1 Introduction
1.1 Motivation
For a domain $\Omega $ in ${\mathbb C}^d$ ( $d\geq 1$ ), let ${\mathbb S}(\Omega )$ denote the set of analytic functions $f:\Omega \to \overline {{\mathbb D}}$ , where ${\mathbb D}$ denotes the open unit disk in ${\mathbb C}$ . Given a function $f\in {\mathbb S}(\Omega )$ , this paper revolves around the question when a given subset ${\mathcal D}$ of $\Omega $ has the property that whenever $g\in {\mathbb S}(\Omega )$ coincides with f on ${\mathcal D}$ , equals to f on whole $\Omega $ . When a subset has this property, we call it a determining set for $(f,\Omega )$ , or just f when the domain is clear from the context. For example, $\{0,1/2\}$ is a determining set for the identity map (by the Schwarz Lemma); any open subset of $\Omega $ is determining for any analytic function on $\Omega $ (by the Identity Theorem). See Rudin [Reference Rudin32, Chapter 5] for some interesting results related to a similar concept for $\Omega ={\mathbb D}^d$ .
The motivation behind the study of determining sets comes from the Pick interpolation problem. It corresponds to the case when ${\mathcal D}$ is a finite set. Given a finite subset ${\mathcal D}=\{\lambda _1,\lambda _2,\dots ,\lambda _N\}$ of $\Omega $ and points $w_1,w_2,\dots ,w_N$ in the open unit disk ${\mathbb D}$ , the Pick interpolation problem asks if there is an analytic function $f:\Omega \to {\mathbb D}$ such that $f(\lambda _j)=w_j$ for $j=1,2,\dots ,N$ . Therefore in this case, ${\mathcal D}$ being a determining set for $(f,\Omega )$ means that the (solvable) Pick problem $\lambda _j\mapsto f(\lambda _j)$ has a unique solution. In view of Pick’s pioneering work [Reference Pick31], it is therefore clear that when $\Omega ={\mathbb D}$ , then ${\mathcal D}$ is determining for f if and only if the Pick matrix
has rank less than N, which is further equivalent to the existence of a Blaschke function of degree less than N solving the data. The classical Pick interpolation problem has seen a wide range of generalizations. To mention a few, a necessary and sufficient condition for the solvability of a given Pick data is known when $\Omega $ is the polydisk ${\mathbb D}^d$ [Reference Agler2], the Euclidian ball ${\mathbb B}_d$ [Reference Kosiński and Zwonek24], the symmetrized bidisk [Reference Agler and Young10, Reference Bhattacharyya and Sau14], an affine variety [Reference Jury, Knese and McCullough20] and in more general setting of test functions [Reference Dritschel and McCullough18, Reference Dritschel, Marcantognini and McCullough17]. However, unlike the classical case, it is rather obscure in higher dimension when it comes to understanding when a given solvable Pick problem has a unique solution, and usually one has to settle with either necessary or sufficient conditions (see, for example, [Reference Agler and McCarthy4, Reference Scheinker33–Reference Scheinker35]).
1.2 The main results
The purpose of this article is to explore this direction where the domain under consideration is the symmetrized bidisk
Following the work [Reference Agler and Young7] of Agler and Young, this domain has remained a field of extensive research in operator theory and complex geometry constituting examples and counter-examples to celebrated problems in these areas such as the rational dilation problem [Reference Agler and Young8, Reference Bhattacharyya, Pal and Shyam Roy13] and the Lempert theorem [Reference Costara16]. In quest of understanding the determining sets, we shall actually consider the following more general situation.
Definition 1.1 Let $\Omega \subset {\mathbb C}^d$ be a domain, $E\subset \Omega $ and $f\in {\mathbb S}(\Omega )$ . We say that a subset ${\mathcal D}$ of E is determining for $(f, E)$ if for every $g\in {\mathbb S}(\Omega )$ , $g=f$ on ${\mathcal D}$ implies $g=f$ on E. If ${\mathcal D}$ is determining for $(f,E)$ for all $f\in {\mathbb S}(\Omega )$ , then we say that ${\mathcal D}$ is determining for E. Moreover, when E is the largest set in $\Omega $ such that ${\mathcal D}$ is determining for $(f,E)$ , we say that E is the uniqueness set for $(f,{\mathcal D})$ , i.e., in this case,
Here, for a function f, we use the standard notation $Z(f)$ for the zero set of f.
Note that if E is the uniqueness set for $(f,{\mathcal D})$ , then for every $z\in \Omega \setminus E$ , there exists a function $g\in {\mathbb S}(\Omega )$ such that $g=f$ on ${\mathcal D}$ but $f(z)\neq g(z)$ . Remarkably, when ${\mathcal D}$ is a finite subset of ${\mathbb G}$ , then for any function $f\in {\mathbb S}({\mathbb G})$ , the uniqueness set for $(f,{\mathcal D})$ is an affine variety (see [Reference Agler and McCarthy6, Reference Krishna Das, Kumar and Sau25]). This is owing to the fact that every solvable Pick data in ${\mathbb G}$ always has a rational inner solution (see [Reference Agler and McCarthy3, Reference Krishna Das, Kumar and Sau25]). Also note that if f and g agree on ${\mathcal D}$ , then ${\mathcal D}$ is determining for $(f,E)$ if and only if ${\mathcal D}$ is determining for $(g,E)$ also. In view of these facts, we shall mostly be concerned with the case when the function f in Definition 1.1 is rational and inner. Here, a function f in ${\mathbb S}({\mathbb G})$ is called inner, if $\lim _{r\to 1-}|f(r\zeta _1+r\zeta _2,r^2\zeta _1\zeta _2)|=1$ for almost all $\zeta _1,\zeta _2$ in ${\mathbb {T}}$ .
Note that ${\mathbb G}$ is the image of ${\mathbb D}^2$ under the (proper) holomorphic map $\pi :(z_1,z_2)\mapsto (z_1+z_2,z_1z_2)$ . The topological boundary of ${\mathbb G}$ is $\partial {\mathbb G}:=\pi (\overline {{\mathbb D}}\times {\mathbb {T}})\cup \pi ({\mathbb {T}}\ \times {\mathbb D})$ and the distinguished boundary of ${\mathbb G}$ is $b{\mathbb G}:=\pi ({\mathbb {T}}\times {\mathbb {T}})$ (see [Reference Agler and Young9]). Here, the distinguished boundary of a bounded domain $\Omega \subset {\mathbb C}^d$ is the $\check {\mbox S}$ ilov boundary with respect to the algebra of complex-valued functions continuous on $\overline {\Omega }$ and holomorphic in $\Omega $ . A special type of algebraic varieties has been prevalent in the study of uniqueness of the solutions of a Pick interpolation problem (see [Reference Agler and McCarthy6, Reference Kosiński22–Reference Krishna Das, Kumar and Sau25, Reference Maciaszek27]). We define it below. Throughout the paper, the notation $\xi $ stands for a polynomial in two variables.
Definition 1.2 An algebraic variety $Z(\xi )$ in ${\mathbb C}^2$ is said to be distinguished with respect to a bounded domain $\Omega $ , if
An example of a distinguished variety with respect to ${\mathbb G}$ is $\{(2z,z^2):z\in {\mathbb C}\}$ . We refer the readers to the papers [Reference Agler and McCarthy6, Reference Bhattacharyya, Kumar and Sau12, Reference Krishna Das, Kumar and Sau25, Reference Krishna Das and Sarkar26, Reference Pal and Shalit29] for results concerning these varieties and their connection to interpolation problems.
We now state the main results of this paper in the order they are proved.
-
(1) In Section 2.1, we reformulate the notion of determining set in the more general setting of reproducing kernel Hilbert spaces and find a sufficient condition for a finite subset of a general domain to be determining. This is Theorem 2.1. We also show by an example that the sufficient condition need not be necessary, in general.
-
(2) Starting with a natural number N, Section 2.2 constructs a finite subset of ${\mathbb G}$ consisting exactly of $N^2-N+1$ many points which is determining for any rational inner function with a natural degree constraint on it. This is Theorem 2.5. Proposition 2.4 is an intermediate step of the construction and is interesting on its own right.
-
(3) Given a distinguished variety ${\mathcal W}=Z(\xi )$ , we investigate in Section 2.3, when the intersection ${\mathcal W}\cap \mathbb {G}$ can be the uniqueness set for $(f,{\mathcal D})$ , where f is a rational inner function and ${\mathcal D}$ a finite subset of ${\mathbb G}$ (see Theorem 2.10). The preparatory results Propositions 2.7 and 2.8 are interesting in their own rights. Proposition 2.7 states that if f is a rational inner function with some regularity assumption, then there is a natural number N depending on f large enough so that any subset of ${\mathcal W}\cap {\mathbb G}$ consisting of N points is determining for $(f,{\mathcal W}\cap {\mathbb G})$ . This section then goes on to find (in Theorem 2.12) a sufficient condition for ${\mathcal W}\cap {\mathbb G}$ to be determining for a rational inner function f with a regularity assumption on it. The condition is just that the inequality
$$ \begin{align*} \quad 2\operatorname{Re}\langle f, \xi h\rangle_{H^2}< \|\xi h\|^2_2 \end{align*} $$holds, whenever h is a nonzero analytic function on ${\mathbb G}$ and $\xi h$ is bounded on ${\mathbb G}$ . Here, the inner product is the Hardy space inner product, briefly discussed in Section 2.3.
-
(4) Section 3 proves a bounded extension theorem for distinguished varieties with no singularities on $b{\mathbb G}$ . More precisely, given a distinguished variety ${\mathcal W}$ , we show that corresponding to every two-variable polynomial f, there is a rational function F on ${\mathbb G}$ such that $F|_{{\mathcal W}\cap {\mathbb G}}=f$ and that $\sup _{\mathbb G}|F(s,p)|\leq \alpha \sup _{{\mathcal W}\cap {\mathbb G}}|f|$ , for some constant $\alpha $ depending only on the distinguished variety ${\mathcal W}$ .
2 Determining and the uniqueness sets
2.1 A result for a general domain
We begin by proving a sufficient condition for a finite subset of a general domain to be determining. The concept of determining set can be formulated in a general setup of reproducing kernel Hilbert spaces. Here, a kernel on a domain $\Omega $ in ${\mathbb C}^d$ ( $d\geq 1$ ) is a function $k:\Omega \times \Omega \to {\mathbb C}$ such that for every choice of points $\lambda _1,\lambda _2,\dots ,\lambda _N$ in $\Omega $ , the $N\times N$ matrix $[k(\lambda _i,\lambda _j)]$ is positive-definite. Given a kernel k, there is a unique Hilbert space $H(k)$ associated with it, called the reproducing kernel Hilbert space; we refer the uninitiated reader to the book [Reference Paulsen and Raghupathi30]. For the purpose of this paper, all that is needed to know is that elements of the form $\{\sum _{j=1}^nc_jk(\cdot ,\lambda _j):c_j\in {\mathbb C}\mbox { and }\lambda _j\in \Omega \}$ constitute a dense set of $H(k)$ . A kernel k is said to be a holomorphic kernel, if it is holomorphic in the first and conjugate holomorphic in the second variable. Note that when k is holomorphic, then so are the elements of $H(k)$ . Let us denote by $\operatorname {Mult}H(k)$ the algebra of all bounded holomorphic functions $\varphi $ on $\Omega $ such that $\varphi \cdot f \in H(k)$ whenever $f\in H(k)$ . Such a holomorphic function is generally referred to as a multiplier for $H(k)$ . Let $\operatorname {Mult}_1H(k)$ denote the set of all multipliers $\varphi $ such that the operator norm of $M_\varphi : f\mapsto \varphi \cdot f$ for all f in $H(k)$ is no greater than one. A subset ${\mathcal D}\subset \Omega $ is said to be determining for a function $\varphi $ in $\operatorname {Mult}_1H(k)$ if whenever $\psi \in \operatorname {Mult}_1H(k)$ such that $\varphi =\psi $ on ${\mathcal D}$ , then $\varphi =\psi $ on $\Omega $ .
Theorem 2.1 Let k be a holomorphic kernel on a domain $\Omega $ in ${\mathbb C}^d$ , $\varphi \in \operatorname {Mult}_1H(k),$ and ${\mathcal D}=\{\lambda _1,\lambda _2,\dots ,\lambda _N\}\subset \Omega $ . If the matrix
is singular, then ${\mathcal D}$ is determining for $\varphi $ .
Proof Since the matrix (2.1) is singular, there is a nonzero vector in its kernel; let us denote it by $\gamma $ . Let $\lambda _{N+1}$ be any point in $\Omega \setminus {\mathcal D}$ , and let $\psi \in \operatorname {Mult}_1H(k)$ be any function such that $\varphi =\psi $ on ${\mathcal D}$ . Since $\psi \in \operatorname {Mult}_1H(k)$ , the operator $M_\psi :f\mapsto \psi \cdot f$ is a contractive operator on $H(k)$ and therefore for every $z\in {\mathbb C}$ ,
Since $\gamma \in \operatorname {Ker}[(1-\varphi (\lambda _i)\overline {\varphi (\lambda _j)})k(\lambda _i,\lambda _j)]$ and $\varphi =\psi $ on ${\mathcal D}$ , the above inequality collapses to
Since the above inequality is true for all $z\in \mathbb {C}$ , we have
which, after a rearrangement of terms, gives
Define for z in $\Omega $ ,
By definition, it is clear that $L\in H(k)$ . Consider the open set ${\mathcal O}=\Omega \setminus Z(L)$ . Note that if $\lambda _{N+1}\in {\mathcal O}$ , then the right-hand side of (2.2) does not vanish, and therefore $\psi (\lambda _{N+1})$ is uniquely determined.
Now suppose $\phi =\psi $ on ${\mathcal O}$ . By the assumption that ${\mathcal O}$ is a set of uniqueness for $\operatorname {Mult}_1(H(k))$ , it follows that $\phi =\psi $ .
The converse of the above result is not true as the simple example below demonstrates.
Example 2.2 Let k be the Bergman kernel on $\Omega =\mathbb {D}$ , i.e., $k(z, w)= (1-z\overline {w})^{-2}$ . Then it is well known that $\operatorname {Mult}_1H(k)={\mathbb S}({\mathbb D})$ (see, for example, [Reference Agler and McCarthy5, Section 2.3]. By the Schwarz lemma, ${\mathcal D}=\{0,1/2\}$ is determining for the identity function. However, the matrix $\left [\begin {smallmatrix} 1 & 1 \\ 1 & 4/3 \end {smallmatrix}\right ]$ is nonsingular.
The rest of the paper specializes to the symmetrized bidisk.
2.2 Finite sets as a determining set
Given a natural number N, this subsection constructs a finite subset ${\mathcal D}$ of ${\mathbb G}$ consisting exactly of $N^2-N+1$ many points, which is determining for any rational inner function on ${\mathbb G}$ with a degree constraint on it. This is inspired by the work of Scheinker [Reference Scheinker34], which extends the following classical result for the unit disk to the polydisks.
Lemma 2.3 (Pick [Reference Pick31])
Let ${\mathcal D}=\{\lambda _1, \lambda _2,\dots ,\lambda _N\}\subset {\mathbb D}$ , and let f be a rational inner function on $\mathbb {D}$ with degree strictly less than N. Then if $g\in {\mathbb S}(\mathbb {D})$ is such that $f=g$ on ${\mathcal D}$ , then $f=g$ on ${\mathbb D}$ .
For $\epsilon>0$ and $z\in {\mathbb C}$ , let $D(z;\epsilon ):=\{w\in {\mathbb {D}}: |z-w|<{\epsilon }\}$ . For $\zeta \in {\mathbb {T}}$ and $a\in {\mathbb D}$ , let $m_{\zeta ,a}$ be the Möbius map
We shall have use of two notions of degree for a polynomial in two variables. The one used in this subsection is the following. For a polynomial $\xi (z,w)=\sum _{i,j}a_{i,j}z^iw^j$ , we define $\deg \xi :=\max (i+j)$ such that $a_{i,j}\neq 0$ . The degree of a rational function in its reduced fractional representation is defined to be the degree of the numerator polynomial. The following is an intermediate step to proving Theorem 2.5.
Proposition 2.4 Let N be a positive integer and for each $j=1,2,\dots ,N$ , let $\beta _j$ be distinct points in ${\mathbb {T}}$ , and let $D_j$ be the analytic disks $D_j=\{(z+\beta _jz, \beta _j z^2):z\in {\mathbb D}\}$ . Then:
-
(a) There exist $\beta \in {\mathbb {T}}$ and $\epsilon>0$ such that for every fixed $\zeta \in D(\beta ;\epsilon )\cap \mathbb {T}$ and $a\in {D(0;\epsilon )}$ , the analytic disk
$$ \begin{align*}{\mathcal D}_{\zeta,a}=\{(z+m_{\zeta,a}(z), zm_{\zeta, a}(z)):z\in{\mathbb D}\} \end{align*} $$intersects each of the analytic disks $D_j$ at a nonzero point.
-
(b) For each $\zeta \in {\mathbb {T}}$ and $\epsilon>0$ , the set
$$ \begin{align*}{\mathcal D}_\zeta=\{(z+m_{\zeta,a}(z), zm_{\zeta, a}(z)):z\in{\mathbb D} \text{ and } a\in{D(0;\epsilon)}\} \end{align*} $$is a determining set for any function in ${\mathbb S}(\mathbb {G})$ .
-
(c) The set
$$ \begin{align*}E=\{(z+\beta_jz, \beta_j z^2): z\in{\mathbb D} \text{ and } j=1,2,\dots,N\}=\cup_{j=1}^ND_j\end{align*} $$is a determining set for any rational inner function of degree less than N.
Proof For part (a), note that given a $\zeta \in {\mathbb {T}}$ and $a\in {\mathbb D}$ , the analytic disk ${\mathcal D}_{\zeta ,a}$ intersects each $D_j$ at a nonzero point if and only if there exist $0\neq z\in {\mathbb D}$ such that for each j, $\beta _jz=m_{\zeta ,a}(z)$ , which is equivalent to having $\overline {a}\beta _jz^2+(\beta _j-\zeta )z-a\zeta =0$ . Therefore, $\zeta $ must belong to ${\mathbb {T}}\setminus \{\beta _j: j=1,2,\dots ,N\}$ . Now fix one such $\zeta $ and j. Let $\lambda _1(a),\lambda _2(a)$ be the roots of the polynomial above. Then clearly $\lambda _1(0)=0=\lambda _2(0)$ . Therefore by continuity of the roots, there exists $\epsilon>0$ such that whenever $a\in D(0;\epsilon )$ , $\lambda _1(a)$ and $\lambda _2(a)$ belong to ${\mathbb D}$ . This $\epsilon $ will of course depend on j but since there are only finitely many j, we can find an $\epsilon>0$ so that (a) holds.
For part (b), we have to show that if $f:\mathbb {G}\to \overline {\mathbb D}$ is any analytic function such that $f|_{{\mathcal D}_\zeta }=0$ , then $f=0$ on ${\mathbb G}$ . Fix $z\in \mathbb {D}$ and consider $f_z:\mathbb D\to \overline {\mathbb D}$ defined by $f_z:w\mapsto f(z+w,zw)$ . Since f vanishes on $\mathcal D_\zeta $ , $f_z$ vanishes on $\{m_{\zeta ,a}(z):a\in D(0;\epsilon )\}$ which shows that $f_z=0$ on ${\mathbb D}$ . Since $z\in {\mathbb D}$ is arbitrary, $f=0$ on ${\mathbb G}$ .
For part (c), let f be a rational inner function of degree less than N and $g\in {{\mathbb S}(\mathbb {G})}$ be such that $g=f$ on each $D_j$ . For each $\zeta $ and a as in part (a), ${\mathcal D}_{\zeta ,a}$ intersects each $D_j$ at say $(s_j,p_j)=(\lambda _j+m_{\zeta ,a}(\lambda _j),\lambda _jm_{\zeta ,a}(\lambda _j))$ . Restrict f and g to ${\mathcal D}_{\zeta ,a}$ to get $f_{\zeta ,a}(z)=f(z+m_{\zeta ,a}(z), zm_{\zeta , a}(z))$ and $g_{\zeta ,a}(z)=g(z+m_{\zeta ,a}(z), zm_{\zeta , a}(z))$ . Then clearly $f_{\zeta ,a}$ is a rational inner function on ${\mathbb D}$ of degree less than N and $g_{\zeta ,a}\in {\mathbb S}({\mathbb D})$ . Then for each $j=1,2,\dots , N$ , $g_{\zeta ,a}(\lambda _j)=f_{\zeta ,a}(\lambda _j)$ . Therefore by Lemma (2.3), we have $g_{\zeta ,a}=f_{\zeta ,a}$ on $\mathbb {D}$ for each $\zeta $ and a as in part (a). Hence $g=f$ on ${\mathcal D}$ , which by part (b) gives $g=f$ on ${\mathbb G}$ . This completes the proof.
Theorem 2.5 For any $N\geq 1$ , there exists a set D consisting of $(N^2-N+1)$ points in $\mathbb {G}$ such that ${\mathcal D}$ is a determining set for any rational inner function of degree less than N.
Proof For $N=1$ , it is trivial because then a rational inner function of degree less than $1$ is identically constant. So suppose $N>1$ . Let $\lambda _1:=0, \lambda _2,\ldots ,\lambda _N$ be distinct points in $\mathbb {D}$ , $\beta _1,\ldots , \beta _{N}$ be distinct points in $\mathbb {T}$ and $D_1,\ldots ,D_{N}$ be the analytic disks as in Proposition 2.4. Consider the set
Since $\beta _j$ and $\lambda _j$ are distinct, ${\mathcal D}$ consists of precisely $N^2-N+1$ many points. Let f be a rational inner function on ${\mathbb G}$ and $g\in {\mathbb S}({\mathbb G})$ be such that g agrees with f on ${\mathcal D}$ . As before, restrict f and g to each $D_k$ to obtain rational inner functions $f_{k}(z)=f(z+\beta _kz,z^2\beta _k)$ and $g_k(z)=g(z+\beta _kz,z^2\beta _k)$ on the unit disk ${\mathbb D}$ . We then have $f_k(\lambda _j)=g_k(\lambda _j)$ for each $j=1,2,\dots ,N$ . Thus by Lemma 2.3, $f_k(z)=g_k(z)$ on ${\mathbb D}$ for each $k=1,2,\dots ,N$ , which is same as saying that $f=g$ on $\cup _{k=1}^ND_k$ . Consequently, by part (c) of Proposition 2.4, $f=g$ on ${\mathbb G}$ .
2.3 Distinguished varieties as a determining and the uniqueness set
A rational function $f=g/h$ with relatively prime polynomials g and h, is called regular if $h\neq {0}$ on $\overline {\mathbb {G}}$ . For example, note that while the rational function $(3p-s)/(3-s)$ is regular, $(2p-s)/(2-s)$ is not.
We first recall the known results that will be used later. Let ${\mathcal W}=Z(\xi )$ be a distinguished variety with respect to $\mathbb {G}$ . Then it follows easily that ${\mathcal V}=Z(\xi \circ \pi )$ defines a distinguished variety with respect to ${\mathbb D}^2$ . Lemma 1.2 of [Reference Agler and McCarthy6] produces a regular Borel measure $\nu $ on $\partial {\mathcal V}:={\mathcal V}\cap {\mathbb {T}}^2$ such that $\nu $ gives rise to a Hardy-type Hilbert function space on ${\mathcal V}\cap {\mathbb D}^2$ , denoted by $H^2(\nu )$ , i.e., $H^2(\nu )$ is the closure in $L^2(\nu )$ of polynomials such that evaluation at every point in ${\mathcal V}\cap {\mathbb D}^2$ is a bounded linear functional on $H^2(\nu )$ . It was then shown in [Reference Pal and Shalit29, Lemma 3.2] that the push-forward measure $\mu (E)=\nu (\pi ^{-1}(E))$ for every Borel subset E of $\partial {\mathcal W}:={\mathcal W}\cap b\Gamma $ has all the properties that $\nu $ has. Furthermore, the spaces $H^2(\mu )$ and $H^2(\nu )$ are unitary equivalent via the isomorphism given by
Note that if $k^\mu $ and $k^\nu $ are the Szegö-type reproducing kernels for $H^2(\mu )$ and $H^2(\nu )$ , respectively, then for every $(z,w)\in {\mathcal V}\cap {\mathbb D}^2$ and $f\in H^2(\mu ),$
We observe the following.
Lemma 2.6 Let ${\mathcal W}$ be a distinguished variety with respect to $\mathbb G$ , and let $\mu $ be the regular Borel measure on $\partial {\mathcal W}$ as in the preceding discussion. Then for every regular rational inner function f on $\mathbb {G}$ , the multiplication operator $M_f$ on $H^2(\mu )$ has a finite dimensional kernel.
Proof We note that for every $(z,w)\in {\mathcal V}\cap {\mathbb D}^2$ ,
Thus, $M_f$ on $H^2(\mu )$ and $M_{f\circ \pi }$ on $H^2(\nu )$ are unitarily equivalent via the unitary U as in 2.3. Now the lemma follows from [Reference Scheinker33, Theorem 3.6], which states that $\operatorname {Ker}M_{f\circ \pi }$ is finite-dimensional.
Proposition 2.7 Let $\mathcal W=Z(\xi )$ be a distinguished variety with respect to $\mathbb G$ , and let f be a regular rational inner function on $\mathbb G$ . If $\dim \operatorname { Ker} M_{f}^* < N$ , then any N distinct points in ${\mathcal W}\cap {\mathbb G}$ is a determining set for $(f, {\mathcal W}\cap \mathbb {G})$ .
Proof Let $\{w_1,w_2,\dots ,w_N\}$ be distinct points in ${\mathcal W}\cap {\mathbb G}$ , and let $g\in {\mathbb S}({\mathbb G})$ be such that $g(w_j)=f(w_j)$ for each $j=1,2,\dots ,N$ . Let ${\mathcal V}=Z(\xi \circ \pi )$ and $\{v_1,v_2,\dots ,v_N\}$ be in ${\mathcal V}\cap {\mathbb D}^2$ such that $\pi (v_j)=w_j$ for all $j=1,2,\ldots ,N$ . Thus, $g\circ \pi (v_j)=f\circ \pi (v_j)$ for each $j=1,2,\dots ,N$ . Theorem 1.7 of [Reference Scheinker33] yields $g\circ \pi =f\circ \pi $ on ${\mathcal V}\cap {\mathbb D}^2$ which is same as $g=f$ on ${\mathcal W}\cap {\mathbb G}$ . This completes the proof.
The $2$ -degree of a two-variable polynomial $\xi \in \mathbb {C}[z,w]$ is defined as $(d_1, d_2)=:\operatorname {2-deg}\xi $ , where $d_1$ and $d_2$ are the largest power of z and w, respectively, in the expansion of $\xi (z,w)$ . The reflection of a two-variable polynomial $\xi \in \mathbb {C}[z,w]$ is defined as
For a rational function $f(z,w)=\xi (z,w)/\eta (z,w)$ with $\xi $ and $\eta $ having no common factor, the $2$ -degree of f is defined to be the $2$ -degree of the numerator. For two pairs of nonnegative integers $(p,q)$ and $(m,n)$ , we write $(p,q)\leq (m,n)$ to indicate that $p\leq m$ and $q\leq n$ .
Proposition 2.8 Let ${\mathcal W}=Z(\xi )$ be an irreducible distinguished variety, and let f be a regular rational inner function on $\mathbb {G}$ of the form
If $\operatorname {2-deg}\xi \circ \pi \leq \operatorname {2-deg}f\circ \pi $ , then for each $(s,p)\in \mathbb {G}\setminus (\mathbb {G}\cap {\mathcal W})$ , there exists a regular rational inner function g on ${\mathbb G}$ such that g coincides with f on ${\mathcal W}\cap {\mathbb G}$ but $g(s,p)\neq f(s,p)$ .
Proof Let $\operatorname {2-deg}\eta \circ \pi =(l,l)$ and $\operatorname {2-deg}\xi \circ \pi =(n,n)$ . The hypothesis then is that $m+l-n$ is nonnegative. For $\epsilon>0$ , define a symmetric function $g_\epsilon $ on ${\mathbb D}^2$ as
Simple computation shows that the reflection of the denominator of $g_\epsilon $ is equal to the numerator of $g_\epsilon $ , which implies that each each $g_\epsilon $ is a rational inner function on ${\mathbb D}^2$ provided that the denominator does not vanish on ${\mathbb D}^2$ . Since $\eta \circ \pi $ does not vanish on $\overline {{\mathbb D}}^2$ , we can always find a sufficiently small $\epsilon $ so that the denominator of each $g_\epsilon $ does not vanish in $\overline {{\mathbb D}}^2$ , thus making $g_\epsilon $ regular.
By Proposition 4.3 of [Reference Knese21], $\xi \circ \pi =c \widetilde {\xi \circ \pi }$ for some $c\in {\mathbb {T}}$ . This ensures that each $g_\epsilon $ coincides with f on ${\mathcal W}\cap {\mathbb G}$ . Now, let $(z_0,w_0)\in {\mathbb D}^2$ be such that $\pi (z_0,w_0)\in {\mathbb G}\setminus {\mathcal W}$ . Then $g_\epsilon (z_0,w_0)=f\circ \pi (z_0,w_0)$ if and only if
which, after cross-multiplication and using the fact that $\xi \circ \pi (z_0,w_0)\neq 0$ , leads to
Since $\eta \circ \pi $ does not vanish on $\overline {{\mathbb D}}^2$ , we have $z_0w_0\neq 0$ . Therefore, the above equation holds if and only if
If $m+l-n=0$ , then f is a constant function. The hypothesis on the 2-degrees of $\xi $ and f then implies that $\xi $ must be constant. This is not possible because $\xi $ defines a distinguished variety. Therefore, $m+l-n\geq 1$ , in which case, equation (2.7) implies that $|f\circ \pi (z_0,w_0)|>1$ . This again is a contradiction because f is a rational inner function and so by the Maximum Modulus Principle, $|f\circ \pi (z)|\leq 1$ for every $(z,w)\in {\mathbb D}^2$ . Consequently, $g_\epsilon (s,p)\neq f(s,p)$ for every $(s,p)\in {\mathbb G}\setminus ({\mathcal W}\cap {\mathbb G})$ .
Remark 2.9 In a forthcoming paper [Reference Bhowmik and Kumar15], it is shown that any rational inner function on ${\mathbb G}$ is of the form (2.4) possibly multiplied by a unimodular constant.
Theorem 2.10 Let ${\mathcal W}=Z(\xi )$ be an irreducible distinguished variety with respect to $\mathbb {G}$ , let f be a regular rational inner function on ${\mathbb G}$ of the form (2.4) such that $\operatorname {2-deg}\xi \circ \pi \leq \operatorname {2-deg}f\circ \pi $ , and let ${\mathcal D}$ be any subset of ${\mathcal W}\cap \mathbb {G}$ consisting of at least $1+\dim \operatorname {Ker}M_f^*$ many points. Then ${\mathcal W}\cap \mathbb {G}$ is the uniqueness set for $(f,{\mathcal D})$ .
Proof Consider the multiplication operator $M_{f}$ on $H^2(\mu )$ , where $H^2(\mu )$ is the Hilbert space corresponding to ${\mathcal W}$ as mentioned in Lemma 2.6. By this lemma, $\dim \operatorname {Ker}(M_{f}^*)$ is finite. So let N be such that $\dim \operatorname {Ker}(M_{f}^*)<N$ and ${\mathcal D}=\{\lambda _1, \lambda _2,\ldots , \lambda _N\}\subset {\mathcal W}\cap \mathbb {G}$ . By Proposition 2.7, ${\mathcal D}$ is determining for $(f,{\mathcal W}\cap \mathbb {G})$ . We use Proposition 2.8 to show that ${\mathcal W}\cap \mathbb {G}$ is the uniqueness set. Toward that end, pick $(s,p)\in {\mathbb G}\setminus {\mathcal W}\cap {\mathbb G}$ . Proposition 2.8 guarantees the existence of a (regular) rational inner function g that coincides with f on ${\mathcal W}\cap {\mathbb G}$ but $g(s,p)\neq f(s,p)$ . This proves that ${\mathcal W}\cap \mathbb {G}$ is the uniqueness set for the interpolation problem. This completes the proof of the theorem.
Remark 2.11 An extremal interpolation problem in ${\mathbb G}$ is a solvable problem with no solution of supremum norm less than $1$ . Let ${\mathcal D}=\{\lambda _1,\lambda _2,\dots ,\lambda _N\}$ be a subset of ${\mathbb G}$ , and letf be a rational inner function on ${\mathbb G}$ such that the N-point Pick problem $\lambda _j\mapsto f(\lambda _j)$ is extremal and that none of the $(N-1)$ -point subproblems is extremel. Then it is shown in [Reference Krishna Das, Kumar and Sau25] that the uniqueness set for $(f,{\mathcal D})$ contains a distinguished variety. Theorem 2.10 can be seen as a converse to this result. Indeed, Theorem 2.10 starts with a distinguished variety ${\mathcal W}=Z(\xi )$ and produces a regular rational inner function f and a finite set ${\mathcal D}$ depending on ${\mathcal W}$ such that ${\mathcal W}\cap {\mathbb G}$ is the uniqueness set for $(f,{\mathcal D})$ . In addition, we note that the problem $\lambda _j\mapsto f(\lambda _j)$ is an extremal problem. This is because if g is any solution of the problem, then by Proposition 2.7, $g=f$ on ${\mathcal W}\cap {\mathbb G}$ . Thus,
The last equality follows because f is a regular rational inner function.
There is a sufficient condition for a distinguished variety to be determining. In the theorem below and in its proof, the inner product $\langle ,\rangle _{H^2}$ for analytic functions $f,g:{\mathbb G}\to {\mathbb C}$ is defined to be
where m is the standard normalized Lebesgue measure on ${\mathbb {T}}\times {\mathbb {T}}$ , and $J(z,w)=z-w$ is the Jacobian of the map $\pi :(z,w)\mapsto (z+w,zw)$ . See the papers [Reference Bhattacharyya, Das and Sau11, Reference Bhattacharyya and Sau14, Reference Misra, Shyam Roy and Zhang28] for some motivation for and operator theory on the spaces of analytic functions for which $\|f\|_2:=\sqrt {\langle f,f\rangle }_{H^2}<\infty $ . Note here that if f is an inner function on ${\mathbb G}$ , then $\|f\|_2=1.$
Theorem 2.12 Let $\mathcal {W}=Z(\xi )$ be a distinguished variety such that $\xi =\xi _1.\xi _2\dots \xi _l$ , where $\xi _i$ are irreducible polynomials with $\xi _i$ and $\xi _j$ are co-prime for each $i\neq j$ , and let f be a regular rational inner function on $\mathbb {G}$ . If for each analytic function $h(\not \equiv 0)$ on $\mathbb {G}$ ,
holds, whenever $\xi h$ is bounded on ${\mathbb G}$ , then ${\mathcal W}\cap {\mathbb G}$ is a determining set for f.
Proof We shall use contrapositive argument. So suppose that there exists $g\in {{\mathbb S}(\mathbb {G})}$ such that g coincides with f on ${\mathcal W}\cap {\mathbb G}$ but $g\neq f$ . Choose an integer N so that $\dim \operatorname {Ker}M_f^*<N$ and pick N distinct points $\lambda _1,\ldots .,\lambda _N\in {\mathcal W}$ . Consider the N-point (solvable) Nevanlinna–Pick problem $\lambda _j\mapsto f(\lambda _j)$ . By Proposition 2.7, all the solutions to this problem agree on ${\mathcal W}\cap {\mathbb G}$ . Since $g\neq f$ , there exists a $\lambda _{N+1}\in \mathbb {G}\setminus {\mathcal W}$ such that $g(\lambda _{N+1})\neq f(\lambda _{N+1})$ . Now consider the $(N+1)$ -point Nevanlinna–Pick problem $\lambda _j\mapsto g(\lambda _j)$ on $\mathbb {G}$ . By [Reference Krishna Das, Kumar and Sau25, Theorem 5.3], every solvable Nevanlinna–Pick problem in ${\mathbb G}$ has a rational inner solution. Let $\psi $ be a rational inner solution to the $(N+1)$ -point problem $\lambda _j\mapsto g(\lambda _j)$ . Since $\psi $ , in particular, solves the problem $\lambda _j\mapsto f(\lambda _j)$ for each $j=1,2,\dots ,N$ , $\psi =f$ on ${\mathcal W}\cap {\mathbb G}$ . But since $\psi (\lambda _{N+1})=g(\lambda _{N+1})\neq f(\lambda _{N+1})$ , $\psi $ is distinct from f. Since $\psi =f$ on ${\mathcal W}\cap {\mathbb G}$ , by the Study Lemma, there exists a rational function h such that $f-\psi =\xi h$ (see [Reference Fischer19, Chapter 1]. Since $\psi $ is inner,
Since f is an inner function, $\|f\|_{2}=1$ , and therefore, the above computation leads to $2\operatorname {Re}\langle f, \xi h\rangle = \|\xi h\|^2_2$ . This contradicts the hypothesis because $\xi h=f-\psi $ is bounded. Consequently, g must coincide with f on ${\mathbb G}$ .
One can easily find examples of distinguished varieties and regular rational inner functions such that the stringent hypothesis of the above result is satisfied.
Example 2.13 Let $f\circ \pi (z,w)=(zw)^d$ and ${\mathcal W}=Z(\xi )$ be such that
where $m,n$ are mutually prime integers bigger than d. Then it follows that ${\mathcal W}$ is a distinguished variety with respect to ${\mathbb G}$ because $Z(z^m-w^n)$ is a distinguished variety with respect to ${\mathbb D}^2$ . For concrete example, one can take $d=1$ and $(m,n)=(2,3)$ – the corresponding distinguished variety then is the Neil parabole. Note that the inner product $\langle ,\rangle $ as defined in (2.8) can be expressed in terms of the inner product on the Hardy space of the bidisk $H^2({\mathbb D}^2)$ as
Let $h:{\mathbb G}\to {\mathbb C}$ be an analytic function such that $\|\xi h\|_2<\infty $ . Since $\{z^iw^j: i,j\geq 0\}$ forms an orthonormal basis for $H^2({\mathbb D}^2)$ , it is easy to read off from (2.9) that $\langle f,\xi h\rangle =0$ . Therefore, by Theorem 2.12, ${\mathcal W}\cap {\mathbb G}$ is a determining set for f as chosen above.
3 A bounded extension theorem
We end with a bounded extension theorem for distinguished varieties with no singularities on the distinguished boundary of $\Gamma $ . Here, singularity of an algebraic variety $Z(\xi )$ at a point means that both the partial derivatives of $\xi $ vanish at that point. Note that the substance of the following theorem is not that there is a rational extension of every polynomial, it is that the supremum of the rational extension over ${\mathbb G}$ does not exceed the supremum of the polynomial over the variety intersected with ${\mathbb G}$ multiplied by a constant that only depends on the variety. See the papers [Reference Adachi, Andersson and Cho1, Reference Knese21, Reference Stout36] for similar results in other contexts.
Theorem 3.1 Let ${\mathcal W}$ be a distinguished variety with respect to $\mathbb {G}$ such that it has no singularities on $b{\Gamma }$ . Then, for every polynomial $f\in \mathbb {C}[s,p]$ , there exists a rational extension F of f such that
for all $(s,p)\in \mathbb {G}$ , where $\alpha $ is a constant depends only on ${\mathcal W}$ .
Proof Let ${\mathcal V}$ be a distinguished variety with respect to $\mathbb {D}^2$ such that ${\mathcal W}=\pi ({\mathcal V})$ . Since ${\mathcal W}$ has no singularities on $b\Gamma $ , it follows that ${\mathcal V}$ has no singularities on $\mathbb {T}^2$ . Invoke Theorem 2.20 of [Reference Knese21] to obtain a rational extension G of the polynomial $f\circ \pi \in \mathbb {C}[z,w]$ such that
for all $(z,w)\in \mathbb {D}^2$ , where $\alpha $ is a constant depends only on ${\mathcal V}$ . Now, define a rational function H on $\mathbb {D}^2$ as follows:
Clearly, H is also a rational extension of $f\circ \pi $ with
Note that H is a symmetric rational function on $\mathbb {D}^2$ . So, there is a rational function F on $\mathbb {G}$ such that
Now, we will show that this F will do our job. It is easy to see that F is a rational extension of f. Let $(s, p)\in \mathbb {G}$ . Then there exists a point $(z,w)\in \mathbb {D}^2$ such that $(s, p)=(z+w, zw)$ . Now,
This complete the proof.
Acknowledgment
P.K. thanks his supervisor Professor Tirthankar Bhattacharyya for some fruitful discussions. The authors thank the anonymous referee for some valuable suggestions.