1 Introduction

In the last twenty years there has been an intense research activity on m-isometries. In this paper, we focus our attention on characterizing m-isometries that have an invertible extension that is also m-isometry.

The notion of m-isometric operator on a Hilbert space was introduced by J. Agler [2] and studied in detail shortly after by J. Agler and M. Stankus in three papers [4,5,6]. These publications can be considered the first ones to initiate this topic of study.

An operator \(T\in L(H)\), the algebra of all bounded linear operators acting on a Hilbert space H, is called an m-isometry, for some positive integer m, if

$$\begin{aligned} \sum _{k=0}^{m} {m \atopwithdelims ()k} (-1)^{k} T^{*k}T^k=0, \end{aligned}$$

where \(T^*\) denotes the adjoint operator of T. When \(m=1\), we obtain an isometry. It is said that T is a strict m-isometry if either \(m=1\) or T is an m-isometry with \(m>1\) but it is not \((m-1)\)-isometry.

As one should expect, m-isometries share many important properties with isometries. For example, the following dichotomy property: the spectrum of an m-isometry is the closed unit disc if it is not invertible or a closed subset of the unit circle if it is invertible [4]. Also, if T is an m-isometry, then T is bounded below; that is, there exists \(M>0\) such that \(\Vert Tx\Vert \ge M\Vert x\Vert \) for every \(x\in H\).

Given an m-isometry \(T \in L(H)\), we are interested in research conditions which guarantee the existence of a Hilbert space K and an operator \(S \in L(K)\), which is an extension of T, such that S is an invertible m-isometry. To say that \(S \in L(K)\) is an extension of \(T \in L(H)\) means that K contains an isometric subspace to H, which we denote also by H, and the restriction \(S_{|H}\) from H to H coincides with T.

Problem 1.1

Characterize those m-isometric operators which have an invertible m-isometrical extension.

In 1969 Douglas [13] obtained that any isometry in a Banach space has an invertible isometric extension, also valid in a Hilbert space context. So, the case \(m=1\) holds. For \(m\ge 2\), first immediate consideration is that m must be odd, since every invertible m-isometry with even m is an \((m-1)\)-isometry by [4, Proposition 1.23].

Our problem is similar to others that arise naturally in Operator Theory and can be formulated in very general terms as follows. Given a class \({{\mathcal {C}} }\) of operators, for example defined on Hilbert spaces, and given a property P relative to those operators, we wish to characterize the operators that have an extension in the class \({{\mathcal {C}}}\) with property P.

Let \(T\in L(H)\) and \(S \in L(K)\) with H a closed subspace of K. Denote by \(P_H\) the orthogonal projection of K onto H and by J the inclusion of H into K. It is said that

  • S is a lifting of T if \(P_HS=TP_H\).

  • S is a dilation of T if \(T^n=P_HS^nJ\), for every \(n\in {\mathbb {N}}\).

Many authors have studied, for a given bounded linear operator \(T \in L(H)\), some additional properties of extension, lifting, or dilation of the operator T. The following results are known and respond to these problems:

  • Every contraction has an extension which in turn has a unitary lifting. Thus, every contraction has a unitary dilation. Also, a contraction has a lifting which is an isometry. See [16].

  • Every isometry has a unitary extension. See [13].

  • Every operator T such that the norms of its powers grow polynomially has an m-isometric lifting for some integer \(m\ge 1\). This lifting can be also extended to an invertible m-isometry. See [9].

Notice that the norms of the powers of an m-isometry have a polynomial behaviour (see part (1) of Proposition 2.1). However, there are operators such that those norms have a polynomial behaviour that are not m-isometries. In [9], the authors study lifting and dilations which are m-isometries. In particular, they obtain that if T is an m-isometry, then T has an \((m+3)\)-isometric lifting with other additional properties.

A special class of m-isometric operators is the \(\ell \)-Jordan isometries; that is, operators which are the sum of an isometry and an \(\ell \)-nilpotent operator which commute. It is known that every \(\ell \)-Jordan isometry is a strict \((2\ell -1)\)-isometry, but the converse is not valid. However, every strict m-isometry on a finite dimensional Hilbert space is an \(\frac{(m+1)}{2}\)-Jordan isometry operator. See [3, 12, 17] for more details.

Another natural and important examples of m-isometries are certain weighted shift operators. In [1, 11], the authors obtained a characterization of weighted shift which are m-isometric.

We summarize the contents of the paper. In Sect. 2, we define a bilateral sequence of operators associated to an m-isometry that allow us to transfer important information of the m-isometry to the bilateral sequence, that it will be an important tool in the paper. In Sect. 3, we present some necessary conditions to obtain an invertible m-isometrical extension. The main results are given in Sect. 4 where we obtain characterizations for an m-isometry to have an invertible m-isometrical extension. Finally, in Sect. 5, we present particular classes of m-isometries for which one can obtain nice results. In particular, we give a useful characterization for a general m-isometrical unilateral weighted shift and for \(\ell \)-Jordan isometries. In particular, every \(\ell \)-Jordan isometry operator has an invertible \((2\ell -1)\)-isometrical extension.

2 Some Previous Results

In this section, we define a bilateral sequence of operators associated to an m-isometry, that allow us to transfer important information of the m-isometry to the bilateral sequence that it will be relevant for obtaining necessary conditions for having an invertible m-isometrical extension.

Any polynomial of degree less or equal to \( m-1\) is uniquely determined by its values at m distinct points. If \(a_0,a_1,\dots ,a_{m-1}\) are given real (or complex) numbers, then the unique polynomial p of degree less or equal to \( m-1\) satisfying \(p(k)=a_k\) for all \(k\in \{ 0,1,\dots ,m-1\}\) is giving by Lagrange interpolating polynomial

$$\begin{aligned} p(z)=\sum _{k=0}^{m-1}a_k \prod _{0\le j\le m-1\atop j\ne k}\frac{z-j}{k-j} . \end{aligned}$$

Note that

$$\begin{aligned} p(n)=\sum _{k=0} ^{m-1} a_k b_k(n) \end{aligned}$$

with

$$\begin{aligned} b_k(n):= \prod _{0\le j\le m-1\atop j\ne k}\frac{n-j}{k-j}= (-1)^{m-k-1} \frac{n (n-1) \ldots \widehat{(n-k)} \cdots (n-m+1)}{k!(m-k-1)!}\;\nonumber \\ \end{aligned}$$
(2.1)

where \(\widehat{(n-k)}\) means that the factor \((n-k)\) is omitted.

Given \(T\in L(H)\), define the bilateral sequence by

$$\begin{aligned} D_n:=\sum _{k=0}^{m-1} b_k(n) T^{*k}T^k , \end{aligned}$$
(2.2)

for every \(n\in {\mathbb {Z}}\). Clearly \(D_n\in L(H)\) and it is self adjoint operator for every \(n\in {\mathbb {Z}}\).

Denote \(p_x(k):=\langle D_kx,x\rangle \) for every \(x\in H\) and \(k\in {\mathbb {Z}}\).

Given \(T\in L(H)\), denote \(T>0\) if \(\langle Tx,x \rangle >0\) for every \(x\in H\setminus \{ 0\}\) and we call it strictly positive operator.

We concentrate now on the family \((D_n)_{n\in {\mathbb {Z}}}\) of operators which arise from a fixed m-isometry. Indeed, the bilateral sequence \((D_n)_{n\in {\mathbb {Z}}}\) has some interesting properties that will be important tools to solve Problem 1.1.

Proposition 2.1

Let \(T\in L(H)\) be an m-isometry and \((D_n)_{n\in {\mathbb {Z}}}\) be operators defined by (2.2). Then

  1. (1)

    [11, Theorem 2.1] & [4] \(D_n=T^{*n}T^n\) and \(p_x(n)= \langle D_nx,x \rangle =\Vert T^nx\Vert ^2>0\) for every \(x\in H{\setminus } \{ 0\} \) and \(n\in {\mathbb {N}}\cup \{ 0\}\). Henceforth, there exists the square root \(D_n^{1/2}\) of \(D_n\), for every \(n\in {\mathbb {N}} \cup \{ 0\}\).

  2. (2)

    \(D_n\) is invertible for every \(n\in {\mathbb {N}} \cup \{ 0\}\).

  3. (3)

    \(T^{*k}D_n T^k=D_{n+k}\) for every \(n\in {\mathbb {Z}}\) and \(k\in {\mathbb {N}}\).

  4. (4)

    Let \(y\in R(T^k)\) for some \(k\in {\mathbb {N}}\). Then \(p_y(-k)=\Vert x\Vert ^2\), where \(y=T^kx\).

  5. (5)

    If \(D_{-n}>0\) and invertible, then \(D_{-k}>0\) and invertible for every \(k\in \{ 1, 2, \cdots , n-1\}\).

Proof

(2) Let \(n\in {\mathbb {N}}\). By [14] Theorem 2.3] & [10], Theorem 3.1] any power of T, \(T^n\) is an m-isometry, so, \(T^n\) is bounded below. Hence

$$\begin{aligned} \Vert D_nx\Vert \Vert x\Vert \ge |\langle D_nx, x\rangle |= \langle D_nx, x\rangle =\Vert T^nx\Vert ^2 \ge M(n)^2 \Vert x\Vert ^2 , \end{aligned}$$

where \(M(n)>0\). That is, \(D_n\) is bounded below. Then trivially \(D_n\) is invertible since \(D_n\) is self adjoint operator. (3) It is enough to prove the required equality for \(k=1\). Observe that

$$\begin{aligned} p_{Tx}(n)=\Vert T^nTx\Vert ^2 = \Vert T^{n+1} x\Vert ^2 =p_x(n+1), \end{aligned}$$

for every \(n\in {\mathbb {N}}\) and

$$\begin{aligned} \langle D_{n+1}x,x \rangle = p_x(n+1) =p_{Tx}(n)= \langle D_nTx, Tx\rangle =\langle T^* D_nTx,x\rangle \end{aligned}$$

for every \( n\in {\mathbb {Z}}\). (4) Let \(y=T^kx\) for some \(k\in {\mathbb {N}}\) and \(x\in H\). Then

$$\begin{aligned} p_{y}(n)= p_{T^kx}(n)=p_x (k+n) , \end{aligned}$$

for every \( n\in {\mathbb {N}} \). Therefore \(p_y(n)=p_x(k+n)\) for every \(n\in {\mathbb {Z}}\). (5) Let \(k\in \{ 1, 2, \cdots , n-1 \}\) and \(x\in H{\setminus }\{ 0\}\). If \(D_{-n}>0\), then by part (3),

$$\begin{aligned} \langle D_{-k}x, x \rangle = \langle T^{*n-k} D_{-n}T^{n-k}x,x \rangle =\langle D_{-n} T^{n-k}x,T^{n-k}x \rangle >0 . \end{aligned}$$
(2.3)

Since \(T^{n-k}\) is bounded below and by (2.3), we have that

$$\begin{aligned} \Vert D_{-k}^{1/2} x\Vert ^2 =\Vert D^{1/2} _{-n}T^{n-k} x\Vert ^2 \ge M \Vert x\Vert ^2 . \end{aligned}$$

So, the result is obtained since \(D_{-k}\) is a self adjoint operator. \(\square \)

We close this section by studying the bilateral sequence \((D_n)_{n\in {\mathbb {Z}}}\) associated to unilateral weighted shift which are m-isometries.

Let H be a Hilbert space with an orthonormal basis \((e_n)_{n\in {\mathbb {N}}}\). Recall that the unilateral weighted shift given by \(S_we_n=w_{n} e_{n+1}\) on H, where \(w_n= \displaystyle \sqrt{ \frac{p(n+1)}{p(n)}}\) with p a polynomial of degree \(m-1\), is a non invertible strict m-isometry, [1]. Also

$$\begin{aligned} p_{e_j}(n)=\Vert S^n_w e_j\Vert ^2= |w_j w_{j+1} \cdots w_{n+j-1}|^2= \frac{p(j+n)}{p(j)} . \end{aligned}$$
(2.4)

The following proposition gives an explicit expression of the operator \(D_n\), when T is an m-isometrical unilateral weighted shift operator.

Proposition 2.2

Let H be a Hilbert space with orthonormal basis \((e_n)_{n\in {\mathbb {N}}}\) and let \(S_w\in L(H)\) be an m-isometrical unilateral weighted shift with weight sequence \(w=(w_n)_{n\in {\mathbb {N}}}\). Then

  1. (1)

    \(D_n\) is a diagonal operator for every \(n\in {\mathbb {Z}}\), with diagonal

    $$\begin{aligned} \lambda _n(j):= \sum _{k=0}^{m-1} b_k(n) \prod _{\ell =j} ^{j+k-1} |w_{\ell } |^2 , \end{aligned}$$

    where \(b_k(n)\) is giving by (2.1).

  2. (2)

    Let \(n\in {\mathbb {Z}}\). The following conditions are equivalent

    1. (a)

      \(D_n\) is invertible.

    2. (b)

      \(D_n>0\).

    3. (c)

      \(\lambda _n(j)>0\) for every \(j\in {\mathbb {N}}\).

Proof

(1) By [1], there exists a polynomial p of degree \(m-1\), such that the weights are given by \(w_n=\sqrt{\frac{p(n+1)}{p(n)}}\). So,

$$\begin{aligned} D_ne_j= & {} \sum _{k=0}^{m-1} b_k(n) S_w^{*k} S_w^k e_j = \sum _{k=0}^{m-1} b_k(n) \prod _{\ell =j}^{j+k-1} |w_\ell |^2 e_j \nonumber \\= & {} \sum _{k=0}^{m-1} b_k(n) \frac{p(j+k)}{p(j )} e_j =\lambda _n(j)e_j\,, \end{aligned}$$
(2.5)

where

$$\begin{aligned} \lambda _n(j)= \sum _{k=0}^{m-1} b_k(n) \frac{p(j+k)}{p(j )} . \end{aligned}$$
(2.6)

(2) It is immediate by (1). \(\square \)

In general, the converse of part (5) of Proposition 2.1 is not valid. A suitable choose of the weight sequence gives an example such that \(D_{-q}>0\) and \(D_{-(q+1)}\) is not positive for some \(q\in {\mathbb {N}}\).

Example 2.3

Let \(q\in {\mathbb {N}}\) and define \(p_q(n):= (n+q)(n+q+1)\). Then \(S_w\) with weight \(w_n=\sqrt{\frac{p_q(n+1)}{p_q(n)}}\) is a 3-isometry and it satisfies that \(D_{-n}>0\) and invertible for \(n\in \{ 1, \cdots , q\}\) and \(D_{-(q+1)}\) is not. In fact,

$$\begin{aligned} \lambda _{-n}(j):= \frac{p_q(j-n)}{p_q(j)} =\frac{(j+q-n)(j+q-n+1)}{(j+q)(j+q+1)} \;, \end{aligned}$$

for \(n\in {\mathbb {N}}\). If \(n\in \{ 1, \cdots , q\}\), then we have that \( -q-1+n< -q+n < 0\). Hence, \(\lambda _{-n}(j) > 0\), for every \(j\in {\mathbb {N}} \). If \(n=q+1\),

$$\begin{aligned} \lambda _{-(q+1)}(j) = \frac{j(j-1)}{(j+q)(j+q+1)} . \end{aligned}$$

Hence \( \lambda _{-(q+1)} (1)=0\) and consequently \(\langle D_{-(q+1)}e_1,e_1\rangle =0\).

3 Necessary Conditions of Having an Invertible m-isometrical Extension

In an attempt towards solution of finding necessary conditions to obtain an invertible m-isometrical extension, we draw upon an interesting connection between \(D_{-1}>0\) and the invertibility of \(D_{-1} \) with the existence of a particular m-isometrical extension. Notice that in the following theorem we do not obtain an invertible m-isometrical extension.

Theorem 3.1

Let \(T\in L(H)\) be an m-isometry. The following statements are equivalent:

  1. (i)

    There exist a Hilbert space \(K\supset H\) and an m-isometry \(S\in L(K)\) such that \(S_{|H}=T\) and \(R(S)=H\).

  2. (ii)

    \(D_{-1}> 0\) and \(D_{-1}\) is invertible.

Proof

\({(i)\Rightarrow (ii)}\): Let \(x\in H\) and \(y=S^{-1}x\in K\). For \(n\in {\mathbb {Z}}\), denote

$$\begin{aligned} {\widetilde{D}}_n:= \sum _{k=0}^{m-1} b_k(n) S^{*k} S^k, \;\; \; {D}_n:= \sum _{k=0}^{m-1} b_k(n) T^{*k} T^k \end{aligned}$$

and for \(n\in {\mathbb {N}}\)

$$\begin{aligned} {\widetilde{p}}_x(n):=\Vert S^n x\Vert ^2, \;\; \; {p}_x(n):=\Vert T^n x\Vert ^2, \end{aligned}$$

where \(b_k(n)\) is given by (2.1). Then

$$\begin{aligned} \begin{aligned} \langle {\widetilde{D}}_{-1}x, x\rangle&= \langle {\widetilde{D}}_{-1} Sy, Sy \rangle =\langle S^* {\widetilde{D}}_{-1} Sy, y\rangle =\langle {\widetilde{D}}_{0} y, y\rangle =\Vert y\Vert ^2\\&= \sum _{k=0}^{m-1} b_{k}(-1) \langle S^{*k} S^k x,x \rangle = \sum _{k=0}^{m-1} b_{k}(-1) \langle T^kx, T^k x\rangle \\&= \langle D_{-1} x, x\rangle . \end{aligned} \end{aligned}$$

Then \(\langle {\widetilde{D}}_{-1} x,x \rangle =\Vert y\Vert ^2=\langle D_{-1}x,x\rangle \ge 0\) for all \(x\in H\). Also

$$\begin{aligned} \Vert D_{-1}x\Vert \Vert x\Vert \ge \langle D_{-1} x,x\rangle =\Vert y\Vert ^2\ge \frac{\Vert Sy\Vert ^2}{\Vert S\Vert ^2}=\frac{\Vert x\Vert ^2}{\Vert S\Vert ^2}. \end{aligned}$$

So, \(D_{-1}>0\) and bounded below. Hence \(D_{-1}\) is invertible since \(D_{-1}\) is self adjoint operator.

\({(ii)\Rightarrow (i)}\): Consider the vector space \(H\times H\) with a new seminorm

$$\begin{aligned} |||(h,h')|||:=\Vert D_{-1}^{1/2}(Th+h')\Vert \end{aligned}$$

and the subspace

$$\begin{aligned} N:=\{(h,h')\in H\times H \;\;: \;\; |||(h,h')|||=0\} . \end{aligned}$$

Let \(K:=(H\times H)/N\) with the quotient norm

$$\begin{aligned} |||(h,h')+ N|||:=\Vert D_{-1}^{1/2}(Th+h')\Vert . \end{aligned}$$

Then K is a normed space. Let us prove that \(|||\cdot |||\) satisfies the parallelogram law. For \(u=(h,h')+N\) and \(v=(g,g')+N\) in K we have

$$\begin{aligned} \begin{aligned} |||u+v|||^2+|||u-v|||^2=&\, \langle D_{-1} (Th+h' +Tg+g'),Th+h'+Tg+g'\rangle \\&+\langle D_{-1} (Th+h' -Tg-g'),Th+h'-Tg-g'\rangle \\ =&\,2\langle D_{-1} (Th+h'),Th+h'\rangle + 2\langle D_{-1} (Tg+g'),Tg+g'\rangle \\ =&\,2|||u|||^2+2|||v|||^2. \end{aligned} \end{aligned}$$

Henceforth, K is a pre-Hilbert space. The linear mapping \(\phi : K \longrightarrow H\) defined by \(\phi ((h,h')+N)= Th+h'\) is an isomorphism. Indeed, \(\phi \) is bounded since \(D_{-1}\) is an invertible operator. It is clear that \(\phi \) is onto and bounded below since the square root of \(D_{-1}\) is a bounded operator. Hence K is complete and so it is a Hilbert space. Moreover,

$$\begin{aligned} |||(h,0)+N|||^2= & {} \Vert D_{-1}^{1/2}(Th)\Vert ^2 =\langle D_{-1} Th,Th\rangle =\langle T^*D_{-1} Th,h\rangle \\= & {} \Vert D_0h\Vert ^2=\Vert h\Vert ^2. \end{aligned}$$

So K contains H as a subspace and we identify \(h\in H\) with \((h,0)+N\in K\).

Define S on K by \(\bigl ((h,h')+N\bigr ):= (Th+h',0)+N \). The operator S is well defined and bounded:

$$\begin{aligned} \begin{aligned} |||S\bigl ( (h,h')+N\bigr )|||^2&=|||(Th+h',0)+N |||^2= \Vert D_{-1}^{1/2}(T(Th+h'))\Vert ^2\\&= \langle D_{-1} (T(Th+h')), T(Th+h')\rangle = \langle D_0 (Th+h'),Th+h' \rangle \\&= \Vert Th+h'\Vert ^2\le \Vert D^{-1/2}_{-1} \Vert ^2 \Vert D_{-1}^{1/2} (Th+h')\Vert ^2 \\&= \Vert D_{-1}^{-1/2}\Vert ^2 |||(h,h')+N|||^2. \end{aligned} \end{aligned}$$

Clearly S is an extension of T. Let \(h\in H\). We have identified h with \((h,0)+N\in K\) and \(S((h,0)+N)=(Th,0)+N\). Also \(SK=H\).

Let us prove that S is an m-isometry. Let \(u=(h,h')+N\in K\) and write \(y:=Th+h'\in H\). We have that \(Su=(y,0)+N\), \(S^ku=(T^{k-1}y,0)+N\) and \(||| S^ku|||^2= \Vert D^{1/2}_{-1}(T^k y) \Vert ^2=\Vert T^{k-1} y\Vert ^2\) for \(k\in {\mathbb {N}}\). So

$$\begin{aligned} \begin{aligned} \sum _{k=0}^m (-1)^k{m\atopwithdelims ()k}|||S^ku|||^2&=|||u|||^2 + \sum _{k=1}^{m} (-1)^k {m \atopwithdelims ()k} |||S^{k}u|||^2\\&= \langle D_{-1}y,y\rangle + \sum _{k=1}^m (-1)^k{m\atopwithdelims ()k}\Vert T^{k-1}y\Vert ^2\\&=\sum _{k=0}^m(-1)^k{m\atopwithdelims ()k} p_y(k-1) =0, \end{aligned} \end{aligned}$$

since \(p_y\) has degree less or equal to \(m-1\). Hence S is an m-isometry. \(\square \)

The following result gives necessary conditions of having an invertible m-isometrical extension.

Proposition 3.2

Let \(T\in L(H)\) be a strict m-isometry.

  1. (1)

    If T is invertible, then \(p_x(n)=\Vert T^{n} x\Vert ^2>0\) for every \(x\in H\setminus \{ 0\}\) and \(n\in {\mathbb {Z}}\).

  2. (2)

    If T has an invertible m-isometrical extension S, then \(p_x(-k){:=}\Vert S^{-k}x\Vert ^2>0\) for every \(x\in H{\setminus } \{ 0\}\) and \(k\in {\mathbb {N}}\), where \(p_x(n):= \Vert T^nx\Vert ^2\) for \(n\in {\mathbb {N}}\). In particular, the degree of \(p_x\) is even for every \(x\in H\setminus \{ 0\}\).

  3. (3)

    If there exists an invertible m-isometrical extension of T, then \(D_{n}>0\) and invertible operator for every \(n\in {\mathbb {Z}}\).

Proof

(1) Part (3) of Proposition 2.1 yields that \(T^{*n}D_{-n}T^n=D_0=I\) for \(n\in {\mathbb {N}}\). So, for every \(x\in H{\setminus }\{ 0\}\) and \(n\in {\mathbb {N}}\),

$$\begin{aligned} p_x(-n)=\langle D_{-n}x,x\rangle =\langle T^{*-n} T^{-n}x,x\rangle =\Vert T^{-n}x\Vert ^2>0, \end{aligned}$$

since \(T^{-1}\) is an m-isometry.

(2) Let \(x\in H\) and \(n\in {\mathbb {N}}\). Denote by

$$\begin{aligned} \begin{aligned} p_x(n):&=\langle D_nx,x \rangle :=\sum _{k=0}^{m-1} b_{k } (n) \Vert T^kx\Vert ^2\\ {\widetilde{p}}_x(n):&=\langle {\widetilde{D}}_nx,x \rangle :=\sum _{k=0}^{m-1} b_{k } (n) \Vert S^kx\Vert ^2, \end{aligned} \end{aligned}$$

where S is an invertible m-isometrical extension of T. Clearly, \(p_x(n)={{\widetilde{p}}} _x(n)\) is a polynomial of degree less or equal to \(m-1\). Observe that \(p_x(-n)={{\widetilde{p}}}_x(-n)=\Vert S^{-n}x\Vert ^2\) for every \(n\in {\mathbb {N}}\). \(\square \)

Remark 3.3

  1. (1)

    Observe that part (2) of the above Proposition implies that the degree of \(p_x\) is even if \(p_x(n)>0\) for every \(n\in {\mathbb {Z}}\). Indeed, this is a different way to prove that there are no invertible strict m-isometries for even m. See also [4, Proposition 1.23].

  2. (2)

    The conditions \(D_n>0\) and invertible operator for every \(n\in {\mathbb {Z}}\) are not sufficient to define an invertible m-isometrical extension of T. Indeed, invertibility of \(D_n\) would suffice to construct an unbounded m-isometrical extension of T with dense range.

Proposition 3.2 allow us to obtain that some m-isometries have not an invertible m-isometrical extension.

Remark 3.4

Let \(T\in L(H)\) be a strict m-isometry. Denote \(p_x(n):=\Vert T^nx\Vert ^2 \), for \(n\in {\mathbb {N}}\) and \(x\in H\setminus \{ 0\}\). Then

  1. (1)

    If \(m=1\), then \(p_x(n)>0\) for every \(x\in H{\setminus } \{0\}\) and \(n\in {\mathbb {Z}}\).

  2. (2)

    If m is even, then there exist \(x_0\in H\) and \(n_0\in {\mathbb {Z}}\) with \(n_0<0\) such that \(p_{x_0} (n_0) \le 0\).

  3. (3)

    If m is odd, then it is possible that \(p_x(n)>0\) for every \(x\in H\setminus \{ 0\}\) and \(n\in {\mathbb {Z}}\) or there exist \(x_0\in H\) and \(n_0\in {\mathbb {Z}}\) with \(n_0<0\) such that \(p_{x_0} (n_0) \le 0\).

In the following examples we present different behaviours of \(p_x(n)\) with negative integer n for unilateral weighted shift.

Example 3.5

Let \(p(n)= n^{m-1}\) with odd m. It is clear that \(p_{e_j}(n):=\Vert S^n_we_j\Vert ^2=\displaystyle \left( \frac{j+n}{j}\right) ^{m-1}\) and \(p_{e_j}(-j)=0\). So, \(S_w\) can not have an invertible m-isometrical extension.

Example 3.6

Let \(p(n):= \prod _{i=1}^{m-1} (mn+i)\) with odd m. It is clear that

$$\begin{aligned} p_{e_j}(n):=\Vert S^n_we_j\Vert ^2= \frac{\displaystyle \prod \nolimits _{i=1}^{m-1} (m(j+n)+i)}{\prod _{i=1}^{m-1} (mj+i)}. \end{aligned}$$

If \(j\ge n\), then \(p_{e_j}(-n)>0\). In other case, \(p_{e_j}(-n)>0\) since \(m-1\) is even. As we will see later, \(S_w\) has an invertible m-isometrical extension by Theorem 5.1.

4 Characterization of Having an Invertible m-isometrical Extension

The main result of this paper is to obtain, for a fixed m-isometry, characterizations of having an invertible m-isometrical extension. In Proposition 3.2, we proved that a necessary condition is that the bilateral sequence of operators \((D_n)_{n\in {\mathbb {Z}}}\) must be strictly positive and invertible.

Now, we are in position to prove the main result.

Theorem 4.1

Let \(T\in L(H)\) be an m-isometry and let \((D_n)_{n\in {\mathbb {Z}}}\) be the bilateral sequence defined by (2.2). Denote \(p_x(n):=\langle D_nx,x\rangle \) for every \(x\in H{\setminus } \{ 0\}\) and \(n\in {\mathbb {Z}}\). The following statements are equivalent:

  1. (i)

    There exist a Hilbert space \(K\supset H\) and an invertible m-isometrical operator \(S\in L(K)\) such that \(S_{|H}=T\).

  2. (ii)

    \(p_x(j)>0\) for every \(x\in H{\setminus } \{0\}\), and \(j\in {\mathbb {Z}}\) and

    $$\begin{aligned} \sup \left\{ \frac{p_x(j+1)}{p_x(j)}: x\in H\setminus \{ 0\}, \;\; j\in {\mathbb {Z}}\right\} <\infty . \end{aligned}$$
    (4.7)
  3. (iii)

    \(D_{n}>0\) and invertible for every \(n\in {\mathbb {Z}} \), and

    $$\begin{aligned} \sup \left\{ \frac{\langle D_{-n+1}x,x\rangle }{\langle D_{-n}x,x\rangle }: x\in H,\;\; \Vert x\Vert =1, \;\; n\in {\mathbb {N}} \right\} <\infty . \end{aligned}$$
    (4.8)

Proof

\({(i) \Rightarrow (ii)}\): Let \(x\in H\setminus \{ 0\}\). Then

$$\begin{aligned} \Vert S^{j+1} x\Vert ^2= \Vert T^{j+1} x\Vert ^2= p_x(j+1)>0 \end{aligned}$$

for \(j\in {\mathbb {Z}}\) and

$$\begin{aligned} \frac{p_x(j+1)}{p_x(j)}=\frac{\Vert S^{j+1}x\Vert ^2}{\Vert S^jx\Vert ^2}\le \Vert S\Vert ^2. \end{aligned}$$

So, we get (4.7).

\({(ii) \Rightarrow (iii)}\): By parts (1) and (2) of Proposition 2.1 we have that \(D_n>0\) and invertible for \(n\in {\mathbb {N}}\). By hypothesis, \(D_j>0\) for \(j\in {\mathbb {Z}}\) since \(p_x(j)=\langle D_jx,x\rangle \). Let us prove that \(D_{-n}\) are bounded below for every \(n\in {\mathbb {N}}\). The condition (4.7) yields that there exists \(M>0\) such that

$$\begin{aligned} p_x(-n)\ge \frac{p_x(-n+1)}{M} \ge \frac{p_x(0)}{M^n}=\frac{\Vert x\Vert ^2}{M^n} \end{aligned}$$

hence

$$\begin{aligned} \Vert D^{1/2} _{-n} x\Vert ^2 \ge \frac{\Vert x\Vert ^2}{M^n}, \end{aligned}$$

for every \(x\in H\setminus \{ 0\}\) and \(n\in {\mathbb {N}}\). Therefore \(D_{-n}\) is bounded below for \(n\in {\mathbb {N}}\) and hence invertible.

It is remained to prove (4.8). Indeed, (4.8) is an immediate consequence of (4.7) using the identification \(p_x(j)=\langle D_{j} x,x \rangle \) for every \(x\in H{\setminus } \{ 0\}\) and \(j\in {\mathbb {Z}}\).

\({(iii)\Rightarrow (i)}\): Let V be the vector space of all sequences \((h_1,h_2,\dots )\) of elements of H with finite support, that is, there exists \(n\in {\mathbb {N}}\) such that \(h_j=0\) for \(j>n\). Define a new seminorm on V by

$$\begin{aligned} |||(h_1,h_2,\dots )|||^2:=\langle D_{-n}y,y\rangle , \end{aligned}$$

where \(n\in {\mathbb {N}}\) is any integer satisfying \(h_j=0\) for \(j>n\) and \(y:=\sum _{j=1}^nT^{n-j}h_j\).

The seminorm \(|||\cdot |||\) does not depend on the choice of n. Indeed, if \(h_j=0\) for \(j>n\), \(r=n+n_0\) with \(n_0\in {\mathbb {N}}\), and \(y=\sum _{j=0}^nT^{n-j}h_j\), then

$$\begin{aligned} \begin{aligned}&\left\langle D_{-r} \sum _{j=1}^{r} T^{r-j }h_j, \sum _{i=1}^{r} T^{r-i }h_i\right\rangle \\&\quad = \left\langle D_{-(n+n_0)} T^{n_0} \left( \sum _{j=1}^{n+n_0} T^{n-j} h_j\right) , T^{n_0}\left( \sum _{i=1}^{n+n_0} T^{n-i }h_i\right) \right\rangle \\&\quad = \left\langle T^{*n_0} D_{-(n+n_0)} T^{n_0} \left( \sum _{j=1}^n T^{n-j} h_j\right) , \sum _{i=1}^n T^{n-i} h_i \right\rangle = \langle D_{-n}y, y \rangle \end{aligned} \end{aligned}$$

where the last equality is by part (3) of Proposition 2.1.

Let \(N:=\{(h_1,h_2,\dots )\in V\;\;:\;\; |||(h_1,h_2,\dots )|||=0\}\) and let K be the completion of V/N.

Let us prove that K is a pre-Hilbert space. For that, it is enough to prove that \(|||\cdot |||\) satisfies the parallelogram law. Let \(u:=(h_1, h_2, \cdots ) +N, \;\; v:=(g_1, g_2, \cdots )+N \in V/N\), \(n\in {\mathbb {N}}\) such that \(h_j=0=g_j\) for \(j>n\) and \(x:= \sum _{j=1}^n T^{n-j} h_j\), \(y:=\sum _{j=1}^n T^{n-j} g_j\). Then

$$\begin{aligned} \begin{aligned} |||u+v|||^2 + |||u-v|||^2&= \langle D_{-n} (x+y),x+y \rangle +\langle D_{-n} (x-y), x-y\rangle \\&= 2(|||u|||^2 + |||v|||^2). \end{aligned} \end{aligned}$$

For each \(h\in H\) we have \(|||(h,0,0,\dots )+N|||^2=\langle D_{-1} Th, Th \rangle =\langle D_0 h,h\rangle =\Vert h\Vert ^2\).

Let L be the closed subspace generated by \((h, 0, \cdots )+N \) with \(h\in H\) and define \(\phi \) on H taking values on L by \( \phi (h):=(h, 0, \cdots )+ N\). Then \(\Vert h\Vert ^2=||| \phi (h)|||^2 \) and \(R(\phi ) = L\). For each \(h\in H\) we can identify h with \((h,0,\dots )+N\in K\). So, K contains H as a subspace.

Define S on V/N by \(S ((h_1, h_2, \cdots )+N):=(Th_1+h_2, h_3, \cdots )+ N\in V/N \). Then the definition of S is correct and S is bounded. Indeed, let \(u:=(h_1, h_2, \cdots ) + N\in V/N\), \(n\in {\mathbb {N}}\) such that \(h_j=0\) for \(j>n\) and \(y:= \sum _{j=1}^n T^{n-j} h_j\). Denote \(({{\widetilde{h}}}_1, {{\widetilde{h}}}_2, \cdots ):=(Th_1+h_2, h_3, \cdots )\). Then

$$\begin{aligned} \begin{aligned} |||Su|||^2&= |||(Th_1+h_2, h_3, \cdots )+N|||^2= \langle D_{-(n-1)} {{\widetilde{y}}}, {{\widetilde{y}}} \rangle \end{aligned} \end{aligned}$$

where

$$\begin{aligned} {{\widetilde{y}}}:= \sum _{j=1}^{n-1} T^{n-1-j} {{\widetilde{h}}}_j= T^{n-1} (Th_1+h_2) + \sum _{j=2} ^{n-1} T^{n-1-j} {\widetilde{h}}_j =y \;. \end{aligned}$$

Then \(|||Su|||^2=\langle D_{-(n-1)} y, y\rangle =p_y (-n+1). \) Repeating the process we have that

$$\begin{aligned} |||S^ku|||^2 = p_y (-n+k), \end{aligned}$$

for \(k=0,\dots m\). Therefore

$$\begin{aligned} \sum _{k=0}^m(-1)^k{m\atopwithdelims ()k}|||S^ku|||^2= \sum _{k=0}^m(-1)^k{m\atopwithdelims ()k} p_y(-n+k)=0, \end{aligned}$$

since \( p_y\) has degree less or equal to \(m-1\). By continuity, S is an m-isometry.

It is easy to see that \(R(S)\supset V+N\). So the range of S is dense, and consequently S is an invertible m-isometry. \(\square \)

Moreover, the invertible extension \(S\in L(K)\) is defined uniquely (up to the unitary equivalence) if we assume that S is minimal, i.e., \(K=\bigvee _{k\ge 0} S^{-k}H\).

We will prove that the converse of part (3) of Proposition 3.2 is not true in general, that is, if \(D_n>0\) and invertible for \(n\in {\mathbb {Z}} \) are not sufficient to have an invertible m-isometrical extension of an m-isometry. Firstly, we need a previous result on m-isometries.

Proposition 4.2

Let \((T_n)_{n\in {\mathbb {N}}}\subset L(H)\) be a uniformly bounded sequence of m-isometries. Then \(T=T_1 \oplus T_2 \oplus \cdots \) is an m-isometry on \(\ell ^2(H)\).

Proof

Since \((T_n)_{n\in {\mathbb {N}}}\) is a uniformly bounded, then \(T=T_1 \oplus T_2 \oplus \cdots \) is well-defined on \(\ell ^2(H)\).

Let \(x=(x_1, x_2, \cdots )\in \ell ^2(H)\). Denote \(p_{x_n}(k):=\Vert T_n^kx_n\Vert ^2\). Since \((T_n)_{n\in {\mathbb {N}}}\) is a sequence of m-isometries, then \((p_{x_n} (k))_{n\in {\mathbb {N}}}\) is a sequence of polynomials of degree less or equal to \(m-1\). Fixed \(k\in {\mathbb {N}}\),

$$\begin{aligned} p_x(k):=\Vert T^kx\Vert ^2= \sum _{n=1} ^\infty \Vert T^k_nx_n\Vert ^2 = \sum _{n=1}^\infty p_{x_n} (k)\; \end{aligned}$$

is a polynomial of degree less or equal to \(m-1\). Hence T is an m-isometry. \(\square \)

It is possible to exhibit an example of m-isometry with odd m such that \(D_n>0\) and invertible for every \(n\in {\mathbb {Z}}\) but not fulfilling the hypothesis of Theorem 4.1. In order to simplify the presentation we include an example with a 3-isometry.

Example 4.3

Let \(q_n(j):= j^2+j(2-\frac{1}{n})+1\). Let H be a Hilbert space with an orthonormal basis \((e_{n,j})_{n,j\in {\mathbb {N}}}\) and \(K:=\ell ^2 (H)\). Define \(T\in L(K)\) by

$$\begin{aligned} Te_{n,j}:= \sqrt{ \displaystyle \frac{q_n(j+1)}{q_n(j)}} e_{n,j+1} \end{aligned}$$

for any \(n,j\in {\mathbb {N}}\). Then

  1. (1)

    T is a 3-isometry on K.

  2. (2)

    \(p_x(k)>0\) for every \(x\in K{\setminus }\{ 0\}\) and \(k\in {\mathbb {Z}}\), where \(p_x(n):=\Vert T^n x\Vert ^2\) for \(n\in {\mathbb {N}}\).

  3. (3)

    \(D_n>0\) and invertible for \(n\in {\mathbb {Z}}\).

  4. (4)

    There is no invertible 3-isometrical extension of T.

Proof: It is clear that \(q_n(j)>0 \) for \(n\in {\mathbb {N}}\) and \(j\in {\mathbb {Z}}\).

Let \(x=(x_1, x_2, \cdots )=\left( \sum _{n=1}^\infty \alpha _{n,1}e_{n,1}, \sum _{n=1}^\infty \alpha _{n,2}e_{n,2}, \cdots \right) \in K\). Then

$$\begin{aligned} T(x_1, x_2, \cdots ):= (0, T_1x_1, T_2,x_2, \cdots ), \end{aligned}$$

where

$$\begin{aligned} T_ix_i:=T_i\left( \sum _{n=1}^{\infty }\alpha _{n,i}e_{n,i}\right) = \sum _{n=1}^\infty \alpha _{n,i} w_{n,i} e_{n,i+1} \end{aligned}$$

and

$$\begin{aligned} w_{n,i}:=\sqrt{\frac{q_n(i+1)}{q_n(i)}} . \end{aligned}$$

By Proposition 4.2, the operator T is a 3-isometry, since \(T_n\) is a 3-isometry for every \(n\in {\mathbb {N}}\) and also \((T_n)_{n\in {\mathbb {N}}}\) is uniformly bounded, that is

$$\begin{aligned} \sup _{n\in {\mathbb {N}}} \Vert T_n\Vert \le \sup _{n,i\in {\mathbb {N}}} \sqrt{\frac{q_n(i+1)}{q_n(i)}}<M \end{aligned}$$

for some positive constant M.

Let us prove that \(p_x(k)>0\) for every \(x\in K{\setminus }\{ 0\}\) and \(k\in {\mathbb {Z}}\). Let \(x=(x_1, x_2, \cdots )=(\sum _{n=1}^\infty \alpha _{n,1}e_{n,1}, \sum _{n=1}^\infty \alpha _{n,2}e_{n,2}, \cdots )\in K{\setminus } \{ 0\}\) and \(k\in {\mathbb {N}}\). Then

$$\begin{aligned} \begin{aligned} p_x(k):=\Vert T^kx\Vert ^2&= \Vert (0, \cdots , 0, T_kT_{k-1} \cdots T_1 x_1, T_{k+1}T_{k} \cdots T_2 x_2, \cdots \Vert ^2\\&= \left\| [\right\| 3]{ (0, \cdots , 0, \sum _{n=1}^\infty \alpha _{n,1} \sqrt{\frac{q_{n}(k+1)}{q_n(1)}} e_{n,k+1}, \cdots )} ^2\\&= \sum _{j=1}^\infty \left\| [\right\| 3]{ \sum _{n=1}^\infty \alpha _{n,j} \sqrt{\frac{q_{n}(k+j)}{q_n(j)}}e_{n,k+j}} ^2 \\&= \sum _{n,j=1}^\infty |\alpha _{n,j} |^2 \frac{q_n(k+j)}{q_n(j)} >0 \end{aligned} \end{aligned}$$

for \(k\in {\mathbb {N}}\). Notice that

$$\begin{aligned} D_{-n}:= \frac{(n+1)(n+2)}{2} I -n(n+2)T^*T+ \frac{n(n+1)}{2} T^{*2} T^2, \end{aligned}$$

is a diagonal operator given by \( D_{-n} e_{m,j}=\lambda _{-n}(k,j)e_{k,j} \) where

$$\begin{aligned} \begin{aligned} \lambda _{-n}(k,j):=&\,\frac{1}{2q_k(j)} \left( (n+1)(n+2)q_k(j)-n(n+2)q_k(j+1)+n(n+1)q_k(j+2)\right) \\ =&\, \frac{1}{2q_k(j)} \left( j^2(n^2+2n+2)+j\left( -\frac{n^2}{k} +4n^2-2\frac{n}{k} +4n -\frac{2}{k} +4\right) \right. \\&\left. -\frac{n^2}{k} +6n^2 +4n +2\right) >0 , \end{aligned} \end{aligned}$$

for \(n,k,j\in {\mathbb {N}}\). So, it is immediate that \(D_{-n}\) is invertible for \(n\in {\mathbb {N}}\).

In order to finish the proof, let us prove that there is no invertible 3-isometrical extension of T. Taking into account that

$$\begin{aligned} \frac{p_{e_{n,1}}(-1)}{p_{e_{n,1}}(-2)}= \frac{q_n(0)}{q_n(-1)}=n, \end{aligned}$$

we have that

$$\begin{aligned} \sup \left\{ \frac{p_x(j+1)}{p_x(j)} \;:\;\; x\in K\setminus \{ 0\}, \;\; j\in {\mathbb {Z}}\right\} =\infty . \end{aligned}$$

\(\Box \)

5 Some Particular Cases

In this section, the goal is to study two different examples of m-isometries, the \(\ell \)-Jordan isometry and unilateral weighted shift that are m-isometries for some m.

In the case of unilateral weighted shift we can obtain a nice characterization of invertible m-isometrical extensions of an m-isometry, as a consequence of Theorem 4.1.

Theorem 5.1

Let H be a Hilbert space with orthonormal basis \((e_n)_{n\in {\mathbb {N}}}\) and let \(S_w\in L(H)\) be an m-isometrical unilateral weighted shift associated to the weight \(w:=(w_n)_{n\in {\mathbb {N}}}\). Then \(S_w\) has an invertible m-isometrical extension if and only if \(p_{e_1}(n)>0 \) for every \(n\in {\mathbb {Z}}\), where \(p_{e_1} (n):=\Vert S^n_w e_1\Vert ^2\) for \(n\in {\mathbb {N}}\).

Proof

If \(S_w\) has an invertible m-isometrical extension S, then \(p_x(n):=\Vert S^nx\Vert ^2>0\) for every \(x\in H{\setminus } \{ 0\}\) and \(n\in {\mathbb {Z}}\), by Proposition 3.2. Hence \(p_{e_1}(n)>0\) for \(n\in {\mathbb {Z}}\).

Let us prove the sufficient condition. Suppose that \(p_{e_{1}}(n)>0\) for \(n\in {\mathbb {Z}}\). A first consequence is that m is odd. By equality (2.4), \(p_{e_1} (n)\) is a polynomial of degree \(m-1\). Hence

$$\begin{aligned} \lim _{n\rightarrow \infty } \frac{ p_{e_1} (-n+1)}{p_{e_1} (-n)}=1, \end{aligned}$$

and

$$\begin{aligned} \inf \left\{ \frac{p_{e_1} (-n+1)}{p_{e_1} (-n)} \;:\;\; n\in {\mathbb {N}} \right\} >0 . \end{aligned}$$

Let K be a Hilbert space with \((e_n)_{n\in {\mathbb {Z}}}\) an orthonormal basis. Define \(T_{\beta } \in L(K)\) by \(T_\beta e_n=\beta _n e_{n+1}\) where \(\beta _n=\sqrt{\displaystyle \frac{p_{e_1}(n)}{p_{e_1}(n-1)}}\) for \(n\in {\mathbb {Z}}\). By [1, Theorem 19] we have that \(T_\beta \) is an m-isometry, since \(p_{e_1}(n)\) is a polynomial of degree \(m-1\) by (2.4). Moreover, \(T_\beta \) is an invertible extension of \( S_w\) and the desired result is proved. \(\square \)

Remark 5.2

In the above theorem, it is possible to obtain the same information with different elements of the orthogonal basis, as a consequence of equality (2.4). Indeed, in the conditions of Theorem 5.1 the following statements are equivalent:

  1. (1)

    \(S_w\) has an invertible m-isometrical extension.

  2. (2)

    \(p_{e_1}(n)>0 \) for \(n\in {\mathbb {Z}}\).

  3. (3)

    \(p_{e_j}(n)>0 \) for \(n\in {\mathbb {Z}}\) and some \(j\in {\mathbb {N}}\).

  4. (4)

    \(p_{e_j}(n)>0 \) for \(n\in {\mathbb {Z}}\) and \(j\in {\mathbb {N}}\).

Let us obtain a first approach to \(\ell \)-Jordan isommetries. In the next result we obtain that any 2-Jordan isometry operator admits an invertible 3-isometric extension, as a particular case of Theorem 4.1.

Corollary 5.3

Let \(T\in L(H)\) be a 2-Jordan isometry operator. Then T has an invertible 2-Jordan isometry extension.

Proof

Let T be a 2-Jordan isometry operator, that is \(T=A+Q\), where A is an isometry and Q is a 2-nilpotent operator such that \(AQ=QA\). By (2.2) we obtain that

$$\begin{aligned} \begin{aligned} D_{-n}&=\frac{(n+1)(n+2)}{2} I -n(n+2) T^*T+\frac{n(n+1)}{2}T^{*2}T^2 \\&= I-n(A^*Q+Q^*A) +n^2 Q^*Q . \end{aligned} \end{aligned}$$

Then

$$\begin{aligned} \langle D_{-n}x,x \rangle = \Vert x\Vert ^2-n(\langle Qx,Ax\rangle + \langle Ax, Qx\rangle ) +n^2 \Vert Qx\Vert ^2 . \end{aligned}$$

Let us prove that \(\langle D_{-n} x,x\rangle >0\) for every \(x\in H\) such that \(\Vert x\Vert =1\) and \(n\in {\mathbb {N}}\). It is enough to prove that

$$\begin{aligned} n^2\Vert Qx\Vert ^2+1> 2nRe(\langle Ax,Qx\rangle ) , \end{aligned}$$
(5.9)

where Re(z) denotes the real part of z. If \(Re (\langle Ax, Qx\rangle ) \le 0\), then (5.9) is clear. Assume that \(Re (\langle Ax, Qx\rangle ) > 0\). Then

$$\begin{aligned} Re (\langle Ax, Qx\rangle )=| Re (\langle Ax, Qx\rangle ) | \le |\langle Ax, Qx\rangle | \le \Vert Ax\Vert \Vert Qx\Vert \le \Vert Q\Vert . \end{aligned}$$

If \(|\langle Ax, Qx \rangle |=\Vert Ax \Vert \Vert Qx\Vert \), then the vectors Ax and Qx are linearly dependent, so there exists \(\lambda \) such that \(Qx=\lambda Ax\). Then \(\lambda =0\), since \(0=\Vert Q^2x\Vert =|\lambda |^2 \Vert A^2x\Vert =|\lambda |^2 \) and therefore \(\Vert Qx\Vert =0\), which is an absurd with \(Re (\langle Ax, Qx\rangle > 0\). If \(|\langle Ax, Qx \rangle | < \Vert Ax\Vert \Vert Qx\Vert \), then

$$\begin{aligned} 2nRe(\langle Ax, Qx \rangle ) <2n \Vert Qx\Vert \le n^2 \Vert Qx\Vert ^2+1. \end{aligned}$$

So, \(\langle D_{-n} x,x\rangle >0\) for every \(x\in H\) such that \(\Vert x\Vert =1\) and all \(n\in {\mathbb {N}}\).

In order to get the result, it is enough to prove that (4.8) is bounded. Let \(x\in H\) such that \(\Vert x\Vert =1 \) and \(n\in {\mathbb {N}}\). Then

$$\begin{aligned} \begin{aligned} \frac{\langle D_{-n+1}x,x\rangle }{\langle D_{-n}x,x\rangle }&= 1+\frac{ 2Re(\langle Ax,Qx \rangle )+(-2n+1)\Vert Qx\Vert ^2}{1-2nRe(\langle Ax, Qx \rangle ) +n^2 \Vert Qx\Vert ^2}\\[1pc]&\le 1+ \left| \frac{ 2Re(\langle Ax,Qx \rangle )+(-2n+1)\Vert Qx\Vert ^2}{1-2nRe(\langle Ax, Qx \rangle ) +n^2 \Vert Qx\Vert ^2} \right| \\[1pc]&\le 1+\frac{ 2 \Vert Q\Vert + (2n-1)\Vert Q\Vert ^2}{1-2n\Vert Q\Vert -n^2 \Vert Q \Vert ^2 } \end{aligned} \end{aligned}$$

converges to zero as n tends to infinity. Hence

$$\begin{aligned} \sup \left\{ \frac{\langle D_{-n+1} x,x\rangle }{\langle D_{-n} x,x\rangle } \;\;: \;\; x\in H, \;\; \Vert x\Vert =1, \;\; n\in {\mathbb {N}} \right\} <\infty . \end{aligned}$$

\(\square \)

Corollary 5.4

Let \(T, \; C\in L(H)\) such that \(TC=CT\).

  1. (1)

    If T is an isometry, then \({\widetilde{T}}:= \left( \begin{array}{ll} T &{} C \\ 0 &{} T \end{array} \right) \) has an invertible 3-isometric extension on \(K\supset H\oplus H\).

  2. (2)

    If \(\lambda T\) is an isometry for some \(\lambda \in {\mathbb {C}}\), then \(\lambda {\widetilde{T}}=\lambda \left( \begin{array}{cc} T &{} C \\ 0 &{} T \end{array} \right) \) has an invertible 3-isometric extension on \(K\supset H\oplus H\).

Proof

(1) It is clear that \({\widetilde{T}} =\left( \begin{array}{cc} T &{} 0\\ 0 &{} T\end{array} \right) +\left( \begin{array}{cc} 0 &{} C\\ 0 &{} 0\end{array} \right) \) is a 2-Jordan isometry operator. Therefore the result is consequence of Corollary 5.3.

Applying (1) to the operator \(\lambda T\) we obtain (2). \(\square \)

A similar result of part (1) of Corollary 5.4 was obtained in [8, Corollary 4.4]. That is, if \(T\in L(H)\) is a contraction and \(C\in L(H)\) such that \(TC=CT\), then \({\widetilde{T}}\) has a 3-isometric lifting on \(K\supset H\oplus H\).

In the next theorem we can improve Corollary 5.3. Indeed, we prove that every \(\ell \)-Jordan isometry has an invertible \(\ell \)-Jordan isometry extension. The first part of our proof is based in the construction by Douglas [13], as it is presented by Laursen and Neumann in the monograph [15, Proposition 1.6,6].

Theorem 5.5

Let \(T \in L(H)\) be an \(\ell \)-Jordan isometry. Then there exist a Hilbert space K and \(S\in L(K)\), such that H is isometrically embedded in K and S is an invertible \(\ell \)-Jordan isometry extension of T.

Proof

As T is an \(\ell \)-Jordan isometry, there are an isometry \(A \in L(H)\) and an \(\ell \)-nilpotent operator \(Q \in L(H)\) such that \(AQ=QA\) and \(T=A+Q\).

Let \(K_0\) be the linear space of all the sequences \(u=(u_n)_{n\in {\mathbb {N}}}\) in H such that there is \(m \in {\mathbb {N}}\) satisfying \(u_{m+k} =A^{k} u_m\), for \(k \in {\mathbb {N}}\). Define, for \(u,v \in K_0\),

$$\begin{aligned} \langle u,v \rangle _0:= \lim _{n \rightarrow \infty } \langle u_n,v_n \rangle \;, \end{aligned}$$

being \(\langle \cdot , \cdot \rangle \) the inner product on H. Note that there exists \(m \in {\mathbb {N}}\) such that \(\langle u_m,v_m \rangle = \langle A^k u_m, A^k v_m \rangle = \langle u_{m+k},v_{m+k} \rangle \), so the sequence \((\langle u_n,v_n \rangle )_{n\in {\mathbb {N}}}\) is eventually constant, that is, there exists \(k_0\in {\mathbb {N}}\) such that \(\langle u_n, v_n\rangle \) is constant for \(n> k_0\). It is routine to verify what \(\langle \cdot ,\cdot \rangle _0\) is a semi-inner product on \(K_0\). Therefore \(K_0\) is a semi pre-Hilbert space. Moreover,

$$\begin{aligned} \Vert u\Vert _0^2:= \langle u,u \rangle _0 = \lim _{n \rightarrow \infty } \langle u_n,u_n \rangle = \lim _{n \rightarrow \infty } \Vert u_n\Vert ^2 \end{aligned}$$

defines a seminorm \(\Vert \cdot \Vert _0\) on \(K_0\).

Let \(M:= \{ u \in K_0: \langle u,u \rangle _0 = \Vert u \Vert _0^2 = 0 \}\). Then M is a closed subspace of \(K_0\) and we consider the quotient space \(K_0/M \). In this space are defined, for \(u,v \in K_0\),

$$\begin{aligned}&\langle u+M,v+M \rangle := \langle u,v \rangle _0 \; \; \text{ and } \\&\Vert u + M\Vert ^2:= \langle u+M,u+M \rangle = \langle u,u \rangle _0 = \Vert u\Vert _0^2 \;, \end{aligned}$$

and we obtain that \(K_0/M\) is a pre-Hilbert space.

Denote by K the Hilbert space what it is the completion of \(K_0/M\). The operator \(J \in L(H,K)\), defined by \(Jx:= (A^nx)_{n\in {\mathbb {N}}}+M\) for \(x \in H\), satisfies that

$$\begin{aligned} \Vert Jx \Vert = \Vert (A^nx)_{n\in {\mathbb {N}}}+M \Vert = \Vert (A^nx)_{n\in {\mathbb {N}}}\Vert _0 = \lim _{n \rightarrow \infty } \Vert A^n x\Vert = \Vert Ax\Vert = \Vert x\Vert \;, \end{aligned}$$

hence J is an isometry. So K contains an isometric copy of H. It is clear that J(H) is a closed subspace of K.

In order to define \(B\in L(K)\), we define an isometry on \(K_0/M\) by

$$\begin{aligned} B((u_n)_{n\in {\mathbb {N}}} +M):= (A u_n)_{n\in {\mathbb {N}}} +M \;, \end{aligned}$$

for every \((u_n)_{n\in {\mathbb {N}}} +M \in K_0/M\). Note that B is a linear isometry whose range contains \(K_0/M\); in fact, given \((v_n)_{n\in {\mathbb {N}}} + M = (v_1,...,v_m,Av_m,A^2 v_m,...)+ M\), we have that

$$\begin{aligned} \begin{aligned} B((\underbrace{0,..., 0}_{m}, v_m,Av_m,A^2 v_m,...)+M)&= (\underbrace{0,..., 0}_{m},Av_m,A^2v_m,A^3 v_m,...)+M \\&= (v_1, \cdots , v_m, Av_m, A^2v_m, \cdots ) +M . \end{aligned} \end{aligned}$$

As \(K_0/M\) is dense in K, we have that B can be extended to an invertible isometry defined on K. Moreover, B can be considered as an extension of A since, for \(x \in H\),

$$\begin{aligned} BJx = B((A^nx)_{n\in {\mathbb {N}}} + M) = (A^{n+1}x)_{n\in {\mathbb {N}}} + M = JAx \;. \end{aligned}$$

That is, \(BJ=JA\).

Define \(P \in L(K)\) in the following way

$$\begin{aligned} P((u_n)_{n\in {\mathbb {N}}} + M) = (Qu_n)_{n\in {\mathbb {N}}} + M \;, \end{aligned}$$

for every \((u_n)_{n\in {\mathbb {N}}} + M \in K_0/M\). It is clear that P is an \(\ell \)-nilpotent. Let us prove that B and P commute. Taking into account that \(AQ=QA\), we have that

$$\begin{aligned} \begin{aligned} BP ((u_n)_{n\in {\mathbb {N}}} + M )&= B ((Qu_n)_{n\in {\mathbb {N}}} + M) = (AQu_n)_{n\in {\mathbb {N}}} + M \\&= (QAu_n)_{n\in {\mathbb {N}}} + M = P ((Au_n)_{n\in {\mathbb {N}}} + M) \\&= PB ((u_n)_{n\in {\mathbb {N}}} + M) \;. \end{aligned} \end{aligned}$$

for every \((u_n)_{n \in {\mathbb {N}}} + M \in K_0/M\). Therefore, \(S:=B+P\in L(K)\) is an \(\ell \)-Jordan isometry that extends T. Moreover, S is an invertible since \(\sigma (S)=\sigma (B)\) and B is an invertible isometry. So the proof is finished. \(\square \)

An operator \(T \in L(H)\) is a doubly \(\ell \)-Jordan isometry if \(T=A+Q\) is an \(\ell \)-Jordan isometry operator such that the \(\ell \)-nilpotent \(Q \in L(H)\) which commutes with A also commutes with \(A^*\). For all scalar \(\lambda \) with \(|\lambda |=1\) and an \(\ell \)-nilpotent operator Q, we have that \(\lambda I + Q\) is a doubly \(\ell \)-Jordan isometry.

Corollary 5.6

Let \(T \in L(H)\) be a doubly \(\ell \)-Jordan isometry. Then there exist a Hilbert space K, such that H is isometrically embedded in K and an invertible doubly \(\ell \)-Jordan isometry extension \(S\in L(K)\) of T.

Remark 5.7

We use the notation of the proof of Theorem 5.5.

  1. (1)

    It is easy to prove that the orthogonal subspace of J(H), \(J(H)^\bot \) is the closure of the subspace of all classes

    $$\begin{aligned} (u_n)_{n\in {\mathbb {N}}} + M = (u_1,...,u_m,Au_m,A^2u_m,...) + M \in K_0/M \end{aligned}$$

    such that \(u_m \in R(A^m)^\bot \).

  2. (2)

    The decomposition \(K=J(H) \oplus J(H)^\bot \) gives rise to the representation of B as a operator matrix:

    $$\begin{aligned} B = \left( \begin{array}{cc} B_1 &{} B_2 \\ 0 &{} B_3 \\ \end{array} \right) \end{aligned}$$
    (5.10)

    being \(B_1 \in L(J(H))\), \(B_2\in L(J(H)^\bot ,J(H))\) and \(B_3 \in L(J(H)^\bot )\). Notice that J(H) is a closed invariant subspace of B.

  3. (3)

    The operator P is defined by the following operator matrix, associated to the decomposition \(K=J(H) \oplus J(H)^\bot \),

    $$\begin{aligned} P=\left( \begin{array}{cc} P_1 &{} P_2 \\ 0 &{} P_3 \\ \end{array} \right) \end{aligned}$$
    (5.11)

    being \(P_1 \in L(J(H))\), \(P_2 \in L(J(H)^\bot ,J(H))\) and \(P_3 \in L(J(H)^\bot )\). Notice that J(H) is a closed invariant subspace of P.

  4. (4)

    If T is a doubly \(\ell \)-Jordan isometry, then \(P_2=0\) in (5.11). For this purpose only it is necessary to prove that if \((u_n)_{n\in {\mathbb {N}}} + M \in J(H)^\bot \), then \(P((u_n)_{n\in {\mathbb {N}}} + M ) \in J(H)^\bot \), and that \(BP^*=P^*B\). In fact, given \(u = (u_1,...,u_m,Au_m, A^2 u_m,...)\) such that \(u_m \in R(A^m)^\bot \), we have that \(Qu_m \in R(A^m)^\bot \) since, for all \(x \in H\),

    $$\begin{aligned} \langle Qu_m,A^mx \rangle = \langle u_m,Q^*A^mx \rangle = \langle u_m,A^m Q^*x \rangle = 0 \;, \end{aligned}$$

    because \(Q^*A= AQ^*\). Therefore \(P((u_n)_{n\in {\mathbb {N}}} + M ) = (Qu_1,...,Qu_m,AQu_m, A^2Qu_m,...)+M \in J(H)^\bot \). Hence \(P(J(H)^\bot ) \subset J(H)^\bot \).