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

mm-isometric composition operators on discrete spaces

Michał Buchała 0000-0001-5272-9600 Doctoral School of Exact and Natural Sciences, Jagiellonian University, Łojasiewicza 11, PL-30348 Kraków, Poland Institute of Mathematics, Jagiellonian University, Łojasiewicza 6, PL-30348 Kraków, Poland [email protected]
Abstract.

In this paper we study composition operators on discrete spaces L2(μ)L^{2}(\mu). We establish the classification of underlying graphs of such operators. We obtain the characterization of mm-isometric composition operators such that the underlying graph has one cycle. The paper contains also the solution of the Cauchy dual subnormality problem in the class of such composition operators.

Key words and phrases:
weighted shifts on graphs, m-isometries, composition operators, completely hyperexpansive operators, Cauchy dual subnormality problem
1991 Mathematics Subject Classification:
Primary 47B33; Secondary 47B20, 47B37

1. Introduction

The composition operators form the important class of operators. The study of such a class of operators is motivated by the Banach-Stone theorem (see [9]), which states that a linear surjective isometry between spaces of continuous functions on compact spaces is in fact a weighted composition operators. The composition operators appear also naturally in ergodic theory (so-called Koopman operator, which is a composition operator induced by measure-preserving transformation). The class of composition operators has been studied extensively for a long time (see e.g. [22], [27]). In [12] Budzyński et.al. presented an example of unbounded non-hyponormal composition operators having graph with one cycle, which generates the Stieltjes moment sequence. In [24] Pietrzycki obtained an analytic model for a weighted composition operator on discrete space under the assumption that the underlying graph of this operator has finite branching index.

mm-isometric operators were introduced by Agler in [2] and investigated in details in [3] (as well as its subsequent parts). From that time on there have been published many papers devoted to study properties of mm-isometries (see e.g. [11], [8], [29], [17], [19]). In [1] and [10] mm-isometric weighted shifts were studied and, eventually, the characterization of weights of mm-isometric shift by real monic polynomials of degree at most m1m-1 was obtained (see [1, Theorem 1]). In [18] Jabłoński and Kośmider investigated mm-isometricity of composition operator on discrete space such that the graph associated to this operator has one cycle and one branching point.

The Cauchy dual T=T(TT)1T^{\prime}=T(T^{\ast}T)^{-1} of a left-invertible operator T𝐁(H)T\in\mathbf{B}(H) was introduced by Shimorin in [26] on the occasion of his study of Wold-type decompositions. From [7, Proposition 6] it follows that the Cauchy dual of a completely hyperexpansive weighted shift is always a subnormal contraction. In [13] there was posed a question whether it is true for every completely hyperexpansive operator. In [5] Chavan et. al. were considering this problem in the more restricted class of 2-isometries and obtained a negative answer. They also provided several sufficient conditions for 2-isometry to have subnormal Cauchy dual. The Cauchy dual subnormality problem was studied also in [6] and [14].

The present paper is devoted to study mm-isometric composition operators on discrete spaces in a more general setting than in [18]. It is organized as follows. In Section 2 we introduce the notation and recall basic definitions and some standard results needed in further parts of the paper. The aim of Section 3 is to fully classify self-maps of a countable set. If we associate a certain graph to such a map, it turns out that if this graph is assumed to be connected, then there are only two posibilities for its geometry: either this graph is a rootless directed tree or it consists of rooted directed trees connected by a cycle (see Theorem 3.5). In Section 5 we investigate the composition operators (viewed as certain weighted shifts) connected to graphs of the second type, that is, the ones having a cycle. We generalize the results presented in [18], where the very particular case of graph with a cycle and the only one branching point was studied. The assumption that the graph has a cycle allows us to obtain certain recursive formulas (see Lemma 4.4), which are key tools in further considerations. The main result of Section 5 is Theorem 5.5, which gives us the characterization of mm-isometricity in terms of the rank of certain matrices. We also prove that the notions of complete hyperexpansivity and 2-isometricity coincide in the class of composition operators having graphs with one cycle. Section 6 is devoted to study the Cauchy dual subnormality problem for 2-isometric composition operators on graphs with one cycle. In Theorem 6.2 the surprising characterization of composition operators having subnormal Cauchy dual is presented. It turns out that if the Cauchy dual is subnormal, then for every vertex vv lying on the cycle the measure representing the sequence (Tnev2)n=0(\lVert T^{n}e_{v}\rVert^{2})_{n=0}^{\infty} is two-atomic. Using this result, we provide an example of a composition operator on graph with one cycle, for which the Cauchy dual is not subnormal.

2. Preliminaries

Denote by \mathbb{N} and \mathbb{Z} the set of non-negative integers and integers, respectively and by \mathbb{R} and \mathbb{C} the field of real and complex numbers, respectively. For pp\in\mathbb{N} we set

p={n:np}.\mathbb{N}_{p}=\{n\in\mathbb{N}\!:n\geqslant p\}.

If 𝕂{,}\mathbb{K}\in\{\mathbb{R},\mathbb{C}\}, then 𝕂n[x]\mathbb{K}_{n}[x] stands for the set of all polynomials of degree at most nn with coefficients in 𝕂\mathbb{K}. From the binomial formula we can derive the following simple fact.

Observation 2.1.

If p𝕂n[x]p\in\mathbb{K}_{n}[x] for some n1n\in\mathbb{N}_{1} and κ𝕂\kappa\in\mathbb{K}, then

p()p(+κ)𝕂n1[x].p(\,\cdot\,)-p(\,\cdot\,+\kappa)\in\mathbb{K}_{n-1}[x].

The next lemma will play a crucial role in our considerations.

Lemma 2.2.

Let m1m\in\mathbb{N}_{1}. Suppose (pi)i=0𝕂m[x](p_{i})_{i=0}^{\infty}\subset\mathbb{K}_{m}[x] and for every ii\in\mathbb{N} write pi(x)=k=0mak(i)xkp_{i}(x)=\sum_{k=0}^{m}a_{k}^{(i)}x^{k}. Assume that for every k[0,m]k\in\mathbb{N}\cap[0,m] the series

bk:=i=0ak(i)b_{k}:=\sum_{i=0}^{\infty}a_{k}^{(i)}

is absolutely convergent. Then the series

(2.1) p(x):=i=0pi(x)p(x):=\sum_{i=0}^{\infty}p_{i}(x)

is absolutely convergent for every x𝕂x\in\mathbb{K} and p𝕂m[x]p\in\mathbb{K}_{m}[x]. Moreover,

p(x)=i=0mbkxk,x.p(x)=\sum_{i=0}^{m}b_{k}x^{k},\qquad x\in\mathbb{R}.
Proof.

If x𝕂x\in\mathbb{K} and nn\in\mathbb{N}, then

i=0n|pi(x)|k=0m|x|ki=0n|ak(i)|.\displaystyle\sum_{i=0}^{n}\lvert p_{i}(x)\rvert\leqslant\sum_{k=0}^{m}\lvert x\rvert^{k}\sum_{i=0}^{n}\lvert a_{k}^{(i)}\rvert.

Hence, the series (2.1) is absolutely convergent. Set q(x)=k=0mbkxkq(x)=\sum_{k=0}^{m}b_{k}x^{k}, x𝕂x\in\mathbb{K}. For every x𝕂x\in\mathbb{K} we have

i=0npi(x)=k=0mxki=0nak(i)nk=0mbkxk=q(x).\sum_{i=0}^{n}p_{i}(x)=\sum_{k=0}^{m}x^{k}\sum_{i=0}^{n}a_{k}^{(i)}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\sum_{k=0}^{m}b_{k}x^{k}=q(x).

Thus, p(x)=q(x)p(x)=q(x) for x𝕂x\in\mathbb{K}. Consequently, p𝕂m[x]p\in\mathbb{K}_{m}[x]. ∎

For a topological space XX we denote by (X)\mathcal{B}(X) the σ\sigma-algebra of Borel subsets of XX; if xXx\in X, then by δx\delta_{x} we denote the Dirac delta measure. For AXA\subset X, χA\chi_{A} stands for characteristic function of AA; we abbreviate χx=χ{x}\chi_{x}=\chi_{\left\{x\right\}} for xXx\in X. If HH is a complex Hilbert space, then 𝐁(H)\mathbf{B}(H) stands for the CC^{\ast}-algebra of all linear and bounded operators on HH.

By a discrete measure space we mean a pair (X,μ)(X,\mu), where XX is an at most countable set and μ:2X[0,]\mu\!:2^{X}\to[0,\infty] is a positive measure on XX such that μ({x})(0,)\mu(\{x\})\in(0,\infty) for every xXx\in X; for simplicity we write μ(x):=μ({x})\mu(x):=\mu(\{x\}) for xXx\in X. If (X,μ)(X,\mu) is a discrete measure space and T:XXT\!:X\to X is a transformation on XX, then μT1\mu\circ T^{-1} is absolutely continuous with respect to μ\mu, where μT1:2X[0,]\mu\circ T^{-1}\!:2^{X}\to[0,\infty] is a positive measure on XX defined by

μT1(A)=μ(T1(A)),AX.\mu\circ T^{-1}(A)=\mu(T^{-1}(A)),\qquad A\subset X.

Denote by hTh_{T} the Radon-Nikodym derivative of μT1\mu\circ T^{-1} with respect to μ\mu; we can easily compute that

(2.2) hT(x)=μ(T1({x}))μ(x),xX.h_{T}(x)=\frac{\mu(T^{-1}(\{x\}))}{\mu(x)},\qquad x\in X.

Assuming additionally that hTL(μ)h_{T}\in L^{\infty}(\mu), we define the composition operator CT𝐁(L2(μ))C_{T}\in\mathbf{B}(L^{2}(\mu)) as follows:

CTf=fT,fL2(μ).C_{T}f=f\circ T,\qquad f\in L^{2}(\mu).

For more details about general theory of composition operators we refer to [22].

We say that S𝐁(H)S\in\mathbf{B}(H) is normal if SS=SSS^{\ast}S=SS^{\ast}; SS is called subnormal if there exists a Hilbert space KK containing (an isometric embedding of) HH and a normal operator N𝐁(K)N\in\mathbf{B}(K) such that N|H=SN|_{H}=S. Recall that S𝐁(H)S\in\mathbf{B}(H) is completely hyperexpansive if

k=0n(1)k(nk)SkSk0,n1.\sum_{k=0}^{n}(-1)^{k}\binom{n}{k}S^{\ast k}S^{k}\leqslant 0,\qquad n\in\mathbb{N}_{1}.

We say that S𝐁(H)S\in\mathbf{B}(H) is mm-isometric (m1m\in\mathbb{N}_{1}), if

k=0m(1)mk(mk)SkSk=0.\sum_{k=0}^{m}(-1)^{m-k}\binom{m}{k}S^{\ast k}S^{k}=0.

For an at most countable set VV we define

2(V)={f:VV|vV|f(v)|2<};\ell^{2}(V)=\left\{f\!:V\to V|\sum_{v\in V}\lvert f(v)\rvert^{2}<\infty\right\};

this is a Hilbert space with the norm given by:

f=vV|f(v)|2,f2(V).\lVert f\rVert=\sqrt{\sum_{v\in V}\lvert f(v)\rvert^{2}},\qquad f\in\ell^{2}(V).

The vectors eve_{v} (vVv\in V) defined as follows:

ev(u)={1, if u=v0, otherwise,e_{v}(u)=\begin{cases}1,&\text{ if }u=v\\ 0,&\text{ otherwise}\end{cases},

form an orthonormal basis of 2(V)\ell^{2}(V).

In the subsequent part of this section we introduce some basic notions of graph theory (for more details see [23] and [4]). By a directed graph we mean a pair G=(V,E)G=(V,E), where VV is a non-empty at most countable set and EV×VE\subset V\times V; elements of VV are called vertices and elements of EE are called edges. Every pair G1=(V1,E1)G_{1}=(V_{1},E_{1}) such that V1VV_{1}\subset V and E1(V1×V1)EE_{1}\subset(V_{1}\times V_{1})\cap E is called a subgraph of GG. For a vertex vVv\in V we define its in-degree, out-degree and degree as follows:

deginv\displaystyle\mathrm{deg}\,\mathrm{in}\,v =#{wV:(w,v)E}\displaystyle=\#\{w\in V\!:(w,v)\in E\}
degoutv\displaystyle\mathrm{deg}\,\mathrm{out}\,v =#{wV:(v,w)E}\displaystyle=\#\{w\in V\!:(v,w)\in E\}
degv\displaystyle\deg v =deginv+degoutv.\displaystyle=\mathrm{deg}\,\mathrm{in}\,v+\mathrm{deg}\,\mathrm{out}\,v.

If v,wVv,w\in V, then a path from vv to ww in GG is a sequence (v0,,vk)V(v_{0},\ldots,v_{k})\subset V such that v0=vv_{0}=v, vk=wv_{k}=w and (vi,vi+1)E(v_{i},v_{i+1})\in E for every i[0,k1]i\in\mathbb{N}\cap[0,k-1]. A path (v0,,vk)(v_{0},\ldots,v_{k}) is called a cycle if v0=vkv_{0}=v_{k}; the cycle is said to be simple if the vertices vjv_{j} (j[0,k1]j\in\mathbb{N}\cap[0,k-1]) are distinct.

Graph GG is said to be connected if for every two distinct vertices v,wVv,w\in V there exists a sequence (v0,,vk)V(v_{0},\ldots,v_{k})\subset V such that v0=vv_{0}=v, vk=wv_{k}=w and {vi,vi+1}E~\{v_{i},v_{i+1}\}\in\tilde{E} for i[0,k1]i\in\mathbb{N}\cap[0,k-1], where

E~={{v,w}V:(v,w)E or (w,v)E}.\tilde{E}=\{\{v,w\}\subset V\!:(v,w)\in E\text{ or }(w,v)\in E\}.

If G=(V,E)G=(V,E) is a directed graph, then its subgraph G1=(V1,E1)G_{1}=(V_{1},E_{1}) is called a connected component if G1G_{1} is connected and for each connected subgraph G1=(V1,E1)G_{1}^{\prime}=(V_{1}^{\prime},E_{1}^{\prime}) of GG satisfying V1V1V_{1}\subset V_{1}^{\prime} and E1E1E_{1}\subset E_{1}^{\prime} we have G1=G1G_{1}=G_{1}^{\prime}. In other words, connected components of GG are maximal connected subgraphs of GG.

A connected graph G=(V,E)G=(V,E) is called a directed tree if GG has no cycles and for every vVv\in V we have deginv1\mathrm{deg}\,\mathrm{in}\,v\leqslant 1. If GG is a directed tree, then there is at most one vertex ωV\omega\in V such that deginω=0\mathrm{deg}\,\mathrm{in}\,\omega=0; if it exists, we call this vertex the root of GG (see [16, Proposition 2.1.1]). The reader can verify that for rooted directed trees the following induction principle holds.

Lemma 2.3.

Let G=(V,E)G=(V,E) be a directed tree with root ω\omega. Let VVV^{\prime}\subset V satisfy the following conditions:

  1. (i)

    ωV\omega\in V^{\prime},

  2. (ii)

    for every vVv\in V^{\prime} we have {uV:(v,u)E}V\{u\in V\!:(v,u)\in E\}\subset V^{\prime}.

Then V=VV^{\prime}=V.

We say that a graph GG is strongly connected if for any two distinct vertices v,wVv,w\in V there is a path from vv to ww and a path from ww to vv (in other words, vv and ww lie on a common cycle). For a directed graph G=(V,E)G=(V,E) we define a relation 𝒮𝒞GV×V\mathcal{SC}_{G}\subset V\times V in the following way: (v,w)𝒮𝒞G(v,w)\in\mathcal{SC}_{G} if and only if either v=wv=w or there exists a path from vv to ww in GG and a path from ww to vv in GG. It is a matter of routine to verify that 𝒮𝒞G\mathcal{SC}_{G} is an equivalence relation on VV. Each equivalence class CVC\subset V of 𝒮𝒞G\mathcal{SC}_{G} induces a subgraph GC=(C,EC)G_{C}=(C,E_{C}), where EC=E(C×C)E_{C}=E\cap(C\times C), which is called a strongly connected component of GG.

3. Classification of graphs of self-maps of countable sets

In [15, Lemma 4.3.1] there was proved that weighted shifts on rootless directed trees with positive weights are unitarily equivalent to composition operators on certain discrete spaces (see also [12, Lemma 3.1.4]). In [12, Theorem 3.2.1] the authors classified composition operators on discrete spaces, for which the associated graph is connected and has at most one vertex of out-degree greater than 1. In [28, Proposition 2.4] composition operators on graphs with the property that every vertex has out-degree 1 are characterized. In this section, using graph-theoretical approach, we give the full classification of graphs of self-maps of at most countable set. This classification was mentioned in [24, p.898]; however, the proof was not provided there.

Suppose XX is an at most countable set and T:XXT\!:X\to X is a transformation on X; for xXx\in X we abbreviate: Tx=T(x)Tx=T(x). Consider the graph GT=(VT,ET)G_{T}=(V_{T},E_{T}), where VT=XV_{T}=X and ET=T1={(Tx,x):xX}E_{T}=T^{-1}=\left\{(Tx,x)\!:x\in X\right\}. In the next few results we investigate the geometry of graph GTG_{T}.

Observation 3.1.

Suppose XX is an at most countable set and T:XXT\!:X\to X. Then, for every vVTv\in V_{T} we have deginv=1\mathrm{deg}\,\mathrm{in}\,v=1. Moreover, if for v,wVTv,w\in V_{T}, vwv\not=w, there exists a path from vv to ww, then there is the unique path from vv to ww consisting of distinct vertices.

Proof.

The first part can be easily derived from the definition of GTG_{T}. Let us prove the ”moreover” part. Obviously, we can always find a path from vv to ww consisting of distinct vertices, so we only have to prove uniqueness. Let v=v0,v1,,vn,vn+1=wv=v_{0},v_{1},\ldots,v_{n},v_{n+1}=w and v=u0,u1,u2,,um,um+1=wv=u_{0},u_{1},u_{2},\ldots,u_{m},u_{m+1}=w be paths connecting vv and ww such that vjvkv_{j}\not=v_{k} for every j,k[0,n+1]j,k\in\mathbb{N}\cap[0,n+1], jkj\not=k and ujuku_{j}\not=u_{k}, j,k[0,m+1]j,k\in\mathbb{N}\cap[0,m+1] for every jkj\not=k; without loss of generality we can assume nmn\leqslant m. Since deginw=1\mathrm{deg}\,\mathrm{in}\,w=1, we have that vn=umv_{n}=u_{m}. Proceeding by induction, we can show that vnk=umkv_{n-k}=u_{m-k} for all k[0,n]k\in\mathbb{N}\cap[0,n]. In particular, we have v0=umnv_{0}=u_{m-n}. This, by the choice of u0,,umu_{0},\ldots,u_{m}, implies that m=nm=n and vj=ujv_{j}=u_{j} for all j[0,n]j\in\mathbb{N}\cap[0,n]. ∎

The proceeding results reveal the structure of strongly connected components of GTG_{T}. Lemma 3.2 shows that every connected component of GTG_{T} may contain at most one non-trivial strongly connected component.

Lemma 3.2.

Suppose XX is an at most countable set and T:XXT\!:X\to X is such that GTG_{T} is connected. Then GTG_{T} contains at most one strongly connected component G0G_{0} with non-empty set of edges (that is, G0({v},)G_{0}\not=(\{v\},\varnothing), where vVTv\in V_{T}).

Proof.

Suppose to the contrary that G0=(V0,E0)G_{0}=(V_{0},E_{0}) and G1=(V1,E1)G_{1}=(V_{1},E_{1}) are two different strongly connected components of GTG_{T} with non-empty sets of edges. Obviously, V0V1=V_{0}\cap V_{1}=\varnothing. Let v0V0v_{0}\in V_{0} and v1V1v_{1}\in V_{1}. We will show that there exists a path from v0v_{0} to v1v_{1}. Since GTG_{T} is connected, there exists a sequence (u0,,un+1)VT(u_{0},\ldots,u_{n+1})\subset V_{T} (nn\in\mathbb{N}) such that u0=v0u_{0}=v_{0}, un+1=v1u_{n+1}=v_{1} and {uk,uk+1}E~T\{u_{k},u_{k+1}\}\in\tilde{E}_{T} for every k[0,n]k\in\mathbb{N}\cap[0,n]. If (u1,v0)ET(u_{1},v_{0})\in E_{T}, then from the fact that G0G_{0} is strongly connected, E0E_{0}\not=\varnothing and deginv0=1\mathrm{deg}\,\mathrm{in}\,v_{0}=1 we obtain that u1V0u_{1}\in V_{0}. This implies that there exists a path from v0v_{0} to u1u_{1} in GTG_{T}. Hence, possibly after extending a sequence (u0,,un+1)(u_{0},\ldots,u_{n+1}), without loss of generality we can assume that (u0,u1)ET(u_{0},u_{1})\in E_{T}. If (uk,uk+1)ET(u_{k},u_{k+1})\in E_{T} for every k[0,n]k\in\mathbb{N}\cap[0,n], then the sequence (u0,,un+1)(u_{0},\ldots,u_{n+1}) is the path from v0v_{0} to v1v_{1}. Suppose that it is not the case; in particular, n1n\geqslant 1. Set

k0=min{k[1,n]:(uk+1,uk)ET}.k_{0}=\min\{k\in\mathbb{N}\cap[1,n]\!:(u_{k+1},u_{k})\in E_{T}\}.

Since (uk0+1,uk0)ET(u_{k_{0}+1,}u_{k_{0}})\in E_{T}, (uk01,uk0)ET(u_{k_{0}-1},u_{k_{0}})\in E_{T} and deginuk0=1\mathrm{deg}\,\mathrm{in}\,u_{k_{0}}=1, it follows that uk0+1=uk01u_{k_{0}+1}=u_{k_{0}-1}. If k0=nk_{0}=n, then we have un1=un+1=v1u_{n-1}=u_{n+1}=v_{1}. Thus, (u0,,un1)(u_{0},\ldots,u_{n-1}) is the path from v0v_{0} to v1v_{1}. If k0<nk_{0}<n, then the sequence

(u0,,un)=(u0,,uk01,uk0+2,,un+1)(u_{0}^{\prime},\ldots,u_{n}^{\prime})=(u_{0},\ldots,u_{k_{0}-1},u_{k_{0}+2},\ldots,u_{n+1})

have the property that {uk,uk+1}E~T\{u_{k}^{\prime},u_{k+1}^{\prime}\}\in\tilde{E}_{T} for every k[0,n1]k\in\mathbb{N}\cap[0,n-1], so the above procedure can be applied to this sequence. Proceeding this way, we eventually obtain the path from v0v_{0} to v1v_{1}. By symmetry, there exists also a path from v1v_{1} to v0v_{0}. Since G0G_{0} and G1G_{1} are strongly connected, we have v1V0v_{1}\in V_{0} and v0V1v_{0}\in V_{1}, which is a contradiction. ∎

The next result shows that non-trivial strongly connected components of GTG_{T} are of very special type.

Lemma 3.3.

Suppose XX is an at most countable set and T:XXT\!:X\to X is such that GTG_{T} is connected and has a unique strongly connected component G0=(V0,E0)G_{0}=(V_{0},E_{0}) with non-empty set of edges. Then G0G_{0} is a simple cycle.

Proof.

If #V0=1\#V_{0}=1, then the conclusion obviously holds. Assume #V02\#V_{0}\geqslant 2. Let v,wV0v,w\in V_{0}, vwv\not=w. By the strong connectedness of G0G_{0}, vv and ww lie on the common cycle; we will prove that this cycle is actually the whole graph G0G_{0}. By Observation 3.1 there exist paths from vv to ww and ww to vv consisting of distinct vertices; call by v=v0,v1,,vn,vn+1=wv=v_{0},v_{1},\ldots,v_{n},v_{n+1}=w the vertices of such a path from vv to ww and by w=w0,w1,,wm,wm+1=vw=w_{0},w_{1},\ldots,w_{m},w_{m+1}=v the vertices of such a path from ww to vv. Set V={v,v1,,vn,w,w1,,wm}V0V^{\prime}=\{v,v_{1},\ldots,v_{n},w,w_{1},\ldots,w_{m}\}\subset V_{0}. It is enough to show that V=V0V^{\prime}=V_{0}. Suppose to the contrary that it is not the case and let uV0Vu\in V_{0}\setminus V^{\prime}. Let u=u0,u1,,u,u+1=wu=u_{0},u_{1},\ldots,u_{\ell},u_{\ell+1}=w be the path connecting uu and ww and consisting of distinct vertices. If n\ell\leqslant n, then the reasoning as in the proof of Observation 3.1 gives us that ui=vi+nu_{i}=v_{i+n-\ell} for i[0,]i\in\mathbb{N}\cap[0,\ell], which implies that u=u0=vnV0u=u_{0}=v_{n-\ell}\in V_{0}. If >n\ell>n, then the same reasoning gives us that v=unv=u_{\ell-n}. Since it necessarily holds that nm\ell-n\leqslant m, we have u=u0=wm+1+nV0u=u_{0}=w_{m+1+n-\ell}\in V_{0}. In both cases we get a contradiction. ∎

The following lemma says that after removing non-trivial strongly connected component from GTG_{T} we actually obtain a family of directed trees.

Lemma 3.4.

Suppose XX is an at most countable set and T:XXT\!:X\to X is such that GTG_{T} is connected and has a unique strongly connected component G0=(V0,E0)G_{0}=(V_{0},E_{0}) with non-empty set of edges. Set G=(VTV0,E)G^{\prime}=(V_{T}\setminus V_{0},E^{\prime}), where E={(u,v)ET:u,vVTV0}E^{\prime}=\{(u,v)\in E_{T}\!:u,v\in V_{T}\setminus V_{0}\}. If G0=(V0,E0)G_{0}^{\prime}=(V_{0}^{\prime},E_{0}^{\prime}) is a connected component of GG^{\prime}, then G0G_{0}^{\prime} is a rooted directed tree.

Proof.

From Lemma 3.2 it follows that G0G_{0}^{\prime} has no cycles. This, together with Observation 3.1, implies that G0G_{0}^{\prime} is a directed tree. It remains to prove that G0G_{0}^{\prime} has a root. Let vV0v\in V_{0} and wV0w\in V_{0}^{\prime}. Proceeding as in the proof of Lemma 3.2 we obtain that there exists a path from vv to ww in GTG_{T}. Let v=v0,v1,,vm,vm+1=wv=v_{0},v_{1},\ldots,v_{m},v_{m+1}=w be the path from vv to ww consisting of distinct vertices. Set M=min{k[0,m+1]:vkV0}M=\min\{k\in\mathbb{N}\cap[0,m+1]\!:v_{k}\notin V_{0}\}. From the fact that deginvM=1\mathrm{deg}\,\mathrm{in}\,v_{M}=1 and that (vM1,vM)ET(v_{M-1},v_{M})\in E_{T} is the only edge coming to vMv_{M} it follows that vMv_{M} has in-degree 0 in G0G_{0}^{\prime}. This implies that vMv_{M} is the root of G0G_{0}^{\prime}. ∎

Now we are in the position to state the main results of this section, which classifies graphs of self-maps T:XXT\!:X\to X.

Theorem 3.5.

Suppose XX is an at most countable set and T:XXT\!:X\to X is such that GTG_{T} is connected. Then exactly one of the following holds:

  1. (i)

    GTG_{T} is a rootless directed tree,

  2. (ii)

    GTG_{T} satisfies the conditions below:

    1. (a)

      there exist κ1\kappa\in\mathbb{N}_{1} and distinct vertices v0,vκ1VTv_{0}\ldots,v_{\kappa-1}\in V_{T} such that (vj,vj+1)ET(v_{j},v_{j+1})\in E_{T} for j[0,κ2]j\in\mathbb{N}\cap[0,\kappa-2] and (vκ1,v0)E(v_{\kappa-1},v_{0})\in E

    2. (b)

      there exists a partition \mathcal{R} of VT{v0,,vκ1}V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\} such that for every RR\in\mathcal{R} the graph (R,ER)(R,E_{R}), where ER={(u,w)ET:u,wR}E_{R}=\{(u,w)\in E_{T}\!:u,w\in R\}, is a rooted directed tree with the root ωR\omega_{R} having the property that there exists exactly one iR[0,κ1]i_{R}\in\mathbb{N}\cap[0,\kappa-1] satisfying (viR,ωR)ET(v_{i_{R}},\omega_{R})\in E_{T}.

Proof.

First, suppose GTG_{T} has no non-trivial strongly connected components. This implies that GTG_{T} has no cycles. By Observation 3.1, GTG_{T} is a rootless directed tree, so in this case (i) holds. Assuming now that GTG_{T} has a non-trivial strongly connected component, by calling Lemmata 3.3 and 3.4, we derive that GTG_{T} satisfies the conditions in (ii). ∎

In Figure 1 we present the example of graph GTG_{T} satisfying Theorem 3.5.(ii); for the convenience, we denote {Ri,k}k=1Ni={R:iR=i}\{R_{i,k}\}_{k=1}^{N_{i}}=\{R\in\mathcal{R}\!:i_{R}=i\} for i[0,κ1]i\in\mathbb{N}\cap[0,\kappa-1], where NiN_{i} is the cardinality of the set on the right hand side.

Refer to caption
Figure 1. Graph GTG_{T} satisfying Theorem 3.5.(ii)

4. Composition operators as weighted shifts on directed graphs

In this section we present how to transform a composition operator on discrete space into a weighted shift on a directed graph; we prove several routine results about this weighted shift. Suppose that (X,μ)(X,\mu) is a discrete measure space and T:XXT\!:X\to X is a transformation on XX such that hTL(μ)h_{T}\in L^{\infty}(\mu). Define the family of weights 𝝀T=(λv)vVT\boldsymbol{\lambda}_{T}=(\lambda_{v})_{v\in V_{T}} by

(4.1) λv=μ(v)μ(Tv),vVT,\lambda_{v}=\frac{\sqrt{\mu(v)}}{\sqrt{\mu(Tv)}},\qquad v\in V_{T},

From (2.2) it can be easily derived that

supuVvVTTv=uλv2<hTL(μ).\sup_{u\in V}\sum_{\begin{subarray}{c}v\in V_{T}\\ Tv=u\end{subarray}}\lambda_{v}^{2}<\infty\iff h_{T}\in L^{\infty}(\mu).

Since we assume that hTL(μ)h_{T}\in L^{\infty}(\mu), the above allows us to define the operator S𝝀T𝐁(2(VT))S_{\boldsymbol{\lambda}_{T}}\in\mathbf{B}(\ell^{2}(V_{T})) as follows:

S𝝀Teu=vVTTv=uλvev,uVT.S_{\boldsymbol{\lambda}_{T}}e_{u}=\sum_{\begin{subarray}{c}v\in V_{T}\\ Tv=u\end{subarray}}\lambda_{v}e_{v},\qquad u\in V_{T}.

It is a matter of routine to prove the following lemma.

Lemma 4.1.

Let (X,μ)(X,\mu) be a discrete measure space. Suppose T:XXT\!:X\to X is such that hTL(μ)h_{T}\in L^{\infty}(\mu). Then the operator U:2(VT)L2(μ)U\!:\ell^{2}(V_{T})\to L^{2}(\mu) given by

Uev=1μ(v)χv,vVT=X,Ue_{v}=\frac{1}{\sqrt{\mu(v)}}\chi_{v},\qquad v\in V_{T}=X,

makes CTC_{T} and S𝛌TS_{\boldsymbol{\lambda}_{T}} unitarily equivalent.

Proof.

First, we show that UU is unitary. For v,wVTv,w\in V_{T} we have

Uev,Uew=1μ(v)χv,1μ(w)χw=1μ(v)μ(w)Xχvχwdμ.\langle Ue_{v},Ue_{w}\rangle=\left\langle\frac{1}{\sqrt{\mu(v)}}\chi_{v},\frac{1}{\sqrt{\mu(w)}}\chi_{w}\right\rangle=\frac{1}{\sqrt{\mu(v)}\cdot\sqrt{\mu(w)}}\int_{X}\chi_{v}\cdot\chi_{w}\,\mathrm{d}\mu.

If vwv\not=w, then Xχvχwdμ=0\int_{X}\chi_{v}\cdot\chi_{w}\,\mathrm{d}\mu=0; if v=wv=w, then

1μ(v)μ(w)Xχvχwdμ=1μ(v)μ(v)=1.\frac{1}{\sqrt{\mu(v)}\cdot\sqrt{\mu(w)}}\int_{X}\chi_{v}\cdot\chi_{w}\,\mathrm{d}\mu=\frac{1}{\mu(v)}\cdot\mu(v)=1.

Hence,

Uev,Uew={0, if vw1, if v=w,\langle Ue_{v},Ue_{w}\rangle=\begin{cases}0,&\text{ if }v\not=w\\ 1,&\text{ if }v=w\end{cases},

which implies that Uev,Uew=ev,ew\langle Ue_{v},Ue_{w}\rangle=\langle e_{v},e_{w}\rangle. Since {ev}vVT\{e_{v}\}_{v\in V_{T}} is the orthonormal basis of 2(VT)\ell^{2}(V_{T}), it follows that Uf,Ug=f,g\langle Uf,Ug\rangle=\langle f,g\rangle for f,g2(VT)f,g\in\ell^{2}(V_{T}), that is, UU is isometry. In turn, Lin{χv:vVT}¯=L2(μ)\overline{\mathrm{Lin}\,\{\chi_{v}\!:v\in V_{T}\}}=L^{2}(\mu), so U(2(VT))=L2(μ)U(\ell^{2}(V_{T}))=L^{2}(\mu). Therefore, UU is unitary. Next, if vVTv\in V_{T}, then

CTUev=1μ(v)χvT=1μ(v)χT1({v})C_{T}Ue_{v}=\frac{1}{\sqrt{\mu(v)}}\cdot\chi_{v}\circ T=\frac{1}{\sqrt{\mu(v)}}\chi_{T^{-1}(\{v\})}

and

US𝝀Tev\displaystyle US_{\boldsymbol{\lambda}_{T}}e_{v} =UuVTTu=vλueu\displaystyle=U\sum_{\begin{subarray}{c}u\in V_{T}\\ Tu=v\end{subarray}}\lambda_{u}e_{u}
=uVTTu=vμ(u)μ(v)1μ(u)χu\displaystyle=\sum_{\begin{subarray}{c}u\in V_{T}\\ Tu=v\end{subarray}}\frac{\sqrt{\mu(u)}}{\sqrt{\mu(v)}}\cdot\frac{1}{\sqrt{\mu(u)}}\chi_{u}
=1μ(v)uVTTu=vχu=1μ(v)χT1({v}).\displaystyle=\frac{1}{\sqrt{\mu(v)}}\sum_{\begin{subarray}{c}u\in V_{T}\\ Tu=v\end{subarray}}\chi_{u}=\frac{1}{\sqrt{\mu(v)}}\chi_{T^{-1}(\{v\})}.

Hence, CTU=US𝝀TC_{T}U=US_{\boldsymbol{\lambda}_{T}}, what completes the proof. ∎

The lemma below states that connected components of GTG_{T} induce reducing subspaces of S𝝀TS_{\boldsymbol{\lambda}_{T}}, so we can restrict our investigation to the case when the graph GTG_{T} is connected.

Lemma 4.2.

Suppose (X,μ)(X,\mu) is a discrete measure space and T:XXT\!:X\to X is a transformation on XX such that hTL(μ)h_{T}\in L^{\infty}(\mu). Let G0=(V0,E0)G_{0}=(V_{0},E_{0}) be a connected component of GTG_{T}. Then the space M=Lin{ev:vV0}¯M=\overline{\mathrm{Lin}\,\{e_{v}\!:v\in V_{0}\}} is a reducing subspace for S𝛌TS_{\boldsymbol{\lambda}_{T}}.

Proof.

Let vV0v\in V_{0}. Since G0G_{0} is a connected component of GTG_{T} we have that for every uVTu\in V_{T} satisfying (v,u)ET(v,u)\in E_{T}, uV0u\in V_{0} and (v,u)E0(v,u)\in E_{0}. Hence, S𝝀TevMS_{\boldsymbol{\lambda}_{T}}e_{v}\in M, which implies that S𝝀TMMS_{\boldsymbol{\lambda}_{T}}M\subset M. Next, we will prove that MM^{\perp} is also invariant for S𝝀TS_{\boldsymbol{\lambda}_{T}}. It can be easily seen that (u,v),(v,u)E(u,v),(v,u)\notin E for every vV0v\in V_{0} and every uVTV0u\in V_{T}\setminus V_{0}. From this it follows that S𝝀TeuMS_{\boldsymbol{\lambda}_{T}}e_{u}\in M^{\perp} for every uVTV0u\in V_{T}\setminus{V_{0}}. This implies that S𝝀TMMS_{\boldsymbol{\lambda}_{T}}M^{\perp}\subset M^{\perp}, what completes the proof. ∎

It turns out that that the operator S𝝀S_{\boldsymbol{\lambda}} preserves the orthogonality of the standard orthonormal basis of 2(VT)\ell^{2}(V_{T}).

Observation 4.3.

Suppose (X,μ)(X,\mu) is a discrete measure space and T:XXT\!:X\to X is a transformation on XX such that hTL(μ)h_{T}\in L^{\infty}(\mu) and that GTG_{T} is connected. Then S𝛌TeuS𝛌TevS_{\boldsymbol{\lambda}_{T}}e_{u}\perp S_{\boldsymbol{\lambda}_{T}}e_{v} for every vertices u,vVTu,v\in V_{T}, vuv\not=u.

Proof.

Let v,uVTv,u\in V_{T} be distinct vertices. By Observation 3.1, we have

{wVT:(v,w)ET}{wVT:(u,w)ET}=.\{w\in V_{T}\!:(v,w)\in E_{T}\}\cap\{w\in V_{T}\!:(u,w)\in E_{T}\}=\varnothing.

Since ew1ew2e_{w_{1}}\perp e_{w_{2}} for every two distinct vertices w1,w2VTw_{1},w_{2}\in V_{T}, we deduce from the above that

S𝝀Teu,S𝝀Tev=w1VT(u,w1)ETw2VT(v,w2)ETλw1λw2¯ew1,ew2=0.\langle S_{\boldsymbol{\lambda}_{T}}e_{u},S_{\boldsymbol{\lambda}_{T}}e_{v}\rangle=\sum_{\begin{subarray}{c}w_{1}\in V_{T}\\ (u,w_{1})\in E_{T}\end{subarray}}\sum_{\begin{subarray}{c}w_{2}\in V_{T}\\ (v,w_{2})\in E_{T}\end{subarray}}\lambda_{w_{1}}\overline{\lambda_{w_{2}}}\langle e_{w_{1}},e_{w_{2}}\rangle=0.

Hence, S𝝀TeuS𝝀TevS_{\boldsymbol{\lambda}_{T}}e_{u}\perp S_{\boldsymbol{\lambda}_{T}}e_{v}. ∎

Set111We stick to the convention that =1\prod\varnothing=1.

Λm,i:=j=m+1m+iλvjmodκ,m[0,κ1],i.\Lambda_{m,i}:=\prod_{j=m+1}^{m+i}\lambda_{v_{j\,\mathrm{mod}\,\kappa}},\qquad m\in\mathbb{N}\cap[0,\kappa-1],i\in\mathbb{N}.

For the further use we emphasize several formulas.

Lemma 4.4.

Suppose (X,μ)(X,\mu) is a discrete space and T:XXT\!:X\to X is such that hTL(μ)h_{T}\in L^{\infty}(\mu) and that GTG_{T} is connected and satisfies Theorem 3.5.(ii).

  1. (i)

    For every m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1], Λm,κ=1\Lambda_{m,\kappa}=1.

  2. (ii)

    For every kk\in\mathbb{N} and every m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1] we have

    (4.2) S𝝀Tkevm\displaystyle S_{\boldsymbol{\lambda}_{T}}^{k}e_{v_{m}} =Λm,kev(m+k)modκ+i=0k1Λm,iRiR=i+mmodκλωRS𝝀Tki1eωR;\displaystyle=\Lambda_{m,k}e_{v_{(m+k)\,\mathrm{mod}\,\kappa}}+\sum_{i=0}^{k-1}\Lambda_{m,i}\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}S_{\boldsymbol{\lambda}_{T}}^{k-i-1}e_{\omega_{R}};
  3. (iii)

    For every nn\in\mathbb{N} and every m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1] the following recursive formula is satisfied:

    (4.3) S𝝀Tn+κevm\displaystyle S_{\boldsymbol{\lambda}_{T}}^{n+\kappa}e_{v_{m}} =S𝝀Tnevm+i=0κ1Λm,iRiR=i+mmodκλωRS𝝀Tn+κi1eωR.\displaystyle=S_{\boldsymbol{\lambda}_{T}}^{n}e_{v_{m}}+\sum_{i=0}^{\kappa-1}\Lambda_{m,i}\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}S_{\boldsymbol{\lambda}_{T}}^{n+\kappa-i-1}e_{\omega_{R}}.
Proof.

(i). It follows from the definition of weights λi\lambda_{i}, i[0,κ1]i\in\mathbb{N}\cap[0,\kappa-1].
(ii). We prove it by induction. For k=0,1k=0,1 the equality (4.2) holds. If (4.2) holds for kk\in\mathbb{N}, then

S𝝀Tk+1evm\displaystyle S^{k+1}_{\boldsymbol{\lambda}_{T}}e_{v_{m}} =S𝝀T(Λm,kev(m+k)modκ+i=0k1Λm,iRiR=i+mmodκλωRS𝝀Tki1eωR)\displaystyle=S_{\boldsymbol{\lambda}_{T}}\left(\Lambda_{m,k}e_{v_{(m+k)\,\mathrm{mod}\,\kappa}}+\sum_{i=0}^{k-1}\Lambda_{m,i}\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}S_{\boldsymbol{\lambda}_{T}}^{k-i-1}e_{\omega_{R}}\right)
=Λm,k+1ev(m+k+1)modκ+Λm,kRiR=k+mmodκλωReωR\displaystyle=\Lambda_{m,k+1}e_{v_{(m+k+1)}\,\mathrm{mod}\,\kappa}+\Lambda_{m,k}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=k+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}e_{\omega_{R}}
+i=0k1Λm,iRiR=i+mmodκλωRS𝝀TkieωR\displaystyle\qquad\qquad\qquad\qquad\qquad+\sum_{i=0}^{k-1}\Lambda_{m,i}\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}S_{\boldsymbol{\lambda}_{T}}^{k-i}e_{\omega_{R}}
=Λm,k+1ev(m+k+1)modκ+i=0kΛm,iRiR=i+mmodκλωRS𝝀TkieωR.\displaystyle=\Lambda_{m,k+1}e_{v_{(m+k+1)\,\mathrm{mod}\,\kappa}}+\sum_{i=0}^{k}\Lambda_{m,i}\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}S_{\boldsymbol{\lambda}_{T}}^{k-i}e_{\omega_{R}}.

Hence, (4.2) holds for k+1k+1.
(iii). It is enough to apply the operator S𝝀TnS^{n}_{\boldsymbol{\lambda}_{T}} to both sides of (4.2) with k=κk=\kappa and use (i). ∎

5. mm-isometric composition operators: graphs with one cycle

In [18] the authors studied composition operators with the property that the associated graphs have one cycle and one branching point. In this paper we also limit our considerations to operators such that related graphs have one cycle, but with no further assumptions on degrees of vertices. From now on we fix a discrete space (X,μ)(X,\mu) and a self-map T:XXT\!:X\to X such that hTL(X)h_{T}\in L^{\infty}(X), so that CTC_{T} (and consequently S𝝀TS_{\boldsymbol{\lambda}_{T}}) is bounded. Since, by Observation 4.1, the composition operator CTC_{T} is unitarily equivalent to S𝝀TS_{\boldsymbol{\lambda}_{T}}, we investigate mm-isometricity of S𝝀TS_{\boldsymbol{\lambda}_{T}}. Whenever we assume that the associated graph GTG_{T} defined in Section 3 satisfies Theorem 3.5.(ii), we stick to the notation introduced there.

Let us start from the general result concerning mm-isometric composition operators on discrete spaces.

Lemma 5.1 (cf. [18, Proposition 2.5]).

Suppose T:XXT\!:X\to X is such that GTG_{T} is connected and such that hTL(μ)h_{T}\in L^{\infty}(\mu). For m1m\in\mathbb{N}_{1} the following conditions are equivalent:

  1. (i)

    S𝝀TS_{\boldsymbol{\lambda}_{T}} is mm-isometric,

  2. (ii)

    for every f2(VT)f\in\ell^{2}(V_{T}) there exists a polynomial pfm1[x]p_{f}\in\mathbb{R}_{m-1}[x] satisfying

    pf(n)=S𝝀Tnf2,n,p_{f}(n)=\lVert S_{\boldsymbol{\lambda}_{T}}^{n}f\rVert^{2},\qquad n\in\mathbb{N},
  3. (iii)

    for every vVTv\in V_{T} there exists a polynomial pvm1[x]p_{v}\in\mathbb{R}_{m-1}[x] satisfying

    (5.1) pv(n)=S𝝀Tnev2,n.p_{v}(n)=\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{v}\rVert^{2},\qquad n\in\mathbb{N}.

Before we state the proof we need the following observation.

Observation 5.2.

Suppose T:XXT\!:X\to X is such that GTG_{T} is connected and such that hTL(μ)h_{T}\in L^{\infty}(\mu). Let m1m\in\mathbb{N}_{1}. Let vVTv\in V_{T} and suppose that pv(x)=i=0m1aixim1[x]p_{v}(x)=\sum_{i=0}^{m-1}a_{i}x^{i}\in\mathbb{R}_{m-1}[x] is polynomial satisfying (5.1). Then there exists a constant C[0,)C\in[0,\infty) depending only on mm and S𝛌TS_{\boldsymbol{\lambda}_{T}} such that |ai|C\lvert a_{i}\rvert\leqslant C for i[0,m1]i\in\mathbb{N}\cap[0,m-1].

Proof.

First, note that the coefficient a0,,am1a_{0},\ldots,a_{m-1} form the unique solution of the following system of linear equations222We stick to the convention that 00=10^{0}=1.:

[00010m110111m1(m1)0(m1)1(m1)m1][a0a1am1]=[S𝝀T0ev2S𝝀T1ev2S𝝀Tm1ev2].\begin{bmatrix}0^{0}&0^{1}&\ldots&0^{m-1}\\ 1^{0}&1^{1}&\ldots&1^{m-1}\\ \vdots&\vdots&\vdots&\vdots\\ (m-1)^{0}&(m-1)^{1}&\ldots&(m-1)^{m-1}\end{bmatrix}\begin{bmatrix}a_{0}\\ a_{1}\\ \vdots\\ a_{m-1}\end{bmatrix}=\begin{bmatrix}\lVert S_{\boldsymbol{\lambda}_{T}}^{0}e_{v}\rVert^{2}\\ \lVert S_{\boldsymbol{\lambda}_{T}}^{1}e_{v}\rVert^{2}\\ \vdots\\ \lVert S_{\boldsymbol{\lambda}_{T}}^{m-1}e_{v}\rVert^{2}\\ \end{bmatrix}.

Applying Cramer’s rule (see [21, Theorem 3.36.(7)]) to the above system of linear equations, we obtain that there exist constants M,j,mM_{\ell,j,m}\in\mathbb{R} (j,[0,m1]j,\ell\in\mathbb{N}\cap[0,m-1]) depending on j,j,\ell (not on the operator S𝝀TS_{\boldsymbol{\lambda}_{T}}) satisfying

a=j=0m1M,j,mS𝝀Tjev2,[0,m1].a_{\ell}=\sum_{j=0}^{m-1}M_{\ell,j,m}\lVert S_{\boldsymbol{\lambda}_{T}}^{j}e_{v}\rVert^{2},\qquad\ell\in\mathbb{N}\cap[0,m-1].

Then, for every [0,m1]\ell\in\mathbb{N}\cap[0,m-1]

|a|j=0m1|M,j,m|S𝝀Tjev2j=0m1|M,j,m|S𝝀Tj2.\lvert a_{\ell}\rvert\leqslant\sum_{j=0}^{m-1}\lvert M_{\ell,j,m}\rvert\cdot\lVert S_{\boldsymbol{\lambda}_{T}}^{j}e_{v}\rVert^{2}\leqslant\sum_{j=0}^{m-1}\lvert M_{\ell,j,m}\rvert\cdot\lVert S_{\boldsymbol{\lambda}_{T}}^{j}\rVert^{2}.

Hence, the constant

C=sup0m1j=0m1|Mi,j,m|S𝝀Tj2C=\sup_{0\leqslant\ell\leqslant m-1}\sum_{j=0}^{m-1}\lvert M_{i,j,m}\rvert\cdot\lVert S_{\boldsymbol{\lambda}_{T}}^{j}\rVert^{2}

satisfies |a|C\lvert a_{\ell}\rvert\leqslant C. Obviously, CC depends only on mm and S𝝀TS_{\boldsymbol{\lambda}_{T}}. ∎

Proof of Lemma 5.1.

The equivalence of (i) and (ii) follows from remarks in [3, p. 389]; the implication (ii)\implies(iii) obviously holds. It remains to prove that (iii) implies (ii). Enumerate VT={vj}j=0NV_{T}=\{v_{j}\}_{j=0}^{N}, where N{}N\in\mathbb{N}\cup\{\infty\}. Let f=j=0Nαvjevj2(VT)f=\sum_{j=0}^{N}\alpha_{v_{j}}e_{v_{j}}\in\ell^{2}(V_{T}). Then, by Observation 4.3,

(5.2) S𝝀Tkf2=j=0N|αvj|2S𝝀Tkevj2,k.\lVert S_{\boldsymbol{\lambda}_{T}}^{k}f\rVert^{2}=\sum_{j=0}^{N}\lvert\alpha_{v_{j}}\rvert^{2}\lVert S_{\boldsymbol{\lambda}_{T}}^{k}e_{v_{j}}\rVert^{2},\qquad k\in\mathbb{N}.

For vVTv\in V_{T} let pv==0m1avxm1[x]p_{v}=\sum_{\ell=0}^{m-1}a_{\ell}^{v}x^{\ell}\in\mathbb{R}_{m-1}[x] be the polynomial satisfying (5.1). Let CC be the constant given by Observation 5.2. Then, for every [0,m1]\ell\in\mathbb{N}\cap[0,m-1],

j=0N|αvj|2|avj|Cj=0N|αvj|2.\sum_{j=0}^{N}\lvert\alpha_{v_{j}}\rvert^{2}\cdot\lvert a_{\ell}^{v_{j}}\rvert\leqslant C\sum_{j=0}^{N}\lvert\alpha_{v_{j}}\rvert^{2}.

Hence, the series j=0N|αvj|2avj\sum_{j=0}^{N}\lvert\alpha_{v_{j}}\rvert^{2}\cdot a_{\ell}^{v_{j}} is absolutely convergent for every [0,m1]\ell\in\mathbb{N}\cap[0,m-1]. By Lemma 2.2,

pf(x):=j=0N|αvj|2pvj(x)m1[x].p_{f}(x):=\sum_{j=0}^{N}\lvert\alpha_{v_{j}}\rvert^{2}p_{v_{j}}(x)\in\mathbb{R}_{m-1}[x].

From (5.2) and (5.1) it follows that pf(n)=S𝝀Tnf2p_{f}(n)=\lVert S_{\boldsymbol{\lambda}_{T}}^{n}f\rVert^{2} for every nn\in\mathbb{N}. ∎

In the next results we characterize mm-isometric composition operators such that the graph GTG_{T} has one cycle. Let us begin with the following lemma.

Lemma 5.3.

Suppose T:XXT\!:X\to X is such that GTG_{T} is connected and satisfies Theorem 3.5.(ii). Assume hTL(μ)h_{T}\in L^{\infty}(\mu) and let m2m\in\mathbb{N}_{2}. Suppose that for every RR\in\mathcal{R} there exists a polynomial pωRm1[x]p_{\omega_{R}}\in\mathbb{R}_{m-1}[x] satisfying (5.1) with v=ωRv=\omega_{R}. Then for every k[0,κ1]k\in\mathbb{N}\cap[0,\kappa-1] the formula

qk(x):=RiR=kλωR2pωR(x),x,q_{k}(x):=\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=k\end{subarray}}\lambda_{\omega_{R}}^{2}p_{\omega_{R}}(x),\qquad x\in\mathbb{R},

defines a polynomial of degree at most m1m-1. Moreover,

(5.3) degqk=max{degpωR:R,iR=k},k[0,κ1].\deg q_{k}=\max\{\deg p_{\omega_{R}}\!:R\in\mathcal{R},\ i_{R}=k\},\qquad k\in\mathbb{N}\cap[0,\kappa-1].
Proof.

Let k[0,κ1]k\in\mathbb{N}\cap[0,\kappa-1]. Enumerate {R:iR=k}={Rj}j=0N\{R\in\mathcal{R}\!:i_{R}=k\}=\{R_{j}\}_{j=0}^{N}, where N{}N\in\mathbb{N}\cup\{\infty\} is the cardinality of the set on the left hand side. Write pωR(x)=i=0m1aiRxip_{\omega_{R}}(x)=\sum_{i=0}^{m-1}a_{i}^{R}x^{i}, xx\in\mathbb{R}, RR\in\mathcal{R}. We will check that the series

(5.4) j=0NλωRj2aiRj\sum_{j=0}^{N}\lambda_{\omega_{R_{j}}}^{2}a_{i}^{R_{j}}

is absolutely convergent for every i[0,m1]i\in\mathbb{N}\cap[0,m-1]. Since, by Observation 4.3, for every [0,m1]\ell\in\mathbb{N}\cap[0,m-1],

S𝝀Tevk2=λk+1modκ2+j=0NλωRj2,\lVert S_{\boldsymbol{\lambda}_{T}}e_{v_{k}}\rVert^{2}=\lambda_{k+1\,\mathrm{mod}\,\kappa}^{2}+\sum_{j=0}^{N}\lambda_{\omega_{R_{j}}}^{2},

it follows that the series j=0NλωRj2\sum_{j=0}^{N}\lambda_{\omega_{R_{j}}}^{2} is convergent. Using Observation 5.2, we obtain that the series j=0NλωRj2aiRj\sum_{j=0}^{N}\lambda_{\omega_{R_{j}}}^{2}a_{i}^{R_{j}} is absolutely convergent. By Lemma 2.2, qkm1[x]q_{k}\in\mathbb{R}_{m-1}[x]. Since pωR(n)0p_{\omega_{R}}(n)\geqslant 0 for every nn\in\mathbb{N}, the leading coefficient of pωRp_{\omega_{R}} is non-negative for every RR\in\mathcal{R}. Hence, (5.3) holds. This completes the proof. ∎

It turns out that if S𝝀TS_{\boldsymbol{\lambda}_{T}} is mm-isometric, then on ’tree parts’ of GTG_{T} it is actually (m1)(m-1)-isometric. The similar result is presented in [18, Theorem 2.10] for composition operators such that the graph GTG_{T} has one cycle and one branching point; however, there is no clear way how to adjust the proof in a way to obtain the same conclusion in a more general situation. In our proof we use a different approach to prove this result in full generality.

Lemma 5.4.

Suppose T:XXT\!:X\to X is such that GTG_{T} is connected and satisfies the condition as in Theorem 3.5.(ii). Assume hTL(μ)h_{T}\in L^{\infty}(\mu) and let m2m\in\mathbb{N}_{2}. Suppose S𝛌T𝐁(2(VT))S_{\boldsymbol{\lambda}_{T}}\in\mathbf{B}(\ell^{2}(V_{T})) is mm-isometric. Then for every vVT{v0,,vκ1}v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\} the polynomial pvp_{v} satisfying (5.1) is of degree at most m2m-2.

Proof.

First, we prove that for every RR\in\mathcal{R}, the polynomial pωRp_{\omega_{R}} satisfying (5.1) with v=ωRv=\omega_{R} is of degree at most m2m-2. By Lemma 5.1, there exist pv0m1[x]p_{v_{0}}\in\mathbb{R}_{m-1}[x] and pωRm1[x]p_{\omega_{R}}\in\mathbb{R}_{m-1}[x] (RR\in\mathcal{R}) satisfying (5.1). Combining (5.1) and (4.3) we obtain that for every nn\in\mathbb{N},

(5.5) pv0(n+κ)pv0(n)=i=0κ1Λ0,i2RiR=iλωR2pωR(n+κi1).p_{v_{0}}(n+\kappa)-p_{v_{0}}(n)=\sum_{i=0}^{\kappa-1}\Lambda_{0,i}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i\end{subarray}}\lambda_{\omega_{R}}^{2}p_{\omega_{R}}(n+\kappa-i-1).

From Observation 2.1 we deduce that the polynomial q()=pv0(+κ)pv0()q(\,\cdot\,)=p_{v_{0}}(\,\cdot\,+\kappa)-p_{v_{0}}(\,\cdot\,) is of degree less than or equal to m2m-2. By Lemma 5.3, the formula

r(x)=i=0κ1Λ0,i2RiR=iλωR2pωR(x+κi1),x.r(x)=\sum_{i=0}^{\kappa-1}\Lambda_{0,i}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i\end{subarray}}\lambda_{\omega_{R}}^{2}p_{\omega_{R}}(x+\kappa-i-1),\qquad x\in\mathbb{R}.

defines a polynomial of degree at most m1m-1. We infer from (5.5) that q=rq=r and, consequently, rm2[x]r\in\mathbb{R}_{m-2}[x]. From Lemma 5.3 it follows that pωRm2[x]p_{\omega_{R}}\in\mathbb{R}_{m-2}[x]. Next, if RR\in\mathcal{R} and vv\in\mathcal{R} is such that pvm2[x]p_{v}\in\mathbb{R}_{m-2}[x], then proceeding as in the proof of Lemma 5.3, we obtain that

qv(x):=uVT(v,u)ETλu2pu(x1),x,q_{v}(x):=\sum_{\begin{subarray}{c}u\in V_{T}\end{subarray}\\ (v,u)\in E_{T}}\lambda_{u}^{2}p_{u}(x-1),\qquad x\in\mathbb{R},

is the polynomial of degree at most m1m-1. For every n1n\in\mathbb{N}_{1},

qv(n)=uVT(v,u)ETλu2pu(n1)=uVT(v,u)ETλu2S𝝀Tn1eu2=S𝝀Tnev2.q_{v}(n)=\sum_{\begin{subarray}{c}u\in V_{T}\end{subarray}\\ (v,u)\in E_{T}}\lambda_{u}^{2}p_{u}(n-1)=\sum_{\begin{subarray}{c}u\in V_{T}\end{subarray}\\ (v,u)\in E_{T}}\lambda_{u}^{2}\lVert S_{\boldsymbol{\lambda}_{T}}^{n-1}e_{u}\rVert^{2}=\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{v}\rVert^{2}.

Hence, qv=pvq_{v}=p_{v}. Since for every uVTu\in V_{T} the leading coefficient of pup_{u} is non-negative, it follows that degpudegpv\deg p_{u}\leqslant\deg p_{v}. Therefore, pum2[x]p_{u}\in\mathbb{R}_{m-2}[x]. The application of Lemma 2.3 completes the proof. ∎

Let us introduce the following notation: for m2m\in\mathbb{N}_{2} and κ1\kappa\in\mathbb{N}_{1} we set

Am,κ=[d1,10000d1,2d2,1000d1,m1d2,m2d3,m3dm1,1000001],A_{m,\kappa}=\begin{bmatrix}d_{1,1}&0&0&\ldots&0&0\\ d_{1,2}&d_{2,1}&0&\ldots&0&0\\ \vdots&\vdots&\vdots&\vdots&\vdots\\ d_{1,m-1}&d_{2,m-2}&d_{3,m-3}&\ldots&d_{m-1,1}&0\\ 0&0&0&\ldots&0&1\end{bmatrix},

where di,j=(mimji)κjd_{i,j}=\binom{m-i}{m-j-i}\kappa^{j}, i,ji,j\in\mathbb{N}, j+imj+i\leqslant m, and if κ>1\kappa>1,

Bm,κ=[1m11m2102m12m220(κ1)m1(κ1)m2(κ1)0].B_{m,\kappa}=\begin{bmatrix}1^{m-1}&1^{m-2}&\ldots&1^{0}\\ 2^{m-1}&2^{m-2}&\ldots&2^{0}\\ \vdots&\vdots&\vdots&\vdots\\ (\kappa-1)^{m-1}&(\kappa-1)^{m-2}&\ldots&(\kappa-1)^{0}\end{bmatrix}.

Now we are in the position to state the main result of this section.

Theorem 5.5.

Suppose T:XXT\!:X\to X is such that GTG_{T} is connected and satisfies Theorem 3.5.(ii). Assume hTL(μ)h_{T}\in L^{\infty}(\mu). For m2m\in\mathbb{N}_{2} the following conditions are equivalent:

  1. (i)

    S𝝀TS_{\boldsymbol{\lambda}_{T}} is mm-isometric,

  2. (ii)

    for every vVT{v0,,vκ1}v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\} there exists the unique polynomial pvm2[x]p_{v}\in\mathbb{R}_{m-2}[x] satisfying (5.1) and if κ>1\kappa>1, m=rankA~m,κ=rankB~m,κm=\mathrm{rank}\,\tilde{A}_{m,\kappa}=\mathrm{rank}\,\tilde{B}_{m,\kappa}, where

    A~m,κ=[Am,κb],B~m,κ=[Am,κbBm,κa],\tilde{A}_{m,\kappa}=\begin{bmatrix}A_{m,\kappa}&b\end{bmatrix},\tilde{B}_{m,\kappa}=\begin{bmatrix}A_{m,\kappa}&b\\ B_{m,\kappa}&a\end{bmatrix},
    b=[bm2bm3b01]T,a=[a1a2aκ1]T,b=\begin{bmatrix}b_{m-2}&b_{m-3}&\ldots&b_{0}&1\end{bmatrix}^{T},\quad a=\begin{bmatrix}a_{1}&a_{2}&\ldots&a_{\kappa-1}\end{bmatrix}^{T},

    ai=S𝝀iev02a_{i}=\lVert S_{\boldsymbol{\lambda}}^{i}e_{v_{0}}\rVert^{2} (ii\in\mathbb{N}, j[1,mi1]j\in\mathbb{N}\cap[1,m-i-1]), and q(x)=i=0m2bixim2[x]q(x)=\sum_{i=0}^{m-2}b_{i}x^{i}\in\mathbb{R}_{m-2}[x] is the polynomial defined by the formula333The fact that this formula defines a polynomial of degree at most m2m-2 follows from Lemma 5.3.

    (5.6) q(x)=i=0κ1Λ0,i2RiR=iλωR2pωR(x+κi1),x.q(x)=\sum_{i=0}^{\kappa-1}\Lambda_{0,i}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i\end{subarray}}\lambda_{\omega_{R}}^{2}p_{\omega_{R}}(x+\kappa-i-1),\qquad x\in\mathbb{R}.

Moreover, if (ii) holds, then pv0(x)=j=0m1cjxjp_{v_{0}}(x)=\sum_{j=0}^{m-1}c_{j}x^{j}, where c=(cm1,,c0)c=(c_{m-1},\ldots,c_{0}) is the solution of

(5.7) Am,κc=b\displaystyle A_{m,\kappa}c=b
Proof.

(i) \Longrightarrow (ii). By Lemma 5.1, for every vVTv\in V_{T} there exists pvm1[x]p_{v}\in\mathbb{R}_{m-1}[x] satisfying (5.1). By Lemma 5.4, for every RR\in\mathcal{R} and every vRv\in R, pvm2[x]p_{v}\in\mathbb{R}_{m-2}[x]. Next, writing pv0(x)=j=0m1cjxjp_{v_{0}}(x)=\sum_{j=0}^{m-1}c_{j}x^{j} and using the binomial formula, we have

pv0(x+κ)pv0(x)==0m2(j=+1m1(j)cjκj)xj.p_{v_{0}}(x+\kappa)-p_{v_{0}}(x)=\sum_{\ell=0}^{m-2}\left(\sum_{j=\ell+1}^{m-1}\binom{j}{\ell}c_{j}\kappa^{j-\ell}\right)x^{j}.

Since, by Lemma 5.3, the formula

q(x)=i=0κ1Λ0,i2RiR=iλωR2pωR(x+κi1)q(x)=\sum_{i=0}^{\kappa-1}\Lambda_{0,i}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i\end{subarray}}\lambda_{\omega_{R}}^{2}p_{\omega_{R}}(x+\kappa-i-1)

defines a polynomial of degree at most m2m-2 satisfying, by (4.3),

q(n)=pv0(n+κ)pv0(n),n,q(n)=p_{v_{0}}(n+\kappa)-p_{v_{0}}(n),\qquad n\in\mathbb{N},

writing q(x)==0m2bxq(x)=\sum_{\ell=0}^{m-2}b_{\ell}x^{\ell} we obtain

(5.8) b=j=+1m1(j)cjκj,[0,m2].b_{\ell}=\sum_{j=\ell+1}^{m-1}\binom{j}{\ell}c_{j}\kappa^{j-\ell},\qquad\ell\in\mathbb{N}\cap[0,m-2].

Hence, (cm1,,c0)(c_{m-1},\ldots,c_{0}) satisfies (5.7). Moreover, since the determinant of Am,κA_{m,\kappa} is non-zero, we infer from Cramer’s rule that c=(cm1,,c0)c=(c_{m-1},\ldots,c_{0}) is the only solution of this system. By the Kronecker-Capelli theorem (see [21, Theorem 3.36.(1)-(3)]) this implies that m=rankA~m,κm=\mathrm{rank}\,\tilde{A}_{m,\kappa}. Assume now κ>1\kappa>1. From the equality pv0(i)=aip_{v_{0}}(i)=a_{i} (i[0,κ1]i\in\mathbb{N}\cap[0,\kappa-1]) we get that cc is also the solution of the following system of linear equations:

b\displaystyle b_{\ell} =j=+1m1(j)cjκj,[0,m2],\displaystyle=\sum_{j=\ell+1}^{m-1}\binom{j}{\ell}c_{j}\kappa^{j-\ell},\qquad\ell\in\mathbb{N}\cap[0,m-2],
a\displaystyle a_{\ell} =j=0m1cjj,[1,κ1];\displaystyle=\sum_{j=0}^{m-1}c_{j}\ell^{j},\qquad\ell\in\mathbb{N}\cap[1,\kappa-1];

the matrix representation of this system takes the form:

(5.9) [Am,κBm,κ]c=[ba]\displaystyle\begin{bmatrix}A_{m,\kappa}\\ B_{m,\kappa}\end{bmatrix}c=\begin{bmatrix}b\\ a\end{bmatrix}

Again, by the Kronecker-Capelli theorem, m=rankB~m,κm=\mathrm{rank}\,\tilde{B}_{m,\kappa}, which gives (ii).
(ii) \Longrightarrow (i). By Lemma 5.1, it is enough to find the polynomials pvk[x]p_{v_{k}}\in\mathbb{R}[x] (k[0,κ1]k\in\mathbb{N}\cap[0,\kappa-1]) satisfying (5.1). First, observe that for every k[1,κ1]k\in\mathbb{N}\cap[1,\kappa-1] the polynomial pvkp_{v_{k}} is uniquely determined by pv0p_{v_{0}} and pvp_{v} (vV{v0,,vκ1}v\in V\setminus\{v_{0},\ldots,v_{\kappa-1}\}). Indeed, assuming we have pv0p_{v_{0}} satisfying (5.1) and using (4.2), we obtain

S𝝀Tnevk2\displaystyle\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{v_{k}}\rVert^{2} =Λ0,k2(pv0(n+k)\displaystyle=\Lambda_{0,k}^{-2}(p_{v_{0}}(n+k)
i=0k1Λ0,i2RiR=imodκλωR2pωR(n+ki1)),n.\displaystyle-\sum_{i=0}^{k-1}\Lambda_{0,i}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i\,\mathrm{mod}\,\kappa\end{subarray}}\lambda_{\omega_{R}}^{2}p_{\omega_{R}}(n+k-i-1)),\quad n\in\mathbb{N}.

By Lemma 5.3,

pvk()\displaystyle p_{v_{k}}(\,\cdot\,) :=Λ0,k2(pv0(+k)i=0k1Λ0,i2RiR=imodκλωR2pωR(+ki1))\displaystyle:=\Lambda_{0,k}^{-2}(p_{v_{0}}(\,\cdot\,+k)-\sum_{i=0}^{k-1}\Lambda_{0,i}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i\,\mathrm{mod}\,\kappa\end{subarray}}\lambda_{\omega_{R}}^{2}p_{\omega_{R}}(\,\cdot+k-i-1))

is a polynomial of degree at most m1m-1, which satisfies (5.1). Hence, we have to find only the polynomial pv0p_{v_{0}}. By Cramer’s rule, (5.7)\eqref{FormLinearSystemRecursion} has only one solution; call c=(cm1,,c0)c=(c_{m-1},\ldots,c_{0}) this unique solution and define pv0(x)=j=0m1cjxjp_{v_{0}}(x)=\sum_{j=0}^{m-1}c_{j}x^{j} for xx\in\mathbb{R}. By (ii) and the Kronecker-Capelli theorem, (cm1,,c0)(c_{m-1},\ldots,c_{0}) has to satisfy also (5.9). Therefore,

pv0(n)=j=0m1cjnj=aj=S𝝀Tnev02,n[0,κ1].p_{v_{0}}(n)=\sum_{j=0}^{m-1}c_{j}n^{j}=a_{j}=\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{v_{0}}\rVert^{2},\qquad n\in\mathbb{N}\cap[0,\kappa-1].

Since (cm1,,c0)(c_{m-1},\ldots,c_{0}) satisfies (5.8), we get that pv0(+κ)pv0()=q()p_{v_{0}}(\,\cdot\,+\kappa)-p_{v_{0}}(\,\cdot\,)=q(\,\cdot\,). Thus, from (4.3) we obtain that pv0(n)=S𝝀Tnev02p_{v_{0}}(n)=\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{v_{0}}\rVert^{2} for every nn\in\mathbb{N}. The ’moreover’ part easily follows from the above reasoning. ∎

In [18] the authors proved that in the class of composition operators such that the graph GTG_{T} has one cycle and one branching point every completely hyperexpansive operator is 2-isometric. Using different approach, with the aid of Theorem 5.5, we are able to prove this result in full generality.

Theorem 5.6 (cf. [18, Corollary 2.15]).

Suppose T:XXT\!:X\to X is such that GTG_{T} is connected and satisfies Theorem 3.5.(ii). Suppose hTL(μ)h_{T}\in L^{\infty}(\mu). The following conditions are equivalent:

  1. (i)

    S𝝀TS_{\boldsymbol{\lambda}_{T}} is 2-isometric,

  2. (ii)

    S𝝀TS_{\boldsymbol{\lambda}_{T}} is completely hyperexpansive.

Proof.

(i)\Longrightarrow(i). By Lemma 5.1, if S𝝀TS_{\boldsymbol{\lambda}_{T}} is 2-isometric, then it is mm-isometric for every m2m\in\mathbb{N}_{2}. Combining this with [3, Proposition 1.5] we obtain that S𝝀TS_{\boldsymbol{\lambda}_{T}} is completely hyperexpansive.
(ii)\Longrightarrow(i). From (ii), by [7, Remark 2], it follows that for every vVTv\in V_{T} there exists the measure τv:([0,1])[0,)\tau_{v}\!:\mathcal{B}([0,1])\to[0,\infty) satisfying

(5.10) S𝝀Tnev2=1+[0,1](1+t++tn1)dτv(t),n1.\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{v}\rVert^{2}=1+\int_{[0,1]}(1+t+\ldots+t^{n-1})\,\mathrm{d}\tau_{v}(t),\qquad n\in\mathbb{N}_{1}.

Inserting (5.10) into (4.3) we obtain that for every n1n\in\mathbb{N}_{1},

1+[0,1](1++tn+κ1)dτvm(t)=1+[0,1](1++tn1)dτvm(t)\displaystyle 1+\int_{[0,1]}(1+\ldots+t^{n+\kappa-1})\,\mathrm{d}\tau_{v_{m}}(t)=1+\int_{[0,1]}(1+\ldots+t^{n-1})\,\mathrm{d}\tau_{v_{m}}(t)
(5.11) +i=0κ1Λm,i2RiR=i+mmodκλωR2(1+[0,1](1++tn+κi2)dτωR(t))\displaystyle+\sum_{i=0}^{\kappa-1}\Lambda_{m,i}^{2}\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}^{2}(1+\int_{[0,1]}(1+\ldots+t^{n+\kappa-i-2})\,\mathrm{d}\tau_{\omega_{R}}(t))

If vVTv\in V_{T}, then for every nn\in\mathbb{N} we have

λu2[0,1]tndτu(t)λu2(1+[0,1](1++tn))dτu(t),uVT,(v,u)ET.\lambda_{u}^{2}\int_{[0,1]}t^{n}\,\mathrm{d}\tau_{u}(t)\leqslant\lambda_{u}^{2}(1+\int_{[0,1]}(1+\ldots+t^{n}))\,\mathrm{d}\tau_{u}(t),\qquad u\in V_{T},(v,u)\in E_{T}.

By (5.10) and Observation 4.3, for every nn\in\mathbb{N},

uVT(v,u)ETλu2(1+[0,1](1++tn))dτu(t)=uVT(v,u)ETλu2S𝝀Tn+1eu2=S𝝀Tn+2ev2.\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{u}^{2}(1+\int_{[0,1]}(1+\ldots+t^{n}))\,\mathrm{d}\tau_{u}(t)=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\!\!\!\!\lambda_{u}^{2}\lVert S_{\boldsymbol{\lambda}_{T}}^{n+1}e_{u}\rVert^{2}=\lVert S_{\boldsymbol{\lambda}_{T}}^{n+2}e_{v}\rVert^{2}.

Thus, the series

uVT(v,u)ETλu2[0,1]tndτu\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{u}^{2}\int_{[0,1]}t^{n}\,\mathrm{d}\tau_{u}

is convergent for every vVTv\in V_{T} and nn\in\mathbb{N}. By (5.10) we have that

[0,1]tn+κdτvm(t)=S𝝀Tn+κ+1evm2S𝝀Tn+κevm2\displaystyle\int_{[0,1]}t^{n+\kappa}\,\mathrm{d}\tau_{v_{m}}(t)=\lVert S_{\boldsymbol{\lambda}_{T}}^{n+\kappa+1}e_{v_{m}}\rVert^{2}-\lVert S_{\boldsymbol{\lambda}_{T}}^{n+\kappa}e_{v_{m}}\rVert^{2}
=(1+[0,1](1++tn+κ)dτvm(t))(1+[0,1](1++tn+κ1)dτvm(t))\displaystyle=(1+\int_{[0,1]}(1+\ldots+t^{n+\kappa})\,\mathrm{d}\tau_{v_{m}}(t))-(1+\int_{[0,1]}(1+\ldots+t^{n+\kappa-1})\,\mathrm{d}\tau_{v_{m}}(t))

Using (5) with nn replaced with n+1n+1, we get that for every n1n\in\mathbb{N}_{1},

[0,1]tn+κdτvm(t)\displaystyle\int_{[0,1]}t^{n+\kappa}\,\mathrm{d}\tau_{v_{m}}(t) =[0,1]tndτvm(t)\displaystyle=\int_{[0,1]}t^{n}\,\mathrm{d}\tau_{v_{m}}(t)
+i=0κ1Λm,i2RiR=i+mmodκλωR2[0,1]tn+κi1dτωR(t),\displaystyle+\sum_{i=0}^{\kappa-1}\Lambda_{m,i}^{2}\!\!\!\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\!\!\lambda_{\omega_{R}}^{2}\int_{[0,1]}t^{n+\kappa-i-1}\,\mathrm{d}\tau_{\omega_{R}}(t),

which implies that

[0,1]tn(tκ1)dτvm(t)=i=0κ1Λm,i2RiR=i+mmodκλωR2[0,1]tn+κi1dτωR(t).\displaystyle\int_{[0,1]}t^{n}(t^{\kappa}-1)\,\mathrm{d}\tau_{v_{m}}(t)=\sum_{i=0}^{\kappa-1}\Lambda_{m,i}^{2}\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}^{2}\int_{[0,1]}t^{n+\kappa-i-1}\,\mathrm{d}\tau_{\omega_{R}}(t).

Since tκ10t^{\kappa}-1\leqslant 0 for every t[0,1]t\in[0,1], the integral on the left hand side is non-positive. On the other hand, we have λv>0\lambda_{v}>0 for every vVTv\in V_{T}, so each term in the sum on the right hand side is non-negative. Therefore, we get that for all m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1], n1n\in\mathbb{N}_{1},

(5.12) [0,1]tn(tκ1)dτvm(t)=0\int_{[0,1]}t^{n}(t^{\kappa}-1)\,\mathrm{d}\tau_{v_{m}}(t)=0

and for all RR\in\mathcal{R},

(5.13) [0,1]tn+κi1dτωR(t)=0,i[0,κ1],i+m=iRmodκ.\int_{[0,1]}t^{n+\kappa-i-1}\,\mathrm{d}\tau_{\omega_{R}}(t)=0,\qquad i\in\mathbb{N}\cap[0,\kappa-1],\ i+m=i_{R}\,\mathrm{mod}\,\kappa.

By (5.13), τωR((0,1])=0\tau_{\omega_{R}}((0,1])=0. This implies that for all RR\in\mathcal{R}, τωR=CRδ0\tau_{\omega_{R}}=C_{R}\delta_{0} with some CR[0,)C_{R}\in[0,\infty). In turn, by (5.12), τvm((0,1))=0\tau_{v_{m}}((0,1))=0. Hence, for all m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1], τvm=Cmδ0+Dmδ1\tau_{v_{m}}=C_{m}\delta_{0}+D_{m}\delta_{1} with some constants Cm,Dm[0,)C_{m},D_{m}\in[0,\infty). Inserting these into (5) we obtain that for every n2n\in\mathbb{N}_{2} and every m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1],

1+Cm+(n+κ)Dm=1+Cm+nDm+i=0κ1Λm,i2RiR=i+mmodκλωR2(1+CR).1+C_{m}+(n+\kappa)D_{m}=1+C_{m}+nD_{m}+\sum_{i=0}^{\kappa-1}\Lambda_{m,i}^{2}\!\!\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\!\!\lambda_{\omega_{R}}^{2}(1+C_{R}).

Hence,

(5.14) κDm=i=0κ1Λm,i2RiR=i+mmodκλωR2(1+CR).\kappa D_{m}=\sum_{i=0}^{\kappa-1}\Lambda_{m,i}^{2}\!\!\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\!\!\lambda_{\omega_{R}}^{2}(1+C_{R}).

If κ>1\kappa>1, then, using (4.3) with n=0n=0 and (5.10) with v=vmv=v_{m}, we get that

1+Cm+κDm\displaystyle 1+C_{m}+\kappa D_{m}
=1+i=0κ2Λm,i2RiR=i+mmodκλωR2(1+CR)+Λm,κ12RiR=m1λωR2\displaystyle=1+\sum_{i=0}^{\kappa-2}\Lambda_{m,i}^{2}\!\!\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\!\!\lambda_{\omega_{R}}^{2}(1+C_{R})+\Lambda_{m,\kappa-1}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=m-1\end{subarray}}\lambda_{\omega_{R}}^{2}
=1+i=0κ1Λm,i2RiR=i+mmodκλωR2(1+CR)Λm,κ12RiR=m1λωR2CR;\displaystyle=1+\sum_{i=0}^{\kappa-1}\Lambda_{m,i}^{2}\!\!\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\!\!\lambda_{\omega_{R}}^{2}(1+C_{R})-\Lambda_{m,\kappa-1}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=m-1\end{subarray}}\lambda_{\omega_{R}}^{2}C_{R};

if κ=1\kappa=1, by (4.2) applied with k=1k=1, we have

1+C0+κD0\displaystyle 1+C_{0}+\kappa D_{0}
=1+RλωR2\displaystyle=1+\sum_{\begin{subarray}{c}R\in\mathcal{R}\end{subarray}}\lambda_{\omega_{R}}^{2}
=1+RλωR2(1+CR)RλωR2CR;\displaystyle=1+\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ \end{subarray}}\lambda_{\omega_{R}}^{2}(1+C_{R})-\sum_{\begin{subarray}{c}R\in\mathcal{R}\end{subarray}}\lambda_{\omega_{R}}^{2}C_{R};

Combining the above with (5.14), we obtain

1+Cm+κDm=1+κDmΛm,κ12RiR=m1λωR2CR,\displaystyle 1+C_{m}+\kappa D_{m}=1+\kappa D_{m}-\Lambda_{m,\kappa-1}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=m-1\end{subarray}}\lambda_{\omega_{R}}^{2}C_{R},

which implies that

Cm=Λm,m+κ12RiR=m1λωR2CR.\displaystyle C_{m}=-\Lambda_{m,m+\kappa-1}^{2}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=m-1\end{subarray}}\lambda_{\omega_{R}}^{2}C_{R}.

Since λv>0\lambda_{v}>0 for all vVTv\in V_{T}, we have that Cm0C_{m}\leqslant 0 for m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1]. On the other hand, Cm0C_{m}\geqslant 0. Therefore, Cm=0C_{m}=0 for all m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1], and, consequently, CR=0C_{R}=0 for every RR\in\mathcal{R}. Hence,

τωR=0,R,\tau_{\omega_{R}}=0,\qquad R\in\mathcal{R},

and

τvm=Dmδ1,m[0,κ1].\tau_{v_{m}}=D_{m}\delta_{1},\qquad m\in\mathbb{N}\cap[0,\kappa-1].

By (5.10), this implies that S𝝀TneωR=1\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{\omega_{R}}\rVert=1 for nn\in\mathbb{N} and that S𝝀nevm2=1+nDm\lVert S_{\boldsymbol{\lambda}}^{n}e_{v_{m}}\rVert^{2}=1+nD_{m} for nn\in\mathbb{N}. It remains to show that

τv=0,vVT{v0,,vκ1},n,\tau_{v}=0,\qquad v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\},n\in\mathbb{N},

which implies that

S𝝀Tnev=1,vVT{v0,,vκ1},n.\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{v}\rVert=1,\qquad v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\},n\in\mathbb{N}.

Since we have already proved this for v=ωRv=\omega_{R}, RR\in\mathcal{R}, by Lemmata 3.4 and 2.3, it is enough to show that if τv=0\tau_{v}=0 for some vVT{v0,,vκ1}v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\}, then τu=0\tau_{u}=0 for every uVTu\in V_{T} satisfying (v,u)ET(v,u)\in E_{T}. Assume vVT{v0,,vκ1}v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\} is such that τv=0\tau_{v}=0. Since

S𝝀Tn+1ev=uVT(v,u)ETλuS𝝀Tneu,n,S_{\boldsymbol{\lambda}_{T}}^{n+1}e_{v}=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{u}S_{\boldsymbol{\lambda}_{T}}^{n}e_{u},\qquad n\in\mathbb{N},

it follows from (5.10) that

(5.15) 1=uVT(v,u)ETλu2(1+[0,1](1++tn1)dτu(t)),n1.1=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{u}^{2}(1+\int_{[0,1]}(1+\ldots+t^{n-1})\,\mathrm{d}\tau_{u}(t)),\qquad n\in\mathbb{N}_{1}.

If we replace nn with n+1n+1 in the above equality, we obtain

1\displaystyle 1 =uVT(v,u)ETλu2(1+[0,1](1++tn)dτu(t))\displaystyle=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{u}^{2}(1+\int_{[0,1]}(1+\ldots+t^{n})\,\mathrm{d}\tau_{u}(t))
=uVT(v,u)ETλu2(1+[0,1](1++tn1)dτu(t))+uVT(v,u)ETλv2[0,1]tndτu(t)\displaystyle=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{u}^{2}(1+\int_{[0,1]}(1+\ldots+t^{n-1})\,\mathrm{d}\tau_{u}(t))+\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{v}^{2}\int_{[0,1]}t^{n}\,\mathrm{d}\tau_{u}(t)
=(5.15)1+uVT(v,u)ETλv2[0,1]tndτu(t).\displaystyle\stackrel{{\scriptstyle\eqref{FormCompletelyAlternatingVertexWithZeroMeasure}}}{{=}}1+\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{v}^{2}\int_{[0,1]}t^{n}\,\mathrm{d}\tau_{u}(t).

Hence,

0=uVT(v,u)ETλv2[0,1]tndτu(t),n1,0=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{v}^{2}\int_{[0,1]}t^{n}\,\mathrm{d}\tau_{u}(t),\qquad n\in\mathbb{N}_{1},

which implies that for every uVTu\in V_{T} satisfying (v,u)ET(v,u)\in E_{T} we have τu((0,1])=0\tau_{u}((0,1])=0. Thus, τu=Cuδ0\tau_{u}=C_{u}\delta_{0} with some Cu[0,)C_{u}\in[0,\infty). Then (5.15) takes the form

1=uVT(v,u)ETλu2(1+Cu),n1.1=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{u}^{2}(1+C_{u}),\qquad n\in\mathbb{N}_{1}.

But

1=S𝝀Tev2=uVT(v,u)ETλu2.1=\lVert S_{\boldsymbol{\lambda}_{T}}e_{v}\rVert^{2}=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}\lambda_{u}^{2}.

Combining these two equalities we obtain that Cu=0C_{u}=0 for every uVTu\in V_{T} satisfying (v,u)ET(v,u)\in E_{T}. ∎

As another corollary of the equality (4.3) we obtain that in our setting graphs of isometric composition operators consist only of cycles.

Corollary 5.7.

Suppose T:XXT\!:X\to X is such that GTG_{T} is connected and satisfies Theorem 3.5.(ii). Suppose hTL(μ)h_{T}\in L^{\infty}(\mu). If S𝛌T𝐁(2(VT))S_{\boldsymbol{\lambda}_{T}}\in\mathbf{B}(\ell^{2}(V_{T})) is isometric, then =\mathcal{R}=\varnothing, that is, VT={v0,,vκ1}V_{T}=\{v_{0},\ldots,v_{\kappa-1}\}. In particular, 2(VT)\ell^{2}(V_{T}) is finite dimensional and S𝛌TS_{\boldsymbol{\lambda}_{T}} is unitary.

Proof.

If S𝝀TS_{\boldsymbol{\lambda}_{T}} is isometric, then so is S𝝀TnS_{\boldsymbol{\lambda}_{T}}^{n} for every nn\in\mathbb{N}. Thus, from (4.3) it follows that

i=0κ1Λm,i2RiR=i+mmodκλωR2S𝝀Tn+κi1eωR2=0,m[0,κ1],n.\sum_{i=0}^{\kappa-1}\Lambda_{m,i}^{2}\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}^{2}\lVert S_{\boldsymbol{\lambda}_{T}}^{n+\kappa-i-1}e_{\omega_{R}}\rVert^{2}=0,\qquad m\in\mathbb{N}\cap[0,\kappa-1],\ n\in\mathbb{N}.

Since Λm,i>0\Lambda_{m,i}>0 for all i,m[0,κ1]i,m\in\mathbb{N}\cap[0,\kappa-1], we have that

RiR=i+mmodκλωR2S𝝀Tn+κi1eωR2=0,i,m[0,κ1],n.\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\lambda_{\omega_{R}}^{2}\lVert S_{\boldsymbol{\lambda}_{T}}^{n+\kappa-i-1}e_{\omega_{R}}\rVert^{2}=0,\qquad i,m\in\mathbb{N}\cap[0,\kappa-1],\ n\in\mathbb{N}.

From this it follows that

RiR=iλωR2=0,i[0,κ1].\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i\end{subarray}}\lambda_{\omega_{R}}^{2}=0,\qquad i\in\mathbb{N}\cap[0,\kappa-1].

Since λωR>0\lambda_{\omega_{R}}>0 for all RR\in\mathcal{R}, the only possibility for the above sum being zero is that the set of indices is empty, that is, =\mathcal{R}=\varnothing. This implies that 2(VT)\ell^{2}(V_{T}) is finite dimensional. Hence, S𝝀TS_{\boldsymbol{\lambda}_{T}} is invertible and, by [3, Proposition 1.23], unitary. ∎

In the end of this section we provide an example of use of Theorem 5.5. We characterize 3-isometric composition operators on graphs, in which the only branching points are the vertices on the cycle. Let κ1\kappa\in\mathbb{N}_{1} and {Ni}i=0κ11{}\{N_{i}\}_{i=0}^{\kappa-1}\subset\mathbb{N}_{1}\cup\{\infty\}. Denote

X={0,1,,κ1}{(i,j,k):i[0,κ1],j[1,Ni],k}.X=\{0,1,\ldots,\kappa-1\}\cup\{(i,j,k)\!:i\in\mathbb{N}\cap[0,\kappa-1],\ j\in\mathbb{N}\cap[1,N_{i}],\ k\in\mathbb{N}\}.

Define T:XXT\!:X\to X by the formula

T(x)={κ1,if x=0x1,if x[1,κ1]i,if x=(i,j,0),0iκ1,0jNi(i,j,k1),if x=(i,j,k),k1,0iκ1,0jNi.T(x)=\begin{cases}\kappa-1,\quad\text{if }x=0\\ x-1,\quad\text{if }x\in\mathbb{N}\cap[1,\kappa-1]\\ i,\quad\text{if }x=(i,j,0),0\leqslant i\leqslant\kappa-1,0\leqslant j\leqslant N_{i}\\ (i,j,k-1),\quad\text{if }x=(i,j,k),k\geqslant 1,0\leqslant i\leqslant\kappa-1,0\leqslant j\leqslant N_{i}\end{cases}.

Then the graph GT=(VT,ET)G_{T}=(V_{T},E_{T}) takes the form

VT\displaystyle V_{T} =X,\displaystyle=X,
ET\displaystyle E_{T} ={(i,(i+1)modκ):i[0,κ1]}\displaystyle=\{(i,(i+1)\,\mathrm{mod}\,\kappa)\!:i\in\mathbb{N}\cap[0,\kappa-1]\}
{(i,(i,j,0)):i[0,κ1],j[1,Ni]}\displaystyle\cup\{(i,(i,j,0))\!:i\in\mathbb{N}\cap[0,\kappa-1],j\in\mathbb{N}\cap[1,N_{i}]\}
{((i,j,k),(i,j,k+1)):i[0,κ1],j[1,Ni],k};\displaystyle\cup\{((i,j,k),(i,j,k+1))\!:i\in\mathbb{N}\cap[0,\kappa-1],\ j\in\mathbb{N}\cap[1,N_{i}],\ k\in\mathbb{N}\};

we present this graph in Figure 2

Refer to caption
Figure 2. Graph GTG_{T}, in which the only branching points are the vertices on the cycle
Theorem 5.8.

Let XX and TT be defined as above and let μ:2X[0,]\mu\!:2^{X}\to[0,\infty] be a measure on XX such that μ(x)(0,)\mu(x)\in(0,\infty) for every xXx\in X. Assume that hTL(μ)h_{T}\in L^{\infty}(\mu). The following conditions are equivalent:

  1. (i)

    CT𝐁(L2(μ))C_{T}\in\mathbf{B}(L^{2}(\mu)) is 3-isometric,

  2. (ii)

    for all i[0,κ1],j[1,Ni]i\in\mathbb{N}\cap[0,\kappa-1],j\in\mathbb{N}\cap[1,N_{i}],

    (5.16) μ((i,j,k))=(μ((i,j,1))μ((i,j,0)))k+μ((i,j,0)),k,\mu((i,j,k))=(\mu((i,j,1))-\mu((i,j,0)))k+\mu((i,j,0)),\quad k\in\mathbb{N},

    and444Note that under the assumption hTL(μ)h_{T}\in L^{\infty}(\mu) we have that for every i[0,κ1]i\in\mathbb{N}\cap[0,\kappa-1], j=1Niμ((i,j,0))=μ(T1({i}))=(2.2)hT(i)μ({i})<\sum_{j=1}^{N_{i}}\mu((i,j,0))=\mu(T^{-1}(\{i\}))\stackrel{{\scriptstyle\eqref{FormRadonNikodymDerivativeDiscrete}}}{{=}}h_{T}(i)\mu(\{i\})<\infty; similar argument shows that j=1Niμ((i,j,1))<\sum_{j=1}^{N_{i}}\mu((i,j,1))<\infty. By (5.16), we get that j=1Niμ((i,j,i1))<\sum_{j=1}^{N_{i}}\mu((i,j,i-\ell-1))<\infty for i[1,κ1]i\in\mathbb{N}\cap[1,\kappa-1], [0,i1]\ell\in\mathbb{N}\cap[0,i-1].

    (5.17) μ(i)=q0(i)=0i1j=1Nμ((,j,i1)),i[1,κ1],\mu(i)=q_{0}(i)-\sum_{\ell=0}^{i-1}\sum_{j=1}^{N_{\ell}}\mu((\ell,j,i-\ell-1)),\quad i\in\mathbb{N}\cap[1,\kappa-1],

    where q0(x)=A2κx2+2BAκ2κx+μ(0)2[x]q_{0}(x)=\frac{A}{2\kappa}x^{2}+\frac{2B-A\kappa}{2\kappa}x+\mu(0)\in\mathbb{R}_{2}[x] and

    A\displaystyle A =i=0κ1j=1Ni(μ((i,j,1))μ((i,j,0))),\displaystyle=\sum_{i=0}^{\kappa-1}\sum_{j=1}^{N_{i}}(\mu((i,j,1))-\mu((i,j,0))),
    B\displaystyle B =i=0κ1j=1Ni[(μ((i,j,1))μ((i,j,0)))(κi1)+μ((i,j,0))].\displaystyle=\sum_{i=0}^{\kappa-1}\sum_{j=1}^{N_{i}}\left[(\mu((i,j,1))-\mu((i,j,0)))(\kappa-i-1)+\mu((i,j,0))\right].
Proof.

(i)\Longrightarrow(ii). If CTC_{T} is 3-isometric, then so is S𝝀TS_{\boldsymbol{\lambda}_{T}}. By Theorem 5.5, this implies that for every i[0,κ1]i\in\mathbb{N}\cap[0,\kappa-1] and j[1,Ni]j\in\mathbb{N}\cap[1,N_{i}] there exists a polynomial pi,j1[x]p_{i,j}\in\mathbb{R}_{1}[x] satisfying (5.1) with v=(i,j,0)v=(i,j,0); it is a matter of routine to verify that pi,jp_{i,j} takes the form

(5.18) pi,j(x)=(λ(i,j,1)21)x+1,x.p_{i,j}(x)=(\lambda_{(i,j,1)}^{2}-1)x+1,\qquad x\in\mathbb{R}.

Since, by (4.1),

pi,j(n)=S𝝀Tne(i,j,0)2=k=1nλi,j,k2=μ((i,j,n))μ((i,j,0)),n1,p_{i,j}(n)=\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{(i,j,0)}\rVert^{2}=\prod_{k=1}^{n}\lambda_{i,j,k}^{2}=\frac{\mu((i,j,n))}{\mu((i,j,0))},\quad n\in\mathbb{N}_{1},

it follows that the polynomial qi,j=μ((i,j,0))pi,jq_{i,j}=\mu((i,j,0))p_{i,j} satisfies the equality qi,j(k)=μ((i,j,k))q_{i,j}(k)=\mu((i,j,k)) for every kk\in\mathbb{N}; moreover, by (5.18) and (4.1), qi,jq_{i,j} takes the form

qi,j(x)=(μ((i,j,1))μ((i,j,0)))x+μ((i,j,0)),x.q_{i,j}(x)=(\mu((i,j,1))-\mu((i,j,0)))x+\mu((i,j,0)),\qquad x\in\mathbb{R}.

Hence, (5.16) holds. Next, by (4.1) and (5.18), the polynomial qq given by (5.6) takes the form

q(x)\displaystyle q(x) =i=0κ1μ(i)μ(0)j=1Niμ((i,j,0))μ(i)pi,j(x+κi1)\displaystyle=\sum_{i=0}^{\kappa-1}\frac{\mu(i)}{\mu(0)}\sum_{j=1}^{N_{i}}\frac{\mu((i,j,0))}{\mu(i)}p_{i,j}(x+\kappa-i-1)
=1μ(0)i=0κ1j=1Ni[(μ((i,j,1))μ((i,j,0)))(x+κi1)+μ((i,j,0))]\displaystyle=\frac{1}{\mu(0)}\sum_{i=0}^{\kappa-1}\sum_{j=1}^{N_{i}}\left[(\mu((i,j,1))-\mu((i,j,0)))(x+\kappa-i-1)+\mu((i,j,0))\right]
=1μ(0)[Ax+B],x.\displaystyle=\frac{1}{\mu(0)}\left[Ax+B\right],\qquad x\in\mathbb{R}.

By Theorem 5.5, the polynomial p02[x]p_{0}\in\mathbb{R}_{2}[x] satisfying (5.1) with v=0v=0 takes the form p0(x)=c2x2+c1x+c0p_{0}(x)=c_{2}x^{2}+c_{1}x+c_{0}, where (c2,c1,c0)(c_{2},c_{1},c_{0}) is the solution of the system

(5.19) [2κ00κ2κ0001][c2c1c0]=[1μ(0)A1μ(0)B1]\begin{bmatrix}2\kappa&0&0\\ \kappa^{2}&\kappa&0\\ 0&0&1\end{bmatrix}\begin{bmatrix}c_{2}\\ c_{1}\\ c_{0}\end{bmatrix}=\begin{bmatrix}\frac{1}{\mu(0)}A\\ \frac{1}{\mu(0)}B\\ 1\end{bmatrix}

Solving the above system of equations, we obtain that p0=1μ(0)q0p_{0}=\frac{1}{\mu(0)}q_{0}. From (4.2) and (4.1) it follows that for every i[1,κ1]i\in\mathbb{N}\cap[1,\kappa-1],

q0(i)=μ(0)S𝝀ie02=μ(i)+=0i1j=1Nμ((,j,i1)).\displaystyle q_{0}(i)=\mu(0)\lVert S_{\boldsymbol{\lambda}}^{i}e_{0}\rVert^{2}=\mu(i)+\sum_{\ell=0}^{i-1}\sum_{j=1}^{N_{\ell}}\mu((\ell,j,i-\ell-1)).

After rearrangement we get (5.17).
(ii)\Longrightarrow(i). Using (5.16), it can be verified that for every i[0,κ1]i\in\mathbb{N}\cap[0,\kappa-1], j[0,Ni]j\in\mathbb{N}\cap[0,N_{i}] the polynomial

pi,j,0(x)=(λ(i,j,1)21)x+1,x,p_{i,j,0}(x)=(\lambda_{(i,j,1)}^{2}-1)x+1,\qquad x\in\mathbb{R},

satisfies (5.1) for v=(i,j,0)v=(i,j,0). Next, from the equality

S𝝀Tne(i,j,k)2=S𝝀Tn+ke(i,j,0)2=1kλ(i,j,)2,k1,\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{(i,j,k)}\rVert^{2}=\frac{\lVert S_{\boldsymbol{\lambda}_{T}}^{n+k}e_{(i,j,0)}\rVert^{2}}{\prod_{\ell=1}^{k}\lambda_{(i,j,\ell)}^{2}},\qquad k\in\mathbb{N}_{1},

it follows that for every i[0,κ1]i\in\mathbb{N}\cap[0,\kappa-1], j[0,Ni]j\in\mathbb{N}\cap[0,N_{i}] and k1k\in\mathbb{N}_{1} the polynomial

pi,j,k(x)=(=1kλi,j,2)1pi,j,0(x+k),x,p_{i,j,k}(x)=\left(\prod_{\ell=1}^{k}\lambda_{i,j,\ell}^{2}\right)^{-1}p_{i,j,0}(x+k),\qquad x\in\mathbb{R},

satisfies (5.1) with v=(i,j,k)v=(i,j,k). Assume now κ>1\kappa>1. We check that rankA~3,κ=rankB~3,κ\mathrm{rank}\,\tilde{A}_{3,\kappa}=\mathrm{rank}\,\tilde{B}_{3,\kappa}. Set p0=1μ(0)q0p_{0}=\frac{1}{\mu(0)}q_{0}. As in the proof of converse implication, we obtain that the coefficients of p0p_{0} satisfies (5.19). Combining (5.17) with (4.2) and (4.1) we deduce that p0p_{0} satisfies

S𝝀Tne02=p0(n),n[0,κ1].\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{0}\rVert^{2}=p_{0}(n),\qquad n\in\mathbb{N}\cap[0,\kappa-1].

This implies that the coefficients of p0p_{0} satisfies also the following system of equations:

[A3,κB3,κ][c2c1c0]=[1μ(0)A1μ(0)B1a1aκ1,]\begin{bmatrix}A_{3,\kappa}\\ B_{3,\kappa}\end{bmatrix}\begin{bmatrix}c_{2}\\ c_{1}\\ c_{0}\end{bmatrix}=\begin{bmatrix}\frac{1}{\mu(0)}A\\ \frac{1}{\mu(0)}B\\ 1\\ a_{1}\\ \vdots\\ a_{\kappa-1},\end{bmatrix}

where a=S𝝀Te02a_{\ell}=\lVert S_{\boldsymbol{\lambda}_{T}}^{\ell}e_{0}\rVert^{2}, [1,κ1]\ell\in\mathbb{N}\cap[1,\kappa-1]. From the Kronecker-Capelli theorem it follows that rankA~3,κ=rankB~3,κ\mathrm{rank}\,\tilde{A}_{3,\kappa}=\mathrm{rank}\,\tilde{B}_{3,\kappa}. The application of Theorem 5.5 completes the proof. ∎

6. Subnormality of Cauchy dual of composition operator

In [18] the authors studied the Cauchy dual subnormality problem for 2-isometric composition operators with the property that the graph GTG_{T} has one cycle and one branching point and obtained the solution in case of graphs with cycle of length 1. In this section we investigate the Cauchy dual subnormality problem for 2-isometric composition operators with graphs having one cycle in full generality. Again, we stick to the notation of Theorem 3.5.(ii). First, we are going to derive formulas for S𝝀TS_{\boldsymbol{\lambda}_{T}}^{\ast} and S𝝀TS_{\boldsymbol{\lambda}_{T}}^{\prime}.

Lemma 6.1.

Suppose T:XXT\!:X\to X is such that GTG_{T} is connected and satisfies Theorem 3.5.(ii). Suppose hTL(μ)h_{T}\in L^{\infty}(\mu). Then

  1. (i)

    S𝝀Tev=λvewS_{\boldsymbol{\lambda}_{T}}^{\ast}e_{v}=\lambda_{v}e_{w} for every vVTv\in V_{T}, where wVTw\in V_{T} is the only vertex satisfying (w,v)ET(w,v)\in E_{T}, that is S𝝀TS_{\boldsymbol{\lambda}_{T}}^{\ast} is the weighted shift on (VT,ET1)(V_{T},E_{T}^{-1}), where ET1={(u,v)VT×VT:(v,u)ET}E_{T}^{-1}=\{(u,v)\in V_{T}\times V_{T}\!:(v,u)\in E_{T}\},

  2. (ii)

    S𝝀TS𝝀Tev=S𝝀Tev2evS_{\boldsymbol{\lambda}_{T}}^{\ast}S_{\boldsymbol{\lambda}_{T}}e_{v}=\lVert S_{\boldsymbol{\lambda}_{T}}e_{v}\rVert^{2}e_{v} for every vVTv\in V_{T},

  3. (iii)

    if S𝝀TS_{\boldsymbol{\lambda}_{T}} is left-invertible, then

    S𝝀Tev=uVT(v,u)ETcvλueu,S_{\boldsymbol{\lambda}_{T}}^{\prime}e_{v}=\sum_{\begin{subarray}{c}u\in V_{T}\\ (v,u)\in E_{T}\end{subarray}}c_{v}\lambda_{u}e_{u},

    where cv=S𝝀Tev2c_{v}=\lVert S_{\boldsymbol{\lambda}_{T}}e_{v}\rVert^{-2}, vVTv\in V_{T}; in particular, S𝝀TS_{\boldsymbol{\lambda}_{T}}^{\prime} is the weighted shift on GTG_{T}.

Proof.

(i). Let vVv\in V. Then

S𝝀Tew,ev={λv,if (w,v)ET0,otherwise.\langle S_{\boldsymbol{\lambda}_{T}}e_{w},e_{v}\rangle=\begin{cases}\lambda_{v},&\text{if }(w,v)\in E_{T}\\ 0,&\text{otherwise}\end{cases}.

Hence, S𝝀Tev,ew0\langle S_{\boldsymbol{\lambda}_{T}}^{\ast}e_{v},e_{w}\rangle\not=0 if and only if (w,v)ET(w,v)\in E_{T}. By Observation 3.1, there is only one vertex wVTw\in V_{T} satisfying (w,v)ET(w,v)\in E_{T}; for this vertex ww we have S𝝀Tev=λvewS_{\boldsymbol{\lambda}_{T}}^{\ast}e_{v}=\lambda_{v}e_{w}. (ii) and (iii) are simple consequences of (i). ∎

Note that if S𝝀TS_{\boldsymbol{\lambda}_{T}} is assumed to be 2-isometric, then, by [25, Lemma 1], it is also an expansion, so that cv(0,1]c_{v}\in(0,1], vVTv\in V_{T}.

Assume that GTG_{T} is connected and satisfies Theorem 3.5.(ii). Suppose S𝝀T𝐁(2(VT))S_{\boldsymbol{\lambda}_{T}}\in\mathbf{B}(\ell^{2}(V_{T})) is 22-isometric. For further applications let us emphasize two formulas (cf. Lemma 4.4). For every m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1] and nn\in\mathbb{N} we have

S𝝀Tnevm\displaystyle S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v_{m}} =Λm,nevm+nmodκ\displaystyle=\Lambda_{m,n}^{\prime}e_{v_{m+n\,\mathrm{mod}\,\kappa}}
+i=0n1cvimodκΛm,iRiR=m+imodκλωRS𝝀Tni1eωR\displaystyle+\sum_{i=0}^{n-1}c_{v_{i\,\mathrm{mod}\,\kappa}}\Lambda_{m,i}^{\prime}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=m+i\,\mathrm{mod}\,\kappa\end{subarray}}\lambda_{\omega_{R}}S_{\boldsymbol{\lambda}_{T}}^{\prime n-i-1}e_{\omega_{R}}

and

S𝝀Tn+κevm\displaystyle S_{\boldsymbol{\lambda}_{T}}^{\prime n+\kappa}e_{v_{m}} =Λm,κS𝝀Tnevm\displaystyle=\Lambda_{m,\kappa}^{\prime}S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v_{m}}
(6.1) +i=0κ1cvimodκΛm,iRiR=m+imodκλωRS𝝀Tn+κi1eωR,\displaystyle+\sum_{i=0}^{\kappa-1}c_{v_{i\,\mathrm{mod}\,\kappa}}\Lambda_{m,i}^{\prime}\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=m+i\,\mathrm{mod}\,\kappa\end{subarray}}\lambda_{\omega_{R}}S_{\boldsymbol{\lambda}_{T}}^{\prime n+\kappa-i-1}e_{\omega_{R}},

where

Λm,i=j=m+1m+iλvjmodκcv(j1)modκ,m[0,κ1],i.\Lambda_{m,i}^{\prime}=\prod_{j=m+1}^{m+i}\lambda_{v_{j\,\mathrm{mod}\,\kappa}}c_{v_{(j-1)\,\mathrm{mod}\,\kappa}},\qquad m\in\mathbb{N}\cap[0,\kappa-1],\ i\in\mathbb{N}.

The above formulas are consequences of Lemma 6.1.(iii) and the fact that for 2-isometric S𝝀TS_{\boldsymbol{\lambda}_{T}} every cvc_{v} (vVT{v0,,vκ1}v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\}) is equal 1 (see Lemma 5.4). Note also that Λm,κ=Λ0,κ\Lambda^{\prime}_{m,\kappa}=\Lambda^{\prime}_{0,\kappa} for every m[1,κ1]m\in\mathbb{N}\cap[1,\kappa-1].
It is well-known (see e.g. [5]) that the Cauchy dual of 2-isometric operator is always a contraction (in our setting it can be easily derived from Lemma 6.1.(iii)). This observation will be crucial in our considerations about subnormality of the Cauchy dual of S𝝀TS_{\boldsymbol{\lambda}_{T}}.

Note that if =\mathcal{R}=\varnothing, then 2(VT)\ell^{2}(V_{T}) is finite dimensional. In such a case every 2-isometric operator is automatically invertible and, by [3, Proposition 1.23], unitary. Since U=UU^{\prime}=U for every unitary operator UU, if =\mathcal{R}=\varnothing and S𝝀TS_{\boldsymbol{\lambda}_{T}} is 2-isometric, S𝝀TS_{\boldsymbol{\lambda}_{T}}^{\prime} is obviously subnormal. In the further investigation we always assume that \mathcal{R}\not=\varnothing. Let us present the main result in this section.

Theorem 6.2.

Let T:XXT\!:X\to X be such that GTG_{T} is connected, satisfies Theorem 3.5.(ii) with additional assumption that \mathcal{R}\not=\varnothing. Assume that hTL(μ)h_{T}\in L^{\infty}(\mu). If S𝛌𝐁(2(VT))S_{\boldsymbol{\lambda}}\in\mathbf{B}(\ell^{2}(V_{T})) is 22-isometric, then the following conditions are equivalent:

  1. (i)

    S𝝀TS_{\boldsymbol{\lambda}_{T}}^{\prime} is subnormal,

  2. (ii)

    for every m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1],

    S𝝀Tnevm2=[0,1]tndμvm(t),n[0,κ1],\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v_{m}}\rVert^{2}=\int_{[0,1]}t^{n}\,\mathrm{d}\mu_{v_{m}}(t),\qquad n\in\mathbb{N}\cap[0,\kappa-1],

with μvm=(1αm)δDκ+αmδ1\mu_{v_{m}}=(1-\alpha_{m})\delta_{\sqrt[\kappa]{D}}+\alpha_{m}\delta_{1}, where

D\displaystyle D =Λ0,κ2=i=0κ1cvi2,\displaystyle=\Lambda_{0,\kappa}^{\prime 2}=\prod_{i=0}^{\kappa-1}c_{v_{i}}^{2},
αm\displaystyle\alpha_{m} =Cm1D,m[0,κ1],\displaystyle=\frac{C_{m}}{1-D},\qquad m\in\mathbb{N}\cap[0,\kappa-1],
Cm\displaystyle C_{m} =i=0κ1cvimodκ2Λm,i2RiR=i+mmodκλωR2,m[0,κ1].\displaystyle=\sum_{i=0}^{\kappa-1}c_{v_{i\,\mathrm{mod}\,\kappa}}^{2}\Lambda_{m,i}^{\prime 2}\!\!\!\!\!\!\!\sum_{\begin{subarray}{c}R\in\mathcal{R}\\ i_{R}=i+m\,\mathrm{mod}\,\kappa\end{subarray}}\!\!\!\!\!\!\!\lambda_{\omega_{R}}^{2},\qquad m\in\mathbb{N}\cap[0,\kappa-1].
Proof.

(i) \Longrightarrow (ii). From [20, Theorems 3.1 and 2.2] and the fact that S𝝀TS_{\boldsymbol{\lambda}_{T}}^{\prime} is contractive it follows that for every vVTv\in V_{T} there exists the measure μv:([0,1])[0,1]\mu_{v}\!:\mathcal{B}([0,1])\to[0,1] such that

S𝝀Tnev2=[0,1]tndμv(t),n.\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v}\rVert^{2}=\int_{[0,1]}t^{n}\,\mathrm{d}\mu_{v}(t),\qquad n\in\mathbb{N}.

In turn, by Theorem 5.5 and Lemma 6.1, we have

S𝝀Tnev=1,vVT{v0,,vκ1},n;\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v}\rVert=1,\quad v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\},\ n\in\mathbb{N};

which implies that μv=δ1\mu_{v}=\delta_{1} for vVT{v0,,vκ1}v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\}. From this and (6) we derive that for every m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1],

S𝝀Tn+κevm2=DS𝝀Tnevm2+Cm,n.\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n+\kappa}e_{v_{m}}\rVert^{2}=D\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v_{m}}\rVert^{2}+C_{m},\qquad n\in\mathbb{N}.

Hence,

[0,1]tn+κdμvm(t)=D[0,1]tndμvm(t)+Cm,n.\int_{[0,1]}t^{n+\kappa}\,\mathrm{d}\mu_{v_{m}}(t)=D\int_{[0,1]}t^{n}\,\mathrm{d}\mu_{v_{m}}(t)+C_{m},\qquad n\in\mathbb{N}.

Since μvm\mu_{v_{m}} is regular, it follows that

(6.2) (tκD)dμvm(t)=Cmδ1,m[0,κ1].(t^{\kappa}-D)\,\mathrm{d}\mu_{v_{m}}(t)=C_{m}\delta_{1},\qquad m\in\mathbb{N}\cap[0,\kappa-1].

Since \mathcal{R}\not=\varnothing, by Corollary 5.7 we have that D<1D<1. This implies that for all m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1], μvm([0,Dκ)(Dκ,1))=0\mu_{v_{m}}([0,\sqrt[\kappa]{D})\cup(\sqrt[\kappa]{D},1))=0. Thus,

(6.3) μvm=(1αm)δDκ+αmδ1\mu_{v_{m}}=(1-\alpha_{m})\delta_{\sqrt[\kappa]{D}}+\alpha_{m}\delta_{1}

with some αm[0,1]\alpha_{m}\in[0,1]. In turn, for every m[0,κ1]m\in\mathbb{N}\cap[0,\kappa-1] we have

D(1αm)+αm\displaystyle D(1-\alpha_{m})+\alpha_{m} =(6.3)[0,1]tκdμvm\displaystyle\stackrel{{\scriptstyle\eqref{ProofFormMeasureOnCycleTwoAtomic}}}{{=}}\int_{[0,1]}t^{\kappa}\,\mathrm{d}\mu_{v_{m}}
=(6.2)Dμvm([0,1])+Cm=D+Cm.\displaystyle\stackrel{{\scriptstyle\eqref{ProofFormRecursiveMeasure}}}{{=}}D\mu_{v_{m}}([0,1])+C_{m}=D+C_{m}.

The above equality implies that αm=Cm1D\alpha_{m}=\frac{C_{m}}{1-D}.
(ii) \Longrightarrow (i). From Theorem 5.5 it follows that

S𝝀Tnev=1,vVT{v0,,vκ1},n,\lVert S_{\boldsymbol{\lambda}_{T}}^{n}e_{v}\rVert=1,\quad v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\},\ n\in\mathbb{N},

which implies, by Lemma 6.1, that

S𝝀Tnev=1,vVT{v0,,vκ1},n.\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v}\rVert=1,\quad v\in V_{T}\setminus\{v_{0},\ldots,v_{\kappa-1}\},\ n\in\mathbb{N}.

We check that

(6.4) S𝝀Tnevm2=[0,1]tndμvm(t),n,m[0,κ1].\displaystyle\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v_{m}}\rVert^{2}=\int_{[0,1]}t^{n}\,\mathrm{d}\mu_{v_{m}}(t),\qquad n\in\mathbb{N},m\in\mathbb{N}\cap[0,\kappa-1].

Since (6.4) holds for n[0,κ1]n\in\mathbb{N}\cap[0,\kappa-1], it is enough to check that if (6.4) holds for some nn\in\mathbb{N}, then it also holds for n+κn+\kappa. By (6) we have

S𝝀Tn+κevm2\displaystyle\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n+\kappa}e_{v_{m}}\rVert^{2} =DS𝝀Tnevm2+Cm=D[0,1]tndμvm(t)+Cm\displaystyle=D\lVert S_{\boldsymbol{\lambda}_{T}}^{\prime n}e_{v_{m}}\rVert^{2}+C_{m}=D\int_{[0,1]}t^{n}\,\mathrm{d}\mu_{v_{m}}(t)+C_{m}
=D(1αm)Dnκ+Dαm+Cm\displaystyle=D(1-\alpha_{m})\sqrt[\kappa]{D^{n}}+D\alpha_{m}+C_{m}
=(1αm)Dn+κκ+DCm1D+Cm\displaystyle=(1-\alpha_{m})\sqrt[\kappa]{D^{n+\kappa}}+\frac{DC_{m}}{1-D}+C_{m}
=(1αm)Dn+κκ+Cm1D\displaystyle=(1-\alpha_{m})\sqrt[\kappa]{D^{n+\kappa}}+\frac{C_{m}}{1-D}
=[0,1]tn+κdμvm(t).\displaystyle=\int_{[0,1]}t^{n+\kappa}\,\mathrm{d}\mu_{v_{m}}(t).

Using [20, Theorems 3.1 and 2.2] we obtain subnormality of S𝝀TS_{\boldsymbol{\lambda}_{T}}^{\prime}. ∎

In [18] the authors proved that the Cauchy dual of 2-isometric composition operator on graph with one cycle of length 1 and one branching point is subnormal. As a corollary, we get the same result in a more general version.

Corollary 6.3 (cf. [18, Theorem 4.6]).

Assume that GTG_{T} is connected and satisfies Theorem 3.5.(ii) with additional assumptions that \mathcal{R}\not=\varnothing and κ=1\kappa=1. If S𝛌T𝐁(2(VT))S_{\boldsymbol{\lambda}_{T}}\in\mathbf{B}(\ell^{2}(V_{T})) is 2-isometric, then S𝛌TS_{\boldsymbol{\lambda}_{T}}^{\prime} is subnormal.

In [5, Example 6.6] the authors gave an example of 2-isometry, for which the Cauchy dual is not subnormal. With the help of Theorems 5.5 and 6.2 we can give another example.

Example 6.4.

Let VT={0,1,2}({0,1,2}×)V_{T}=\{0,1,2\}\cup(\{0,1,2\}\times\mathbb{N}) and

ET\displaystyle E_{T} ={(0,1),(1,2),(2,0)}{(i,(i,0)):i{0,1,2}}\displaystyle=\{(0,1),(1,2),(2,0)\}\cup\{(i,(i,0))\!:i\in\{0,1,2\}\}
{((i,j),(i,j+1)):i{0,1,2},j}.\displaystyle\cup\{((i,j),(i,j+1))\!:i\in\{0,1,2\},j\in\mathbb{N}\}.

The graph GT=(VT,ET)G_{T}=(V_{T},E_{T}) is exactly the graph defined at the end of Section 5 taken with κ=3\kappa=3 and N0=N1=N2=1N_{0}=N_{1}=N_{2}=1 (see Figure 2). Set 𝛌={λ0,λ1,λ2}{λi,j:i{0,1,2},j}\boldsymbol{\lambda}=\{\lambda_{0},\lambda_{1},\lambda_{2}\}\cup\{\lambda_{i,j}\!:i\in\{0,1,2\},j\in\mathbb{N}\} as follows

λ1=a,λ2=1,λ0=1a,\displaystyle\lambda_{1}=a,\ \lambda_{2}=1,\ \lambda_{0}=\frac{1}{a},
λ0,0=λ1,0=1,λ2,0=z,\displaystyle\lambda_{0,0}=\lambda_{1,0}=1,\ \lambda_{2,0}=z,
λi,j=1,i{0,1,2},j1,\displaystyle\lambda_{i,j}=1,\qquad i\in\{0,1,2\},j\in\mathbb{N}_{1},

where a,z(0,)a,z\in(0,\infty). It is easy to verify that S𝛌𝐁(2(VT))S_{\boldsymbol{\lambda}}\in\mathbf{B}(\ell^{2}(V_{T})); moreover, S𝛌nev=1\lVert S_{\boldsymbol{\lambda}}^{n}e_{v}\rVert=1 for every v({0,1,2})×v\in(\{0,1,2\})\times\mathbb{N} and nn\in\mathbb{N}, so the polynomial pv=1p_{v}=1 satisfies (5.1) with v({0,1,2})×v\in(\{0,1,2\})\times\mathbb{N}. The polynomial q0[x]q\in\mathbb{R}_{0}[x] given by the formula (5.6) takes the form

q(t)=1+a2+a2z2,t.q(t)=1+a^{2}+a^{2}z^{2},\qquad t\in\mathbb{R}.

Hence, in view of Theorem 5.5, S𝛌S_{\boldsymbol{\lambda}} is 2-isometric if and only if rankA~2,3=rankB~2,3\mathrm{rank}\,\tilde{A}_{2,3}=\mathrm{rank}\,\tilde{B}_{2,3}. The latter holds if and only if the polynomial

p0(t)=13(1+a2+a2z2)t+1,t,p_{0}(t)=\frac{1}{3}\left(1+a^{2}+a^{2}z^{2}\right)t+1,\qquad t\in\mathbb{R},

satisfies the equalities p0(1)=S𝛌e02p_{0}(1)=\lVert S_{\boldsymbol{\lambda}}e_{0}\rVert^{2} and p0(2)=S𝛌2e02p_{0}(2)=\lVert S_{\boldsymbol{\lambda}}^{2}e_{0}\rVert^{2}, which is equivalent to the following system of equations

(6.5) a2+1\displaystyle a^{2}+1 =13(1+a2+a2z2)+1\displaystyle=\frac{1}{3}\left(1+a^{2}+a^{2}z^{2}\right)+1
(6.6) 2a2+1\displaystyle 2a^{2}+1 =23(1+a2+a2z2)+1\displaystyle=\frac{2}{3}\left(1+a^{2}+a^{2}z^{2}\right)+1

Assume S𝛌S_{\boldsymbol{\lambda}} is 2-isometric. By Theorem 6.2, if S𝛌S_{\boldsymbol{\lambda}}^{\prime} is subnormal, then there exists α(0,1)\alpha\in(0,1) satisfying

{αD3+1α=S𝝀e02αD23+1α=S𝝀2e02,\begin{cases}\alpha\sqrt[3]{D}+1-\alpha=\lVert S_{\boldsymbol{\lambda}}^{\prime}e_{0}\rVert^{2}\\ \alpha\sqrt[3]{D^{2}}+1-\alpha=\lVert S_{\boldsymbol{\lambda}}^{\prime 2}e_{0}\rVert^{2}\end{cases},

which implies that

(6.7) S𝝀e021D31=S𝝀2e021D231.\frac{\lVert S_{\boldsymbol{\lambda}}^{\prime}e_{0}\rVert^{2}-1}{\sqrt[3]{D}-1}=\frac{\lVert S_{\boldsymbol{\lambda}}^{\prime 2}e_{0}\rVert^{2}-1}{\sqrt[3]{D^{2}}-1}.

Note that

S𝝀e02\displaystyle\lVert S_{\boldsymbol{\lambda}}^{\prime}e_{0}\rVert^{2} =1a2+1\displaystyle=\frac{1}{a^{2}+1}
S𝝀2e02\displaystyle\lVert S_{\boldsymbol{\lambda}}^{\prime 2}e_{0}\rVert^{2} =a22(a2+1)2+1(a2+1)2\displaystyle=\frac{a^{2}}{2(a^{2}+1)^{2}}+\frac{1}{(a^{2}+1)^{2}}

and

D=14(a2+1)2(a2+z2)2.D=\frac{1}{4(a^{2}+1)^{2}(a^{-2}+z^{2})^{2}}.

To obtain 2-isometric operator S𝛌S_{\boldsymbol{\lambda}}, the Cauchy dual of which is not subnormal, it is enough to find a,z(0,)a,z\in(0,\infty), which satisfy (6.5) and (6.6) and does not satisfy (6.7); this holds if we take, for instance, a=27a=\frac{2}{\sqrt{7}}, z=12z=\frac{1}{2}.

Statements and Declarations

Funding. The author of the publication received an incentive scholarship from the funds of the program Excellence Initiative - Research University at the Jagiellonian University in Kraków.

Competing interests. The authors have no conflicts of interest to declare that are relevant to the content of this article.

Acknowledgements

The author would like to thank Zenon Jabłoński for his helpful suggestions concerning the final version of this paper. We are also grateful to anonymous reviewer for the comments improving the exposition quality of the paper.

References

  • [1] Belal Abdullah and Trieu Le. The structure of mm-isometric weighted shift operators. Oper. Matrices, 10(2):319–334, 2016.
  • [2] Jim Agler. A disconjugacy theorem for Toeplitz operators. Amer. J. Math., 112(1):1–14, 1990.
  • [3] Jim Agler and Mark Stankus. mm-isometric transformations of Hilbert space. I. Integral Equations Operator Theory, 21(4):383–429, 1995.
  • [4] A.V. Aho, J.E. Hopcroft, and J.D. Ullman. Data Structures and Algorithms. Adison-Wesley, Reading, 1983.
  • [5] Akash Anand, Sameer Chavan, Zenon Jan Jabłoński, and Jan Stochel. A solution to the Cauchy dual subnormality problem for 2-isometries. J. Funct. Anal., 277(12):108292, 51, 2019.
  • [6] Akash Anand, Sameer Chavan, Zenon Jan Jabłoński, and Jan Stochel. The Cauchy dual subnormality problem for cyclic 2-isometries. Adv. Oper. Theory, 5(3):1061–1077, 2020.
  • [7] Ameer Athavale. On completely hyperexpansive operators. Proc. Amer. Math. Soc., 124(12):3745–3752, 1996.
  • [8] Catalin Badea and Laurian Suciu. The Cauchy dual and 2-isometric liftings of concave operators. J. Math. Anal. Appl., 472(2):1458–1474, 2019.
  • [9] S. Banach. Théorie des opérations linéaires, volume 1 of Monogr. Mat., Warszawa. PWN - Panstwowe Wydawnictwo Naukowe, Warszawa, 1932.
  • [10] Teresa Bermúdez, Antonio Martinón, and Emilio Negrín. Weighted shift operators which are mm-isometries. Integral Equations Operator Theory, 68(3):301–312, 2010.
  • [11] Teresa Bermúdez, Antonio Martinón, and Juan Agustín Noda. Products of mm-isometries. Linear Algebra Appl., 438(1):80–86, 2013.
  • [12] Piotr Budzyński, Zenon Jan Jabłoński, Il Bong Jung, and Jan Stochel. Subnormality of unbounded composition operators over one-circuit directed graphs: exotic examples. Adv. Math., 310:484–556, 2017.
  • [13] Sameer Chavan. On operators Cauchy dual to 2-hyperexpansive operators. Proc. Edinb. Math. Soc. (2), 50(3):637–652, 2007.
  • [14] Sameer Chavan, Soumitra Ghara, and Md. Ramiz Reza. The Cauchy dual subnormality problem via de Branges–Rovnyak spaces. Studia Math., 265(3):315–341, 2022.
  • [15] Zenon Jan Jabłoński, Il Bong Jung, and Jan Stochel. A non-hyponormal operator generating Stieltjes moment sequences. J. Funct. Anal., 262(9):3946–3980, 2012.
  • [16] Zenon Jan Jabłoński, Il Bong Jung, and Jan Stochel. Weighted shifts on directed trees. Mem. Amer. Math. Soc., 216(1017):viii+106, 2012.
  • [17] Zenon Jan Jabłoński, Il Bong Jung, and Jan Stochel. mm-isometric operators and their local properties. Linear Algebra Appl., 596:49–70, 2020.
  • [18] Zenon Jan Jabłoński and Jakub Kośmider. mm-isometric composition operators on directed graphs with one circuit. Integral Equations Operator Theory, 93(5):Paper No. 50, 26, 2021.
  • [19] Eungil Ko and Jongrak Lee. On mm-isometric Toeplitz operators. Bull. Korean Math. Soc., 55(2):367–378, 2018.
  • [20] Alan Lambert. Subnormality and weighted shifts. J. London Math. Soc. (2), 14(3):476–480, 1976.
  • [21] Singh A. Nair, M. Thamban. Linear Algebra. Springer, 2018.
  • [22] Eric A. Nordgren. Composition operators on Hilbert spaces. In Hilbert space operators (Proc. Conf., Calif. State Univ., Long Beach, Calif., 1977), volume 693 of Lecture Notes in Math., pages 37–63. Springer, Berlin, 1978.
  • [23] O. Ore. Theory of Graphs, volume XXXVIII of Amer. Math. Soc. Colloq. Publ. American Mathematical Society, Providence, 1962.
  • [24] PawełPietrzycki. A Shimorin-type analytic model on an annulus for left-invertible operators and applications. J. Math. Anal. Appl., 477(2):885–911, 2019.
  • [25] Stefan Richter. Invariant subspaces of the Dirichlet shift. J. Reine Angew. Math., 386:205–220, 1988.
  • [26] Sergei Shimorin. Wold-type decompositions and wandering subspaces for operators close to isometries. J. Reine Angew. Math., 531:147–189, 2001.
  • [27] R. K. Singh and J. S. Manhas. Composition operators on weighted spaces of vector-valued continuous functions. Acta Sci. Math. (Szeged), 58(1-4):453–472, 1993.
  • [28] J.B. Stochel. Weighted quasishifts, generalized commutation relation and subnormality. Ann. Univ. Sarav. Ser. Math., 3(3):i–iv and 109–128, 1990.
  • [29] Laurian Suciu. Operators with expansive mm-isometric liftings. Monatsh. Math., 198(1):165–187, 2022.