Hostname: page-component-848d4c4894-wg55d Total loading time: 0 Render date: 2024-04-30T13:28:07.978Z Has data issue: false hasContentIssue false

Corrigendum: Von Neumann algebras and extensions of inverse semigroups

Published online by Cambridge University Press:  22 August 2023

Allan P. Donsig
Affiliation:
Department of Mathematics, University of Nebraska-Lincoln, Lincoln, NE, 68588-0130 USA (adonsig@unl.edu; dpitts2@unl.edu)
Adam H. Fuller
Affiliation:
Department of Mathematics, Ohio University, Athens, OH, 45701, Greece USA (fullera@ohio.edu)
David R. Pitts
Affiliation:
Department of Mathematics, University of Nebraska-Lincoln, Lincoln, NE, 68588-0130 USA (adonsig@unl.edu; dpitts2@unl.edu)
Rights & Permissions [Opens in a new window]

Abstract

Type
Corrigendum
Copyright
© The Author(s), 2023. Published by Cambridge University Press on Behalf of The Edinburgh Mathematical Society.

Definition 2.1 of [Reference Donsig, Fuller and Pitts1] is incomplete and should be replaced by the following.

Corrected Definition 2.1. Let ${\mathcal{L}}$ be a Boolean algebra. A representation of ${\mathcal{L}}$ is a map $\pi:{\mathcal{L}}\rightarrow \text{proj}({\mathcal{B}})$ of ${\mathcal{L}}$ into the projection lattice of a unital $C^*$-algebra ${\mathcal{B}}$ such that for every $s,t\in{\mathcal{L}}$,

\begin{equation*}\pi(s\wedge t)=\pi(s)\pi(t) \quad\text{and}\quad \pi(\neg s)=I-\pi(s).\end{equation*}

With this corrected definition, the statement and proof of [Reference Donsig, Fuller and Pitts1, Proposition 2.2] are unchanged with the exception that the $C^*$-algebra ${\mathcal{B}}$ is to be assumed unital:

Corrected Proposition 2.2. Let ${\mathcal{L}}$ be a Boolean algebra with character space $\widehat{{\mathcal{L}}}$. For each $s\in{\mathcal{L}}$, let $\widehat{s}\in C(\widehat{{\mathcal{L}}})$ be the Gelfand transform, $\widehat{s}(\rho)=\rho(s)$. Then $C(\widehat{{\mathcal{L}}})$ has the following universal property: if ${\mathcal{B}}$ is a unital $C^*$-algebra and $\theta:{\mathcal{L}}\rightarrow {\mathcal{B}}$ is a representation such that $\theta({\mathcal{L}})$ generates ${\mathcal{B}}$ as a $C^*$-algebra, then there exists a unique $*$-epimorphism $\alpha:C(\widehat{{\mathcal{L}}})\rightarrow {\mathcal{B}}$ such that for every $s\in{\mathcal{L}}$,

\begin{equation*}\theta(s)=\alpha(\widehat{s}).\end{equation*}

The discussion on [Reference Donsig, Fuller and Pitts1, Page 63] uses [Reference Donsig, Fuller and Pitts1, Proposition 2.2] in preparation for [Reference Donsig, Fuller and Pitts1, Definition 2.9]. This discussion is unaffected by the corrections given above.

Unfortunately, the incomplete definition of representation given in [Reference Donsig, Fuller and Pitts1, Definition 2.1] leads to a gap in the work presented in [Reference Donsig, Fuller and Pitts1]. To describe the issue, given a Cartan inverse monoid ${\mathcal{S}}$, [Reference Donsig, Fuller and Pitts1, Section 4.2] defines ${\mathcal{D}}:=C(\widehat{{\mathcal{E}}({\mathcal{S}})})$, constructs a ${\mathcal{D}}$-valued reproducing kernel $K:{\mathcal{S}}\times{\mathcal{S}}\rightarrow {\mathcal{D}}$ ([Reference Donsig, Fuller and Pitts1, Defintion 4.7]), and constructs a right Hilbert ${\mathcal{D}}$-module ${\mathfrak{A}}$ ([Reference Donsig, Fuller and Pitts1, Proposition 4.12]). Further [Reference Donsig, Fuller and Pitts1, Theorem 4.16] shows there is a map $\lambda:{\mathcal{G}}\rightarrow {\mathcal{L}}({\mathfrak{A}})$ so that for $v\in {\mathcal{G}}$ and $s\in {\mathcal{S}}$,

\begin{equation*}\lambda(v)k_s=k_{q(v)s}\sigma(v,s),\end{equation*}

where (as in [Reference Donsig, Fuller and Pitts1, Definition 4.13]),

\begin{equation*} \sigma(v,s)=j(q(v)s)^\dagger vj(s)=j(s^\dagger q(v^\dagger))vj(s) \end{equation*}

and $j:{\mathcal{S}}\rightarrow {\mathcal{G}}$ is an order-preserving section (see [Reference Donsig, Fuller and Pitts1, Definition 4.1 and Proposition 4.6]). Also, the proof of [Reference Donsig, Fuller and Pitts1, Theorem 4.16] establishes:

  1. (a) λ is one-to-one;

  2. (b) for $v, w\in {\mathcal{G}}$,

    \begin{equation*}\lambda(v)^*=\lambda(v^\dagger)\quad\text{and}\quad \lambda(vw)=\lambda(v)\lambda(w).\end{equation*}

To this point, all is well. However, on page 84 of [Reference Donsig, Fuller and Pitts1] we wrote,

“Let

\begin{equation*} {\mathfrak{B}}=\overline{\operatorname{span}} \{k_e \colon e\in{\mathcal{E}}({\mathcal{S}})\}\subseteq{\mathfrak{A}}. \end{equation*}

Note that ${\mathfrak{B}}$ is a right Hilbert ${\mathcal{D}}$-submodule of ${\mathfrak{A}}$. Proposition 2.2 shows that $\lambda|_{{\mathcal{E}}({\mathcal{G}})}$ extends to a $*$-monomorphism $\alpha_\ell{}:{\mathcal{D}}\rightarrow {\mathcal{L}}({\mathfrak{A}})$.”

There is a gap here, because we did not establish that $\lambda|_{{\mathcal{E}}({\mathcal{G}})}:{\mathcal{E}}({\mathcal{G}})\rightarrow {\mathcal{D}}$ is a representation in the sense of Corrected Definition 2.1 given above, and therefore, we have not shown that we can apply Corrected Proposition 2.2. To fill this gap, we must establish the following fact.

Lemma 1. For $e\in{\mathcal{E}}({\mathcal{G}})$,

(2)\begin{equation} \lambda(\neg e)=I-\lambda(e). \end{equation}

Before verifying this, we make some preliminary remarks. Observe that since

is an idempotent separating extension, ${\mathcal{E}}({\mathcal{G}})=\iota({\mathcal{E}}({\mathcal{P}})).$ Therefore, we may identify ${\mathcal{E}}({\mathcal{G}})$ with ${\mathcal{E}}({\mathcal{P}})$ via the map ι.

Since $q\circ \iota=\pi$, we find $q|_{{\mathcal{E}}({\mathcal{G}})}:{\mathcal{E}}({\mathcal{G}})\rightarrow {\mathcal{E}}({\mathcal{S}})$ is an isomorphism. As $q\circ j={\operatorname{id}}|_{\mathcal{S}}$,

\begin{equation*}j|_{{\mathcal{E}}({\mathcal{S}})}=\left(q|_{{\mathcal{E}}({\mathcal{G}})}\right)^{-1}=\left(\pi|_{{\mathcal{E}}({\mathcal{P}})}\right)^{-1}.\end{equation*}

Hence for $e\in{\mathcal{E}}({\mathcal{G}})$ and $s\in {\mathcal{S}}$, [Reference Donsig, Fuller and Pitts1, Lemma 4.2] gives

\begin{equation*}\sigma(e,s)=j(s^\dagger q(e))ej(s)= j(s^\dagger)j(q(e))ej(s)=j(s^\dagger)ej(s)=j(s^\dagger q(e)s).\end{equation*}

Thus, for $e\in {\mathcal{E}}({\mathcal{G}})$ and $s\in{\mathcal{S}}$,

\begin{equation*}\lambda(e)k_s=k_{q(e)s}j(s^\dagger q(e)s)\overset{\text{[1, Cor. 4.9]}}=k_{q(e)s(s^\dagger q(e)s)}=k_{q(e)s}.\end{equation*}

By [Reference Donsig, Fuller and Pitts1, Proposition 4.12], $\{k_s: s\in{\mathcal{S}}\}$ has dense span in ${\mathfrak{A}}$. Thus to establish (2), it suffices to show that for every $s\in {\mathcal{S}}$ and $e\in {\mathcal{E}}({\mathcal{S}})$,

(3)\begin{equation} k_{es}+k_{(\neg e) s}=k_s. \end{equation}

To establish (3), we require the following.

Fact 4. Let ${\mathcal{S}}$ be a Boolean inverse monoid. For $r, s, t\in {\mathcal{S}}$ with s and t orthogonal,

(5)\begin{equation} (s\vee t)\wedge r=(s\wedge r)\vee (t\wedge r). \end{equation}

Proof. Note that $s\wedge r\leq (s\vee t)\wedge r$ because $s\leq s\vee t$; a similar inequality holds for t. Therefore,

(6)\begin{equation}(s\wedge r)\vee (t\wedge r)\leq (s\vee t)\wedge r. \end{equation}

Since $(s\vee t)\wedge r\leq s\vee t$, there is $f\in {\mathcal{E}}({\mathcal{S}})$ such that $(s\vee t)\wedge r=(s\vee t)f$. Now multiplication distributes over finite orthogonal joins in a Boolean inverse monoid [Reference Lawson2, p. 386], so

\begin{equation*}(s\vee t)\wedge r=sf\vee tf.\end{equation*}

As the left side is a meet and the right a join, $r \ge sf$ and $r \ge tf$, and so

\begin{equation*}sf\leq s\wedge r\quad\text{and}\quad tf\leq t\wedge r.\end{equation*}

Hence

(7)\begin{equation}(s\vee t)\wedge r=sf\vee tf\leq (s\wedge r)\vee (t\wedge r). \end{equation}

Combining (6) and (7) gives (5).

We now complete the proof of Lemma 1. As noted, it suffices to establish (3), and this is what we shall do.

Fix $s\in{\mathcal{S}}$ and $e\in{\mathcal{E}}({\mathcal{S}})$. Again using the fact that multiplication distributes over orthogonal joins in a Boolean inverse monoid, we see that for $t\in {\mathcal{S}}$,

\begin{equation*}s^\dagger t=(s^\dagger et)\vee (s^\dagger (\neg e) t).\end{equation*}

Fact 4 gives

\begin{equation*}((s^\dagger et)\wedge 1) \vee (s^\dagger (\neg e)t) \wedge 1)=((s^\dagger et) \vee (s^\dagger(\neg e)t))\wedge 1=s^\dagger t\wedge 1.\end{equation*}

Notice that $(s^\dagger et)\wedge 1$ and $(s^\dagger (\neg e)t) \wedge 1$ are orthogonal idempotents in ${\mathcal{E}}({\mathcal{S}})$. Since $j|_{{\mathcal{E}}({\mathcal{S}})}$ is a Boolean algebra isomorphism onto ${\mathcal{E}}({\mathcal{P}})=\operatorname{proj}({\mathcal{D}})$, we get

\begin{equation*}j(s^\dagger t\wedge 1)=j((s^\dagger et)\wedge 1)\vee j((s^\dagger(\neg e)t)\wedge1)=j((s^\dagger et)\wedge 1) + j((s^\dagger(\neg e)t)\wedge1).\end{equation*}

In other words, this shows that for $t\in {\mathcal{S}}$,

\begin{equation*}k_s(t)=k_{es}(t)+k_{(\neg e)s}(t),\end{equation*}

which is (3).

References

Donsig, A. P., Fuller, A. H. and Pitts, D. R., Von Neumann algebras and extensions of inverse semigroups, Proc. Edinb. Math. Soc. (2) 60(1): (2016), 5797.CrossRefGoogle Scholar
Lawson, M. V., A noncommutative generalization of Stone duality, J. Aust. Math. Soc. 88(3): (2010), 385404.CrossRefGoogle Scholar