Hostname: page-component-7bb8b95d7b-cx56b Total loading time: 0 Render date: 2024-09-28T21:22:27.319Z Has data issue: false hasContentIssue false

Dilations of Markovian semigroups of measurable Schur multipliers

Published online by Cambridge University Press:  05 April 2023

Cédric Arhancet*
Affiliation:
Lycée Lapérouse, 13 Rue Didier Daurat, 81000 Albi, France
Rights & Permissions [Opens in a new window]

Abstract

Using probabilistic tools, we prove that any weak* continuous semigroup $(T_t)_{t \geqslant 0}$ of self-adjoint unital completely positive measurable Schur multipliers acting on the space $\mathrm {B}({\mathrm {L}}^2(X))$ of bounded operators on the Hilbert space ${\mathrm {L}}^2(X)$, where X is a suitable measure space, can be dilated by a weak* continuous group of Markov $*$-automorphisms on a bigger von Neumann algebra. We also construct a Markov dilation of these semigroups. Our results imply the boundedness of the McIntosh’s ${\mathrm {H}}^\infty $ functional calculus of the generators of these semigroups on the associated Schatten spaces and some interpolation results connected to ${\mathrm {BMO}}$-spaces. We also give an answer to a question of Steen, Todorov, and Turowska on completely positive continuous Schur multipliers.

Type
Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

The study of dilations of operators is of central importance in operator theory and has a long tradition in functional analysis. Indeed, dilations are powerful tools which allow to reduce general studies of operators to more tractable ones.

Suppose $1<p<\infty $ . In the spirit of Sz.-Nagy’s dilation theorem for contractions on Hilbert spaces, Fendler [Reference FendlerFen97] proved a dilation result for any strongly continuous semigroup $(T_t)_{t \geqslant 0}$ of positive contractions on an ${\mathrm {L}}^p$ -space ${\mathrm {L}}^p(\Omega )$ . More precisely, this theorem says that there exist a measure space $\Omega '$ , two positive contractions $J \colon {\mathrm {L}}^p(\Omega ) \to {\mathrm {L}}^p(\Omega ')$ and $P \colon {\mathrm {L}}^p(\Omega ') \to {\mathrm {L}}^p(\Omega )$ , and a strongly continuous group of positive invertible isometries $(U_t)_{t \in \mathbb {R}}$ on ${\mathrm {L}}^p(\Omega ')$ such that

(1.1) $$ \begin{align} T_t = PU_tJ, \end{align} $$

for any $t \geqslant 0$ (see also [Reference FendlerFen98]). Note that, in this situation, the map $J \colon {\mathrm {L}}^p(\Omega ) \to {\mathrm {L}}^p(\Omega ')$ is an isometric embedding, whereas the map $JP \colon {\mathrm {L}}^p(\Omega ') \to {\mathrm {L}}^p(\Omega ')$ is a contractive projection.

In the noncommutative setting, measure spaces and ${\mathrm {L}}^p$ -spaces are replaced by von Neumann algebras and noncommutative ${\mathrm {L}}^p$ -spaces and positive maps by completely positive maps. In their remarkable paper [Reference Junge and Le MerdyJLM07], Junge and Le Merdy essentiallyFootnote 1 showed that there is no hope to have a “reasonable” analog of Fendler’s theorem for semigroups of completely positive contractions acting on noncommutative ${\mathrm {L}}^p$ -spaces. It is a striking difference with the world of classical (=commutative) ${\mathrm {L}}^p$ -spaces of measure spaces.

Let X be a $\sigma $ -finite measure space. In this paper, our first main result (Theorem 5.2) gives a dilation on weak* continuous semigroups of self-adjoint unital completely positive measurable Schur multipliers acting on $\mathrm {B}({\mathrm {L}}^2(X))$ in the spirit of (1.1) but at the level $p=\infty $ . Our construction heavily relies on the use of Gaussian processes. Our dilation induces an isometric dilation similar to the one of Fendler’s theorem for the strongly continuous semigroup induced by the semigroup $(T_t)_{t \geqslant 0}$ on the Schatten space $S^p_X \overset {\mathrm {def}}{=} S^p({\mathrm {L}}^2(X))$ for any $1 \leqslant p <\infty $ . Finally, note that, in this paper, the weak* continuity of the semigroup means that the map $\mathbb {R}^+ \to \mathbb {C}$ , $t \mapsto \langle T_t(x),y \rangle _{\mathrm {B}({\mathrm {L}}^2(X)),S^1_X}$ is continuous for any $x \in \mathrm {B}({\mathrm {L}}^2(X))$ and any $y \in S^1_X$ .

Moreover, this result is analogous to the result of [Reference ArhancetArh20], which provides a dilation of weak* continuous semigroups of self-adjoint unital completely positive Fourier multipliers acting on the group von Neumann algebra $\mathrm {VN}(G)$ of a locally compact group G. Note also that in the case where X is a finite set equipped with the counting measure, it is possible to use the considerably more complicated approach of [Reference ArhancetArh19] (relying on the use of ultrafilters) whose assumptions are satisfied by the results of [Reference ArhancetArh13a] and [Reference SpronkRic08] for obtaining a dilation. But even in this case, our new and more general approach gives an explicit and a more useful dilation since the von Neumann algebra where the group $(U_t)_{t \in \mathbb {R}}$ acts is injective. This point is important for applications in vector-valued noncommutative ${\mathrm {L}}^p$ -spaces since we need injective von Neumann algebras in this context (see [Reference SakaiPis98]). We refer to the papers [Reference Arhancet and KrieglerAFM17, Reference Arhancet, Fackler and Le MerdyALM14, Reference ArhancetArh13b, Reference HaaseHaM11] for related things. Finally, note that Schur multipliers are crucial operators in noncommutative analysis and are connected to a considerable number of topics as harmonic analysis, double operator integrals, perturbation theory, and Grothendieck’s theorem. More recently, the semigroups of Schur multipliers considered in this paper were connected to noncommutative geometry in the memoir [Reference Arhancet and KrieglerArK22a].

One of the important consequences of Fendler’s theorem is the boundedness, for the generator of a strongly continuous semigroup $(T_t)_{t \geqslant 0}$ of positive contractions, of a bounded ${\mathrm {H}}^\infty $ functional calculus which is a fundamental tool in various areas: harmonic analysis of semigroups, multiplier theory, Kato’s square root problem, maximal regularity in parabolic equations, and control theory. For detailed information, we refer the reader to [Reference Haagerup and MusatHaa06, Reference Junge and MeiJMX06, Reference Kunstmann and WeisKuW04], to the survey [Reference Le MerdyLeM07], and to the recent book [Reference Hytönen, van Neerven, Veraar and WeisHvNVW18] and the references therein. Our results also give a similar result on ${\mathrm {H}}^\infty $ functional calculus in the noncommutative context in the case of semigroups of measurable Schur multipliers.

Recall that a locally integrable function $f \colon \mathbb {R}^n \to \mathbb {C}$ is said to have bounded mean oscillation if

$$ \begin{align*}\| f \|_{{\mathrm{BMO}}} \overset{\mathrm{def}}{=} \sup_{Q \in \mathcal{Q}} \frac{1}{|Q|} \int_Q \left| f(y) - f_Q \right|{}^2 \mathop{}\mathopen{}\mathrm{d} y < \infty, \end{align*} $$

where $f_Q \overset {\mathrm {def}}{=} \frac {1}{|Q|} \int _Q f$ is the average of f over Q, $|Q|$ is the volume of Q, and $\mathcal {Q}$ is the set of all cubes in $\mathbb {R}^n$ . The quotient space of bounded mean oscillation functions modulo the space of constant functions with the previous seminorm is a Banach space. The importance of the ${\mathrm {BMO}}$ -norm and the ${\mathrm {BMO}}$ -space lies in the fact that they arise as end-point spaces for estimating the bounds of linear maps on function spaces on $\mathbb {R}^n$ . This includes many Calderón–Zygmund operators. By interpolation, ${\mathrm {BMO}}$ -spaces are an effective tool to obtain ${\mathrm {L}}^p$ -bounds of Fourier multipliers. In this sense, the ${\mathrm {BMO}}$ -space is a natural substitute of the space ${\mathrm {L}}^\infty (\mathbb {R}^n)$ .

Spaces of functions with bounded means oscillation can also be studied through semigroups. If $\mathcal {T}=(T_t)_{t \geqslant 0}$ is a Markov semigroup on an ${\mathrm {L}}^\infty $ -space ${\mathrm {L}}^\infty (\Omega )$ of a (finite) measure space $\Omega $ , then we can define a ${\mathrm {BMO}}$ -seminorm by

$$ \begin{align*}\| f \|_{{\mathrm{BMO}}_{\mathcal{T}}} \overset{\mathrm{def}}{=} \sup_{t \geqslant 0} \big\| T_t\left(|f - T_t(f)|^2 \right) \big\|_\infty^{\frac{1}{2}}. \end{align*} $$

More generally, Junge and Mei introduced in [Reference Junge and MeiJM12] several noncommutative semigroup ${\mathrm {BMO}}$ -spaces starting from a Markov semigroup on a semifinite von Neumann algebra (i.e., a noncommutative ${\mathrm {L}}^\infty $ -space). Relations between these spaces are studied, and interpolation results are obtained. A crucial ingredient of their approach is Markov dilations of semigroups that allow one to use martingale theory and derive results from this probabilistic setting. See also [Reference Ferguson, Mei and SimanekFMS19, Reference Jiao and WangJiW17, Reference Junge, Le Merdy and XuJM10, Reference Junge and ZengJuZ15] for other applications of Markov dilations.

The second main result of this paper is Theorem 6.1, which gives the construction of a Markov dilation of each weak* continuous semigroup of self-adjoint unital completely positive measurable Schur multipliers acting on $\mathrm {B}({\mathrm {L}}^2(X))$ . Again, our construction relies on probabilistic tools. From this result, we obtain interpolation results with ${\mathrm {BMO}}$ -spaces as end-point for these semigroups (see Theorem 7.2 and Remark 7.3). Probably, our results will be useful for estimating norms of Schur multipliers on Schatten spaces. Finally, we equally refer to [Reference CaspersCas19, Reference Caspers, Junge, Sukochev and ZaninCJSZ20] for strongly related papers.

1.1 Structure of the paper

The paper is organized as follows. Section 2 gives background. In Section 3, we give useful observations on measurable positive-definite kernels and negative-definite kernels. In Section 4, we give complements on measurable Schur multipliers. We answer a question of Steen, Todorov, and Turowska in Remark 4.9. We also give a precise description of weak* continuous semigroups of unital self-adjoint completely positive measurable Schur multipliers on some suitable measure spaces. Section 5 gives a proof of our main result of dilation of these semigroups. Next, in Section 6, we also construct a Markov dilation for these semigroups. In Section 7, we describe some applications of our results to functional calculus and interpolation.

2 Preliminaries

In this paper, we suppose that the considered measure spaces are complete (i.e., every subset of every null set is measurable).

2.1 Isonormal processes

Let H be a real Hilbert space. An H-isonormal process on a probability space $(\Omega ,\mu )$ ([Reference PisierNua06, Definition 1.1.1], [Reference NualartNeer07, Definition 6.5]) is a linear mapping $\mathrm {W} \colon H \to {\mathrm {L}}^0(\Omega )$ from H into the space ${\mathrm {L}}^0(\Omega )$ of measurable functions on $\Omega $ with the following properties:

(2.1) $$ \begin{align} & \text{For any }h \in H\text{, the random variable }\mathrm{W}(h)\text{ is a centered real Gaussian.} \end{align} $$
(2.2) $$ \begin{align} & \text{For any } h_1, h_2 \in H, \text{ we have } \mathbb{E}\big(\mathrm{W}(h_1) \mathrm{W}(h_2)\big)= \langle h_1, h_2\rangle_H. \end{align} $$
(2.3) $$ \begin{align} & \text{The linear span of the products } \mathrm{W}(h_1)\mathrm{W}(h_2)\cdots \mathrm{W}(h_m), \text{ with } m \geqslant 0 \text{ and } h_1,\ldots, h_m \\ &\nonumber\text{in }H, \text{ is dense in the real Hilbert space }{\mathrm{L}}^2_{\mathbb{R}}(\Omega).\end{align} $$

Here, we make the convention that the empty product, corresponding to $m=0$ in (2.3), is the constant function $1$ . Moreover, $\mathbb {E}$ is used to denote the expected value.

If $(e_i)_{i \in I}$ is an orthonormal basis of H and if $(\gamma _i)_{i \in I}$ is a family of independent standard Gaussian random variables on a probability space $\Omega $ , then for any $h \in H$ , the family $(\gamma _i \langle h,e_i\rangle _H)_{i \in I}$ is summable in ${\mathrm {L}}^2(\Omega )$ and

(2.4) $$ \begin{align} \mathrm{W}(h) \overset{\mathrm{def}}{=} \sum_{i \in I} \gamma_i \langle h,e_i\rangle_H, \quad h \in H, \end{align} $$

defines an H-isonormal process.

Recall that the span of elements $\mathrm {e}^{{\mathrm {i}}\mathrm {W}(h)}$ is weak* dense in ${\mathrm {L}}^\infty (\Omega )$ by [Reference JansonJan97, Remark 2.15]. Using [Reference Hytönen, van Neerven, Veraar and WeisHvNVW18, Proposition E.2.2] with t instead of $\xi $ and by observing by (2.2) that the variance $\mathbb {E}(\mathrm {W}(h)^2)$ of the Gaussian variable $\mathrm {W}(h)$ is equal to $ \| h \|_H^2$ , we see that

(2.5) $$ \begin{align} \mathbb{E}\big(\mathrm{e}^{{\mathrm{i}} t\mathrm{W}(h)}\big) =\mathrm{e}^{-\frac{t^2}{2} \| h \|_H^2}, \quad t \in \mathbb{R}, h \in H. \end{align} $$

If $u \colon H \to H$ is a contraction, we denote by $\Gamma _\infty (u) \colon {\mathrm {L}}^\infty (\Omega ) \to {\mathrm {L}}^\infty (\Omega )$ the (symmetric) second quantization of u acting on the complex Banach space ${\mathrm {L}}^\infty (\Omega )$ . Recall that the map $\Gamma _\infty (u) \colon {\mathrm {L}}^\infty (\Omega ) \to {\mathrm {L}}^\infty (\Omega )$ preserves the integral.Footnote 2 If u is a surjective isometry, we have

(2.6) $$ \begin{align} \Gamma_\infty(u) \big(\mathrm{e}^{{\mathrm{i}}\mathrm{W}(h)}\big) =\mathrm{e}^{{\mathrm{i}}\mathrm{W}(u(h))}, \quad h \in H, \end{align} $$

and $\Gamma _\infty (u) \colon {\mathrm {L}}^\infty (\Omega ) \to {\mathrm {L}}^\infty (\Omega )$ is a $*$ -automorphism of the von Neumann algebra ${\mathrm {L}}^\infty (\Omega )$ .

Furthermore, the second quantization functor $\Gamma $ satisfies the following elementary result [Reference ArhancetArh20, Lemma 2.1]. In the part 1, we suppose that the constructionFootnote 3 is given by the concrete representation (2.4).

Lemma 2.1

  1. (1) If ${\mathrm {L}}^\infty (\Omega )$ is equipped with the weak* topology, then the map $H \to {\mathrm {L}}^\infty (\Omega )$ , $h \mapsto \mathrm {e}^{{\mathrm {i}}\mathrm {W}(h)}$ is continuous.

  2. (2) If $(U_t)_{t \in \mathbb {R}}$ is a strongly continuous group of orthogonal operators acting on the Hilbert space H, then $(\Gamma _\infty (U_t))_{t \in \mathbb {R}}$ is a weak* continuous group of operators acting on the Banach space ${\mathrm {L}}^\infty (\Omega )$ .

Let H be a real Hilbert space. Following [Reference PalmerNVW15, Definition 2.2] and [Reference NualartNeer07, Definition 6.11], we say that an ${\mathrm {L}}^2_{\mathbb {R}}(\mathbb {R}^+,H)$ -isonormal process $\mathrm {W}$ is an H-cylindrical Brownian motion. In this case, for any $t \geqslant 0$ and any $h \in H$ , we let

(2.7) $$ \begin{align} \mathrm{W}_t(h)\overset{\mathrm{def}}{=}\mathrm{W}\big(1_{[0,t]} \otimes h\big). \end{align} $$

We introduce the filtration $({\mathscr {F}}_t)_{t \geqslant 0}$ defined by

(2.8) $$ \begin{align} \mathscr{F}_t\overset{\mathrm{def}}{=} \sigma\big(\mathrm{W}_r(h) : r \in [0,t], h \in H\big), \end{align} $$

that is, the $\sigma $ -algebra generated by the random variables $\mathrm {W}_r(h)$ for $r \in [0,t]$ and $h \in H$ .

By [Reference NualartNeer07, p. 77], for any fixed $h \in H$ , the family $(\mathrm {W}_t(h))_{t \geqslant 0}$ is a Brownian motion. This means that [Reference NualartNeer07, Definition 6.2]

(2.9) $$ \begin{align} & \mathrm{W}_0(h) = 0\text{ almost surely,} \end{align} $$
(2.10) $$ \begin{align} & \mathrm{W}_t(h)-\mathrm{W}_s(h)\text{ is Gaussian with variance }(t-s) \| h \|_H^2\text{ for any }0 \leqslant s \leqslant t, \end{align} $$
(2.11) $$ \begin{align} & \mathrm{W}_t(h)-\mathrm{W}_s(h)\text{ is independent of }\{\mathrm{W}_r(h) : r \in [0 ,s] \}\text{ for any }0 \leqslant s \leqslant t. \end{align} $$

Indeed, by [Reference NualartNeer07, p. 163],

(2.12) $$ \begin{align} &\text{the increment }\mathrm{W}_t(h)-\mathrm{W}_s(h)\text{ is independent of the }\sigma\text{-algebra }\mathscr{F}_s. \end{align} $$

Moreover, by [Reference NualartNeer07, p. 163], the family $(\mathrm {W}_t(h))_{t \geqslant 0}$ is a martingale with respect to $(\mathscr {F}_t)_{t \geqslant 0}$ . In particular, the random variable $\mathrm {W}_t(h)$ is $\mathscr {F}_t$ -measurable. If $0 \leqslant s \leqslant t$ , note that

$$ \begin{align*}\| 1_{]s,t]} \otimes h \|_{{\mathrm{L}}^2_{\mathbb{R}}(\mathbb{R}^+,H)}^2 = \big\| 1_{]s,t]} \big\|_{{\mathrm{L}}^2_{\mathbb{R}}(\mathbb{R}^+)}^2 \| h \|_H^2 =(t-s) \| h \|_H^2. \end{align*} $$

Using (2.5) together with the previous computation, we obtain

(2.13) $$ \begin{align} \mathrm{E}\big(\mathrm{e}^{{\mathrm{i}} \mathrm{W}(1_{]s,t]} \otimes h)}\big) =\mathrm{e}^{-\frac{t-s}{2} \| h \|_H^2}, \quad 0 \leqslant s \leqslant t,\ h \in H. \end{align} $$

2.2 Probabilities

Let $\Omega $ be a probability space, and let X be a Banach space. If $f \in {\mathrm {L}}^1(\Omega ,X)$ is independent of the sub- $\sigma $ -algebra $\mathscr {F}$ , then, by [Reference Hytönen, van Neerven, Veraar and WeisHvNVW16, Proposition 2.6.35], its conditional expectation $\mathbb {E}(f|\mathscr {F})$ with respect to $\mathscr {F}$ is given by the constant function:

(2.14) $$ \begin{align} \mathbb{E}(f|\mathscr{F}) = \mathbb{E}(f). \end{align} $$

2.3 Hilbert–Schmidt operators

Let X be a $\sigma $ -finite measure space. We will use the space $S^\infty _X \overset {\mathrm {def}}{=} S^\infty ({\mathrm {L}}^2(X))$ of compact operators, its dual $S^1_X$ , and the space $\mathrm {B}({\mathrm {L}}^2(X))$ of bounded operators on the Hilbert space ${\mathrm {L}}^2(X)$ . If $f \in {\mathrm {L}}^2(X \times X)$ , we denote the associated Hilbert–Schmidt operator by

(2.15) $$ \begin{align} \begin{array}{cccc} K_f \colon & {\mathrm{L}}^2(X) & \longrightarrow & {\mathrm{L}}^2(X) \\ & \xi & \longmapsto & \int_{X} f(\cdot,y)\xi(y) \mathop{}\mathopen{}\mathrm{d} y \\ \end{array}. \end{align} $$

We will also use the notation $K((x,y) \mapsto f(x,y))$ for $K_f$ . With the notation $\check {f}(x,y) \overset {\mathrm {def}}{=} f(y,x)$ , we have

(2.16) $$ \begin{align} (K_f)^* =K_{\check{\overline{f}}}. \end{align} $$

The operator $K_f$ is self-adjoint if and only if $f(x,y)=\overline {f(y,x)}$ almost everywhere.

For any $f,g \in {\mathrm {L}}^2(X \times X)$ , we have $K_fK_g=K_h$ where $h \in {\mathrm {L}}^2(X \times X)$ is defined by

(2.17) $$ \begin{align} h(x,y) =\int_X f(x,z)g(z,y) \mathop{}\mathopen{}\mathrm{d} z \end{align} $$

with $ \| h \|_{2} \leqslant \| f \|_{2} \| g \|_{2}$ .

2.4 Kernels

Let X be a set. A function $\varphi \colon X \times X \to \mathbb {C}$ is called Hermitian if $\varphi (y,x) = \overline {\varphi (x,y)}$ for any $x,y \in X$ . A function $\varphi \colon X \times X \to \mathbb {C}$ is a positive-definite kernel if for any integer $n \geqslant 1$ , any $c_1,\ldots ,c_n \in \mathbb {C}$ , and any $x_1,\ldots ,x_n \in X$ , we have

(2.18) $$ \begin{align} \sum_{i,j=1}^n c_i\overline{c_j} \varphi(x_i,x_j) \geqslant 0. \end{align} $$

In this definition, it is enough to consider mutually different elements $x_1,\ldots ,x_n$ of X. By [Reference Bekka, de la Pierre and ValetteBCR84, Section 1.5, p. 68], any positive-definite kernel is Hermitian.

Example 2.2 If X is a finite set, then the function $\varphi \colon X \times X \to \mathbb {C}$ is a positive-definite kernel if and only if the matrix $[\varphi (x_i,x_j)]$ is positive-definite.

Example 2.3 Let H be a Hilbert space, and let $\alpha \colon X \to H$ be a map. By [Reference BożejkoBoz87, p. 9], the function $\varphi \colon X \times X \to \mathbb {C}$ defined by

(2.19) $$ \begin{align} \varphi(x,y) = \langle \alpha_x,\alpha_y\rangle_H, \quad x,y \in X, \end{align} $$

is a positive-definite kernel.

Conversely, by [Reference Bekka, de la Pierre and ValetteBCR84, p. 82], every positive-definite kernel $\varphi $ is of this form for some suitable Hilbert space H and some map $\alpha \colon X \to H$ . Moreover, if $\varphi $ is real-valued, we can take a real Hilbert space.

We say that $\varphi \colon X \times X \to \mathbb {C}$ is a negative-definite kernel if $\varphi $ is Hermitian and if for any integer $n \geqslant 1$ , any $c_1,\ldots ,c_n \in \mathbb {C}$ with $c_1+\cdots +c_n=0$ , and any $x_1,\ldots ,x_n \in X$ , we have

$$ \begin{align*}\sum_{i,j=1}^n c_i \overline{c_j} \varphi(x_i,x_j) \leqslant 0. \end{align*} $$

Example 2.4 Let H be a real Hilbert space, and let $\alpha \colon X \to H$ be a map. The map $\psi \colon X \times X \to \mathbb {R}$ defined by $\psi (x,y) = \| \alpha _x - \alpha _y \|_H^2$ is a negative-definite kernel.

By [Reference Bekka, de la Pierre and ValetteBCR84, Proposition 3.2, p. 82], every real-valued negative-definite kernel $\psi $ which vanishes on the diagonal $\{(x,x) : x \in X\}$ is of this form for a real Hilbert space H.

The following is the famous Schoenberg’s theorem [Reference Bekka, de la Pierre and ValetteBCR84, Theorem 2.2, p. 74], which relates negative-definite kernels and positive-definite kernels.

Theorem 2.5 Let X be a set, and let $\psi \colon X \times X \to \mathbb {C}$ be a map. Then $\psi $ is a negative-definite kernel if and only if for any $t>0$ the map $\mathrm {e}^{-t\psi } \colon X \times X \to \mathbb {C}$ is a positive-definite kernel.

3 Complements on measurable kernels

The following lemma is folklore. For the sake of completeness, we give an argument following Cartier [Reference CartierCar81, Lemme, p. 287].

Lemma 3.1 Let H be a Hilbert space. Suppose that X is a set equipped with a $\sigma $ -algebra $\mathcal {A}$ . Let $\alpha \colon X \to H$ be a function. For any $x,y \in X$ , we define $\varphi \colon X \times X \to \mathbb {C}$ by (2.19). If $\varphi $ is measurable with respect to $\mathcal {A} \otimes \mathcal {A}$ , then the map $\alpha $ is weak measurable, i.e., for any $h \in H$ , the map $X \to \mathbb {C}$ , $x \mapsto \langle \alpha _x,h \rangle _H$ is measurable.

Proof We denote by $H_1$ the set of vectors h of H such that the function $X \to \mathbb {C}$ , $x \mapsto \langle \alpha _x ,h \rangle _H$ is measurable. It is obvious that $H_1$ is a closed subspace of H. It is clear that any vector of H which is orthogonal to the range of $\alpha $ belongs to $H_1$ , i.e., $\alpha (X)^\perp \subset H_1$ . We infer that $\overline {\mathop {\mathrm {span}\,}\nolimits \alpha (X)}^\perp \subset H_1$ . Since the partial functions of $\varphi $ are measurable, for any $y \in X$ , the map $X \to \mathbb {C}$ , $x \mapsto \langle \alpha _x,\alpha _{y} \rangle _H$ is measurable. So we also have the inclusion $\alpha (X) \subset H_1$ , hence $\overline {\mathop {\mathrm {span}\,}\nolimits \alpha (X)} \subset H_1$ . We conclude that $H \subset H_1$ .

Lemma 3.2 Suppose that X is a set equipped with a $\sigma $ -algebra $\mathcal {A}$ . Let $\psi \colon X \times X \to \mathbb {R}$ be a measurable real-valued negative-definite kernel which vanishes on the diagonal $\{ (x,x) : x \in X \}$ . Then there exist a real Hilbert space H and a weak measurable map $\alpha \colon X \to H$ such that

$$ \begin{align*}\psi(x,y) = \| \alpha_x - \alpha_y \|_H^2, \quad x,y \in X. \end{align*} $$

Proof Fix $x_0 \in X$ . Recall that by [Reference Bekka, de la Pierre and ValetteBCR84, Lemma 2.1, p. 74], the function ${\varphi \colon X \times X \to \mathbb {R}}$ defined by

(3.1) $$ \begin{align} \varphi(x,y) =\frac{1}{2}\big[\psi(x,x_0)+\psi(x_0,y)-\psi(x,y)\big], \quad x,y \in X, \end{align} $$

is a positive-definite kernel. Since $\psi $ is measurable, this kernel is measurable. So there exist a real Hilbert space H and a function $\alpha \colon X \to H$ satisfying (2.19), which is weak measurable by Lemma 3.1. Now, for any $x,y \in X$ , we have

$$ \begin{align*} \| \alpha_x - \alpha_y \|_H^2 = \| \alpha_x \|_H^2+ \| \alpha_y \|_H^2-2 \langle \alpha_x,\alpha_y \rangle_H \overset{(2.19)}{=} \varphi(x,x)+\varphi(y,y)-2\varphi(x,y) \\ &\overset{(3.1)}{=}\frac{1}{2}\big(\psi(x,x_0)+\psi(x_0,x)-\psi(x,x)\big)+\frac{1}{2}\big(\psi(y,x_0)+\psi(x_0,y)-\psi(y,y)\big)\\ &\qquad-\big(\psi(x,x_0)+\psi(x_0,y)-\psi(x,y)\big) =\psi(x,y).\\[-37pt] \end{align*} $$

We introduce the following definition in the spirit of [Reference MercerMer21, p. 25].

Definition 3.1 Let X be a measure space. We say that a measurable map $\varphi \colon X \times X \to \mathbb {C}$ is an integrally positive-definite kernel if for any $\xi \in {\mathrm {L}}^1(X)$ the function $(x,y) \mapsto \varphi (x,y) \overline {\xi (x)} \xi (y)$ is integrable on $X \times X$ and

(3.2) $$ \begin{align} \iint_{X \times X} \varphi(x,y) \overline{\xi(x)} \xi(y) \mathop{}\mathopen{}\mathrm{d} x \mathop{}\mathopen{}\mathrm{d} y \geqslant 0. \end{align} $$

Let $\varphi \colon X \times X \to \mathbb {C}$ be a measurable positive-definite kernel on a $\sigma $ -finite measure space X. Cartier showed in [Reference CartierCar81, p. 302] that (3.2) is satisfied for any measurable function $\xi \colon X \to \mathbb {C}$ if the previous integral is absolutely convergent. Moreover, Cartier proved the following result [Reference CartierCar81, Corollaire, p. 301].

Proposition 3.3 Suppose that X is a $\sigma $ -finite measure space. Let $\varphi \colon X \times X \to \mathbb {C}$ be a measurable and bounded function. Then $\varphi $ is equal almost everywhere to a bounded and measurable positive-definite kernel on $X \times X$ if and only if $\varphi $ is an integrally positive-definite kernel.

Remark 3.4 The author naively thought that he proved the “if” part of an ${\mathrm {L}}^2(X)$ -variant of this result in a previous version of this paper relying on the version of Mercer’s theorem for finite measure spaces stated in [Reference KönigKon86, Theorem 3.a.1, p. 145]. Unfortunately, the proof of [Reference KönigKon86, Theorem 3.a.1, p. 145] has a large gap.Footnote 4 In particular, the estimate $\sup _n \| f_n \|_\infty <\infty $ is not proved and seems doubtful, and consequently the arguments of convergence of [Reference KönigKon86] are inexact.

Remark 3.5 It would be interesting to study variants of Definition 3.1 by replacing the space ${\mathrm {L}}^1(X)$ by another space of functions in the spirit of [Reference Husain and WarsiHuW75] and the references therein.

The following is [Reference CartierCar81, Théorème, p. 283]Footnote 5 and would be used in the next section.

Proposition 3.6 Let X be a locally compact space equipped with a Radon measure with support X, and let $\varphi \colon X \times X \to \mathbb {C}$ be a continuous function. Then $\varphi $ is a positive-definite kernel if and only if for any continuous function $\xi \colon X \to \mathbb {C}$ with compact support, we have (3.2).

4 A description of Markov semigroups of measurable Schur multipliers

Let X be a $\sigma $ -finite measure space. We say that a function $\varphi \in {\mathrm {L}}^\infty (X \times X)$ induces a measurable Schur multiplier on $\mathrm {B}({\mathrm {L}}^2(X))$ if the map $S^2_X \mapsto \mathrm {B}({\mathrm {L}}^2(X))$ , $K_{f} \mapsto K_{\varphi f}$ induces a bounded operator from $S^\infty _X$ into $\mathrm {B}({\mathrm {L}}^2(X))$ . In this case, the operator $S^\infty _X \mapsto \mathrm {B}({\mathrm {L}}^2(X))$ , $K_{f}\mapsto K_{\varphi f}$ admits by [Reference Blecher and Le MerdyBLM04, Lemma A.2.2, p. 360] a unique weak* extension $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ called the measurable Schur multiplier associated with $\varphi $ . It is known that in this case

(4.1) $$ \begin{align} \| \varphi \|_{{\mathrm{L}}^\infty(X \times X)} \leqslant \| M_\varphi \|_{\mathrm{B}({\mathrm{L}}^2(X)) \to \mathrm{B}({\mathrm{L}}^2(X))}. \end{align} $$

We refer to the surveys [Reference van NeervenTod15, Reference Todorov and TurowskaToT10] for more information. See also [Reference StratilaSpr04].

Example 4.1 If the set $X=\{1,\ldots ,n\}$ is equipped with the counting measure, we can identify the space $\mathrm {B}({\mathrm {L}}^2(X))$ with the matrix algebra $\mathrm {M}_n$ . Then each operator $K_{f}$ identifies to the matrix $[f(i,j)]$ . A Schur multiplier is given by a map $M_\varphi \colon \mathrm {M}_n \to \mathrm {M}_n$ , $[f(i,j)] \mapsto [\varphi (i,j)f(i,j)]$ .

We say that a measurable Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ is self-adjoint if its restriction $S^2_X \mapsto S^2_X$ , $K_{f} \mapsto K_{\varphi f}$ on the Hilbert space $S^2_X$ is self-adjoint. See [Reference Junge and MeiJMX06, Section 5] for more information on a related notion for operators acting on von Neumann algebras.

Proposition 4.2 Suppose that X is a $\sigma $ -finite measure space. A measurable Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ is self-adjoint if and only if the function $\varphi $ is real-valued almost everywhere.

Proof The self-adjointness property is equivalent to

$$ \begin{align*}\operatorname{\mathrm{Tr}}\big(M_\varphi(K_f)K_g^*\big) =\operatorname{\mathrm{Tr}}\big(K_f(M_\varphi(K_g))^*\big), \quad \text{i.e.,} \quad \operatorname{\mathrm{Tr}}\big(K_{\varphi f}K_g^*\big) =\operatorname{\mathrm{Tr}}\big(K_f(K_{\varphi g})^*\big) \end{align*} $$

for any $f,g \in {\mathrm {L}}^2(X \times X)$ . Since ${\mathrm {L}}^2(X \times X) \to S^2_X$ , $f \mapsto K_f$ is an isometry from the Hilbert space ${\mathrm {L}}^2(X \times X)$ onto the Hilbert space $S^2_X$ of Hilbert–Schmidt operators on ${\mathrm {L}}^2(X)$ , that means that

$$ \begin{align*}\iint_{X \times X} \varphi(x,y)f(x,y)\overline{g(x,y)} \mathop{}\mathopen{}\mathrm{d} x\mathop{}\mathopen{}\mathrm{d} y =\iint_{X \times X} f(x,y) \overline{\varphi(x,y)g(x,y)} \mathop{}\mathopen{}\mathrm{d} x\mathop{}\mathopen{}\mathrm{d} y. \end{align*} $$

Note that $f\overline {g}$ belongs to ${\mathrm {L}}^1(X \times X)$ and that each function of ${\mathrm {L}}^1(X \times X)$ has this form. By duality, we deduce that $\varphi =\overline {\varphi }$ almost everywhere.

We will show that a measurable Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ is unital if and only if its restriction on $S^1_X$ is trace-preserving. We need some information on Arens products to prove some preliminary result. Let A be a Banach algebra. Recall that we can equip its bidual $A^{**}$ with two natural products, the left and right Arens products defined in [Reference Blecher and Le MerdyBLM04, pp. 78–79] on $A^{**}$ . We say A is Arens regular if these products coincide on $A^{**}$ . Note that the algebra $S^\infty _X$ of compact operators is Arens regular by [Reference RicardPal01, Theorem 9.1.39, p. 838].

Proposition 4.3 Let X be a $\sigma $ -finite measure space. The map $S^2_X \to S^2_X$ , $K_f \mapsto K_{\check {f}}$ extends to an involutive normal $*$ -antiautomorphism $R \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ .

Proof For any $f \in {\mathrm {L}}^2(X \times X)$ , we let $R(K_f) \overset {\mathrm {def}}{=} K_{\check {f}}$ . On the one hand, for any $f,g \in {\mathrm {L}}^2(X \times X)$ , the Hilbert–Schmidt operator $K_{f}K_g \colon {\mathrm {L}}^2(X) \to {\mathrm {L}}^2(X)$ is associated with the function h defined by (2.17), and for almost all $(x,y) \in X \times X$ , we have

$$ \begin{align*}\check{h}(x,y)= \int_X f(y,z)g(z,x) \mathop{}\mathopen{}\mathrm{d} z. \end{align*} $$

On the other hand, by (2.17), the Hilbert–Schmidt operator $K_{\check {g}}K_{\check {f}} \colon {\mathrm {L}}^2(X) \to {\mathrm {L}}^2(X)$ is associated with the function

$$ \begin{align*}(x,y) \mapsto \int_X \check{g}(x,z)\check{f}(z,y) \mathop{}\mathopen{}\mathrm{d} z=\int_X f(y,z) g(z,x) \mathop{}\mathopen{}\mathrm{d} z. \end{align*} $$

We obtain that $R(K_fK_g)=R(K_g)R(K_f)$ . Moreover, we have

$$ \begin{align*}(R(K_{f}))^* =\big(K_{\check{f}}\big)^* \overset{(2.16)}{=} K_{\overline{f}} =R(K_{\check{\overline{f}}}) \overset{(2.16)}{=} R((K_f)^*). \end{align*} $$

We conclude that the map $S^2_X \to S^2_X$ , $K_f \mapsto K_{\check {f}}$ is an involutive $*$ -antiautomorphism. From the formula $(K_f)^*=K_{\check {\overline {f}}}$ , it is not difficult to see that the Banach adjoint of $K_f$ is $K_{\check {f}}$ . Hence, $ \| K_f \|_{\mathrm {B}({\mathrm {L}}^2(X))}= \| K_{\check {f}} \|_{\mathrm {B}({\mathrm {L}}^2(X))}$ . We deduce an involutive $*$ -antiautomorphism $R \colon S^\infty _X \to S^\infty _X$ . By [Reference Blecher and Le MerdyBLM04, Lemma A.2.2, p. 360], we deduce a unique weak* continuous extension $R \colon (S^\infty _X)^{**} \to \mathrm {B}({\mathrm {L}}^2(X))$ . Moreover, by [Reference Blecher and Le MerdyBLM04, Section 2.5.5, p. 79], this map is a $*$ -antiautomorphism where the double dual $(S^\infty _X)^{**}$ is equipped with the Arens product. Finally, note that by [Reference RicardPal01, Theorem 9.1.39, p. 838], the algebra $(S^\infty _X)^{**}$ identifies with the algebra $\mathrm {B}({\mathrm {L}}^2(X))$ .

We introduce the following duality bracket:

(4.2) $$ \begin{align} \langle z,y \rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} \overset{\mathrm{def}}{=} \operatorname{\mathrm{Tr}}(R(z)y), \quad z \in \mathrm{B}({\mathrm{L}}^2(X)), y \in S^1_X. \end{align} $$

Proposition 4.4 Let X be a $\sigma $ -finite measure space. The preadjoint $(M_{\varphi })_* \colon S^1_X\to S^1_X$ of a measurable Schur multiplier $M_{\varphi } \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ coincides with its restriction on $S^1_X$ , i.e., for any $z \in \mathrm {B}({\mathrm {L}}^2(X))$ and $y \in S^1_X$ , we have

(4.3) $$ \begin{align} \big\langle M_{\varphi}(z),y \big\rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} =\big\langle z,M_{\varphi}(y) \big\rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X}. \end{align} $$

Proof Recall that the map ${\mathrm {L}}^2(X \times X) \to S^2_X$ , $f \mapsto K_f$ is an isometry from the Hilbert space ${\mathrm {L}}^2(X \times X)$ onto the Hilbert space $S^2_X$ of Hilbert–Schmidt operators on ${\mathrm {L}}^2(X)$ . For any $K_f,K_g \in S^1_X$ , we deduce that

$$ \begin{align*} \big\langle M_{\varphi}(K_f),K_g \big\rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} =\big\langle K_{\varphi f},K_g \big\rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} \overset{(4.2)}{=} \operatorname{\mathrm{Tr}}(R(K_{\varphi f})K_g) \\ &=\operatorname{\mathrm{Tr}}(K_{\check{\varphi} \check{f}}K_g) =\iint_{X \times X} \varphi(y,x)f(y,x) g(y,x) \mathop{}\mathopen{}\mathrm{d} x\mathop{}\mathopen{}\mathrm{d} y. \end{align*} $$

Moreover, we have

$$ \begin{align*} \big\langle K_f,K_{\varphi g} \big\rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} \overset{(4.2)}{=} \operatorname{\mathrm{Tr}}(R(K_f) K_{\varphi g})=\operatorname{\mathrm{Tr}}(K_{\check{f}} K_{\varphi g}) \\ &=\iint_{X \times X} f(y,x)\varphi(y,x) g(y,x) \mathop{}\mathopen{}\mathrm{d} x\mathop{}\mathopen{}\mathrm{d} y. \end{align*} $$

We conclude by density.

Now, we prove the following equivalence by a standard argument.

Proposition 4.5 Suppose that X is a $\sigma $ -finite measure space. A measurable Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ is unital if and only if it is trace-preserving when restricted to the space $S^1_X$ .

Proof It is elementary to check that a $*$ -antiautomorphism is necessarily unital. If the measurable Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ is unital, then for any $y \in S^1_X$ , we have $\big \langle M_{\varphi }(1),y \big \rangle _{\mathrm {B}({\mathrm {L}}^2(X)), S^1_X} \overset {(4.3)}{=} \big \langle 1,M_{\varphi }(y) \big \rangle _{\mathrm {B}({\mathrm {L}}^2(X)), S^1_X}$ , that is, $\operatorname {\mathrm {Tr}} y=\operatorname {\mathrm {Tr}} M_{\varphi }(y)$ . We conclude that $M_\varphi $ is trace-preserving.

If the measurable Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ is trace-preserving, then for any $x \in S^1_X$ , we have

$$ \begin{align*} \operatorname{\mathrm{Tr}} (R(M_\varphi(1))x) \overset{(4.2)}{=} \big\langle M_{\varphi}(1),x \big\rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} \overset{(4.3)}{=} \big\langle 1,M_{\varphi}(x) \big\rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} \\ &\overset{(4.2)}{=} \operatorname{\mathrm{Tr}}(R(1)M_\varphi(x)) =\operatorname{\mathrm{Tr}}(R(1)x). \end{align*} $$

We conclude by duality that $M_\varphi (1)=1$ .

It is proved in [Reference DuquetDuq22, Theorem 3.5] that a measurable Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ is unital if and only if $\varphi (x,x)=1$ almost everywhere on X. We deduce the following result.

Corollary 4.6 Suppose that X is a $\sigma $ -finite measure space. Let $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ be a measurable Schur multiplier. The following conditions are equivalent.

  1. (1) $M_\varphi $ is trace-preserving when restricted to the space $S^1_X$ .

  2. (2) $M_\varphi $ is unital.

  3. (3) $\varphi (x,x)=1$ almost everywhere on X.

In the following statement, we say that $\varphi $ induces a (completely) positive measurable Schur multiplier if the linear map $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ is (completely) positive.

Proposition 4.7 Let X be a $\sigma $ -finite measure space. Let $\varphi \in {\mathrm {L}}^\infty (X \times X)$ . Consider the following properties.

  1. (1) $\varphi $ induces a completely positive measurable Schur multiplier.

  2. (2) $\varphi $ induces a positive measurable Schur multiplier.

  3. (3) The function $\varphi $ is an integrally positive-definite kernel.

We have $1 \iff 2 \Rightarrow 3$ .

Proof 1. $\Rightarrow $ 2. It is obvious.

2. $\Rightarrow $ 1. That is essentially a standard application of [Reference TodorovSTT14, Lemma 4.3]. See also [Reference van Neerven, Veraar and WeisTod14, Exercise 4.15].

2. $\Rightarrow $ 3. Let $\xi \in {\mathrm {L}}^1(X)$ . We can write $\xi =\xi _1\xi _2$ for some functions $\xi _1,\xi _2 \in {\mathrm {L}}^2(X)$ . Note that the operator $K_{\overline {\xi _1} \otimes \xi _1} \colon {\mathrm {L}}^2(X) \to {\mathrm {L}}^2(X)$ is positive. Indeed, for any $\eta \in {\mathrm {L}}^2(X)$ , we have

$$ \begin{align*} \big\langle K_{\overline{\xi_1} \otimes \xi_1}(\eta),\eta \big\rangle_{{\mathrm{L}}^2(X)} \overset{(2.15)}{=} \bigg(\int_X \xi_1 \eta \bigg)\big\langle \overline{\xi_1} ,\eta \big\rangle_{{\mathrm{L}}^2(X)} \\ &=\overline{\big\langle \overline{\xi_1} ,\eta \big\rangle_{{\mathrm{L}}^2(X)}} \big\langle \overline{\xi_1} ,\eta \big\rangle_{{\mathrm{L}}^2(X)} =\big|\big\langle \overline{\xi_1} ,\eta \big\rangle_{{\mathrm{L}}^2(X)} \big|^2 \geqslant 0. \end{align*} $$

We deduce that $\langle M_\varphi (K_{\overline {\xi _1} \otimes \xi _1})\xi _2,\xi _2 \rangle _{{\mathrm {L}}^2(X)} \geqslant 0$ . Finally, it suffices to observe that

$$ \begin{align*} &\big\langle M_\varphi(K_{\overline{\xi_1} \otimes \xi_1})\xi_2,\xi_2 \big\rangle_{{\mathrm{L}}^2(X)} =\Big\langle K_{\varphi(\overline{\xi_1} \otimes \xi_1)} \xi_2,\xi_2 \Big\rangle_{{\mathrm{L}}^2(X)} =\int_X \big(K_{\varphi(\overline{\xi_1} \otimes \xi_1)} \xi_2\big)(x)\overline{\xi_2(x)} \mathop{}\mathopen{}\mathrm{d} x \\ &\!\!\!\quad\overset{(2.15)}{=} \int_X \bigg(\int_X \varphi(x,y)\overline{\xi_1(x)} \xi_1(y)\xi_2(y)\mathop{}\mathopen{}\mathrm{d} y\bigg)\overline{\xi_2(x)} \mathop{}\mathopen{}\mathrm{d} x \\ &\,\quad=\iint_{X \times X} \varphi(x,y) \overline{\xi_1(x)} \xi_1(y)\xi_2(y) \overline{\xi_2(x)} \mathop{}\mathopen{}\mathrm{d} x \mathop{}\mathopen{}\mathrm{d} y =\iint_{X \times X} \varphi(x,y) \overline{\xi(x)} \xi(y) \mathop{}\mathopen{}\mathrm{d} x \mathop{}\mathopen{}\mathrm{d} y. \end{align*} $$

So (3.2) is satisfied.

Remark 4.8 Let X be a locally compact space equipped with a Radon measure with full support. Let $\varphi \in {\mathrm {L}}^\infty (X \times X)$ be a function which induces a measurable Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ . The statement “3” implies 2” is true. Suppose that $\varphi $ is an integrally positive-definite kernel. By Proposition 3.3, we can suppose that $\varphi $ is equal to a bounded and measurable positive-definite kernel on $X \times X$ . Let $K_f \colon {\mathrm {L}}^2(X) \to {\mathrm {L}}^2(X)$ be a positive Hilbert–Schmidt operator with a continuous symbol $f \colon X \times X \to \mathbb {C}$ . Then the inequality

$$ \begin{align*}\iint_{X \times X} f(x,y) \overline{\xi(x)} \xi(y) \mathop{}\mathopen{}\mathrm{d} x \mathop{}\mathopen{}\mathrm{d} y \geqslant 0 \end{align*} $$

is clearly satisfied for any $\xi \in {\mathrm {L}}^2(X)$ , in particular for any function $\xi \in \mathrm {C}_c(X)$ . By Proposition 3.6, we deduce that f is the class of a positive-definite kernel. Using the stability under pointwise product of the set of definite-positive kernels [Reference Bekka, de la Pierre and ValetteBCR84, Theorem 1.12, p. 69], we see that the function $\varphi f$ of ${\mathrm {L}}^2(X \times X)$ is also the class of a positive-definite kernel. By the observation following Definition 3.1, we deduce that for any $\xi \in {\mathrm {L}}^2(X)$ , we have

$$ \begin{align*}\iint_{X \times X} \varphi(x,y)f(x,y) \overline{\xi(x)} \xi(y) \mathop{}\mathopen{}\mathrm{d} x \mathop{}\mathopen{}\mathrm{d} y \geqslant 0. \end{align*} $$

Consequently $M_\varphi (K_f)=K_{\varphi f}$ is a positive operator on the Hilbert space ${\mathrm {L}}^2(X)$ . We conclude with Kaplansky’s theorem using the cone of positive Hilbert–Schmidt operators with continuous kernels.

Remark 4.9 Let X be a locally compact space equipped with a Radon measure with full support. Here, we explain how to give an answer to [Reference TodorovSTT14, Question 4.13(ii)]. Consider a positive continuous Schur multiplier $M_\varphi \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ , i.e., a measurable Schur multiplier with a continuous symbol $\varphi \colon X \times X \to \mathbb {C}$ . By Proposition 4.7, the map $\varphi $ is an integrally positive-definite kernel. In particular, for any continuous function $ \xi \colon X \to \mathbb {C}$ with compact support, we have (3.2). By Proposition 3.6, we infer that the continuous function $\varphi $ is a positive-definite kernel. Consequently, by [Reference Bekka, de la Pierre and ValetteBCR84, p. 82], there exist a Hilbert space H and some map $\alpha \colon X \to H$ satisfying (2.19). Since the function $\varphi $ is continuous, we conclude by the (classical) result [Reference CartierCar81, Lemme, p. 279]Footnote 6 that the map $\alpha \colon X \to H$ is necessarily continuous (here the Hilbert space H is equipped with its usual topology). Since $\varphi \in {\mathrm {L}}^\infty (X \times X)$ , the function $x \mapsto \varphi (x,x)=\langle \alpha _x,\alpha _x \rangle _H$ is bounded. We conclude that $\alpha $ is bounded.

With sharp contrast, we think that the answer to the similar question [Reference TodorovSTT14, Question 4.13(i)] for (not necessarily positive) continuous Schur multipliers is negative.

The following is a generalization of [Reference ArhancetArh13a, Proposition 5.4]. Recall that in this paper, the weak* continuity of a semigroup $(T_t)_{t \geqslant 0}$ means that the map $\mathbb {R}^+ \to \mathbb {C}$ , $t \mapsto \langle T_t(x),y \rangle _{\mathrm {B}({\mathrm {L}}^2(X)),S^1_X}$ is continuous for any $x \in \mathrm {B}({\mathrm {L}}^2(X))$ and any $y \in S^1_X$ .

Proposition 4.10

  1. (1) Let X be a second countable Radon measure space. If $(T_t)_{t \geqslant 0}$ is a weak* continuous semigroup of self-adjoint unital completely positive measurable Schur multipliers on $\mathrm {B}({\mathrm {L}}^2(X))$ with continuous symbols, then there exist a real Hilbert space H and a continuous function $\alpha \colon X \to H$ such that the symbol of $T_t$ is

    (4.4) $$ \begin{align} \phi_t(x,y) =\mathrm{e}^{-t \| \alpha_x-\alpha_y \|_{H}^2}, \quad \text{for almost all }x,y \in X. \end{align} $$
  2. (2) Let X be a (complete) $\sigma $ -finite measure space such that if $(T_t)_{t \geqslant 0}$ is a weak* continuous semigroup of self-adjoint unital completely positive measurable Schur multipliers on $\mathrm {B}({\mathrm {L}}^2(X))$ , then there exist a real Hilbert space H and a weak measurable function $\alpha \colon X \to H$ such that the symbol of $T_t$ is given by (4.4).

Proof 1. For any $t \geqslant 0$ , let $\phi _t \colon X \times X \to \mathbb {C}$ be the continuous symbol of the Schur multiplier $T_t$ . Since the map $T_t$ is self-adjoint, the function $\phi _t$ is real-valued by Proposition 4.2. By Corollary 4.6, we have $\phi _t(x,x)=1$ for any $x \in X$ and any $t \geqslant 0$ . By Proposition 4.7, each $\phi _t$ is an integrally positive-definite kernel. We deduce by Proposition 3.6 that each $\phi _t$ is a positive-definite kernel. Finally, we have $\phi _0=1$ .

For any $t,t' \geqslant 0$ and any $f \in {\mathrm {L}}^2(X \times X)$ , the relation $T_tT_{t'}(K_{f})=T_{t+t'}(K_f)$ gives the equality $K_{\phi _{t}\phi _{t'} f}=K_{\phi _{t+t'}f}$ . We infer that $\phi _t\phi _{t'}f=\phi _{t+t'}f$ in the space ${\mathrm {L}}^2(X \times X)$ . It is apparent that $(\phi _t)_{t \geqslant 0}$ defines a semigroup $(\mathrm {Mult}_t)_{t \geqslant 0}$ of multiplication operators acting on ${\mathrm {L}}^2(X \times X)$ defined by

(4.5) $$ \begin{align} \mathrm{Mult}_t(f) \overset{\mathrm{def}}{=} \phi_tf, \quad t \geqslant 0, f \in {\mathrm{L}}^2(X \times X). \end{align} $$

Note that (4.1) implies that $(\mathrm {Mult}_t)_{t \geqslant 0}$ is a contraction semigroup. We will show the weak continuity of this semigroup. For any functions $f,g \in {\mathrm {L}}^2(X \times X)$ such that $K_g \in S^1_X$ , we have using the weak* continuity of the semigroup $(T_t)_{t \geqslant 0}$

$$ \begin{align*} \big\langle \mathrm{Mult}_t(f),g \big\rangle_{{\mathrm{L}}^2(X \times X)} \overset{(4.5)}{=} \langle \phi_tf,g \rangle_{{\mathrm{L}}^2(X \times X)} =\big\langle K_{\phi_t f}, K_g \big\rangle_{S^2_X} \\ &=\big\langle T_{t}(K_f), K_g \big\rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} \xrightarrow[t \to 0]{} \langle K_f, K_g \rangle_{\mathrm{B}({\mathrm{L}}^2(X)), S^1_X} =\langle f,g \rangle_{{\mathrm{L}}^2(X \times X)} \end{align*} $$

(here, we make a slight abuse of notation for the definition of the brackets). Note that the set of functions $g \in {\mathrm {L}}^2(X \times X)$ such that $K_g \in S^1_X$ is dense in ${\mathrm {L}}^2(X \times X)$ since $S^1_X$ is dense in the space $S^2_X$ . Now, with [Reference MegginsonMeg98, Exercise 2.71(a), p. 234], it is clear (using the contractivity of the semigroup) that this semigroup is weak continuous. By [Reference Engel and NagelEnN00, Theorem 5.8, p. 40], this semigroup is even strongly continuous. By [Reference Engel and NagelEnN00, Proposition 4.12, p. 32], there exists a measurable function $\psi \colon X \times X \to \mathbb {R}$ such that for almost all $x,y \in X$ ,

$$ \begin{align*}\phi_t(x,y) =\mathrm{e}^{-t\psi(x,y)}, \quad t \geqslant 0. \end{align*} $$

We deduce that for almost all $x,y \in X$ , we have if $t>0$ ,

$$ \begin{align*}\psi(x,y) =-\frac{1}{t}\log \phi_t(x,y). \end{align*} $$

Since $\phi _t$ is continuous, we can suppose that $\psi $ is continuous. Moreover, for any $x \in X$ , we have

$$ \begin{align*}\psi(x,x) =-\frac{1}{t}\log \phi_t(x,x) =0. \end{align*} $$

By Schoenberg’s theorem (Theorem 2.5), the function $\psi $ defines a real-valued negative definite kernel which vanishes on the diagonal of $X \times X$ . Finally, we use the characterization of continuous real-valued negative-definite kernels of [Reference Berg, Christensen and ResselBHV08, Theorem C.2.3].

2. For any $t \geqslant 0$ , let $\phi _t \colon X \times X \to \mathbb {C}$ be the symbol of $T_t$ . Since the map $T_t$ is self-adjoint, the function $\phi _t$ is real-valued almost everywhere by Proposition 4.2. Consequently, we can suppose that each function $\phi _t$ is real-valued. By Proposition 4.7, each $\phi _t$ is an integrally positive-definite kernel. Consequently, we can suppose by Proposition 3.3 that the function $\phi _t$ is a real-valued positive-negative kernel and bounded (by 1).

By Corollary 4.6, we have $\phi _t(x,x)=1$ for almost all $x \in X$ and any $t \geqslant 0$ . For any $t>0$ , let $A_t$ be the subset of X such that $\phi _t(x,x) \not =1$ . The subset $\{ (x,x) : x \in A_t \}$ is a subset of $A_t \times A_t$ , hence negligible. By modifying $\phi _t$ on $A_t$ , we can suppose that $\phi _t(x,x) =1$ for any $x \in X$ . Indeed, it is clear by observing the definition (2.18) that the property positive-definite is not lost by this change.

The end of the proof is similar to the proof of the first point replacing [Reference Berg, Christensen and ResselBHV08, Theorem C.2.3] by Lemma 3.2.

Remark 4.11 In the opposite direction, consider a separable real Hilbert space H and a weak measurable function $\alpha \colon X \to H$ . Let $(e_n)_{n \geqslant 1}$ be an orhonormal basis of H. We can write

(4.6) $$ \begin{align} \langle \alpha_x,\alpha_y \rangle_H =\sum_{n \geqslant 1} \langle \alpha_x, e_n \rangle_H \langle e_n, \alpha_y \rangle_H, \quad x,y \in X. \end{align} $$

So the map $X \times X \to \mathbb {C}$ , $(x,y) \mapsto \langle \alpha _x,\alpha _y \rangle _H$ is measurable with respect to the $\sigma $ -algebra $\mathcal {A} \otimes \mathcal {A}$ . Since we have

$$ \begin{align*}\| \alpha_x - \alpha_y \|_H^2 = \| \alpha_x \|_H^2+ \| \alpha_y \|_H^2-2 \langle \alpha_x,\alpha_y \rangle_H, \end{align*} $$

we see that the symbol $\phi _t$ defined in (4.4) is measurable.

Note that $\phi _t$ converges to the function 1 when $t \to 0$ for the weak* topology of ${\mathrm {L}}^\infty (X \times X)$ . Indeed, for any $f \in {\mathrm {L}}^1(X \times X)$ , we have using dominated convergence theorem

$$ \begin{align*} \langle\phi_t,f \rangle_{{\mathrm{L}}^\infty(X \times X),{\mathrm{L}}^1(X \times X)} =\int_{X \times X} \mathrm{e}^{-t \| \alpha_x-\alpha_y \|_{H}^2} f(x,y)\mathop{}\mathopen{}\mathrm{d} x \mathop{}\mathopen{}\mathrm{d} y \xrightarrow[t \to 0]{} \int_{X \times X} f(x,y)\mathop{}\mathopen{}\mathrm{d} x \mathop{}\mathopen{}\mathrm{d} y \\ &=\langle 1 ,f \rangle_{{\mathrm{L}}^\infty(X \times X),{\mathrm{L}}^1(X \times X)}. \end{align*} $$

With this convergence, it is left to the reader to show with [Reference Arhancet and Le MerdyArK22b, Lemma 4.25] that the symbols (4.4) define a weak* continuous semigroup $(T_t)_{t \geqslant 0}$ .

If H is not separable, we can obtain the measurability of the symbols (4.4) if we make the stronger assertion that $\alpha $ is strongly measurable [Reference Hytönen, van Neerven, Veraar and WeisHvNVW16, Definition 1.1.14, p. 8], which is equivalent by the Pettis measurability theorem [Reference Hytönen, van Neerven, Veraar and WeisHvNVW16, Theorem 1.1.20, p. 10] to say that $\alpha $ is weak measurable and essentially separably valued, i.e., there exists a negligible subset N of X such that $\alpha (X-N)$ is separable. In this case, we can find a separable Hilbert space H containing $\alpha (X-N)$ . Consequently, with an orhonormal basis $(e_n)_{n \geqslant 1}$ of H (4.6) is true on $(X-N) \times (X-N)$ . We conclude as previously (if X is complete).

5 Dilations of semigroups of measurable Schur multipliers

Let X be a $\sigma $ -finite measure space. If $g \in {\mathrm {L}}^\infty (X)$ , we will use the multiplication operator $\mathrm {Mult}_{g} \colon {\mathrm {L}}^2(X) \to {\mathrm {L}}^2(X)$ , $\xi \mapsto g\xi $ . Recall that $(\mathrm {Mult}_{g})^*=\mathrm {Mult}_{\overline {g}}$ . We will use the following elementary result, which describes the composition of a Hilbert–Schmidt operator $K_f \colon {\mathrm {L}}^2(X) \to {\mathrm {L}}^2(X)$ with multiplication operators. Recall that the space of Hilbert–Schmidt operators is an ideal of the algebra $\mathrm {B}({\mathrm {L}}^2(X))$ of bounded operators on the Hilbert space ${\mathrm {L}}^2(X)$ .

Lemma 5.1 For any $\xi \in {\mathrm {L}}^2(X)$ , any $g,h \in {\mathrm {L}}^\infty (X)$ , and any $f \in {\mathrm {L}}^2(X \times X)$ , we have for almost all $x \in X$

(5.1) $$ \begin{align} \big((\mathrm{Mult}_{g} K_f \mathrm{Mult}_{h})(\xi)\big)(x) =\int_{X} g(x)h(y)f(x,y)\xi(x) \mathop{}\mathopen{}\mathrm{d} y. \end{align} $$

So the Hilbert–Schmidt operator $\mathrm {Mult}_{g} K_f \mathrm {Mult}_{h} \colon {\mathrm {L}}^2(X) \to {\mathrm {L}}^2(X)$ is associated with the function $(x,y) \mapsto g(x)h(y)f(x,y)\xi (x)$ .

Proof For almost all $x \in X$ , we have

$$ \begin{align*} &\big((\mathrm{Mult}_{g} K_f \mathrm{Mult}_{h})(\xi)\big)(x) =\big((\mathrm{Mult}_{g} K_f)(h\xi)\big)(x) \\ &\quad\!\!\!\overset{(2.15)}{=} \bigg(\mathrm{Mult}_{g}\bigg(\int_{X} f(\cdot,y)h(y)\xi(y) \mathop{}\mathopen{}\mathrm{d} y\bigg)\bigg)(x)\\ &\,\quad=\bigg(g(\cdot)\int_{X} f(\cdot,y)h(y)\xi(y) \mathop{}\mathopen{}\mathrm{d} y\bigg)(x) =\int_{X} g(x)h(y)f(x,y)\xi(x) \mathop{}\mathopen{}\mathrm{d} x.\\[-40pt] \end{align*} $$

The following is the first main result of this paper.

Theorem 5.2 Let X be a $\sigma $ -finite measure space. Let H be a real Hilbert space, and let $\alpha \colon X \to H$ be a strongly measurable map. Then the symbols (4.4) for $t \geqslant 0$ define a weak* continuous semigroup $(T_t)_{t \geqslant 0}$ of self-adjoint unital completely positive Schur multipliers on $\mathrm {B}({\mathrm {L}}^2(X))$ . Moreover, there exist an injective von Neumann algebra M equipped with a normal semifinite faithful trace, a weak* continuous group $(U_t)_{t \in \mathbb {R}}$ of trace-preserving $*$ -automorphisms of M, and a unital trace-preserving injective normal $*$ -homomorphism $J \colon \mathrm {B}({\mathrm {L}}^2(X)) \to M$ such that

(5.2) $$ \begin{align} T_t =\mathbb{E} U_t J, \end{align} $$

for any $t \geqslant 0$ , where $\mathbb {E} \colon M \to \mathrm {B}({\mathrm {L}}^2(X))$ is the canonical trace-preserving normal faithful conditional expectation associated with J.

Proof Let $\mathrm {W} \colon {\mathrm {L}}^2_{\mathbb {R}}(\mathbb {R},H) \to {\mathrm {L}}^0(\Omega )$ be an ${\mathrm {L}}^2_{\mathbb {R}}(\mathbb {R},H)$ -isonormal process on a probability space $(\Omega ,\mu )$ as in (2.4). We define the von Neumann algebra $M \overset {\mathrm {def}}{=} {\mathrm {L}}^\infty (\Omega ) \overline \otimes \mathrm {B}({\mathrm {L}}^2(X))$ . Note that M is an injective von Neumann algebra. We equip this von Neumann algebra with the normal semifinite faithful trace $\tau _M \overset {\mathrm {def}}{=} \int _{\Omega } \cdot \otimes \operatorname {\mathrm {Tr}}$ . By [Reference Steen, Todorov and TurowskaSak71, Theorem 1.22.13], we have a $*$ -isomorphism $M={\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)))$ . We define the canonical unital normal injective $*$ -homomorphism

(5.3) $$ \begin{align} \begin{array}{cccc} J \colon & \mathrm{B}({\mathrm{L}}^2(X)) & \longrightarrow & {\mathrm{L}}^\infty(\Omega) \overline\otimes \mathrm{B}({\mathrm{L}}^2(X)) \\ & z & \longmapsto & 1 \otimes z \\ \end{array}. \end{align} $$

It is clear that the map J preserves the traces. We denote by $\mathbb {E} \colon {\mathrm {L}}^\infty (\Omega ) \overline \otimes \mathrm {B}({\mathrm {L}}^2(X)) \to \mathrm {B}({\mathrm {L}}^2(X))$ the canonical trace-preserving normal faithful conditional expectation of M onto $\mathrm {B}({\mathrm {L}}^2(X))$ . Note that $\mathbb {E}=\int _\Omega \cdot \otimes \mathrm {Id}_{\mathrm {B}({\mathrm {L}}^2(X))}$ . For almost all $\omega \in \Omega $ and any $t \geqslant 0$ , we introduce with (2.7)

(5.4) $$ \begin{align} k_{t,\omega}(x) \overset{\mathrm{def}}{=} \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}_t(\alpha_x))(\omega)}, \quad x \in X. \end{align} $$

Let $(f_i)$ be an orthonormal basis of the separable Hilbert space ${\mathrm {L}}^2_{\mathbb {R}}(\mathbb {R}^+)$ . By the Pettis measurability theorem [Reference Hytönen, van Neerven, Veraar and WeisHvNVW16, Theorem 1.1.20, p. 10], there exists a negligible subset N of X such that $\alpha (X-N)$ is subset of a separable closed subspace $H_0$ of H. Let $(e_j)$ be an orthonormal basis of $H_0$ . For any $t>0$ fixed, we can write almost everywhere (with [Reference Hytönen, van Neerven, Veraar and WeisHvNVW18, Corollary 6.4.4, pp. 35–36])

$$ \begin{align*} (\mathrm{W}_t(\alpha_x))(\omega) \overset{(2.4)(2.7)}{=} \sum_{i,j \in I} \gamma_{i,j}(\omega) \big\langle 1_{[0,t]} \otimes \alpha_x,f_i \otimes e_j \big\rangle_{{\mathrm{L}}^2_{\mathbb{R}}(\mathbb{R}^+,H)} \\ &=\sum_{i,j \in I} \gamma_{i,j}(\omega) \big\langle 1_{[0,t]},f_i \big\rangle_{} \langle \alpha_x, e_j \rangle_H. \end{align*} $$

We conclude that the function $X \times \Omega \to \mathbb {C}$ , $(x,\omega ) \mapsto k_{t,\omega }(x)$ is measurable.

If $t<0$ , we define $k_{t,\omega }(x)$ by the same formula but with $\mathrm {W}_t(\alpha _x) \overset {\mathrm {def}}{=} \mathrm {W}(-1_{[t,0]} \otimes \alpha _x)$ . For any $t \in \mathbb {R}$ , we can define an element $V_t$ of ${\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)))$ by

(5.5) $$ \begin{align} V_t(\omega) \overset{\mathrm{def}}{=} \mathrm{Mult}_{k_{t,\omega}}. \end{align} $$

Indeed, the function $V_t$ is weak* measurable since if we consider a normal linear form $f \colon \mathrm {B}({\mathrm {L}}^2(X)) \to \mathbb {C}$ , $z \mapsto \sum _{n \geqslant 1} \langle z(\xi _n), \eta _n \rangle _{{\mathrm {L}}^2(X)}$ with $\sum _{n \geqslant 1} ( \| \xi _n \|_{{\mathrm {L}}^2(X)}+ \| \eta _n \|_{{\mathrm {L}}^2(X)}) < \infty $ [Reference LiLi92, Proposition 1.2.2, p. 9], we see that

$$ \begin{align*}f\big(\mathrm{Mult}_{k_{t,\omega}}\big) =\sum_{n \geqslant 1} \big\langle \mathrm{Mult}_{k_{t,\omega}}\xi_n, \eta_n \big\rangle_{{\mathrm{L}}^2(X)} =\sum_{n \geqslant 1} \int_X k_{t,\omega} \xi_n \eta_n \end{align*} $$

is measurable since each integral defines a measurable function $\omega \to \int _X k_{t,\omega } \xi _n \eta _n$ by Fubini’s theorem. Note that $V_t$ is an unitary element of ${\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)))$ . Now, for any $t \in \mathbb {R}$ , we define the linear map

(5.6) $$ \begin{align} \begin{array}{cccc} \mathcal{U}_t \colon & {\mathrm{L}}^\infty\big(\Omega,\mathrm{B}({\mathrm{L}}^2(X))\big) & \longrightarrow & {\mathrm{L}}^\infty\big(\Omega,\mathrm{B}({\mathrm{L}}^2(X))\big) \\ & g & \longmapsto & V_t gV_t^* \\ \end{array}. \end{align} $$

It is obvious that each map $\mathcal {U}_t$ is a trace-preserving $*$ -automorphism of the von Neumann algebra M. Moreover, for any elements $\xi ,\eta $ of the Hilbert space ${\mathrm {L}}^2(\Omega ,{\mathrm {L}}^2(X))={\mathrm {L}}^2(\Omega \times X)$ , we obtain usingFootnote 7 the first part of Lemma 2.1 since $\overline {\xi } \eta \in {\mathrm {L}}^1(\Omega \times X)$

$$ \begin{align*} \big\langle V_t(\xi), \eta \big\rangle_{{\mathrm{L}}^2(\Omega,{\mathrm{L}}^2(X))} =\int_{\Omega \times X} \overline{V_t(\xi)(\omega,x)} \eta(\omega,x) \mathop{}\mathopen{}\mathrm{d}\mu(\omega) \mathop{}\mathopen{}\mathrm{d} x \\ &\!\!\!\overset{(5.5)}{=} \int_{\Omega \times X} \overline{k_{t,\omega}(x)\xi(\omega,x)} \eta(\omega,x) \mathop{}\mathopen{}\mathrm{d}\mu(\omega) \mathop{}\mathopen{}\mathrm{d} x \\ &\!\!\!\overset{(5.4)}{=} \int_{\Omega \times X} \mathrm{e}^{-\sqrt{2}{\mathrm{i}} (\mathrm{W}_t(\alpha_x))(\omega)} \overline{\xi(\omega,x)} \eta(\omega,x) \mathop{}\mathopen{}\mathrm{d}\mu(\omega) \mathop{}\mathopen{}\mathrm{d} x \\ &\xrightarrow[t \to 0]{} \int_{\Omega \times X} \mathrm{e}^{-\sqrt{2}{\mathrm{i}} (\mathrm{W}_{t_0}(\alpha_x))(\omega)}\overline{\xi(\omega,x)} \eta(\omega,x)\mathop{}\mathopen{}\mathrm{d}\mu(\omega)\mathop{}\mathopen{}\mathrm{d} x =\langle V_{t_0}(\xi), \eta \rangle_{{\mathrm{L}}^2(\Omega,{\mathrm{L}}^2(X))}. \end{align*} $$

So $(V_t)_{t \in \mathbb {R}}$ is a weak continuous family of unitaries hence a strongly continuous family since by [Reference TodorovStr81, p. 41] the weak operator topology and the strong operator topology coincide on the unitary group. By composition on bounded subsets, the map $\mathbb {R} \to {\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)))$ , $t \mapsto \mathcal {U}_t(g)$ is strong operator continuous for any $g \in {\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)))$ and clearly even strong* operator continuous, hence weak* continuous. We conclude that $(\mathcal {U}_t)_{t \in \mathbb {R}}$ is a point-weak* continuous family of $*$ -automorphisms. For any $f \in {\mathrm {L}}^2(X \times X)$ and any $t \in \mathbb {R}$ , we have

(5.7) $$ \begin{align} \mathcal{U}_t(1 \otimes K_f) \overset{(5.6)}{=} \omega \mapsto (V_t (1 \otimes K_f) V_t^*)(\omega) =\omega \mapsto V_t(\omega) K_f V_t(\omega)^* \\ &\overset{(5.5)}{=} \omega \mapsto \mathrm{Mult}_{k_{t,\omega}} K_f (\mathrm{Mult}_{k_{t,\omega}})^* \nonumber \overset{(5.1)}{=} \omega \mapsto K\big((x,y) \mapsto k_{t,\omega}(x)\overline{k_{t,\omega}(y)}f(x,y) \big) \nonumber \\ &\overset{(5.4)}{=} \omega \mapsto K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}_t(\alpha_x-\alpha_y))(\omega)} f(x,y) \big). \nonumber \end{align} $$

In particular, for almost all $\omega \in \Omega $ , any $\xi \in {\mathrm {L}}^2(X)$ , and almost all $y \in X$ , we have

(5.8) $$ \begin{align} \big(\big[(\mathcal{U}_t(1 \otimes K_f))(\omega)\big](\xi)\big)(y) \overset{(2.15)}{=} \int_X \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}_t(\alpha_x-\alpha_y))(\omega)} f(x,y) \xi(x) \mathop{}\mathopen{}\mathrm{d} x. \end{align} $$

For any $t \in \mathbb {R}$ , we introduce the right shift $\mathcal {S}_t \colon {\mathrm {L}}^2_{\mathbb {R}}(\mathbb {R}) \to {\mathrm {L}}^2_{\mathbb {R}}(\mathbb {R})$ . We consider the operator $S_t \overset {\mathrm {def}}{=} \Gamma _\infty (\mathcal {S}_t) \otimes \mathrm {Id}_{\mathrm {B}({\mathrm {L}}^2(X))} \colon {\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X))) \to {\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)))$ . Essentially, by the second part of Lemma 2.1, $(S_t)_{t \in \mathbb {R}}$ is a weak* continuous semigroup of $*$ -automorphisms. We have

(5.9) $$ \begin{align} S_t\big(\mathrm{e}^{{\mathrm{i}} \mathrm{W}(1_{[s,s'[} \otimes h)} \otimes z\big) \overset{(2.6)}{=} \mathrm{e}^{{\mathrm{i}} \mathrm{W}(1_{[s+t,s'+t[} \otimes h)} \otimes z, \quad t \in \mathbb{R}, h \in H, z \in \mathrm{B}({\mathrm{L}}^2(X)), s<s'. \end{align} $$

For any $t \in \mathbb {R}$ , we define the $*$ -automorphism $U_t \overset {\mathrm {def}}{=} \mathcal {U}_t S_t$ . We will prove that $(U_t)_{t \in \mathbb {R}}$ is a group of operators. On the one hand, for any $t,t' \in \mathbb {R}$ , we have

$$ \begin{align*} U_{t'}U_t\big(\mathrm{e}^{{\mathrm{i}}\mathrm{W}(1_{[s,s']} \otimes h)} \otimes 1\big) =\mathcal{U}_{t'} S_{t'} \mathcal{U}_t S_t\big(\mathrm{e}^{{\mathrm{i}}\mathrm{W}(1_{[s,s']} \otimes h)} \otimes 1\big) \\ &\!\!\!\!\overset{(5.9)}{=} \mathcal{U}_{t'} S_{t'} \mathcal{U}_t\big(\mathrm{e}^{{\mathrm{i}}\mathrm{W}(1_{[s+t,s'+t]} \otimes h)} \otimes 1\big) \\ &=\mathcal{U}_{t'} S_{t'} \big(\mathrm{e}^{{\mathrm{i}}\mathrm{W}(1_{[s+t,s'+t]} \otimes h)} \otimes 1\big) \overset{(5.9)}{=} \mathcal{U}_{t'}\big(\mathrm{e}^{{\mathrm{i}}\mathrm{W}(1_{[s+t+t',s'+t+t']} \otimes h)} \otimes 1\big) \\ &=\mathrm{e}^{{\mathrm{i}}\mathrm{W}(1_{[s+t+t',s'+t+t']} \otimes h)} \otimes 1 =U_{t+t'}\big(\mathrm{e}^{{\mathrm{i}}\mathrm{W}(1_{[s,s']} \otimes h)} \otimes 1\big). \end{align*} $$

On the other hand, for any $t,t' \geqslant 0$ and any $\xi \in {\mathrm {L}}^2(X)$ , we have

$$ \begin{align*} U_{t'}U_t(1 \otimes K_f) =\mathcal{U}_{t'} S_{t'}\mathcal{U}_tS_t(1 \otimes K_f) \\ &\!\!\!\overset{(5.9)}{=} \mathcal{U}_{t'} S_{t'}\mathcal{U}_t (1 \otimes K_f) \\ &\!\!\!\overset{(5.7)}{=} \mathcal{U}_{t'} S_{t'} \big(\omega \mapsto K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}_t(\alpha_x-\alpha_y))(\omega)} f(x,y)\big)\big)\\ &=\mathcal{U}_{t'} \big(\omega \mapsto K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(1_{[t',t+t']} \otimes (\alpha_x-\alpha_y)))(\omega)} f(x,y)\big)\big)\\ &\!\!\!\!\!\!\!\overset{(5.6)(5.5)}{=} \omega \mapsto \mathrm{Mult}_{k_{t',\omega}} K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(1_{[t',t+t']} \otimes (\alpha_x-\alpha_y)))(\omega)} f(x,y)\big) \\ &(\mathrm{Mult}_{k_{t',\omega}})^*\\ &\!\!\!\overset{(5.1)}{=} \omega \mapsto K\big((x,y) \mapsto k_{t',\omega}(x)\overline{k_{t',\omega}}(y)\mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(1_{[t',t+t']} \otimes (\alpha_x-\alpha_y)))(\omega)} f(x,y) \big) \\ &\!\!\!\overset{(5.4)}{=} \omega \mapsto K\big((x,y) \mapsto\mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}_{t'}(\alpha_x-\alpha_y))(\omega)}\mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(1_{[t',t+t']} \otimes (\alpha_x-\alpha_y)))(\omega)} f(x,y) \big)\\ &= \omega \mapsto K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}_{t+t'}(\alpha_x-\alpha_y))(\omega)} f(x,y) \big) \\ &\!\!\!\overset{(5.8)}{=} \mathcal{U}_{t+t'}(1 \otimes K_f) =U_{t+t'}(1 \otimes K_f). \end{align*} $$

Similarly for any $t \leqslant 0$ and any $t' \geqslant 0$ such that $t' \leqslant -t$ and any $\xi \in {\mathrm {L}}^2(X)$ , we have

$$ \begin{align*} U_{t'}U_t(1 \otimes K_f) =\mathcal{U}_{t'} S_{t'}\mathcal{U}_tS_t(1 \otimes K_f) \overset{(5.9)}{=} \mathcal{U}_{t'} S_{t'}\mathcal{U}_t (1 \otimes K_f) \\ &\!\!\!\overset{(5.7)}{=} \mathcal{U}_{t'} S_{t'} \big(\omega \mapsto K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(-1_{[t,0]} \otimes (\alpha_x-\alpha_y)))(\omega)} f(x,y)\big)\big)\\ &=\mathcal{U}_{t'} \big(\omega \mapsto K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(-1_{[t'+t,t']} \otimes (\alpha_x-\alpha_y)))(\omega)} f(x,y)\big)\big)\\ &\!\!\!\!\!\!\!\!\!\overset{(5.6)(5.5)}{=} \omega \mapsto \mathrm{Mult}_{k_{t',\omega}} K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(-1_{[t'+t,t']} \otimes (\alpha_x-\alpha_y)))(\omega)} f(x,y)\big) \\ &\!\!\!(\mathrm{Mult}_{k_{t',\omega}})^*\\ &\!\!\!\overset{(5.1)}{=} \omega \mapsto K\big((x,y) \mapsto k_{t',\omega}(x)\overline{k_{t',\omega}}(y)\mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(-1_{[t'+t,t']} \otimes (\alpha_x-\alpha_y)))(\omega)} f(x,y) \big) \\ &\!\!\!\overset{(5.4)}{=} \omega \mapsto K\big((x,y) \mapsto\mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(1_{[0,t']} \otimes (\alpha_x-\alpha_y)(\omega)}\mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(-1_{[t'+t,t']} \otimes (\alpha_x-\alpha_y)))(\omega)} \\ &\!\!\!f(x,y) \big)\\ &=\omega \mapsto K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}(-1_{[t'+t,0]} \otimes (\alpha_x-\alpha_y)(\omega)} f(x,y) \big) \quad \text{since } t+t' \leqslant 0 \leqslant t' \\ &\!\!\!\overset{(5.8)}{=} \mathcal{U}_{t+t'}(1 \otimes K_f) =U_{t+t'}(1 \otimes K_f). \end{align*} $$

The remaining cases are left to the reader. With these computations, it is easy to check that $(U_t)_{t \in \mathbb {R}}$ is a weak* continuous group of $*$ -automorphisms (consider the preadjoint maps of the $U_t$ ’s and use [Reference Engel and NagelEnN00, Lemme, B.15] to prove the weak* continuity of the group).

For any $f \in {\mathrm {L}}^2(X \times X)$ , any $\xi \in {\mathrm {L}}^2(X)$ , all $t \geqslant 0$ , and almost all y of X, we finally haveFootnote 8

$$ \begin{align*} \big(\big[\mathbb{E} U_tJ(K_f)\big](\xi)\big)(y) \overset{(5.3)}{=} \big(\big[\mathbb{E} U_t(1 \otimes K_f)\big](\xi)\big)(y) =\big(\big[\mathbb{E} \mathcal{U}_t S_t(1 \otimes K_f)\big](\xi)\big)(y) \\ &= \big(\big[\mathbb{E} \mathcal{U}_t(1 \otimes K_f)\big](\xi)\big)(y) =\bigg(\bigg[\int_{\Omega} \big(\mathcal{U}_t(1 \otimes K_f)\big)(\omega) \mathop{}\mathopen{}\mathrm{d} \mu(\omega)\bigg](\xi)\bigg)(y) \\ &=\int_{\Omega} \big(\big[ (\mathcal{U}_t(1 \otimes K_f))(\omega)\big](\xi)\big)(y) \mathop{}\mathopen{}\mathrm{d} \mu(\omega)\\ &\!\!\kern-2pt\overset{(5.8)}{=} \int_{\Omega}\bigg(\int_{X} \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}_t(\alpha_x-\alpha_y))(\omega)} f(x,y)\xi(x) \mathop{}\mathopen{}\mathrm{d} x\bigg) \mathop{}\mathopen{}\mathrm{d} \mu(\omega)\\ &=\int_{X} \bigg(\int_{\Omega} \mathrm{e}^{\sqrt{2}{\mathrm{i}} (\mathrm{W}_t(\alpha_x-\alpha_y))(\omega)} \mathop{}\mathopen{}\mathrm{d} \mu(\omega) \bigg) f(x,y)\xi(x) \mathop{}\mathopen{}\mathrm{d} x \\ &\!\!\!\!\kern-1pt\overset{(2.13)}{=} \int_{X} \mathrm{e}^{-t \| \alpha_x-\alpha_y \|_{H}^2} f(x,y)\xi(x) \mathop{}\mathopen{}\mathrm{d} x \\ &\!\!\kern-2pt\overset{(4.4)}{=} \int_{X} \phi_t(x,y)f(x,y)\xi(x) \mathop{}\mathopen{}\mathrm{d} x \overset{(2.15)}{=}\big(K_{\phi_t f}(\xi)\big)(y). \end{align*} $$

Hence, we have $\big (\mathbb {E} U_tJ(K_f)\big )(\xi )=\big (K_{\phi _t f}(\xi )$ and finally $\mathbb {E} U_tJ(K_f)=K_{\phi _t f}$ . By weak* density, we deduce the existence of $T_t$ and (5.2). For any $t,t' \geqslant 0$ and all $x,y \in X$ , we have

$$ \begin{align*}\mathrm{e}^{-t \| \alpha_x-\alpha_y \|_{H}^2}\mathrm{e}^{-t' \| \alpha_x-\alpha_y \|_{H}^2} =\mathrm{e}^{-(t+t') \| \alpha_x-\alpha_y \|_{H}^2}. \end{align*} $$

We deduce that $(T_t)_{t \geqslant 0}$ is a semigroup. Furthermore, the function (4.4) is real-valued. By Proposition 4.2, we conclude that each $T_t$ is self-adjoint. The factorization (5.2) shows that each $T_t$ is completely positive and unital. The weak* continuity is a consequence of the one of the group $(U_t)_{t \in \mathbb {R}}$ .

Remark 5.3 It would be interesting to give a direct proof of the first part, i.e., without the factorization (5.2).

6 Markov dilations of semigroups of measurable Schur multipliers

Our second main result is the following theorem, which gives a standard Markov dilation.

Theorem 6.1 Let X be a $\sigma $ -finite measure space. Let H be a real Hilbert space, and let $\alpha \colon X \to H$ be a strongly measurable map. Consider the weak* continuous semigroup $(T_t)_{t \geqslant 0}$ of self-adjoint unital completely positive Schur multipliers on $\mathrm {B}({\mathrm {L}}^2(X))$ defined in Theorem 5.2 by the symbols (4.4). There exist an injective von Neumann algebra N equipped with a normal semifinite faithful trace, an increasing filtration $(N_s)_{s \geqslant 0}$ of N with associated trace-preserving conditional normal faithful expectations $\mathbb {E}_s \colon N \to N_s$ , and trace-preserving unital normal injective $*$ -homomorphisms $\pi _s \colon \mathrm {B}({\mathrm {L}}^2(X)) \to N_s$ such that

(6.1) $$ \begin{align} \mathbb{E}_s\pi_t =\pi_sT_{t-s}, \quad 0 \leqslant s \leqslant t. \end{align} $$

Proof Let $\mathrm {W} \colon {\mathrm {L}}^2_{\mathbb {R}}(\mathbb {R}^+,H) \to {\mathrm {L}}^0(\Omega )$ be an H-cylindrical Brownian motion on a probability space $(\Omega ,\mu )$ (see Section 2). As in the proof of Theorem 5.2, for almost all $\omega \in \Omega $ and any $t \geqslant 0$ , we introduce the element (5.4) and for any $t \geqslant 0$ the element $V_t$ of ${\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)))$ defined by (5.5). Recall that $V_t$ is a unitary element of the von Neumann algebra $N \overset {\mathrm {def}}{=} {\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)))$ , which is clearly injective. Again with the $*$ -homomorphism J defined in (5.3) and the $*$ -automorphism $\mathcal {U}_t$ defined in (5.6), we define for any $t \geqslant 0$ the unital normal injective $*$ -homomorphism

(6.2) $$ \begin{align} \begin{array}{cccc} \pi_t\overset{\mathrm{def}}{=} \mathcal{U}_tJ \colon & \mathrm{B}({\mathrm{L}}^2(X)) & \longrightarrow & {\mathrm{L}}^\infty\big(\Omega,\mathrm{B}({\mathrm{L}}^2(X))\big) \\ & z & \longmapsto & V_t (1 \otimes z)V_t^* \\ \end{array}. \end{align} $$

It is obvious that $\pi _t$ is trace-preserving. For any $t \geqslant 0$ , we also define the canonical normal conditional expectations $\mathbb {E}_{\mathscr {F}_t} \colon {\mathrm {L}}^\infty (\Omega ) \to {\mathrm {L}}^\infty (\Omega ,\mathscr {F}_t)$ and $\mathbb {E}_t \overset {\mathrm {def}}{=}\mathbb {E}_{\mathscr {F}_t} \otimes \mathrm {Id}_{\mathrm {B}({\mathrm {L}}^2(X))} \colon {\mathrm {L}}^\infty (\Omega ,\mathrm {B}({\mathrm {L}}^2(X)) \to {\mathrm {L}}^\infty (\Omega ,\mathscr {F}_t,\mathrm {B}({\mathrm {L}}^2(X))$ where the $\sigma $ -algebra $\mathscr {F}_t$ is defined in (2.8). We introduce the von Neumann algebra $N_t \overset {\mathrm {def}}{=} {\mathrm {L}}^\infty (\Omega ,\mathscr {F}_t,\mathrm {B}({\mathrm {L}}^2(X))$ . Note that the range of $\pi _t$ is included in $N_t$ .

For almost all $\omega \in \Omega $ , the Hilbert–Schmidt operator

$$ \begin{align*}\big(\pi_t(K_f)\big)(\omega) \overset{(6.2)(5.3)}{=} \big(\mathcal{U}_t(1 \otimes K_f)\big)(\omega) \end{align*} $$

is associated by the computation (5.7) with the function

(6.3) $$ \begin{align} (x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}(\mathrm{W}_t(\alpha_x-\alpha_y))(\omega)} f(x,y). \end{align} $$

Similarly, for any $0 \leqslant s \leqslant t$ and almost all $\omega \in \Omega $ , the operator

$$ \begin{align*} \big(\pi_s(T_{t-s}(K_f))\big)(\omega) =\big(\pi_s(K_{\phi_{t-s}f}) \big)(\omega) \end{align*} $$

is Hilbert–Schmidt and associated with the function

(6.4) $$ \begin{align} (x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}(\mathrm{W}_s(\alpha_x-\alpha_y))(\omega)} \mathrm{e}^{-(t-s) \| \alpha_x-\alpha_y \|_{H}^2}f(x,y). \end{align} $$

Now, recall that $(\mathrm {W}_t(h))_{t \geqslant 0}$ is a Brownian motion for any fixed $h \in H$ . Hence, for any $0 \leqslant s \leqslant t$ and any $x,y \in X$ , the random variable

(6.5) $$ \begin{align} \mathrm{W}\big(1_{]s,t]} \otimes (\alpha_x-\alpha_y)\big) =\mathrm{W}\big(1_{[0,t]} \otimes (\alpha_x-\alpha_y)\big)-\mathrm{W}\big(1_{[0,s]} \otimes (\alpha_x-\alpha_y)\big) \\ &\overset{(2.7)}{=} \mathrm{W}_t(\alpha_x-\alpha_y)-\mathrm{W}_s(\alpha_x-\alpha_y) \nonumber \end{align} $$

is independent by (2.12) from the $\sigma $ -algebra $\mathscr {F}_s \overset {(2.8)}{=} \sigma \big (\mathrm {W}_r(h) : r \in [0,s], h \in H\big )$ . Consequently, the random variable $\mathrm {e}^{\sqrt {2}{\mathrm {i}}\mathrm {W}(1_{]s,t]}\otimes (\alpha _x-\alpha _y))}$ is also independent from the $\sigma $ -algebra $\mathscr {F}_s$ . Using [Reference Hytönen, van Neerven, Veraar and WeisHvNVW16, Proposition 2.6.31] in the fourth equality since each random variable $\mathrm {W}_s(\alpha _x)$ is $\mathscr {F}_s$ -measurable, we finally obtain, for any $0 \leqslant s \leqslant t$ and any $f \in {\mathrm {L}}^2(X \times X)$ ,

$$ \begin{align*} \mathbb{E}_s\pi_t(K_f) \overset{(6.3)}{=} \mathbb{E}_s\Big(\omega \mapsto \big[K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}} \mathrm{W}_t(\alpha_x-\alpha_y)(\omega)} f(x,y)\big)\big]\Big) \\ &\overset{(6.5)}{=} \mathbb{E}_s\Big(\omega \mapsto \big[K\big((x,y) \mapsto\mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}_s(\alpha_x-\alpha_y)(\omega)}\mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}(1_{]s,t]}\otimes (\alpha_x-\alpha_y))(\omega)}f(x,y)\big)\big]\Big)\\ &\overset{(5.1)}{=} \mathbb{E}_s\Big(\omega \mapsto \mathrm{Mult}_{k_{s,\omega}}\big[K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}(1_{]s,t]}\otimes (\alpha_x-\alpha_y))(\omega)}f(x,y)\big)\big]\mathrm{Mult}_{\overline{k_{s,\omega}}}\Big)\\ &=\big[\omega \mapsto \mathrm{Mult}_{k_{s,\omega}}\big] \mathbb{E}_s\big(\omega \mapsto \big[K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}(1_{]s,t]}\otimes (\alpha_x-\alpha_y))(\omega)}f(x,y)\big)\big) \\ &\big[\omega \mapsto \mathrm{Mult}_{\overline{k_{s,\omega}}} \big]\\ &\overset{(2.14)}{=}\big[\omega \mapsto \mathrm{Mult}_{k_{s,\omega}}\big] \mathbb{E}\big(\omega \mapsto \big[ K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}(1_{]s,t]}\otimes (\alpha_x-\alpha_y))(\omega)}f(x,y)\big)\big]\big) \\ &\big[\omega \mapsto \mathrm{Mult}_{\overline{k_{s,\omega}}} \big]. \end{align*} $$

Now, for any $\xi \in {\mathrm {L}}^2(X)$ and almost all $y \in Y$ , we see that

$$ \begin{align*} \mathbb{E}\big(\omega \mapsto \big[K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}(1_{]s,t]}\otimes (\alpha_x-\alpha_y))(\omega)}f(x,y)\big)\big]\big) (\xi)(y) \\ &=\bigg(\int_\Omega K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}(1_{]s,t]}\otimes (\alpha_x-\alpha_y))(\omega)}f(x,y)\big)\big) \mathop{}\mathopen{}\mathrm{d} \mu(\omega)\bigg) (\xi)(y) \\ &=\int_X \bigg(\int_\Omega \mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}(1_{]s,t]}\otimes (\alpha_x-\alpha_y))(\omega)}\mathop{}\mathopen{}\mathrm{d} \mu(\omega)\bigg)f(x,y)\xi(x) \mathop{}\mathopen{}\mathrm{d} x \\ &\!\!\!\!\kern-1pt\overset{(2.13)}{=} \int_X \mathrm{e}^{-(t-s) \| \alpha_x-\alpha_y \|_{H}^2}f(x,y)\xi(x) \mathop{}\mathopen{}\mathrm{d} x. \end{align*} $$

Consequently, the previous expectation is

$$ \begin{align*} \mathbb{E}\big(\omega \mapsto \big[K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}(1_{]s,t]}\otimes (\alpha_x-\alpha_y))(\omega)}f(x,y)\big)\big]\big) \\ &=K\big((x,y) \mapsto \mathrm{e}^{-(t-s) \| \alpha_x-\alpha_y \|_{H}^2}f(x,y)\big). \end{align*} $$

Finally, we obtain

$$ \begin{align*} \mathbb{E}_s\pi_t(K_f) =\big[\omega \mapsto \mathrm{Mult}_{k_{s,\omega}}\big] K\big((x,y) \mapsto \mathrm{e}^{-(t-s) \| \alpha_x-\alpha_y \|_{H}^2}f(x,y)\big) \big[\omega \mapsto \mathrm{Mult}_{\overline{k_{s,\omega}}} \big]\\ &\overset{(5.1)}{=} \omega \mapsto K\big((x,y) \mapsto \mathrm{e}^{\sqrt{2}{\mathrm{i}}\mathrm{W}_s(\alpha_x-\alpha_y)(\omega)}\mathrm{e}^{-(t-s) \| \alpha_x-\alpha_y \|_{H}^2}f(x,y)\big) \\ &\overset{(6.4)}{=} \pi_s(T_{t-s}(K_f)). \end{align*} $$

By weak* density, the proof is complete.

Similarly, we can prove the following reversed Markov dilation.

Theorem 6.2 Let X be a $\sigma $ -finite measure space. Let H be a real Hilbert space, and let $\alpha \colon X \to H$ be a strongly measurable map. Consider the weak* continuous semigroup $(T_t)_{t \geqslant 0}$ of self-adjoint unital completely positive Schur multipliers on $\mathrm {B}({\mathrm {L}}^2(X))$ defined in Theorem 5.2 by the symbols (4.4). There exist an injective von Neumann algebra N equipped with a normal semifinite faithful trace, a decreasing filtration $(N_s)_{s \geqslant 0}$ of N with associated trace-preserving normal faithful conditional expectations $\mathbb {E}_s \colon N \to N_s$ , and trace-preserving unital normal injective $*$ -homomorphisms $\pi _s \colon \mathrm {B}({\mathrm {L}}^2(X)) \to N_s$ such that

(6.6) $$ \begin{align} \mathbb{E}_s\pi_t =\pi_sT_{s-t}, \quad 0 \leqslant t \leqslant s. \end{align} $$

7 Applications to functional calculus and interpolation

We refer to [Reference Haagerup and MusatHaa06, Reference Hytönen, van Neerven, Veraar and WeisHvNVW18, Reference Junge and MeiJMX06, Reference Kunstmann and WeisKuW04, Reference Le MerdyLeM07] for background on sectoriality and ${\mathrm {H}}^\infty $ functional calculus. In the spirit of the result of [Reference ArhancetArh20, Theorem 4.1], we can prove the following theorem with a similar argument.

Theorem 7.1 Let X be a $\sigma $ -finite measure space. Let H be a real Hilbert space, and let $\alpha \colon X \to H$ be a strongly measurable map. Consider the weak* continuous semigroup $(T_t)_{t \geqslant 0}$ of self-adjoint unital completely positive Schur multipliers on $\mathrm {B}({\mathrm {L}}^2(X))$ defined in Theorem 5.2 by the symbols (4.4). Suppose $1<p<\infty $ . We let $-A_p$ be the generator of the induced strongly continuous semigroup $(T_{t,p})_{t \geqslant 0}$ on the Banach space $S^p_X$ . Then, for any $\theta>\pi |\frac {1}{p}-\frac {1}{2}|$ , the operator $A_p$ has a completely bounded ${\mathrm {H}}^{\infty }(\Sigma _\theta )$ functional calculus.

Let $\mathcal {T}=(T_t)_{t \geqslant 0}$ be a Markov semigroup on a von Neumann algebra $\mathcal {M}$ equipped with a normal semifinite faithful trace. We set $ \mathcal {M}_0 \overset {\mathrm {def}}{=} \left \{ x \in \mathcal {M} : \lim _{t \rightarrow \infty } T_t(x) = 0 \right \}$ where the limit is a weak* limit. For $1 \leqslant p < \infty $ , we let $ {\mathrm {L}}^p_0(\mathcal {M}) \overset {\mathrm {def}}{=} \{x \in {\mathrm {L}}^p(\mathcal {M}) : \lim _{t \to \infty } T_t(x)=0 \}$ where the limit is for the ${\mathrm {L}}^p$ -norm.

Recall that the column-BMO-seminorm is defined by

$$ \begin{align*}\| x \|_{{\mathrm{BMO}}^c_{\mathcal{T}}} \overset{\mathrm{def}}{=} \sup_{t \geqslant 0} \big\| T_t\left(|x - T_t(x)|^2 \right) \big\|_\infty^{\frac{1}{2}}, \quad x \in {\mathrm{L}}^2_0(\mathcal{M}). \end{align*} $$

Then we can define the row-BMO-seminorm $ \| x \|_{{\mathrm {BMO}}^r_{\mathcal {T}}} \overset {\mathrm {def}}{=} \| x^* \|_{{\mathrm {BMO}}^c_{\mathcal {T}}}$ and finally the BMO-seminorm

$$ \begin{align*}\| x \|_{{\mathrm{BMO}}_{\mathcal{T}}} \overset{\mathrm{def}}{=} \max\big\{ \| x \|_{{\mathrm{BMO}}^c_{\mathcal{T}}}, \| x \|_{{\mathrm{BMO}}^r_{\mathcal{T}}} \big\}. \end{align*} $$

In this context, we can introduce a Banach space ${\mathrm {BMO}}_{\mathcal {T}}$ and we refer to [Reference MeiMei08, Section 3] and [Reference CaspersCas19, Reference Junge and MeiJM12] for a precise definition. Combining Theorem 6.2 and [Reference Junge and MeiJM12, Theorem 5.12], we obtain the following interpolation formula by observing that each operator $T_t$ of the semigroup is trace-preserving by the factorization of Theorem 5.2.

Theorem 7.2 Let X be a $\sigma $ -finite measure space. Let H be a real Hilbert space, and let $\alpha \colon X \to H$ be a strongly measurable map. Consider the weak* continuous semigroup $(T_t)_{t \geqslant 0}$ of self-adjoint unital completely positive Schur multipliers on $\mathrm {B}({\mathrm {L}}^2(X))$ defined in Theorem 5.2 by the symbols (4.4). Suppose $1\leqslant p<\infty $ . Then

$$ \begin{align*}\big({\mathrm{BMO}}_{\mathcal{T}},S^1_{X,0}\big)_{\frac{1}{p}} \approx S^p_{X,0} \end{align*} $$

with equivalence of norms up to a constant $\approx p$ .

Remark 7.3 We can replace the space ${\mathrm {BMO}}_{\mathcal {T}}$ by other ${\mathrm {BMO}}$ -spaces (see [Reference Junge and MeiJM12, Theorem 5.12]).

Finally, we could give applications in ergodic theory in the spirit of the paper [Reference Jiao and WangJiW17].

Acknowledgment

I would like to thank Julio Delgado for an interesting discussion. I am grateful to Lyudmyla Turowska for short discussions and the reference [Reference TodorovSTT14] and Hermann König for a very instructive discussion on [Reference KönigKon86, Theorem 3.a.1, p. 145]. Finally, I would also like to thank the referees for corrections.

Footnotes

The author acknowledges support by the grant ANR-18-CE40-0021 (project HASCON) of the French National Research Agency ANR.

1 The authors prove that there exists no “reasonable” analog of a variant of Fendler’s result for a discrete semigroup $(T^k)_{k \geqslant 0}$ of completely positive contractions.

2 That means that for any $f \in {\mathrm {L}}^\infty (\Omega )$ , we have $\int _{\Omega } \Gamma _\infty (u)f \mathop {}\mathopen {}\mathrm {d}\mu =\int _{\Omega } f \mathop {}\mathopen {}\mathrm {d}\mu $ .

3 The existence of a proof of Lemma 2.1 without (2.4) is unclear.

4 It was confirmed by email by the author of the book.

5 The “only if” part is true without the assumption “with support X.”

6 If (2.19) is satisfied on a topological space X, then $\varphi $ is continuous if and only if $\alpha $ is continuous.

7 We have $1_{[0,t]} \xrightarrow [t \to t_0]{} 1_{[0,t_0]}$ and $1_{[t,0]} \xrightarrow [t \to t_0]{} 1_{[t_0,0]}$ in ${\mathrm {L}}^2_{\mathbb {R}}(\mathbb {R})$ .

8 Note that the function $f(\cdot ,y)$ belongs to ${\mathrm {L}}^2(X)$ for almost all $y \in X$ .

References

Arhancet, C., On Matsaev’s conjecture for contractions on noncommutative ${L}^p$ -spaces . J. Operator Theory 69(2013), no. 2, 387421.10.7900/jot.2010dec29.1905CrossRefGoogle Scholar
Arhancet, C., Analytic semigroups on vector valued noncommutative ${L}^p$ -spaces . Stud. Math. 216(2013), no. 3, 271290.CrossRefGoogle Scholar
Arhancet, C., Dilations of semigroups on von Neumann algebras and noncommutative ${L}^p$ -spaces . J. Funct. Anal. 276(2019), no. 7, 22792314.CrossRefGoogle Scholar
Arhancet, C., Dilations of Markovian semigroups of Fourier multipliers on locally compact groups . Proc. Amer. Math. Soc. 148(2020), no. 6, 25512563.10.1090/proc/14938CrossRefGoogle Scholar
Arhancet, C., Fackler, S., and Le Merdy, C., Isometric dilations and ${H}^{\infty }$ calculus for bounded analytic semigroups and Ritt operators . Trans. Amer. Math. Soc. 369(2017), 68996933.CrossRefGoogle Scholar
Arhancet, C. and Kriegler, C., Riesz transforms, Hodge–Dirac operators and functional calculus for multipliers, Lecture Notes in Mathematics, 2304, Springer, Cham, 2022.10.1007/978-3-030-99011-4CrossRefGoogle Scholar
Arhancet, C. and Kriegler, C.. Fourier–Stieltjes algebras, decomposable Fourier multipliers and amenability. Preprint, 2022. arXiv:2205.13823 Google Scholar
Arhancet, C. and Le Merdy, C., Dilation of Ritt operators on ${L}^p$ -spaces . Israel J. Math. 201(2014), no. 1, 373414.CrossRefGoogle Scholar
Bekka, B., de la Pierre, P., and Valette, A., Kazhdan’s property (T), New Mathematical Monographs, 11, Cambridge University Press, Cambridge, 2008.CrossRefGoogle Scholar
Berg, C., Christensen, J., and Ressel, P., Harmonic analysis on semigroups. Theory of positive definite and related functions, Springer, New York, 1984.10.1007/978-1-4612-1128-0CrossRefGoogle Scholar
Blecher, D. and Le Merdy, C., Operator algebras and their modules—an operator space approach, London Mathematical Society Monographs. New Series, 30, Oxford Science Publications–Clarendon Press–Oxford University Press, Oxford, 2004.CrossRefGoogle Scholar
Bożejko, M., Positive and negative definite kernels on discrete groups. Lecture notes, Heidelberg University, 1987.Google Scholar
Cartier, P., Une étude des covariances mesurables (in French) [A study of measurable covariances]. In: Mathematical analysis and applications, part A, Adv. in Math. Suppl. Stud., 7a, Academic Press, New York–London, 1981, pp. 267316.Google Scholar
Caspers, M., Harmonic analysis and BMO-spaces of free Araki-woods factors . Stud. Math. 246(2019), no. 1, 71107.CrossRefGoogle Scholar
Caspers, M., Junge, M., Sukochev, F. and Zanin, D., BMO-estimates for non-commutative vector valued Lipschitz functions . J. Funct. Anal. 278(2020), no. 3, Article no. 108317, 39 pp.CrossRefGoogle Scholar
Duquet, C., Dilation properties of measurable Schur multipliers and Fourier multipliers. Preprint, 2022. arXiv:2202.06575 CrossRefGoogle Scholar
Engel, K.-J. and Nagel, R., One-parameter semigroups for linear evolution equations, Graduate Texts in Mathematics, 194, Springer, New York, 2000.Google Scholar
Fendler, G., Dilations of one parameter semigroups of positive contractions on ${L}^p$ spaces . Can. J. Math. 49(1997), no. 4, 736748.CrossRefGoogle Scholar
Fendler, G., On dilations and transference for continuous one-parameter semigroups of positive contractions on ${\mathbf{\mathcal{L}}}^p$ -spaces . Ann. Univ. Sarav. Ser. Math. 9(1998), no. 1, 108317.Google Scholar
Ferguson, T., Mei, T., and Simanek, B., ${H}^{\infty }$ -calculus for semigroup generators on BMO . Adv. Math. 347(2019), 408441.10.1016/j.aim.2019.02.027CrossRefGoogle Scholar
Haagerup, U. and Musat, M., Factorization and dilation problems for completely positive maps on von Neumann algebras . Commun. Math. Phys. 303(2011), no. 2, 555594.CrossRefGoogle Scholar
Haase, M., The functional calculus for sectorial operators, Operator Theory: Advances and Applications, 169, Birkhäuser, Basel, 2006.CrossRefGoogle Scholar
Husain, T. and Warsi, S. A., Positive definite functions for the class ${L}_p(G)$ . Canadian J. Math. 27(1975), no. 5, 11491156.CrossRefGoogle Scholar
Hytönen, T., van Neerven, J., Veraar, M., and Weis, L., Analysis in Banach spaces: martingales and Littlewood–Paley theory. Vol. I, Springer, Cham, 2016.10.1007/978-3-319-48520-1CrossRefGoogle Scholar
Hytönen, T., van Neerven, J., Veraar, M., and Weis, L., Analysis in Banach spaces: probabilistic methods and operator theory. Vol. II, Springer, Cham, 2018.Google Scholar
Janson, S., Gaussian Hilbert spaces, Cambridge Tracts in Mathematics, 129, Cambridge University Press, Cambridge, 1997.CrossRefGoogle Scholar
Jiao, Y. and Wang, M., Noncommutative harmonic analysis on semigroups . Indiana Univ. Math. J. 66(2017), no. 2, 401417.CrossRefGoogle Scholar
Junge, M. and Le Merdy, C., Dilations and rigid factorisations on noncommutative ${L}^p$ -spaces . J. Funct. Anal. 249(2007), 220252.CrossRefGoogle Scholar
Junge, M., Le Merdy, C., and Xu, Q., ${H}^{\infty }$ functional calculus and square functions on noncommutative ${L}^p$ -spaces . Astérisque 305(2006), vi+138 pp.Google Scholar
Junge, M. and Mei, T., Noncommutative Riesz transforms—a probabilistic approach . Amer. J. Math. 132(2010), no. 3, 611680.CrossRefGoogle Scholar
Junge, M. and Mei, T., BMO spaces associated with semigroups of operators . Math. Ann. 352(2012), no. 3, 691743.10.1007/s00208-011-0657-0CrossRefGoogle Scholar
Junge, M. and Zeng, Q., Noncommutative martingale deviation and Poincaré type inequalities with applications . Probab. Theory Relat. Fields 161(2015), nos. 3–4, 449507.CrossRefGoogle Scholar
König, H., Eigenvalue distribution of compact operators, Operator Theory: Advances and Applications, 16, Birkhäuser, Basel, 1986.CrossRefGoogle Scholar
Kunstmann, P. C. and Weis, L., Maximal   $L_p$ -regularity for parabolic equations, Fourier multiplier theorems and $H^{\infty}$ -functional calculus. In: Functional analytic methods for evolution equations, Lecture Notes in Mathematics, 1855, Springer, Berlin–Heidelberg, 2004, pp. 65311.CrossRefGoogle Scholar
Le Merdy, C., Square functions, bounded analytic semigroups, and applications . In: Perspectives in operator theory, Banach Center Publications, 75, Polish Acad. Sci. Inst. Math., Warsaw, 2007, pp. 191220.CrossRefGoogle Scholar
Li, B.-R., Introduction to operator algebras, World Scientific, Singapore, 1992.Google Scholar
Megginson, R. E., An introduction to Banach space theory, Graduate Texts in Mathematics, 183, Springer, New York, 1998.10.1007/978-1-4612-0603-3CrossRefGoogle Scholar
Mei, T., Tent spaces associated with semigroups of operators . J. Funct. Anal. 255(2008), no. 12, 33563406.CrossRefGoogle Scholar
Mercer, J., Linear transformations and functions of positive type . Proc. R. Soc. Lond. A 99(1921), no. 696, 1938.Google Scholar
Nualart, D., The Malliavin calculus and related topics. 2nd ed., Springer, Berlin, 2006.Google Scholar
Palmer, T. W., Banach algebras and the general theory of *-algebras . Vol. 2, Encyclopedia of Mathematics and Its Applications, 79, Cambridge University Press, Cambridge, 2001.10.1017/CBO9780511574757CrossRefGoogle Scholar
Pisier, G., Non-commutative vector valued ${L}_p$ -spaces and completely $p$ -summing maps . Astérisque, 247(1998), vi+131 pp.Google Scholar
Ricard, É., A Markov dilation for self-adjoint Schur multipliers . Proc. Amer. Math. Soc. 136(2008), no. 12, 43654372.CrossRefGoogle Scholar
Sakai, S., C*-algebras and W*-algebras, Springer, New York, 1971.Google Scholar
Spronk, N., Measurable Schur multipliers and completely bounded multipliers of the Fourier algebras . Proc. Lond. Math. Soc. 89(2004), no. 1, 161192.10.1112/S0024611504014650CrossRefGoogle Scholar
Steen, N. M., Todorov, I. G., and Turowska, L., Local operator multipliers and positivity . J. Funct. Anal. 267(2014), no. 1, 80111.10.1016/j.jfa.2014.04.007CrossRefGoogle Scholar
Stratila, S., Modular theory in operator algebras, Editura Academiei Republicii Socialiste România, Bucharest; Abacus Press, Tunbridge Wells, 1981, translated from the Romanian by the author.Google Scholar
Todorov, I. G., Herz–Schur multipliers. Lectures at “Banach and Operator Algebras over Groups,” 2014.Google Scholar
Todorov, I. G., Interactions between harmonic analysis and operator theory . Serdica Math. J. 41(2015), no. 1, 1334.Google Scholar
Todorov, I. G. and Turowska, L., Schur and operator multipliers . In: Banach algebras, Banach Center Publications, 91, Polish Acad. Sci. Inst. Math., Warsaw, 2009 (2010), pp. 385410.Google Scholar
van Neerven, J., Stochastic evolution equations. ISEM lecture notes, 2007/08.Google Scholar
van Neerven, J., Veraar, M., and Weis, L., Stochastic integration in Banach spaces—a survey . In: Stochastic analysis: a series of lectures, Progress in Probability, 68, Birkhäuser, Basel, 2015, pp. 297332.10.1007/978-3-0348-0909-2_11CrossRefGoogle Scholar