1 Orthogonal polynomials on the real line and tridiagonal operators

Let \(\mu \) be a probability measure on the real line. We denote the support of \(\mu \) by \(\mathcal {S}\) and assume its cardinality \(\#\mathcal {S}=\infty \). Let \((p_n)_{n\in {\mathbb {N}}_0}\) denote the unique orthonormal polynomial sequence with respect to \(\mu \), that is \(\deg p_n =n\), \(\int p_n p_m d\mu =\delta _{n,m}\), and \(p_n\) has a positive leading coefficient for all \(n,m\in {\mathbb {N}}_0\). The orthonormal polynomial sequence \((p_n)_{n\in {\mathbb {N}}_0}\) satisfies a recurrence relation

$$\begin{aligned} xp_n(x)=\lambda _n p_{n+1}(x)+\beta _np_n(x)+\lambda _{n-1} p_{n-1}(x) \end{aligned}$$
(1)

with \(p_{-1}(x)=0\), \(p_0(x)=1\), \(\lambda _{-1}=0\), \(\lambda _n>0\) and \(\beta _n \in {\mathbb {R}}\) for all \(n \in {\mathbb {N}}_0\).

Conversely, if \((p_n)_{n\in {\mathbb {N}}_0}\) is defined by (1), there is a probability measure \(\mu \) such that \((p_n)_{n\in {\mathbb {N}}_0}\) is the orthonormal polynomial sequence with respect to \(\mu \), see e.g. [2].

In the case \((\lambda _n)_{n\in {\mathbb {N}}_0}\) and \((\beta _n)_{n\in {\mathbb {N}}_0}\) are bounded \(\mathcal {S}\) is compact and vice versa. The boundedness implies also that the orthogonalization measure \(\mu \) is uniquely determined. The smallest interval containing \(\mathcal {S}\) is called the true interval of orthogonality, see e.g. [2].

Now, let \(x_0\in {\mathbb {R}}{\setminus } \mathcal {N}\), where \(\mathcal {N}=\{x\in {\mathbb {C}}: \exists n\in {\mathbb {N}} \text{ with } p_n(x)=0\}\) is the set of zeros of all orthonormal polynomials. It is well known that \(\mathcal {N}\subset {\mathbb {R}}\), see e.g. [2]. The normalized polynomials

$$\begin{aligned} P_n(x)={p_n(x)\over p_n(x_0)} \end{aligned}$$
(2)

form an orthogonal polynomial sequence \((P_n)_{n\in {\mathbb {N}}_0}\) with respect to \(\mu \), that is

$$\begin{aligned} \int P_n P_m d\mu ={\delta _{n,m} \over h_n} \end{aligned}$$
(3)

with \(h_n>0\). We call \(x_0\) a normalizing point. The corresponding three-term recurrence relation is

$$\begin{aligned} xP_n(x)=\gamma _n P_{n+1}(x)+\beta _nP_n(x)+\alpha _n P_{n-1}(x) \end{aligned}$$
(4)

with \(P_{-1}(x)=0\), \(P_0(x)=1\),

$$\begin{aligned} \gamma _n= & {} {p_{n+1}(x_0) \over p_n(x_0)} \lambda _n,\end{aligned}$$
(5)
$$\begin{aligned} \alpha _{n}= & {} {p_{n-1}(x_0) \over p_n(x_0)} \lambda _{n-1}, \quad \text{ and }\end{aligned}$$
(6)
$$\begin{aligned} \alpha _n+\beta _n+\gamma _n= & {} x_0 \quad \text {for all}\quad n \in {\mathbb {N}}_0. \end{aligned}$$
(7)

Note that (6) implies \(\alpha _0=0\). It is also important to emphasize that (7) applies if and only if \(x_0\) is a normalization point and that our investigations heavily depend on Eq. (7).

Moreover, \(\gamma _n\alpha _{n+1}=\lambda _n^2>0\). One easily shows

$$\begin{aligned} h_{n+1} \alpha _{n+1}=h_n\gamma _n \end{aligned}$$
(8)

which implies

$$\begin{aligned} h_n={ \gamma _0 \ldots \gamma _{n-1} \over \alpha _1 \ldots \alpha _n}=p_n^2(x_0)\quad \text{ for } \text{ all } n \in {\mathbb {N}}_0. \end{aligned}$$
(9)

Note that (9) also applies in the case \(n=0\), where the nominator and denominator are empty products that means they are set equal 1 by default. Therefore, (3) as well as (9) yields \(h_0=1\).

The so called Christoffel–Darboux formula is given by

$$\begin{aligned} \sum _{k=0}^n P_k(x)P_k(y)h_k=\gamma _n h_n {P_{n+1}(x)P_n(y)-P_n(x)P_{n+1}(y) \over x-y}, \end{aligned}$$
(10)

see [2]. Hence,

$$\begin{aligned} \sum _{k=0}^n P_k(x)^2 h_k=\gamma _n h_n (P_{n+1}'(x)P_n(x)-P_n'(x)P_{n+1}(x)), \end{aligned}$$
(11)

and in particular setting \(x=x_0\) we get

$$\begin{aligned} P_{n+1}'(x_0)-P_n'(x_0) = {H_n \over \gamma _nh_n} \end{aligned}$$
(12)

with

$$\begin{aligned} H_n=\sum _{k=0}^n h_k\quad \text{ for } \text{ all }\quad n \in {\mathbb {N}}_0. \end{aligned}$$
(13)

Definition 1.1

If \(\{P_{n+1}'(x_0)-P_n'(x_0): n\in {\mathbb {N}}_0\}\) is bounded, then we call \(x_0\) a normalizing point with bounded growth of derivatives.

Note that further on speaking about \(x_0\) as a normalizing point of bounded growth of derivatives is the same as to speak about the boundedness of \(\{{H_n \over \gamma _nh_n}: n\in {\mathbb {N}}_0\}\).

Subsequently we deal with the case \(\mathcal {S}=\text{ supp }\mu \) is compact which is equivalent with \((\gamma _n\alpha _{n+1})_{n\in {\mathbb {N}}_0}\) and \((\beta _n)_{n\in {\mathbb {N}}_0}\) are bounded sequences. Then the true interval of orthogonality is \([\min \mathcal {S},\max \mathcal {S}]\).

Lemma 1.1

In the case \(x_0\ge \max \mathcal {S}\) we have \(\alpha _{n+1}, \gamma _n >0\) for all \(n\in {\mathbb {N}}_0\) and in the case \(x_0\le \min \mathcal {S}\) we have \(\alpha _{n+1}, \gamma _n <0\) for all \(n\in {\mathbb {N}}_0\).

Proof

Since \(\mathcal {N}\subset (\min \mathcal {S}, \max \mathcal {S})\) and the leading coefficient of all orthonormal polynomials is positive we have in the case \(x_0\ge \max \mathcal {S}\) that \(p_n(x_0)>0\) for all \(n\in {\mathbb {N}}_0\). Whereas in the case \(x_0\le \min \mathcal {S}\) the sign of \(p_n(x_0)\) is alternating. \(\square \)

On the set of complex-valued sequences there acts a linear operator \(T: {\mathbb {C}}^{{\mathbb {N}}_0}\rightarrow {\mathbb {C}}^{{\mathbb {N}}_0}\) determined by the recurrence relation (4). More precisely, for \(\xi \in {\mathbb {C}}^{{\mathbb {N}}_0}\) put

$$\begin{aligned} (T\xi )_n=T\xi _n=\gamma _n\xi _{n+1}+\beta _n\xi _n+\alpha _n\xi _{n-1} \quad \text{ for } \text{ all }\quad n\in {\mathbb {N}}_0, \end{aligned}$$
(14)

where \(\xi _{-1}=0\). Written as tridiagonal matrix the operator T has the form

$$\begin{aligned} T=\left( \begin{array}{ccccccc} \beta _0&{}\gamma _0&{}0&{}0&{}0&{}\cdots &{}\\ \alpha _1&{}\beta _1&{}\gamma _1&{}0&{}0&{}\cdots &{}\\ 0&{}\alpha _2&{}\beta _2&{}\gamma _2&{}0&{}\cdots &{}\\ 0&{}0&{}\alpha _3&{}\beta _3&{}\gamma _3&{}\cdots &{}\\ \vdots &{}\vdots &{}\vdots &{}\ddots &{}\ddots &{}\ddots \end{array}\right) . \end{aligned}$$
(15)

Note that in our investigations T acts on the different spaces \({\mathbb {C}}^{{\mathbb {N}}_0}\), \(l^1(h)\) and \(l^2(h)\) which is clear from the respective context. First let us study T as an operator on

$$\begin{aligned} l^1(h)=\left\{ \xi \in {\mathbb {C}}^{{\mathbb {N}}_0}:\sum _{n=0}^\infty \mid \xi _n \mid h_n<\infty \right\} \end{aligned}$$
(16)

with norm \(\Vert \xi \Vert _1=\sum _{n=0}^\infty \mid \xi _n\mid h_n\) for all \(\xi \in l^1(h)\).

Proposition 1.1

In the case \(\mid \alpha _n\mid \), \(\mid \beta _n\mid \) and \(\mid \gamma _n\mid \le B\) for all \(n\in {\mathbb {N}}_0\) the operator T: \(l^1(h)\rightarrow l^1(h)\) is well defined and continuous. Especially we have

$$\begin{aligned} \sum _{n=0}^\infty T\xi _nh_n=x_0\sum _{n=0}^\infty \xi _nh_n \quad \text{ and }\quad \Vert T\xi \Vert _1 \le C\Vert \xi \Vert _1 \end{aligned}$$
(17)

for all \(\xi \in l^1(h)\), where \(C=\min (3B,\mid x_0\mid +2B)\).

Proof

Set \(\gamma _{-1}=\xi _{-1}=h_{-1}=0\).

Applying (8) and the assumed absolute convergence of the series we obtain

$$\begin{aligned} \sum _{n=0}^\infty T\xi _nh_n&=\sum _{n=0}^\infty (\gamma _n\xi _{n+1}+\beta _n\xi _n+\alpha _n\xi _{n-1})h_n\\&=\sum _{n=0}^\infty \alpha _{n+1}\xi _{n+1}h_{n+1}+\beta _n\xi _nh_n+\gamma _{n-1}\xi _{n-1}h_{n-1}\\&=\sum _{n=0}^\infty (\alpha _{n}+\beta _n+\gamma _{n})\xi _{n}h_{n} =x_0\sum _{n=0}^\infty \xi _nh_n.\\ \sum _{n=0}^\infty \mid T\xi _n\mid h_n&\le \sum _{n=0}^\infty (\mid \gamma _n\mid \mid \xi _{n+1}\mid +\mid \beta _n\mid \mid \xi _n\mid +\mid \alpha _n\mid \mid \xi _{n-1}\mid )h_n\\&=\sum _{n=0}^\infty \mid \alpha _{n+1}\mid \mid \xi _{n+1}\mid h_{n+1}+\mid \beta _n\mid \mid \xi _n\mid h_n\\&\quad +\mid \gamma _{n-1}\mid \mid \xi _{n-1}\mid h_{n-1}\\&=\sum _{n=0}^\infty (\mid \alpha _{n}\mid +\mid \beta _n\mid +\mid \gamma _{n}\mid )\mid \xi _{n}\mid h_{n} \le 3B \sum _{n=0}^\infty \mid \xi _{n}\mid h_{n}. \end{aligned}$$

At least two of the coefficients in \(\mid \alpha _{n}\mid +\mid \beta _n\mid +\mid \gamma _{n}\mid \) do have the same sign. For instance, if \(\text {sign}\alpha _n=\text {sign}\beta _n\), then \(\mid \alpha _{n}\mid +\mid \beta _n\mid +\mid \gamma _{n}\mid =\mid \alpha _{n}+\beta _n+\gamma _n-\gamma _n\mid +\mid \gamma _{n}\mid \le \mid x_0\mid +2\mid \gamma _n\mid \le \mid x_0\mid +2B\). Proceeding the same way with all the other possibilities one gets alternatively \(\sum _{n=0}^\infty \mid T\xi _n\mid h_n \le (\mid x_0\mid +2B)\sum _{n=0}^\infty \mid \xi _{n}\mid h_{n}\), which completes the proof. \(\square \)

We focus on the weighted Hilbert space

$$\begin{aligned} l^2(h)=\left\{ \xi \in {\mathbb {C}}^{{\mathbb {N}}_0}:\sum _{n=0}^\infty \mid \xi _n\mid ^2h_n<\infty \right\} \end{aligned}$$
(18)

with scalar product \(\langle \xi ,\upsilon \rangle =\sum _{n=0}^\infty \xi _n\overline{\upsilon _n}h_n\) and norm \(\Vert \xi \Vert _2=\sqrt{\langle \xi ,\xi \rangle }\) for all \(\xi ,\upsilon \in l^2(h)\).

Proposition 1.2

In the case \(\mid \alpha _n\mid \), \(\mid \beta _n\mid \) and \(\mid \gamma _n\mid \le B\) for all \(n\in {\mathbb {N}}_0\) the operator T: \(l^2(h)\rightarrow l^2(h)\) is a well defined, self-adjoint and continuous operator with

$$\begin{aligned} \Vert T\xi \Vert _2 \le C\Vert \xi \Vert _2, \end{aligned}$$
(19)

where \(C=\min (3B,\mid x_0\mid +2B)\).

Proof

Set \(\gamma _{-1}=\xi _{-1}=h_{-1}=\upsilon _{-1}=0\).

Now let \(\xi \in l^2(h)\). Since

$$\begin{aligned} \mid T\xi _n\mid&\le \sqrt{\mid \gamma _n\mid }\sqrt{\mid \gamma _n\mid }\mid \xi _{n+1}\mid +\sqrt{\mid \beta _n\mid }\sqrt{\mid \beta _n\mid }\mid \xi _n\mid \\&\quad +\sqrt{\mid \alpha _n\mid }\sqrt{\mid \alpha _n\mid }\mid \xi _{n-1}\mid \end{aligned}$$

the Cauchy–Schwarz inequality implies

$$\begin{aligned} \mid T\xi _n\mid ^2 \le (\mid \gamma _{n}\mid +\mid \beta _n\mid +\mid \alpha _{n}\mid )(\mid \gamma _{n}\mid \mid \xi _{n+1}\mid ^2+\mid \beta _n\mid \mid \xi _n\mid ^2+\mid \alpha _{n}\mid \mid \xi _{n-1}\mid ^2). \end{aligned}$$

Therefore, proceeding like in the proof of Proposition 1.1

$$\begin{aligned} \sum _{n=0}^\infty \mid T\xi _n\mid ^2 h_n&\le C \sum _{n=0}^\infty (\mid \gamma _n\mid \mid \xi _{n+1}\mid ^2+\mid \beta _n\mid \mid \xi _n\mid ^2+\mid \alpha _n\mid \mid \xi _{n-1}\mid ^2)h_n\\&= C\sum _{n=0}^\infty \mid \alpha _{n+1}\mid \mid \xi _{n+1}\mid ^2 h_{n+1}+\mid \beta _n\mid \mid \xi _n\mid ^2 h_n\\&\quad +\mid \gamma _{n-1}\mid \mid \xi _{n-1}\mid ^2 h_{n-1}\\&= C\sum _{n=0}^\infty \; (\mid \alpha _{n}\mid +\mid \beta _n\mid +\mid \gamma _{n}\mid )\mid \xi _{n}\mid ^2 h_{n} \le C^2 \sum _{n=0}^\infty \mid \xi _{n}\mid ^2 h_{n}, \end{aligned}$$

which implies \( \Vert T\xi \Vert _2 \le C \Vert \xi \Vert _2, \) where \(C=\min (3B,\mid x_0\mid +2B)\).

For arbitrary \(\xi ,\upsilon \in l^2(h)\) one gets due to the absolute convergence

$$\begin{aligned} \langle T\xi ,\upsilon \rangle= & {} \sum _{n=0}^\infty (\gamma _n\xi _{n+1}+\beta _n\xi _n+\alpha _n \xi _{n-1})\overline{\upsilon _n}h_n\\= & {} \sum _{n=0}^\infty \xi _n(\gamma _{n-1}\overline{\upsilon _{n-1}}h_{n-1}+\beta _n\overline{\upsilon _n}h_n+\alpha _{n+1}\overline{\upsilon _{n+1}}h_{n+1})\\= & {} \sum _{n=0}^\infty \xi _n(\alpha _n\overline{\upsilon _{n-1}}h_n+\beta _n\overline{\upsilon _n}h_n+\gamma _n\overline{\upsilon _{n+1}}h_n) = \langle \xi ,T\upsilon \rangle . \end{aligned}$$

\(\square \)

Corollary 1.1

In the case \(\mid \alpha _n\mid \), \(\mid \beta _n\mid \) and \(\mid \gamma _n\mid \le B\) for all \(n\in {\mathbb {N}}_0\) the spectrum \(\sigma (T)\) is a subset of \([-C,C]\), where \(C=\min (3B,\mid x_0\mid +2B)\).

The numerical range of T is the set

$$\begin{aligned} W(T)=\left\{ \langle T\xi ,\xi \rangle :\xi \in l^2(h),\Vert \xi \Vert _2=1\right\} . \end{aligned}$$
(20)

Since T is self-adjoint we have

$$\begin{aligned} \{ m(T),M(T)\}\subseteq \sigma (T) \subseteq \,\text{ co }(\sigma (T))\subseteq \overline{W(T)}=[m(T),M(T)], \end{aligned}$$
(21)

where \(\text{ co }(\sigma (T))\) is the convex hull of \(\sigma (T)\), \(m(T)=\inf W(T)\) and \(M(T)=\sup W(T)\), see [5, Intro]. Moreover, \(\Vert T\Vert = \max (\mid m(T)\mid ,\mid M(T)\mid )\).

Proposition 1.3

In the case \(\mid \alpha _n\mid \), \(\mid \beta _n\mid \) and \(\mid \gamma _n\mid \le B\) for all \(n\in {\mathbb {N}}_0\) one gets

$$\begin{aligned} \langle (x_0\text{ id } -T)\xi ,\xi \rangle =\sum _{n=0}^\infty \gamma _n\mid \xi _n-\xi _{n+1}\mid ^2h_n \quad \text{ for } \text{ all }\quad \xi \in l^2(h). \end{aligned}$$
(22)

Proof

Set \(\gamma _{-1}=\xi _{-1}=h_{-1}=0\). Using (8) and the absolute convergence of the series one gets for an arbitrary \(\xi \in l^2(h)\) that

$$\begin{aligned} \sum _{n=0}^\infty (x_0 \xi _n-T\xi _n)\overline{\xi _n}h_n&=\sum _{n=0}^\infty ((\alpha _n+\beta _n+\gamma _n)\xi _n-\alpha _n\xi _{n-1}\\&\quad -\beta _n\xi _n-\gamma _n\xi _{n+1})\overline{\xi _n}h_n\\&=\sum _{n=0}^\infty (\gamma _n\xi _n\overline{\xi _n} -\gamma _n\xi _{n+1}\overline{\xi _n})h_n\\&\quad +(\gamma _{n-1}\xi _n\overline{\xi _n}-\gamma _{n-1}\xi _{n-1}\overline{\xi _n})h_{n-1}\\&=\sum _{n=0}^\infty \gamma _n(\xi _n\overline{\xi _n} -\xi _{n+1}\overline{\xi _n}+\xi _{n+1}\overline{\xi _{n+1}}-\xi _{n}\overline{\xi _{n+1}})h_{n}\\&=\sum _{n=0}^\infty \gamma _n\mid \xi _n-\xi _{n+1}\mid ^2h_n. \end{aligned}$$

\(\square \)

Lemma 1.2

The following statements apply.

  1. (i)

    \(\sum _{n=0}^\infty \gamma _n\mid \xi _n-\xi _{n+1}\mid ^2h_n \ge 0\) for all \(\xi \in l^2(h)\) with \(\Vert \xi \Vert _2=1\) if and only if \(\gamma _n>0\) for all \(n\in {\mathbb {N}}_0\).

  2. (ii)

    \(\sum _{n=0}^\infty \gamma _n\mid \xi _n-\xi _{n+1}\mid ^2h_n \le 0\) for all \(\xi \in l^2(h)\) with \(\Vert \xi \Vert _2=1\) if and only if \(\gamma _n<0\) for all \(n\in {\mathbb {N}}_0\).

Proof

If \(\gamma _n>0\) for all \(n\in {\mathbb {N}}_0\) then \(\sum _{n=0}^\infty \gamma _n\mid \xi _n-\xi _{n+1}\mid ^2h_n \ge 0\) for all \(\xi \in l^2(h)\). In the case we have not \(\gamma _n>0\) for all \(n\in {\mathbb {N}}_0\) there is an index \(m\in {\mathbb {N}}_0\) such that \(\gamma _m<0\) and \(\gamma _n>0\) for all \(n\in \{0,\ldots ,m-1\}\). Define \(\zeta \in l^2(h)\) by \(\zeta _n = (\sum _{k=0}^m h_k)^{-1/2}\) for all \(n\in \{0,\ldots ,m\}\) and \(\zeta _n = 0\) for all \(n\in \{m+1,m+2\ldots \}\). Then \(\Vert \zeta \Vert _2=1\) and \(\sum _{n=0}^\infty \gamma _n\mid \zeta _n-\zeta _{n+1}\mid ^2h_n = \gamma _m \mid \zeta _m\mid ^2 h_m <0\).

The second statement is shown quite analogue. \(\square \)

Corollary 1.2

If \(\mid \alpha _n\mid \), \(\mid \beta _n\mid \), \(\mid \gamma _n\mid \le B\) for all \(n\in {\mathbb {N}}_0\) and \(C=\min (3B, \mid x_0\mid +2B)\), then the following statements apply.

  1. (i)

    If \(\gamma _n>0\) for all \(n\in {\mathbb {N}}_0\), then \(\overline{W(T)}\subseteq [-C,x_0]\). In particular, \(\sigma (T)\subseteq [-C,x_0]\).

  2. (ii)

    If \(\gamma _n<0\) for all \(n\in {\mathbb {N}}_0\), then \(\overline{W(T)}\subseteq [x_0,C]\). In particular, \(\sigma (T)\subseteq [x_0,C]\).

  3. (iii)

    If there exist \(k,l\in {\mathbb {N}}_0\) with \(\gamma _k\gamma _l<0\), then \(x_0 \in (\min \mathcal {S}, \max \mathcal {S})\).

Note that in the following \(L^2({\mathbb {R}},\mu )\) is as usual a set of equivalence classes and a function used in this context represents an equivalence class. This is also expressed by using the formulation ’for \(\mu \)-almost all \(x\in {\mathbb {R}}\)’.

Define \(\epsilon ^{(k)} \in l^2(h)\) by

$$\begin{aligned} \epsilon ^{(k)}_n=\frac{\delta _{k,n}}{h_k}\quad \text{ for } \text{ all }\quad n,k\in {\mathbb {N}}_0. \end{aligned}$$
(23)

Then obviously

$$\begin{aligned} \Vert \epsilon ^{(k)}\Vert ^2_2=\frac{1}{h_k}=\int P_k ^2d\mu \quad \text{ for } \text{ all }\quad k\in {\mathbb {N}}_0. \end{aligned}$$
(24)

Extending the map \(\epsilon ^{(k)} \mapsto P_k\) linearly to the linear span of \(\{\epsilon ^{(k)}:k\in {\mathbb {N}}_0\}\) and finally to the closure of the linear span we get the so-called Plancherel isomorphism

$$\begin{aligned} \mathcal {P}: l^2(h)\rightarrow L^2({\mathbb {R}},\mu ), \end{aligned}$$

which is an isometric isomorphism from \(l^2(h)\) onto \(L^2({\mathbb {R}},\mu )\). It is completely determined by

$$\begin{aligned} \mathcal {P}(\epsilon ^{(k)})=P_k\quad \text{ for } \text{ all }\quad k\in {\mathbb {N}}_0. \end{aligned}$$

Note that

$$\begin{aligned} T\epsilon ^{(k)}=\alpha _k\epsilon ^{(k-1)}+\beta _k\epsilon ^{(k)}+\gamma _k\epsilon ^{(k+1)}\quad \text{ for } \text{ all }\quad k\in {\mathbb {N}}_0, \end{aligned}$$
(25)

where \(\epsilon ^{(-1)}_n=0\) for all \(n\in {\mathbb {N}}_0\). Now we define an operator M on \(L^2({\mathbb {R}},\mu )\) by

$$\begin{aligned} M(f)=\mathcal {P}\circ T\circ \mathcal {P}^{-1}(f)\quad \text{ for } \text{ all }\quad f \in L^2({\mathbb {R}},\mu ), \end{aligned}$$
(26)

where \(\mathcal {P}^{-1}\) denotes the inverse operator of \(\mathcal {P}\). Then \(M\in B(L^2({\mathbb {R}},\mu ))\) with \(\Vert M\Vert \,\le \min (3B,\mid x_0\mid +2B)\). Taking into account the three-term recurrence relation (4) we deduce that

$$\begin{aligned} M(P_k)(x)=\mathcal {P}(T\epsilon ^{(k)})(x)=\mathcal {P}(\alpha _k\epsilon ^{(k-1)}+\beta _k\epsilon ^{(k)}+\gamma _k\epsilon ^{(k+1)})(x)=xP_k(x) \end{aligned}$$
(27)

for \(\mu \)-almost all \(x\in {\mathbb {R}}\) and for all \(k\in {\mathbb {N}}_0\). If g is a function in the linear span of \(\{P_k: k\in {\mathbb {N}}_0\}\), then the linearity of M yields

$$\begin{aligned} M(g)(x)=x g(x) \quad \text{ for }\quad \mu \text {-almost all}\; x\in {\mathbb {R}}. \end{aligned}$$
(28)

Since M is bounded and the closure of the linear span of \(\{P_k: k\in {\mathbb {N}}_0\}\) is \(L^2({\mathbb {R}},\mu )\) we get by standard arguments that

$$\begin{aligned} M(f)(x)=xf(x)\ \text{ for }\ \mu \text {-almost all}\; x \in {\mathbb {R}}\;\text{ and } \text{ for } \text{ all }\; f\in L^2({\mathbb {R}},\mu ).\qquad \end{aligned}$$
(29)

By [4, Definition 2.61 and Corollary 4.24] the spectrum \(\sigma (M)\) is exactly the essential range

$$\begin{aligned} \mathcal {R}=\left\{ \lambda \in {\mathbb {R}}: \mu (\{x \in {\mathbb {R}}:\mid x-\lambda \mid <\epsilon \})>0 \text{ for } \text{ all } \epsilon >0\right\} , \end{aligned}$$
(30)

Obviously \(\mathcal {R}=\text{ supp }\mu \) and \(\sigma (M)=\sigma (T)\). Hence, we can add to Corollary 1.2 the following result.

Corollary 1.3

For orthogonal polynomials \((P_n)_{n\in {\mathbb {N}}_0}\) which are defined by (4) with \(\mid \alpha _n\mid \), \(\mid \beta _n\mid \) and \(\mid \gamma _n\mid \le B\) for all \(n\in {\mathbb {N}}_0\) we have

$$\begin{aligned} \mathcal {S}=\text{ supp }\mu =\sigma (T). \end{aligned}$$
(31)

2 A characterization of \(x_0 \notin \mathcal {S}\)

In the whole section we assume that \(\mid \alpha _n\mid \), \(\mid \beta _n\mid \) and \(\mid \gamma _n\mid \le B\) for all \(n\in {\mathbb {N}}_0\).

The main result of this paper will be a necessary and sufficient condition for \(x_0 \in \mathcal {S}\). Moreover, in the case of \(x_0 \notin \mathcal {S}\) we will present an explicit form of the inverse \((x_0\text{ id }-T)^{-1}\), which is based on a weighted Cesàro operator \(C\in B(l^2(h))\).

Define \(C\eta =((C\eta )_n)_{n\in {\mathbb {N}}_0}=(C\eta _n)_{n\in {\mathbb {N}}_0}\) by

$$\begin{aligned} C\eta _n=\frac{1}{H_n}\sum _{k=0}^n \eta _kh_k \quad \text{ for } \text{ all }\quad \eta \in l^2(h). \end{aligned}$$
(32)

Then C is a bounded linear operator on \(l^2(h)\) with \(\Vert C\Vert \le 2\), see [3, Theorem A]. It is straightforward to show that the adjoint operator \(C^* \in B(l^2(h))\) is defined by

$$\begin{aligned} C^*\eta _n=\sum _{k=n}^\infty \eta _k\frac{h_k}{H_k}\quad \text{ for } \text{ all }\quad \eta \in l^2(h). \end{aligned}$$
(33)

Theorem 2.1

If \(x_0 \notin \mathcal {S}= \sigma (T)\), then \(x_0\) is a normalizing point with bounded growth of derivatives.

Proof

Given \({n \in {\mathbb {N}}}_0\) denote by \(\chi ^{(n)}\) the sequence with \(\chi ^{(n)}_k=1\) for \(k\in \{0,\ldots ,n\}\) and \(\chi ^{(n)}_k=0\) for \(k\in \{n+1,n+2,\ldots \}\). An easy computation shows that

$$\begin{aligned} (x_0\text{ id }-T)\chi ^{(n)}_k= & {} 0\quad \text{ for } \text{ all }\quad k \in {\mathbb {N}}_0 \backslash \{n,n+1\},\\ (x_0\text{ id }-T)\chi ^{(n)}_n= & {} \gamma _n,\quad \text{ and }\\ (x_0\text{ id }-T)\chi ^{(n)}_{n+1}= & {} -\alpha _{n+1}. \end{aligned}$$

Hence,

$$\begin{aligned} \Vert (x_0\text{ id }-T)\chi ^{(n)}\Vert _2^2&=\gamma _n^2h_n+\alpha _{n+1}^2h_{n+1}=\gamma _n(\gamma _n+\alpha _{n+1})h_n\\&=\mid \gamma _n\mid \mid \gamma _n+\alpha _{n+1}\mid h_n. \end{aligned}$$

Since \(x_0\notin \sigma (T)\), there exists \(A = (x_0\text{ id }-T)^{-1}\in B(l^2(h))\). Then

$$\begin{aligned} \Vert A \circ (x_0\text{ id }-T)\chi ^{(n)}\Vert _2^2&= \Vert \chi ^{(n)}\Vert _2^2=\sum _{k=0}^nh_k=H_n,\quad \text{ and }\\ \Vert A \circ (x_0\text{ id }-T)\chi ^{(n)}\Vert _2^2&\le \Vert A\Vert ^2\Vert (x_0\text{ id }-T)\chi ^{(n)})\Vert _2^2\\&=\Vert A\Vert ^2\mid \gamma _n\mid h_n\mid \gamma _n+\alpha _{n+1}\mid \\&\le 2B\Vert A\Vert ^2\mid \gamma _n\mid h_n, \end{aligned}$$

which implies

$$\begin{aligned} H_n\le 2B\Vert A\Vert ^2\mid \gamma _n\mid h_n\quad \text{ for } \text{ all }\quad n\in {\mathbb {N}}_0. \end{aligned}$$

Therefore, \(\left\{ \frac{H_n}{\gamma _nh_n}: n\in {\mathbb {N}}_0\right\} \) is bounded. \(\square \)

In order to prove the converse implication we start with determining a sequence \(\omega =(\omega _n)_{n\in {\mathbb {N}}_0}\) such that \((x_0\text{ id }-T)(\omega )=\epsilon ^{(0)}\). Note that in the following lemma the operator T acts on \({\mathbb {C}}^{{\mathbb {N}}_0}\).

Lemma 2.1

A sequence \(\omega =(\omega _n)_{n\in {\mathbb {N}}_0} \in {\mathbb {C}}^{{\mathbb {N}}_0}\) satisfies \((x_0 \text{ id }-T)(\omega )=\epsilon ^{(0)}\) if and only if

$$\begin{aligned} \omega _{n+1}=\omega _0-\sum _{k=0}^n \frac{1}{\gamma _kh_k} \quad \text{ for } \text{ all }\quad n\in {\mathbb {N}}_0. \end{aligned}$$
(34)

Proof

We have \(((x_0 \text{ id }-T)\omega )_0=1/h_0=1\) if and only if \(\omega _0-\omega _1=\frac{1}{\gamma _0}\). For \(n\ge 1\) we see that \(((x_0\text{ id }-T)\omega )_n=\omega _n-(\gamma _n\omega _{n+1}+\beta _n\omega _n+\alpha _n\omega _{n-1})=0\) if and only if \(\gamma _n(\omega _{n+1}-\omega _n)=\alpha _n(\omega _n-\omega _{n-1})\). Now, by iteration we get

$$\begin{aligned} \omega _{n+1}-\omega _n=\frac{\alpha _n}{\gamma _n}(\omega _n-\omega _{n-1})=\frac{\alpha _n\alpha _{n-1}\cdots \alpha _1}{\gamma _n\gamma _{n-1}\cdots \gamma _1}\frac{-1}{\gamma _0}=\frac{-1}{\gamma _nh_n}. \end{aligned}$$

\(\square \)

Next we investigate under which assumptions a sequence \(\omega =(\omega _n)_{n\in {\mathbb {N}}_0}\) of Lemma 2.1 is a member of \(l^2(h)\).

Lemma 2.2

If \(x_0\) is a normalizing point with bounded growth of derivatives, then

$$\begin{aligned} \sum _{k=0}^\infty \frac{1}{\mid \gamma _k\mid h_k}<\infty , \end{aligned}$$
(35)

and consequently the series \( \sum _{k=0}^\infty \frac{1}{\gamma _kh_k} \) is convergent.

Proof

Due to the assumption there exists a \(D>0\) with

$$\begin{aligned} \sum _{k=0}^n\left( \frac{1}{\gamma _kh_k}\right) ^2h_k\le D \sum _{k=0}^n\left( \frac{1}{H_k}\right) ^2h_k \quad \text{ for } \text{ all }\quad n \in {\mathbb {N}}_0. \end{aligned}$$

Since \(C\epsilon ^{(0)}=\left( \frac{1}{H_n}\right) _{n\in {\mathbb {N}}_0} \in l^2(h)\), we have \(\left( \frac{1}{\gamma _nh_n}\right) _{n\in {\mathbb {N}}_0}\in l^2(h)\), that is \(\sum _{k=0}^\infty \frac{1}{\gamma _k^2h_k}<\infty .\) Finally, \(\gamma _k^2 \le B\mid \gamma _k\mid \) yields \(\sum _{k=0}^\infty \frac{1}{\mid \gamma _k\mid h_k}<\infty ,\) which implies the series \(\sum _{k=0}^\infty \frac{1}{\gamma _k h_k}\) is convergent. \(\square \)

Now with respect to Lemma 2.1, if the series \(\sum _{k=0}^\infty \frac{1}{\gamma _kh_k}\) is convergent, then the sequence \(\omega =(\omega _n)_{n\in {\mathbb {N}}_0}\) is defined by

$$\begin{aligned} \omega _n=\sum _{k=n}^\infty \frac{1}{\gamma _kh_k}\quad \text{ for } \text{ all }\quad {n \in {\mathbb {N}}}_0. \end{aligned}$$
(36)

In order to prove that \(\omega \in l^2(h)\) whenever \(\left\{ \frac{H_n}{\gamma _nh_n}:n \in {\mathbb {N}}_0\right\} \) is bounded, we use the adjoint weighted Cesàro operator \(C^*\in B(l^2(h))\). Define a sequence \(\eta =(\eta _n)_{n\in {\mathbb {N}}_0}\) by

$$\begin{aligned} \eta _n=\frac{H_n}{\gamma _nh_n^2}\quad \text{ for } \text{ all }\quad n \in {\mathbb {N}}_0. \end{aligned}$$
(37)

Lemma 2.3

If \(x_0\) is a normalizing point with bounded growth of derivatives, then \(\eta \in l^2(h)\).

Proof

We have \(\mid \eta _n\mid \le D \frac{1}{h_n}\) for all \(n \in {\mathbb {N}}_0\). Hence, we have to show that \(\left( \frac{1}{h_n}\right) _{n\in {\mathbb {N}}_0} \in l^2(h)\). According to Lemma 2.2 it follows

$$\begin{aligned} \sum _{k=0}^\infty \frac{1}{h_k^2} h_k\le B \sum _{k=0}^\infty \frac{1}{\mid \gamma _k\mid h_k}<\infty . \end{aligned}$$

\(\square \)

Since \(C^*\eta =\omega ,\) Lemmas 2.3 and 2.1 yield the following proposition.

Proposition 2.1

If \(x_0\) is a normalizing point with bounded growth of derivatives, then \(\omega \in l^2(h)\) (defined by (36)) satisfies \((x_0\text{ id }-T)\omega =\epsilon ^{(0)}\).

Assuming that \(x_0\) is a normalizing point with bounded growth of derivatives our next goal is to find sequences \(\omega ^{(m)}\in l^2(h)\) with \((x_0\text{ id }-T)\omega ^{(m)}=\epsilon ^{(m)}\) for all \(m \in {\mathbb {N}}.\)

To that end, we introduce a sequence of operators \(S_m \in B(l^2(h))\) by setting

$$\begin{aligned} S_{m+1}=\frac{1}{\gamma _m}\left( T\circ S_m-\beta _mS_m-\alpha _mS_{m-1}\right) \quad \text{ for } \text{ all }\quad m \in {\mathbb {N}}_0, \end{aligned}$$
(38)

where \(S_{-1} = 0\) and \(S_0 = \text{ id }\).

Proposition 2.2

The following two statements apply.

  1. (i)
    $$\begin{aligned} S_m\epsilon ^{(0)}=\epsilon ^{(m)}\quad \text{ for } \text{ all }\quad m \in {\mathbb {N}}_0. \end{aligned}$$
    (39)
  2. (ii)
    $$\begin{aligned} (S_m\omega )_k=\left\{ \begin{array}{lll} \omega _m &{}\quad \textrm{if}&{}\quad k=0,\ldots ,m, \\ \omega _k &{}\quad \textrm{if} &{}\quad k=m+1,m+2,\ldots \end{array}\right. \quad \text{ for } \text{ all }\; m \in {\mathbb {N}}_0. \end{aligned}$$
    (40)

Proof

In any case the proof is done by induction.

(i): By trivial means we have \(S_0\epsilon ^{(0)}=\epsilon ^{(0)}\). Since \(S_1=\frac{1}{\gamma _0}\left( T-\beta _0\text{ id }\right) \) we have \((S_1\epsilon ^{(0)})_0= \frac{1}{\gamma _0}(\beta _0-\beta _0)=0\), \((S_1\epsilon ^{(0)})_1= \frac{\alpha _1}{\gamma _0}\frac{1}{h_0}=\frac{1}{h_1}\) and \((S_1\epsilon ^{(0)})_k=0\) for all \(k\ge 2\). Therefore, \(S_1\epsilon ^{(0)}=\epsilon ^{(1)}\).

Assume that \(S_m\epsilon ^{(0)}=\epsilon ^{(m)}\) and \(S_{m-1}\epsilon ^{(0)}=\epsilon ^{(m-1)}\) for \(m\in {\mathbb {N}}_0\) is already shown. Then

$$\begin{aligned} S_{m+1}\epsilon ^{(0)}= & {} \frac{1}{\gamma _m}\left( T\circ S_m-\beta _mS_m-\alpha _mS_{m-1}\right) \epsilon ^{(0)}\\= & {} \frac{1}{\gamma _m}\left( T\epsilon ^{(m)}-\beta _m\epsilon ^{(m)}-\alpha _m\epsilon ^{(m-1)}\right) \\= & {} \frac{1}{\gamma _m}\left( \alpha _m\epsilon ^{(m-1)} +\beta _m\epsilon ^{(m)}+\gamma _m\epsilon ^{(m+1)} -\beta _m\epsilon ^{(m)}-\alpha _m\epsilon ^{(m-1)}\right) \\= & {} \frac{1}{\gamma _m}\left( \gamma _m\epsilon ^{(m+1)}\right) =\epsilon ^{(m+1)}. \end{aligned}$$

(ii): By trivial means we have \((S_0\omega )_k=\omega _k\) for all \(k \in {\mathbb {N}}_0\). Moreover, since \(S_1={1\over \gamma _0}(T-\beta _0id)\), \(T\omega =x_0id-\epsilon ^{(0)}\) and \(x_0=\beta _0+\gamma _0\) we get

$$\begin{aligned} S_1\omega = \frac{1}{\gamma _0}T\omega -\frac{\beta _0}{\gamma _0}\omega =\frac{1}{\gamma _0}((\beta _0 + \gamma _0)\omega -\epsilon ^{(0)})-\frac{\beta _0}{\gamma _0}\omega = \omega -\frac{\epsilon ^{(0)}}{\gamma _0}. \end{aligned}$$

Hence, \((S_1\omega )_0=\omega _1\), \((S_1\omega )_1=\omega _1\), and \((S_1\omega )_k=\omega _k\) for all \(k\ge 2\).

Assume again that the statement is already shown for \(m \in {\mathbb {N}}\) and \(m-1\). Then for \(k=0,\ldots ,m-1\) we have

$$\begin{aligned} (S_{m+1}\omega )_k= & {} \frac{1}{\gamma _m}((T\circ S_m \omega )_k-\beta _m\omega _m-\alpha _m \omega _{m-1})\\= & {} \frac{1}{\gamma _m}(x_0\omega _m-\beta _m\omega _m-\alpha _m \omega _{m-1}) =\omega _m+\frac{\alpha _m}{\gamma _m}(\omega _m-\omega _{m-1})\\= & {} \omega _m-\frac{\alpha _m}{\gamma _m}\frac{1}{\gamma _{m-1}h_{m-1}} =\omega _m-\frac{1}{\gamma _m h_m}=\omega _{m+1}. \end{aligned}$$

For \(k=m\) it follows

$$\begin{aligned} (S_{m+1}\omega )_k=\frac{1}{\gamma _m}(\alpha _m\omega _m+\beta _m\omega _m+\gamma _m\omega _{m+1}-\beta _m\omega _m-\alpha _m \omega _m)=\omega _{m+1}. \end{aligned}$$

Finally, for \(k=m+1,m+2,\ldots \) we have

$$\begin{aligned} (S_{m+1}\omega )_k&=\frac{1}{\gamma _m}\left( \alpha _k\omega _{k-1}+\beta _k \omega _k+\gamma _k \omega _{k+1}-\beta _m\omega _k-\alpha _m\omega _k\right) \\&=\frac{1}{\gamma _m}\left( \gamma _k\left( \omega _k-\frac{1}{\gamma _kh_k}\right) +\beta _k\omega _k\right. \nonumber \\&\left. \quad +\alpha _k \left( \omega _k+\frac{1}{\gamma _{k-1}h_{k-1}}\right) -\beta _m\omega _k-\alpha _m\omega _k\right) \\&=\frac{1}{\gamma _m}\left( x_0 \omega _k-\frac{1}{h_k}+\frac{1}{h_k}-\beta _m\omega _k-\alpha _m\omega _k\right) =\omega _k. \end{aligned}$$

\(\square \)

Now our goal is met by setting \(\omega ^{(m)} = S_m\omega \) for all \(m \in {\mathbb {N}}_0\).

Proposition 2.3

If \(x_0\) is a normalizing point with bounded growth of derivatives, then \(S_m\omega \in l^2(h)\) satisfies \((x_0\text{ id }-T)S_m\omega =\epsilon ^{(m)}\) for all \(m \in {\mathbb {N}}_0.\)

Proof

Obviously \(S_m\) commutes with \(x_0\text{ id }-T\). Hence,

$$\begin{aligned} (x_0\text{ id }-T)S_m\omega =S_m(x_0\text{ id }-T)\omega =S_m\epsilon ^{(0)}=\epsilon ^{(m)}\quad \text{ for } \text{ all }\quad m \in {\mathbb {N}}_0. \end{aligned}$$

\(\square \)

For \(m \in {\mathbb {N}}_0\) define the sequence \(\eta ^{(m)}\) by

$$\begin{aligned} \eta ^{(m)}_k=\left\{ \begin{array}{lll} 0 &{}\quad \textrm{if}&{}\quad k=0,\ldots ,m-1, \\ \frac{H_k}{\gamma _k h_k^2}&{}\quad \textrm{if} &{}\quad k=m,m+1,\ldots \end{array}\right. . \end{aligned}$$
(41)

Note that \(\eta ^{(0)}=\eta \). If \(\left\{ \frac{H_n}{\gamma _nh_n}:n \in {\mathbb {N}}_0\right\} \) is bounded, then according to Lemma 2.3 we know that \(\eta ^{(m)}\in l^2(h)\) for all \(m \in {\mathbb {N}}_0\). Moreover,

$$\begin{aligned} C^*\eta ^{(m)}_n=\left\{ \begin{array}{lll} \sum _{k=m}^\infty \frac{\eta ^{(m)}_kh_k}{H_k} =\sum _{k=m}^\infty \frac{1}{\gamma _kh_k}=\omega _m &{}\quad \textrm{if}&{}\quad n\le m, \\ \sum _{k=n}^\infty \frac{1}{\gamma _kh_k}=\omega _n &{}\quad \textrm{if}&{}\quad n>m \end{array}\right. . \end{aligned}$$
(42)

By Proposition 2.2(ii) we have \(C^*\eta ^{(m)} =S_m\omega \) for all \(m \in {\mathbb {N}}_0\). Now we can combine the results above to determine the inverse operator of \(x_0\text{ id }-T\). Define a sequence \(\varphi =(\varphi _n)_{n \in {\mathbb {N}}_0}\) by

$$\begin{aligned} \varphi _n=\frac{H_n^2}{\gamma _nh_n^2}\quad \text{ for } \text{ all }\quad n \in {\mathbb {N}}_0. \end{aligned}$$
(43)

Note that \(\frac{H_n}{\mid \gamma _n\mid h_n} \le D\) for all \(n \in {\mathbb {N}}_0\) implies \(\mid \varphi _n\mid \le D^2B\) for all \(n \in {\mathbb {N}}_0\).

The multiplication with \(\varphi \in l^\infty \) defines a bounded operator \(M_\varphi \) on \(l^2(h)\), where \(M_\varphi (\xi )_n=\varphi _n\xi _n\) for all \(\xi \in l^2(h)\), \(n \in {\mathbb {N}}_0\).

Theorem 2.2

If \(x_0\) is a normalizing point with bounded growth of derivatives, then \(C^*\circ M_\varphi \circ C\) is the inverse of the operator \(x_0\text{ id }-T\), where \(\varphi \) is the sequence in (43).

Proof

Let \(m \in {\mathbb {N}}_0\). We know that \(C\epsilon ^{(m)}_k=0\) for all \(k=0,\ldots ,m-1\) and \(C\epsilon ^{(m)}_k=\frac{1}{H_k}\) for all \(k=m,m+1,\ldots \). Hence, \(M_\varphi \circ C\epsilon ^{(m)}=\eta ^{(m)}\) and \(C^* \circ M_\varphi \circ C\epsilon ^{(m)}=S_m\omega \). In particular

$$\begin{aligned} (x_0\text{ id }-T)\circ (C^*\circ M_\varphi \circ C)\epsilon ^{(m)}=\epsilon ^{(m)}\quad \text{ for } \text{ all }\quad m \in {\mathbb {N}}_0. \end{aligned}$$

Furthermore, we obtain

$$\begin{aligned}&C^*\circ M_\varphi \circ C\circ (x_0\text{ id }-T)\epsilon ^{(m)}\\&\quad = (C^* \circ M_\varphi \circ C)(x_0 \epsilon ^{(m)}-(\gamma _m \epsilon ^{(m+1)}+\beta _m\epsilon ^{(m)}+\alpha _m\epsilon ^{(m-1)}))\\&\quad = x_0S_m\omega -(\gamma _m S_{m+1}\omega +\beta _m S_m\omega +\alpha _m S_{m-1}\omega )\\&\quad = x_0S_m\omega -T \circ S_m\omega =(x_0\text{ id }-T)S_m\omega =\epsilon ^{(m)}\quad \text{ for } \text{ all }\quad m \in {\mathbb {N}}_0. \end{aligned}$$

Therefore,

$$\begin{aligned} C^*\circ M_\varphi \circ C\circ (x_0\text{ id }-T)= \text{ id } =(x_0\text{ id }-T)\circ C^* \circ M_\varphi \circ C, \end{aligned}$$

i.e. \((x_0 \text{ id }-T)^{-1}=C^*\circ M_\varphi \circ C.\) \(\square \)

Summing up the results we gain the following theorem.

Theorem 2.3

\(x_0 \notin \text{ supp }\mu = \sigma (T)\) if and only if \(x_0\) is a normalizing point with bounded growth of derivatives.

Finally, we want to show the relationship of the results here with [1, Theorem 2.3]. For this we use the terms \(\mathcal {A}\), A and \(\mathcal {D}(A)\) with the same meaning as in [1]. Let

$$\begin{aligned} \mathcal {A}=\left( \begin{array}{ccccccc} \beta _0&{}\lambda _0&{}0&{}0&{}0&{}\cdots &{}\\ \lambda _0&{}\beta _1&{}\lambda _1&{}0&{}0&{}\cdots &{}\\ 0&{}\lambda _1&{}\beta _2&{}\lambda _2&{}0&{}\cdots &{}\\ 0&{}0&{}\lambda _2&{}\beta _3&{}\lambda _3&{}\cdots &{}\\ \vdots &{}\vdots &{}\vdots &{}\ddots &{}\ddots &{}\ddots \end{array}\right) , \end{aligned}$$
(44)

where \((\beta _n)_{n\in {\mathbb {N}}_0}\) and \((\lambda _n)_{n\in {\mathbb {N}}_0}\) are the coefficients of (1). Then \(\mathcal {A}\) can be regarded as a linear operator \(\mathcal {A}: {\mathbb {C}}^{{\mathbb {N}}_0}\rightarrow {\mathbb {C}}^{{\mathbb {N}}_0}\), \(\xi \mapsto \mathcal {A}\xi =(\mathcal {A}\xi _n)_{n\in {\mathbb {N}}_0}\), where

$$\begin{aligned} \mathcal {A}\xi _n=\lambda _{n-1}\xi _{n-1}+\beta _n\xi _n+\lambda _n\xi _{n+1}\quad \text{ for } \text{ all }\quad n\in {\mathbb {N}}_0. \end{aligned}$$
(45)

Note that \(\lambda _{-1}=0\) and \(\xi _{-1}\) can be chosen arbitrary.

Moreover, let \(l^2=l^2(h)\) with \(h_n=1\) for all \(n\in {\mathbb {N}}_0\), and

$$\begin{aligned} c_{00}=\{\xi \in {\mathbb {C}}^{{\mathbb {N}}_0}: \#\{n\in {\mathbb {N}}_0: \xi _n\ne 0\}<\infty \}. \end{aligned}$$
(46)

Of course, \((c_{00},\Vert \Vert _2)\) is a subspace of the Hilbert space \((l^2,\Vert \Vert _2)\). As mentioned in [1] the linear operator

$$\begin{aligned} \mathcal {A}: c_{00}\rightarrow l^2, \quad \xi \mapsto \mathcal {A}\xi \end{aligned}$$
(47)

is closable and its closure is given by

$$\begin{aligned}{} & {} A: \mathcal {D}(A)\rightarrow l^2, \quad \xi \mapsto A \xi ,\quad \text{ where }\end{aligned}$$
(48)
$$\begin{aligned}{} & {} \mathcal {D}(A)=\left\{ \xi \in l^2: \exists (\xi ^{(k)})_{k\in {\mathbb {N}}_0} \subset c_{00}: \lim _{k\rightarrow \infty }\xi ^{(k)}=\xi \wedge \lim _{k\rightarrow \infty }\mathcal {A}\xi ^{(k)}\,\text{ exists }\right\} \nonumber \\ \end{aligned}$$
(49)
$$\begin{aligned}{} & {} \text{ and }\quad A \xi =\lim _{k\rightarrow \infty }\mathcal {A}\xi ^{(k)}. \end{aligned}$$
(50)

According to [1, Theorem 2.3.], the following statements hold true.

If \(x_0\in \Omega (A)={\mathbb {R}}{\setminus }\sigma (A)\) then

$$\begin{aligned} \sup _{n\ge 0}{\sum _{k=0}^n p_k^2(x_0) \over \lambda _n^2 (p_n^2(x_0)+p_{n+1}^2(x_0))} < \infty . \end{aligned}$$
(51)

Provided A is bounded, also the converse is true.

Note that the assumptions made at the beginning of Sect. 2 imply the boundedness of \((\lambda _n)_{n\in {\mathbb {N}}_0}\) and \((\beta _n)_{n\in {\mathbb {N}}_0}\). One can show that the boundedness of \((\lambda _n)_{n\in {\mathbb {N}}_0}\) and \((\beta _n)_{n\in {\mathbb {N}}_0}\) imply that A is a bounded operator.

The relationship with our result can be derived from

$$\begin{aligned}{} & {} (\alpha _{n+1} +\gamma _n) {\sum _{k=0}^n p_k^2(x_0) \over \lambda _n^2 (p_n^2(x_0)+p_{n+1}^2(x_0))}\nonumber \\{} & {} \quad = \alpha _{n+1}{ p_n^2 (x_0)+p_{n+1}^2(x_0)\over p_n^2 (x_0)} {\sum _{k=0}^n p_k^2(x_0) \over \lambda _n^2 (p_n^2(x_0)+p_{n+1}^2(x_0))}\nonumber \\{} & {} \quad ={\sum _{k=0}^n p_k^2(x_0) \over {\lambda _n^2 \over \alpha _{n+1}}p_n^2(x_0)}= {\sum _{k=0}^n p_k^2(x_0) \over \gamma _n p_n^2(x_0) }={H_n \over \gamma _n h_n } \end{aligned}$$
(52)

and \(\text {sign}\, \alpha _{n+1}= \text {sign}\,\gamma _n\).

In [1] there is no formula of the inverse as in Theorem 2.2 but there is no restriction \(x_0\in {\mathbb {R}}{\setminus } \mathcal {N}\).