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

Monomial Operators

Jim Agler and John E. McCarthy Partially supported by National Science Foundation Grant DMS 2054199
Abstract

We study monomial operators on L2[0,1]L^{2}[0,1], that is bounded linear operators that map each monomial xnx^{n} to a multiple of xpnx^{p_{n}} for some pnp_{n}. We show that they are all unitarily equivalent to weighted composition operators on a Hardy space. We characterize what sequences pnp_{n} can arise. In the case that pnp_{n} is a fixed translation of nn, we give a criterion for boundedness of the operator.

1 Introduction

When studying polynomial approximation in L2[0,1]L^{2}[0,1], the following class of operators arises naturally.

Definition 1.1.

A monomial operator is a bounded linear operator T:L2[0,1]L2[0,1]T:L^{2}[0,1]\to L^{2}[0,1] with the property that there exist constants cnc_{n} and pnp_{n} so that

T:xncnxpnn.T:x^{n}\mapsto c_{n}\,x^{p_{n}}\qquad\forall n\in{\mathbb{N}}. (1.2)

If, in addition, there is some τ\tau so that pn=n+τp_{n}=n+\tau for every nn, we call it a flat monomial operator.

The powers pnp_{n} may be complex, but must lie in the half plane

𝕊:={s:Re(s)>12},{\mathbb{S}}:=\{s\in\mathbb{C}:\operatorname{Re}(s)>-\frac{1}{2}\},

in order for xpnx^{p_{n}} to lie in L2L^{2}. Well-known examples of monomial operators include the Hardy operator HH

H:f\displaystyle H:f \displaystyle\mapsto 1x0xf(t)𝑑t\displaystyle\frac{1}{x}\int_{0}^{x}f(t)dt
Hxn\displaystyle Hx^{n} =\displaystyle= 1n+1xn;\displaystyle\frac{1}{n+1}x^{n};

the operator MxM_{x} of multiplication by xx; and the Volterra operator

V=MxH:f\displaystyle V=M_{x}H:f \displaystyle\mapsto 0xf(t)𝑑t\displaystyle\int_{0}^{x}f(t)dt
Vxn\displaystyle Vx^{n} =\displaystyle= 1n+1xn+1.\displaystyle\frac{1}{n+1}x^{n+1}.

In [2] the authors studied flat monomial operators, and showed that they all leave invariant every subspace of the form

{fL2[0,1]:f=0on[0,t]}.\{f\in L^{2}[0,1]:f=0{\rm\ on\ }[0,t]\}.

It was shown independently by Brodskii [6] and Donoghue [10] that these subspaces are exactly the invariant subspaces for the Volterra operator.

It is the purpose of this note to examine general monomial operators of the form (1.2). To describe them, it is convenient to introduce a Hardy space associated with 𝕊{\mathbb{S}}. The Hardy space of the unit disk, which we denote H2(𝔻)H^{2}(\mathbb{D}), is the Hilbert space of holomorphic functions on 𝔻\mathbb{D} with finite norm, where the norm is given by

ϕH2(𝔻)2=sup0<r<112π02π|ϕ(reiθ)|2𝑑θ.\|\phi\|^{2}_{H^{2}(\mathbb{D})}\ =\ \sup_{0<r<1}\frac{1}{2\pi}\int_{0}^{2\pi}|\phi(re^{i\theta})|^{2}d\theta.

There are two distinct definitions for the Hardy space of a half-plane. Let λ\lambda be the linear fractional transformation give by

λ(s)=ss+1,\lambda(s)=\frac{s}{s+1}, (1.3)

that takes 𝕊{\mathbb{S}} onto 𝔻\mathbb{D}.

Definition 1.4.

By H2(𝕊)H^{2}({\mathbb{S}}) we mean {ϕλ:ϕH2(𝔻)}\{\phi\circ\lambda:\phi\in H^{2}(\mathbb{D})\}, and the norm is defined so that ϕϕλ\phi\mapsto\phi\circ\lambda is unitary.

An equivalent definition of H2(𝕊)H^{2}({\mathbb{S}}) is the set of functions ff that are holomorphic in 𝕊{\mathbb{S}} and such that |f|2|f|^{2} has a harmonic majorant there. The norm squared is equal to the value of this harmonic majorant at 0. If fH2(𝕊)f\in H^{2}({\mathbb{S}}), then it has boundary values a.e. (see e.g. [11]), and the norm is given by

f2=12π|f(12+it)|2t2+14𝑑t.\|f\|^{2}\ =\ \frac{1}{2\pi}\int_{-\infty}^{\infty}\frac{|f(-\frac{1}{2}+it)|^{2}}{t^{2}+\frac{1}{4}}dt. (1.5)

The reproducing kernel for H2(𝕊)H^{2}({\mathbb{S}}) is given by

k(s,u)=ku,ks=(1+s)(1+u¯)1+s+u¯.k(s,u)\ =\ \langle k_{u},k_{s}\rangle\ =\ \frac{(1+s)(1+\bar{u})}{1+s+\bar{u}}. (1.6)

In light of (1.6), the map

U:L2[0,1]\displaystyle U:L^{2}[0,1] \displaystyle\to H2(𝕊)\displaystyle H^{2}({\mathbb{S}})
xs\displaystyle x^{s} \displaystyle\mapsto 11+sks¯\displaystyle\frac{1}{1+s}k_{\bar{s}}

extends to a unitary from L2[0,1]L^{2}[0,1] onto H2(𝕊)H^{2}({\mathbb{S}}). In [1] we show that for any fL2f\in L^{2}, we have

Uf(s)=(1+s)01f(x)xs𝑑x.Uf(s)=(1+s)\int_{0}^{1}f(x)x^{s}dx.

If TT is a monomial operator, we shall define

T^=UTU:H2(𝕊)H2(𝕊).\widehat{T}\ =\ UTU^{*}:H^{2}({\mathbb{S}})\to H^{2}({\mathbb{S}}). (1.7)

One might wonder whether the definition of monomial operator should require that it take xsx^{s} to some multiple of a monomial for every s𝕊s\in{\mathbb{S}}; our first theorem asserts that this always happens, just by assuming it on the natural numbers. It also shows that, after moving to the Hardy space as in (1.7), monomial operators correspond to the adjoints of weighted composition operators.

If gg is a holomorphic function on some domain, let us define

g(s)=g(s¯)¯.g^{\cup}(s)\ =\ \overline{g(\overline{s})}.

By (1.5) we see that if fH2(𝕊)f\in H^{2}({\mathbb{S}}), then fH2(𝕊)=fH2(𝕊)\|f^{\cup}\|_{H^{2}({\mathbb{S}})}=\|f\|_{H^{2}({\mathbb{S}})}.

Theorem 1.8.

Let T:L2[0,1]L2[0,1]T:L^{2}[0,1]\to L^{2}[0,1] be a monomial operator given by (1.2). Then there exists a holomorphic map β:𝕊𝕊\beta:{\mathbb{S}}\to{\mathbb{S}} and a function gH2(𝕊)g\in H^{2}({\mathbb{S}}) so that, for every s𝕊s\in{\mathbb{S}}, we have

T(xs)=1+β(s)1+sg(s)xβ(s).T(x^{s})\ =\ \frac{1+\beta(s)}{1+s}g(s)\ x^{\beta(s)}. (1.9)

Moreover, we have

T^f(s)=g(s)f(β(s))fH2(𝕊).\widehat{T}^{*}f(s)\ =\ g^{\cup}(s)f(\beta^{\cup}(s))\qquad\forall f\in H^{2}({\mathbb{S}}). (1.10)

Equation (1.10) says

T^=MgCβ,\widehat{T}^{*}\ =\ M_{g^{\cup}}C_{\beta^{\cup}},

where CβC_{\beta} denotes the composition operator ffβf\mapsto f\circ\beta, and MgM_{g} denotes the multiplication operator fgff\mapsto gf.

Weighted composition operators have been studied for some time. See e.g. [12, 8, 15, 9, 5, 4, 7]. By Littlewood’s subordination principle, the operator CβC_{\beta} is always bounded whenever β\beta is a holomorphic self-map of 𝕊{\mathbb{S}} [16, Thm. 10.4.2]. A multiplication operator MhM_{h} is bounded if and only if hH(𝕊)h\in H^{\infty}({\mathbb{S}}). It was observed in [14] that it is possible for the product MhCβM_{h}C_{\beta} to be bounded even when MhM_{h} is not.

A consequence of Theorem 1.8 is that it allows us to specify what sequences pnp_{n} can occur in (1.2).

Corollary 1.11.

Let (pn)(p_{n}) be a sequence in 𝕊{\mathbb{S}}. Then there exists some choice of scalars cnc_{n}, not all zero, so that T:xncnxpnT:x^{n}\mapsto c_{n}x^{p_{n}} extends to be a bounded linear operator on L2[0,1]L^{2}[0,1] if and only if

[pm+pn¯+1m+n+1] 0.\left[\frac{p_{m}+\overline{p_{n}}+1}{m+n+1}\right]\ \geq\ 0.

In Section 3 we study which pairs of functions β\beta and gg give rise to a bounded operator in (1.9). We answer this question only for flat monomial operators, i.e. when β(s)=s+τ\beta(s)=s+\tau for some constant τ\tau. If (τ)<0\Re(\tau)<0, it follows from Corollary 1.11 that such a TT can never be bounded.

If (τ)0\Re(\tau)\geq 0, we have:

Theorem 1.12.

Let (τ)0\Re(\tau)\geq 0, and let TT be defined by

T(xn)=1+n+τ1+ng(n)xn+τT(x^{n})\ =\ \frac{1+n+\tau}{1+n}g(n)\ x^{n+\tau}

for some function gH2(𝕊)g\in H^{2}({\mathbb{S}}).

(i) If (τ)>0\Re(\tau)>0, then TT extends to be a bounded linear operator from L2[0,1]L^{2}[0,1] to L2[0,1]L^{2}[0,1] if and only if the Poisson integral of |g|2|g|^{2} is bounded on all half-planes that are strictly contained in 𝕊{\mathbb{S}}.

(ii) If (τ)=0\Re(\tau)=0, then TT is bounded if and only if gg is bounded on 𝕊{\mathbb{S}}.

This theorem is proved in Theorems 3.3 and 3.6.

2 Proof of Theorem 1.8

Proof of Theorem 1.8. Step 1. Let T^\widehat{T} be given by (1.7). Define gg by g=T^1g^{\cup}=\widehat{T}^{*}1. We have

T^kn\displaystyle\widehat{T}k_{n} =\displaystyle\ =\ UT(1+n)xn\displaystyle UT(1+n)x^{n}
=\displaystyle= (1+n)Ucnxpn\displaystyle(1+n)Uc_{n}x^{p_{n}}
=\displaystyle= cn1+n1+pnkpn¯.\displaystyle c_{n}\frac{1+n}{1+p_{n}}k_{\overline{p_{n}}}.

Since k0=1k_{0}=1, we have

g,kn\displaystyle\langle g^{\cup},k_{n}\rangle =\displaystyle\ =\ 1,cn1+n1+pnkpn¯\displaystyle\langle 1,c_{n}\frac{1+n}{1+p_{n}}k_{\overline{p_{n}}}\rangle
=\displaystyle= cn¯1+n1+pn¯.\displaystyle\overline{c_{n}}\frac{1+n}{1+\overline{p_{n}}}.

This gives

g(n)=cn1+n1+pn,g(n)\ =\ c_{n}\frac{1+n}{1+p_{n}},

and so we can write

T^kn=g(n)kpn¯.\widehat{T}k_{n}\ =\ g(n)k_{\overline{p_{n}}}. (2.1)

For any u𝕊u\in{\mathbb{S}}, let hu=T^kuh_{u}=\widehat{T}^{*}k_{u}. We get from (2.1)

hu,kn\displaystyle\langle h_{u},k_{n}\rangle =\displaystyle\ =\ g(n)ku,kpn¯\displaystyle g^{\cup}(n)\langle k_{u},k_{\overline{p_{n}}}\rangle
hu(n)\displaystyle\Rightarrow\quad h_{u}(n) =\displaystyle= g(n)(1+u¯)(1+pn¯)1+u¯+pn¯\displaystyle g^{\cup}(n)\frac{(1+\bar{u})(1+\overline{p_{n}})}{1+\bar{u}+\overline{p_{n}}} (2.2)
hu(n)\displaystyle\Rightarrow\quad h_{u}^{\cup}(n) =\displaystyle\ =\ g(n)(1+u)(1+pn)1+u+pn.\displaystyle g(n)\frac{(1+u)(1+{p_{n}})}{1+u+{p_{n}}}. (2.3)

Define β\beta by

β(s)=(1+u)(hu(s)g(s))(1+u)g(s)hu(s).\beta(s)\ =\ \frac{(1+u)(h_{u}^{\cup}(s)-g(s))}{(1+u)g(s)-h_{u}^{\cup}(s)}. (2.4)

As gg and huh_{u} are both in the Hardy space H2(𝕊)H^{2}({\mathbb{S}}), a priori we know that β\beta is in the Nevanlinna class of meromorphic functions on 𝕊{\mathbb{S}}, the class of quotients of H2H^{2} functions. Moreover, it follows from (2.3) that

pn=(1+u)(hu(n)g(n))(1+u)g(n)hu(n)=β(n).p_{n}\ =\ \frac{(1+u)(h_{u}^{\cup}(n)-g(n))}{(1+u)g(n)-h_{u}^{\cup}(n)}\ =\ \beta(n).

Observe that {\mathbb{N}} is not a zero set for H2(𝕊)H^{2}({\mathbb{S}})—indeed, (λ(n)=nn+1)(\lambda(n)=\frac{n}{n+1}) is not a Blaschke sequence for H2(𝔻)H^{2}(\mathbb{D}). Therefore it is a set of uniqueness for the Nevanlinna class, and hence β\beta is the unique function in the class that satisfies β(n)=pn\beta(n)=p_{n}. In particular, the definition (2.4) is actually independent of uu.

We can write (2.2) as

(T^ku)(n)=g(n)ku(β(n))n.(\widehat{T}^{*}k_{u})(n)\ =\ g^{\cup}(n)k_{u}(\beta^{\cup}(n))\qquad\forall n\in{\mathbb{N}}. (2.5)

Since {\mathbb{N}} is a set of uniqueness, (2.5) holds everywhere

(T^ku)(s)=g(s)ku(β(s))s𝕊.(\widehat{T}^{*}k_{u})(s)\ =\ g^{\cup}(s)k_{u}(\beta^{\cup}(s))\qquad\forall s\in{\mathbb{S}}. (2.6)

Step 2. We must show that β:𝕊𝕊\beta:{\mathbb{S}}\to{\mathbb{S}}. Suppose there is some point s𝕊s\in{\mathbb{S}} where g(s)0g(s)\neq 0 and β(s)=w\beta^{\cup}(s)=w is in 𝕊\mathbb{C}\setminus{\mathbb{S}}. Then there is a sequence qnq_{n} such that each qnq_{n} is a finite linear combination of kernel functions, qn1\|q_{n}\|\leq 1, and qn(w)q_{n}(w)\to\infty. By (2.6), we have

(T^qn)(s)=g(s)qn(w).(\widehat{T}^{*}q_{n})(s)\ =\ g^{\cup}(s)q_{n}(w). (2.7)

The right-hand side of (2.7) tends to infinity, the left-hand side is bounded by Tks\|T\|\|k_{s}\|, a contradiction. Since β\beta^{\cup} is meromorphic, we conclude that whenever g(s)0g^{\cup}(s)\neq 0 then β(s)𝕊\beta(s)\in{\mathbb{S}}. Therefore any singularities of β\beta^{\cup} on the zero set of gg^{\cup} are removable, so we conclude that β\beta^{\cup}, and hence also β\beta, is a self-map of 𝕊{\mathbb{S}}.

Thus we have proved (1.10) for any finite linear combination of kernel functions, and so, by a limiting argument, it is true for all fH2(𝕊)f\in H^{2}({\mathbb{S}}).

Step 3. From (2.6), we have

T^ks,ku\displaystyle\langle\widehat{T}k_{s},k_{u}\rangle =\displaystyle= T^ku,ks¯\displaystyle\overline{\langle\widehat{T}^{*}k_{u},k_{s}\rangle}
=\displaystyle\ =\ g(s¯)ku(β(s))¯\displaystyle g(\bar{s})\overline{k_{u}(\beta^{\cup}(s))}
=\displaystyle= g(s¯)(1+u)(1+β(s¯))1+u+β(s¯))\displaystyle g(\bar{s})\frac{(1+u)(1+\beta(\bar{s}))}{1+u+\beta(\bar{s}))}
=\displaystyle= g(s¯)kβ(s)(u).\displaystyle g(\bar{s})k_{\beta^{\cup}(s)}(u).

So

T^ks=g(s¯)kβ(s).\widehat{T}k_{s}\ =\ g(\bar{s})k_{\beta^{\cup}(s)}. (2.8)

Therefore

Txs\displaystyle Tx^{s} =\displaystyle\ =\ UT^U[xs]\displaystyle U^{*}\widehat{T}U[x^{s}]
=\displaystyle= UT^[11+sks¯]\displaystyle U^{*}\widehat{T}[\frac{1}{1+s}k_{\bar{s}}]
=\displaystyle= U[11+sg(s)kβ(s)¯]\displaystyle U^{*}[\frac{1}{1+s}g(s)k_{\overline{\beta(s)}}]
=\displaystyle= 1+β(s)1+sg(s)xβ(s).\displaystyle\frac{1+\beta(s)}{1+s}g(s)x^{\beta(s)}.

This proves (1.9). \Box

Proof of Corollary 1.11. By Theorem 1.8, a necessary condition for the existence of a non-zero bounded TT that maps each xnx^{n} to a multiple of xpnx^{p_{n}} is that there be some holomorphic self-map β\beta of 𝕊{\mathbb{S}} that maps nn to pnp_{n}. This condition is also sufficient, since choosing

cn=1+pn1+nc_{n}\ =\ \frac{1+p_{n}}{1+n}

gives UTUUTU^{*} is the adjoint of CβC_{\beta^{\cup}}, which is bounded.

When is there a map β:𝕊𝕊\beta:{\mathbb{S}}\to{\mathbb{S}} that interpolates nn to pnp_{n}? Composing with the Riemann map λ\lambda from (1.3), this is equivalent to asking when there exists ϕ=λβλ1\phi=\lambda\beta\lambda^{-1} from 𝔻\mathbb{D} to 𝔻\mathbb{D} that maps nn+1\frac{n}{n+1} to pnpn+1\frac{p_{n}}{p_{n}+1}. By Pick’s theorem [3, Thm. 1.81], this occurs if and only if

[1pmpm+1pn¯pn¯+11mm+1n¯n¯+1] 0.\left[\frac{1-\frac{p_{m}}{p_{m}+1}\frac{\overline{p_{n}}}{\overline{p_{n}}+1}}{1-\frac{m}{m+1}\frac{\bar{n}}{\bar{n}+1}}\right]\ \geq\ 0. (2.9)

Rearranging (2.9), we get

[m+1pm+1n¯+1pn¯+1pm+pn¯+1m+n¯+1] 0.\left[\frac{m+1}{p_{m}+1}\frac{\bar{n}+1}{\overline{p_{n}}+1}\frac{p_{m}+\overline{p_{n}}+1}{m+\bar{n}+1}\right]\ \geq\ 0.

As conjugating by the rank one operator

[m+1pm+1n¯+1pn¯+1]\left[\frac{m+1}{p_{m}+1}\frac{\bar{n}+1}{\overline{p_{n}}+1}\right]

does not affect positivity, and n=n¯n=\bar{n}, we get that an interpolating β\beta exists if and only if

[pm+pn¯+1m+n+1] 0.\left[\frac{p_{m}+\overline{p_{n}}+1}{m+n+1}\right]\ \geq\ 0.

\Box

3 Flat Monomial Operators

In this section, we shall consider operators of the form

T:xncnxn+τ,T:x^{n}\mapsto c_{n}x^{n+\tau}, (3.1)

and in particular when they are bounded. Note first that if τ=τ0+iτ1\tau=\tau_{0}+i\tau_{1}, with τ0\tau_{0} and τ1\tau_{1} real, then the effect of τ1\tau_{1} is to multiply everything in the range by xiτ1x^{i\tau_{1}}, which is unimodular. So without loss of generality we can assume that τ\tau is real. Moreover, by Corollary 1.11, TT can only be bounded if τ0\tau\geq 0.

Let us first handle the case τ=0\tau=0. In the terminology of Theorem 1.8, β(s)=s\beta(s)=s, so TT is bounded if and only if there is a bounded function gg on 𝕊{\mathbb{S}} that satisfies g(n)=cng(n)=c_{n}. By Pick’s theorem, this happens if and only if

[1cmcn¯1+m+n] 0.\left[\frac{1-c_{m}\overline{c_{n}}}{1+m+n}\right]\ \geq\ 0. (3.2)

So we get:

Theorem 3.3.

The map T:xncnxnT:x^{n}\mapsto c_{n}x^{n} extends to be a bounded linear map if and only if (3.2) holds. Moreover, if it is non-zero, it is never compact.

When τ>0\tau>0, how do we determine whether (3.1) extends to be bounded? Let

γ(s)=s+τ.\gamma(s)\ =\ s+\tau. (3.4)

We must find gH2(𝕊)g\in H^{2}({\mathbb{S}}) that satisfies

g(n)=cn¯1+n1+n+τ.g(n)\ =\ \overline{c_{n}}\ \frac{1+n}{1+n+\tau}.

Then TT is bounded if and only if MgCγM_{g}C_{\gamma} is bounded. (We have changed notation slightly from Theorem 1.8 to avoid the use of gg^{\cup} and to emphasize that we have a fixed choice of β\beta). To investigate MgCγM_{g}C_{\gamma}, we turn to the Poisson kernel.

The Poisson kernel for 𝕊{\mathbb{S}} at a point s=12+σ+its=-\frac{1}{2}+\sigma+it is given by

P12+σ+it(12+iy)=1πσσ2+(yt)2.P_{-\frac{1}{2}+\sigma+it}(-\frac{1}{2}+iy)\ =\ \frac{1}{\pi}\frac{\sigma}{\sigma^{2}+(y-t)^{2}}.

The Poisson integral of some function ff defined on the line {(s)=12}\{\Re(s)=-\frac{1}{2}\} is given by

P[f](12+σ+it)=1πσσ2+(yt)2f(12+iy)𝑑yP[f](-\frac{1}{2}+\sigma+it)\ =\ \frac{1}{\pi}\int_{-\infty}^{\infty}\frac{\sigma}{\sigma^{2}+(y-t)^{2}}f(-\frac{1}{2}+iy)dy

It is convenient to introduce another Hardy space, 2(𝕊){\mathcal{H}}^{2}({{\mathbb{S}}}).

Definition 3.5.

The space 2(𝕊){\mathcal{H}}^{2}({{\mathbb{S}}}) consists of all functions ff that are holomorphic in 𝕊{\mathbb{S}} and satisfy

supx>12|f(x+iy)|2𝑑y<.\sup_{x>-\frac{1}{2}}\int_{-\infty}^{\infty}|f(x+iy)|^{2}dy<\infty.

Every function f2(𝕊)f\in{\mathcal{H}}^{2}({{\mathbb{S}}}) has non-tangential boundary limits almost everywhere, and its norm is then given by (see [13]):

f2(𝕊)2=|f(12+iy)|2𝑑y.\|f\|^{2}_{{\mathcal{H}}^{2}({{\mathbb{S}}})}\ =\ \int_{-\infty}^{\infty}|f(-\frac{1}{2}+iy)|^{2}dy.

Let us write

ρ:={s:(s)>ρ}.{\mathbb{H}}_{\rho}:=\{s\in\mathbb{C}:\Re(s)>\rho\}.

So 𝕊=12{\mathbb{S}}={\mathbb{H}}_{\frac{1}{2}} in this notation.

Theorem 3.6.

The operator MgCγM_{g}C_{\gamma} is bounded if and only if the Poisson integral of |g|2|g|^{2} is bounded on some (and hence every) half-plane ρ{\mathbb{H}}_{\rho} for ρ>12\rho>-\frac{1}{2}.

Proof of Thm. 3.6. The map

W:f(s)12π1s+1f(s)W:f(s)\mapsto\frac{1}{\sqrt{2\pi}}\frac{1}{s+1}f(s)

is a unitary from H2(𝕊)H^{2}({\mathbb{S}}) onto 2(𝕊){\mathcal{H}}^{2}({{\mathbb{S}}}). One checks

WMgCγW=Ms+τ+1s+1g(s)Cγ.WM_{g}C_{\gamma}W^{*}\ =\ M_{\frac{s+\tau+1}{s+1}g(s)}C_{\gamma}. (3.7)

For ss in 𝕊{\mathbb{S}} we have

1|s+τ+1s+1|1+2τ,1\leq|\frac{s+\tau+1}{s+1}|\leq 1+2\tau,

so for boundedness purposes we can drop this factor in (3.7) and conclude that MgCγM_{g}C_{\gamma} is bounded on H2(𝕊)H^{2}({\mathbb{S}}) if and only if it is bounded on 2(𝕊){\mathcal{H}}^{2}({{\mathbb{S}}}). Thus we wish to know for which gg does

|g(12+iy)|2|f(12+τ+iy)|2𝑑y|f(12+iy)|2𝑑y\int_{-\infty}^{\infty}|g(-\frac{1}{2}+iy)|^{2}|f(-\frac{1}{2}+\tau+iy)|^{2}dy\ \lesssim\ \int_{-\infty}^{\infty}|f(-\frac{1}{2}+iy)|^{2}dy

hold? This is the same as asking when integrating along the vertical line {(s)=12+τ}\{\Re(s)=-\frac{1}{2}+\tau\} with weight |g(12+iy)|2|g(-\frac{1}{2}+iy)|^{2} is a Carleson measure for 2(𝕊){\mathcal{H}}^{2}({{\mathbb{S}}}). By [13, VI.3], this happens if and only if

supσ>0,tσ(τ+σ)2+(yt)2|g(12+iy)|2𝑑y=C<.\sup_{\sigma>0,t\in\mathbb{R}}\int_{-\infty}^{\infty}\frac{\sigma}{(\tau+\sigma)^{2}+(y-t)^{2}}|g(-\frac{1}{2}+iy)|^{2}dy\ =C\ <\ \infty. (3.8)

Taking σ=τ\sigma=\tau in (3.8), we get that MgCγM_{g}C_{\gamma} is bounded implies

suptτ(2τ)2+(yt)2|g(12+iy)|2𝑑yK,\sup_{t\in\mathbb{R}}\int_{-\infty}^{\infty}\frac{\tau}{(2\tau)^{2}+(y-t)^{2}}|g(-\frac{1}{2}+iy)|^{2}dy\ \leq\ K,

for some constant KK, so

suptτ(τ)2+(yt)2|g(12+iy)|2𝑑y 4K.\sup_{t\in\mathbb{R}}\int_{-\infty}^{\infty}\frac{\tau}{(\tau)^{2}+(y-t)^{2}}|g(-\frac{1}{2}+iy)|^{2}dy\ \leq\ 4K.

If u=12+τ+σ+itu=-\frac{1}{2}+\tau+\sigma+it is any point in the half-plane 12+τ{\mathbb{H}}_{-\frac{1}{2}+\tau}, we have

Pu(12+iy)=1πσ+τ(σ+τ)2+(yt)2.P_{u}(-\frac{1}{2}+iy)\ =\ \frac{1}{\pi}\frac{\sigma+\tau}{(\sigma+\tau)^{2}+(y-t)^{2}}.

As

σ+τ(τ+σ)2+(yt)2σ(τ+σ)2+(yt)2+τ(τ)2+(yt)2,\frac{\sigma+\tau}{(\tau+\sigma)^{2}+(y-t)^{2}}\ \leq\frac{\sigma}{(\tau+\sigma)^{2}+(y-t)^{2}}+\frac{\tau}{(\tau)^{2}+(y-t)^{2}},

we conclude that (3.8) implies that the Poisson integral of |g|2|g|^{2} is bounded on 12+τ{\mathbb{H}}_{-\frac{1}{2}+\tau}. Conversely, the boundedness of the Poisson integral of |g|2|g|^{2} on 12+τ{\mathbb{H}}_{-\frac{1}{2}+\tau} implies (3.8).

Finally, to see that boundedness of the Poisson integral of |g|2|g|^{2} on 12+σ{\mathbb{H}}_{-\frac{1}{2}+\sigma} implies boundedness on 12+ρ{\mathbb{H}}_{-\frac{1}{2}+\rho} for any 0<ρ<σ0<\rho<\sigma, we just observe that

1πρρ2+(yt)2(σ2ρ2)1πσσ2+(yt)2.\frac{1}{\pi}\frac{\rho}{\rho^{2}+(y-t)^{2}}\ \leq\ (\frac{\sigma^{2}}{\rho^{2}})\frac{1}{\pi}\frac{\sigma}{\sigma^{2}+(y-t)^{2}}.

\Box


Example 3.9  Let τ>0\tau>0, and take g(s)=1(1+s)(s+12)cg(s)=\frac{1}{(1+s)(s+\frac{1}{2})^{c}} for some 0<c<120<c<\frac{1}{2}. Then gg is not bounded on 𝕊{\mathbb{S}}, so MgM_{g} is not bounded. But

P[|g|2](12+σ+it)\displaystyle P[|g|^{2}](-\frac{1}{2}+\sigma+it) =\displaystyle\ =\ 1πσσ2+(yt)21|y|2c114+y2𝑑y\displaystyle\frac{1}{\pi}\int_{-\infty}^{\infty}\frac{\sigma}{\sigma^{2}+(y-t)^{2}}\frac{1}{|y|^{2c}}\frac{1}{\frac{1}{4}+y^{2}}dy
\displaystyle\leq 1πσ1|y|2c114+y2𝑑y\displaystyle\frac{1}{\pi\sigma}\int_{-\infty}^{\infty}\frac{1}{|y|^{2c}}\frac{1}{\frac{1}{4}+y^{2}}dy
\displaystyle\leq 1πσ[114|y|2c𝑑y+114+y2𝑑y]\displaystyle\frac{1}{\pi\sigma}\left[\int_{-1}^{1}\frac{4}{|y|^{2c}}dy+\int_{-\infty}^{\infty}\frac{1}{\frac{1}{4}+y^{2}}dy\right]
=\displaystyle\ =\ 1πσ[812c+2π]\displaystyle\frac{1}{\pi\sigma}\left[\frac{8}{1-2c}+2\pi\right]

This is bounded in each half-plane ρ{\mathbb{H}}_{\rho}, so by Theorem 3.6 MgCγM_{g}C_{\gamma} is bounded.

4 Unitary Monomial Operators

Bourdon and Narayan characterized unitary weighted composition operators [5]. Their theorem (translated to 𝕊{\mathbb{S}}) is:

Theorem 4.1.

(Bourdon-Narayan) The operator MgCβM_{g}C_{\beta} is unitary on H2(𝕊)H^{2}({\mathbb{S}}) if and only if β\beta is an automorphism of 𝕊{\mathbb{S}} and g(s)=eiθks0/ks0g(s)=e^{i\theta}k_{s_{0}}/\|k_{s_{0}}\|, where s0=β1(0)s_{0}=\beta^{-1}(0).

We shall give a proof of their theorem directly in the context of monomial operators. First, let us describe the automorphisms of 𝕊{\mathbb{S}} in a convenient way.

Lemma 4.2.

The function β:𝕊\beta:{\mathbb{S}}\to\mathbb{C} is an automorphism of 𝕊{\mathbb{S}} if and only if there is a function ϕ:𝕊\phi:{\mathbb{S}}\to\mathbb{C} so that

1+β(s)+β(t)¯1+s+t¯=ϕ(s)ϕ(t)¯.\frac{1+\beta(s)+\overline{\beta(t)}}{1+s+\overline{t}}\ =\ \phi(s)\overline{\phi(t)}. (4.3)

Proof: Notice that β\beta is an automorphism of 𝕊{\mathbb{S}} if and only if b:=λβλ1b:=\lambda\circ\beta\circ\lambda^{-1} is an automorphism of 𝔻\mathbb{D}. By Pick’s theorem (see [3, Sec. 2.6]) the latter occurs if and only if

1b(w)¯b(z)1w¯z=ψ(w)¯ψ(z)\frac{1-\overline{b(w)}b(z)}{1-\bar{w}z}\ =\ \overline{\psi(w)}\psi(z) (4.4)

for some function ψ\psi on 𝔻\mathbb{D}. Doing some algebra, (4.4) becomes (4.3) with

ϕ(s)=1+β(s)1+sψ(s1+s).\phi(s)\ =\ \frac{1+\beta(s)}{1+s}\psi(\frac{s}{1+s}).

\Box

Consequently, we have the following characterization of unitary monomial operators (which can also be derived from Bourdon-Narayan and Theorem 1.8).

Theorem 4.5.

The operator

T:xsc(s)xβ(s)T:x^{s}\mapsto c(s)x^{\beta(s)}

is unitary on L2[0,1]L^{2}[0,1] if and only if β\beta is an automorphism of 𝕊{\mathbb{S}} and c(s)c(s) is defined by

c(s)=eiθ1+2β(0)1+β(0)¯+β(s)1+s.c(s)\ =\ \frac{e^{i\theta}}{\sqrt{1+2\Re\beta(0)}}\frac{1+\overline{\beta(0)}+\beta(s)}{1+s}. (4.6)

Proof: Since β\beta is holomorphic by Theorem 1.8 and has to be non-constant for TT to be bounded, the range of TT must be dense. Therefore it is unitary if and only if it is isometric. It is isometric if and only if for every s,ts,t we have

xs,xt\displaystyle\langle x^{s},x^{t}\rangle =\displaystyle\ =\ c(s)xβ(s),c(t)xβ(t)\displaystyle\langle c(s)x^{\beta(s)},c(t)x^{\beta(t)}\rangle
11+s+t¯\displaystyle\Leftrightarrow\quad\frac{1}{1+s+\bar{t}} =\displaystyle\ =\ c(s)c(t)¯1+β(s)+β(t)¯.\displaystyle\frac{c(s)\overline{c(t)}}{1+\beta(s)+\overline{\beta(t)}}.

That means (4.4) holds, so β\beta is an automorphism. Letting t=0t=0 we get

c(0)=eiθ1+2β(0),c(0)\ =\ e^{i\theta}\sqrt{1+2\Re\beta(0)},

and solving for c(s)c(s) we get (4.6).

Conversely, suppose (4.6) holds and β\beta is an automorphism. By Lemma 4.2, we have (4.3) for some ϕ\phi. Letting t=0t=0 and solving, we get that ϕ\phi is given by (4.6). \Box

5 Questions

Question 5.1.

If MgCγM_{g}C_{\gamma} is bounded, with γ\gamma as in (3.4), can one approximate it in norm by operators of the form MgnCγM_{g_{n}}C_{\gamma} where gnH(𝕊)g_{n}\in H^{\infty}({\mathbb{S}})?

Question 5.2.

A generalization of the previous question is for non-flat monomial operators. Can every bounded operator of the form MgCβM_{g}C_{\beta} be approximated by operators MgnCβM_{g_{n}}C_{\beta} with bounded gng_{n}?

Question 5.3.

What are necessary and sufficient conditions for the functions β\beta and gg so that the operator TT defined by (1.9) is bounded? Compact?

On behalf of all authors, the corresponding author states that there is no conflict of interest.

References

  • [1] Jim Agler and John E. McCarthy. The asymptotic Müntz-Szász theorem. Preprint, 2022.
  • [2] Jim Agler and John E. McCarthy. A generalization of Hardy’s operator and an asymptotic Müntz-Szász theorem. Preprint, 2022.
  • [3] Jim Agler, John E. McCarthy, and N. J. Young. Operator Analysis: Hilbert Space Methods in Complex Analysis. Cambridge Tracts in Mathematics. Cambridge University Press, 2020.
  • [4] Wolfgang Arendt, Isabelle Chalendar, Mahesh Kumar, and Sachi Srivastava. Asymptotic behaviour of the powers of composition operators on Banach spaces of holomorphic functions. Indiana Univ. Math. J., 67(4):1571–1595, 2018.
  • [5] Paul S. Bourdon and Sivaram K. Narayan. Normal weighted composition operators on the Hardy space H2(𝕌)H^{2}(\mathbb{U}). J. Math. Anal. Appl., 367(1):278–286, 2010.
  • [6] M. S. Brodskii. On a problem of I. M. Gelfand. Uspehi Mat. Nauk (N.S.), 12(2(74)):129–132, 1957.
  • [7] Isabelle Chalendar and Jonathan R. Partington. Weighted composition operators: isometries and asymptotic behaviour. J. Operator Theory, 86(1):189–201, 2021.
  • [8] Carl C. Cowen and Eva A. Gallardo-Gutiérrez. A new class of operators and a description of adjoints of composition operators. J. Funct. Anal., 238(2):447–462, 2006.
  • [9] Carl C. Cowen and Eungil Ko. Hermitian weighted composition operators on H2H^{2}. Trans. Amer. Math. Soc., 362(11):5771–5801, 2010.
  • [10] W. F. Donoghue, Jr. The lattice of invariant subspaces of a completely continuous quasi-nilpotent transformation. Pacific J. Math., 7:1031–1035, 1957.
  • [11] Peter L. Duren. Theory of HpH^{p} Spaces. Academic Press, New York, 1970.
  • [12] Frank Forelli. The isometries of HpH^{p}. Canadian J. Math., 16:721–728, 1964.
  • [13] John B. Garnett. Bounded Analytic Functions. Academic Press, New York, 1981.
  • [14] Gajath Gunatillake. Compact weighted composition operators on the Hardy space. Proc. Amer. Math. Soc., 136(8):2895–2899, 2008.
  • [15] Thomas Kriete and Jennifer Moorhouse. Linear relations in the Calkin algebra for composition operators. Trans. Amer. Math. Soc., 359(6):2915–2944, 2007.
  • [16] K. Zhu. Operator Theory in Function Spaces, volume 139 of Monographs and textbooks in pure and applied mathematics. Marcel Dekker, Inc., New York, 1990.