Acceso abierto

Transfunctions Applied to Plans, Markov Operators and Optimal Transport

 y   
29 oct 2024

Cite
Descargar portada

Introduction

Let ℳX and ℳY be vector spaces of finite signed measures defined on measurable spaces (X, ∑X) and (Y, ∑Y ), respectively. For finite positive measure µ on (X, ∑X) and for real-valued function f ∈ ℒ1(X, µ), let denote the measure A ↦ ∫A f dµ and define μp,+={fμ:fp(X,μ),f0}andμp:={fμ:fp(X,μ)}, {\cal M}_\mu ^{p, + } = \{ f\mu :\,f \in \,{\cal L^p}(X,\mu ),\,f \ge 0\} \,{\rm{and}}\,{\cal M}_\mu ^p: = \{ f\mu :\,f \in {\cal L^p}(X,\mu )\} , for p ∈ [1, ∞]. We define ℳµ to be the set of all finite signed measures absolutely continuous with respect to µ. By the Radon–Nikodym Theorem, μ=μ1 {\cal M_\mu } = {\cal M}_\mu ^1 . Similarly, we define vp,+ {\cal M}_v^{p, + } and vp {\cal M_v^p} for finite positive measure ν on (Y, ∑Y ).

A transfunction is any function Φ: ℳX → ℳY , [8]. Strongly σ-additive transfunctions are those which are linear and continuous with respect to total variation. We will sometimes call an operator between Banach spaces strongly σ-additive if it is linear and norm-continuous.

Plans have applications for finding weak solutions for optimal transport problems, [10]. Markov operators, defined in Section 2, have some similarities to stochastic matrices, [4]. Plans and Markov operators have a bijective correspondence as described in [9] and in Section 2. We assign to any corresponding Markov operator/plan pair (T, κ) with marginals µ, ν a unique transfunction Φ: ℳµ → ℳν – called a Markov transfunction. However, each Markov transfunction corresponds to a family of Markov operators (resp. plans) which have different marginals but follow the same “instructions”. Φ, T , and κ are related via the equalities Φ(1Aμ)(B)=BT(1A)dv=κ(A×B),YgdΦ(fμ)=YT(f)d(gv)=X×Y(fg)dκ \eqalign{ & \Phi ({1_A}\mu )(B) = \int_B {T({1_A})\,dv = \kappa (A\, \times \,B)} , \cr & \int_Y {g\,d\Phi (f\mu )} = \int_Y {T(f)\,d(gv)} = \,\int_{X \times Y} {(f \otimes g)\,d\kappa } \cr} which hold for all AX, BY , f ∈ ℒ(X), and g ∈ ℒ (Y ). The first set of equalities, although simpler, imply the second set of equalities by strong σ-additivity of Φ, bounded-linearity of T , and σ-additivity of κ.

In our investigation of transfunctions we are motivated by the theory developed for the Monge-Kantorovich transportation problems and their far-reaching outcomes; see [1], [5], [6], and [10].

Let ℱX and ℱY be spaces of measurable functions which are integrable by measures in ℳX and ℳY , respectively. If {ℱX, ℳX} and {ℱY , ℳY } are separating pairs with respect to integration as defined in Section 3, then we define the Radon adjoint of Φ: ℳX → ℳY (if it exists) to be the unique linear bounded operator Φ* : ℱY → ℱX such that XΦ(g)dλ=YgdΦ(λ) \int_X {{\Phi ^ * }} \,(g)\,d\lambda = \int_Y {\,g\,d\Phi (\lambda )} for all g ∈ ℱY and λ ∈ ℳX.

If X and Y are second-countable locally compact Hausdorff spaces, if ℱX and ℱY are Banach spaces of bounded continuous functions (uniform norm) and if ℳX and ℳY are Banach spaces of finite regular signed measures (total variation), then any strongly σ-additive weakly-continuous transfunction Φ: ℳX → ℳY has an adjoint Φ* which is a linear, uniformly-continuous and bounded-pointwise-continuous operator (and vice versa) such that ||Φ|| = ||Φ*||. When X and Y are also non-atomic, ℳX and ℳY include measures which are non-atomic and strictly-positive; see [2].

In future research, we wish to develop functional analysis on transfunctions, and adjoints may be utilized to this end. In contexts where operators on functions are more appropriate or preferable, the adjoint may prove crucial.

A simple transfunction Φ: ℳX → ℳY is one which has the form Φ(λ):=i=1mfi,λρi \Phi (\lambda )\,: = \sum\limits_{i = 1}^m {\left\langle {{f_i},\,\lambda } \right\rangle } {\rho _i} for f1, . . . , fm X and ρ1, . . . , ρm ∈ ℳY , where 〈fi, λ〉 := ∫X fi dλ. Simple transfunctions are weakly-continuous and strongly σ-additive. When working with locally compact Polish (metric) spaces, simple Markov transfunctions have two advantages: they weakly approximate all Markov transfunctions, and a subclass of them can be utilized to approximate the optimal cost between two marginals with respect to a transport cost c(x, y) that is bounded by αd(x, y)p for constants α, p > 0.

In [3] the notions of localization of a transfunction and the graph of a transfunction are introduced and studied. They give us an insight into which transfunctions arise from continuous functions or measurable functions or are close to such functions.

This paper is based on part of Bentley’s PhD dissertation.

Markov transfunctions

In this section, we describe a class of transfunctions in which each transfunction corresponds to a family of plans and a family of Markov operators. First, we introduce these concepts. All measurable or continuous functions shall be real-valued in this text. Note that the following definitions allow for all finite positive measures rather than all probability measures.

Definition 2.1

Let µ and ν be finite positive measures on (X, ∑X) and (Y, ∑Y ) respectively with ||µ|| = ||ν||. Let κ be a finite positive measure on the product measurable space (X × Y, ∑X×Y ). We say that κ is a plan with marginals µ and ν if κ(A × Y ) = µ(A) and κ(X × B) = ν(B) for all A ∈ ∑X and B ∈ ∑Y . We define Π(µ, ν) to be the set of all plans with marginals µ and ν.

If random variables X, Y have laws µ, ν, then any coupling of X, Y has a law κ which is a plan in Π(µ, ν).

Definition 2.2

Let µ and ν be finite positive measures on (X, ∑X) and (Y, ∑Y ) respectively with ||µ|| = ||ν||, and let p ∈ [1, ∞]. We say that a function T : ℒp(X, µ) → ℒp(Y, ν) is a Markov operator if:

T is linear with T 1X = 1Y ;

f ≥ 0 implies T f ≥ 0 for all f ∈ ℒp(X, µ);

X f dµ = ∫Y T f dν for all f ∈ ℒp(X, µ).

Notice that the definition of Markov operators depends on underlying measures µ and ν on X and Y respectively, even when p = ∞. We now define some properties for transfunctions that are analogous to (ii) and (iii) from Definition 2.2.

Definition 2.3

Let Φ: ℳX Y be a transfunction.

Φ is positive if λ ≥ 0 implies that Φ λ ≥ 0 for all λ ∈ ℳX.

Φ is measure-preserving if (Φλ)(Y ) = λ(X) for all λ ∈ ℳX.

Φ is Markov if it is strongly σ-additive, positive and measure-preserving.

By [9], there is a bijective relationship between plans and Markov operators. We will show soon that a relationship between Markov operators and Markov transfunctions exists, which will imply that all three concepts are connected.

Lemma 2.4

Let µ be a finite positive measure on (X, ∑X), and define the map Jµ : ℒ1(X, µ) → ℳµ via Jµf := fµ. Then Jµ (hence Jμ-1 J_\mu ^{ - 1} ) is a positive linear isometry.

Proof

Positivity and linearity of integrals with respect to µ ensure that Jµ is positive and linear. Surjectivity of Jµ is the statement of the Radon–Nikodym Theorem. Injectivity and isometry hold because Jμf=Jμ(f+)Jμ(f)=Xf+dμ+Xfdμ=Xfdμ=f1. \eqalign{ & \left\| {{J_\mu }f} \right\| & = \left\| {{J_\mu }({f^ + }) - {J_\mu }({f^ - })} \right\| = \int_X {{f^ + }d\mu + \int_X {{f^ - }d\mu } } \cr & & = \int_X {\left| f \right|} d\mu = {\left\| f \right\|_1}. \cr}

Theorem 2.5

Let µ and ν be finite positive measures on X and Y respectively, with ||µ|| = ||ν|| and let s ∈ [1, ∞]. For every Markov operator T : ℒs(X, µ) → ℒs(Y, ν), there exists a unique Markov transfunction Φ:μsνs \Phi \:\,{\cal M}_\mu ^s \to {\cal M}_\nu ^s such that BT(1A)dν=Φ(1Aμ)(B) \int_B {T({1_A})} d\nu = \Phi ({1_A}\mu )(B) for all A ∈ ∑X and B ∈ ∑Y .

Every Markov transfunction Φ:μsνs \Phi :\,{\cal M}_\mu ^s \to {\cal M}_\nu ^s corresponds to a family of Markov operators {Tλ,ρ : ℒ(X, λ) → ℒ(Y, ρ) | λμs,+ \lambda \in {\cal M_\mu ^{s, + }} , ρ = Φλ } which satisfies BTλ,ρ(1A)dρ=Φ(1Aλ)(B) \int_B {{T_{\lambda ,\rho }}({1_A})d\rho = \Phi ({1_A}\lambda )} (B) for all A ∈ ∑X and B ∈ ∑Y .

Proof

First, we prove each statement for s = 1, then extend the argument to other values of s. Let T : ℒ1(X, µ) → ℒ1(Y, ν) be a Markov operator. Define Φ=JνTJμ1 \Phi = {J_\nu }\,T\,J_\mu ^{ - 1} . Since all three operators in the definition of Φ are positive and strongly σ-additive, we see that Φ is also positive and strongly σ-additive. Next, if λ ∈ ℳµ, then (Φλ)(Y)=Jν(TJμ1λ)(Y)=YT(Jμ1λ)dν=XJμ1(λ)dμ=λ(X) (\Phi \lambda )(Y) = {J_\nu }(TJ_\mu ^{ - 1}\lambda )(Y) = \int_Y {T(J_\mu ^{ - 1}\lambda )d\nu = } \int_X {J_\mu ^{ - 1}} (\lambda )d\mu = \lambda (X) by the definitions of isometries Jμ1 J_\mu ^{ - 1} and Jν, and by property (iii) of T , so Φ is measure-preserving. Finally, notice that Φ(1Aμ)(B)=JνT(Jμ1(1Aμ))(B)=Jν(T1A)(B)=BT(1A)dν \Phi ({1_A}\mu )(B) = {J_\nu }T(J_\mu ^{ - 1}({1_A}\mu ))(B) = {J_\nu }(T{1_A})(B) = \int_B {T({1_A})d\nu } for all A ∈ ∑X and B ∈ ∑Y , hence the relation holds.

Now let s ∈ (1, ∞] and let T : ℒs(X, µ) → ℒs(Y, ν) be a Markov operator. By Theorem 1 from [9], T can be uniquely extended to a Markov operator on ℒ1(X, µ). By our previous argument, corresponds to a Markov transfunction Φ̂ defined on ℳµ. We define Φ to be the restriction of Φ̂ to μs {\cal M}_\mu ^s . The necessary properties are inherited from the previous argument.

Now we prove the second statement. Let s ∈ [1, ∞], let Φ:μsνs \Phi :\,{\cal M}_\mu ^s \to {\cal M}_\nu ^s be a Markov transfunction, let λμs \lambda \in {\cal M}_\mu ^s be positive, and define ρ:=Φ(λ)νs \rho \,: = \,\Phi (\lambda ) \in {\cal M}_\nu ^s , which is also positive. Define T=Tλ,ρ:=Jρ1ΦJλ T = {T_{\lambda ,\rho }}\,: = J_\rho ^{ - 1}\,\Phi \,{J_\lambda } with domain L(X, λ). Then T(1X)=Jρ1Φ(Jλ(1X))=Jρ1(Φλ)=Jρ1ρ=1Y. T({1_X}) = J_\rho ^{ - 1}\Phi ({J_\lambda }({1_X})) = J_\rho ^{ - 1}(\Phi \lambda ) = J_\rho ^{ - 1}\rho = {1_Y}.

Since all three operators in the definition of T are positive and strongly σ-additive, we see that T is also positive and strongly σ-additive, satisfying parts (i) and (ii) of Definition 2.2. Next, if f ∈ ℒ(X, λ), then YTfdρ=YJρ1(ΦJλf)dρ=(Φ(Jλf))(Y)=(Jλf)(X)=Xfdλ, \int_Y {T\,f\,d\rho = \int_Y {J_\rho ^{ - 1}} (\Phi {J_\lambda }f)} d\rho = (\Phi ({J_\lambda }f))(Y) = ({J_\lambda }f)(X) = \int_X {f\,d\lambda } , so (iii) of Definition 2.2 is met. Finally, notice that BT(1A)dρ=BJρ1(ΦJλ(1A))dρ=Φ(Jλ(1A))(B)=Φ(1Aλ)(B) \int_B {T({1_A})d\rho } = \int_B {J_\rho ^{ - 1}} (\Phi {J_\lambda }({1_A}))d\rho = \Phi ({J_\lambda }(1A))(B) = \Phi ({1_A}\lambda )(B) for all A ∈ ∑X and B ∈ ∑Y , so the relation holds.

One consequence from Theorem 2.5 is that any Markov transfunction defined on μs {\cal M}_\mu ^s for s ∈ [1, ∞] uniquely extends or restricts to μs {\cal M}_\mu ^{s'} for all s′ ∈ [1, ∞], thus the value of s is insignificant. This is analogous to a similar property held by Markov operators, as in [9].

The remainder of this section aims to emphasize the importance of Theorem 2.5. For any p ∈ [1, ∞], a transfunction Φ:μpνp \Phi :{\cal M}_\mu ^p \to {\cal M}_\nu ^p , a Markov operator T : ℒp(X, µ) → ℒp(Y, ν ) and a plan κ ∈ Π(µ, ν) that satisfy the equalities Φ(1Aμ)(B)=BT(1A)dν=κ(A×B) \Phi ({1_A}\mu )(B) = \int_B {T({1_A})\,d\nu = \kappa (A\, \times \,B)} for all AX and BY contain the same information (transportation method), but convey it differently. By extending the equalities above for all f ∈ ℒp(X, µ) and g ∈ ℒq(Y, ν ) with 1/p + 1/q = 1, we have YgdΦ(fμ)=YT(f)d(gν)=X×Y(fg)dκ. \int_Y {g\,d\Phi (f\mu )} = \int_Y {T(f)\,d(g\nu ) = \int_{X \times Y} {(f \otimes g)\,d\kappa .} }

Note that if some positive measure µ′ also generates μp {\cal M}_\mu ^p , and if we define ν′ = Φ(µ′), then the same transfunction Φ:μpνp \Phi :{\cal M}_\mu ^p \to {\cal M}_\nu ^p corresponds to a Markov operator T′ : ℒp(X, µ′) → ℒp(Y, ν′) and it corresponds to a plan κ′ with marginals µ′ and ν′. Therefore T and T′ are different Markov operators, κ and κ′ are different plans, yet they follow the same “instructions” encoded by Φ. In this regard, Φ is a global way to describe a transportation method independent of marginals. If µ′ instead generates a smaller space than ℳµ, then Φ restricted to ℳµ contains part but not all of the instructions. Regardless, Φ will be Markov on this restriction. Notably, if µ′ = , then Φ: ℳµ → ℳν has associated Markov operator Th(f) := T (hf) and associated plan κ′ = (h ⊗ 1Y )κ.

Radon adjoints of transfunctions

Let (X, ∑X) be a Borel measurable space, let ℱX be a subset of bounded measurable real-valued functions on X and let ℳX be a subset of finite signed measures on X. Analogously, we have Y , ℱY and ℳY . For f ∈ ℱX and λ ∈ ℳX, define 〈f, λ〉 := ∫X f dλ. Similarly, for g ∈ ℱY and ρ ∈ ℳY , define 〈g, ρ 〉 := ∫Y g dρ. Occasionally, the elements within angular brackets shall be written in reverse order.

We say that {ℱX, ℳX} is a separating pair if 〈f1, λ〉 = 〈f2, λ〉 for all λ ∈ ℳX implies that f1 = f2, and if 〈f, λ1〉 = 〈f, λ2〉 for all f ∈ ℱX implies that λ1 = λ2. In this section, we shall develop some theory for two choices of the collections {ℱX, ℳX} and {ℱY , ℳY }, which we call the continuous setting and the measurable setting.

Definition 3.1

Let {ℱX, ℳX} and {ℱY , ℳY } each be a separating pair, let Φ: ℳX → ℳY be a transfunction, and let S : ℱY → ℱX be a function. Then Φ and S are Radon adjoints of each other if the equation YgdΦ(λ)=XS(g)dλ,i.e.g,Φ(λ)=S(g),λ \int_Y {g\,d\Phi (\lambda ) = \int_X {S(g)\,d\lambda ,\,{\rm{i}}.{\rm{e}}.} } \,\left\langle {g,\,\Phi \,(\lambda )} \right\rangle = \left. {S(g),\lambda } \right\rangle holds for all g ∈ ℱY and λ ∈ ℳX.

By utilizing the separation properties of 〈·, ·〉, Radon adjoints of both kinds are unique if they exist. We shall denote the Radon adjoint of Φ by Φ* and of S by S*.

If (Φ, S) is a Radon adjoint pair, then for all g ∈ ℱY , g,Φiλi=Sg,iλi=iSg,λi=ig,Φλi=g,iΦλi, \left\langle {g,\,\Phi \sum\nolimits_i {{\lambda _i}} } \right\rangle = \left\langle {Sg,\sum\nolimits_i {{\lambda _i}} } \right\rangle = \sum\nolimits_i {\left\langle {Sg,\,{\lambda _i}} \right\rangle } = \sum\nolimits_i {\left\langle {g,\,\Phi {\lambda _i}} \right\rangle } = \left\langle {g,\,\sum\nolimits_i {\Phi {\lambda _i}} } \right\rangle , meaning that Φ is linear. Similarly, for all λ ∈ ℳX, Sigi,λ=igi,Φλ=igi,Φλ=iSgi,λ=iSgi,λ, \left\langle {S\sum\nolimits_i {{g_i},\,\lambda } } \right\rangle = \,\left\langle {\sum\nolimits_i {{g_i},\,\Phi \lambda } } \right\rangle = \sum\nolimits_i {\left\langle {{g_i},\,\Phi \lambda } \right\rangle } = \sum\nolimits_i {\left\langle {S{g_i},\,\lambda } \right\rangle } = \left\langle {\sum\nolimits_i {S{g_i},\,\lambda } } \right\rangle , meaning that S is linear.

Example 3.2

If Φ = f# (the push-forward operator) for some measurable f : XY , then Φ*(g) = gf = f*g (the pull-back operator acting on g). This is because ∫Y g d(f#λ) = ∫X gf dλ for all g ∈ ℱY, λ ∈ℳX.

Example 3.3

If X = Y and Φλ:= f λ for some continuous (or measurable) f : X → ℝ, then Φ* (g) = gf. This is because ∫X g d() = ∫X g f dλ for all g ∈ ℱX, λ∈ ℳx.

Definition 3.4

Let {ℱX, ℳX} and {ℱY , ℳY } each be a separating pair.

(fn) weakly converges to f in ℱX, notated as fnwf f_n {\underrightarrow w}f , if every finite regular measure λ on X yields 〈fn, λ〉 → 〈f, λ〉 as n → ∞.

λnn=1 \left( {{\lambda _n}} \right)_{n = 1}^\infty weakly converges to λ in ℳX, notated as λnwλ \lambda _n \underrightarrow {\,w\,}\lambda , if every bounded continuous f : X → ℝ yields 〈f, λn〉 → 〈f, λ〉 as n → ∞.

An operator S : ℱY → ℱX is weakly continuous if gnwg g_n \underrightarrow {\,w\,}g in ℱY implies that SgnwSg Sg_n \underrightarrow {\,w}\,Sg in ℱX.

A transfunction Φ: ℳX → ℳY is weakly continuous if λnwλ \lambda _n \underrightarrow {\,w}\,\lambda in ℳX implies that ΦλnwΦλ \Phi \lambda _n \,\underrightarrow w\,\Phi \lambda in ℳY.

Note that weak convergence of (fn) in Definition 3.4 (i) is the same notion as bounded-pointwise convergence.

Approximations of identity
Definition 4.1

For a metric space (X, d) with xX, AX and δ > 0, define B(x; δ) := {zX : d(x, z) < δ} to be the δ-ball around x and define B(A; δ) := ∪xAB(x; δ) to be the δ-inflation around A.

The following two lemmas aid in showing Proposition 4.5.

Lemma 4.2

Let (X, d) be a locally compact metric space. The positive function c: X → (0, ∞] defined via c(x):=sup{δ>0:B(x:δ)isprecompact} c(x)\,: = \,\sup \{ \delta > 0\,:\,B(x:\delta )\,is\,precompact\} is either identicallyor it is finite and continuous on X. It follows that every compact set K has a precompact inflation B(K; δ) for some δ > 0.

Lemma 4.3

Let (X, d) be a locally compact Polish metric space. Then there exists a pair of sequences (xi)i=1 ({x_i})_{i = 1}^\infty from X and (βi)i=1 ({\beta _i})_{i = 1}^\infty from (0, 1] and there exists a function p: ℕ → ℕ such that for all n ∈ ℕ, Kn:=i=1p(n)B(xi,βi/n)¯ {K_n}\,: = \bigcup\limits_{i = 1}^{p(n)} {\overline {B({x_i},\,{\beta _i}/n)} } is compact with Kn+1Kn+1Kn {K_{n + 1}}\, \supseteq K_{n + 1}^\circ \, \supseteq {K_n} and n=1Kn=X \bigcup\nolimits_{n = 1}^\infty {{K_n}\, = \,X} .

Using the setup from Lemma 4.3, we define the collection of sets Cn,i:=B(xi,βi/n)¯j<1B(xj,βj/n)¯ {C_{n,i}}\,: = \,\overline {B({x_i},{\beta _i}/n)} - \bigcup\limits_{j < 1} {\overline {B({x_j},\,{\beta _j}/n)} } for all n, i ∈ ℕ. It follows for any n ∈ ℕ that i=1p(n)Cn,i=Kn \bigcup\nolimits_{i = 1}^{p(n)} {{C_{n,i}}} = {K_n} .

Definition 4.4

A measure µ is called a point-mass measure at x if µ(A) = 1 when xA and µ(A) = 0 when xA. A finite linear combination of point-mass measures is called a simple measure.

It is straightforward to show that simple measures are regular. The following proposition suggests a method to create approximations of identity, which shall be discussed in their respective sections below.

Proposition 4.5

Simple measures on a second-countable locally compact Hausdorff space form a dense subset of all finite regular measures with respect to weak convergence.

Proof

Construct sequences (xi)i=1 ({x_i})_{i = 1}^\infty , (βi)i=1 ({\beta _i})_{i = 1}^\infty , p: ℕ → ℕ and (Cn,i) via Lemma 4.3. Fix some positive finite measure λX+ \lambda \in {\cal M}_X^ + . Construct a sequence (λn)n=1 ({\lambda _n})_{n = 1}^\infty of positive simple measures via λn:=i=1p(n)λ(Cn,i)δxi {\lambda _n}: = \sum\nolimits_{i = 1}^{p(n)} {\lambda ({C_{n,i}})} {\delta _{{x_i}}} . We will show that λnwλ \lambda _n \,\underrightarrow w\,\lambda . In doing so, we fix some function fCb(X) and show that 〈f, λn〉 → 〈f, λ〉. For density of signed measures, one utilizes the Jordan decomposition and applies a similar argument for each component.

Let ε > 0. Define η := ε/(3||f|| + 3||λ|| + 1) so that ||f|| η < ε/3 and that ||λ||η < ε/3. Choose some natural M such that λ(KMc)<η \lambda (K_M^c) < \eta . Apply Lemma 4.2 to obtain some α > 0 with L:B(KM;α)¯ L:\overline {B({K_M};\alpha )} being compact. By uniform continuity of f|L, choose some natural N > M such that 2/N < α and for all xL, f(B(x; 2/N) ∩ L) ⊆ B(f(x); η).

Now let n > N. Define ρn,M:=i=1p(n)λ(Cn,iKM)δxi {\rho _{n,M}}: = \sum\nolimits_{i = 1}^{p(n)} {\lambda ({C_{n,i}} \cap {K_M}){\delta _{{x_i}}}} . Notice that Cn,iKM ≠ Ø implies that xiB(KM ; 1/n) and that Cn,iB(KM ; 2/n) ⊆ L, resulting in f(Cn,i) ⊆ B(f(xi); η). Three observations can be made:

f,λ1KMλfλ(KMc)<fη \left| {\left\langle {f,\lambda - {1_{{K_M}}}\lambda } \right\rangle } \right| \le \left\| f \right\| \cdot \lambda (K_M^c) < \left\| f \right\|\eta ;

f,1KMλρn,MKMfdλi=1p(n)f(xi)λ(Cn,iKM)<λη \left| {\left\langle {f,{1_{{K_M}}}\lambda - {\rho _{n,M}}} \right\rangle } \right| \le \left| {\int_{{K_M}} {f\,d\lambda - \sum\nolimits_{i = 1}^{p(n)} {f({x_i})\lambda ({C_{n,i}} \cap {K_M})} } } \right| < \left\| \lambda \right\|\eta ;

f,ρn,Mλnfi=1p(n)λ(Cn,iKMc)fλ(KMc)<fη \left| {\left\langle {f,{\rho _{n,M}} - {\lambda _n}} \right\rangle } \right| \le \left\| f \right\|\sum\nolimits_{i = 1}^{p(n)} {\lambda ({C_{n,i}} \cap K_M^c) \le \left\| f \right\|\lambda (K_M^c) < \left\| f \right\|\eta } .

Therefore, | 〈f, λλn〉| < 3(ε/3) = ε.

For any finite signed measure λ on X, the sequence (λn) of simple measures from Proposition 4.5 weakly converges to λ, hence the sequence of transfunctions (In) given by In:λλn=i=1p(n)1Cn,i,λδxi {I_n}:\lambda \mapsto {\lambda _n} = \sum\nolimits_{i = 1}^{p(n)} {\left\langle {{1_{{C_{n,i}}}},\lambda } \right\rangle {\delta _{{x_i}}}} is an approximation of identity.

The approximation of identity above is simply described with characteristic functions (1Cn,i) and point-mass measures (δxi). However, in each of the two settings below, either the characteristic functions must be replaced by bounded continuous functions or the point-mass measures must be replaced by compactly-supported measures that are absolutely continuous with respect to some underlying measure. With the correct choice of replacements, the same argument as given in Proposition 4.5 can be applied, yielding valid approximations of identities for the respective settings.

For the remainder of this paper, let X and Y be locally-compact Polish spaces, and pick any complete metric for each of them when needed.

Continuous setting: ℱ = Cb, ℳ = ℳfr

Let ℱX = Cb(X) denote the Banach space of all bounded continuous functions on X with the uniform norm and let ℳX = ℳfr(X) denote the Banach space of all finite (hence, regular) signed measures on X with the total variation norm. Develop Y , ℱY , and ℳY analogously. It is known that {ℱX, ℳX} is a separating pair in this setting.

An approximation of identity can be formed in this setting: keep the point-mass measures ρn,i := δxi, then for each natural n, replace the characteristic functions {1Cn,i : 1 ≤ ip(n)} used in Proposition 4.5 with positive compactly supported continuous functions {fn,i : 1 ≤ i ≤ p(n)} such that fn,i 1B(Cn,i;1/n) and that 1Kni=1p(n)fn,i1B(Kn;1/n) {1_{{K_n}}} \le \sum\nolimits_{i = 1}^{p(n)} {{f_{n,i}} \le {1_{B({K_n};1/n)}}} . Then an approximation of identity in the continuous setting is given by the sequence (In), where In:λi=1p(n)fn,i,λρn,i=i=1p(n)fn,i,λδxi. {I_n}:\lambda \mapsto \sum\limits_{i = 1}^{p(n)} {\left\langle {{f_{n,i}},\lambda } \right\rangle \,{\rho _{n,i}} = \sum\limits_{i = 1}^{p(n)} {\left\langle {{f_{n,i}},\lambda } \right\rangle \,{\delta _{{x_i}}}.} }

Theorem 5.1

Every strongly σ-additive and weakly-continuous transfunction Φ: ℳfr(X) → ℳfr(Y ) has a strongly σ-additive and weakly-continuous Radon adjoint S : Cb(Y ) → Cb(X). Conversely, every strongly σ-additive and weakly-continuous operator S : Cb(Y ) → Cb(X) has a strongly σ-additive and weakly-continuous Radon adjoint Φ: ℳfr(X) → ℳfr(Y ). When the Radon adjoint pair exists, their operator norms are equal (with respect to total-variation and uniform-convergence).

Proof

For the first claim, define S(g)(x) := 〈g, Φ(δx)〉 for all gCb(Y ) and for all xX so that 〈S(g), δx〉 = 〈g, Φ(δx)〉. Let xnx on X, so that δxnwδx \delta _{x_n } \,\underrightarrow w\,\delta _x , which means that Φ(δxn)wΦ(δx) \Phi (\delta _{x_n } )\,\underrightarrow w\,\Phi (\delta _x ) . Also let gng bounded-pointwise in Cb(Y ) (i.e. gnwg g_n \,\underrightarrow w\,g ). Then the statements below ensure that S(g) ∈ Cb(Y ), that S is bounded (hence uniform-continuous) and that S is bounded-pointwise-continuous (via the Dominated Convergence Theorem): S(g)(xn)=S(g),δxn=g,Φ(δxn)g,Φ(δx)=S(g),δx=S(g)(x);S(g)=supxXS(g)(x)=supxXS(g),δx=supxXg,Φ(δx)gΦ;S(gn)(x)=S(gn),δx=gn,Φδxg,Φδx=S(g),δx=S(g)(x). \eqalign{ & S(g)\,({x_n}) = \left\langle {S(g),{\delta _{{x_n}}}} \right\rangle = \left\langle {g,\Phi ({\delta _{{x_n}}})} \right\rangle \to \left\langle {g,\Phi ({\delta _x})} \right\rangle = \left\langle {S(g),{\delta _x}} \right\rangle = S(g)(x); \cr & \left\| {S(g)} \right\| = \mathop {\sup }\limits_{x \in X} \left| {S(g)\,(x)} \right| = \mathop {\sup }\limits_{x \in X} \left| {\left\langle {S(g),{\delta _x}} \right\rangle } \right| = \mathop {\sup }\limits_{x \in X} \left| {\left\langle {g,\Phi ({\delta _x})} \right\rangle } \right| \le \left\| g \right\| \cdot \left\| \Phi \right\|; \cr & S({g_n})\,(x) = \left\langle {S({g_n}),{\delta _x}} \right\rangle = \left\langle {{g_n},\Phi {\delta _x}} \right\rangle \to \left\langle {g,\Phi {\delta _x}} \right\rangle = \left\langle {S(g),{\delta _x}} \right\rangle = S(g)\,(x). \cr}

Since countable linear combinations of point-mass measures are weakly dense in ℳfr(X), the linearity and weak-continuity of the second coordinate in the 〈·, ·〉 structure and the weak-continuity of Φ yields that 〈S(g), λ〉 = 〈g, Φ λ〉 for all gCb(Y ) and λ ∈ ℳfr(X). Hence, S is the Radon adjoint of Φ with the desired properties.

For the second claim, note that for every λ ∈ ℳfr(X), the continuous functional g ↦ 〈S(g), λ〉 defined on C0(Y ) has Riesz representation 〈·,Φ (λ)〉 for some unique signed measure Φ(λ) ∈ ℳfr(Y ). Defining Φ in this manner for all λ, we obtain the equation 〈S(g), λ〉 = 〈g, Φλ〉 for all gC0(Y ) and λ ∈ ℳfr(X). C0(Y ) is dense in Cb(Y ) with respect to bounded-pointwise convergence, so with C0(Y)gnwgCb(Y) C_0 (Y) \mathrel\ni g_n \,\underrightarrow w\,g \in C_b (Y) , it follows that 〈gn, Φλ〉 → 〈g, Φλ〉 by the Dominated Convergence Theorem. Similarly, bounded-pointwise-continuity of S ensures that S(gn)wS(g) S(g_n )\,\underrightarrow w\,S(g) , which means that 〈S(gn), λ〉 → 〈S(g), λ 〉 by the Dominated Convergence Theorem. Therefore, 〈S(g), λ〉 = 〈g, Φλ〉 for all gCb(Y ) and λ ∈ ℳfr(X), implying that Φ is the Radon adjoint of S. To see that Φ is weakly-continuous, let λnwλ \lambda _n \,\underrightarrow w\,\lambda . Then 〈g, Φλn〉 = ⌣S(g), λn〉 → 〈S(g), λ〉 = 〈g, Φλ〉. Therefore, ΦλnwΦλ \Phi \lambda _n \,\underrightarrow w\,\Phi \lambda . Finally, ||Φ|| ≤ ||S||, hence ||Φ|| = ||S||, follows via Φλ=supg=1g,Φλ=supg=1S(g),λSλ. \left\| {\Phi \lambda } \right\| = \mathop {\sup }\limits_{\left\| g \right\| = 1} \,\left| {\left\langle {g,\Phi \lambda } \right\rangle } \right| = \mathop {\sup }\limits_{\left\| g \right\| = 1} \,\left| {\left\langle {S(g),\lambda } \right\rangle } \right| \le \left\| S \right\| \cdot \left\| \lambda \right\|.

Measurable setting: ℱ = ℒ, ℳ = ℳ

In this setting, let (X, ∑X, µ) be a finite measure space, let ℱX := ℒ(X, µ) and let X:=μ {{\cal M}_X}: = {\cal M}_\mu ^\infty . Define (Y, ∑Y , ν), ℱY , ℳY analogously. Then it is straightforward to verify that ℱX and ℳX separate each other.

An approximation of identity can be formed in this setting: for each natural n and 1 ≤ i ≤ p(n), replace each point-mass measure δxi used in Proposition 4.5 with the measure ρn,i := 1Cn,iµ and define fn,i := 1Cn,i(Cn,i) when µ(Cn,i) > 0; otherwise, define fn,i = 0. That is, an approximation of identity in the measurable setting is given by the sequence (In), where In:λi=1p(n)fn,i,λρn,i=i=1μ(Cn,i)>0p(n)1Cn,iμ(Cn,i),λ1Cn,iμ. I_n :\lambda \mapsto \sum\limits_{i = 1}^{p(n)} {\left\langle {f_{n,i} ,\lambda } \right\rangle } \,\rho _{n,i} = \sum\limits_{ i = 1 \atop \mu (C_{n,i} ) \gt 0} ^{p(n)} {\left\langle {{{1C_{n,i} } \over {\mu (C_{n,i} )}},\lambda } \right\rangle } \,1C_{n,i} \mu . The following lemma will be used in the proof of the next theorem:

Lemma 6.1

For every strongly σ-additive transfunction Φ:μν \Phi :{\cal M}_\mu ^\infty \to {\cal M}_\nu ^\infty , there is a unique strongly σ-additive transfunction Φ:νμ {\Phi ^\dagger }:{\cal M}_\nu ^\infty \to {\cal M}_\mu ^\infty such that Φ(1B ν)(A) = Φ(1Aµ)(B) for all measurable AX, BY , which implies thatf, Φ(g ν)〉 = 〈Φ(), gfor all f ∈ ℒ(X, µ), g ∈ ℒ(Y, ν). Also, Φ†† = Φ.

Proof

Let Φ be strongly σ-additive. For fixed BY , it follows by strong σ-additivity of Φ that the set function A ↦ Φ(1Aµ)(B) is a measure. Define this measure to be Ψ(1B ν). Then Ψ, defined on {1B ν| BY } is a strongly σ-additive transfunction that behaves like Φ in the equality above. Ψ can be linearly extended to ν {\cal M}_\nu ^\infty according to the following equalities for AX, g ≅ ∑j βj1Bj with ∑j|βj| < ∞: Ψ(gν)(A)=j=1βjΨ(1Bjν)(A)=j=1βjΦ(1Aμ)(Bj)=YgdΦ(1Aμ). \Psi (g\nu )\,(A) = \sum\limits_{j = 1}^\infty {{\beta _j}} \Psi (1{B_j}\nu )\,(A) = \sum\limits_{j = 1}^\infty {{\beta _j}\Phi ({1_A}\mu )} \,({B_j}) = \int_Y {g\,d\Phi \,({1_A}\mu ).}

The extended Ψ is strongly σ-additive on ν {\cal M}_\nu ^\infty . A similar calculation shows that XfdΨ(gν)=YgdΦ(fμ) \int_X {f\,d\Psi \,(g\nu )} = \int_Y {g\,d\Phi \,(f\,\mu )} for all fL(X, µ) and g ∈ ℒ(Y, ν). Therefore Φ is uniquely determined to be Ψ. Finally, Φ††(1Aµ)(B) = Φ(1B ν)(A) = Φ(1Aµ)(B) for all AX and BY , so Φ†† = Φ.

If Φ is Markov, then Φ is also Markov. Furthermore, the plans κ, κ corresponding to Φ, Φ respectively are dual to each other: that is, κ(A × B) = κ(B × A) for all measurable sets AX, BY . However, Φ is sensitive to the choice of measures µ and ν, which is not ideal when working with non-injective extensions of Markov transfunction Φ.

Theorem 6.2

Every strongly σ-additive Φ:μν \Phi :{\cal M}_\mu ^\infty \to {\cal M}_\nu ^\infty has a linear and bounded Radon adjoint S : ℒ(Y, ν) → ℒ(X, µ). Conversely, every linear and bounded operator S : ℒ(Y, ν) → ℒ(X, µ) has a strongly σ-additive Radon adjoint Φ:μν \Phi :{\cal M}_\mu ^\infty \to {\cal M}_\nu ^\infty .

Proof

Assume that Φ:μν \Phi :{\cal M}_\mu ^\infty \to {\cal M}_\nu ^\infty is strongly σ-additive and define S:Jμ1ΦJν S:\,J_\mu ^{ - 1}{\Phi ^\dagger }{J_\nu } with domain ℒ(Y, ν). Then S is linear and bounded. S = Φ* follows because for any f ∈ ℒ(X, µ) and g ∈ ℒ(Y, ν), g,Φ(fμ)=Φ(gν),f=(JμS)g,f=S(g),fμ. \left\langle {g,\Phi \,(f\mu )} \right\rangle = \left\langle {{\Phi ^\dagger }(g\nu ),\,f} \right\rangle = \left\langle {({J_\mu }S)g,f} \right\rangle = \left\langle {S(g),\,f\mu } \right\rangle .

On the other hand, assume that S : ℒ(Y, ν) → ℒ(X, µ) is linear and bounded. Then define Ψ:=JμSJν1 \Psi : = {J_\mu }SJ_\nu ^{ - 1} and Φ := Ψ. Then Φ is strongly σ-additive. Φ = S* follows because for any f ∈ ℒ(X, µ) and g ∈ ℒ(Y, ν), S(g),fμ=(Jμ1Ψ)(gν),fμ=Φ(gν),f=g,Φ(fμ). \left\langle {S(g),f\mu } \right\rangle = \left\langle {(J_\mu ^{ - 1}\Psi )\,(g\nu ),f\mu } \right\rangle = \left\langle {{\Phi ^\dagger }(g\nu ),f} \right\rangle = \left\langle {g,\Phi (f\mu )} \right\rangle .

Simple transfunctions

Let ℱX, ℱY , ℳX, and ℳY be defined in either the continuous setting or the measurable setting.

Definition 7.1

A transfunction Φ: ℳX → ℳY is simple if there exist functions (fi)i=1m ({f_i})_{i = 1}^m from ℱX and there exist measures (ρi)i=1m ({\rho _i})_{i = 1}^m from ℳY such that λX,Φλ=i=1mfi,λρi. \forall \lambda \in {{\cal M}_X},\Phi \lambda = \sum\limits_{i = 1}^m {\left\langle {{f_i},\lambda } \right\rangle } {\rho _i}.

It is straightforward to verify that simple transfunctions are strongly σ-additive. In the continuous setting, simple transfunctions are also weakly-continuous. Therefore by Theorem 5.1 in the continuous setting or Theorem 6.2 in the measurable setting, the Radon adjoint Φ* exists and satisfies gY,Φg=i=1mg,ρifi. \forall g \in {{\cal F}_Y},\,{\Phi ^ * }g = \sum\limits_{i = 1}^m {\left\langle {g,\,{\rho _i}} \right\rangle } {f_i}.

Note that the approximations of identity covered in both the continuous setting and the measurable setting involve sequences of simple transfunctions.

Theorem 7.2

In both the continuous setting and the measurable setting, linear weakly-continuous transfunctions can be approximated by simple transfunctions with respect to weak convergence; that is, simple transfunctions form a dense subset of linear weakly-continuous transfunctions with respect to weak convergence.

Proof

Let Φ: ℳX → ℳY be weakly-continuous transfunction and fix λ ∈ ℳX. Define Φn := Φ In, where In:λi=1p(n)fn,i,λρn,i {I_n}:\lambda \mapsto \sum\nolimits_{i = 1}^{p(n)} {\left\langle {{f_{n,i}},\lambda } \right\rangle } {\rho _{n,i}} forms the approximation of identity as defined in either Subsections 3.2 (continuous setting) or 3.3 (measurable setting). Then Φnλ=i=1p(n)fn,i,λΦρn,i {\Phi _n}\lambda = \sum\nolimits_{i = 1}^{p(n)} {\left\langle {{f_{n,i}},\lambda } \right\rangle \Phi {\rho _{n,i}}} , implying that Φn is a simple transfunction. It follows by Inλwλ I_n \lambda \,\underrightarrow w\,\lambda and by weak-continuity of Φ that Φnλ=Φ(Inλ)wΦλ \Phi _n \lambda = \Phi (I_n \lambda )\,\underrightarrow w\,\Phi \lambda .

Applications: optimal transport

Markov transfunctions provide a new perspective to optimal transport theory.

Definition 8.1

Let (X, ∑X, µ) and (Y, ∑Y , ν) be Polish measure spaces with finite positive measures µ and ν, respectively, with ||µ|| = ||ν||. A cost function is any continuous function c: X×Y → [0, ∞). A plan κ ∈ Π(µ, ν) is c-optimal if ∫X×Y c dκ ≤X×Y c dπ for all π∈ Π(µ, ν). A Markov transfunction Φ: ℳX → ℳY is c-optimal on µ if the corresponding plan κ with marginals µ and Φµ is c-optimal, and Φ is simply c-optimal if it is c-optimal on ℳX.

The next proposition implies that optimal inputs for Φ form a large class of measures.

Proposition 8.2

Let (X, ∑X), (Y, ∑Y ) be Polish spaces, let c be a cost function, and let Φ: ℳX → ℳY be a Markov transfunction. If Φ is c-optimal on µ ∈ ℳX, then Φ is c-optimal on μ {\cal M}_\mu ^\infty .

Proof

The proof follows easily from Theorem 4.6 in [10] on the inheritance of optimality of plans by restriction.

In the next theorem, we provide a “warehouse strategy” which approximates the optimal cost between fixed marginals with respect to some cost function. First, we subdivide the input marginal by local regions, and send the subdivided measures to point mass measures – warehouses – within their respective regions. Second, we transfer mass between warehouses via the discrete transport problem. Finally, the warehouses locally redistribute to form the output marginal. The overall cost of transport via the warehouse strategy approaches the optimal cost as the size of the regions decreases.

Theorem 8.3

Let (X, ∑X) be a locally compact Polish measurable space with complete metric d, let λ and ρ be finite positive compactly-supported measures with ||λ|| = ||ρ||, and let c: X × X → [0, ∞) be a cost function with c(x, y) ≤ αd(x, y)p for some constants α, p > 0. The optimal cost between marginals λ, ρ with respect to c can be sufficiently approximated by the costs of simple Markov transfunctions.

Proof

Consider the approximation of identity (In) from the continuous setting in Section 5. For large n, we create a composition of three simple Markov transfunctions: λ first maps to Inλ=i=1p(n)fn,i,λδxi {I_n}\lambda = \sum\nolimits_{i = 1}^{p(n)} {\left\langle {{f_{n,i}},\lambda } \right\rangle } {\delta _{{x_i}}} , which maps to Inρ=i=1p(n)fn,i,ρδxi {I_n}\rho = \sum\nolimits_{i = 1}^{p(n)} {\left\langle {{f_{n,i}},\rho } \right\rangle } {\delta _{{x_i}}} , which finally maps to ρ. These steps are measure-preserving because Kn (from Lemma 4.3) contains the supports of λ and ρ for large n. The most crucial goal is to determine the optimal simple Markov transfunction for the middle step.

The first and last steps cost no more than αn−p||λ|| each, which reduces to 0 as n → ∞. This means that the optimal cost between marginals λn and ρn approaches the optimal cost between marginals λ and ρ as n → ∞. Solving the former optimal cost is the well-known discrete version of the Monge-Kantorovich transport problem.

By approximating each of the values 〈fn,i, λ 〉 ≈ an,i/z and 〈fn,i〉 ≈ bn,i/z for natural numbers an,i, bn,i, z with 1 ≤ i ≤ p(n), the middle step can approximately be interpreted as the Assignment Problem on a weighted bipartite graph between vertex sets P and Q, where P denotes a set created by forming an,i copies of a vertex corresponding to each δxi in λn, Q denotes the set created by forming bn,j copies of a vertex corresponding to each δxj in ρn, and drawing edges between these vertices with weight c(xi, xj). This problem has been studied, and can be solved in polynomial time of P=i=1p(n)an,iλz \left| P \right| = \sum\nolimits_{i = 1}^{p(n)} {{a_{n,i}} \approx } \left\| \lambda \right\|z ; the Hungarian method is one well-known algorithm [7].

Although Theorem 8.3 provides a sequence of simple transfunctions that approximate the optimal cost between fixed marginals, the sequence is not expected to converge weakly to an optimal Markov transfunction, as the solutions to the middle step could vary greatly as n increases. Consequently, we can find a Markov transfunction whose cost between marginals is sufficiently close to the optimal cost, but Theorem 8.3 does not provide an optimal Markov transfunction.

However, for any Markov transfunction between fixed marginals, the next theorem yields an approximation by simple Markov transfunctions with respect to weak convergence. Consequently, the cost between the marginals of the constructed sequence of simple Markov transfunctions approaches the cost for the original transfunction.

Theorem 8.4

Let (X, ∑X, µ) and (Y, ∑Y , ν) be locally compact Polish measure spaces with finite compactly-supported positive measures µ and ν such that ||µ|| = ||ν||. Any Markov transfunction Φ: ℳµ → ℳν can be approximated by simple Markov transfunctions with respect to weak convergence.

Proof

Consider the approximation of identity (In) with respect to µ from the measurable setting from Section 6. Let n be large so that Kn (from Lemma 4.3) contains the supports of µ and ν.

Let κ be the plan corresponding to Markov transfunction Φ from Theorem 2.5. For 1 ≤ i, j ≤ p(n), the quantity κ(Cn,i × Cn,j) represents how much mass transfers from 1Cn,iµ to 1Cn,jν. If µ(Cn,i)ν(Cn,j) > 0, then we can approximate nonzero measure (1Cn,i ⊗ 1Cn,j ) κ with κn,i,j:=κ(Cn,i×Cn,j)1Cn,iμμ(Cn,i)×1Cn,jνν(Cn,j). {\kappa _{n,i,j}}: = \kappa ({C_{n,i}}\, \times \,{C_{n,j}}){{1{C_{n,i}}\mu } \over {\mu ({C_{n,i}})}} \times {{1{C_{n,j}}\nu } \over {\nu ({C_{n,j}})}}. Otherwise, we define κn,i,j := 0. Then κn := ∑ij κn,i,j is a plan from Π(µ, ν) which corresponds to a Markov transfunction Φn from Theorem 2.5.

Next, we show that κnwκ \kappa _n \,\underrightarrow w\,\kappa as n → ∞. Let cCb(X × Y ) with ||c|| 1, and for 1 ≤ i, j ≤ p(n), let βn,i,j := sup c(Cn,i × Cn,j) inf c(Cn,i × Cn,j). By uniform continuity of c on Kn × Kn, we have that βn := max{ βn,i,j |1 ≤ i, j ≤ p(n)} → 0 as n → ∞, which implies that c,κκnijβn,i,jκ(Cn,i×Cn,j)βnκ0. \left| {\left\langle {c,\kappa - {\kappa _n}} \right\rangle } \right| \le \sum\limits_i {\sum\limits_j {{\beta _{n,i,j}}\,\kappa ({C_{n,i}}\, \times {C_{n,j}})} \le {\beta _n}\left\| \kappa \right\| \to 0.}

There are some properties of Φn worth noting: Φn maps μ {\cal M}_\mu ^\infty to span{1Cn,j ν}; Φn behaves as a matrix when applied to span{1Cn,iµ}; the structure of κn guarantees that Φn = ΦnIn. If we choose bases (1Cn,iµ) and (1Cn,j ν), the matrix Mn representing Φn has entries Mn(j, i) := κ(Cn,i × Cn,j)(Cn,j).

Let λμ \lambda \in {\cal M}_\mu ^\infty and for 1 ≤ i ≤ p(n). Then Φnλ=ΦnInλ=jiκ(Cn,i×Cn,j)ν(Cn,j)1Cn,iμ(Cn,i),λ1Cn,jν, {\Phi _n}\lambda = {\Phi _n}{I_n}\lambda = \sum\limits_j {\left\langle {\sum\limits_i {{{\kappa ({C_{n,i}}\, \times \,{C_{n,j}})} \over {\nu ({C_{n,j}})}}{{{1_C}_{_{n,i}}} \over {\mu ({C_{n,i}})}},} \lambda } \right\rangle {1_C}_{_{n,j}}\nu ,} showing that Φn is simple.

We now show that ΦnλwΦλ \Phi _n \lambda \,\underrightarrow w\,\Phi \lambda as n → ∞. Let gCb(Y ) with ||g|| ≤ 1.

Let ε > 0. Since λ = for some f ∈ ℒ(X, µ), choose some Cb(X) such that ||(f − f̃)µ|| < ε/3. Since Φ=Φn=1 \left\| {{\Phi ^ * }} \right\| = \left\| {\Phi _n^ * } \right\| = 1 , we have that g,Φ(ff˜)μ=Φg,(ff˜)μΦg(ff˜)με/3, \left| {\left\langle {g,\Phi (f - \tilde f)\mu } \right\rangle } \right| = \left| {\left\langle {{\Phi ^ * }g,(f - \tilde f)\mu } \right\rangle } \right| \le \left\| {{\Phi ^ * }g} \right\| \cdot \left\| {(f - \tilde f)\mu } \right\| \le \varepsilon /3, and that g,Φn(ff˜)μ=Φng,(ff˜)μΦng(ff˜)με/3. \left| {\left\langle {g,{\Phi _n}(f - \tilde f)\mu } \right\rangle } \right| = \left| {\left\langle {\Phi _n^ * g,\,(f - \tilde f)\mu } \right\rangle } \right| \le \left\| {\Phi _n^ * g} \right\| \cdot \left\| {(f - \tilde f)\mu } \right\| \le \varepsilon /3. Since κnwκ \kappa _n \,\underrightarrow w\,\kappa as n → ∞ and gCb(X × Y ), there is some natural N so that for all n ≥ N, g,(ΦΦn)f˜μ=f˜g,κκn<ε/3. \left| {\left\langle {g,(\Phi - {\Phi _n})\tilde f\mu } \right\rangle } \right| = \left| {\left\langle {\tilde f \otimes g,\kappa - {\kappa _n}} \right\rangle } \right| < \varepsilon /3. It follows from above that |〈g, (Φ Φn) λ〉| ≤ ε for n ≥ N by the triangle inequality.

Theorem 8.4 can be strengthened by removing the assumptions that µ and ν are compactly supported; the approximation of identity may not capture all of µ nor λ, and κn ∈ Π(1Knµ, 1Kn ν) may not belong to Π(µ, ν), but the rest of the analysis holds. Notably, to show κnwκ \kappa _n \,\underrightarrow w\,\kappa as n → ∞, the inequalities become c,κκnκ(Knc)+ijβn,i,jκ(Cn,i×Cn,j)κ(Knc)+βnκn0. \eqalign{ & \left| {\left\langle {c,\kappa - {\kappa _n}} \right\rangle } \right| & \le \kappa (K_n^c) + \sum\limits_i {\sum\limits_j {{\beta _{n,i,j}}\,\kappa ({C_{n,i}}\, \times \,{C_{n,j}})} } \cr & & \le \kappa (K_n^c) + {\beta _n}\left\| {{\kappa _n}} \right\| \to 0. \cr}

Idioma:
Inglés
Calendario de la edición:
2 veces al año
Temas de la revista:
Matemáticas, Matemáticas generales