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

Smooth Functional Calculus and Spectral Theorem in Banach Spaces

Luis A. Cedeño-Pérez and Hernando Quevedo Instituto de Ciencias Nucleares, Universidad Nacional Autónoma de México, AP 70543, Mexico City, Mexico Dipartimento di Fisica and Icra, Università di Roma “La Sapienza”, Roma, Italy Al-Farabi Kazakh National University, Al-Farabi av. 71, 050040 Almaty, Kazakhstan
Abstract

The notion of projection families generalizes the the classical notions of vector and operator-valued measures. We show that projection families are general enough to extend the Spectral Theorem to Banach algebras and operators between Banach spaces. To this end, we first develop a Smooth Functional Calculus in Banach algebras using the Cauchy-Pompeiu Formula, which is further extended to a Continuous Functional Calculus. We also show that these theorems are proper generalizations of the usual result for operators between Hilbert spaces.

preprint: AIP/123-QED

I Introduction

In a previous work [4], we defined a new kind of measures called projection families and developed their theory of integration. In the present article, we show that the properties of projection families permit to generalize the Spectral Theorem, both to Banach algebras and operators between Banach spaces.

Let XX be a Banach algebra and xXx\in X. If ff is an holomorphic function in a neighborhood Ω\Omega of σ(x)\sigma(x), we define f(x)f(x) as the element of XX given by

f(x)=12πiγf(λ)Rx(λ)𝑑λ,f(x)=\frac{1}{2\pi i}\int_{\gamma}f(\lambda)\;R_{x}(\lambda)\;d\lambda,

where γ\gamma is a simple closed curve Ω\Omega that surrounds σ(x)\sigma(x) and Rx(λ)R_{x}(\lambda) is the resolvent function of xx. This is known as the Holomorphic Functional Calculus. Despite being defined for every element of the Banach algebra, this functional calculus has the defect that the class of holomorphic functions is too small.

The Holomorphic Functional Calculus is clearly based on the Cauchy Integral Formula. If we wish to extend the Holomorphic Functional Calculus to the continuous functions in σ(x)\sigma(x), it is natural to first look for a generalization of the Cauchy Integral Formula to use as a starting point. We use the less known Cauchy-Pompeiu Formula, which generalizes the Cauchy Integral Formula to smooth functions. This formula states that if Ω\Omega\subset\mathbb{C} is an open set with compact closure and smooth boundary and ff is smooth in Ω¯\overline{\Omega} then

f(λ)=12πiωf(z)zλ𝑑z1πωzf(z)zλ𝑑x𝑑yf(\lambda)=\frac{1}{2\pi i}\int_{\partial\omega}\frac{f(z)}{z-\lambda}\;dz-\frac{1}{\pi}\int_{\omega}\frac{\partial_{z^{\ast}}f(z)}{z-\lambda}\;dxdy

for any λΩ\lambda\in\Omega (see [10]). Based on the Cauchy-Pompeiu Formula, the Smooth Functional Calculus should be defined as

f(x)=12πiγf(λ)Rx(λ)𝑑γ(λ)1πintγσ(x)zf(λ)Rx(λ)d(λ).f(x)=\frac{1}{2\pi i}\int_{\gamma}f(\lambda)R_{x}(\lambda)\;d\gamma(\lambda)-\frac{1}{\pi}\int_{int\;\gamma\setminus\sigma(x)}\partial_{z^{\ast}}f(\lambda)R_{x}(\lambda)\;d\ell(\lambda).

The first integral exists by continuity and compactness, however, the second one is much more complicated since the resolvent function Rx(λ)R_{x}(\lambda) diverges close to the spectrum. The first integral ignores this as the curve γ\gamma is never in the spectrum, but this can not be avoided in the second integral since the integration is on the plane. The second integral is, essentially, a singular integral and as such may not define a continuous operator. Two things are needed to deal with this singular integral. Firstly, we require xx to have additional properties in terms of its resolvent RxR_{x}. This leads to the notion of regular elements, which are precisely those singular integrals with a continuous behavior. The second is to allow f(x)f(x) to be an element of the larger space XX^{\ast\ast} instead of XX. The previous formula gives the definition of the Smooth Functional Calculus when both conditions are met.

Given that smooth functions are dense in the continuous functions, one would expect the extension to the continuous case to be rather straightforward. This is not the case since the Smooth Functional Calculus is easily seen to be continuous with respect to the norm of C1C^{1}, but the continuity required to extend by density is with respect to the uniform norm. This difficulty comes from, essentially, the appearance of the first derivatives in the Smooth Functional Calculus. To deal with this it is necessary to manipulate the definition of the Smooth Functional Calculus in such a way that the resulting expression does not have a dependence on first derivatives of the function. Once this is done, the continuous extension to the space of continuous functions will be possible, resulting in the Continuous Functional Calculus.

The existence of the Continuous Functional Calculus will allow us to prove two versions of the Spectral Theorem, the first for Banach algebras and the second for operators between Banach spaces. The Banach algebra version is clearly valid in the Banach algebra B(X)B(X), however, it turns out to be too restrictive, thus the second version is developed under less restrictive hypotheses which we show are enough for operators. The central idea of our constructions is to study the function

(f,Λ,x)Λ(f(T)(x)),(f,\Lambda,x)\longmapsto\Lambda(f(T)(x)),

for a fixed operator TB(X)T\in B(X), and study its continuity properties. For fixed Λ\Lambda and xx, the continuity with respect to ff leads to the existence of a measure μΛ,xT\mu^{T}_{\Lambda,x} which determines the functional. The continuity with respect to Λ\Lambda and xx implies that the family of measures

μT={μΛ,xT|ΛX,xX}\mu^{T}=\{\mu^{T}_{\Lambda,x}\;|\;\Lambda\in X^{\ast},\;x\in X\}

is an operator projection family. This allows us to extend the Continuous Functional Calculus to the space of integrable functions L1(μT)L^{1}(\mu^{T}). Finally, we show that this theorem is a strict generalization of the usual Spectral Theorem for operators between Hilbert spaces.

The first section of this work is to prove the Cauchy-Pompeiu Formula and other results in geometric integration that will be useful throughout the work. In the second section, we state the basic results on vector integration and projection families, as well as the analytical tools we will require. In the third section, we develop the Smooth and Continuous Functional Calculi executing the program described in this section. Finally, in the last section, we prove the Spectral Theorem in the contexts of Banach algebras and operators between Banach spaces.

II Geometric Integration

II.1 Cauchy-Pompeiu Formula

The complex plane \mathbb{C} has two coordinate differential forms dxdx and dydy, which induce linearly independent differential forms

dz=dx+idydz=dx+i\;dy

and

dz=dxidy,dz^{\ast}=dx-i\;dy,

and thus span every 11-form in \mathbb{C}. A simple computation shows that

dzdz=2idxdy,dz\wedge dz^{\ast}=-2i\;dx\wedge dy, (1)

and every 22-form is obtained by multiplying this form by a scalar function. These differential forms have associated tangent vectors

2z=xiy2\partial_{z}=\partial_{x}-i\partial_{y}

and

2z=x+iy,2\partial_{z^{\ast}}=\partial_{x}+i\partial_{y},

given by the conditions dz(z)=1dz(\partial_{z})=1, dz(z)=1dz^{\ast}(\partial_{z^{\ast}})=1 and dz(z)=dz(z)=0dz(\partial_{z^{\ast}})=dz^{\ast}(\partial_{z})=0.

If uu is a smooth function defined in an open set ω\omega with smooth boundary ω\partial\omega, then the complex integral along ω\partial\omega coincides with the integral of the 11-form udzu\;dz along ω\partial\omega. Furthermore, the Stokes Theorem implies that

ωu𝑑z\displaystyle\int_{\partial\omega}u\;dz =ωd(udz)\displaystyle=\int_{\omega}d(u\;dz)
=ω𝑑udz\displaystyle=\int_{\omega}du\wedge dz
=ω(zudz+zudz)dz\displaystyle=\int_{\omega}(\partial_{z}u\;dz+\partial_{z^{\ast}}u\;dz^{\ast})\wedge dz
=ωzudzdz.\displaystyle=\int_{\omega}\partial_{z^{\ast}}u\;dz^{\ast}\wedge dz. (2)

It is convenient to note that the function

λ1λ\lambda\longmapsto\frac{1}{\lambda}

is integrable near the origin, since the dimension is two. This simplifies the convergence of certain integrals.

Theorem II.1 (Cauchy-Pompeiu Formula).

Let Ω\Omega\subset\mathbb{C} be an open set with compact closure, uC(Ω)u\in C^{\infty}(\Omega) and ωΩ\omega\subset\Omega an open set with smooth boundary such that ωΩ\partial\omega\subset\Omega. The formula

u(λ)\displaystyle u(\lambda) =12πiωuzλ𝑑z+12πiωzuzλ𝑑zdz\displaystyle=\frac{1}{2\pi i}\int_{\partial\omega}\frac{u}{z-\lambda}\;dz+\frac{1}{2\pi i}\int_{\omega}\frac{\partial_{z^{\ast}}u}{z-\lambda}dz^{\ast}\wedge dz
=12πiωuzλ𝑑z1πωzuzλ𝑑xdy\displaystyle=\frac{1}{2\pi i}\int_{\partial\omega}\frac{u}{z-\lambda}\;dz-\frac{1}{\pi}\int_{\omega}\frac{\partial_{z^{\ast}}u}{z-\lambda}\;dx\wedge dy

is valid for any λω\lambda\in\omega.

Proof.

Since λω\lambda\in\omega and ω\omega is open, there exists ϵ>0\epsilon>0 such that Bϵ(λ)ωB_{\epsilon}(\lambda)\subset\omega. We define

ωϵ=ωBϵ(λ).\omega_{\epsilon}=\omega\setminus B_{\epsilon}(\lambda).

The function

uzλ\frac{u}{z-\lambda}

is CC^{\infty} in ωϵ\omega_{\epsilon}, thus equation (II.1) implies that

ωϵzuzλ𝑑zdz\displaystyle\int_{\omega_{\epsilon}}\frac{\partial_{z^{\ast}}u}{z-\lambda}\;dz^{\ast}\wedge dz =ωϵuzλ𝑑z\displaystyle=\int_{\partial\omega_{\epsilon}}\frac{u}{z-\lambda}\;dz
=ωuzλ𝑑zBϵ(λ)uzλ𝑑z.\displaystyle=\int_{\partial\omega}\frac{u}{z-\lambda}\;dz-\int_{\partial B_{\epsilon}(\lambda)}\frac{u}{z-\lambda}\;dz. (3)

Parametrize Bϵ(λ)\partial B_{\epsilon}(\lambda) as λ+ϵeit\lambda+\epsilon e^{it} with t[0,2π]t\in[0,2\pi], hence

dz=iϵeitdt.dz=i\epsilon e^{it}\;dt.

It follows that

Bϵ(λ)uzλ𝑑z\displaystyle\int_{\partial B_{\epsilon}(\lambda)}\frac{u}{z-\lambda}\;dz =i02πu(λ+ϵeit)𝑑t.\displaystyle=i\int_{0}^{2\pi}u(\lambda+\epsilon e^{it})\;dt.

Substituting in equation (II.1) we find that

ωϵzuzλ𝑑zdz=ωuzλ𝑑zi02πu(λ+ϵeit)𝑑t.\int_{\omega_{\epsilon}}\frac{\partial_{z^{\ast}}u}{z-\lambda}\;dz^{\ast}\wedge dz=\int_{\partial\omega}\frac{u}{z-\lambda}\;dz-i\int_{0}^{2\pi}u(\lambda+\epsilon e^{it})\;dt.

The Monotone Convergence Theorem on the left-hand side and the Dominated Convergence Theorem on the right-hand side imply that the limit as ϵ0\epsilon\to 0 is

ωzuzλ𝑑zdz=ωuzλ𝑑z2πiu(λ).\int_{\omega}\frac{\partial_{z^{\ast}}u}{z-\lambda}\;dz^{\ast}\wedge dz=\int_{\partial\omega}\frac{u}{z-\lambda}\;dz-2\pi i\;u(\lambda).

The result follows from rearranging terms. ∎

The language of differential topology is useful to deduce the Cauchy-Pompeiu Formula, however, in the rest of the work the language of measure theory will be more useful.

Corollary II.1.1.

Under the hypotheses of the Cauchy-Pompeiu Formula, if γ=ω\gamma=\partial\omega, dγd\gamma is the measure induced by γ\gamma and \ell is the Lebesgue measure then

u(λ)=12πiωuzλ𝑑γ1πωzuzλ𝑑.u(\lambda)=\frac{1}{2\pi i}\int_{\partial\omega}\frac{u}{z-\lambda}\;d\gamma-\frac{1}{\pi}\int_{\omega}\frac{\partial_{z^{\ast}}u}{z-\lambda}\;d\ell.
Proof.

It follows from the Cauchy-Pompeiu Formula and equation (1). ∎

Just as for the Cauchy Integral Formula and analytic functions, the Cauchy-Pompeiu Formula is equivalent to other statements on the nullity of integrals and the independence of trajectories.

Theorem II.2 (Cauchy-Pompeiu).

Let Ω\Omega be an open subset of \mathbb{C} with compact closure and ff a smooth function in Ω\Omega. The following are satisfied:

  1. 1.

    For any simple closed curve γ\gamma in Ω\Omega we have that

    12πiγf𝑑γ1πintγzfd=0,\frac{1}{2\pi i}\int_{\gamma}f\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma}\partial_{z^{\ast}}f\;d\ell=0,

    where intγint\;\gamma is the geometric interior of γ\gamma and not the topological interior.

  2. 2.

    If γ1\gamma_{1} and γ2\gamma_{2} are closed curves homotopic in Ω\Omega then

    12πiγ1f𝑑γ1πintγ1zfd=12πiγ2f𝑑γ1πintγ2zfd.\frac{1}{2\pi i}\int_{\gamma_{1}}f\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma_{1}}\partial_{z^{\ast}}f\;d\ell=\frac{1}{2\pi i}\int_{\gamma_{2}}f\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma_{2}}\partial_{z^{\ast}}f\;d\ell.
Proof.

For the first statement, we apply the Cauchy-Pompeiu Formula to the smooth function f(z)(zλ)f(z)(z-\lambda) with λintγ\lambda\in int\;\gamma. For the second statement, we apply the first one to an appropriate homology class. ∎

II.2 Coarea Formula

The last fact from geometric integration that we require is the Coarea Formula, which relates the integrals on sublevel sets and integrals on level sets (see [2]). This formula will allow us to study the singular behavior of the resolvent function RxR_{x}, which in turn is related to the singular behavior of reciprocals of distance-to-a-set functions.

Theorem II.3 (Coarea Formula).

Let Ω\Omega\subset\mathbb{C} be an open set and H:Ω[0,1]H\colon\Omega\to[0,1] a CC^{\infty} function. For each t[0,1]t\in[0,1] define

Ωt=H1([0,t))\Omega_{t}=H^{-1}([0,t))

and

γt=H1({t}).\gamma_{t}=H^{-1}(\{t\}).

If Φt\Phi_{t} is the flow of H\nabla H then

d=1|H|dγtdtΦ,d\ell=\frac{1}{|\nabla H|}d\gamma_{t}\wedge d_{t}\Phi,

thus, if f:Ωf\colon\Omega\to\mathbb{R} is integrable then

Ωf𝑑=01γtf1|H|𝑑γt𝑑t.\int_{\Omega}f\;d\ell=\int_{0}^{1}\int_{\gamma_{t}}f\frac{1}{|\nabla H|}\;d\gamma_{t}\;dt.
Proof.

We first note that by Sard’s Theorem, the set of critical values of HH has zero Lebesgue measure, thus the curves γt\gamma_{t} are smooth except for a null set. The Chain Rule implies that the differential form dtΦd_{t}\Phi satisfies the equations

dtΦ(γt˙)=0d_{t}\Phi(\dot{\gamma_{t}})=0

and

|dtΦ|=|H|,|d_{t}\Phi|=|\nabla H|,

hence if γt\gamma_{t} is parametrized by arc-length and v\vec{v} is a vector parallel to H\nabla H then

1|H|dtΦdγt(v,γ˙t)\displaystyle\frac{1}{|\nabla H|}d_{t}\Phi\wedge d\gamma_{t}(v,\dot{\gamma}_{t}) =1|H|dtΦ(v)dγt(γ˙t)\displaystyle=\frac{1}{|\nabla H|}d_{t}\Phi(v)d\gamma_{t}(\dot{\gamma}_{t})
=1,\displaystyle=1,

thus this is the volume form of the plane. The result follows from this. ∎

Let σ\sigma be a closed subset of the plane. Recall that the function

d(λ,σ)=inf{|λz||zσ}d(\lambda,\sigma)=\inf\{|\lambda-z|\;|\;z\in\sigma\}

vanishes only if λσ\lambda\in\sigma and is Lipschitz, thus the Radamacher Theorem implies that it is differentiable almost everywhere. Furthermore, these functions are characterized by the condition

|H|=1.|\nabla H|=1.

We now use the Coarea Formula to study the integrability of the reciprocals of this kind of functions.

Corollary II.3.1.

Let σ\sigma\subset\mathbb{C} be a compact set and Ω\Omega a neighbourhood of σ\sigma. If (σ)=0\ell(\sigma)=0 then

λ1d(λ,σ)\lambda\longmapsto\frac{1}{d(\lambda,\sigma)}

is integrable in Ωσ\Omega\setminus\sigma.

Proof.

We first study the integrals without the assumption that σ\sigma has a null Lebesgue measure. The function

dσ(λ)=d(λ,σ)d_{\sigma}(\lambda)=d(\lambda,\sigma)

satisfies dσ1(0)=σd_{\sigma}^{-1}(0)=\sigma, thus, when used as the function HH in the Coarea formula, we have that

Ωσ1d(λ,σ)𝑑(λ)\displaystyle\int_{\Omega\setminus\sigma}\frac{1}{d(\lambda,\sigma)}\;d\ell(\lambda) =01γt1d(λ,σ)𝑑γt𝑑t.\displaystyle=\int_{0}^{1}\int_{\gamma_{t}}\frac{1}{d(\lambda,\sigma)}\;d\gamma_{t}\;dt.

Note that γt\gamma_{t} is the set of points at a distance tt to σ\sigma, hence

Ω0σ1d(λ,σ)𝑑(λ)\displaystyle\int_{\Omega_{0}\setminus\sigma}\frac{1}{d(\lambda,\sigma)}\;d\ell(\lambda) =01γt1t𝑑γt𝑑t\displaystyle=\int_{0}^{1}\int_{\gamma_{t}}\frac{1}{t}\;d\gamma_{t}\;dt
=01l(γt)t𝑑t,\displaystyle=\int_{0}^{1}\frac{l(\gamma_{t})}{t}\;dt,

where l(γt)l(\gamma_{t}) is the length of γt\gamma_{t}. We now assume that (σ)=0\ell(\sigma)=0, which is the same as (Ω0)=0\ell(\Omega_{0})=0. Define the truncation function

I(ϵ)=ϵRl(γt)t𝑑tI(\epsilon)=\int_{\epsilon}^{R}\frac{l(\gamma_{t})}{t}\;dt

with RϵR\geq\epsilon and compute the integral

01I(ϵ)𝑑ϵ\displaystyle\int_{0}^{1}I(\epsilon)\;d\epsilon =01ϵRl(γt)t𝑑t𝑑ϵ\displaystyle=\int_{0}^{1}\int_{\epsilon}^{R}\frac{l(\gamma_{t})}{t}\;dt\;d\epsilon
010tl(γt)t𝑑ϵ𝑑t\displaystyle\leq\int_{0}^{1}\int_{0}^{t}\frac{l(\gamma_{t})}{t}\;d\epsilon\;dt
=0Rl(γt)𝑑t\displaystyle=\int_{0}^{R}l(\gamma_{t})\;dt
=(ΩR),\displaystyle=\ell(\Omega_{R}),

where the last equality follows from the Coarea Formula. It follows that

I(0)\displaystyle I(0) =limR01R0RI(ϵ)𝑑ϵ\displaystyle=\lim_{R\to 0}\frac{1}{R}\int_{0}^{R}I(\epsilon)\;d\epsilon
limR01R(ΩR)\displaystyle\leq\lim_{R\to 0}\frac{1}{R}\ell(\Omega_{R})
=dt((γt))|t=0\displaystyle=d_{t}(\ell(\gamma_{t}))|_{t=0}
=l(γ0)\displaystyle=l(\gamma_{0})
<,\displaystyle<\infty,

where the last equality follows from the Coarea Formula and the previous one from the fact that (Ω0)=0\ell(\Omega_{0})=0. ∎

By means of the integral

01l(γt)t𝑑t\int_{0}^{1}\frac{l(\gamma_{t})}{t}\;dt

it is easy to see that the integral of 1dσ\frac{1}{d_{\sigma}} can diverge for sets with non-null measure, such as BR(0)¯\overline{B_{R}(0)}. This does not imply that 1dσ\frac{1}{d_{\sigma}} diverges for sets with non-null measure, as care has to be taken not to integrate over sets where the functions are infinite, in the same way that this does not imply that 1|λ|\frac{1}{|\lambda|} has a divergent integral near 0 in 2\mathbb{R}^{2}.

The Cauchy-Pompeiu Formula can also be used to obtain information on the integral over codimension 11 manifolds from the original integral.

Theorem II.4.

Let (γn)n(\gamma_{n})_{n\in\mathbb{N}} a sequence of smooth curves such that intγnΩint\;\gamma_{n}\searrow\Omega. If fC1(intγ1Ω)f\in C^{1}(int\;\gamma_{1}\setminus\Omega) then the limit

limnγnf𝑑γn\lim_{n\to\infty}\int_{\gamma_{n}}f\;d\gamma_{n}

exists.

Proof.

By the Cauchy-Pompeiu Formula, there exists a number cc such that

c=12πiγnf𝑑γn1πintγ1Ωzfdc=\frac{1}{2\pi i}\int_{\gamma_{n}}f\;d\gamma_{n}-\frac{1}{\pi}\int_{int\;\gamma_{1}\setminus\Omega}\partial_{z^{\ast}}f\;d\ell

for every nn\in\mathbb{N}. The left-hand side clearly has a limit as nn\to\infty and the second integral in the right-hand side vanishes on the same limit. This implies that the first integral on the right-hand side has a limit as nn\to\infty. ∎

Note that the limit in the previous result is precisely cc, which is the common value of the integrals

12πiγf𝑑γ1πintγΩzfd\frac{1}{2\pi i}\int_{\gamma}f\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma\setminus\Omega}\partial_{z^{\ast}}f\;d\ell

along any of the curves.

III Projection Families and Integration

III.1 Vector Integration

We now provide a brief summary of the most important integrals of vector-valued functions. See [6], [7] or [8]. First of all, two notions of measurability arise naturally in this setting.

Definition.

Let (Ω,Σ,μ)(\Omega,\Sigma,\mu) be a measure space, XX a Banach space and f:ΩXf\colon\Omega\to X.

  1. 1.

    ff is weakly measurable if the function Λf\Lambda\circ f is measurable for each ΛX\Lambda\in X^{\ast}.

  2. 2.

    ff is strongly measurable if there exists a sequence of simple functions (sn)n(s_{n})_{n\in\mathbb{N}} such that sna.e.fs_{n}\xrightarrow[]{a.e.}f.

The first type of integrals comes from considering Riemann-type sums.

Definition.

Let (Ω,Σ,μ)(\Omega,\Sigma,\mu) be a measure space, XX a Banach space, and f:ΩXf\colon\Omega\to X a strongly measurable function. ff is Bochner integrable if there exists a sequence of simple functions (sn)n(s_{n})_{n\in\mathbb{N}} such that

limn|snf|𝑑μ=0,\lim_{n\to\infty}\int|s_{n}-f|\;d\mu=0,

in which case we define the Bochner integral as

f𝑑μ=limnsn𝑑μ.\int f\;d\mu=\lim_{n\to\infty}\int s_{n}\;d\mu.
Theorem III.1 (Bochner’s Integrability Theorem).

Let (Ω,Σ,μ)(\Omega,\Sigma,\mu) be a measure space, XX a Banach space and f:ΩXf\colon\Omega\to X. ff is Bochner integrable if and only if ff is stringly measurable and

|f|𝑑μ<.\int|f|\;d\mu<\infty.

The class of Bochner integrable functions is not large enough and its computation is not easy. The following two integrals are improvements in both senses.

Definition.

Let (Ω,Σ,μ)(\Omega,\Sigma,\mu) be a measure space, XX a Banach space, and f:ΩXf\colon\Omega\to X a weakly measurable function. We say that ff is scalarly integrable if ΛfL1(μ)\Lambda\circ f\in L^{1}(\mu) for each ΛX\Lambda\in X^{\ast}.

Definition.

Let (Ω,Σ,μ)(\Omega,\Sigma,\mu) be a measure space, XX a Banach space, and f:ΩXf\colon\Omega\to X a scalarly integrable function. The function ff is Pettis integrable if there exists f𝑑μX\int f\;d\mu\in X such that

Λ(Pf𝑑μ)=PΛf𝑑μ\Lambda\left(P\int f\;d\mu\right)=P\int\Lambda\circ f\;d\mu

for each ΛX\Lambda\in X^{\ast}.

Theorem III.2 (Dunford’s Lemma).

Let (Ω,Σ,μ)(\Omega,\Sigma,\mu) be a measure space, XX a Banach space, and f:ΩXf\colon\Omega\to X a scalarly integrable function. The functional

Df𝑑μ:ΛΛf𝑑μ\begin{array}[]{ccc}D\int f\;d\mu\colon&\longrightarrow&\mathbb{C}\\ \Lambda&\longmapsto&\int\Lambda\circ f\;d\mu\end{array}

defines an element of XX^{\ast\ast}.

Definition.

Let (Ω,Σ,μ)(\Omega,\Sigma,\mu) be a measure space, XX a Banach space, and f:ΩXf\colon\Omega\to X a scalarly integrable function. We define the Dunford integral as the functional Df𝑑μXD\int f\;d\mu\in X^{\ast\ast} given by

(Df𝑑μ)(Λ)=Λf𝑑μ.\left(D\int f\;d\mu\right)(\Lambda)=\int\Lambda\circ f\;d\mu.

Schematically, the different kinds of integrability we have discussed so far are related in the following way.

Bochner \Rightarrow Pettis \Rightarrow Dunford.

If ff is Bochner integrable then the Bochner and Pettis integrals are equal. If ff is Pettis integrable then the Pettis and Dunford integrals are related by a slightly more complicated equation:

Df𝑑μ=J(Pf𝑑μ).D\int f\;d\mu=J\left(P\int f\;d\mu\right).

III.2 Projection Families

We now summarize the most basic properties of vector and operator projection families, as developed in [4].

Definition.

Let XX be a Banach space and (Ω,Σ)(\Omega,\Sigma) a measurable space. An operator projection family is a family of measures in Ω\Omega, denoted by

μ={μΛ,x|ΛX,xX},\mu=\{\mu_{\Lambda,x}\;|\;\Lambda\in X^{\ast},\;x\in X\},

with the following two properties.

  1. 1.

    The function

    X×X(Ω)(Λ,x)μΛ,x\begin{array}[]{ccc}X^{\ast}\times X&\longrightarrow&\mathcal{M}(\Omega)\\ (\Lambda,x)&\longmapsto&\mu_{\Lambda,x}\end{array}

    is bilinear.

  2. 2.

    If (Λi)iI(\Lambda_{i})_{i\in I} and (xj)jJ(x_{j})_{j\in J} are nets such that ΛiΛ\Lambda_{i}\to\Lambda and xjxx_{j}\to x then

    μΛi,xsetμΛ,x\mu_{\Lambda_{i},x}\xrightarrow[]{set}\mu_{\Lambda,x}

    and

    μΛ,xjsetμΛ,x.\mu_{\Lambda,x_{j}}\xrightarrow[]{set}\mu_{\Lambda,x}.
Definition.

Given an operator projection family μ\mu and fL1(μ)f\in L^{1}(\mu) we define the integral of ff with respect to μ\mu as the operator f𝑑μB(X,X)\int f\;d\mu\in B(X,X^{\ast\ast}) induced by the bilinear bounded form given by

f𝑑μ:X×X(Λ,x)f𝑑μΛ,x.\begin{array}[]{cccc}\int f\;d\mu\colon&X^{\ast}\times X&\longrightarrow&\mathbb{C}\\ &(\Lambda,x)&\longmapsto&\int f\;d\mu_{\Lambda,x}\end{array}.
Definition.

Let XX be a Banach space, (Ω,Σ)(\Omega,\Sigma) a measurable space, and μ\mu an operator projection family. We say that fL(μ)f\in L^{\infty}(\mu) is properly integrable if

f𝑑μJ(X).\int f\;d\mu\in J(X).

We will use the following weak compactness theorem further ahead. See [9] for its proof.

Theorem III.3 (Grothendieck’s Lemma).

Let (X,||)(X,|\cdot|) be a normed space and AXA\subset X a bounded set. AA is relatively compact in the weak topology if and only if for every pair of sequences (xn)n(x_{n})_{n\in\mathbb{N}} in AA and (Λm)m(\Lambda_{m})_{m\in\mathbb{N}} in BXB_{X^{\ast}} such that the iterated limits

limnlimmΛm(xn)\lim_{n\to\infty}\lim_{m\to\infty}\Lambda_{m}(x_{n})

and

limmlimnΛm(xn),\lim_{m\to\infty}\lim_{n\to\infty}\Lambda_{m}(x_{n}),

exist the equality of iterated limits is satisfied, that is,

limmlimnΛm(xn)=limnlimmΛm(xn).\lim_{m\to\infty}\lim_{n\to\infty}\Lambda_{m}(x_{n})=\lim_{n\to\infty}\lim_{m\to\infty}\Lambda_{m}(x_{n}).

IV Smooth and Continuous Functional Calculi

IV.1 Vector Cauchy-Pompeiu Formula

We now extend the Cauchy-Pompeiu formula to the Banach-valued case.

Proposition IV.0.1 (Vector Cauchy-Pompeiu Formula).

Let XX be a Banach space and f:ΩXf\colon\Omega\subset\mathbb{C}\to X a CC^{\infty} function. The Cauchy-Pompeiu Formula is valid for ff in the following sense

f(λ)=12πiPωfzλ𝑑γ1πPωzfzλ𝑑.f(\lambda)=\frac{1}{2\pi i}P\int_{\partial\omega}\frac{f}{z-\lambda}\;d\gamma-\frac{1}{\pi}P\int_{\omega}\frac{\partial_{z^{\ast}}f}{z-\lambda}\;d\ell.
Proof.

The functions inside the integrals are continuous in the punctured set ωϵ=ωBϵ(0)\omega_{\epsilon}=\omega\setminus B_{\epsilon}(0), which has compact closure, hence the Pettis integrals on the right-hand side exist, over the set ωϵ\omega_{\epsilon}. A limiting procedure similar to the one used in the proof of the Cauchy-Pompeiu Formula (theorem II.1) shows that the integrals in the right-hand side exist as elements of XX^{\ast\ast}. This is because, despite the Dominated Convergence Theorem for the Pettis integral implies that the integral exists in XX, the Monotone Convergence Theorem for Pettis integral only concludes existence as an element of XX^{\ast\ast}. If ΛX\Lambda\in X^{\ast} then Λf:Ω\Lambda\circ f\colon\Omega\subset\mathbb{C}\to\mathbb{C} is a CC^{\infty} function, thus the Cauchy-Pompeiu Formula is valid and implies that

Λf(λ)=12πiωΛfzλ𝑑γ1πωzΛfzλ𝑑.\Lambda\circ f(\lambda)=\frac{1}{2\pi i}\int_{\partial\omega}\frac{\Lambda\circ f}{z-\lambda}\;d\gamma-\frac{1}{\pi}\int_{\omega}\frac{\partial_{z^{\ast}}\Lambda\circ f}{z-\lambda}\;d\ell.

The definition of the Dunford integral and this equation imply that

Λ(f(λ))=(12πiDωfzλ𝑑γ1πDωzfzλ𝑑)(Λ).\Lambda(f(\lambda))=\left(\frac{1}{2\pi i}D\int_{\partial\omega}\frac{f}{z-\lambda}\;d\gamma-\frac{1}{\pi}D\int_{\omega}\frac{\partial_{z^{\ast}}f}{z-\lambda}\;d\ell\right)(\Lambda).

Since ΛX\Lambda\in X^{\ast} is arbitrary, we conclude that

J(f(λ))=12πiJ(Pωfzλ𝑑γ)1πDωzfzλ𝑑.J(f(\lambda))=\frac{1}{2\pi i}J\left(P\int_{\partial\omega}\frac{f}{z-\lambda}\;d\gamma\right)-\frac{1}{\pi}D\int_{\omega}\frac{\partial_{z^{\ast}}f}{z-\lambda}\;d\ell.

This implies that

1πDωzfzλ𝑑\displaystyle\frac{1}{\pi}D\int_{\omega}\frac{\partial_{z^{\ast}}f}{z-\lambda}\;d\ell =12πiJ(Pωfzλ𝑑γ)J(f(λ))\displaystyle=\frac{1}{2\pi i}J\left(P\int_{\partial\omega}\frac{f}{z-\lambda}\;d\gamma\right)-J(f(\lambda))
=J(12πiPωfzλ𝑑γf(λ)),\displaystyle=J\left(\frac{1}{2\pi i}P\int_{\partial\omega}\frac{f}{z-\lambda}\;d\gamma-f(\lambda)\right),

hence the function zfzλ\frac{\partial_{z^{\ast}}f}{z-\lambda} is Pettis integrable with respect to \ell, thus we have

f(λ)=12πiPωfzλ𝑑γ1πPωzfzλ𝑑.f(\lambda)=\frac{1}{2\pi i}P\int_{\partial\omega}\frac{f}{z-\lambda}\;d\gamma-\frac{1}{\pi}P\int_{\omega}\frac{\partial_{z^{\ast}}f}{z-\lambda}\;d\ell.

The immediate corollary is that the integrals in the previous formula do not depend on the trajectory.

Corollary IV.0.1.

Let XX be a Banach space and f:ΩXf\colon\Omega\subset\mathbb{C}\to X a CC^{\infty} function. Given smooth curves γ1,2\gamma_{1,2} such that γ1,2Ω\gamma_{1,2}\subset\Omega and λintγ1,2\lambda\in int\;\gamma_{1,2} then

12πiγ1fzλ𝑑γ1πintγ1zfzλ𝑑=12πiγ2fzλ𝑑γ1πintγ2zfzλ𝑑,\frac{1}{2\pi i}\int_{\gamma_{1}}\frac{f}{z-\lambda}\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma_{1}}\frac{\partial_{z^{\ast}}f}{z-\lambda}\;d\ell=\frac{1}{2\pi i}\int_{\gamma_{2}}\frac{f}{z-\lambda}\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma_{2}}\frac{\partial_{z^{\ast}}f}{z-\lambda}\;d\ell,

that is, the integrals in the Vector Cauchy-Pompeiu Formula do not depend on the trajectory.

Once again, there is another immediate consequence.

Theorem IV.1 (Vector Cauchy-Pompeiu).

Let Ω\Omega be an open subset of \mathbb{C} with compact closure and f:ΩXf\colon\Omega\to X a CC^{\infty} function. The following statements are true:

  1. 1.

    For any γ\gamma simple closed curve in Ω\Omega we have

    12πiγf𝑑γ1πintγzfd=0.\frac{1}{2\pi i}\int_{\gamma}f\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma}\partial_{z^{\ast}}f\;d\ell=0.
  2. 2.

    If γ1\gamma_{1} and γ2\gamma_{2} are closed curves homotopic in Ω\Omega then

    12πiγf𝑑γ1πintγzfd=12πiγf𝑑γ1πintγzfd.\frac{1}{2\pi i}\int_{\gamma}f\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma}\partial_{z^{\ast}}f\;d\ell=\frac{1}{2\pi i}\int_{\gamma}f\;d\gamma-\frac{1}{\pi}\int_{int\;\gamma}\partial_{z^{\ast}}f\;d\ell.

    In particular, All integrals exist in XX^{\ast\ast}.

IV.2 Regular Elements and Smooth Functional Calculus

We now develop the functional calculi we will use to obtain the Spectral Theorem. We do this in the context of Banach algebras. Further ahead we apply a similar but less restrictive procedure to the case of operators in a Banach space, which, essentially, consists in applying these results in a pointwise manner, except for certain modifications. We trust the reader can provide these changes. We take this risk since the methods we develop also provide a version of the Spectral Theorem for Banach algebras.

Definition.

Let XX be a Banach algebra and xXx\in X. We say that xx is regular if for any Ω\Omega precompact neighborhood of σ(x)\sigma(x) the resolvent function RxR_{x} is Dunford integrable.

We will work with the class C(x)C^{\infty}(x) of CC^{\infty} functions defined in a precompact neighborhood of σ(x)\sigma(x).

Definition.

Let XX be a Banach algebra, xXx\in X regular, fC(x)f\in C^{\infty}(x) and γ\gamma a smooth curve such that σ(x)intγ\sigma(x)\subset int\;\gamma and γDomf\gamma\subset Dom\;f. We define f(x)f(x) as the element of XX^{\ast\ast} given by

f(x)=12πiDγf(λ)Rx(λ)𝑑γ(λ)1πDintγσ(x)zf(λ)Rx(λ)d(λ).f(x)=\frac{1}{2\pi i}D\int_{\gamma}f(\lambda)R_{x}(\lambda)\;d\gamma(\lambda)-\frac{1}{\pi}D\int_{int\;\gamma\setminus\sigma(x)}\partial_{z^{\ast}}f(\lambda)R_{x}(\lambda)\;d\ell(\lambda).

The application

C(x)Xff(x)\begin{array}[]{ccc}C^{\infty}(x)&\longrightarrow&X^{\ast\ast}\\ f&\longmapsto&f(x)\end{array}

is the Smooth Functional Calculus.

The Vector Cauchy-Pompeiu Formula implies that the previous definition does not depend on the trajectory. Despite f(x)Xf(x)\in X^{\ast\ast} we will write Λ(f(x))\Lambda(f(x)) instead of f(x)(Λ)f(x)(\Lambda), understanding that f(x)f(x) is actually an element of the bidual space.

The Smooth Functional Calculus possesses a natural continuity property.

Proposition IV.1.1.

Let XX be a Banach algebra and xXx\in X regular. If Ω\Omega is a neighbourhood of σ(x)\sigma(x) then the application

(C(Ω),||C1)fΛ(f(x))\begin{array}[]{ccc}(C^{\infty}(\Omega),|\cdot|_{C^{1}})&\longrightarrow&\mathbb{C}\\ f&\longmapsto&\Lambda(f(x))\end{array}

is continuous for each ΛX\Lambda\in X^{\ast}.

Proof.

It is enough to establish the continuity of the second integral. If (fn)n(f_{n})_{n\in\mathbb{N}} is a sequence in C(Ω)C^{\infty}(\Omega) such that fnC1ff_{n}\xrightarrow[]{C^{1}}f then

|intγσ(x)zfnΛRxdintγσ(x)zfΛRxd|\displaystyle\left|\int_{int\;\gamma\setminus\sigma(x)}\partial_{z^{\ast}}f_{n}\Lambda\circ R_{x}\;d\ell-\int_{int\;\gamma\setminus\sigma(x)}\partial_{z^{\ast}}f\Lambda\circ R_{x}\;d\ell\right| intγσ(x)|zfnzf||ΛRx|𝑑\displaystyle\leq\int_{int\;\gamma\setminus\sigma(x)}|\partial_{z^{\ast}}f_{n}-\partial_{z^{\ast}}f||\Lambda\circ R_{x}|\;d\ell
|zfnzf|intγσ(x)|ΛRx|𝑑\displaystyle\leq|\partial_{z^{\ast}}f_{n}-\partial_{z^{\ast}}f|_{\infty}\int_{int\;\gamma\setminus\sigma(x)}|\Lambda\circ R_{x}|\;d\ell
|fnf|C1intγσ(x)|ΛRx|𝑑.\displaystyle\leq|f_{n}-f|_{C^{1}}\int_{int\;\gamma\setminus\sigma(x)}|\Lambda\circ R_{x}|\;d\ell.

Since the last integral is finite the result follows. ∎

This continuity property is natural, because of the appearance of first derivatives, but the Spectral Theorem will require a stronger property. Essentially, we require an expression for the Smooth Functional Calculus that does not depend on derivatives. Recall that the Goldstine Theorem implies that J(X)J(X) is dense in XX^{\ast\ast}, thus f(x)f(x) can be realized as a limit of evaluations in XX in the topology τω\tau_{\omega^{\ast}}. To this end, we make the following construction: Given xXx\in X the spectrum σ(x)\sigma(x) is compact, hence, by a known result of differential topology, there exists a CC^{\infty} function HH such that

H1({0})=σ(x).H^{-1}(\{0\})=\sigma(x).

Sard’s Theorem implies that the set of critical values has null Lebesgue measure, thus there exists a sequence of real numbers (tn)n(t_{n})_{n\in\mathbb{N}} such that tn0t_{n}\to 0 and each tnt_{n} is a regular value. Define open sets

Ωn=H1([0,tn)),\Omega_{n}=H^{-1}([0,t_{n})),

which are neighbourhoods of σ(x)\sigma(x), satisfy Ωnσ(x)\Omega_{n}\searrow\sigma(x) and their boundaries

γn\displaystyle\gamma_{n} =Ωn\displaystyle=\partial\Omega_{n}
=H1({tn})\displaystyle=H^{-1}(\{t_{n}\})

are smooth curves. We also have that intγn=Ωnint\;\gamma_{n}=\Omega_{n}. Denote the measure induced by the volume form of γn\gamma_{n} by dγnd\gamma_{n}. From now on we will use this notation.

Proposition IV.1.2.

Let xXx\in X be regular and fC(x)f\in C^{\infty}(x). The equation

Λ(f(x))=12πilimnγnfΛRx𝑑γn\Lambda(f(x))=\frac{1}{2\pi i}\lim_{n\to\infty}\int_{\gamma_{n}}f\;\Lambda\circ R_{x}\;d\gamma_{n}

is satisfied for each ΛX\Lambda\in X^{\ast}. In particular, the limit on the right-hand side exists.

Proof.

By definition,

Λ(f(x))=12πiγnf(λ)ΛRx(λ)𝑑γ(λ)1πΩnσ(x)zf(λ)ΛRx(λ)d(λ)\Lambda(f(x))=\frac{1}{2\pi i}\int_{\gamma_{n}}f(\lambda)\Lambda\circ R_{x}(\lambda)\;d\gamma(\lambda)-\frac{1}{\pi}\int_{\Omega_{n}\setminus\sigma(x)}\partial_{z^{\ast}}f(\lambda)\Lambda\circ R_{x}(\lambda)\;d\ell(\lambda)

and does not depend on the trajectory γn\gamma_{n}. Since xx is regular the second integral in the right-hand side vanishes as nn\to\infty by the Monotone Convergence Theorem, which implies the desired conclusion. ∎

Notice the similarity of this result with theorem II.4. Before using this result we shift our attention to another fundamental concern. In order for the Smooth Functional Calculus to be of interest we need there to be enough regular elements in a Banach algebra. The following proposition provides a sufficient condition for regularity.

Proposition IV.1.3.

Let xXx\in X be an element such that (σ(x))=0\ell(\sigma(x))=0 and

|Rx(λ)|=1d(λ,σ(x)).|R_{x}(\lambda)|=\frac{1}{d(\lambda,\sigma(x))}.

From the fact that Rx(λ)R_{x}(\lambda) is Bochner integrable, it follows that xx is regular.

The hypothesis on the norm is satisfied, for example, by normal operators in Hilbert spaces. Thus, this result is valid, in particular, for normal operators in Hilbert spaces with null measure spectrum.

Proof.

It follows from the proposition II.3.1 that 1d(λ,σ(x))\frac{1}{d(\lambda,\sigma(x))} is integrable, hence our hypothesis implies that |Rx(λ)||R_{x}(\lambda)| is integrable, therefore Rx(λ)R_{x}(\lambda) is Bochner integrable. ∎

Another relatively common situation is that of analytic C0C_{0}-semigroup generators, since in this case there exists a non-negative function M0M\geq 0 that depends only on the argument of λ\lambda such that

|Rx(λ)|M(argλ)|λ|,|R_{x}(\lambda)|\leq\frac{M(arg\;\lambda)}{|\lambda|},

and in this case Rx(λ)R_{x}(\lambda) is Bochner integrable if MM is integrable, even if σ(x)\sigma(x) has non-null Lebesgue measure.

IV.3 Continuous Functional Calculus

The key to establishing the Spectral Theorem is to extend the Smooth Functional Calculus to continuous functions. We will do this through proposition IV.1.2. If xXx\in X is regular and fC(x)f\in C(x) then there exists a sequence of CC^{\infty} functions (fn)n(f_{n})_{n\in\mathbb{N}}, all defined in the same open subset as ff, such that fnunifff_{n}\xrightarrow[]{unif}f. The natural definition for f(x)f(x) is

Λ(f(x))=limnΛ(fn(x)).\Lambda(f(x))=\lim_{n\to\infty}\Lambda(f_{n}(x)).

Since the Cauchy-Pompeiu formula involves the derivative zfn\partial_{z^{\ast}}f_{n} this may not make sense. For this reason the formula of proposition IV.1.2, which does not involve derivatives, plays a crucial role.

Theorem IV.2 (Continuous Functional Calculus).

The Smooth Functional Calculus is continuous when its domain C(Ω)C^{\infty}(\Omega) is equipped with the supremum norm |||\cdot|_{\infty} and the codomain is equipped with the topology τω\tau_{\omega^{\ast}}. In consequence, the Smooth Functional Calculus admits a linear and continuous extension to C(Ω)C(\Omega) by density, that is, if fC(Ω)f\in C(\Omega) and (fn)n(f_{n})_{n\in\mathbb{N}} is a sequence in C(Ω)C^{\infty}(\Omega) such that fnunifff_{n}\xrightarrow[]{unif}f in Ω\Omega then

Λ(f(x))=limnΛ(fn(x)).\Lambda(f(x))=\lim_{n\to\infty}\Lambda(f_{n}(x)).
Proof.

Let fC(x)f\in C^{\infty}(x) and (fm)m(f_{m})_{m\in\mathbb{N}} be a sequence in C(x)C^{\infty}(x), all with a common domain Ω\Omega, such that fmunifff_{m}\xrightarrow[]{unif}f in Ω\Omega. By proposition IV.1.1 the limit

limnγnΛRx𝑑\lim_{n\to\infty}\int_{\gamma_{n}}\Lambda\circ R_{x}\;d\ell

exists, hence the sequence is bounded, say by a constant MM. For ΛX\Lambda\in X^{\ast} it follows that

2π|Λ(fm(x))Λ(f(x))|\displaystyle 2\pi|\Lambda(f_{m}(x))-\Lambda(f(x))| limnsupxγn|fm(x)f(x)|γnΛRx𝑑\displaystyle\leq\lim_{n\to\infty}\sup_{x\in\gamma_{n}}|f_{m}(x)-f(x)|\int_{\gamma_{n}}\Lambda\circ R_{x}\;d\ell
Mlimnsupxγn|fm(x)f(x)|\displaystyle\leq M\lim_{n\to\infty}\sup_{x\in\gamma_{n}}|f_{m}(x)-f(x)|
M|fmf|.\displaystyle\leq M|f_{m}-f|_{\infty}.

The right-hand side converges to zero, hence Λ(fm(x))Λ(f(x))\Lambda(f_{m}(x))\to\Lambda(f(x)). This implies that fm(x)ωf(x)f_{m}(x)\xrightarrow[]{\omega^{\ast}}f(x). ∎

The previous constructions have the defect that the function needs to be defined in a neighborhood of σ(x)\sigma(x). We now verify that only the values of the function in σ(x)\sigma(x) are important.

Lemma IV.2.1.

Let f,gC(x)f,g\in C(x) be such that fσ(x)=gσ(x)f\restriction_{\sigma(x)}=g\restriction_{\sigma(x)}, then f(x)=g(x)f(x)=g(x).

Proof.

It is enough to establish this for smooth functions. We have the estimate

2π|Λ(f(x))Λ(g(x))|\displaystyle 2\pi|\Lambda(f(x))-\Lambda(g(x))| =|limnγn(fg)ΛRx𝑑|\displaystyle=\left|\lim_{n\to\infty}\int_{\gamma_{n}}(f-g)\;\Lambda\circ R_{x}\;d\ell\right|
limnsupsγn|f(x)g(x)|γn|ΛRx|𝑑\displaystyle\leq\lim_{n\to\infty}\sup_{s\in\gamma_{n}}|f(x)-g(x)|\int_{\gamma_{n}}|\Lambda\circ R_{x}|\;d\ell
Mlimnsupsγn|f(x)g(x)|.\displaystyle\leq M\lim_{n\to\infty}\sup_{s\in\gamma_{n}}|f(x)-g(x)|.

Since ff and gg are continuous and equal in σ(x)\sigma(x) we have that the last limit vanishes by uniform continuity. Since ΛX\Lambda\in X^{\ast} is arbitrary we have that f(x)=g(x)f(x)=g(x). ∎

If fC(σ(x))f\in C(\sigma(x)) the Tietze Extension Theorem implies that ff can be continuously extended to any neighborhood of σ(x)\sigma(x) with the same norm and the previous lemma implies that if FF and GG are two such extensions then F(x)=G(x)F(x)=G(x), hence we can define f(x)f(x) as any of them. The function obtained this way is still continuous, since if fnunifff_{n}\xrightarrow[]{unif}f in σ(x)\sigma(x) then the functions ffnf-f_{n} can be extended to a common neighborhood in such a way that the extension has the same norm. Continuity in C(x)C(x) implies that fn(x)ωf(x)f_{n}(x)\xrightarrow[]{\omega^{\ast}}f(x) and continuity is established. The following proposition summarizes the previous results and gathers the different expressions that we have obtained for the Smooth and Continuous Functional Calculi.

Proposition IV.2.1 (Smooth and Continuous Functional Calculi).

Let XX be a Banach algebra and xXx\in X a regular element.

  1. 1.

    The Smooth Functional Calculus is the map C(x)XC^{\infty}(x)\to X^{\ast\ast} given by

    Λ(f(x))\displaystyle\Lambda(f(x)) =12πiγfΛRx𝑑12πintγσ(x)fΛRx𝑑\displaystyle=\frac{1}{2\pi i}\int_{\gamma}f\;\Lambda\circ R_{x}\;d\ell-\frac{1}{2\pi}\int_{int\;\gamma\setminus\sigma(x)}f\;\Lambda\circ R_{x}\;d\ell
    =limn12πiγnfΛRx.\displaystyle=\lim_{n\to\infty}\frac{1}{2\pi i}\int_{\gamma_{n}}f\;\Lambda\circ R_{x}.

    This map is continuous if the codomain is equipped with the topology τω\tau_{\omega^{\ast}}.

  2. 2.

    The Continuous Functional Calculus is the map C(σ(x))XC(\sigma(x))\to X^{\ast\ast} given by

    f(x)\displaystyle f(x) =limnfn(x),\displaystyle=\lim_{n\to\infty}f_{n}(x),

    where (fn)n(f_{n})_{n\in\mathbb{N}} is any sequence of smooth functions with a neighbourhood of σ(x)\sigma(x) as a common domain such that fnunifff_{n}\xrightarrow[]{unif}f. This maps is continuous when considered as a funciton (C(σ(x)),||)(X,τω)(C(\sigma(x)),|\cdot|_{\infty})\to(X,\tau_{\omega^{\ast}}).

V Spectral Theorem in Banach Spaces

We can finally establish our generalizations of the Spectral Theorem. The first version is for regular elements of a Banach algebra. This applies, in particular, to the case of bounded linear operators in Banach spaces. Regularity is too restrictive of a condition in this context, hence our second version of the Spectral Theorem requires a less restrictive condition for operators in Banach spaces. In this case, the result will be valid for a certain class of operators called pointwise regular, which includes normal operators in Hilbert spaces.

V.1 Spectral Theorem in Banach Algebras

The previous result implies that the functional

C(σ(x))fΛ(f(x))\begin{array}[]{ccc}C(\sigma(x))&\longrightarrow&\mathbb{C}\\ f&\longmapsto&\Lambda(f(x))\end{array}

defines an element of C(σ(x))C(\sigma(x))^{\ast} for each ΛX\Lambda\in X^{\ast}. The Riesz-Markov-Kakutani Theorem implies that there exists a unique Borel measure μΛx\mu^{x}_{\Lambda} in σ(T)\sigma(T) such that

Λ(f(x))=σ(x)f𝑑μΛx.\Lambda(f(x))=\int_{\sigma(x)}f\;d\mu^{x}_{\Lambda}.

We have proven the following.

Theorem V.1.

Let XX be a Banach algebra and xXx\in X regular. There exists a unique family of measures

μx={μΛx|ΛX}\mu_{x}=\{\mu^{x}_{\Lambda}\;|\;\Lambda\in X^{\ast}\}

in σ(x)\sigma(x) such that

Λ(f(x))=σ(T)f𝑑μΛx\Lambda(f(x))=\int_{\sigma(T)}f\;d\mu^{x}_{\Lambda} (4)

for each ΛX\Lambda\in X^{\ast} and fC(σ(x))f\in C(\sigma(x)). In particular,

Λ(x)\displaystyle\Lambda(x) =σ(T)I𝑑dμΛx\displaystyle=\int_{\sigma(T)}Id\;d\mu^{x}_{\Lambda}
=σ(T)λ𝑑μΛx(λ).\displaystyle=\int_{\sigma(T)}\lambda\;d\mu^{x}_{\Lambda}(\lambda).

We will call μx\mu^{x} the spectral family of xx. We now show that μx\mu^{x} is a vector projection family.

Proposition V.1.1.

Let XX be a Banach algebra, xXx\in X regular and μx\mu^{x} its spectral family. The function

ΛμΛx\Lambda\longmapsto\mu^{x}_{\Lambda}

is linear.

Proof.

Given Λ,ΦX\Lambda,\Phi\in X^{\ast} and cc\in\mathbb{C}, we have the following estimate for each fC(σ(x))f\in C(\sigma(x))

σ(x)f𝑑μΛ+cΦx\displaystyle\int_{\sigma(x)}f\;d\mu^{x}_{\Lambda+c\Phi} =(Λ+cΦ)(f(x))\displaystyle=(\Lambda+c\Phi)(f(x))
=Λ(f(x))+cΦ(f(x))\displaystyle=\Lambda(f(x))+c\Phi(f(x))
=σ(x)f𝑑μΛx+cσ(x)f𝑑μΦx\displaystyle=\int_{\sigma(x)}f\;d\mu^{x}_{\Lambda}+c\int_{\sigma(x)}f\;d\mu^{x}_{\Phi}
=σ(x)fd(μΛx+cμΦx),\displaystyle=\int_{\sigma(x)}f\;d(\mu^{x}_{\Lambda}+c\mu^{x}_{\Phi}),

that is, for each fC(σ(x))f\in C(\sigma(x)) we have that

σ(x)fd(μΛ+cΦxμΛxcμΦx)=0.\int_{\sigma(x)}f\;d(\mu^{x}_{\Lambda+c\Phi}-\mu^{x}_{\Lambda}-c\mu^{x}_{\Phi})=0.

Since this is a Borel measure, this can only happen if

μΛ+cΦx=μΛx+cμΦx.\mu^{x}_{\Lambda+c\Phi}=\mu^{x}_{\Lambda}+c\mu^{x}_{\Phi}.

It remains only to prove the continuity of the spectral family μx\mu^{x}. This will follow from the continuity of the Continuous Functional Calculus.

Theorem V.2.

Let XX be a Banach algebra, xXx\in X regular and μx\mu^{x} its spectral family. The spectral family is a vector projection family.

Proof.

If gC(σ(x))g\in C(\sigma(x)) then the function ΛΛ(f(x))\Lambda\longmapsto\Lambda(f(x)) is linear and bounded since

σ(x)g𝑑μΛx=Λ(g(x))\int_{\sigma(x)}g\;d\mu^{x}_{\Lambda}=\Lambda(g(x))

and the right-hand side is linear and bounded as a function of Λ\Lambda. This implies that the same is valid for elements of L(μx)L^{\infty}(\mu^{x}), therefore the family μx\mu^{x} is a vector projection family. ∎

We can finally state the Spectral Theorem, the proof of which is given by all of the previous results.

Theorem V.3 (Spectral Theorem in Banach Algebras).

Let XX be a Banach algebra and xXx\in X regular. There exists a unique vector projection family μx\mu^{x} in σ(x)\sigma(x) such that

f(x)=σ(x)f𝑑μxf(x)=\int_{\sigma(x)}f\;d\mu^{x}

for each fC(σ(x))f\in C(\sigma(x)). In particular,

x\displaystyle x =σ(x)I𝑑dμx\displaystyle=\int_{\sigma(x)}Id\;d\mu^{x}
=σ(x)λ𝑑μx(λ).\displaystyle=\int_{\sigma(x)}\lambda\;d\mu^{x}(\lambda).

In virtue of the previous theorem, we can apply the theory of vector projection families to the spectral family, In particular, if fL1(μ)f\in L^{1}(\mu) the integral

σ(x)f𝑑μx\int_{\sigma(x)}f\;d\mu^{x}

defines an element of XX^{\ast\ast}. This makes the following definition possible.

Definition.

Let XX be a Banach algebra, xXx\in X regular and μx\mu^{x} its spectral family. We define the Borel Functional Calculus associated to xx as the function that to each fL1(μx)f\in L^{1}(\mu^{x}) associates the element of XX^{\ast\ast} given by its integral σ(T)f𝑑μx\int_{\sigma(T)}f\;d\mu^{x}, that is,

f(x)=σ(x)f𝑑μx.f(x)=\int_{\sigma(x)}f\;d\mu^{x}.

Note that the Borel Functional Calculus is an extension of the Holomorphic Functional Calculus. Furthermore, if XX is reflexive we have the following.

Proposition V.3.1.

Let XX be a Banach algebra, xXx\in X regular and μx\mu^{x} its spectral family. If XX is reflexive then each element L1(μx)L^{1}(\mu^{x}) is properly integrable.

Proposition V.3.2.

Let xXx\in X be a regular element and μx\mu^{x} its spectral family. The following conditions are equivalent:

  1. 1.

    The family μx\mu^{x} is induced by a vector measure νx\nu^{x}, in the sense that μΛx=Λνx\mu^{x}_{\Lambda}=\Lambda\circ\nu^{x} for each ΛX\Lambda\in X^{\ast}.

  2. 2.

    Every function in L(μx)L^{\infty}(\mu^{x}) is properly integrable.

  3. 3.

    For each measurable set EE the function ΛμΛx(E)\Lambda\longmapsto\mu^{x}_{\Lambda}(E) is continuous with respect to τω\tau_{\omega^{\ast}}.

V.2 Spectral Theorem for Operators in Banach Spaces

The Spectral Theorem for Banach algebras can be directly applied to the Banach algebra B(X)B(X). If TB(X)T\in B(X) then TT is regular if RT:ρ(T)B(X)R_{T}\colon\rho(T)\to B(X) is Dunford integrable, which may be too restrictive for most cases. For this reason, we modify the previous construction to improve the Spectral Theorem in this setting.

Definition.

Let XX be a Banach space and TB(X)T\in B(X). We will say that TT is pointwise regular if for each xXx\in X the function RTx:ρ(T)XR_{T}^{x}\colon\rho(T)\to X given by

RTx(λ)=RT(λ)(x)R_{T}^{x}(\lambda)=R_{T}(\lambda)(x)

is Dunford integrable.

Note that the function RTxR_{T}^{x} is always analytic in its domain and singular in σ(T)\sigma(T). Also note that if TT is regular then it is pointwise regular. When we developed the Smooth and Continuous Functional Calculi we only used the fact that XX was a Banach algebra to define RxR_{x} and the fact that it is analytic in its domain. Therefore, we can define the Smooth and Continuous Functional Calculi for pointwise regular operators.

Definition (Smooth and Continuous Functional Calculi).

Let XX be a Banach space and TB(X)T\in B(X) pointwise regular.

  1. 1.

    The Smooth Functional Calculus is the map C(T)L(X,X)C^{\infty}(T)\to L(X,X^{\ast\ast}) given by

    Λ(f(T)(x))\displaystyle\Lambda(f(T)(x)) =12πiγfΛRTx𝑑γ12πintγσ(x)fΛRTx𝑑\displaystyle=\frac{1}{2\pi i}\int_{\gamma}f\;\Lambda\circ R_{T}^{x}\;d\gamma-\frac{1}{2\pi}\int_{int\;\gamma\setminus\sigma(x)}f\;\Lambda\circ R_{T}^{x}\;d\ell
    =limn12πiγnfΛRTx𝑑γn.\displaystyle=\lim_{n\to\infty}\frac{1}{2\pi i}\int_{\gamma_{n}}f\;\Lambda\circ R_{T}^{x}\;d\gamma_{n}.
  2. 2.

    The Continuous Functional Calculus is the map C(σ(x))L(X,X)C(\sigma(x))\to L(X,X^{\ast\ast}) given by

    Λ(f(T)(x))\displaystyle\Lambda(f(T)(x)) =limnΛ(fn(x)),\displaystyle=\lim_{n\to\infty}\Lambda(f_{n}(x)),

    where (fn)n(f_{n})_{n\in\mathbb{N}} is any sequence of smooth functions with a neighbourhood of σ(x)\sigma(x) as their common domain such that fnunifff_{n}\xrightarrow[]{unif}f.

The proof that the images belong to XX^{\ast\ast} is formally identical to the corresponding result for Banach algebras, replacing RxR_{x} by RTxR_{T}^{x}. Naturally, one must first define the Smooth Functional Calculus and then define the Continuous Functional Calculus. Once again, the proofs are identical. The only difficulty not present in the previous setting is continuity with respect to xx.

Theorem V.4.

If fC(σ(T))f\in C(\sigma(T)) then f(T)B(X,X)f(T)\in B(X,X^{\ast\ast}).

Proof.

If ff is continuous then f(T)f(T) is the limit of a sequence of smooth functions applied to TT, it is thus enough to show the result is valid if ff is smooth. The proof consists in applying the Closed Graph Theorem twice.

We first show that the application xΛRTxx\longmapsto\Lambda\circ R_{T}^{x} is continuous from XX to C(Ω)C(\Omega) for any Ω\Omega neighbourhood of σ(T)\sigma(T). Let (xm)m(x_{m})_{m\in\mathbb{N}} be a sequence in XX such that xmxx_{m}\to x and ΛRTxnunifh\Lambda\circ R_{T}^{x_{n}}\xrightarrow[]{unif}h in Ω\Omega. Given λρ(T)\lambda\in\rho(T) we have that RTxm(λ)=RT(λ)(x)R_{T}^{x_{m}}(\lambda)=R_{T}(\lambda)(x) and RT(λ)B(X)R_{T}(\lambda)\in B(X), hence RT(λ)(xm)RT(λ)(x)R_{T}(\lambda)(x_{m})\to R_{T}(\lambda)(x) and RTxmpwRTxR_{T}^{x_{m}}\xrightarrow[]{pw}R_{T}^{x}. It follows that ΛRTxmpwΛRTx\Lambda\circ R_{T}^{x_{m}}\xrightarrow[]{pw}\Lambda\circ R_{T}^{x}. Since ΛRTxmunifh\Lambda\circ R_{T}^{x_{m}}\xrightarrow[]{unif}h in Ω\Omega we have that h=ΛRTxh=\Lambda\circ R_{T}^{x}. The Closed Graph Theorem implies the continuity of this map.

We now show that f(T)B(X,X)f(T)\in B(X,X^{\ast\ast}). Suppose that xnxx_{n}\to x and f(T)(xm)XΦf(T)(x_{m})\xrightarrow[]{X^{\ast\ast}}\Phi for a certain ΦX\Phi\in X^{\ast\ast}. We must show that Φ=f(T)(x)\Phi=f(T)(x). On the one hand, we have that

Φ(Λ)\displaystyle\Phi(\Lambda) =limmΛ(f(T)(xm))\displaystyle=\lim_{m\to\infty}\Lambda(f(T)(x_{m}))
=limmlimnγnfΛRTxm𝑑γn\displaystyle=\lim_{m\to\infty}\lim_{n\to\infty}\int_{\gamma_{n}}f\;\Lambda\circ R_{T}^{x_{m}}\;d\gamma_{n}

and this iterated limit exists. On the other hand, the continuity of the map xΛRTxx\longmapsto\Lambda\circ R_{T}^{x} from XX to C(Ω)C(\Omega) implies that fΛRTxmuniffΛRTxf\;\Lambda\circ R_{T}^{x_{m}}\xrightarrow[]{unif}f\;\Lambda\circ R_{T}^{x} en Ω\Omega. It follows that

limnlimmγnfΛRTxm𝑑γn\displaystyle\lim_{n\to\infty}\lim_{m\to\infty}\int_{\gamma_{n}}f\;\Lambda\circ R_{T}^{x_{m}}\;d\gamma_{n} =limnγnfΛRTx𝑑γn\displaystyle=\lim_{n\to\infty}\int_{\gamma_{n}}f\;\Lambda\circ R_{T}^{x}\;d\gamma_{n}
=Λ(f(T)(x))\displaystyle=\Lambda(f(T)(x))

and this iterated limit exists. The sequence of functions (fΛRTxm)m(f\;\Lambda\circ R_{T}^{x_{m}})_{m\in\mathbb{N}} has precompact image as it is convergent. We define functionals in C(Ω)C(\Omega)^{\ast} as

Φn(g)=γnfg𝑑γn.\Phi_{n}(g)=\int_{\gamma_{n}}fg\;d\gamma_{n}.

The function ff is smooth, thus theorem II.4 implies that the pointwise limit of the functionals (Φn)n(\Phi_{n})_{n\in\mathbb{N}} exists and defines an element of C(Ω)C(\Omega)^{\ast}. In particular, the sequence of functionals (Φn)n(\Phi_{n})_{n\in\mathbb{N}} is bounded.

The previous considerations imply that Grothendieck’s Lemma can be applied to these sequences to provide the equality of iterated limits, therefore

Φ(Λ)=Λ(f(T)(x)).\Phi(\Lambda)=\Lambda(f(T)(x)).

This implies that Φ=f(T)(x)\Phi=f(T)(x) and we conclude by the Closed Graph Theorem. ∎

The continuity of the Continuous Functional Calculus is now verified pointwise, that is, for each ΛX\Lambda\in X^{\ast} and xXx\in X the functional

C(σ(T))fΛ(f(T)(x))\begin{array}[]{ccc}C(\sigma(T))&\longrightarrow&\mathbb{C}\\ f&\longmapsto&\Lambda(f(T)(x))\end{array} (5)

is continuous. In the case f(T)B(X)f(T)\in B(X), this is just continuity in the weak operator topology. The proof is once again identical to the one for Banach algebras. Even though this property is weaker than the one for Banach algebras, it suffices to establish the Spectral Theorem.

Theorem V.5 (Spectral Theorem for Operators in Banach Spaces).

Let XX be a Banach space and TB(X)T\in B(X) a pointwise regular operator. There exists a unique operator projection family μT\mu^{T} in σ(T)\sigma(T) such that

f(T)=σ(T)f𝑑μTf(T)=\int_{\sigma(T)}f\;d\mu^{T}

for each fC(σ(T))f\in C(\sigma(T)). In particular,

T\displaystyle T =σ(T)I𝑑dμT\displaystyle=\int_{\sigma(T)}Id\;d\mu^{T}
=σ(T)λ𝑑μT.\displaystyle=\int_{\sigma(T)}\lambda\;d\mu^{T}.

The existence of the measures follows from the continuity of the application

(f,Λ,x)Λ(f(T)(x))(f,\Lambda,x)\longmapsto\Lambda(f(T)(x))

in the first variable. The fact μT\mu^{T} is an operator projection family follows from the continuity of the last two variables. At this moment we trust that it is clear that all of our results follow from the linearity and separate continuity of the previous function, the proofs for which we have already provided in a slightly different setting.

Definition.

Let XX be a Banach and TB(X)T\in B(X) pointwise regular. We define the Borel Functional Calculus as the application that to each fL1(μT)f\in L^{1}(\mu^{T}) associates the operator

f(T)=σ(T)f𝑑μTf(T)=\int_{\sigma(T)}f\;d\mu^{T}

in B(X,X)B(X,X^{\ast\ast}).

Once again, if XX is reflexive then f(T)f(T) is actually an operator in B(X)B(X).

Proposition V.5.1.

Let TB(X)T\in B(X) be a pointwise regular operator and μT\mu^{T} its spectral family. The following conditions are equivalent:

  1. 1.

    The family μT\mu^{T} is induced by an operator-valued measure νT\nu^{T}, in the sense that μΛ,xT(E)=Λ(νT(E)(x))\mu^{T}_{\Lambda,x}(E)=\Lambda(\nu^{T}(E)(x)) for each ΛX\Lambda\in X^{\ast} and xXx\in X.

  2. 2.

    Each function in L(μT)L^{\infty}(\mu^{T}) is properly integrble.

  3. 3.

    For each measurable set EE the function (Λ,x)μΛ,xT(E)(\Lambda,x)\longmapsto\mu^{T}_{\Lambda,x}(E) is continuos with respect to τω\tau_{\omega^{\ast}} in the first variable and with respect to τω\tau_{\omega} in the second one.

Lastly, we show that our second version of the Spectral Theorem is an extension of the results for Hilbert spaces.

Proposition V.5.2.

Let HH be a Hilbert space and TB(H)T\in B(H). If TT is normal then TT is pointwise regular.

Proof.

The Spectral Theorem for operators in Hilbert spaces implies that there exists a unique resolution of the identity ETE^{T} such that

y,f(T)(x)=σ(T)f𝑑Ey,xT\langle y,f(T)(x)\rangle=\int_{\sigma(T)}f\;dE^{T}_{y,x}

for each fC(σ(T))f\in C(\sigma(T)). It follows that

Ωσ(T)|y,RT(λ)(x)|𝑑λ\displaystyle\int_{\Omega\setminus\sigma(T)}|\langle y,R_{T}(\lambda)(x)\rangle|\;d\lambda =Ωσ(T)|σ(T)1zλ𝑑Ey,xT(z)|𝑑λ\displaystyle=\int_{\Omega\setminus\sigma(T)}\left|\int_{\sigma(T)}\frac{1}{z-\lambda}\;dE^{T}_{y,x}(z)\right|\;d\lambda
σ(T)Ωλ1|zλ|𝑑λd|Ey,xT|(z)\displaystyle\leq\int_{\sigma(T)}\int_{\Omega\setminus\lambda}\frac{1}{|z-\lambda|}\;d\lambda\;d|E^{T}_{y,x}|(z)
=C|Ey,xT|(σ(T)),\displaystyle=C|E^{T}_{y,x}|(\sigma(T)),

proving the integral is finite. It follows that RTxR_{T}^{x} is Dunford integrable for each xHx\in H and TT, hence pointwise regular. ∎

Acknowledgements

This work was supported by DGAPA-UNAM, grant No. IN108225.

Data Availability Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Declarations

Conflict of interest The authors declare that they have no conflict of interest.

References

  • Bar [56] Robert G. Bartle. A general bilinear vector integral. Studia Mathematica, 15(3):337–352, 1956.
  • Cha [84] Isaac Chavel. Eigenvalues in Riemannian Geometry. Academic Press, 1984.
  • Con [90] John B. Conway. A Course in Functional Analysis. Springer-Verlag, 1990.
  • CQ [24] Luis A. Cedeño and Hernando Quevedo. A general theory of operator-valued measures, 2024.
  • Dav [76] E. B. Davies. Quantum Theory of Open Systems. Academic Press, 1976.
  • Die [77] Joseph Diestel. Vector Measures. American Mathematical Society, 1977.
  • Din [67] N. Dinculeanu. Vector Measures. Pergamon Press, 1967.
  • Gra [77] William H. Graves. On the theory of vector measures. Memoirs of the American Mathematical Society, 12(195), 1977.
  • Hol [75] Richard B. Holmes. Geometric Functional Analysis and its Applications. Springer-Verlag, 1975.
  • Hor [90] Lars Hormander. An Introduction to Complex Analysis in Several Variables. North-Holland, 1990.
  • Lew [70] Daniel Ralph Lewis. Integration with respect to vector measures. Pacific Journal of Mathematics, 33(1):157–165, 1970.
  • Rud [91] Walter Rudin. Functional Analysis. McGrawhill, 1991.