrix}=
=(2)ja0(1)j(j1)2|a1a3a5a2j10000a0a2a4a2j200000a1a3a2j300000a0a2a2j40000000aj+10000000aj00000000a1a3a2j5a2j30000a0a2a2j6a2j4000000aj2aj000000aj3aj1|==(-2)^{j}a_{0}(-1)^{\frac{j(j-1)}{2}}\begin{vmatrix}a_{1}&a_{3}&a_{5}&\dots&a_{2j-1}&0&0&\dots&0&0\\ a_{0}&a_{2}&a_{4}&\dots&a_{2j-2}&0&0&\dots&0&0\\ 0&a_{1}&a_{3}&\dots&a_{2j-3}&0&0&\dots&0&0\\ 0&a_{0}&a_{2}&\dots&a_{2j-4}&0&0&\dots&0&0\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&\dots&a_{j+1}&0&0&\dots&0&0\\ 0&0&0&\dots&a_{j}&0&0&\dots&0&0\\ 0&0&0&\dots&0&a_{1}&a_{3}&\dots&a_{2j-5}&a_{2j-3}\\ 0&0&0&\dots&0&a_{0}&a_{2}&\dots&a_{2j-6}&a_{2j-4}\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&\dots&0&0&0&\dots&a_{j-2}&a_{j}\\ 0&0&0&\dots&0&0&0&\dots&a_{j-3}&a_{j-1}\\ \end{vmatrix}=
=(1)j(j+1)22ja0Δj1(p)Δj(p).=(-1)^{\frac{j(j+1)}{2}}2^{j}a_{0}\Delta_{j-1}(p)\Delta_{j}(p).

Using this lemma it is easy to prove the equivalence of the conditions 1)1) and 2)2) of Theorem 4.7.

Theorem 4.20.

The polynomial pp defined in (4.34) is self-interlacing of type I if and only if

Δn1(p)>0,Δn3(p)>0,,\Delta_{n-1}(p)>0,\ \Delta_{n-3}(p)>0,\dots, (4.38)
(1)[n+12]Δn(p)>0,(1)[n+12]1Δn2(p)>0,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]}\Delta_{n}(p)>0,\displaystyle(-1)^{\left[\tfrac{n+1}{2}\right]-1}\Delta_{n-2}(p)>0,\ldots (4.39)

where the Hurwitz minors Δi(p)\Delta_{i}(p) are defined in (2.11).

Proof.

In the proof of Theorem 4.3, it was established that if pp is self-interlacing of type I, then the function p(z)p(z)-\dfrac{p(-z)}{p(z)} is an R-function with exactly nn poles.

Let n=2l+1n=2l+1. Then the function RR defined in (4.35) is an R-function. By Theorem 1.18, we have Dj(R)>0D_{j}(R)>0 for j=1,,nj=1,\ldots,n. These inequalities together with the formulæ (4.36) imply

(1)j(j+1)2Δj1(p)Δj(p)>0,j=1,,n.(-1)^{\frac{j(j+1)}{2}}\Delta_{j-1}(p)\Delta_{j}(p)>0,\qquad j=1,\ldots,n. (4.40)

Multiplying the inequalities (4.40) for j=2mj=2m and j=2m1j=2m-1, we obtain

Δ2m12Δ2mΔ2m2>0.\Delta_{2m-1}^{2}\Delta_{2m}\Delta_{2m-2}>0.

Consequently, the minors Δ2i(p)\Delta_{2i}(p), i=1,,li=1,\ldots,l, are positive, so the inequalities (4.38) are proved for odd nn.

If we multiply the inequalities (4.40) for j=2mj=2m and j=2m+1j=2m+1, we get

Δ2m2(p)Δ2m1(p)Δ2m+1(p)>0.-\Delta_{2m}^{2}(p)\Delta_{2m-1}(p)\Delta_{2m+1}(p)>0.

These inequalities imply (4.39) for odd nn.

The converse assertion can be proved in the same way. That is, the inequalities (4.38)–(4.39) imply the inequalities (4.40) which in turn imply the inequalities Dj(R)>0D_{j}(R)>0, j=1,,nj=1,\ldots,n, according to (4.36). By Theorem 1.18, the function p(z)p(z)-\dfrac{p(-z)}{p(z)} is an R-function, so pp is a self-interlacing polynomial of type I (see the proof of Theorem 4.3).

If nn is even, the inequalities (4.38)–(4.39) can be proved analogously using the function FF defined in (4.35) instead of the function RR. ∎

From Lemma 4.19 and from Hurwitz’s stability criterion (see Theorem 3.5) it follows one more stability criterion.

Theorem 4.21.

  • 1)

    The polynomial pp is Hurwitz stable if and only if for the function RR defined in (4.35), the following inequalities hold

    (1)j(j+1)2Dj(R)>0,j=1,2,,n.(-1)^{\frac{j(j+1)}{2}}D_{j}(R)>0,\quad j=1,2,\ldots,n.
  • 2)

    The polynomial pp is Hurwitz stable if and only if for the function FF defined in (4.35), the following inequalities hold

    (1)j(j1)2Dj(F)>0,j=1,2,,n.(-1)^{\frac{j(j-1)}{2}}D_{j}(F)>0,\quad j=1,2,\ldots,n.

Finally, we establish a relationships between the Hurwitz minors Δj(p)\Delta_{j}(p) of a given polynomial pp and the Hankel minors D^j(R)\widehat{D}_{j}(R) and D^j(F)\widehat{D}_{j}(F) (for the definition of the minors D^j\widehat{D}_{j} see (1.8)). To do this, consider the functions

zR(z)=z+h1(z)p(z)=z+2a1zn2a3zn22a5zn4a0zn+a1zn1++an=z+s0+s1z+s2z2+s3z3+zR(z)=z+\dfrac{h_{1}(z)}{p(z)}=z+\dfrac{-2a_{1}z^{n}-2a_{3}z^{n-2}-2a_{5}z^{n-4}-\dots}{a_{0}z^{n}+a_{1}z^{n-1}+\dots+a_{n}}=z+s_{0}+\frac{s_{1}}{z}+\frac{s_{2}}{z^{2}}+\frac{s_{3}}{z^{3}}+\dots (4.41)
zF(z)=z+h2(z)(1)np(z)=z+2a1zn+2a3zn2+2a5zn4+a0zna1zn1++(1)nan=z+t0+t1z+t2z2+t3z3+zF(z)=z+\dfrac{h_{2}(z)}{(-1)^{n}p(-z)}=z+\dfrac{2a_{1}z^{n}+2a_{3}z^{n-2}+2a_{5}z^{n-4}+\dots}{a_{0}z^{n}-a_{1}z^{n-1}+\dots+(-1)^{n}a_{n}}=z+t_{0}+\frac{t_{1}}{z}+\frac{t_{2}}{z^{2}}+\frac{t_{3}}{z^{3}}+\dots (4.42)

where s0=t0=1s_{0}=t_{0}=1, and note that D^j(R)=Dj(zR)\widehat{D}_{j}(R)=D_{j}(zR) and D^j(F)=Dj(zF)\widehat{D}_{j}(F)=D_{j}(zF). This allows us to establish the following lemma.

Lemma 4.22.

Given the polynomial pp, for the functions RR and FF defined in (4.35), the following relationships hold

2j(p,h1)=a02jD^j(R)=(1)j(j1)22jΔj2(p),j=1,2,,\nabla_{2j}(p,h_{1})=a_{0}^{2j}\widehat{D}_{j}(R)=(-1)^{\frac{j(j-1)}{2}}2^{j}\Delta_{j}^{2}(p),\quad j=1,2,\ldots, (4.43)

and

2j(p,h2)=a02jD^j(F)=(1)j(j1)22jΔj2(p),j=1,2,,\nabla_{2j}(p,h_{2})=a_{0}^{2j}\widehat{D}_{j}(F)=(-1)^{\frac{j(j-1)}{2}}2^{j}\Delta_{j}^{2}(p),\quad j=1,2,\ldots, (4.44)

where the determinants121212The polynomials h1h_{1} and h2h_{2} are defined in (4.41) and (4.42), respectively. 2j(p,h1)\nabla_{2j}(p,h_{1}) and 2j(p,h2)\nabla_{2j}(p,h_{2}) are the leading principal minors of order 2j2j of the matrices 2n+1(p,h1)\mathcal{H}_{2n+1}(p,h_{1}) and 2n+1(p,h2)\mathcal{H}_{2n+1}(p,h_{2}) respectively, and Δj(p)\Delta_{j}(p) are the Hurwitz minors of the polynomial pp, Δ0(p)1\Delta_{0}(p)\equiv 1.

Proof.

Assuming aj=0a_{j}=0 for j>nj>n, from (4.41) we obtain

2j(p,h1)=(2)j|a0a1a2a3aj1aja2j2a2j1a10a300aj+1a2j100a0a1a2aj2aj1a2j3a2j20a10a3aj100a2j100a0a1aj3aj2a2j4a2j300a100aj1a2j300000a0a1aj1aj0000a100aj+1|=\nabla_{2j}(p,h_{1})=(-2)^{j}\begin{vmatrix}a_{0}&a_{1}&a_{2}&a_{3}&\dots&a_{j-1}&a_{j}&\dots&a_{2j-2}&a_{2j-1}\\ a_{1}&0&a_{3}&0&\dots&0&a_{j+1}&\dots&a_{2j-1}&0\\ 0&a_{0}&a_{1}&a_{2}&\dots&a_{j-2}&a_{j-1}&\dots&a_{2j-3}&a_{2j-2}\\ 0&a_{1}&0&a_{3}&\dots&a_{j-1}&0&\dots&0&a_{2j-1}\\ 0&0&a_{0}&a_{1}&\dots&a_{j-3}&a_{j-2}&\dots&a_{2j-4}&a_{2j-3}\\ 0&0&a_{1}&0&\dots&0&a_{j-1}&\dots&a_{2j-3}&0\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&a_{0}&a_{1}&\dots&a_{j-1}&a_{j}\\ 0&0&0&0&\dots&a_{1}&0&\dots&0&a_{j+1}\\ \end{vmatrix}=
=2j|a10a300aj+1a2j10a00a200aja2j200a10a3aj100a2j10a00a2aj200a2j200a100aj1a2j3000a000aj2a2j4000000a2aj00000a100aj+10000a0a1aj1aj|==2^{j}\begin{vmatrix}a_{1}&0&a_{3}&0&\dots&0&a_{j+1}&\dots&a_{2j-1}&0\\ a_{0}&0&a_{2}&0&\dots&0&a_{j}&\dots&a_{2j-2}&0\\ 0&a_{1}&0&a_{3}&\dots&a_{j-1}&0&\dots&0&a_{2j-1}\\ 0&a_{0}&0&a_{2}&\dots&a_{j-2}&0&\dots&0&a_{2j-2}\\ 0&0&a_{1}&0&\dots&0&a_{j-1}&\dots&a_{2j-3}&0\\ 0&0&a_{0}&0&\dots&0&a_{j-2}&\dots&a_{2j-4}&0\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&0&\dots&0&a_{2}&\dots&a_{j}&0\\ 0&0&0&0&\dots&a_{1}&0&\dots&0&a_{j+1}\\ 0&0&0&0&\dots&a_{0}&a_{1}&\dots&a_{j-1}&a_{j}\\ \end{vmatrix}=
=2j(1)j(j1)2|a1a3a5a2j10000a0a2a4a2j200000a1a3a2j300000a0a2a2j40000000aj+10000000aj00000000a1a3a2j3a2j10000a0a2a2j4a2j2000000aj1aj+1000aj100aj2aj|==2^{j}(-1)^{\frac{j(j-1)}{2}}\begin{vmatrix}a_{1}&a_{3}&a_{5}&\dots&a_{2j-1}&0&0&\dots&0&0\\ a_{0}&a_{2}&a_{4}&\dots&a_{2j-2}&0&0&\dots&0&0\\ 0&a_{1}&a_{3}&\dots&a_{2j-3}&0&0&\dots&0&0\\ 0&a_{0}&a_{2}&\dots&a_{2j-4}&0&0&\dots&0&0\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&\dots&a_{j+1}&0&0&\dots&0&0\\ 0&0&0&\dots&a_{j}&0&0&\dots&0&0\\ 0&0&0&\dots&0&a_{1}&a_{3}&\dots&a_{2j-3}&a_{2j-1}\\ 0&0&0&\dots&0&a_{0}&a_{2}&\dots&a_{2j-4}&a_{2j-2}\\ \dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots&\dots\\ 0&0&0&\dots&0&0&0&\dots&a_{j-1}&a_{j+1}\\ 0&0&0&\dots&a_{j-1}&0&0&\dots&a_{j-2}&a_{j}\\ \end{vmatrix}=
=(1)j(j1)22jΔj2(p).=(-1)^{\frac{j(j-1)}{2}}2^{j}\Delta_{j}^{2}(p).

The relationships (4.44) can be proved analogously. ∎

Thus, we obtain that for any real polynomial pp of degree nn with nonzero Hurwitz minors and, in particular, for Hurwitz stable and for self-interlacing polynomials, the following inequalities hold

(1)j(j1)2D^j(R)=(1)j(j1)2D^j(F)>0,j=1,2,,n.(-1)^{\frac{j(j-1)}{2}}\widehat{D}_{j}(R)=(-1)^{\frac{j(j-1)}{2}}\widehat{D}_{j}(F)>0,\quad j=1,2,\ldots,n.

Last, let us note that the formulæ (4.36)–(4.37) and (4.43)–(4.44) can be obtained (overcoming certain difficulties) from some theorems of the book [41]. But it was more simple to deduce them directly as we did in Lemmata 4.19 and 4.22.

4.5 Almost and quasi- self-interlacing polynomials

In this section, we describe polynomials, which are dual (in the sense of Theorem 4.8) to the quasi-stable polynomials. Because of the mentioned duality we just give the definition of these polynomials.

Definition 4.23.

A polynomial p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) of degree nn is called quasi-self-interlacing of type I (of type II) with degeneracy index mm, 0mn0\leqslant m\leqslant n, if the polynomial p0(z2)zp1(z2)p_{0}(-z^{2})-zp_{1}(-z^{2}) (resp. the polynomial p0(z2)+zp1(z2)p_{0}(-z^{2})+zp_{1}(-z^{2})) is quasi-stable with degeneracy index mm.

In other words, the polynomial pp is quasi-self-interlacing if non-common zeroes f the polynomials p(z)p(z) and p(z)p(-z) are real, simple and interlacing.

Obviously, all results regarding quasi-stable polynomials can be easily reformulated for the quasi-stable polynomials. So we leave such reformulation to the reader.

In Section 5 we use the quasi-stable polynomials with degeneracy index 11, that is, the polynomials of the form p(z)=zq(z)p(z)=zq(z), where q(z)q(z) is a self-interlacing polynomial. Such polynomials we call almost self-interlacing polynomials. Note that if p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) is almost self-interlacing, then the function Φ=p1p0\Phi=\dfrac{p_{1}}{p_{0}} is an R-function with all poles positive except one, which is zero.

4.6 ”Strange” polynomials

This section is devoted to short description of some numeric results on the distribution of zeroes of some polynomials closely connected to the Hurwitz stable and self-interlacing polynomials. We call these polynomials ”strange” because of their curious zero location.

Let p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) be a Hurwitz stable polynomial of degree nn. Consider the polynomial q(z)=p0(z2)+zp1(z2)q(z)=p_{0}(-z^{2})+zp_{1}(z^{2}). It is evident that q(0)=p(0)0q(0)=p(0)\neq 0. Calculations with Maple Software showed that the polynomial qq has exactly [n+12]\left[\dfrac{n+1}{2}\right] simple zeroes in the open right half-plane and exactly [n2]\left[\dfrac{n}{2}\right] simple zeroes in the open left half-plane, but it has no zeroes on the imaginary axis. At least one zero of qq is nonreal.

Let us denote by μ1,μ2,\mu_{1},\mu_{2},\ldots the distinct absolute values of the zeroes of the polynomial qq lying in the open left half-plane such that

0<μ1<μ2<μ3<0<\mu_{1}<\mu_{2}<\mu_{3}<\ldots

And denote by λ1,λ2,\lambda_{1},\lambda_{2},\ldots the distinct absolute values of the zeroes of the polynomial qq lying in the open right half-plane such that

0<λ1<λ2<λ3<0<\lambda_{1}<\lambda_{2}<\lambda_{3}<\ldots

So numerical experiments showed that

0<λ1<μ1<λ2<μ2<λ3<0<\lambda_{1}<\mu_{1}<\lambda_{2}<\mu_{2}<\lambda_{3}<\ldots

The polynomial p0(z2)+zp1(z2)p_{0}(z^{2})+zp_{1}(-z^{2}) possesses similar properties according to calculations.

However, it is not known for sure if all the polynomials of the type p0(z2)+zp1(z2)p_{0}(-z^{2})+zp_{1}(z^{2}), where the polynomial p0(z2)+zp1(z2)p_{0}(z^{2})+zp_{1}(z^{2}) is Hurwitz stable, have the zero location described above.

4.7 Matrices with self-interlacing spectrum

In this section, we consider some classes of real matrices with self-interlacing spectrum and develop a method of constructing such kind of matrices from a given totally positive matrix. But at first, we recall some definitions and statements from the book [10].

Definition 4.24 ([10]).

A square matrix A=aij1nA=\|a_{ij}\|_{1}^{n} is called sign definite of class nn if for any knk\leqslant n, all the non-zero minors of order kk have the same sign εk\varepsilon_{k}. The sequence {ε1,ε2,,εn}\{\varepsilon_{1},\varepsilon_{2},\ldots,\varepsilon_{n}\} is called the signature sequence of the matrix AA.

A sign definite matrix of class nn is called strictly sign definite of class nn if all its minors are different from zero.

Definition 4.25 ([10]).

A square sign definite matrix A=aij1nA=\|a_{ij}\|_{1}^{n} of class nn is called the matrix of class n+n^{+} if some its power is a strictly sign definite matrix of class nn.

Note that a sign definite (strictly sign definite) matrix of class nn with the signature sequence ε1=ε2==εn=1\varepsilon_{1}=\varepsilon_{2}=\ldots=\varepsilon_{n}=1 is totally nonnegative (strictly totally positive). Also a sign definite matrix of class n+n^{+} with the signature sequence ε1=ε2==εn=1\varepsilon_{1}=\varepsilon_{2}=\ldots=\varepsilon_{n}=1 is an oscillating matrix (see [10]), that is, a totally nonnegative matrix whose certain power is strictly totally positive. It is clear from the Binet-Cauchy formula that the square of a sign definite matrix is totally nonnegative.

In [10] it was established the following lemma.

Lemma 4.26.

Let the matrix A=aij1nA=\|a_{ij}\|_{1}^{n} be totally nonnegative. Then the matrices B=ani+1,j1nB=\|a_{n-i+1,j}\|_{1}^{n} and C=ai,nj+11nC=\|a_{i,n-j+1}\|_{1}^{n} are sign definite of class nn. Moreover, the signature sequence of the matrices BB and CC is as follows:

εk=(1)k(k1)2,k=1,2,,n.\varepsilon_{k}=(-1)^{\tfrac{k(k-1)}{2}},\quad k=1,2,\ldots,n. (4.45)

Note that the matrices BB and CC can be represented as follows

B=JA,andC=AJ,B=JA,\qquad\text{and}\qquad C=AJ,

where

J=(0001001001001000).J=\begin{pmatrix}0&0&\dots&0&1\\ 0&0&\dots&1&0\\ \vdots&\vdots&\cdot&\vdots&\vdots\\ 0&1&\dots&0&0\\ 1&0&\dots&0&0\\ \end{pmatrix}. (4.46)

It is easy to see that the matrix JJ is sign definite of class nn (but not of class n+n^{+}) with the signature sequence of the form (4.45). So by the Binet-Cauchy formula [11] we obtain the following statement.

Theorem 4.27.

The matrix A=aij1nA=\|a_{ij}\|_{1}^{n} is totally nonnegative if and only if the matrix JAJA (or the matrix AJAJ) is sign definite of class nn with the signature sequence (4.45). The matrix JJ is defined in (4.46).

Obviously, the converse statement is also true.

Theorem 4.28.

The matrix A=aij1nA=\|a_{ij}\|_{1}^{n} is a sign definite matrix of class nn with the signature sequence (4.45) if and only if the matrix JAJA (or the matrix AJAJ) is totally nonnegative.

In the sequel, we need the following two theorems established in the book [10].

Theorem 4.29.

Let the matrix A=aij1nA=\|a_{ij}\|_{1}^{n} be a sign definite of class n+n^{+} with the signature sequence εk\varepsilon_{k}, k=1,2,,nk=1,2,\ldots,n. Then all the eigenvalues λk\lambda_{k}, k=1,2,,nk=1,2,\ldots,n, of the matrix AA are nonzero real and simple, and if

|λ1|>|λ2|>>|λn|>0,|\lambda_{1}|>|\lambda_{2}|>\ldots>|\lambda_{n}|>0, (4.47)

then

signλk=εkεk1,k=1,2,,n,ε0=1.\emph{sign}\,\lambda_{k}=\dfrac{\varepsilon_{k}}{\varepsilon_{k-1}},\qquad k=1,2,\ldots,n,\,\varepsilon_{0}=1. (4.48)
Theorem 4.30.

A totally nonnegative matrix A=aij1nA=\|a_{ij}\|_{1}^{n} is oscillating if and only if AA is nonsingular and the following inequalities hold

aj,j+1>0,andaj+1,j>0,j=1,2,,n1.a_{j,j+1}>0,\qquad\text{and}\qquad a_{j+1,j}>0,\qquad j=1,2,\ldots,n-1.

Let us introduce the following definition.

Definition 4.31.

A matrix AA is said to have the self-interlacing spectrum if its eigenvalues are real and simple and satisfy the following inequalities

λ1>λ2>λ3>>(1)n1λn>0,\lambda_{1}>-\lambda_{2}>\lambda_{3}>\ldots>(-1)^{n-1}\lambda_{n}>0, (4.49)

or

λ1>λ2>λ3>>(1)nλn>0.-\lambda_{1}>\lambda_{2}>-\lambda_{3}>\ldots>(-1)^{n}\lambda_{n}>0. (4.50)

Now we are in a position to complement Lemma 4.26.

Theorem 4.32.

Let all the entries of a nonsingular matrix A=aij1nA=\|a_{ij}\|_{1}^{n} be nonnegative131313The matrices with nonnegative entries are usually called nonnegative matrices. and for each ii, i=1,2,,n1i=1,2,\ldots,n-1, there exist number r1r_{1} and r2r_{2}, 1r1,r2n1\leqslant r_{1},r_{2}\leqslant n, such that

ani,r1an+1r1,i>0,an+1i,r2an+1r2,i+1>0a_{n-i,r_{1}}\cdot a_{n+1-r_{1},i}>0,a_{n+1-i,r_{2}}\cdot a_{n+1-r_{2},i+1}>0 (4.51)
(orai,n+1r1ar1,ni>0,ai+1,n+1r2ar2,n+1i>0).(\text{or}\qquad a_{i,n+1-r_{1}}\cdot a_{r_{1},n-i}>0,a_{i+1,n+1-r_{2}}\cdot a_{r_{2},n+1-i}>0). (4.52)

The matrix AA is totally nonnegative if and only if the matrix B=JA=bij1nB=JA=\|b_{ij}\|_{1}^{n} (or, respectively, the matrix C=AJC=AJ) is sign definite of class n+n^{+} with the signature sequence defined in (4.45). Moreover, the matrix BB (or, respectively, the matrix CC) possesses a self-interlacing spectrum of the form (4.49).

Proof.

We prove the theorem in the case when the condition (4.51) holds. The case of the condition (4.52) can be established analogously.

Let AA be a nonsingular totally nonnegative matrix and the condition (4.51) holds. From Theorem 4.27 it follows that the matrix B=JAB=JA is sign definite of class nn with the signature sequence (4.45). In order to the matrix be be sign definite of class n+n^{+} it is necessary and sufficient that a certain power of this matrix BB be strictly sign definite of class nn. Since the entries of the matrix JJ have the form

(J)ij={1,i=n+1j;0,in+1j;(J)_{ij}=\begin{cases}&1,\qquad i=n+1-j;\\ &0,\qquad i\neq n+1-j;\end{cases}

the entries of the matrix BB can be represented as follows:

bij=k=1n(J)ikakj=an+1i,j.b_{ij}=\sum\limits_{k=1}^{n}(J)_{ik}a_{kj}=a_{n+1-i,j}.

Consider the totally nonnegative matrix B2B^{2}. Its entries have the form

(B2)ij=k=1nbikbkj=k=1nan+1i,kan+1k,j.(B^{2})_{ij}=\sum\limits_{k=1}^{n}b_{ik}b_{kj}=\sum\limits_{k=1}^{n}a_{n+1-i,k}a_{n+1-k,j}.

From these formulæ and from (4.51) it follows that all the entries of the matrix B2B^{2} above and under the main diagonal are positive, that is, (B2)i,i+1>0(B^{2})_{i,i+1}>0 (B2)i+1,i>0(B^{2})_{i+1,i}>0, i=1,2,,n1i=1,2,\ldots,n-1. By Theorem 4.30, B2B^{2} is an oscillating matrix. According to the definition of oscillating matrices, a certain power of B2B^{2} is strictly totally positive. Thus, we proved that a certain power of the matrix BB is strictly sign definite, so BB is a sign definite matrix of class n+n^{+} with the signature sequence (4.45) according to Lemma 4.26. By Theorem 4.29, all eigenvalues of the matrix BB are nonzero real and simple. Moreover, if we enumerate the eigenvalues in order of decreasing absolute values as in (4.47), then from (4.48) and (4.45) we obtain that the spectrum of BB is of the form (4.49).

The converse assertion of the theorem follows from Theorem 4.27. ∎

Remark 4.33.

For n=2l+1n=2l+1 (for n=2ln=2l), the characteristic polynomial of the matrix BB described in Theorem 4.32 is self-interlacing of type I (of type II).

Remark 4.34.

If the matrix A-A is totally nonnegative and the conditions (4.51) (or the conditions (4.52)) hold, then the matrix B=JAB=JA (or, respectively, C=AJC=AJ) has a spectrum of the form (4.50).

Remark 4.35.

One can obtain another types of totally nonnegative matrices which result matrices with self-interlacing spectra after multiplication by the matrix JJ. To do this we need to change the conditions (4.51)–(4.52) by another ones such that, for instance, the matrix B4B^{4}, or B6B^{6}, (or B8B^{8} etc.) becomes oscillating.

Theorem 4.32 implies the following corollary.

Corollary 4.36.

A nonsingular matrix AA with positive entries is totally nonnegative if and only if the matrix B=JAB=JA (or the matrix C=AJC=AJ) is sign definite of class n+n^{+} with the signature sequence (4.45). Moreover, the matrix BB has a self-interlacing spectrum of the form (4.49).

Consider a particular case of conditions (4.51). Suppose that all diagonal entries if the matrix AA be positive ajj>0a_{jj}>0, j=1,2,,nj=1,2,\ldots,n. It is easy to see that in this case the conditions (4.51) hold if aj,j+1>0a_{j,j+1}>0, j=1,2,,n1j=1,2,\ldots,n-1. If all remaining entries of the matrix AA are nonnegative, then by Theorem 4.32 the matrices B=JAB=JA and C=AJC=AJ have self-interlacing spectra if AA is totally nonnegative.

If all other entries of the matrix AA (that is, all entries except ajja_{jj} and aj,j+1a_{j,j+1}, which are positive) equal zero, then AA is a bidiagonal matrix with positive entries on and under the main diagonal. Clearly, AA is totally nonnegative. Then by Theorem 4.32 the matrix B=JAB=JA has a self-interlacing spectrum. Thus we obtain the following statement established in [15]:

Theorem 4.37.

Any anti-bidiagonal n×nn\times n matrix with positive entries

B=(00000bn0000bn2bn1000bn4bn3000cn40000cn2cn3000cncn10000)B=\begin{pmatrix}0&0&0&\dots&0&0&b_{n}\\ 0&0&0&\dots&0&b_{n-2}&b_{n-1}\\ 0&0&0&\dots&b_{n-4}&b_{n-3}&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&c_{n-4}&\dots&0&0&0\\ 0&c_{n-2}&c_{n-3}&\dots&0&0&0\\ c_{n}&c_{n-1}&0&\dots&0&0&0\\ \end{pmatrix} (4.53)

has a self-interlacing spectrum of the form (4.49). Here the entries bj>0b_{j}>0, j=2,3,,nj=2,3,\ldots,n, lie above the main diagonal, the entries cj>0c_{j}>0, j=2,3,,nj=2,3,\ldots,n, lie under the main diagonal and the only entry on the main diagonal is a1>0a_{1}>0.

In [15] it was also proved that the spectrum of the matrix (4.53) coincides with the spectrum of the following tridiagonal matrix

K=(a1b2000c20b3000c30000000bn000cn0).K=\begin{pmatrix}a_{1}&b_{2}&0&\dots&0&0\\ c_{2}&0&b_{3}&\dots&0&0\\ 0&c_{3}&0&\dots&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&\dots&0&b_{n}\\ 0&0&0&\dots&c_{n}&0\\ \end{pmatrix}.

It is well-known [10] that the spectrum of this matrix does not depend on the entries bjb_{j} and cjc_{j} separately. It depends on products bjcjb_{j}c_{j}, j=2,3,,nj=2,3,\ldots,n. So in order to the matrices (4.53) and KK to have self-interlacing spectra, it is sufficient that the inequalities a1>0a_{1}>0 and bjcj>0b_{j}c_{j}>0, j=2,3,,nj=2,3,\ldots,n, hold.

Finally, consider a tridiagonal matrix

MJ=(a1b1000c1a2b2000c2a300000an1bn1000cn1an),M_{J}=\begin{pmatrix}a_{1}&b_{1}&0&\dots&0&0\\ c_{1}&a_{2}&b_{2}&\dots&0&0\\ 0&c_{2}&a_{3}&\dots&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&\dots&a_{n-1}&b_{n-1}\\ 0&0&0&\dots&c_{n-1}&a_{n}\\ \end{pmatrix},

where ak,bk,cka_{k},b_{k},c_{k}\in\mathbb{R} and ckbk0c_{k}b_{k}\neq 0. In [10], there was proved the following fact.

Theorem 4.38.

The matrix MJM_{J} is oscillating if and only if all the entries bkb_{k} and ckc_{k} are positive and all the leading principal minors of MJM_{J} are also positive:

a1>0,|a1b1c1a2|>0,|a1b10c1a2b20c2a3|>0,|a1b1000c1a2b2000c2a300000an1bn1000cn1an|>0.a_{1}>0,\;\begin{vmatrix}a_{1}&b_{1}\\ c_{1}&a_{2}\end{vmatrix}>0,\;\begin{vmatrix}a_{1}&b_{1}&0\\ c_{1}&a_{2}&b_{2}\\ 0&c_{2}&a_{3}\end{vmatrix}>0,\;\ldots\;\begin{vmatrix}a_{1}&b_{1}&0&\dots&0&0\\ c_{1}&a_{2}&b_{2}&\dots&0&0\\ 0&c_{2}&a_{3}&\dots&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&\dots&a_{n-1}&b_{n-1}\\ 0&0&0&\dots&c_{n-1}&a_{n}\\ \end{vmatrix}>0. (4.54)

This theorem together with Theorem 4.32 implies the following statement.

Theorem 4.39.

The anti-tridiagonal matrix

AJ=(0000b1a1000b2a2c1000a3c200bn2an2000bn1an1cn2000ancn10000),A_{J}=\begin{pmatrix}0&0&0&\dots&0&b_{1}&a_{1}\\ 0&0&0&\dots&b_{2}&a_{2}&c_{1}\\ 0&0&0&\dots&a_{3}&c_{2}&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&b_{n-2}&a_{n-2}&\dots&0&0&0\\ b_{n-1}&a_{n-1}&c_{n-2}&\dots&0&0&0\\ a_{n}&c_{n-1}&0&\dots&0&0&0\\ \end{pmatrix},

where aj,bj,cj>0a_{j},b_{j},c_{j}>0 for j=1,2,,n1j=1,2,\ldots,n-1, is sign definite of class n+n^{+} and has a self-interlacing spectrum (4.49) if and only if the following inequalities hold:

(1)k(k1)2AJ(12kn+1kn+2kn)>0,k=1,2,,n.(-1)^{\frac{k(k-1)}{2}}A_{J}\begin{pmatrix}1&2&\dots&k\\ n+1-k&n+2-k&\dots&n\\ \end{pmatrix}>0,\quad k=1,2,\ldots,n. (4.55)
Proof.

If the matrix AJA_{J} is sign definite of class n+n^{+} and has a self-interlacing spectrum of the form (4.49), then according to Theorem 4.29, the signs of its nonzero minors can be calculated by the formula (4.45). This implies the inequalities (4.55).

Conversely, let the inequalities (4.55) hold. Then we have that the inequalities (4.54) hold for the matrix MJ=JAJM_{J}=JA_{J}. By Theorem 4.38, the matrix MJM_{J} is oscillating and, in particular, totally nonnegative. Now notice that (MJ)ii>0(M_{J})_{ii}>0, i=1,,ni=1,\ldots,n, and (MJ)k,k+1>0(M_{J})_{k,k+1}>0, k=1,,k1k=1,\ldots,k-1, so MJM_{J} satisfies the condition (4.51) of Theorem 4.32. Therefore, the matrix AJA_{J} is sign definite of class n+n^{+} and has a self-interlacing spectrum of the form (4.49). ∎

5 Generalized Hurwitz polynomials

In this section, we describe the class of real polynomials whose associated function Φ\Phi is an R-function. In particular, this class includes all Hurwitz stable and self-interlacing polynomials as extremal cases.

5.1 General theory

Let us consider a polynomial

p(z)=a0zn+a1zn1++an,a1,,an,a0>0.p(z)=a_{0}z^{n}+a_{1}z^{n-1}+\dots+a_{n},\qquad a_{1},\dots,a_{n}\in\mathbb{R},\ a_{0}>0. (5.1)
Definition 5.1.

The polynomial pp is called generalized Hurwitz polynomial of type I of order kk, where 1k[n+12]1\leqslant k\leqslant\left[\dfrac{n+1}{2}\right], if it has exactly kk zeroes in the closed right half-plane, all of which are nonnegative and simple:

0μ1<μ2<<μk,0\leqslant\mu_{1}<\mu_{2}<\cdots<\mu_{k}, (5.2)

such that p(μi)0p(-\mu_{i})\neq 0, i=1,,ki=1,\ldots,k, and pp has an odd number of zeroes, counting multiplicities, on each interval (μk,μk1),,(μ3,μ2),(μ2,μ1)(-\mu_{k},-\mu_{k-1}),\ldots,(-\mu_{3},-\mu_{2}),(-\mu_{2},-\mu_{1}). Moreover, the number of zeroes of pp on the interval (μ1,0)(-\mu_{1},0) (if any) is even, counting multiplicities. The other real zeroes lie on the interval (,μk)(-\infty,-\mu_{k}): an odd number of zeroes, counting multiplicities, when n=2ln=2l, and an even number of zeroes, counting multiplicities, when n=2l+1n=2l+1. All nonreal zeroes of pp (if any) are located in the open left half-plane of the complex plane.

Definition 5.2.

If p(z)p(z) is a generalized Hurwitz polynomials of type I, then the polynomial p(z)p(-z) is called generalized Hurwitz polynomial of type II.

It clear that all results obtained for the generalized Hurwitz polynomials of type I can be easily reformulated for the generalized Hurwitz polynomials of type I. Thus, in the rest of the section we consider only generalized Hurwitz polynomials of type I unless explicitly stipulated otherwise.

Since the generalized Hurwitz polynomials of order kk have exactly kk zeroes in the closed right half-plane, the generalized Hurwitz polynomials of order 0 have no zeroes in the closed right half-plane, so they are Hurwitz stable.

Analogously, if pp is a generalized Hurwitz polynomial of order k=[n+12]k=\left[\dfrac{n+1}{2}\right] without a root at zero, then pp is self-interlacing of type I. In fact, let 0<μ1<μ2<<μk0<\mu_{1}<\mu_{2}<\ldots<\mu_{k} be its positive zeroes. Then by definition, pp has an odd number of zeroes, counting multiplicities, (at least one) on each interval (μj+1,μj)(-\mu_{j+1},-\mu_{j}), j=1,,k1j=1,\ldots,k-1. Moreover, if n=2ln=2l, then pp has at least one zero on the interval (,μk)(-\infty,-\mu_{k}). It is easy to see that pp can not have nonreal zeroes and has exactly one simple zero on each interval (μj+1,μj)(-\mu_{j+1},-\mu_{j}), j=1,,k1j=1,\ldots,k-1, and on the interval (,μk)(-\infty,-\mu_{k}) for n=2ln=2l. Besides, the zeroes of pp are distributed as in (4.1).

Thus, Hurwitz stable and self-interlacing polynomials are generalized Hurwitz polynomials of minimal and maximal orders, respectively. We note that generalized Hurwitz polynomials of maximal order with a root at zero are almost self-interlacing.

Now we are in a position to establish the main theorem of the theory of generalized Hurwitz polynomials. This theorem is a direct generalization of Theorems 3.3 and 4.3.

Theorem 5.3.

Let pp be a given real polynomial of degree n1n\geqslant 1 as in (5.1). The polynomial pp is generalized Hurwitz if and only if its associated function Φ\Phi defined in (2.7) is an R-function with exactly l=[n2]l=\left[\dfrac{n}{2}\right] (for a10a_{1}\neq 0) or l1l-1 (for a1=0a_{1}=0) poles141414Let us note that if n=2ln=2l and Φ\Phi is an R-function, then a10a_{1}\neq 0, so Φ\Phi has exactly ll poles.. Moreover,

  • if n=2ln=2l or if n=2l+1n=2l+1 with

    a0a1>0,a_{0}a_{1}>0, (5.3)

    then the number of nonnegative poles of the function Φ\Phi equals order of pp;

  • if n=2l+1n=2l+1 with

    a0a10,a_{0}a_{1}\leqslant 0, (5.4)

    then the number of nonnegative poles of the function Φ\Phi equals order of pp minus one.

Proof.

As we said above, for k=0k=0 and k=[n+12]k=\left[\dfrac{n+1}{2}\right] this theorem is true.

Sufficiency. Let the function Φ\Phi be an R-function with exactly rr nonpositive poles, 0r[n2]0\leqslant r\leqslant\left[\dfrac{n}{2}\right].

At first, assume that a10a_{1}\neq 0 and the function Φ\Phi has no pole at zero. Then Φ\Phi has exactly l=[n2]l=\left[\dfrac{n}{2}\right] poles and can be represented in the following form by Theorem 1.18:

Φ(u)=α+j=1rαjuωj+j=r+1lαju+ωj,\Phi(u)=\alpha+\sum_{j=1}^{r}\frac{\alpha_{j}}{u-\omega_{j}}+\sum_{j=r+1}^{l}\frac{\alpha_{j}}{u+\omega_{j}}, (5.5)

where α\alpha\in\mathbb{R} (α=0\alpha=0 if and only if n=2ln=2l), αj>0,ωj>0\alpha_{j}>0,\omega_{j}>0 for j=1,,lj=1,\ldots,l, all poles are distinct. Let the positive poles are enumerated as follows 0<ω1<ω2<<ωr0<\omega_{1}<\omega_{2}<\ldots<\omega_{r}.

As in the proofs of Theorems 3.3 and 4.3, we note that the set of zeroes of the polynomial pp coincides with the set of roots of the following equation

Φ(z2)=1z.\Phi(z^{2})=-\dfrac{1}{z}. (5.6)

So if we suppose that p(λ)=0p(\lambda)=0 and Imλ0\operatorname{Im}\lambda\neq 0 (that is, λ\lambda is a nonreal zero of pp), then from (5.6) we obtain

ImΦ(λ2)Im(λ2)=1|λ|2ImλImλ2=12|λ|21Reλ<0,\displaystyle\frac{\operatorname{Im}\Phi(\lambda^{2})}{\operatorname{Im}(\lambda^{2})}=\frac{1}{|\lambda|^{2}}\frac{\operatorname{Im}\lambda}{\operatorname{Im}\lambda^{2}}=\frac{1}{2|\lambda|^{2}}\frac{1}{\operatorname{Re}\lambda}<0, (5.7)

since Φ\Phi is an R-function by assumption. From (5.7) it follows that all nonreal zeroes of pp lie in the open left half-plane. Thus, all zeroes of pp are located in the open left half-plane or on the nonnegative real half-axis.

As in the proof of Theorem 4.3, we investigate the distribution of the real zeroes of the polynomial pp. By assumption, |Φ(0)|<|\Phi(0)|<\infty, so the polynomial pp does not vanish at zero. In fact, if p(0)=0p(0)=0, then p0(0)=0p_{0}(0)=0, but it contradicts with |Φ(0)|<|\Phi(0)|<\infty. Thus, we suppose that z{0}z\in\mathbb{R}\setminus\{0\} and change the variables as follows z2=u>0z^{2}=u>0. Then for real nonzero zz, the equation (5.6) is equivalent to the following system

{Φ(u)=1u,Φ(u)=1u,u>0.\begin{cases}\ \Phi(u)=-\dfrac{1}{\sqrt{u}},\\ \ \Phi(u)=\dfrac{1}{\sqrt{u}},\end{cases}\quad u>0. (5.8)

We also note (as in the proof of Theorem 4.3) that all positive roots (if any) of the first equation (5.8) are squares of positive roots of the equation (5.6), and all positive roots (if any) of the second equation (5.8) are squares of negative roots of the equation (5.6). The equation (5.6) may have nonreal roots, but as we showed above all these roots are located in the open left half-plane.

Let n=2ln=2l. Then limu±Φ(u)=α=0\lim\limits_{u\to\pm\infty}\Phi(u)=\alpha=0. The set of zeroes the function F1(u)=Φ(u)+1uF_{1}(u)=\Phi(u)+\dfrac{1}{\sqrt{u}} coincides with the set of roots of the first equation (5.8). In the same way as in the proof of Theorem 4.3, one can show that the function F1F_{1} has exactly one simple zero, say μi2\mu_{i}^{2} (μi>0\mu_{i}>0), on each interval (ωi1,ωi)(\omega_{i-1},\omega_{i}), i=2,,ri=2,\ldots,r and exactly one simple zero μ12\mu_{1}^{2} (μ1>0\mu_{1}>0) on the interval (0,ω1)(0,\omega_{1}). Moreover, denoting by ξi\xi_{i} a unique zero151515Recall that Φ\Phi is an R-function by assumption. Consequently, it has exactly one simple zero on each interval (ωi,ωi+1)(\omega_{i},\omega_{i+1}). of the function Φ\Phi on the interval (ωi,ωi+1)(\omega_{i},\omega_{i+1}), i=1,,r1i=1,\ldots,r-1, we have

μi2<ωi<ξi<μi+12<ωi+1,i=1,,r1,\mu_{i}^{2}<\omega_{i}<\xi_{i}<\mu_{i+1}^{2}<\omega_{i+1},\qquad i=1,\ldots,r-1, (5.9)

since F1(ξi)=1ξi>0F_{1}(\xi_{i})=\dfrac{1}{\sqrt{\xi_{i}}}>0, and F1F_{1} is decreasing between its poles on the positive half-axis as a sum of decreasing functions.

Thus, the first equation (5.8) has exactly rr positive roots μi2\mu_{i}^{2} all of which are simple and satisfy the inequalities (5.9).

Now consider the function F2(u)=Φ(u)1uF_{2}(u)=\Phi(u)-\dfrac{1}{\sqrt{u}} whose set of zeroes coincides with the set of roots of the second equation (5.8). In the same way as in the proof of Theorem 4.3, one can show that the function F2F_{2} has an odd number of zeroes, counting multiplicities, on each interval (ωi,μi+12)(\omega_{i},\mu^{2}_{i+1}), i=1,,r1i=1,\ldots,r-1, and on the interval (ωr,+)(\omega_{r},+\infty). Moreover, F2(u)<0F_{2}(u)<0 on the intervals [μi+12,ωi+1)[\mu^{2}_{i+1},\omega_{i+1}), i=1,,r1i=1,\ldots,r-1. Let us now turn our attention to the interval (0,ω1)(0,\omega_{1}). Since |Φ(0)|<0|\Phi(0)|<0 by assumption and 1u-\dfrac{1}{\sqrt{u}}\to-\infty as u+0u\to+0, we have F2(u)F_{2}(u)\to-\infty as u+0u\to+0. Besides, F2(u)=F1(u)2uF_{2}(u)=F_{1}(u)-\dfrac{2}{\sqrt{u}}, so F2(μ12)=2μ1<0F_{2}(\mu_{1}^{2})=-\dfrac{2}{\mu_{1}}<0. Therefore, F2F_{2} has an even number of zeroes, counting multiplicities, on the interval (0,μ12)(0,\mu_{1}^{2}). However, since F1(u)F_{1}(u) is monotone decreasing on (0,ω1)(0,\omega_{1}) as a sum of two monotone decreasing functions, we have F1(u)<0F_{1}(u)<0 on the interval (μ12,ω1)(\mu_{1}^{2},\omega_{1}). Consequently, F2(u)F_{2}(u) is also negative for u(μ12,ω1)u\in(\mu_{1}^{2},\omega_{1}), so its zeroes on the interval (0,ω1)(0,\omega_{1}) (if any) lie, indeed, on the interval (0,μ12)(0,\mu_{1}^{2}).

Thus, we obtain that in the closed right half-plane, the polynomial pp has only rr zeroes, all of which are positive and simple:

0<μ1<μ2<<μr.0<\mu_{1}<\mu_{2}<\ldots<\mu_{r}.

Moreover, pp has an even number of zeroes, counting multiplicities, on the interval (μ1,0)(-\mu_{1},0) and an odd number of zeroes, counting multiplicities, on each interval (μi+1,μi)(-\mu_{i+1},-\mu_{i}), i=1,,r1i=1,\ldots,r-1. Also p(μi)0p(-\mu_{i})\neq 0 for i=1,,ri=1,\ldots,r. As we showed above, all nonreal zeroes of pp are in the open left-half-plane, so pp is a generalized Hurwitz polynomial of order rr.

Let n=2l+1n=2l+1, and the condition (5.3) holds. The only difference between this case and the case n=2ln=2l is behaviour of the function Φ\Phi at infinity. Thus, as in the case n=2ln=2l, F1F_{1} has a unique simple zero, say μi2\mu_{i}^{2} (μi>0\mu_{i}>0), i=1,,ri=1,\ldots,r, on each interval (0,ω1)(0,\omega_{1}), (ω1,ω2)(\omega_{1},\omega_{2}), …, (ωr1,ωr)(\omega_{r-1},\omega_{r}). These zeroes are distributed as in (5.9). Furthermore, we have F1(u)+F_{1}(u)\to+\infty as uωru\searrow\omega_{r} and F1(u)α=a0a1>0F_{1}(u)\to\alpha=\dfrac{a_{0}}{a_{1}}>0 as u+u\to+\infty, so F1(u)>0F_{1}(u)>0 for u(ωr,+)u\in(\omega_{r},+\infty) (recall that Φ\Phi is a monotone decreasing function on the interval (ωr,+)(\omega_{r},+\infty) and Φ(u)+\Phi(u)\to+\infty as uωru\searrow\omega_{r}). As in the case n=2ln=2l, the function F2(u)F_{2}(u) has an even number of zeroes, counting multiplicities, on the interval (0,μ12)(0,\mu_{1}^{2}) and an odd number of zeroes, counting multiplicities, on each interval (ωi,μi+12)(\omega_{i},\mu^{2}_{i+1}) for i=1,,r1i=1,\ldots,r-1. Moreover, F2F_{2} has no zeroes on the intervals (μ12,ω1)(\mu_{1}^{2},\omega_{1}) and (μi+12,ωi+1)(\mu^{2}_{i+1},\omega_{i+1}), i=1,,r1i=1,\ldots,r-1. At the same time, we have F2(u)+F_{2}(u)\to+\infty as uωru\searrow\omega_{r} and F2(u)α>0F_{2}(u)\to\alpha>0 as u+u\to+\infty. Since the function F2(u)F_{2}(u) is not monotone, it has an even number of zeroes, counting multiplicities, on the interval (ωr,+)(\omega_{r},+\infty). Recall that these zeroes are squares of positive zeroes of the polynomial pp on the interval (ωr,+)(\omega_{r},+\infty). Since the other zeroes of the functions F1F_{1} and F2F_{2} are distributed as in the case n=2ln=2l, we obtain that pp is a generalized Hurwitz polynomial of order rr.

Let now n=2l+1n=2l+1, and let a0a1<0a_{0}a_{1}<0. This case also differs from the case n=2ln=2l by behaviour of the function Φ\Phi at infinity: limu+Φ(u)=α=a0a1<0\lim\limits_{u\to+\infty}\Phi(u)=\alpha=\dfrac{a_{0}}{a_{1}}<0. As in the case n=2ln=2l, the function F1(u)=Φ(u)+1uF_{1}(u)=\Phi(u)+\dfrac{1}{\sqrt{u}} has a unique simple zero, say μi2\mu_{i}^{2} (μi>0\mu_{i}>0), i=1,,ri=1,\ldots,r, on each interval (0,ω1)(0,\omega_{1}), (ω1,ω2)(\omega_{1},\omega_{2}), …, (ωr1,ωr)(\omega_{r-1},\omega_{r}). These zeroes are distributed as in (5.9). Furthermore, we have F1(u)+F_{1}(u)\to+\infty as uωru\searrow\omega_{r} and F1(u)α<0F_{1}(u)\to\alpha<0 as u+u\to+\infty. Since F1(u)F_{1}(u) is decreasing on the interval (ωr,+)(\omega_{r},+\infty) as a sum of two decreasing functions on this interval, it has a unique simple zero, say μr+12\mu_{r+1}^{2} (μr+1>0\mu_{r+1}>0) on (ωr,+)(\omega_{r},+\infty) such that

ωr<ξr<μr+12,\omega_{r}<\xi_{r}<\mu^{2}_{r+1},

where ξr\xi_{r} is a zero161616Recall that this zero is unique, since Φ\Phi is a monotone function on (ωr,+)(\omega_{r},+\infty). of Φ\Phi on the interval (ωr,+)(\omega_{r},+\infty). As in the case n=2ln=2l, the function F2(u)F_{2}(u) has an even number of zeroes, counting multiplicities, on the interval (0,μ12)(0,\mu_{1}^{2}) and an odd number of zeroes, counting multiplicities, on each interval (ωi,μi+12)(\omega_{i},\mu^{2}_{i+1}) for i=1,,r1i=1,\ldots,r-1. Moreover, F2F_{2} has no zeroes on the intervals (μi2,ωi)(\mu^{2}_{i},\omega_{i}), i=1,,ri=1,\ldots,r. Now consider the interval (ωr,+)(\omega_{r},+\infty). Since F2(u)+F_{2}(u)\to+\infty as uωru\searrow\omega_{r} and F2(μr+12)=1μr+1<0F_{2}(\mu^{2}_{r+1})=-\dfrac{1}{\mu_{r+1}}<0, we obtain that F2F_{2} has an odd number of zeroes, counting multiplicities, on the interval (ωr,μr+12)(\omega_{r},\mu^{2}_{r+1}). Moreover, F2(u)=F1(u)2u<0F_{2}(u)=F_{1}(u)-\dfrac{2}{\sqrt{u}}<0 for (μr+12,+)(\mu^{2}_{r+1},+\infty), because F1F_{1} is negative on that interval. Since the positive roots of the first equation (5.8) are squares of the positive zeroes of the polynomial pp and the roots of the second equation (5.8) are squares of the negative zeroes of the polynomial pp, we obtain that that pp has only r+1r+1 zeroes in the closed right half-plane, all of which are positive and simple:

0<μ1<μ2<<μr+1.0<\mu_{1}<\mu_{2}<\ldots<\mu_{r+1}.

Moreover, pp has an even number of zeroes, counting multiplicities, on the interval (μ1,0)(-\mu_{1},0) and an odd number of zeroes, counting multiplicities, on each interval (μi+1,μi)(-\mu_{i+1},-\mu_{i}), i=1,,ri=1,\ldots,r. All nonreal zeroes of pp are in the open left half-plane, and p(μi)0p(-\mu_{i})\neq 0 for i=1,,r+1i=1,\ldots,r+1, so pp is a generalized Hurwitz polynomial of order r+1r+1.

If n=2l+1n=2l+1, and a1=0a_{1}=0, then by Theorem 1.18, the function Φ\Phi has exactly l1l-1 poles and can be represented in the form

Φ(u)=βu+γ+j=1rαjuωj+j=r+1l1αju+ωj,\Phi(u)=\beta u+\gamma+\sum_{j=1}^{r}\frac{\alpha_{j}}{u-\omega_{j}}+\sum_{j=r+1}^{l-1}\frac{\alpha_{j}}{u+\omega_{j}}, (5.10)

where β<0\beta<0, γ\gamma\in\mathbb{R}, αj>0,ωj>0\alpha_{j}>0,\omega_{j}>0 for j=1,,lj=1,\ldots,l, all poles are distinct. In the same way as in the case n=2l+1n=2l+1 with a0a1<0a_{0}a_{1}<0, one can show that pp is a generalized Hurwitz polynomial of order r+1r+1, since Φ(u)\Phi(u) is also negative for sufficiently large uu.

Suppose now that the function Φ\Phi has a pole at zero. Then p0(0)=0p_{0}(0)=0, so p(0)=an=0p(0)=a_{n}=0. At the same time, p1(0)=an1=p(0)0p_{1}(0)=a_{n-1}=p^{\prime}(0)\neq 0, since Φ\Phi is an R-function by assumption, so it has interlacing poles and zeroes (see Theorem 1.18). Thus, pp has a simple root at zero. Note that the set of zeroes of the polynomial pp, but the root at zero, coincides with the set of roots of the equation (5.6).

If n=2ln=2l or n=2l+1n=2l+1 with (5.3), then as above, one can show that in the closed right half-plane pp has exactly rr zeroes, say μi\mu_{i}, i=1,ri=1\ldots,r, all of which are nonnegative and simple:

0=μ1<μ2<μ2<<μr,0=\mu_{1}<\mu_{2}<\mu_{2}<\ldots<\mu_{r},

and p(μi)0p(-\mu_{i})\neq 0 for i=2,,ri=2,\ldots,r. Also pp has an odd number of zeroes, counting multiplicities, on each interval (μi,μi1)(-\mu_{i},-\mu_{i-1}) and an odd (even) number of zeroes on the interval (,μr)(-\infty,-\mu_{r}) if n=2ln=2l (n=2l+1n=2l+1). All nonreal zeroes of pp lie in the open left half-plane. Thus, pp is generalized Hurwitz of order rr.

If n=2l+1n=2l+1 and the condition (5.4) holds, then in the same way as above, one can show that pp is generalized Hurwitz of order r+1r+1.


Necessity. Let pp be a generalized Hurwitz polynomial of order kk. At first, suppose that p(0)0p(0)\neq 0.

Let n=2ln=2l. Then by definition, p(z)p(z) has only kk zeroes in the closed right half-plane, all of which are positive and simple: 0<μ1<μ2<<μk0<\mu_{1}<\mu_{2}<\ldots<\mu_{k}, and p(μi)0p(-\mu_{i})\neq 0. Moreover, pp has an odd number of zeroes, counting multiplicities, on each interval (,μk)(-\infty,-\mu_{k}), (μk,μk1)(-\mu_{k},-\mu_{k-1}),…, (μ2,μ1)(-\mu_{2},-\mu_{1}), and it has an even number of zeroes, counting multiplicities, on the interval (μ1,0)(-\mu_{1},0). All nonreal zeroes of pp lie in the open left half-plane. Thus, we can represent the polynomial pp as a product p(z)=h(z)g(z)p(z)=h(z)g(z), where hh is a self-interlacing polynomial171717More exactly, hh is self-interlacing of type I., of degree 2k2k whose positive zeroes are μi\mu_{i}, i=1,,ki=1,\ldots,k, and gg is a Hurwitz stable polynomial of degree 2(lk)2(l-k), which has an even number of zeroes, counting multiplicities, on each interval (,μk)(-\infty,-\mu_{k}), (μk,μk1)(-\mu_{k},-\mu_{k-1}),…, (μ2,μ1)(-\mu_{2},-\mu_{1}), (μ1,0)(-\mu_{1},0). Let

h(z)=h0(z2)+zh1(z2),h(z)=h_{0}(z^{2})+zh_{1}(z^{2}),
g(z)=g0(z2)+zg1(z2).g(z)=g_{0}(z^{2})+zg_{1}(z^{2}).

Then by Theorem 4.3, the function H:=h1h0H:=\dfrac{h_{1}}{h_{0}} is an R-function with exactly kk poles, all of which are positive. Analogously, by Theorem 3.3, the function G:=g1g0G:=\dfrac{g_{1}}{g_{0}} is an RR function with exactly lkl-k poles, all of which are negative. If p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}), then

p1(u)=h0(u)g1(u)+h1(u)g0(u),p_{1}(u)=h_{0}(u)g_{1}(u)+h_{1}(u)g_{0}(u), (5.11)
p0(w)=h0(u)g0(u)+uh1(u)g1(u).p_{0}(w)=h_{0}(u)g_{0}(u)+uh_{1}(u)g_{1}(u). (5.12)

Thus,

Φ(u)=p1(u)p0(u)=h0(u)g1(u)+h1(u)g0(u)h0(u)g0(u)+uh1(u)g1(u)=H(u)+G(u)1+uH(u)G(u).\displaystyle\Phi(u)=\frac{p_{1}(u)}{p_{0}(u)}=\frac{h_{0}(u)g_{1}(u)+h_{1}(u)g_{0}(u)}{h_{0}(u)g_{0}(u)+uh_{1}(u)g_{1}(u)}=\frac{H(u)+G(u)}{1+uH(u)G(u)}.

Note that the function

h1(u)h0(u)+g1(u)g0(u)=p1(u)h0(u)g0(u)\displaystyle\frac{h_{1}(u)}{h_{0}(u)}+\frac{g_{1}(u)}{g_{0}(u)}=\frac{p_{1}(u)}{h_{0}(u)g_{0}(u)} (5.13)

is an R-function as a sum of R-functions. Consequently, all zeroes of p1p_{1} are real and simple and interlace zeroes of the polynomial h0g0h_{0}g_{0}, which has kk positive simple zeroes and lkl-k negative simple zeroes.

Furthermore, the function

F(u)=h0(u)h1(u)ug1(u)g0(u)=1H(u)uG(u)=p0(u)h1(u)g0(u)\displaystyle F(u)=-\frac{h_{0}(u)}{h_{1}(u)}-\frac{ug_{1}(u)}{g_{0}(u)}=-\dfrac{1}{H(u)}-uG(u)=-\frac{p_{0}(u)}{h_{1}(u)g_{0}(u)} (5.14)

is an R-function by Theorem 1.23, since the functions HH and GG are R-functions. Consequently, all zeroes of the polynomial p0p_{0} are real and simple and interlace zeroes of the polynomial h1g0h_{1}g_{0} (see Theorem 1.18). Thus, from (5.13)–(5.14) it follows that all poles and zeroes of the function Φ=p1p0\Phi=\dfrac{p_{1}}{p_{0}} are real and simple. Therefore, Φ\Phi can be represented as follows

Φ(u)=p1(u)p0(u)=j=1lαjuωj,\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}=\sum\limits_{j=1}^{l}\dfrac{\alpha_{j}}{u-\omega_{j}}, (5.15)

where ωi\omega_{i}\in\mathbb{R}, ωiωj\omega_{i}\neq\omega_{j} for iji\neq j, and

αj=p1(ωj)p0(ωj)0,j=1,2,,l.\displaystyle\alpha_{j}=\frac{p_{1}(\omega_{j})}{p^{\prime}_{0}(\omega_{j})}\neq 0,\qquad j=1,2,\ldots,l.

According to Theorem 1.18, now it is sufficient to establish positivity of the numbers αj\alpha_{j} to prove that Φ\Phi is an R-function.

Let ωj\omega_{j} be a zero of the polynomial p0p_{0}. Since p(0)0p(0)\neq 0 by assumption, we have p0(0)0p_{0}(0)\neq 0, so ωj0\omega_{j}\neq 0, j=1,2,,lj=1,2,\ldots,l. It is clear that h1(ωj)g0(ωj)0h_{1}(\omega_{j})g_{0}(\omega_{j})\neq 0, because the zeroes of the polynomial p0p_{0} interlace the zeroes of h1g0h_{1}g_{0}. From (5.12) it follows that

h0(ωj)=ωjh1(ωj)g1(ωj)g0(ωj),\displaystyle h_{0}(\omega_{j})=-\frac{\omega_{j}h_{1}(\omega_{j})g_{1}(\omega_{j})}{g_{0}(\omega_{j})},

so

p1(ωj)=h1(ωj)g0(ωj)ωjh1(ωj)g12(ωj)g0(ωj)==h1(ωj)g0(ωj)[g02(ωj)ωjg12(ωj)]=h1(ωj)g0(ωj)g(ωj)g(ωj).\begin{array}[]{l}\displaystyle p_{1}(\omega_{j})=h_{1}(\omega_{j})g_{0}(\omega_{j})-\frac{\omega_{j}h_{1}(\omega_{j})g_{1}^{2}(\omega_{j})}{g_{0}(\omega_{j})}=\\ =\dfrac{h_{1}(\omega_{j})}{g_{0}(\omega_{j})}\left[g_{0}^{2}(\omega_{j})-\omega_{j}g_{1}^{2}(\omega_{j})\right]=\dfrac{h_{1}(\omega_{j})}{g_{0}(\omega_{j})}g(\sqrt{\omega_{j}})g(-\sqrt{\omega_{j}}).\end{array}

According to (5.14), we have p0(u)=h1(u)g0(u)F(u)p_{0}(u)=-h_{1}(u)g_{0}(u)F(u), and F(ωj)=0F(\omega_{j})=0, therefore

p0(ωj)=h1(ωj)g0(ωj)F(ωj).p_{0}^{\prime}(\omega_{j})=-h_{1}(\omega_{j})g_{0}(\omega_{j})F^{\prime}(\omega_{j}).

So we obtain

αj=p1(ωj)p0(ωj)=h1(ωj)g0(ωj)g(ωj)g(ωj)h1(ωj)g0(ωj)F(ωj)=g(ωj)g(ωj)g02(ωj)F(ωj).\alpha_{j}=\dfrac{p_{1}(\omega_{j})}{p^{\prime}_{0}(\omega_{j})}=\dfrac{h_{1}(\omega_{j})}{g_{0}(\omega_{j})}\cdot\dfrac{g(\sqrt{\omega_{j}})g(-\sqrt{\omega_{j}})}{-h_{1}(\omega_{j})g_{0}(\omega_{j})F^{\prime}(\omega_{j})}=-\dfrac{g(\sqrt{\omega_{j}})g(-\sqrt{\omega_{j}})}{g_{0}^{2}(\omega_{j})F^{\prime}(\omega_{j})}.

Since FF is an R-function, we have F(u)<0F^{\prime}(u)<0, so g02(ωj)F(ωj)<0g_{0}^{2}(\omega_{j})F^{\prime}(\omega_{j})<0. Thus, we obtain that

sign(αj)=sign(g(ωj)g(ωj)).\textrm{sign}(\alpha_{j})=\textrm{sign}\left(g(\sqrt{\omega_{j}})g(-\sqrt{\omega_{j}})\right).

If ωj<0\omega_{j}<0, then g(ωj)g(ωj)=g02(ωj)ωjg12(ωj)>0g(\sqrt{\omega_{j}})g(-\sqrt{\omega_{j}})=g_{0}^{2}(\omega_{j})-\omega_{j}g_{1}^{2}(\omega_{j})>0, since g0g_{0} and g1g_{1} are real polynomials. So αj>0\alpha_{j}>0 in this case.

Let now ωj>0\omega_{j}>0. The polynomial gg is Hurwitz stable with positive leading coefficient by construction, so g(z)>0g(z)>0 for all z0z\geqslant 0. Therefore, sign(αj)=sign(g(ωj))\textrm{sign}(\alpha_{j})=\textrm{sign}\left(g(-\sqrt{\omega_{j}})\right). Since degp0=l\deg p_{0}=l, degg0=lk\deg g_{0}=l-k, degh1=k1\deg h_{1}=k-1, the function FF defined in (5.14) has exactly ll zeroes and l1l-1 poles. Consequently, the maximal positive zero of the polynomial p0p_{0} is greater than the maximal positive zero of the polynomial h1h_{1} (recall that all zeroes of h1h_{1} are positive). Besides, FF is decreasing between its poles, so by (5.12), we obtain F(0)=p0(0)h1(0)g0(0)=h0(0)h1(0)>0F(0)=-\dfrac{p_{0}(0)}{h_{1}(0)g_{0}(0)}=-\dfrac{h_{0}(0)}{h_{1}(0)}>0, since hh is self-interlacing by construction, and (4.20) holds for it. This means that the function FF (and the polynomial p0p_{0}) has a positive zero, which is less than the minimal positive zero of h1h_{1}. Thus, the polynomial p0p_{0} has exactly kk positive zeroes and, respectively, exactly lkl-k negative zeroes.

Let us enumerate the positive zeroes of p0p_{0} as follows

0<ω1<ω2<<ωk.0<\omega_{1}<\omega_{2}<\ldots<\omega_{k}.

Then denoting the zeroes of the polynomial h1h_{1} by γj\gamma_{j} (j=1,2,,k1)(j=1,2,\ldots,k-1), we obtain

0<ω1<γ1<ω2<γ2<<ωk<γk1<ωk.0<\omega_{1}<\gamma_{1}<\omega_{2}<\gamma_{2}<\ldots<\omega_{k}<\gamma_{k-1}<\omega_{k}.

Let us now show that ω1>μ12\omega_{1}>\mu_{1}^{2}. Indeed, since μ1\mu_{1} is the minimal positive zero of hh and h(0)0h(0)\neq 0, the polynomial hh does not change its sign on the interval [0,μ1)[0,\mu_{1}). Moreover, both polynomials h0h_{0} and h1h_{1} have no zeroes on the interval [0,μ12][0,\mu_{1}^{2}] (see the proof of Theorem 4.3). Consequently, we get sign(h(0))=sign(h0(0))=sign(h1(0))=sign(h0(μ12))=sign(h1(μ12))\textrm{sign}(h(0))=\textrm{sign}\left(h_{0}(0)\right)=-\textrm{sign}\left(h_{1}(0)\right)=\textrm{sign}\left(h_{0}(\mu_{1}^{2})\right)=-\textrm{sign}\left(h_{1}(\mu_{1}^{2})\right) that follows from the self-interlacing of hh. Note that sign(p0(0))=sign(h0(0)g0(0))=sign(h0(0))\textrm{sign}\left(p_{0}(0)\right)=\textrm{sign}\left(h_{0}(0)g_{0}(0)\right)=\textrm{sign}\left(h_{0}(0)\right), since gg is Hurwitz stable with the positive leading coefficient, so g0(0)>0g_{0}(0)>0. The number μ1\mu_{1} is a zero of the polynomial hh, therefore h0(μ12)=μ1h1(μ12)h_{0}(\mu_{1}^{2})=-\mu_{1}h_{1}(\mu_{1}^{2}), so from (5.12) we have

p0(μ12)=h0(μ12)g0(μ12)μ1h0(μ12)g1(μ12)=h0(μ12)g(μ1).p_{0}(\mu_{1}^{2})=h_{0}(\mu_{1}^{2})g_{0}(\mu_{1}^{2})-\mu_{1}h_{0}(\mu_{1}^{2})g_{1}(\mu_{1}^{2})=h_{0}(\mu_{1}^{2})g(-\mu_{1}).

By construction, the polynomial gg has an even number of zeroes, counting multiplicities, on each interval (,μk),(μk,μk1),,(μ2,μ1),(μ1,0)(-\infty,-\mu_{k}),(-\mu_{k},-\mu_{k-1}),\ldots,(-\mu_{2},-\mu_{1}),(-\mu_{1},0), therefore sign(g(0))=sign(g(μj))\textrm{sign}\left(g(0)\right)=\textrm{sign}\left(g(-\mu_{j})\right), j=1,2,,kj=1,2,\ldots,k, that is, g(μj)>0g(-\mu_{j})>0. This implies that sign(p0(μ12))=sign(h0(μ12))=sign(h0(0))=sign(p0(0))\textrm{sign}\left(p_{0}(\mu_{1}^{2})\right)=\textrm{sign}\left(h_{0}(\mu_{1}^{2})\right)=\textrm{sign}\left(h_{0}(0)\right)=\textrm{sign}\left(p_{0}(0)\right). Consequently, the polynomial p0p_{0} has an even number of zeroes, counting multiplicities, on the interval (0,μ12)(0,\mu_{1}^{2}). But the positive zeroes of the polynomials p0p_{0} and h1h_{1} interlace, and h1h_{1} has no zeroes in the interval (0,μ12)(0,\mu_{1}^{2}), therefore p0p_{0} has also no zeroes on that interval, that is, ω1>μ12\omega_{1}>\mu_{1}^{2}.

Note that (see the proof of Theorem 4.3)

0<μ12<γ1<μ22<γ2<<μk12<γk1<μk2.0<\mu_{1}^{2}<\gamma_{1}<\mu_{2}^{2}<\gamma_{2}<\ldots<\mu_{k-1}^{2}<\gamma_{k-1}<\mu_{k}^{2}.

Thus, we already established that 0<μ12<ω1<γ1<μ220<\mu_{1}^{2}<\omega_{1}<\gamma_{1}<\mu_{2}^{2}. Now we prove that the polynomial p0p_{0} has no zeroes on each interval (γj,μj+12)(\gamma_{j},\mu_{j+1}^{2}), j=1,2,,k1j=1,2,\ldots,k-1. In fact, since sign(p0(0))=sign(h0(0))\textrm{sign}\left(p_{0}(0)\right)=\textrm{sign}\left(h_{0}(0)\right) and both polynomials p0p_{0} and h0h_{0} change their signs on each interval (γj,γj+1)(\gamma_{j},\gamma_{j+1}), j=1,2,,k1j=1,2,\ldots,k-1, we obtain sign(p0(γj))=sign(h0(γj))\textrm{sign}\left(p_{0}(\gamma_{j})\right)=\textrm{sign}\left(h_{0}(\gamma_{j})\right) for all j=1,2,,k1j=1,2,\ldots,k-1. By Theorem 4.3, the following inequalities hold

0<μ12<β1<γ1<μ22<β2<γ2<<μk12<βk1<γk1<μk2<βk,0<\mu_{1}^{2}<\beta_{1}<\gamma_{1}<\mu_{2}^{2}<\beta_{2}<\gamma_{2}<\ldots<\mu_{k-1}^{2}<\beta_{k-1}<\gamma_{k-1}<\mu_{k}^{2}<\beta_{k}, (5.16)

where βj\beta_{j} are the zeroes of the polynomial h0h_{0}. From (5.16) it follows that h0h_{0} does not change its sign on each interval (γj,μj+12)(\gamma_{j},\mu_{j+1}^{2}), j=1,2,,k1j=1,2,\ldots,k-1, that is, sign(h0(γj))=sign(h0(μj+12))\textrm{sign}\left(h_{0}(\gamma_{j})\right)=\textrm{sign}\left(h_{0}(\mu_{j+1}^{2})\right) for all j=1,2,,k1j=1,2,\ldots,k-1, in particular. Since h(μj+1)=0h(\mu_{j+1})=0 and therefore h0(μj+12)=μjh1(μj+12)h_{0}(\mu_{j+1}^{2})=-\mu_{j}h_{1}(\mu_{j+1}^{2}), we have

p0(μj+12)=h0(μj+12)g0(μj+12)μj+1h0(μj+12)g1(μj+12)=h0(μj+12)g(μj+1).p_{0}(\mu_{j+1}^{2})=h_{0}(\mu_{j+1}^{2})g_{0}(\mu_{j+1}^{2})-\mu_{j+1}h_{0}(\mu_{j+1}^{2})g_{1}(\mu_{j+1}^{2})=h_{0}(\mu_{j+1}^{2})g(-\mu_{j+1}).

The inequalities g(μj+1)>0g(-\mu_{j+1})>0, j=1,,k1j=1,\ldots,k-1, proved above imply sign(p0(μj+12))=sign(h0(μj+12))=sign(h0(γj))=sign(p0(γj))\textrm{sign}\left(p_{0}(\mu_{j+1}^{2})\right)=\textrm{sign}\left(h_{0}(\mu_{j+1}^{2})\right)=\textrm{sign}\left(h_{0}(\gamma_{j})\right)=\textrm{sign}\left(p_{0}(\gamma_{j})\right). Consequently, the polynomial p0p_{0} has an even number of zeroes, counting multiplicities, on each interval (γj,μj+12)(\gamma_{j},\mu_{j+1}^{2}), j=1,,k1j=1,\ldots,k-1. But the zeroes of the polynomials p0p_{0} and h1h_{1} interlace, and h1h_{1} has no zeroes on the intervals (γj,μj+12)(\gamma_{j},\mu_{j+1}^{2}), j=1,,k1j=1,\ldots,k-1, therefore p0p_{0} also has no zeroes on those intervals. So we proved the following inequalities

μj2<ωj<γj<μj+12forj=1,,k1.\mu_{j}^{2}<\omega_{j}<\gamma_{j}<\mu_{j+1}^{2}\qquad\text{for}\qquad j=1,\ldots,k-1. (5.17)

Let us consider the intervals (μj2,ωj)(\mu_{j}^{2},\omega_{j}), j=1,,kj=1,\ldots,k and prove that the polynomial g(u)g(-\sqrt{u}) has no zeroes on those interval. Suppose that it is not true. Then there exists a number β(μj2,ωj)\beta\in(\mu_{j}^{2},\omega_{j}) such that g(β)=0g(-\sqrt{\beta})=0. Then g0(β)=βg1(β)>0g_{0}(\beta)=\sqrt{\beta}g_{1}(\beta)>0, and from the equality

p0(β)=h0(β)g0(β)+βh1(β)g1(β)=g0(β)h(β),p_{0}(\beta)=h_{0}(\beta)g_{0}(\beta)+\beta h_{1}(\beta)g_{1}(\beta)=g_{0}(\beta)h(\sqrt{\beta}),

we obtain that sign(p0(β))=sign(h(β))\textrm{sign}\left(p_{0}(\beta)\right)=\textrm{sign}\left(h(\sqrt{\beta})\right). It contradicts the equality sign(p0(β))=sign(h(β))\textrm{sign}\left(p_{0}(\beta)\right)=-\textrm{sign}\left(h(\sqrt{\beta})\right), which follows from (5.17) and from the equalities sign(p0(0))=sign(h0(0))=sign(h(0))\textrm{sign}\left(p_{0}(0)\right)=\textrm{sign}\left(h_{0}(0)\right)=\textrm{sign}\left(h(0)\right) established above.

Thus, we obtain that the polynomial gg has real zeroes (en even number of zeroes, counting multiplicities) only on the intervals (,ωk)(-\infty,\sqrt{\omega_{k}}), (μk,ωk1)(-\mu_{k},-\sqrt{\omega_{k-1}}), …, (μ2,ω1),(μ1,0)(-\mu_{2},-\sqrt{\omega_{1}}),(-\mu_{1},0). Consequently, the following equalities hold sign(g(0))=sign(g(μj))=sign(g(ωj))=sign(αj)>0\textrm{sign}\left(g(0)\right)=\textrm{sign}\left(g(-\mu_{j})\right)=\textrm{sign}\left(g(-\sqrt{\omega_{j}})\right)=\textrm{sign}(\alpha_{j})>0 for all j=1,2,,kj=1,2,\ldots,k. So we showed that if n=2ln=2l and p(0)0p(0)\neq 0, then the function Φ(u)=p1p0\Phi(u)=\dfrac{p_{1}}{p_{0}} can be represented as in (5.15) with positive αj\alpha_{j} and with exactly kk positive poles, as required.


Let now n=2l+1n=2l+1 and p(0)0p(0)\neq 0. Then by definition, the polynomial pp has only kk zeroes in the closed right half-plane, all of which are positive: 0<μ1<μ2<<μk0<\mu_{1}<\mu_{2}<\ldots<\mu_{k}. Moreover, pp has an odd number of zeroes of pp, counting multiplicities, on each interval (μk,μk1)(-\mu_{k},-\mu_{k-1}), …, (μ3,μ2)(-\mu_{3},-\mu_{2}), (μ2,μ1)(-\mu_{2},-\mu_{1}), and an even number of zeroes, counting multiplicities, on the intervals (,μk)(-\infty,-\mu_{k}) and (μ1,0)(-\mu_{1},0). All nonreal zeroes of pp are located in the open left half-plane.

As in the case n=2ln=2l, we represent the polynomial pp as a product p(z)=g(z)h(z)p(z)=g(z)h(z), where hh is self-interlacing polynomial (of type I) of degree 2k12k-1 whose positive zeroes are μ1,μ2,,μk\mu_{1},\mu_{2},\ldots,\mu_{k}, and gg is Hurwitz stable polynomial of degree 2(l+1k)2(l+1-k), which has an even number of zeroes, counting multiplicities, on each interval (,μk),(μk,μk1),(-\infty,-\mu_{k}),(-\mu_{k},-\mu_{k-1}),\ldots, (μ2,μ1),(μ1,0)(-\mu_{2},-\mu_{1}),(-\mu_{1},0). In the same way as above one can prove that the function Φ=p1p0\Phi=\dfrac{p_{1}}{p_{0}} is an R-function. The only difference between this case and the case of n=2ln=2l is the number of positive poles of the function Φ\Phi. In fact, let us consider the function FF defined in (5.14). In this case, we have degp0=l\deg p_{0}=l if a10a_{1}\neq 0 (or degp0=l1\deg p_{0}=l-1 if a1=0a_{1}=0), degh1=k1\deg h_{1}=k-1, degg0=l+1k\deg g_{0}=l+1-k, therefore deg(h1g0)=l\deg(h_{1}g_{0})=l. Thus, the number of poles of the function FF is equal to (or greater than) the number of its zeroes.

Let

g(z)=b0z2(lk)+2+c1z2(lk)+1+,g(z)=b_{0}z^{2(l-k)+2}+c_{1}z^{2(l-k)+1}+\ldots,
h(z)=c0z2k1+c1z2k2+h(z)=c_{0}z^{2k-1}+c_{1}z^{2k-2}+\ldots

Then the leading coefficient of the polynomial h1h_{1} equals c0c_{0}, and the leading coefficient of the polynomial g0g_{0} equals b0b_{0} according to (2.4)–(2.5). Consequently, the leading coefficient of the polynomial h1g0h_{1}g_{0} is equal to c0b0=a0c_{0}b_{0}=a_{0}. Moreover, the leading coefficient of the polynomial p0p_{0} is equal to a1a_{1} or181818If a1=0a_{1}=0. a3a_{3} by (2.5), so we have limuF(u)=a1a0\displaystyle\lim_{u\to\infty}F(u)=-\frac{a_{1}}{a_{0}} or a3a0-\dfrac{a_{3}}{a_{0}}. The function FF is an R-function (that can be proved in the same way as above), so it is decreasing between its poles, and F(0)=h0(0)h1(0)>0F(0)=-\dfrac{h_{0}(0)}{h_{1}(0)}>0 (see (4.20)). Therefore, the minimal positive zero of the polynomial p0p_{0} is less than the minimal positive zero of the polynomial h1h_{1}.

If the condition (5.3) holds, then limuF(u)=a1a0<0\displaystyle\lim_{u\to\infty}F(u)=-\dfrac{a_{1}}{a_{0}}<0, so the maximal positive zero of p0p_{0} (the maximal positive zero of the function FF) is greater than the maximal positive zero of the polynomial h1h_{1}. This means that the polynomial p0p_{0} has exactly degh1+1=k\deg h_{1}+1=k positive zeroes, so the function Φ\Phi has exactly kk positive poles.

If the condition (5.4) holds, then either F(u)+0F(u)\to+0 as u+u\to+\infty if a1=0a_{1}=0 or F(u)a1a0>0F(u)\to-\dfrac{a_{1}}{a_{0}}>0 as u+u\to+\infty if a10a_{1}\neq 0, so we have limu+F(u)>0\displaystyle\lim_{u\to+\infty}F(u)>0. This implies that the maximal positive zero of h1h_{1} is greater than the maximal positive zero of p0p_{0}, so the polynomial p0p_{0} has exactly k1=degh1k-1=\deg h_{1} positive zeroes. Consequently, Φ\Phi has exactly k1k-1 positive poles. Thus, for n=2l+1n=2l+1 and p(0)0p(0)\neq 0, we obtain that the function Φ=p1p0\Phi=\dfrac{p_{1}}{p_{0}} can be represented as in (5.15) with positive αj\alpha_{j} and with exactly kk (if (5.3) holds) or k1k-1 (if (5.4) holds) positive poles, as required.


Let n=2ln=2l and p(0)=0p(0)=0. Then by definition, the polynomial pp has exactly kk zeroes in the closed right half-plane: 0=μ1<μ2<<μk0=\mu_{1}<\mu_{2}<\ldots<\mu_{k}. Moreover, pp has an odd number of zeroes on each interval (,μk)(-\infty,-\mu_{k}), (μk,μk1)(-\mu_{k},-\mu_{k-1}), …, (μ3,μ2)(-\mu_{3},-\mu_{2}), (μ2,0)(-\mu_{2},0), counting multiplicities. All nonreal zeroes of pp are located in the open left half-plane.

As above, we represent the polynomial pp as a product: p(z)=g(z)h(z)=p0(z2)+zp1(z2)p(z)=g(z)h(z)=p_{0}(z^{2})+zp_{1}(z^{2}), where h(z)=h0(z2)+zh1(z2)h(z)=h_{0}(z^{2})+zh_{1}(z^{2}) is an almost self-interlacing polynomial191919See Section 4.5. (of type I) of degree 2k2k whose nonnegative zeroes are μ1,μ2,,μk\mu_{1},\mu_{2},\ldots,\mu_{k}, and g(z)=g0(z2)+zg1(z2)g(z)=g_{0}(z^{2})+zg_{1}(z^{2}) is a Hurwitz stable polynomial of degree 2(lk)2(l-k), which has an even number of zeroes, counting multiplicities, on each interval (,μk)(-\infty,-\mu_{k}), (μk,μk1)(-\mu_{k},-\mu_{k-1}), …, (μ3,μ2)(-\mu_{3},-\mu_{2}), (μ2,0)(-\mu_{2},0).

In the same way as above, one can establish that Φ\Phi is an R-function. However, p0p_{0} has a root at zero in this case. The polynomial h0h_{0} also has a root at zero, and all its other roots are positive. Since degh=2k\deg h=2k, we have degh0=degh1+1=k\deg h_{0}=\deg h_{1}+1=k. Besides, the zeroes of h0h_{0} interlace the zeroes of h1h_{1}, so all zeroes of the polynomial h1h_{1} are positive. Consequently, the R-function FF defined in (5.14) has k1k-1 positive poles and lkl-k negative poles. It is easy to see that deg(g0h1)=l1\deg(g_{0}h_{1})=l-1 and degp0=l\deg p_{0}=l, so the maximal positive zero of the polynomial p0p_{0} (and of the function FF) is greater than the maximal positive zero of the polynomial h1h_{1}. Also the minimal positive of p0p_{0} is also greater than the minimal positive zero of h1h_{1}, since FF is decreasing between its poles and F(0)=0F(0)=0. Thus, we obtain that p0p_{0} has exactly k1k-1 positive zeroes, since the zeroes of p0p_{0} interlace the zeroes of g0h1g_{0}h_{1}. Consequently, the function Φ\Phi has lkl-k negative poles, k1k-1 positive poles and a pole at zero.


If n=2l+1n=2l+1 and the polynomial p(z)=p0(z2)+zp1(z2)p(z)=p_{0}(z^{2})+zp_{1}(z^{2}) has a root at zero, then in the same as above, one can prove that Φ\Phi is an R-function with either exactly k1k-1 positive poles (if the condition (5.3) holds), or exactly k2k-2 positive poles (if the condition (5.4) holds), since we have F(0)=0F(0)=0 for the R-function FF defined in (5.14). ∎

Theorem 5.3 allow us to use properties of R-functions to obtain additional criteria for polynomials from the class of generalized the Hurwitz polynomials. In particular, we can generalize Hurwitz and Liénard and Chipart criteria.

Let us consider the Hankel matrix S(Φ)=si+ji,j=0S(\Phi)=\|s_{i+j}\|^{\infty}_{i,j=0} constructed with the coefficients of the series (2.13) or (2.18). From Theorems 5.31.18, and 1.24 we obtain the following criteria for generalized Hurwitz polynomials.

Theorem 5.4.

The polynomial pp defined in (5.1), with a10a_{1}\neq 0 if n=2l+1n=2l+1, is generalized Hurwitz if and only if

Dj(Φ)>0,j=1,,l.D_{j}(\Phi)>0,\qquad j=1,\ldots,l. (5.18)

Its order kk equals:

  • 1)

    if n=2ln=2l, then

    k=lV(1,D^1(Φ),D^2(Φ),,D^r(Φ)),k=l-\operatorname{{\rm V}}(1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi)), (5.19)
  • 2)

    if n=2l+1n=2l+1, then

    k=lV(s1,1,D^1(Φ),D^2(Φ),,D^r(Φ))+1,k=l-\operatorname{{\rm V}}(-s_{-1},1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi))+1, (5.20)

where r=lr=l for D^l(Φ)0\widehat{D}_{l}(\Phi)\neq 0, and r=l1r=l-1 for202020By Corollary 1.5, if D^l(Φ)=0\widehat{D}_{l}(\Phi)=0, then D^l1(Φ)0\widehat{D}_{l-1}(\Phi)\neq 0. D^l(Φ)=0\widehat{D}_{l}(\Phi)=0. Here s1=a0a1s_{-1}=\dfrac{a_{0}}{a_{1}}.

The number of sign changes in the sequence 1,D^1(Φ),D^2(Φ),,D^r(Φ)1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi) must be calculated according to the Frobenius rule provided by Theorem 1.6.

Proof.

According to Theorem 5.3, the polynomial pp, with a10a_{1}\neq 0 if n=2l+1n=2l+1, is generalized Hurwitz if and only if its associated function Φ\Phi is an R-function with exactly l=[n2]l=\left[\dfrac{n}{2}\right] poles. But by Theorem 1.18, this is equivalent to the inequalities (5.18).

If n=2ln=2l, or if n=2l+1n=2l+1 with (5.3), then by Theorem 5.3 the number of nonnegative poles of the function Φ\Phi is equal to order of the polynomial pp. From Theorem 1.24 we obtain (5.19), which is equivalent to (5.20) if n=2l+1n=2l+1 and (5.3) holds.

If n=2l+1n=2l+1 and a0a1<0a_{0}a_{1}<0, then by Theorem 5.3, the number of nonnegative poles of Φ\Phi is equal to order of pp minus one. Thus, by Theorem 1.24 we have

k1=lV(1,D^1(Φ),D^2(Φ),,D^r(Φ)),k-1=l-\operatorname{{\rm V}}(1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi)),

that is equivalent to (5.20) in this case. ∎

In the same way as above one can establish the corresponding theorem for polynomials of odd degree with a1=0a_{1}=0.

Theorem 5.5.

Let the polynomial pp of degree n=2l+1n=2l+1 be defined in (5.1) and let a1=0a_{1}=0. The polynomial pp is generalized Hurwitz if and only if a0a3<0a_{0}a_{3}<0 and

Dj(Φ)>0,j=1,,l1.D_{j}(\Phi)>0,\qquad j=1,\ldots,l-1. (5.21)

Its order kk equals:

k=lV(1,D^1(Φ),D^2(Φ),,D^r(Φ)),k=l-\operatorname{{\rm V}}(1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi)), (5.22)

where r=l1r=l-1 for D^l1(Φ)0\widehat{D}_{l-1}(\Phi)\neq 0, and r=l2r=l-2 for212121By Corollary 1.5, if D^l1(Φ)=0\widehat{D}_{l-1}(\Phi)=0, then D^l2(Φ)0\widehat{D}_{l-2}(\Phi)\neq 0. D^l1(Φ)=0\widehat{D}_{l-1}(\Phi)=0.

The number of sign changes in the sequence 1,D^1(Φ),D^2(Φ),,D^r(Φ)1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi) must be calculated according to the Frobenius rule provided by Theorem 1.6.

Proof.

In fact, let the polynomial pp is generalized Hurwitz of order kk. Then by Theorem 5.3, the function Φ=p1p0\Phi=\dfrac{p_{1}}{p_{0}} is an R-function. In particular, this means that |degp0degp1|1|\deg p_{0}-\deg p_{1}|\leqslant 1 (see e.g. [7, 16]), therefore, degp1=l\deg p_{1}=l and degp0=l1\deg p_{0}=l-1, since degp=2l+1\deg p=2l+1 and a1=0a_{1}=0 by assumption. So by Theorems 1.18 and 5.3 the function Φ\Phi has the form

Φ(u)=a0a3u+a2a3a0a5a32+j=1k1αjuωj+j=kl1αju+νj,\Phi(u)=\dfrac{a_{0}}{a_{3}}\,u+\dfrac{a_{2}a_{3}-a_{0}a_{5}}{a_{3}^{2}}+\sum\limits_{j=1}^{k-1}\dfrac{\alpha_{j}}{u-\omega_{j}}+\sum\limits_{j=k}^{l-1}\dfrac{\alpha_{j}}{u+\nu_{j}}, (5.23)

where a0a3<0\dfrac{a_{0}}{a_{3}}<0, αj>0\alpha_{j}>0 for j=1,,l1j=1,\ldots,l-1, ωj0\omega_{j}\geqslant 0, j=1,k1j=1\ldots,k-1, and νj>0\nu_{j}>0 for j=k,,l1j=k,\ldots,l-1. The inequalities (5.21) and the formula (5.22) now follow from Theorems 1.18 and 1.24.

Conversely, from the conditions (5.21)–(5.22), a1=0a_{1}=0 and from the inequality a0a3<0a_{0}a_{3}<0 it follows that the function Φ\Phi has the form (5.23) by Theorems 1.18 and 1.24. Now Theorem 5.3 implies that pp is a generalized Hurwitz polynomial of order kk. ∎

From Theorems 5.4 and 5.5 and from the formulæ (2.14)–(2.15) and (2.21) we obtain the following theorem, which is an analogue of Hurwitz stability criterion for generalized Hurwitz polynomials.

Theorem 5.6 (Generalized Hurwitz theorem).

The polynomial pp given in (5.1) is generalized Hurwitz if and only if

Δn1(p)>0,Δn3(p)>0,Δn5(p)>0,\Delta_{n-1}(p)>0,\ \Delta_{n-3}(p)>0,\ \Delta_{n-5}(p)>0,\ldots (5.24)

The order kk of the polynomial pp equals

k=V(Δn(p),Δn2(p),,1)ifp(0)0,k=\operatorname{{\rm V}}(\Delta_{n}(p),\Delta_{n-2}(p),\ldots,1)\qquad\text{if}\quad p(0)\neq 0, (5.25)

or

k=V(Δn2(p),Δn4(p),,1)+1ifp(0)=0.k=\operatorname{{\rm V}}(\Delta_{n-2}(p),\Delta_{n-4}(p),\ldots,1)+1\qquad\text{if}\quad p(0)=0. (5.26)

Here the number of sign changes in (5.25) and (5.26) must be calculated according to the Frobenius rule provided by Theorem 1.6.

Proof.

Let the real polynomial pp be of degree nn such that the coefficient a1a_{1} is nonzero for n=2l+1n=2l+1. By Theorem 5.4, the polynomial pp is generalized Hurwitz if and only if the inequalities (5.18) hold. These inequalities are equivalent to the inequalities (5.24) according to the formulæ (2.14)–(2.15).

If the polynomial pp is of odd degree and a1=0a_{1}=0, then by Theorem 5.5, the polynomial pp is generalized Hurwitz if and only if the inequalities (5.21) and a0a3>0-a_{0}a_{3}>0 hold. Those inequalities are equivalent to the inequalities (5.24) according to the formulæ (2.21)–(2.22).

We now prove that order of the polynomial can be calculated by the formulæ (5.25)–(5.26).

Let n=2ln=2l, then from (5.19) and (2.14) we obtain

k=lV(1,D^1(Φ),D^2(Φ),,D^r(Φ))=lV(1,Δ2(p),Δ4(p),,(1)rΔ2r(p))==V(1,Δ2(p),Δ4(p),,Δ2r(p))+lr.\begin{array}[]{c}k=l-\operatorname{{\rm V}}(1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi))=l-\operatorname{{\rm V}}(1,-\Delta_{2}(p),\Delta_{4}(p),\ldots,(-1)^{r}\Delta_{2r}(p))=\\ \\ =\operatorname{{\rm V}}(1,\Delta_{2}(p),\Delta_{4}(p),\ldots,\Delta_{2r}(p))+l-r.\end{array}

This is exactly (5.25) for r=lr=l and exactly (5.26) for r=l1r=l-1.

Let now n=2l+1n=2l+1 and a0a1>0a_{0}a_{1}>0. The formulæ (5.20) and (2.14) imply

k=lV(s1,1,D^1(Φ),D^2(Φ),,D^r(Φ))+1==l+1V(s1,1)+V(1,D^1(Φ),D^2(Φ),,D^r(Φ))==lV(1,a13Δ3(p),a15Δ5(p),,(1)ra12r1Δ2r+1(p))==V(1,Δ1(p),Δ3(p),Δ5(p),,Δ2r+1(p))+lr,\begin{array}[]{c}k=l-\operatorname{{\rm V}}(-s_{-1},1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi))+1=\\ \\ =l+1-\operatorname{{\rm V}}(-s_{-1},1)+\operatorname{{\rm V}}(1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi))=\\ \\ =l-\operatorname{{\rm V}}(1,-a_{1}^{-3}\Delta_{3}(p),a_{1}^{-5}\Delta_{5}(p),\ldots,(-1)^{r}a_{1}^{-2r-1}\Delta_{2r+1}(p))=\\ \\ =\operatorname{{\rm V}}(1,\Delta_{1}(p),\Delta_{3}(p),\Delta_{5}(p),\ldots,\Delta_{2r+1}(p))+l-r,\end{array} (5.27)

since a1=Δ1(p)>0a_{1}=\Delta_{1}(p)>0. One can see that (5.27) implies (5.25) for r=lr=l and (5.26) for r=l1r=l-1.

Let n=2l+1n=2l+1 and a0a1<0a_{0}a_{1}<0. The formulæ (5.20) and (2.14) imply

k=lV(s1,1,D^1(Φ),D^2(Φ),,D^r(Φ))+1==lV(s1,1)+V(1,D^1(Φ),D^2(Φ),,D^r(Φ))+1==1+lV(1,a13Δ3(p),a15Δ5(p),,(1)ra12r1Δ2r+1(p))==1+V(a1,a12Δ3(p),a14Δ5(p),,a12rΔ2r+1(p))+lr==V(1,Δ1(p))+V(a1,a12Δ3(p),a14Δ5(p),,a12rΔ2r+1(p))+lr==V(1,Δ1(p),Δ3(p),Δ5(p),,Δ2r+1(p))+lr,\begin{array}[]{c}k=l-\operatorname{{\rm V}}(-s_{-1},1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi))+1=\\ \\ =l-\operatorname{{\rm V}}(-s_{-1},1)+\operatorname{{\rm V}}(1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi))+1=\\ \\ =1+l-\operatorname{{\rm V}}(1,-a_{1}^{-3}\Delta_{3}(p),a_{1}^{-5}\Delta_{5}(p),\ldots,(-1)^{r}a_{1}^{-2r-1}\Delta_{2r+1}(p))=\\ \\ =1+\operatorname{{\rm V}}(a_{1},a_{1}^{-2}\Delta_{3}(p),a_{1}^{-4}\Delta_{5}(p),\ldots,a_{1}^{-2r}\Delta_{2r+1}(p))+l-r=\\ \\ =\operatorname{{\rm V}}(1,\Delta_{1}(p))+\operatorname{{\rm V}}(a_{1},a_{1}^{-2}\Delta_{3}(p),a_{1}^{-4}\Delta_{5}(p),\ldots,a_{1}^{-2r}\Delta_{2r+1}(p))+l-r=\\ \\ =\operatorname{{\rm V}}(1,\Delta_{1}(p),\Delta_{3}(p),\Delta_{5}(p),\ldots,\Delta_{2r+1}(p))+l-r,\end{array} (5.28)

since a1=Δ1(p)<0a_{1}=\Delta_{1}(p)<0. Now from (5.28) we obtain (5.25) for r=lr=l and (5.26) for r=l1r=l-1.

Finally, let n=2l+1n=2l+1 and a1=Δ1(p)=0a_{1}=\Delta_{1}(p)=0. The formulæ (5.22) and (2.21)–(2.22) yield

k=lV(1,D^1(Φ),D^2(Φ),,D^r(Φ))==lV(1,Δ5(p)Δ3(p),Δ7(p)Δ3(p),,(1)rΔ2r+3(p)Δ3(p))==l1r+V(1,0,a0a32)+V(1,Δ5(p)Δ3(p),Δ7(p)Δ3(p),,Δ2r+3(p)Δ3(p))==l1r+V(1,Δ1(p),Δ3(p))+V(Δ3(p),Δ5(p),Δ7(p),,Δ2r+3(p))==l1r+V(1,Δ1(p)Δ3(p),Δ5(p),Δ7(p),,Δ2r+3(p)),\begin{array}[]{c}k=l-\operatorname{{\rm V}}(1,\widehat{D}_{1}(\Phi),\widehat{D}_{2}(\Phi),\ldots,\widehat{D}_{r}(\Phi))=\\ \\ =l-\operatorname{{\rm V}}\left(1,-\dfrac{\Delta_{5}(p)}{\Delta_{3}(p)},\dfrac{\Delta_{7}(p)}{\Delta_{3}(p)},\ldots,(-1)^{r}\dfrac{\Delta_{2r+3}(p)}{\Delta_{3}(p)}\right)=\\ \\ =l-1-r+\operatorname{{\rm V}}(1,0,-a_{0}a_{3}^{2})+\operatorname{{\rm V}}\left(1,\dfrac{\Delta_{5}(p)}{\Delta_{3}(p)},\dfrac{\Delta_{7}(p)}{\Delta_{3}(p)},\ldots,\dfrac{\Delta_{2r+3}(p)}{\Delta_{3}(p)}\right)=\\ \\ =l-1-r+\operatorname{{\rm V}}(1,\Delta_{1}(p),\Delta_{3}(p))+\operatorname{{\rm V}}\left(\Delta_{3}(p),\Delta_{5}(p),\Delta_{7}(p),\ldots,\Delta_{2r+3}(p)\right)=\\ \\ =l-1-r+\operatorname{{\rm V}}(1,\Delta_{1}(p)\Delta_{3}(p),\Delta_{5}(p),\Delta_{7}(p),\ldots,\Delta_{2r+3}(p)),\end{array} (5.29)

where r=l1r=l-1 for p(0)0p(0)\neq 0, and r=l2r=l-2 for p(0)=0p(0)=0. Now (5.29) is equivalent to (5.25) for r=l1r=l-1 and to (5.26) for r=l2r=l-2. ∎

Remark 5.7.

Using the formulæ (2.14)–(2.15) and (2.21)–(2.22), one can easily reformulate Theorem 5.6 in terms of the minors ηj(p)\eta_{j}(p) defined in Definition 2.2.

Now from Theorems 1.28 and 5.3 we obtain the following theorem, which is a generalization of the famous Liénard and Chipart criterion of stability for the class of generalized Hurwitz polynomials.

Theorem 5.8 (Generalized Liénard and Chipart theorem).

The polynomial pp given in (5.1) is generalized Hurwitz if and only if the inequalities (5.24) hold. The order kk of the polynomial pp equals

k=v(an,an2,,1)=v(an,an1,an3,,1)ifan0,k=v(a_{n},a_{n-2},\ldots,1)=v(a_{n},a_{n-1},a_{n-3},\ldots,1)\qquad\text{if}\qquad a_{n}\neq 0, (5.30)

or222222If an=0a_{n}=0, then always an10a_{n-1}\neq 0, since the polynomials p1p_{1} and p0p_{0}, the numerator and the denominator of the R-function Φ\Phi, have no common roots.

k=v(an2,an4,,1)+1=v(an1,an3,,1)+1,ifan=0,k=v(a_{n-2},a_{n-4},\ldots,1)+1=v(a_{n-1},a_{n-3},\ldots,1)+1,\qquad\text{if}\qquad a_{n}=0, (5.31)

where vv is the number of strong sign changes (see Definition 1.27).

Proof.

The fact that the polynomial pp is generalized Hurwitz if and only if the inequalities (5.24) hold follow from Theorem 5.6. Now we prove the formulæ (5.30)–(5.31).

Let n=2ln=2l. By Theorem 5.3, pp is a generalized Hurwitz polynomial if and only if the function

Φ(u)=p1(u)p0(u)=a1ul1+a3ul2++a2l3u+a2l1a0ul+a2ul1++a2l2u+a2l\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}=\dfrac{a_{1}u^{l-1}+a_{3}u^{l-2}+\ldots+a_{2l-3}u+a_{2l-1}}{a_{0}u^{l}+a_{2}u^{l-1}+\ldots+a_{2l-2}u+a_{2l}}

is an R-function. The number of its nonnegative poles is equal to order kk of the polynomial pp. Let Φ\Phi has exactly mm positive poles232323Clearly, m=km=k if an0a_{n}\neq 0, and m=k1m=k-1 if an=0a_{n}=0.. We prove that

m=v(an,an2,,1).m=v(a_{n},a_{n-2},\ldots,1). (5.32)

It will be equivalent to to the first equalities in (5.30)–(5.31), because vv is the number of strong sign changes, so v(an,an2,,1)=v(an2,an4,,1)v(a_{n},a_{n-2},\ldots,1)=v(a_{n-2},a_{n-4},\ldots,1) if an=0a_{n}=0. Now Theorem 1.28 implies

m=v(a2l,a2l2,,a0)=v(a2l,a2l2,,a0)+v(a0,1)=v(a2l,a2l2,,a0,1),m=v(a_{2l},a_{2l-2},\ldots,a_{0})=v(a_{2l},a_{2l-2},\ldots,a_{0})+v(a_{0},1)=v(a_{2l},a_{2l-2},\ldots,a_{0},1),

which is exactly (5.32).

Let us now consider the function

Ψ(u)=1Φ(u)=p0(u)p1(u)=a0ul+a2ul1++a2l2u+a2la1ul1+a3ul2++a2l3u+a2l1.\Psi(u)=-\dfrac{1}{\Phi(u)}=-\dfrac{p_{0}(u)}{p_{1}(u)}=-\dfrac{a_{0}u^{l}+a_{2}u^{l-1}+\ldots+a_{2l-2}u+a_{2l}}{a_{1}u^{l-1}+a_{3}u^{l-2}+\ldots+a_{2l-3}u+a_{2l-1}}.

This function is also an R-function by Theorem 1.23, since Φ\Phi is an R-function with exactly kk nonnegative poles, say 0ω1<,ωk0\leqslant\omega_{1}<\ldots,\omega_{k}. Note that Φ(u)>0\Phi(u)>0 for u(ωk,+)u\in(\omega_{k},+\infty), so it can have at least k1k-1 and at most kk positive zeroes. There can be only four possibilities:

  • I.

    Φ(0)=a2l1a2l<0\Phi(0)=\dfrac{a_{2l-1}}{a_{2l}}<0. Then v(a2l,a2l1)=1v(a_{2l},a_{2l-1})=1, and Φ\Phi has no zeroes on the interval (0,ω1)(0,\omega_{1}), since it is decreasing on this interval and Φ(u)\Phi(u)\to-\infty as uω1u\nearrow\omega_{1}. Therefore, Φ\Phi has exactly k1k-1 positive zeroes, so Ψ\Psi has exactly k1k-1 positive poles. By Theorem 1.28 and Remark 1.29 applied to the function Ψ\Psi, we have

    k1=v(a2l1,a2l3,,a1)=1+v(a2l,a2l1)+v(a2l1,a2l3,,a1,1)+v(a1,1)==1+v(a2l,a2l1,a2l3,,a1,1),\begin{array}[]{c}k-1=v(a_{2l-1},a_{2l-3},\ldots,a_{1})=-1+v(a_{2l},a_{2l-1})+v(a_{2l-1},a_{2l-3},\ldots,a_{1},1)+v(a_{1},1)=\\ \\ =-1+v(a_{2l},a_{2l-1},a_{2l-3},\ldots,a_{1},1),\end{array}

    which is exactly

    k=v(an,an1,an3,,1).k=v(a_{n},a_{n-1},a_{n-3},\ldots,1). (5.33)

    Here we used the inequality Δ1(p)=a1>0\Delta_{1}(p)=a_{1}>0, which follows from the inequalities (5.24) for n=2ln=2l.

  • II.

    Φ(0)=a2l1a2l>0\Phi(0)=\dfrac{a_{2l-1}}{a_{2l}}>0. Then v(a2l,a2l1)=0v(a_{2l},a_{2l-1})=0, and Φ\Phi has exactly kk positive zeroes, so Ψ\Psi has exactly kk positive poles. Thus, from Theorem 1.28 and Remark 1.29 applied to the function Ψ\Psi, we obtain

    k=v(a2l1,a2l3,,a1)=v(a2l,a2l1)+v(a2l1,a2l3,,a1)+v(a1,1)==v(a2l,a2l1,a2l3,,a1,1),\begin{array}[]{c}k=v(a_{2l-1},a_{2l-3},\ldots,a_{1})=v(a_{2l},a_{2l-1})+v(a_{2l-1},a_{2l-3},\ldots,a_{1})+v(a_{1},1)=\\ \\ =v(a_{2l},a_{2l-1},a_{2l-3},\ldots,a_{1},1),\end{array}

    which is exactly (5.33).

  • III.

    Φ(0)=0\Phi(0)=0. Then Ψ\Psi has a pole at zero, that is, a2l1=0a_{2l-1}=0. Since Ψ\Psi is an R-function, it has positive residues at each its pole, so limu0uΨ(u)=a2la2l3>0\lim\limits_{u\to 0}u\Psi(u)=-\dfrac{a_{2l}}{a_{2l-3}}>0. Thus, v(a2l,a2l3)=v(a2l,a2l1,a2l3)=1v(a_{2l},a_{2l-3})=v(a_{2l},a_{2l-1},a_{2l-3})=1. As in Case I, Φ\Phi has exactly k1k-1 positive zeroes, so the function Ψ\Psi has exactly k1k-1 positive poles. Therefore, by Theorem 1.28 and Remark 1.29 applied to the function Ψ\Psi, we have

    k1=v(a2l3,a2l5,,a1)=1+v(a2l,a2l1,a2l3)+v(a2l3,a2l5,,a1)+v(a1,1)==1+v(a2l,a2l1,a2l3,,a1,1),\begin{array}[]{c}k-1=v(a_{2l-3},a_{2l-5},\ldots,a_{1})=-1+v(a_{2l},a_{2l-1},a_{2l-3})+v(a_{2l-3},a_{2l-5},\ldots,a_{1})+v(a_{1},1)=\\ \\ =-1+v(a_{2l},a_{2l-1},a_{2l-3},\ldots,a_{1},1),\end{array}

    which is precisely (5.33).

  • IV.

    Ψ(0)=0\Psi(0)=0. Then p(0)=an=a2l=0p(0)=a_{n}=a_{2l}=0, and the function Φ\Phi has a pole at zero, so it has exactly k1k-1 positive zeroes. Thus, the function Ψ\Psi has exactly k1k-1 positive poles, and by Theorem 1.28 and Remark 1.29 applied to the function Ψ\Psi, we have

    k1=v(a2l1,a2l3,,a1)=v(a2l1,a2l3,,a1,1),\begin{array}[]{c}k-1=v(a_{2l-1},a_{2l-3},\ldots,a_{1})=v(a_{2l-1},a_{2l-3},\ldots,a_{1},1),\end{array}

    therefore,

    k=v(an1,an3,,1)+1.k=v(a_{n-1},a_{n-3},\ldots,1)+1. (5.34)

Let now n=2l+1n=2l+1. By Theorem 5.3, pp is a generalized Hurwitz polynomial if and only if Φ\Phi is an R-function with exactly ll zeroes and ll or l1l-1 poles. If kk is order of the polynomial pp, then Φ\Phi has exactly kk nonnegative poles for a1>0a_{1}>0 and exactly k1k-1 nonnegative poles for a10a_{1}\leqslant 0.

At first, we establish the formula (5.32), where mm is the number of positive poles of the function Φ\Phi. As we mentioned above, this formula is equivalent to the first equalities in (5.30)–(5.31).

If a1>0a_{1}>0, then from Theorem 1.28 applied to the function Φ\Phi we obtain

m=v(a2l+1,a2l1,,a1)=v(a2l+1,a2l1,,a1,1).m=v(a_{2l+1},a_{2l-1},\ldots,a_{1})=v(a_{2l+1},a_{2l-1},\ldots,a_{1},1).

This is exactly (5.32).

Let a10a_{1}\leqslant 0. As we established above, if a1=0a_{1}=0, then a3<0a_{3}<0. Therefore, for a10a_{1}\leqslant 0, we always have v(a3,a1)=v(a3,a1,1)1v(a_{3},a_{1})=v(a_{3},a_{1},1)-1. Thus, Theorem 1.28 implies

m=v(a2l+1,a2l1,,a1)=v(a2l+1,a2l1,,a1,1)1,m=v(a_{2l+1},a_{2l-1},\ldots,a_{1})=v(a_{2l+1},a_{2l-1},\ldots,a_{1},1)-1,

which is again (5.32), since for a10a_{1}\leqslant 0, m=k1m=k-1 if p(0)0p(0)\neq 0, and m=k2m=k-2 if p(0)=0p(0)=0.

As above, consider the function

Ψ(u)=1Φ(u)=p0(u)p1(u)=a1ul+a3ul1++a2l1u+a2l+1a0ul+a2ul1++a2l2u+a2l.\Psi(u)=-\dfrac{1}{\Phi(u)}=-\dfrac{p_{0}(u)}{p_{1}(u)}=-\dfrac{a_{1}u^{l}+a_{3}u^{l-1}+\ldots+a_{2l-1}u+a_{2l+1}}{a_{0}u^{l}+a_{2}u^{l-1}+\ldots+a_{2l-2}u+a_{2l}}.

This function is also an R-function by Theorem 1.23. As above, there can be only four possibilities:

  • I.

    Φ(0)=a2la2l+1<0\Phi(0)=\dfrac{a_{2l}}{a_{2l+1}}<0, so v(a2l+1,a2l)=1v(a_{2l+1},a_{2l})=1. If a1>0a_{1}>0, then by Theorem 5.3, Φ\Phi has exactly kk positive poles, say 0<ω1<<ωk0<\omega_{1}<\ldots<\omega_{k}. Since Φ(u)a0a1>0\Phi(u)\to\dfrac{a_{0}}{a_{1}}>0 as u+u\to+\infty and Φ\Phi is decreasing between its poles, the function Φ\Phi has no zeroes on the interval (ωk,+)(\omega_{k},+\infty). Also Φ\Phi has no zeroes on the interval (0,ω1)(0,\omega_{1}), since Φ(0)<0\Phi(0)<0 by assumption. Thus, Φ\Phi has exactly k1k-1 positive zeroes, so Ψ\Psi has exactly k1k-1 positive poles.

    If a10a_{1}\leqslant 0, then by Theorem 5.3, the function Φ\Phi has exactly k1k-1 positive poles, say 0<ω1<<ωk10<\omega_{1}<\ldots<\omega_{k-1}. Since Φ(u)\Phi(u) is decreasing between its poles, it is negative for sufficiently large positive uu. Moreover, Φ(u)+\Phi(u)\nearrow+\infty as uωk1u\searrow\omega_{k-1}. Thus, Φ\Phi has a zero on the interval (ωk1,+)(\omega_{k-1},+\infty), but it has no zeroes (0,ω1)(0,\omega_{1}), since Φ(0)<0\Phi(0)<0 by assumption. Consequently, Φ\Phi has exactly k1k-1 positive zeroes, so Ψ\Psi has exactly k1k-1 positive poles.

    Thus, we showed that the function Ψ\Psi has exactly k1k-1 positive poles regardless of the sign of a1a_{1}. Now from Theorem 1.28 and Remark 1.29 applied to the function Ψ\Psi, we obtain

    k1=v(a2l,a2l2,,a0)=1+v(a2l+1,a2l)+v(a2l,a2l2,,a0,1)==1+v(a2l+1,a2l,a2l2,,a0,1),\begin{array}[]{c}k-1=v(a_{2l},a_{2l-2},\ldots,a_{0})=-1+v(a_{2l+1},a_{2l})+v(a_{2l},a_{2l-2},\ldots,a_{0},1)=\\ \\ =-1+v(a_{2l+1},a_{2l},a_{2l-2},\ldots,a_{0},1),\end{array}

    which is exactly (5.33).

  • II.

    Φ(0)=a2la2l+1>0\Phi(0)=\dfrac{a_{2l}}{a_{2l+1}}>0, so v(a2l+1,a2l)=0v(a_{2l+1},a_{2l})=0. In the same way as above, one can show that the function Ψ\Psi has exactly kk positive poles regardless of the sign of a1a_{1}. Thus, from Theorem 1.28 and Remark 1.29 applied to the function Ψ\Psi, we obtain

    k=v(a2l,a2l2,,a0)=v(a2l+1,a2l)+v(a2l,a2l2,,a0,1)==v(a2l+1,a2l,a2l2,,a0,1).\begin{array}[]{c}k=v(a_{2l},a_{2l-2},\ldots,a_{0})=v(a_{2l+1},a_{2l})+v(a_{2l},a_{2l-2},\ldots,a_{0},1)=\\ \\ =v(a_{2l+1},a_{2l},a_{2l-2},\ldots,a_{0},1).\end{array}

    So the formula (5.33) is also valid in this case.

  • III.

    Φ(0)=0\Phi(0)=0. Then Ψ\Psi has a pole at zero, that is, a2l=0a_{2l}=0. Since Ψ\Psi is an R-function, it has positive residues at each its pole, so limu0uΨ(u)=a2l+1a2l2>0\lim\limits_{u\to 0}u\Psi(u)=-\dfrac{a_{2l+1}}{a_{2l-2}}>0. Thus, v(a2l+1,a2l2)=v(a2l+1,a2l,a2l2)=1v(a_{2l+1},a_{2l-2})=v(a_{2l+1},a_{2l},a_{2l-2})=1. As above, one can show that Ψ\Psi has exactly k1k-1 positive poles regardless of the sign of a1a_{1}. Therefore, by Theorem 1.28 and Remark 1.29 applied to the function Ψ\Psi, we have

    k1=v(a2l2,a2l4,,a0)=1+v(a2l+1,a2l,a2l2)+v(a2l2,a2l4,,a0,1)==1+v(a2l+1,a2l,a2l2,,a0,1),\begin{array}[]{c}k-1=v(a_{2l-2},a_{2l-4},\ldots,a_{0})=-1+v(a_{2l+1},a_{2l},a_{2l-2})+v(a_{2l-2},a_{2l-4},\ldots,a_{0},1)=\\ \\ =-1+v(a_{2l+1},a_{2l},a_{2l-2},\ldots,a_{0},1),\end{array}

    which is precisely (5.33).

  • IV.

    Ψ(0)=0\Psi(0)=0. Then p(0)=an=a2l+1=0p(0)=a_{n}=a_{2l+1}=0, and the function Φ\Phi has a pole at zero. In the same way as above, one can prove that the function Ψ\Psi has exactly k1k-1 positive poles regardless of the sign of a1a_{1}, so Theorem 1.28 and Remark 1.29 applied to the function Ψ\Psi imply

    k1=v(a2l,a2l2,,a0),\begin{array}[]{c}k-1=v(a_{2l},a_{2l-2},\ldots,a_{0}),\end{array}

    which is equivalent to (5.34), since a0>0a_{0}>0.

This theorem implies a necessary condition for polynomials to be generalized Hurwitz. This condition plays a role of Stodola’s necessary condition for Hurwitz stable polynomials (Theorem 3.2).

Corollary 5.9.

If the polynomial

p(z)=a0zn+a1zn1++an,a1,,an,a0>0,p(z)=a_{0}z^{n}+a_{1}z^{n-1}+\dots+a_{n},\qquad a_{1},\dots,a_{n}\in\mathbb{R},\ a_{0}>0, (5.35)

is generalized Hurwitz (of type I) of order kk, then

k=v(an,an2,,1)=v(an,an1,an3,,1)ifp(0)0,k=v(a_{n},a_{n-2},\ldots,1)=v(a_{n},a_{n-1},a_{n-3},\ldots,1)\qquad\text{if}\quad p(0)\neq 0,

or

k1=v(an,an2,,1)=v(an,an1,an3,,1)ifp(0)=0,k-1=v(a_{n},a_{n-2},\ldots,1)=v(a_{n},a_{n-1},a_{n-3},\ldots,1)\qquad\text{if}\quad p(0)=0,

This corollary implies Theorem 3.2 (Stodola’s theorem) for k=0k=0, and it implies Theorem 4.10 for k=[n+12]k=\left[\dfrac{n+1}{2}\right] and an0a_{n}\neq 0.

Note that G. F. Korsakov [21] made an attempt to prove Theorem 5.8. However, his methods did not allow him to prove the whole theorem. He proved only sufficiency, that is, he proved that if the inequalities (5.24) hold for the polynomial (5.35), then pp is generalized Hurwitz, and the number kk of its positive zeroes can be found by the formulæ

  • for n=2ln=2l,

    k=v(an,an2,,a2,a0);k=v(a_{n},a_{n-2},\ldots,a_{2},a_{0}); (5.36)
  • for n=2l+1n=2l+1,

    k=v(an,an2,,a3,a1,a0).k=v(a_{n},a_{n-2},\ldots,a_{3},a_{1},a_{0}). (5.37)

Since in [21], it was assumed that an0a_{n}\neq 0, the number kk in (5.36)–(5.37) is exactly order of the polynomial pp. Nevertheless, the generalized Hurwitz polynomials with a root at zero are also partially described in [21]. In fact, there was proved that if the following inequalities hold for the polynomial pp defined in (5.35)

Δn(p)>0,Δn2(p)>0,Δn4(p)>0,,\Delta_{n}(p)>0,\ \Delta_{n-2}(p)>0,\ \Delta_{n-4}(p)>0,\ldots, (5.38)

then pp has exactly kk (defined by (5.36)–(5.37)) zeroes on the positive real half-axis, and other its zeroes are located in the open left half-plane. Generally speaking, the polynomial pp satisfied the inequalities (5.38) is not a generalized Hurwitz polynomial. However, if pp satisfies (5.38), then the polynomial q(z)=zp(z)q(z)=zp(z) of degree n+1n+1 is a generalized Hurwitz polynomial of order k+1k+1. Indeed, it is easy to see that the following equalities are true:

Δi(q)=Δi(p),i=1,,n.\Delta_{i}(q)=\Delta_{i}(p),\qquad i=1,\ldots,n.

So the inequalities (5.38) for the polynomial pp of degree nn are exactly the inequalities (5.24) hold for the polynomial qq of degree n+1n+1, so qq is a generalized Hurwitz polynomial having k+1k+1 nonnegative zeroes.

At last, we establish a relation between generalized Hurwitz polynomials and continued fractions of Stieltjes type (if any).

Theorem 5.10.

If the polynomial pp of degree nn defined in (5.35) is generalized Hurwitz (of type I) of order kk and Δn2j(p)0\Delta_{n-2j}(p)\neq 0, i=1,,li=1,\ldots,l, where Δ0(p)1\Delta_{0}(p)\equiv 1, then the function Φ\Phi has the following Stieltjes type continued fraction expansion:

Φ(u)=p1(u)p0(u)=c0+1c1u+1c2+1c3u+1+1T,T={c2lifp(0)0,c2l1uifp(0)=0,\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}=c_{0}+\dfrac{1}{c_{1}u+\cfrac{1}{c_{2}+\cfrac{1}{c_{3}u+\cfrac{1}{\ddots+\cfrac{1}{T}}}}},\qquad T=\begin{cases}&c_{2l}\qquad\text{if}\;\;\;p(0)\neq 0,\\ &c_{2l-1}u\quad\text{if}\;\;\;p(0)=0,\end{cases} (5.39)

where c0=0c_{0}=0 if nn is even, and c00c_{0}\neq 0 if nn is odd, and

c2j1>0,j=1,,l.c_{2j-1}>0,\qquad j=1,\ldots,l. (5.40)

Moreover, if p(0)0p(0)\neq 0, then the number of negative coefficients c2jc_{2j}, j=0,1,,lj=0,1,\ldots,l, equals kk, and if p(0)=0p(0)=0, then the number of negative coefficients c2jc_{2j}, j=0,1,,l1j=0,1,\ldots,l-1, equals k1k-1.

Conversely, if for given polynomial pp of degree nn, its associated function Φ\Phi has a Stieltjes continued fraction (5.39)–(5.40), then pp is a generalized Hurwitz polynomial (with Δn2j(p)0\Delta_{n-2j}(p)\neq 0, i=1,,li=1,\ldots,l), where ll defined in (2.2). Its order equals the number of negative coefficients c2jc_{2j}, j=0,1,,lj=0,1,\ldots,l, if p(0)0p(0)\neq 0, or the number of negative coefficients c2jc_{2j}, j=0,1,,l1j=0,1,\ldots,l-1, plus one if p(0)=0p(0)=0.

Proof.

The theorem follows from Theorem 5.6 and from the formulæ (3.13)–(3.14) ∎

Theorem 5.10 does not cover the case of a1=0a_{1}=0. The following theorem fills this gap.

Theorem 5.11.

If the polynomial pp of degree n=2l+1n=2l+1 defined in (5.35) with a1=0a_{1}=0 is generalized Hurwitz (of type I) of order kk and Δ2j+1(p)0\Delta_{2j+1}(p)\neq 0, i=1,,l1i=1,\ldots,l-1, then the function Φ\Phi has the following Stieltjes continued fraction expansion:

Φ(u)=p1(u)p0(u)=c1u+c0+1c1u+1c2+1c3u+1+1T,T={c2l2ifp(0)0,c2l3uifp(0)=0,\Phi(u)=\dfrac{p_{1}(u)}{p_{0}(u)}=-c_{-1}u+c_{0}+\dfrac{1}{c_{1}u+\cfrac{1}{c_{2}+\cfrac{1}{c_{3}u+\cfrac{1}{\ddots+\cfrac{1}{T}}}}},\qquad T=\begin{cases}&c_{2l-2}\qquad\text{if}\;\;\;p(0)\neq 0,\\ &c_{2l-3}u\quad\text{if}\;\;\;p(0)=0,\end{cases} (5.41)

where c0c_{0}\in\mathbb{R}, and

c2j1>0,j=0,1,,l1.c_{2j-1}>0,\qquad j=0,1,\ldots,l-1. (5.42)

Moreover, if p(0)0p(0)\neq 0, then the number of negative coefficients c2jc_{2j}, j=1,,l1j=1,\ldots,l-1, equals k1k-1. But if p(0)=0p(0)=0, then the number of negative coefficients c2jc_{2j}, j=0,1,,l2j=0,1,\ldots,l-2, equals k2k-2.

Conversely, if for given polynomial pp of degree n=2l+1n=2l+1, its associated function Φ\Phi has a Stieltjes continued fraction (5.41)–(5.42), then pp is a generalized Hurwitz polynomial (with Δ2j+1(p)0\Delta_{2j+1}(p)\neq 0, i=1,,l1i=1,\ldots,l-1 and Δ1(p)=a1=0\Delta_{1}(p)=a_{1}=0). Its order equals the number of negative coefficients c2jc_{2j}, j=1,,l1j=1,\ldots,l-1, plus one if p(0)0p(0)\neq 0, or the number of negative coefficients c2jc_{2j}, j=1,,l2j=1,\ldots,l-2, plus two if p(0)=0p(0)=0.

Proof.

Let the polynomial pp of degree n=2l+1n=2l+1 with a1=0a_{1}=0 be generalized Hurwitz (of type I) of order kk and Δ2j+1(p)0\Delta_{2j+1}(p)\neq 0, i=1,,l1i=1,\ldots,l-1. By Theorem 5.5 and by formulæ (2.21), we obtain a0a3<0a_{0}a_{3}<0 and

{Dj(Φ)>0,j=1,,l1,D^j(Φ)0j=1,,l2.\begin{cases}&D_{j}(\Phi)>0,\qquad j=1,\ldots,l-1,\\ &\widehat{D}_{j}(\Phi)\neq 0\qquad j=1,\ldots,l-2.\end{cases} (5.43)

It is easy to see that for the function F(u):=Φ(u)+c1uF(u):=\Phi(u)+c_{-1}u, where c1=a0a3c_{-1}=-\dfrac{a_{0}}{a_{3}}, the following equalities hold

{Dj(F)=Dj(Φ),D^j(F)=D^j(Φ)0,j=1,,l1,\begin{cases}&D_{j}(F)=D_{j}(\Phi),\\ &\widehat{D}_{j}(F)=\widehat{D}_{j}(\Phi)\neq 0,\end{cases}\qquad j=1,\ldots,l-1, (5.44)

so the function FF has a Stiltjes type continued fraction (1.33) with c2j1>0c_{2j-1}>0, j=1,,l1j=1,\ldots,l-1, that follows from (5.43)–(5.44) and (1.38). Also from (5.22), (5.43)–(5.44) and (1.37) we obtain that the number of negative coefficients c2jc_{2j}, j=1,,l1j=1,\ldots,l-1, equals k1k-1 if p(0)0p(0)\neq 0, or k2k-2 if p(0)=0p(0)=0.

The converse assertion of the theorem follows from the formulæ (5.43)–(5.44), from Theorem 1.36 applied to the function F(u):=Φ(u)a0a3uF(u):=\Phi(u)-\dfrac{a_{0}}{a_{3}}u, and from Theorem 5.5. ∎

Theorem 1.22 implies the following theorem.

Theorem 5.12.

Let the polynomial pp of degree n2n\geqslant 2 as in (5.35) be generalized Hurwitz polynomial. Then all polynomials

pj(z)=i=0n2j[ni2]([ni2]1)([ni2]+j1)aizn2ji,k=1,,[n2]1,p_{j}(z)=\sum\limits_{i=0}^{n-2j}\left[\dfrac{n-i}{2}\right]\left(\left[\dfrac{n-i}{2}\right]-1\right)\cdots\left(\left[\dfrac{n-i}{2}\right]+j-1\right)a_{i}z^{n-2j-i},\quad k=1,\ldots,\left[\dfrac{n}{2}\right]-1,

are also generalized Hurwitz.

5.2 Application to bifurcation theory. Polynomials dependent on parameters.

Let us consider a system of ordinary differential equations dependent on a real parameter, say α\alpha:

{x˙(t)=F(x(t),α)=A(α)x(t)+o(x(t)),x(t0)=x0F(x),xn,\begin{cases}&\dot{x}(t)=F(x(t),\alpha)=A(\alpha)x(t)+o(x(t)),\\ &x(t_{0})=x_{0}\end{cases}\qquad F(x),\,\,x\in\mathbb{R}^{n}, (5.45)

where the n×nn\times n matrix A(α)A(\alpha) is the linear part of the function F(x,α)F(x,\alpha), and F(0,α)0F(0,\alpha)\equiv 0. Suppose that for some value of the parameter α0\alpha_{0} all the eigenvalues of the matrix A(α0)A(\alpha_{0}) are located in the open left half-plane of the complex plane. In this case, the stability theory says [26, 28, 18] that the zero solution of the system (5.45) is (asymptotically) stable, that is, all the solutions of the system (5.45), whose initial conditions x0x_{0} are sufficiently small, tend to zero as t+t\to+\infty.

When the parameter α\alpha arrives at its critical value, say α1\alpha_{1}^{*} the following two basic cases can occur, generically: either a pair of complex conjugate eigenvalues (of multiplicity one) of the matrix A(α1)A(\alpha_{1}^{*}) appear on the imaginary axis or an eigenvalue (of multiplicity one) becomes 0. In the former case, for α=α1+ε\alpha=\alpha_{1}^{*}+\varepsilon, where ε>0\varepsilon>0 is sufficiently small, there appears a small-amplitude limit cycle branching from the zero solution (a fixed point) of the system (5.45), but the zero solution loses its stability242424Here we assume that for α>α1\alpha>\alpha_{1}^{*}, the matrix A(α)A(\alpha) has an eigenvalue with a positive real part.. This type of bifurcation is called Hopf bifurcation [26, 28, 18]. In hydrodynamics [42], such type of bifurcation is called an oscillatory loss of stability or oscillatory instability. In the case, when the matrix A(α1)A(\alpha_{1}^{*}) has all the eigenvalues in the open left half-plane except one, which is equal to zero and is of multiplicity one, new fixed points of the system (5.45) branch from the zero fixed point. In hydrodynamics [42], such type of bifurcation is called a monotone loss of stability or monotone instability. To determine the type of bifurcation of a fixed point of a given system is a very important problem in hydrodynamics and mechanics (see [28, 42] and references there).

Let us consider the characteristic polynomial p(z,α)=det(zEA(α))p(z,\alpha)=\det(zE-A(\alpha)) of the matrix A(α)A(\alpha). By assumption p(z,α0)p(z,\alpha_{0}) is Hurwitz stable. Recall that α1>α0\alpha_{1}^{*}>\alpha_{0} is the critical value of the parameter α\alpha such that the zero solution of the system (5.45) loses its stability. The monotonicity principle takes place if and only if the polynomial p(z,α1)p(z,\alpha_{1}^{*}) is a generalized Hurwitz polynomial, that is, the inequalities (5.24) hold. This fact follows from Theorem 5.6. Moreover, the polynomial p(z,α)p(z,\alpha) is generalized Hurwitz of order 0 (Hurwitz stable) for α<α1\alpha<\alpha_{1}*, and p(z,α1)p(z,\alpha_{1}^{*}) is a generalized Hurwitz polynomial of order 11.

Now we describe further changes of zero location of the polynomial p(z,α)p(z,\alpha) when α\alpha goes beyond α1\alpha_{1}^{*} until p(z,α)p(z,\alpha) becomes self-interlacing. Let {α2,α3,,αm}\{\alpha_{2}^{*},\alpha_{3}^{*},\ldots,\alpha_{m}^{*}\}, 2m[n+12]2\leqslant m\leqslant\left[\dfrac{n+1}{2}\right], be the next critical values of the parameter α\alpha at each of which order of the generalized Hurwitz polynomial p(z,α)p(z,\alpha) increases. This means that p(z,αi)p(z,\alpha_{i}^{*}) has a simple root at zero, and for α>αi\alpha>\alpha_{i}^{*} sufficiently close to αi\alpha_{i}^{*}, this root becomes a positive simple root of p(z,α)p(z,\alpha) with the minimal absolute value (among real roots of p(z,α)p(z,\alpha)). Therefore, when α\alpha transfers αm\alpha_{m}^{*}, the polynomial p(z,α)p(z,\alpha) has exactly mm positive simple zeroes for α>αm\alpha>\alpha_{m}^{*} sufficiently close to αm\alpha_{m}^{*}. Denote the minimal positive zero of p(z,α)p(z,\alpha) by μm\mu_{m}. Obviously, if the difference ααm>0\alpha-\alpha_{m}^{*}>0 is sufficiently small, then the number μm>0\mu_{m}>0 is close to zero, so p(z,α)p(z,\alpha) has no zeroes on the interval (μm,0)(-\mu_{m},0). Let α\alpha increase. One more simple positive zero of the polynomial p(z,α)p(z,\alpha) can appear252525It is possible only if m<[n+12]m<\left[\dfrac{n+1}{2}\right]., when two zeroes of a pair of complex conjugate zeroes of p(z,α)p(z,\alpha) (all its nonreal zeroes lie in the open left half-plane) face each other on the interval (μm,0)(-\mu_{m},0) to become a real zero of p(z,α)p(z,\alpha) of multiplicity two. Further, when α\alpha grows, this zero splits into to real simple roots on the interval (μm,0)(-\mu_{m},0). One of those roots goes toward zero and crosses it, when α\alpha transfers through the next critical value αm+1\alpha_{m+1}^{*} of the parameter α\alpha. Thus, for sufficiently small difference ααm+1>0\alpha-\alpha_{m+1}^{*}>0, this zero becomes the smallest positive zero μm+1\mu_{m+1}. The second zero of the mentioned pair of zeroes remains on the interval (μm,μm+1)(-\mu_{m},-\mu_{m+1}), since it cannot cross the value μm-\mu_{m} and cannot go away from the real axis.

Let us show that there are no any other scenarios for positive zeroes to appear if p(z,α)p(z,\alpha) is generalized Hurwitz for all α\alpha. In fact, suppose that for α>αm\alpha>\alpha_{m}^{*}, some negative zeroes of p(z,α)p(z,\alpha) come into the interval (μm,0)(-\mu_{m},0) from the real axis, that is, from the interval (,μm)(-\infty,-\mu_{m}). But in this case, the polynomial p(z,α)p(z,\alpha) will have zeroes ±μm\pm\mu_{m} at some value of α\alpha, so by the Orlando’s formula [12, 16], we have Δn+1(p)=0\Delta_{n+1}(p)=0. According to Theorem 5.6, this contradicts the assumption that p(z,α)p(z,\alpha) is generalized Hurwitz.

Thus, when α\alpha increases from α0\alpha_{0}, at each interval (μj,μj+1)(-\mu_{j},-\mu_{j+1}) the polynomial p(z,α)p(z,\alpha) has an odd number of zeroes. One of those zeroes appears on (μj,μj+1)(-\mu_{j},-\mu_{j+1}) as it was described above, while the other zeroes appear from complex conjugate pairs. The maximal number of positive zeroes of p(z,α)p(z,\alpha) equals [n+12]\left[\dfrac{n+1}{2}\right]. In this case, p(z,α)p(z,\alpha) becomes self-interlacing.

So the monotonicity principle takes place for the system (5.45) if and only if for the first critical value α1\alpha_{1}^{*} of the parameter α\alpha the characteristic polynomial p(z,α1)=det(zEA(α1))p(z,\alpha_{1}^{*})=\det(zE-A(\alpha_{1}^{*})) is generalized Hurwitz, that is, the inequalities (5.24) hold.

Acknowledgement

I am grateful to Yu.S. Barkovsky and O. Holtz for helpful discussions and for Barkovsky’s detailed suggestions for the improvement of some results in the paper.

References

  • [1] N. I. Akhiezer. The classical moment problem and some related questions in analysis. Translated by N. Kemmer. Hafner Publishing Co., New York, 1965.
  • [2] N. I. Akhiezer and M. G. Krein. Some questions in the theory of moments. translated by W. Fleming and D. Prill. Translations of Mathematical Monographs, Vol. 2. American Mathematical Society, Providence, R.I., 1962.
  • [3] B. A. Asner, Jr. On the total nonnegativity of the Hurwitz matrix. SIAM J. Appl. Math., 18:407–414, 1970.
  • [4] F. V. Atkinson. Discrete and continuous boundary problems. Mathematics in Science and Engineering, Vol. 8. Academic Press, New York, 1964.
  • [5] Yu. S. Barkovsky. Private communication, 2004.
  • [6] Yu. S. Barkovsky. Lectures on the Routh-Hurwitz problem. arXiv:0802.1805, 2008.
  • [7] N. G. Čebotarev and N. N. Meĭman. The Routh-Hurwitz problem for polynomials and entire functions. Real quasipolynomials with r=3r=3, s=1s=1. Trudy Mat. Inst. Steklov., 26:331, 1949. Appendix by G. S. Barhin and A. N. Hovanskiĭ.
  • [8] S. Fisk. Polynomials, roots, and interlacing. arXiv:0612833, 2006.
  • [9] G. Frobenius. Über das Trägheitsgesetz der quadratischen Formen. Sitz.-Ber. Acad. Wiss. Phys.-Math. Klasse, Berlin, pages 241–256; 407–431, 1894.
  • [10] F. P. Gantmacher and M. G. Krein. Oscillation matrices and kernels and small vibrations of mechanical systems. AMS Chelsea Publishing, Providence, RI, revised edition, 2002. Translation based on the 1941 Russian original, Edited and with a preface by Alex Eremenko.
  • [11] F. R. Gantmacher. The theory of matrices. Vol. 1. Translated by K. A. Hirsch. Chelsea Publishing Co., New York, 1959.
  • [12] F. R. Gantmacher. The theory of matrices. Vol. 2. Translated by K. A. Hirsch. Chelsea Publishing Co., New York, 1959.
  • [13] C. Hermite. On the number of roots of an algebraic equation contained between given limits. Internat. J. Control, 26(2):183–195, 1977. Translated from the French original by P. C. Parks, Routh centenary issue.
  • [14] O. Holtz. Hermite-Biehler, Routh-Hurwitz, and total positivity. Linear Algebra Appl., 372:105–110, 2003.
  • [15] O. Holtz. The inverse eigenvalue problem for symmetric anti-bidiagonal matrices. Linear Algebra Appl., 408:268–274, 2005.
  • [16] O. Holtz and Tyaglov M. Structured matrices, continued fractions, and root localization of polynomials. arXiv:0912.4703, 2009.
  • [17] A. Hurwitz. Über die Bedingungen, unter welchen eine Gleichung nur Wurzeln mit negativen reellen Theilen besitzt. In Stability theory (Ascona, 1995), volume 121 of Internat. Ser. Numer. Math., pages 239–249. Birkhäuser, Basel, 1996. Reprinted from Math. Ann. 46 (1895), 273–284 [JFM 26.0119.03].
  • [18] G. Iooss and D. Joseph. Elementary stability and bifurcation theory. Undergraduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1990.
  • [19] I. V. Kac, , and M. G. Krein. On the spectral functions of the string, volume 103 of Amer. Math. Soc. Transl. American Mathematical Society, 1974.
  • [20] J. H. B. Kemperman. A Hurwitz matrix is totally positive. SIAM J. Math. Anal., 13(2):331–341, 1982.
  • [21] G. F. Korsakov. A generalization of a theorem of Liénard and Chipart. Mat. Zametki, 22(1):13–21, 1977.
  • [22] M. G. Kreĭn and M. A. Naĭmark. The method of symmetric and Hermitian forms in the theory of the separation of the roots of algebraic equations. Linear and Multilinear Algebra, 10(4):265–308, 1981. Translated from the Russian by O. Boshko and J. L. Howland.
  • [23] M. G. Kreĭn and A. A. Nudelman. The Markov moment problem and extremal problems. American Mathematical Society, Providence, R.I., 1977. Ideas and problems of P. L. Čebyšev and A. A. Markov and their further development, Translated from the Russian by D. Louvish, Translations of Mathematical Monographs, Vol. 50.
  • [24] L. Kronecker. Algebraische Reduction der Schaaren bilinearer Formen. S.-B. Akad. Berlin, 1890.
  • [25] A. Liénard and M. Chipart. Sur la signe de la partie réelle des racines d’une équation algébraique.
  • [26] A. M. Lyapunov. The general problem of the stability of motion. Taylor & Francis Ltd., London, 1992. Translated from Edouard Davaux’s French translation (1907) of the 1892 Russian original and edited by A. T. Fuller, With an introduction and preface by Fuller, a biography of Lyapunov by V. I. Smirnov, and a bibliography of Lyapunov’s works compiled by J. F. Barrett, Lyapunov centenary issue, Reprint of Internat. J. Control 55 (1992), no. 3 [ MR1154209 (93e:01035)], With a foreword by Ian Stewart.
  • [27] A. A. Markov. Collected works (Russian). Academy of Scines USSR, Moscow, 1948.
  • [28] J. E. Marsden and M. McCracken. The Hopf bifurcation and its applications. Springer-Verlag, New York, 1976. With contributions by P. Chernoff, G. Childs, S. Chow, J. R. Dorroh, J. Guckenheimer, L. Howard, N. Kopell, O. Lanford, J. Mallet-Paret, G. Oster, O. Ruiz, S. Schecter, D. Schmidt and S. Smale, Applied Mathematical Sciences, Vol. 19.
  • [29] G. Pick. Über die Beschränkungen analytische Funktionen, welche durch vorgegebene Funktionswerte bewirkt werden. Math. Ann., 77:7–23, 1916.
  • [30] V. Pivovarchik. On spectra of a certain class of quadratic operator pencils with one-dimensional linear part. Ukraïn. Mat. Zh., 59(5):702–716, 2007.
  • [31] V. Pivovarchik. Symmetric Hermite-Biehler polynomials with defect. In Operator theory in inner product spaces, volume 175 of Oper. Theory Adv. Appl., pages 211–224. Birkhäuser, Basel, 2007.
  • [32] V. Pivovarchik and C. van der Mee. The inverse generalized Regge problem. Inverse Problems, 17(6):1831–1845, 2001.
  • [33] V. Pivovarchik and H. Woracek. Shifted Hermite-Biehler functions and their applications. Integral Equations Operator Theory, 57(1):101–126, 2007.
  • [34] V. Pivovarchik and H. Woracek. The square-transform of Hermite-Biehler functions. A geometric approach. Methods Funct. Anal. Topology, 13(2):187–200, 2007.
  • [35] E. J. Routh. A treatise on the stability of a given state of motion. London, 1877.
  • [36] T. Sheil-Small. Complex polynomials, volume 75 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2002.
  • [37] B. Simon. The classical moment problem as a self-adjoint finite difference operator. Adv. Math., 137(1):82–203, 1998.
  • [38] T. J. Stieltjes. Recherches sur les fractions continues. Ann. Fac. Sci. Toulouse Math. (6), 4(3):J76–J122, 1995. Reprint of Ann. Fac. Sci. Toulouse 8 (1894), J76–J122.
  • [39] T. J. Stieltjes. Recherches sur les fractions continues. Ann. Fac. Sci. Toulouse Math. (6), 4(4):A5–A47, 1995. Reprint of Ann. Fac. Sci. Toulouse 9 (1895), A5–A47.
  • [40] C. van der Mee and V. Pivovarchik. The Sturm-Liouville inverse spectral problem with boundary conditions depending on the spectral parameter. Opuscula Math., 25(2):243–260, 2005.
  • [41] H. S. Wall. Analytic Theory of Continued Fractions. D. Van Nostrand Company, Inc., New York, 1948.
  • [42] V. I. Yudovich. Eleven great problems of mathematical hydrodynamics. Mosc. Math. J., 3(2):711–737, 746, 2003. Dedicated to Vladimir I. Arnold on the occasion of his 65th birthday.