3. Consequences

In this section we will recover several known results, some in greater generality than in their original statements, dealing with different algebraic structures on classes of continuous functions and their isomorphisms.

The general procedure we will use is the following: Suppose that $X$ and $Y$ are locally compact Hausdorff spaces, $H_X$ and $H_Y$ are Hausdorff spaces, $\theta_X\in C(X,H_X)$, $\theta_Y\in C(Y,H_Y)$ and $\mathcal{A}(X)$ and $\mathcal{A}(Y)$ regular subsets of $C_c(X,\theta_X)$ and $C_c(Y,\theta_Y)$, respectively.

  1. Describe the relation $\perpp$ with the algebraic structure at hand. In general, we will first describe $\perp$ and use it in order to describe $\perpp$.

  2. Given an algebraic isomorphism $T\colon\mathcal{A}(X)\to\mathcal{A}(Y)$ for appropriate classes of functions $\mathcal{A}(X)$ and $\mathcal{A}(Y)$, the first item ensures that $T$ is a $\perpp$-isomorphism, so let $\phi\colon Y\to X$ be the $T$-homeomorphism.

At this point, we have proved that the algebraic isomorphism $T\colon\mathcal{A}(X)\to\mathcal{A}(Y)$ determines an isomorphism $\phi\colon Y\to X$. Since we moreover would like to describe $T$ in terms of $\phi$, we proceed as follows using the theory of Section 2:

  1. Prove that $T$ is $\phi$-basic, as in Definition 2.2. Let $\chi$ be the $(\phi,T)$-transform.

  2. Items 1.-3. also apply to $T^{-1}$, which is thus also a basic $\perpp$-isomorphism.

  3. Items 3. and 4. imply, by Proposition 2.6(a) and (b), that the sections $\chi(y,\cdot)$ are injective and in the case that $\mathcal{A}(Y)$ is regular, item (d) implies that $\chi(y,\cdot)$ is also surjective.

  4. If we have a classification of algebraic isomorphisms between $H_X$ and $H_Y$, this classification will apply to each section $\chi(y,\cdot)$ of the $(\phi,T)$-transform, by Proposition 2.7. This will in turn describe $T$ completely.

3.1. Milgram's Theorem

In this subsection, we will always assume that $X$ (and similarly $Y$) is a locally compact Hausdorff space. We first generalize Milgram's theorem (and by consequence, Gelfand‒Kolmogorov and Gelfand‒Naimark) as follows: Let $S_X$ be a non-trivial path-connected Hausdorff topological monoid with zero $0$ and unit $1$, and which is both:

Common examples of such semigroups are the following, under usual product: $\mathbb{R}$, $\mathbb{C}$, $[-1,1]$, the closed complex unit disc, $\operatorname{Gl}(n,\mathbb{R})\cup\left\{0\right\}$, and other variations.

We consider $\theta_X=0$, the zero map from $X$ to $S_X$. Let us also write $C_c(X,S_X)$ for $C_c(X,0)$ (to make the counter-domain explicit). Consider $Y$, $\theta_Y$ and $S_Y$ similarly. Urysohn's Lemma and path-connectedness of $S_X$ implies that $C_c(X,S_X)$ is regular.

Following the procedure in the beginning of this section, we first describe $\perpp$ in multiplicative terms:

Lemma 3.1.

If $f,g\in C_c(X,S_X)$, then

\begin{equation*} f\perpp g\iff \exists h\in C_c(X,S_X)\text{ such that }hf=f\text{ and }hg=0.\tag{3.1} \end{equation*}

Proof.

The condition in the right-hand side of (3.1) states that $h=1$ on $\supp(f)$ and $h\perp g$, so the implication $\Leftarrow$ is immediate. The implication $\Rightarrow$ (i.e., the existence of such $h$ provided $f\perpp g$) follows from Urysohn's Lemma and path-connectedness of $S_X$.

As a consequence, any multiplicative isomorphism $C_c(X,S_X)\to C_c(Y,S_Y)$ is a $\perpp$-isomorphism.

Corollary 3.2.

If $T\colon C(X,S_X)\to C_c(Y,S_Y)$ is a multiplicative isomorphism, then $X$ and $Y$ are homeomorphic.

To recover Milgram's original theorem in full generality, we restrict now to the case $S_X=S_Y=\mathbb{R}$ in order to obtain an explicit description of $T$ as above.

First, we will need to recall the classification of multiplicative isomorphisms of $\mathbb{R}$. The general case may be found in [MR2467621, Theorem 3.1.3], but for our purposes it will suffice to consider only the continuous isomorphisms (which were already described in [MR0029476, Lemma 4.3], for example).

Proposition 3.3.

Let $\tau\colon\mathbb{R}\to\mathbb{R}$ be a multiplicative isomorphism. Then $\tau$ is continuous if and only if $0\smallerthan x\smallerthan 1$ implies $0\smallerthan \tau(x)\smallerthan 1$. In this case, $\tau$ has the form $\tau(x)=\operatorname{sgn}(x)|x|^p$ for some $p\greaterthan 0$.

We will now classify multiplicative isomorphisms from $C_c(X,\mathbb{R})$ and $C_c(Y,\mathbb{R})$. Note that the following theorem is a generalization of [MR0029476, Theorem A] to the locally compact setting.

Theorem 3.4 (Milgram's Theorem [MR0029476, Theorem A], for locally compact spaces).

Let $X$ and $Y$ be locally compact Hausdorff spaces and let $T\colon C_c(X,\mathbb{R})\to C_c(Y,\mathbb{R})$ be a multiplicative isomorphism. Then there exists a homeomorphism $\phi\colon Y\to X$, a closed, discrete and isolated subset $F\subseteq Y$, and a continuous positive function $p\colon Y\setminus F\to(0,\infty)$ satisfying

\begin{equation*} Tf(y)=\operatorname{sgn}(f(\phi(y)))|f(\phi (y))|^{p(y)} \end{equation*}
for all $f\in C_c(X,\mathbb{R})$ and $y\in Y\setminus F$.

Proof.

By Lemma 3.1, $T$ is a $\perpp$-isomorphism, so let $\phi$ be the $T$-homeomorphism. We prove that $T$ is $\phi$-basic in a few simple steps.

  1. If $f,h\in C_c(X,\mathbb{R})$, then $f=1$ on $\supp(h)$ (i.e., $fh=h$) if and only if $Tf=1$ on $\supp(Th)$ (i.e., $(Tf)(Th)=(Th)$). This implies that

  2. If $f,h\in C_c(X,\mathbb{R})$, then $f\neq 0$ on $\supp(h)$ if and only if $Tf\neq 0$ on $\supp(Th)$.

    Indeed, if $f\neq 0$ on $\supp(h)$, we can find $g\in C_c(X,\mathbb{R})$ such that $g=1/f$ on $\supp(h)$. Item 1. implies that $TfTg=1$ on $\supp(Th)$, and in particular $Tf\neq 0$ on $\supp(Th)$. As the sets of the form $\sigma(h)$ form a basis of $X$, we conclude that

  3. If $f\in C_c(X,\mathbb{R})$ and $y\in Y$, then $f(\phi(y))\neq 0$ if and only if $Tf(y)\neq 0$.

    A similar argument to that of item 2. also proves that

  4. If $f,g,h\in C_c(X,\mathbb{R})$, then $f$ and $g$ coincide and are nonzero on $\supp(h)$ if and only if $Tf$ and $Tg$ coincide and are nonzero on $\supp(h)$.

  5. In particular, from 3., $f(\phi(y))=0$ if and only if $Tf(y)=0$.

  6. If $f\in C_c(X,\mathbb{R})$ and $y\in Y$, then $f(\phi(y))=1$ if and only if $Tf(y)=1$:

    Suppose this was not the case, say $f(\phi(y))=1$ but $Tf(y)\neq 1$. From item 1. above we know that $y$ is not isolated. If necessary, as $T$ is a $\perp$-isomorphism, Theorem 1.16 allows us to change $f$ by a function which equals $1/f$ on some neighbourhood of $\phi(y)$ and assume that $Tf(y)\greaterthan 1$. The rest of the argument is similar to that of [MR0029476, Lemma 4.1]: We first take a sequence of distinct points $y_n$, all contained in some compact neighbourhood of $y$, satisfying

    • $f(\phi(y_n))^n\to 1=f(\phi(y))$; and

    • $Tf(y_n)\to Tf(y)$, so in particular $Tf(y_n)^n\to\infty$.

    Using Propositions 2.9 and 2.10(b), we may find $g\in C_c(X,\mathbb{R})$ which coincides with $f^n$ on some neighbourhood of $\phi(y_n)$. But then $Tg$ coincides with $(Tf)^n$ on a neighbourhood $y_n$ (because $T$ is a $\perp$-isomorphism), contradicting the fact that $Tg$ is bounded.

We conclude, from 5. and 6., that $T$ is $\phi$-basic. Let $\chi\colon Y\times\mathbb{R}\to\mathbb{R}$ be the $T$-transform.

Let $F=\left\{y\in Y:\chi(y,\cdot)\text{ is discontinuous}\right\}$. Let us prove that $F$ is closed and discrete, or equivalently that $F\cap K$ is finite for all compact $K\subseteq Y$. Otherwise, by Proposition 3.3, there would be distinct $y_1,y_2,\ldots\in K$ and a strictly decreasing sequence $t_n\to 0$ such that $\chi(y_n,t_n)\greaterthan n$. Going to a subsequence if necessary, we can construct, by Proposition 2.10(b), $f\in C_c(X,\mathbb{R})$ with $f(\phi(y_n))=t_n$, so $Tf(y_n)\greaterthan n$ for all $n$, a contradiction to the boundedness of $Tf$.

To prove that $F$ is isolated we prove that it is open: If $z\in F$, then there is $t\in(0,1)$ with $\chi(z,t)\greaterthan 1$. Take $f\in C_c(X,\mathbb{R})$ such that $f=t$ on a neighbourhood $U$ of $\phi(z)$. In particular $Tf(z)\greaterthan 1$, so there is a neighbourhood $W$ of $z$ such that $Tf\greaterthan 1$ on $W$. Then for all $y\in\phi^{-1}(U)\cap W$, $\chi(y,t)=Tf(y)\greaterthan 1$, so $y\in F$.

Therefore, $F$ consists of isolated points, since $Y$ is locally compact. For $y\not\in F$, Proposition 2.7 and Proposition 3.3 imply that $\chi(y,\cdot)$ has the form $\chi(y,t)=\operatorname{sgn}(t)|t|^{p(y)}$ for some $p(y)\greaterthan 0$. As for the continuity of $p$, Let $U$ be any relatively compact open subset of $Y$ not intersecting $F$, and $f\in C_c(X,\mathbb{R})$ with $f=2$ on $\phi(U)$. Then $Tf(y)=f(y)^{p(y)}=2^p(y)$, so $p=\log_{2}\circ Tf$ on $U$. This proves that $T$ is continuous on $Y\setminus F$.

3.2. Group-Valued functions; the Hernández‒Ródenas Theorem

Given topological groups $G$ and $H$, denote by $\operatorname{AbsIso}(G,H)$ the set of algebraic group isomorphisms from $G$ to $H$, and by $\operatorname{TopIso}(G,H)$ the set of topological (i.e., homeomorphisms which are) group isomorphisms.

Let $X$ and $Y$ be compact Hausdorff spaces and $G$ a Hausdorff topological group. In [MR2324919, Theorem 3.7], Hernández and Ródenas classified non-vanishing group morphisms (not necessarily isomorphisms) $T\colon C(X,G)\to C(Y,G)$ which satisfy the following properties:

  1. There exists a continuous group morphism $\psi\colon G\to C(X,G)$, where $C(X,G)$ is endowed with the topology of pointwise convergence, such that for all $\alpha\in G$ and all $y\in Y$, $T(\psi(\alpha))(y)=\alpha$;

  2. For every continuous endomorphism $\theta\colon G\to G$ and every $f\in C(X,G)$, $T(\theta\circ f)=\theta\circ (Tf)$.

If $T$ is a group isomorphism and $T^{-1}$ is continuous (with respect to uniform convergence) then condition (i) is immediately satisfied, however this is not true for (ii): For example, if $\operatorname{TopIso}(G,G)$ is non-abelian and $\rho\in\operatorname{TopIso}(G,G)$ is any non-central element, then the map $T\colon C(X,G)\to C(X,G)$ given by $Tf=\rho\circ f$ is a group isomorphism, and a self-homeomorphism of $C(X,G)$ with the topology of uniform convergence, which does not satisfy (ii).

In the next theorem we obtain the same type of classification as in [MR2324919, Theorem 3.7], without assuming condition (ii), however we consider only non-vanishing group isomorphisms. Given a compact Hausdorff space $X$, a topological group $G$ and $\alpha\in G$, denote by $\overline{\alpha}$ the constant function $X\to G$, $x\mapsto \alpha$. We endow $C(X,G)$ with the topology of pointwise convergence.

Theorem 3.5.

Suppose that $G$ and $H$ are Hausdorff topological groups, and $X$ and $Y$ are compact Hausdorff spaces for which $(X,\overline{1_G},C(X,G))$ and $(Y,\overline{1_H},C(Y,H))$ are regular.

Let $T\colon C(X,G)\to C(Y,H)$ be a non-vanishing group isomorphism (Definition 2.12). Then there exist a homeomorphism $\phi\colon Y\to X$ and a map $w\colon Y\to\operatorname{AbsIso}(H,G)$ such that $Tf(y)=w(y)(f(\phi(y))$ for all $y\in Y$ and $f\in C(X,G)$.

If $T$ is continuous on the constant functions then each $w(y)$ is continuous and $T$ is continuous on $C(X,G)$. If both $T$ and $T^{-1}$ are continuous on the constant functions then $w(y)\in\operatorname{TopIso}(H,G)$ and $T$ is a homeomorphism for the topologies of pointwise convergence.

Proof.

By Theorem 2.14 and Proposition 2.8, there is a homeomorphism $\phi\colon Y\to X$ such that $T$ is $\phi$-basic, and the sections $\chi(y,\cdot)\colon G\to H$ of the $(\phi,T)$-transform $\chi$ are group morphisms by Proposition 2.7. Similar facts hold for $T^{-1}$, so Proposition 2.6 implies that each section $\chi(y,\cdot)$ is bijective. Letting $w(y)=\chi(y,\cdot)$ we are done with the first part.

Now note that for all $\alpha\in G$ and $y\in Y$,

\begin{equation*} w(y)(\alpha)=w(y)(\overline{\alpha}(\phi(y)))=T\overline{\alpha}(y) \end{equation*}
which implies that every $w(y)$ is continuous if and only if $T$ is continuous on the constant functions (because the map $\alpha\mapsto\overline{\alpha}$ from $G$ to $C(X,G)$ is a homeomorphism onto its image). In this case, from the equality
\begin{equation*} Tf(y)=w(y)(f(\phi(y))\qquad\text{for all}\quad f\in C(X,G)\quad\text{and}\quad y\in Y, \end{equation*}
we can readily see that $T$ is continuous. The last part, assuming also that $T^{-1}$ is continuous on the constant functions, is similar, using $T^{-1}$ and $w(y)^{-1}$ in place of $T$ and $w(y)$.

3.3. Kaplansky's Theorem

Let $R$ be a totally ordered set without supremum nor infimum, regarded as a topological space with the order topology, and let $X$ be a locally compact Hausdorff space. We consider the pointwise order on $C(X,R)$: $f\leq g$ if and only if $f(x)\leq g(x)$ for all $x$, which makes $C(X,R)$ a lattice: for all $f,g\in C(X,R)$ and $x\in X$,

\begin{equation*} (f\lor g)(x)=\max\left\{f(x),g(x)\right\},\qquad\text{and}\qquad(f\land g)(x)=\min\left\{f(x),g(x)\right\}. \end{equation*}
Denote by $C_b(X,R)$ the sublattice of bounded continuous functions from $X$ to $R$.

In [MR0020715] Kaplansky proved that if $X$ is compact and $R$-normal (this is a stronger version of strong regularity; see [MR0020715, p. 618]), then the lattice $C(X,R)$ determines $X$ completely, and in [MR0026240], classified additive lattice isomorphisms between these lattices of functions in the case that $R=\mathbb{R}$. We improve on these results in the following ways: We allow $X$ to be non-compact (only locally compact), obtain a recovery theorem for $X$ from a subcollection $\mathcal{A}$ of $C_b(X,R)$ (Theorem 3.11), and classify lattice isomorphisms in the case of non-real-valued functions for first-countable spaces (Theorem 3.13).

We will consider sublattices $\mathcal{A}$ of $C_b(X,R)$ which satisfy

  1. for all $f,g\in\mathcal{A}$, $\overline{[f\neq g]}$ is compact;

  2. for all $f\in\mathcal{A}$, every open set $U\subseteq X$, every $x\in U$ and $\alpha\in R$, there exists $g\in\mathcal{A}$ such that $g(x)=\alpha$ and $[g\neq f]\subseteq U$.

Condition (L1) is equivalent to saying that $\mathcal{A}\subseteq C_c(X,\theta)$, where $\theta$ is some (any) element of $\mathcal{A}$, and condition (L2) is a form of regularity. These conditions are satisfied in the settings that have been previously studied.

Example 3.6 (Kaplansky, [MR0020715]).

Suppose that $X$ is compact and $\mathcal{A}$ is an $R$-normal sublattice of $C(X,R)$. Condition (L1) is trivial, so let us check that $\mathcal{A}$ satisfies (L2): Suppose $f$, $U$, $x$ and $\alpha$ are as in that condition. For the sake of the argument we can assume $f(x)\leq\alpha$. Let $\beta$ be any lower bound of $f(X)$, and from $R$-normality find $h\in\mathcal{A}$ such that $h(x)=\alpha$ and $h=\beta$ outside $U$. Then $g=f\lor h$ has the desired properties.

Example 3.7 (Li‒Wong, [MR3162258]).

Suppose that $X$ is compact, $R=\mathbb{R}$, and $\mathcal{A}$ is a regular additive subgroup of $C(X,\mathbb{R})$, so (L1) is also trivial and again we need to verify condition (L2): Let $f$, $U$, $x$ and $\alpha$ be as in (L2). By regularity, take $h$ such that $\supp(h)\subseteq U$ and $h(x)=\alpha-f(x)$. Then $g=f+h$ has the desired properties

We will now recover the main result of [MR0020715], The following lemma is based on [MR0052760].

Lemma 3.8.

Let $\mathcal{A}$ be a sublattice of $C_b(X,R)$ satisfying (L1) and (L2), and let $f_0$ be any element of $\mathcal{A}$. Let $\mathcal{A}_{\geq f_0}=\left\{f\in\mathcal{A}:f\geq f_0\right\}$. Then $(X,f_0,\mathcal{A}_{\geq f_0})$ is weakly regular and for $f,g\in\mathcal{A}_{\geq f_0}$,

  1. $f\perp g\iff f\land g=f_0$ (which is the minimum of $\mathcal{A}_{\geq f_0})$;

  2. $f\Subset g$ is equivalent to the following statement:

    \begin{equation*} \begin{split} \text{``for every bounded subset $\mathcal{H}\subseteq\mathcal{A}$ such that $h\subseteq f$ for all $h\in\mathcal{H}$,}\\ \text{there is an upper bound $k$ of $\mathcal{H}$ such that $k\subseteq g$.''} \end{split}\tag{K} \end{equation*}

Proof.

Weak regularity is immediate from (L2) and the fact that $R$ does not have a supremum, and item (a) is trivial. Let us prove (b). First suppose $f\Subset g$ and $\mathcal{H}\subseteq\mathcal{A}$ is a bounded subset such that $h\subseteq f$ for all $h\in \mathcal{H}$. Let $\alpha\in R$ be an upper bound of $\bigcup_{h\in\mathcal{H}}h(X)$. From weak regularity and compactness of $\supp(f)$, we can take finitely many functions $k_1,\ldots,k_n$ such that $k_i\Subset g$, and for every $x\in\supp(f)$ there is some $i$ with $k_i(x)\greaterthan \alpha$. Letting $k=\bigvee_{i=1}^n k_i$ we obtain the desired properties.

Conversely, suppose that condition (K) holds. Let $\alpha$ be any upper bound of $f_0(X)$ and again take $\beta\greaterthan \alpha$. Let $\mathscr{H}=\left\{h\in\mathcal{A}_{\geq f_0}:h\leq \beta, h\subseteq f\right\}$. Let $k$ be an upper bound of $\mathcal{H}$ with $k\subseteq g$. By Property (L2), we have $\sigma(f)=\bigcup_{h\in\mathscr{H}}\sigma(h)$, so $k\geq\beta$ on $\sigma(f)$ and thus also on $\supp(f)$, which implies $f\Subset k$. Since $k\subseteq g$ then $f\Subset g$.

As an immediate consequence of Lemma 3.8 and Theorem 1.17 we have the following generalization of Kaplansky's Theorem:

Theorem 3.9 (Kaplansky [MR0020715]).

Suppose $R$ has no supremum nor infimum, $\mathcal{A}(X)$ and $\mathcal{A}(Y)$ are sublattices of $C_b(X,R)$ and $C_b(Y,R)$, respectively, satisfying conditions (L1) and (L2), and $T\colon\mathcal{A}(X)\to\mathcal{A}(Y)$ is a lattice isomorphism. Then for any $f_0\in\mathcal{A}$, $T$ restricts to a $\perpp$-isomorphism of the regular sublattices $\mathcal{A}(X)_{\geq f_0}$ and $\mathcal{A}(Y)_{\geq Tf_0}$. In particular, $X$ and $Y$ are homeomorphic.

Our immediate goal is to prove that the homeomorphism between $X$ and $Y$ given by Theorem 3.9 does not depend on the choice of function $f_0$ in Lemma 3.8.

Lemma 3.10.

Under the conditions of Theorem 3.9, let $\phi\colon Y\to X$ be the $T|_{\mathcal{A}(X)_{\geq f_0}}$-homeomorphism. If $f,g\in\mathcal{A}(X)_{\geq{f_0}}$ then $\phi^{-1}(\operatorname{int}([f=g]))=\operatorname{int}([Tf=Tg])$.

Proof.

We will use the superscript $f_0$ as in Definition 1.1, that is, for all $f\geq f_0$.

\begin{equation*} \sigma^{f_0}(f)=\operatorname{int}(\overline{[f\neq f_0]}),\qquad\text{and}\qquad Z^{f_0}(f)=\operatorname{int}([f=f_0]). \end{equation*}
and similarly with $Tf_0$ in place of $f_0$. Then $\phi$ is the only homeomorphism satisfying $\sigma^{f_0}(f)=\phi(\sigma^{Tf_0}(Tf))$, or equivalently $Z^{f_0}(f)=\phi(Z^{Tf_0}(Tf))$, for all $f\geq f_0$.

First, assume that $f\leq g$, and let $U=\operatorname{int}([f=g])$. For all $x\in[f\smallerthan g]$, choose a function $k_x\in\mathcal{A}(X)_{\geq f_0}$ such that

Then $g=\sup\left\{f,k_x:x\in[f\smallerthan g]\right\}$, so $Tg=\sup\left\{Tf,Tk_x:x\in[f\smallerthan g]\right\}$..

Let us prove that $Tf(y)=Tg(y)$ for all $y\in\phi(U)$. Since $U\subseteq\bigcap_{x\in[f\smallerthan g]}Z^{f_0}(k_x)$, then $\phi^{-1}(U)\subseteq\bigcap_{x\in[x\smallerthan g]}Z^{Tf_0}(Tk_x)$.

Given $y\in\phi(U)$, use property (L2) to find $h\in\mathcal{A}(Y)$ such that $h(y)=Tf(y)$ and $h=Tg$ outside of $\phi(U)$. Then $h'=(Tf\lor h)\land Tg$ is an upper bound of $\left\{Tf,Tk_x:x\in[f\smallerthan g]\right\}$, so it is also an upper bound of $Tg$, and in particular

\begin{align*} Tg(y)\leq ((Tf\lor h)\land Tg)(y)=Tf(y)\leq (Tf\lor h)(y)=Tf(y), \end{align*}
so $Tg(y)=Tf(y)$.

In the general case, if $f=g$ on an open set $U$ then $f=f\land g=g$ on $U$, so the previous case implies that $Tf$ and $Tg$ both coincide with $T(f\land g)$ on $\phi^{-1}(U)$.

Theorem 3.11.

Under the conditions of Theorem 3.9, there exists a unique homeomorphism $\phi\colon Y\to X$ such that $\phi(\operatorname{int}([Tf=Tg]))=\operatorname{int}([f=g])$ for all $f,g\in\mathcal{A}(X)$. (In this case we will still call $\phi$ the $T$-homeomorphism.)

Proof.

For each $f_0\in \mathcal{A}(X)$, let $\phi^{f_0}\colon Y\to X$ be the $T|_{\mathcal{A}(X)_{\geq f_0}}$-homeomorphism. Given $f_0,g_0\in\mathcal{A}(X)$, Lemma 3.10 implies that $\phi^{f_0\land g_0}$ satisfies the property of both the $T|_{\mathcal{A}(X)_{\geq f_0}}$ and the $T|_{\mathcal{A}(X)_{\geq g_0}}$-homeomorphisms, so $\phi^{f_0}=\phi^{f_0\land g_0}=\phi^{g_0}$. We are done by letting $\phi=\phi^{f_0}$ for some arbitrary $f_0\in\mathcal{A}(X)$.

A natural goal now is to classify the lattice isomorphisms as given in Theorem 3.11, which is possible when we consider first-countable spaces. A similar argument to that of Theorem 2.6 appears in [MR2995073], although in a different context (considering lattices of possibly unbounded real-valued continuous functions on complete metric spaces). See also [MR2999998], MR3404615].

Let us reinforce the assumptions, assumed throughout this subsection, that $X$ denotes a locally compact Hausdorff space and $R$ is a totally ordered set without maximum or minimum with the order topology. The following is a version of Proposition 2.10 in this setting, and is also proven by cutting and pasting (aided by the lattice operations of $R$), so we omit the details.

Proposition 3.12.

Assume further that $X$ and $R$ are first-countable, and that $\theta\in C_b(X,R)$ is such that $C_c(X,\theta)$ satisfies (L2). Suppose that $\left\{x_n\right\}_n$ is an injective sequence in $X$, converging to $x_\infty\in X$. Let $g_n\in C(X,R)$ be functions such that $g_n(x_n)\to \theta(x)$.

Then there exists $f\in C_c(X,\theta)$ such that $f=g_n$ on a neighbourhood of $x_n$ for each $n$, and that $f=\theta$ outside of a compact containing $\left\{x_n:n\in\mathbb{N}\right\}$ (which may be taken as small as desired with this property).

Theorem 3.13.

Suppose that $X$, $Y$ and $R$ are first-countable, that $C_c(X,\theta_X)$ and $C_c(Y,\theta_Y)$ satisfy (L2), and that $T\colon C_c(X,\theta_X)\to C_c(Y,\theta_Y)$ is a lattice isomorphism. Then there are a unique homeomorphism $\phi\colon Y\to X$ and a continuous function $\chi\colon Y\times R\to R$ such that

\begin{equation*} Tf(y)=\chi(y,f(\phi(y)))\qquad\text{ for all }y\in Y\text{ and }f\in C_c(X,0)\tag{3.2} \end{equation*}
and $\chi(y,\cdot)\colon R\to R$ is an increasing bijection for each $y\in Y$.

Proof.

Let $\phi\colon Y\to X$ be the $T$-homeomorphism. We just need to prove that $T$ is $\phi$-basic, so assume $y\in Y$ and $f(\phi(y))=g(\phi(y))$. In order to prove that $Tf(y)=Tg(y)$, we may assume that $f\leq g$, by considering the auxiliary function $f\land g$.

If $y$ is isolated in $Y$, then $f$ and $g$ coincide on the open set $\left\{\phi(y)\right\}$, so $Tf$ and $Tg$ coincide on the open set $\left\{y\right\}$.

Assume then that $y$ is not isolated. Since $Y$ is first-countable, let $(y_n)_n$ be an injective sequence in $Y$ converging to $y$. By Proposition 3.12, there is $h\in C_c(X,\theta)$ such that

Then $\phi(y)\in\overline{\operatorname{int}[f=h]}$, so $t\in\overline{\operatorname{int}[Tf=Th]}$ and so $Tf(y)=Th(y)$. Similarly, $Tg(y)=Th(y)=Tf(y)$. This proves that $T$ is $\phi$-basic. Let $\chi$ be the $(\phi,T)$-transform.

Proposition 2.7, applied to the signature of lattices (with the binary symbol $\lor$ interpreted as join) implies that the sections $\chi(y,\cdot)$ are lattice isomorphisms of $R$ for all $n$, and in particular homeomorphisms. The proof that $\chi$ is continuous is similar to that of implication (ii)$\Rightarrow$(i) of Theorem 2.11 (using Proposition 3.12 instead of 2.10).

In the case of additive lattice isomorphisms of spaces of real-valued functions, we do not require the first-countability hypothesis.

Theorem 3.14.

Suppose $R=\mathbb{R}$, and $T\colon C_c(X)\to C_c(Y)$ is an additive lattice isomorphism. Then there are a unique homeomorphism $\phi\colon Y\to X$ and a unique positive continuous function $p\colon Y\to (0,\infty)$ such that $Tf(y)=p(y)f(\phi(y))$ for all $f\in C_c(X)$ and $y\in Y$.

Proof.

First note that for all $f\in C_c(X)$, $|f|=(f\lor 0)-(f\land 0)$, so $T|f|=|Tf|$. Let $\phi\colon Y\to X$ be the $T$-homeomorphism, given by Theorem 3.11.

The proof that $T$ is $\phi$-basic is similar to that of item 6. of the proof of Theorem 3.4. Let $\chi$ be the $T$-transform. Each section $\chi(y,\cdot)$ is an additive order-preserving bijection (Propositions 2.7 and 2.6) and hence has the form $\chi(y,t)=p(y)t$ for some $p(y)\greaterthan 0$. If $Tf(y)\neq 0$, then $f(\phi(y))\neq 0$ as well and $p=Tf/(f\circ \phi)$ on a neighbourhood of $y$, thus $p$ is continuous.

3.4. Li‒Wong Theorem

We will now recover a theorem of Li and Wong, [MR3162258, Theorem 2.2], which can be seen as a generalization of Theorem 3.14. We will proceed in the opposite direction, i.e., by proving their result instead as a consequence of the more general Theorem 3.9. Let $\mathbb{K}=\mathbb{R}$ or $\mathbb{C}$.

Theorem 3.15 (Li‒Wong [MR3162258]).

Let $X$ and $Y$ be compact Hausdorff spaces, and $\mathcal{A}(X)$ and $\mathcal{A}(Y)$ be two regular vector sublattices of $C(X,\mathbb{K})$ and $C(Y,\mathbb{K})$, respectively. Suppose that $T\colon \mathcal{A}(X)\to \mathcal{A}(Y)$ is a $\mathbb{K}$-linear isomorphism which preserves non-vanishing functions, that is, for all $f\in\mathcal{A}(X)$,

\begin{equation*} 0\in f(X)\iff 0\in Tf(Y). \end{equation*}
Then there is a homeomorphism $\phi\colon Y\to X$ and a continuous non-vanishing function $p\colon Y\to\mathbb{K}$ such that $Tf(y)=p(y)f(\phi(y))$ for all $f\in\mathcal{A}(X)$ and $y\in Y$.

The following technical lemma is the main necessary tool of the proof. We do not assume that $\mathcal{A}(X)$ contains the constant functions, however since it is a regular lattice then it contains a strictly positive function $F$ satisfying $0\smallerthan F\smallerthan 1/2$ (see the beginning of the proof of Theorem 3.15 below). The use of the constant function $1/2$ in the proof of [MR3162258, Lemma 2.3] can be replaced by $F$.

Lemma 3.16 ([MR3162258, Lemma 2.3]).

Any $T$ as in Theorem 3.15 is a $\perp$-isomorphism.

Proof of Theorem 3.15.

We will use Theorem 3.11. First we need to modify $T$ to obtain a lattice isomorphism. Since $\mathcal{A}(X)$ is a sublattice, then for all $f\in\mathcal{A}(X)$,

\begin{equation*} f^+=\max(f,0),\quad f^-=\max(-f,0)\quad\text{and}\quad|f|=f^++f^-\text{ belong to }\mathcal{A}(X). \end{equation*}

As $\mathcal{A}(X)$ is regular and $X$ is compact, we can take finitely many functions $f_1,\ldots,f_n\in\mathcal{A}(X)$ such that for all $x\in X$, $f_i(x)\neq 0$ for some $i$, and therefore the function $F=\sum_{i=1}^n|f_i|$ belongs to $\mathcal{A}(X)$ and is non-vanishing, so $TF$ is also non-vanishing. We define new classes of functions

\begin{equation*} \mathcal{B}(X)=\left\{f/F:f\in\mathcal{A}(X)\right\},\qquad\mathcal{B}(Y)=\left\{g/TF:g\in\mathcal{A}(Y)\right\}. \end{equation*}
It is immediate to see that $\mathcal{B}(X)$ and $\mathcal{B}(Y)$ are regular, and contain the constant functions of $X$ and $Y$, respectively. Define a linear isomorphism $S\colon \mathcal{B}(X)\to\mathcal{B}(Y)$, $S(f)=T(fF)/TF$, which preserves non-vanishing functions and satisfies $S(1)=1$. Given a scalar $\lambda$, linearity and the non-vanishing property of $S$ imply that, for all $f\in\mathcal{B}(X)$,
\begin{align*} \lambda\not\in f(X)&\iff f-\lambda\text{ is non-vanishing}\\ &\iff Sf-\lambda\text{ is non-vanishing}\iff \lambda\not\in Sf(Y), \end{align*}
so $f(X)=Sf(Y)$, i.e., $S$ preserves images of functions.

However, in order to apply Theorem 3.11 we also need to make sure that $\mathcal{B}(X)$ and $\mathcal{B}(Y)$ are lattices. As $F\greaterthan 0$, it readily follows that $\mathcal{B}(X)$ is a (self-adjoint) sublattice of $C(X,\mathbb{K})$, however this is not so immediate for $\mathcal{B}(Y)$.

As $S$ preserves images of functions, it preserves real functions. If $f\in\mathcal{B}(X)$, then $S(\operatorname{Re}(f))$ and $S(\operatorname{Im}(f))$ are real functions such that $Sf=S(\operatorname{Re}(f))+iS(\operatorname{Im}(f))$. As $T$ is a $\perp$-isomorphism then $S$ is also a $\perp$-isomorphism, so we also obtain $S(\operatorname{Re}(f))\perp S(\operatorname{Im}(f))$. This is enough to conclude that $S$ preserves real and imaginary parts of functions, from which it follows that $\mathcal{B}(Y)$ is self-adjoint. Similarly, $S$ preserves positive and negative parts of functions. In particular, if $f\in\mathcal{B}(Y)$ then $f^+\in\mathcal{B}(Y)$, and this is enough to conclude that $\mathcal{B}(Y)$ is a sublattice of $C(Y,\mathbb{K})$, and that $S$ is an order-preserving isomorphism.

We may then consider only real-valued functions, and the complex case will follow by linearity (and since $S$ preserves real and imaginary parts). By Kaplansky's Theorem (3.11), we can construct the $S$-homeomorphism $\phi\colon Y\to X$. Now we need to prove that $S$ is $\phi$-basic.

Suppose $f(x)\neq 0$ for a given $x\in X$, and let us assume, without loss of generality, that $f(x)\greaterthan 0$. Then $f\greaterthan 0$ on some neighbourhood $U$ of $x$. Again using compactness of $X\setminus U$ and regularity of the sublattice $\mathcal{B}(X)$ we can construct a function $g\in\mathcal{B}(X)$ such that $g=0$ on some neighbourhood of $x$ and $g\greaterthan 0$ on $X\setminus U$. Letting $\widetilde{f}=f\lor g$, we have $\widetilde{f}=f$ on some neighbourhood of $x$, so $S\widetilde{f}=Sf$ on some neighbourhood of $\phi^{-1}(x)$. But $\widetilde{f}$ is non-vanishing, so $S\widetilde{f}$ is also non vanishing and in particular $Sf(\phi^{-1}(x))\neq 0$. This proves that $S$ is basic with respect to $\phi$.

Letting $\chi\colon Y\times\mathbb{R}\to\mathbb{R}$ be the $S$-transform, we have that all sections $\chi(y,\cdot)$ are linear and increasing (Theorem 2.7), hence of the form $\chi(y,t)=P(y)t$ for a certain $P(y)\greaterthan 0$. Since $S(1)=1$ then $P=1$, that is, $\chi(y,t)=t$ for all $t\in\mathbb{R}$.

Finally, for all $f\in\mathcal{A}(X)$ and $y\in Y$,

\begin{equation*} Tf(y)=(TF)(y)\left[S\left(\frac{f}{F}\right)(y)\right]=(TF)(y)\chi\left(y,\frac{f}{F}(\phi(y))\right)=\frac{TF(y)}{F(\phi(y))}f(\phi(y)), \end{equation*}
as we wanted.

3.5. Jarosz' Theorem

Throughout this subsection, we fix $\mathbb{K}=\mathbb{R}$ or $\mathbb{C}$. Given a locally compact Hausdorff space $X$, we let $C_c(X)=C_c(X,0)$, the vector space of compactly supported $\mathbb{K}$-valued functions.

Theorem 3.17 (Jarosz [MR1060366]).

If $T\colon C_c(X)\to C_c(Y)$ is a linear $\perp$-isomorphism, then there exist a homeomorphism $\phi\colon Y\to X$ and a continuous non-vanishing function $p\colon Y\to\mathbb{K}$ such that $Tf(y)=p(y)f(\phi(y))$ for all $f\in C_c(X)$ and $y\in Y$.

Proof.

First assume that $X$ and $Y$ are compact, and let us show that $f\neq 0$ everywhere if and only if $Tf\neq 0$ everywhere. Suppose otherwise, say $f(x)=0$, and we have two cases: first, if $f$ is constant on a neighbourhood of $x$, this means that $Z(f)\neq \varnothing$, and Theorem 1.16 implies that $Z(Tf)\neq \varnothing$, and in particular $Tf(y)=0$ for any $y\in Z(Tf)$.

In the second case, if $f$ is not constant on any neighbourhood of $x$, an argument similar to the one in the proof of Theorem 3.4 yields a contradiction to $Tf$ being bounded, so $Tf(y)=0$ for some $y\in Y$.

The result follows in this case from the Li‒Wong Theorem (Theorem 3.15).

Now let $X$ and $Y$ be arbitrary locally compact Hausdorff. Given $b\in C_c(X)$, set $T_b\colon C(\supp(b))\to C(\supp(Tb))$ as $T_bf=(Tf')|_{\supp(Tb)}$, where $f'$ is any element of $C_c(X)$ extending $f$. Note that $T_bf$ does not depend on the choice of $f'$, because $T$ is an additive $\perp$-isomorphism (and Theorem 1.16).

Since $f\perpp g$ if and only if $f|_{\supp(b)}\perpp g|_{\supp(b)}$ for all $b$, the previous case allows us to obtain functions $p^b$ and $\phi^b$ such that $Tf(y)=p^b(y)f(\phi^b(y))$ for all $y\in\supp(Tb)$. Clearly, if $b\subseteq b'$ then $p^{b'}|_{\supp(b)}=p^b$ and $\phi^{b'}|_{\supp(b)}=\phi^{b'}$. Thus defining $p$ and $\phi$ as $p(y)=p^b(y)$ and $\phi(y)=\phi^b(y)$ where $b\in C_c(X)$ is such that $y\in\supp(b)$ we obtain the desired maps.

3.6. Banach‒Stone Theorem

We use the same notation as in the previous subsection. Given a locally compact Hausdorff space $X$, endow $C_c(X)$ with the supremum norm: $\Vert f\Vert_\infty=\sup_{x\in X}|f(x)|$.

Recall that, by the Riesz‒Markov‒Kakutani Representation Theorem ([MR584266, Theorem 2.14]), continuous linear functionals on $C_c(X)$ correspond to (integration with respect to) regular Borel measures on $X$. As a consequence, the extremal points $T$ of the unit ball of the dual of $C_c(X)$ have the form $T(f)=\lambda f(x)$ for some $x\in X$ and $|\lambda|=1$.

Given $f\in C_c(X)$, denote by $N(f)$ the set of extremal points $T$ in the unit ball of the dual space $C_c(X)^*$ such that $T(f)\neq 0$. From the previous paragraph we obtain

\begin{equation*} f\perp g\iff N(f)\cap N(g)=\varnothing\tag{BS}, \end{equation*}
and the Banach‒Stone Theorem is an immediate consequence of Jarosz' Theorem.

Theorem 3.18 (Banach‒Stone [MR1501905]).

Let $X$ and $Y$ be locally compact Hausdorff spaces and let $T\colon C_c(X)\to C_c(Y)$ be an isometric linear isomorphism. Then there exists a homeomorphism $\phi\colon Y\to X$ and a continuous function $p\colon Y\to\mathbb{S}^1$ for which

\begin{equation*} Tf(y)=p(y)f(\phi(y))\qquad\forall f\in C(X),\ \forall y\in Y. \end{equation*}

3.7. $L^1$-spaces

Let $\mathbb{K}=\mathbb{R}$ or $\mathbb{C}$ be fixed. Given a topological space $X$, $C_c(X)$ will denote the space of $\mathbb{K}$-valued compactly supported continuous functions on $X$, where supports are the usual ones: $\supp(f)=\overline{[f\neq 0]}$.

Following [MR924157], a Borel measure $\mu$ on $X$ will be called regular if

and recall that the support of $\mu$ is the set of points $x\in X$ whose neighbourhoods always have positive measure. We say that $\mu$ is fully supported (on $X$) is the support of $\mu$ coincides with $X$, i.e., if every nonempty open subset has positive measure.

We will now prove that the $C_c(X)$ endowed with the $L^1$-norm of a fully supported measure $\mu$ completely determine both $X$ and $\mu$. The following lemma describes the $\perp$ relation in terms of the $L^1$-norm, and follows from elementary measure-theoretic considerations.

Lemma 3.19.

Let $X$ be a locally compact Hausdorff space and $\mu$ a fully supported, locally finite Borel measure on $X$. If $\Vert\cdot\Vert_1$ denotes the corresponding $L^1$-norm, then for all $f,g\in C_c(X)$, $f\perp g$ if and only if

\begin{equation*} \Vert Af+Bg\Vert_1=|A|\Vert f\Vert_1+|B|\Vert g\Vert_1\qquad\forall A,B\in\mathbb{K}\tag{3.3} \end{equation*}

Theorem 3.20.

Let $X$ and $Y$ be locally compact Hausdorff spaces with fully supported regular Borel measures $\mu_X$ and $\mu_Y$, and let $T\colon C_c(X)\to C_c(Y)$ be a linear isomorphism which is isometric with respect to the $L^1$-norms. Then there exists a homeomorphism $\phi\colon Y\to X$ and a continuous function $p\colon Y\to\mathbb{S}^1$ such that

\begin{equation*} Tf(y)=p(y)\frac{d\mu_X}{d(\phi_*\mu_Y)}(\phi(y))f(\phi(y)) \end{equation*}
for all $f\in C_c(X)$ and $y\in Y$.

Proof.

By the previous lemma, $T$ is a $\perp$-isomorphism, so Jarosz' Theorem (3.17) implies that there are a homeomorphism $\phi\colon Y\to X$ and a non-vanishing continuous function $P\colon Y\to\mathbb{C}$ such that $T(f)(y)=P(y)f(\phi(y))$ for all $f\in C_c(X)$ and $y\in Y$.

Now using the fact that $T$ is isometric, we have, for every $f\in C_c(X)$,

\begin{equation*} \int_X|f|d\mu_X=\int_Y|Tf|d\mu_Y=\int_Y|P||f\circ\phi|d\mu_Y=\int_X|P\circ\phi^{-1}||f|d(\phi_*\mu_Y) \end{equation*}
which means that $|P\circ\phi^{-1}|$ is a continuous instance of the Radon‒Nikodym derivative $d\mu_X/d(\phi_*\mu_Y)$. Since $p=P/|P|\colon Y\to\mathbb{S}^1$ is continuous, we obtain the result.

3.8. Measured groupoid convolution algebras

In the next three results, we will focus on convolution algebras of topological groupoids. First, we will consider measured groupoids in the sense of Hahn. See [MR496796], MR496797], MR584266], MR1444088], MR0427598]. Note that throughout this section we consider only regular measures.

Recall that a groupoid $\mathcal{G}$ is a small category with inverses, and a topological groupoid is a groupoid endowed with a topology making the product and inversion maps continuous.

The source and range maps on $\mathcal{G}$ are defined as $\so(a)=a^{-1}a$ and $\ra(a)=aa^{-1}$, respectively. The unit space of $\mathcal{G}$ is $\mathcal{G}^{(0)}=\so(\mathcal{G})$, and is identified with the object space of $\mathcal{G}$. We denote by $\mathcal{G}^{(2)}=\left\{(a,b)\in\mathcal{G}\times\mathcal{G}:\so(a)=\ra(b)\right\}$ the set of composable pairs, i.e., pairs $(a,b)$ for which the product $ab$ is defined. Given $x,y\in\mathcal{G}^{(0)}$, we denote $\mathcal{G}^y=\ra^{-1}(x)$, $\mathcal{G}_x=\so^{-1}(x)$, and $\mathcal{G}_x^y=\mathcal{G}^y\cap\mathcal{G}_x$. We call $\mathcal{G}_x^x$ the isotropy group at $x$. The product of two subsets $A,B\subseteq\mathcal{G}$ is $AB=\left\{ab:(a,b)\in(A\times B)\cap\mathcal{G}^{(2)}\right\}$.

Common examples of topological groupoids are: Equivalence relations, topological groups (where $\mathcal{G}^{(0)}$ is a singleton) and topological spaces (where $\mathcal{G}=\mathcal{G}^{(0)}$). More generally, every continuous group action induces a transformation groupoid.

Initially, given a locally compact Hausdorff topological groupoid $\mathcal{G}$, we consider $C_c(\mathcal{G})$, the space of real or complex-valued, compactly supported, continuous functions on $\mathcal{G}$, simply as a vector space (with pointwise operations). Recall the notion of a Haar system:

Definition 3.21 ([MR584266, Definition 2.2]).

A (continuous) left Haar system for a locally compact Hausdorff topological groupoid $\mathcal{G}$ is a collection of regular Borel measures $\lambda=\left\{\lambda^x:x\in\mathcal{G}^{(0)}\right\}$ on $\mathcal{G}$ such that

  1. For each $x\in\mathcal{G}^{(0)}$, $\lambda^x$ has support contained in $\mathcal{G}^x$;

  2. (left invariance) For each $a\in\mathcal{G}$, $\lambda^{\ra(a)}(aE)=\lambda^{\so(a)}(E)$ for every compact $E\subseteq \mathcal{G}^{\so(a)}$;
  3. (continuity) For each $f\in C_c(\mathcal{G})$, the map $\mathcal{G}^{(0)}\to\mathbb{C}$, $x\mapsto \int fd\lambda^x$, is continuous.

We will not make any distinction of whether each $\lambda^x$ is considered as a measure on $\mathcal{G}$ or as a measure on $\mathcal{G}^x$. We say that $\lambda$ is fully supported if the support of $\lambda^x$ is all of $\mathcal{G}^x$ for all $x\in\mathcal{G}^{(0)}$.

Left invariance of $\lambda$ implies that for all $a\in\mathcal{G}$ and $f\in C_c(\mathcal{G}^{\ra(a)})$

\begin{equation*} \int f(s)d\lambda^{\ra(a)}(s)=\int f(at)d\lambda^{\so(a)}(t) \end{equation*}
and we endow $C_c(\mathcal{G})$ with convolution product
\begin{equation*} (fg)(a)=\int f(s)g(s^{-1}a)\lambda^{\ra(a)}(s)=\int f(at)g(t^{-1})\lambda^{\so(a)}(t), \end{equation*}
which makes $C_c(\mathcal{G})$ an algebra. It follows that for all $f,g\in C_c(\mathcal{G})$,
\begin{equation*} \supp(fg)\subseteq\supp(f)\supp(g).\tag{3.4} \end{equation*}

Definition 3.22.

Let $\mathcal{G}$ be a locally compact Hausdorff topological groupoid with a left Haar system $\lambda$. Given a regular Borel measure $\mu$ on $\mathcal{G}^{(0)}$, the measure induced by $\mu$ and $\lambda$ is the unique regular Borel measure $(\lambda\circ\mu)$ on $\mathcal{G}$ which satisfies

\begin{equation*} (\lambda\circ\mu)(E)=\int_{\mathcal{G}^{(0)}}\lambda^x(E)d\mu(x) \end{equation*}
for every compact $E\subseteq\mathcal{G}$. (The existence of $(\lambda\circ\mu)$ is guaranteed by the Riesz‒Markov‒Kakutani Representation Theorem.)

If $\mu$ is fully supported on $\mathcal{G}^{(0)}$ and $\lambda$ is a fully supported Haar system on a locally compact Hausdorff groupoid $\mathcal{G}$, then $(\lambda\circ\mu)$ is fully supported on $\mathcal{G}$.

The following lemma will allow us to verify if certain maps are groupoid morphisms.

Lemma 3.23.

Given a topological groupoid $\mathcal{G}$ with $\mathcal{G}^{(0)}$ Hausdorff and $a,b\in\mathcal{G}$, we have $\so(a)=\ra(b)$ if and only if for every pair of neighbourhoods $U$ of $a$ and $V$ of $b$ the product $UV$ is nonempty.

Proof.

From the second condition one can construct two nets $(a_i)_i$ and $(b_i)_i$ (over the same ordered set) converging to $a$ and $b$, respectively, such that $\so(a_i)=\ra(b_i)$, and so $\so(a)=\ra(b)$ because $\mathcal{G}^{(0)}$ is Hausdorff. The reverse implication is trivial.

Lemma 3.24.

If $\lambda$ and $\mu$ are continuous Haar systems on a locally compact Hausdorff topological groupoid $\mathcal{G}$ such that the Radon‒Nikodym derivatives $D^x=\frac{d\lambda^x}{d\mu^x}$ exist for all $x\in\mathcal{G}^{(0)}$, then $D$ is invariant in the sense that for all $a\in\mathcal{G}$ and $\mu^{\so(a)}$-almost every $g\in\mathcal{G}^{\so(a)}$, $D^{\ra(a)}(ag)=D^{\so(a)}(g)$.

Proof.

Using invariance of $\mu$ and $\lambda$, we have, for every $f\in C_c(\mathcal{G}^{\so(a)})$,

\begin{align*} \int f(t) D^{\ra(a)}(at)d\mu^{\so(a)}(t)&=\int f(a^{-1}s)D^{\ra(a)}(s)d\mu^{\ra(a)}(s)\\ &=\int f(a^{-1}s)d\lambda^{\ra(a)}(s)=\int f(t)d\lambda^{\so(a)}(t). \end{align*}
Thus $t\mapsto D^{\ra(a)}(at)$ satisfies the property of the Radon‒Nikodym derivative $d\lambda^{\so(a)}/d\mu^{\so(a)}=D^{\so(a)}$, hence these functions coincide $\mu^{\so(a)}$-a.e.

Now we prove that the convolution algebra $C_c(\mathcal{G})$ together with the $L^1$-norm coming from $\lambda\circ\mu$, where $\lambda$ is a fully supported Haar system on $\mathcal{G}$ and $\mu$ is a fully supported measure on $\mathcal{G}^{(0)}$ completely determines the triple $(\mathcal{G},\lambda,\mu)$, up to isomorphism (compare to [MR2102633]). We denote by $\mathbb{S}^1$ the circle group (of complex numbers with absolute value $1$ under multiplication).

Theorem 3.25.

Let $\mathcal{G}$ and $\mathcal{H}$ be locally compact Hausdorff groupoids. For each $Z\in\left\{\mathcal{G},\mathcal{H}\right\}$, let $\lambda_Z$ be a fully supported Haar system on $Z$, and $\mu_Z$ a fully supported regular Borel measure on $Z^{(0)}$.

If $T\colon C_c(\mathcal{G})\to C_c(\mathcal{H})$ is an algebra isomorphism which is isometric with respect to the $L^1$-norms of $(\lambda_Z\circ \mu_Z)$ ($Z\in\left\{\mathcal{G},\mathcal{H}\right\}$), then there are a topological groupoid isomorphism $\phi:\mathcal{H}\to\mathcal{G}$ and a continuous morphism $p:\mathcal{H}\to\mathbb{S}^1$ such that

\begin{equation*} Tf(h)=p(h)D(\phi(h))f(\phi(h)) \end{equation*}
where $D$ is a continuous instance of the Radon‒Nikodym derivative
\begin{equation*} D(a)=\frac{d\lambda_{\mathcal{G}}^{\ra(a)}}{d(\phi_*\lambda_{\mathcal{H}}^{\phi^{-1}(\ra(a))})}(a) \end{equation*}
and in this case, $\mu_{\mathcal{G}}=\phi_*\mu_{\mathcal{H}}$.

Proof.

Again applying Lemma 3.19 and Jarosz' Theorem (3.17), we can find a homeomorphism $\phi\colon\mathcal{H}\to\mathcal{G}$ and a continuous non-vanishing scalar function $P$ such that

\begin{equation*} Tf(h)=P(h) f(\phi(h))\qquad\text{for all } f\in C_c(\mathcal{G})\text{ and }h\in\mathcal{H}. \end{equation*}

Let us check that $\phi$ is a groupoid morphism. Suppose $(a,b)\in\mathcal{H}^{(2)}$, and consider neighbourhoods $U$ and $V$ of $\phi(a)$ and $\phi(b)$, respectively.

Choose non-negative functions $f_U,f_V\in C_c(\mathcal{H})$ such that

\begin{equation*} \supp(f_a)\subseteq\phi^{-1}(U),\quad\supp(f_b)\subseteq \phi^{-1}(V)\quad\text{and}\quad f_a(a)=f_b(b)=1. \end{equation*}
Then $ab\in\supp(f_af_b)$, because $\lambda_{\mathcal{H}}$ has full support, and so $\phi(ab)\in\supp(T^{-1}(f_af_b))$. As $\phi$ is the $T$-homeomorphism and $T$ is an isomorphism, the inclusion in (3.4) implies $\phi(ab)\subseteq UV$. By Lemma 3.23, the product $\phi(a)\phi(b)$ is defined, and moreover, continuity of the product implies that every neighbourhood of $\phi(a)\phi(b)$ contains $\phi(ab)$. Since $\mathcal{G}$ is Hausdorff then $\phi(ab)=\phi(a)\phi(b)$. Therefore $\phi$ is a morphism and a homeomorphism, thus a topological groupoid isomorphism.

We now need to rewrite the function $P$ as $P(h)=p(h)D(\phi(h))$ as in the statement of the theorem, and to this end we use multiplicativity of $T$ and compare integrals. If $f,g\in C_c(\mathcal{G})$ and $c\in\mathcal{H}$, then on one hand

\begin{align*} T(fg)(c)&=(TfTg)(c)=\int_{\mathcal{G}^{\ra(c)}} Tf(t)Tg(t^{-1}c)\lambda_{\mathcal{H}}^{\ra(c)}(t)\\ &=\int_{\mathcal{H}^{\ra(c)}} P(t)f(\phi(t))P(t^{-1}c)g(\phi(t^{-1}c))d\lambda_{\mathcal{H}}^{\ra(c)}(t)\\ &=\int_{\mathcal{G}^{\phi(\ra(c))}} P(\phi^{-1}(s))f(s)P(\phi^{-1}(s)^{-1}c)g(s^{-1}\phi(c))d(\phi_*\lambda_{\mathcal{H}}^{\ra(c)})(s)\tag{3.5} \end{align*}
and on the other
\begin{align*} T(fg)(c)&=P(c)(fg)(\phi(c))=P(c)\int_{\mathcal{G}^{\phi(\ra(c))}} f(t)g(t^{-1}\phi(c))d\lambda_{\mathcal{G}}^{\phi(\ra(c))}(t)\tag{3.6} \end{align*}

Now let $f\in C_c(\mathcal{G}^{\phi(\ra(c))})$ be an arbitrary non-negative function. Define $g\in C_c(\mathcal{G}_{\phi(\so(c))})$ by $g(t)=f(\phi(c)t^{-1})$. Extending $f$ and $g$ arbitrarily to elements of $C_c(\mathcal{G})$, Equations (3.5) and (3.6) become

\begin{equation*} \int_{\mathcal{G}^{\phi(\ra(c))}} P(\phi^{-1}(s))P(\phi^{-1}(s)^{-1}c)|f(s)|^2d\phi_*\lambda_{\mathcal{H}}^{\ra(c)}(s)=P(c)\int_{\mathcal{G}^{\phi(\ra(c))}} |f(s)|^2d\lambda_{\mathcal{G}}^{\phi(\ra(c))}(s)\tag{3.7} \end{equation*}
for all non-negative $f\in C_c(\mathcal{G}^{\phi(\ra(c))})$ and all $c\in\mathcal{H}$. Define $D\colon\mathcal{G}\to\mathbb{C}$ (or $\mathbb{R}$ in the real case) by
\begin{equation*} D(s)=\frac{P(\phi^{-1}(s))P(\phi^{-1}(s)^{-1})}{P(\phi^{-1}(\ra(s)))}. \end{equation*}
Using Equation (3.7) with $y=\ra(c)$ in place of $c$, we obtain
\begin{equation*} \int_{\mathcal{G}^{\phi(y)}} D(s)|f(s)|^2d(\phi_*\lambda_{\mathcal{H}}^{y})(s)=\int_{\mathcal{G}^{\phi(y)}}|f(s)|^2d\lambda_{\mathcal{G}}^{\phi(y)}(s) \end{equation*}
for all $f\in C_c(\mathcal{G}^{\phi(y)})$, thus $D$ is a continuous instance of the Radon‒Nikodym derivative
\begin{equation*} D(s)=\frac{d\lambda_{\mathcal{G}}^{\phi(y)}}{d(\phi_*\lambda_{\mathcal{H}}^{y})}(s)=\frac{d\lambda_{\mathcal{G}}^{\phi(\ra(c))}}{d(\phi_*\lambda_{\mathcal{H}}^{\ra(c)})}(s). \end{equation*}
Now applying this to Equation (3.7), and using regularity of all measures involved, we conclude that for $\lambda_{\mathcal{G}}^{\phi(\ra(c))}$-a.e. $s\in\mathcal{G}^{\phi(\ra(c))}$
\begin{equation*} P(\phi^{-1}(s))P(\phi^{-1}(s)^{-1}c)=D(s)P(c) \end{equation*}
and since all functions involved are continuous, and $\lambda_{\mathcal{G}}^{\phi(\ra(c))}$ has full support, the same equality is actually valid for all $s\in\mathcal{G}^{\phi(\ra(c))}$. Equivalently, for all $c\in\mathcal{H}$ and all $t\in\mathcal{H}^{\ra(c)}$, $P(t)P(t^{-1}c)=D(\phi(t))P(c)$. Together with Lemma 3.24, this implies that the map $p=P/(D\circ\phi)$ is a continuous groupoid morphism from $\mathcal{H}$ to the group of non-zero scalars.

Now let us verify that $\mu_{\mathcal{G}}$ and $\phi_*\mu_{\mathcal{H}}$ are equivalent measures. By regularity of the measures, we may extend the formula $T(f)=P\cdot (f\circ\phi)$ to obtain a linear isomorphism between the classes of measurable functions on $\mathcal{G}$ and on $\mathcal{H}$, and which restricts to an isometry from $L^1(\lambda_{\mathcal{G}}\circ\mu_{\mathcal{G}})$ to $L^1(\lambda_{\mathcal{H}}\circ\mu_{\mathcal{H}})$. Suppose that $K\subseteq\mathcal{G}^{(0)}$ has positive $\mu_{\mathcal{G}}$-measure, and let $f=1_{\ra^{-1}(K)}$ be the characteristic function of $\ra^{-1}(K)$. Then $Tf=P\cdot 1_{\phi^{-1}(\ra(K))}$. As $\Vert f\Vert_{\mathcal{G}}\greaterthan 0$, then $\Vert Tf\Vert_{\mathcal{H}}\greaterthan 0$ as well, so the support of $Tf$ has positive $(\lambda_{\mathcal{H}}\circ\mu_{\mathcal{H}})$-measure, and this implies that $\ra(\supp(Tf))=\phi^{-1}(K)$ has positive $\mu_{\mathcal{H}}$-measure. Therefore $\mu_{\mathcal{G}}$ is absolutely continuous with respect to $\phi_*\mu_{\mathcal{H}}$. The reverse implication is similar.

To prove that $p$ takes value in $\mathbb{S}^1$, let us denote by $\Vert\cdot\Vert_Z$ the $L^1$-norm with respect to $(\lambda_Z\circ\mu_Z)$ when $Z\in\left\{\mathcal{G},\mathcal{H}\right\}$. For all $f\in C_c(\mathcal{G})$ we have

\begin{align*} \Vert Tf\Vert_{\mathcal{H}}&=\int_{\mathcal{H}^{(0)}}\left(\int_{\mathcal{H}^y}|Tf|d\lambda_{\mathcal{H}}^y\right)d\mu_{\mathcal{H}}(y)=\int_{\mathcal{H}^{(0)}}\left(\int_{\mathcal{H}^y}D|p(f\circ\phi)|d\lambda_{\mathcal{H}}^y\right)d\mu_{\mathcal{H}}(y)\\ &=\int_{\mathcal{G}^{(0)}}\left(\int_{\mathcal{G}^{x}}|(p\circ\phi^{-1})f|d\lambda_{\mathcal{G}}^x\right)d(\phi_*\mu_{\mathcal{H}})(x)\\ &=\int_{\mathcal{G}^{(0)}}\left(\int_{\mathcal{G}^{x}}|(p\circ\phi^{-1})(s)|\left(\frac{d(\phi_*\mu_{\mathcal{H}})}{d\mu_{\mathcal{G}}}(x)\right)|f(s)|d\lambda_{\mathcal{G}}^x(s)\right)d\mu_{\mathcal{G}}(x)\\ &=\int_{\mathcal{G}^{(0)}}\left(\int_{\mathcal{G}^{x}}|(p\circ\phi^{-1})(s)|\left(\frac{d(\phi_*\mu_{\mathcal{H}})}{d\mu_{\mathcal{G}}}(\ra(s))\right)|f(s)|d\lambda_{\mathcal{G}}^x(s)\right)d\mu_{\mathcal{G}}(x)\\ &=\int_{\mathcal{G}}|p\circ\phi^{-1}|\left(\frac{d(\phi_*\mu_{\mathcal{H}})}{d\mu_{\mathcal{G}}}\circ\ra\right)|f(s)|d(\lambda_{\mathcal{G}}\circ\mu_{\mathcal{G}}) \end{align*}
and since $\Vert Tf\Vert_{\mathcal{H}}=\Vert f\Vert_{\mathcal{G}}=\int_{\mathcal{G}}|f|d(\lambda_{\mathcal{G}}\circ\mu_{\mathcal{G}})$, we obtain
\begin{equation*} |p\circ\phi^{-1}|=\frac{d\mu_{\mathcal{G}}}{d(\phi_*\mu_{\mathcal{H}})}\circ\ra\qquad(\lambda_{\mathcal{G}}\circ\mu_{\mathcal{G}})\text{-a.e.}\tag{3.8} \end{equation*}
Since $p$ is a morphism then $p(\mathcal{H}^{(0)})=\left\{1\right\}$, which, along with Equation (3.8) and continuity of $p$, yields $|p|=|p\circ\ra|=1$ on $\mathcal{H}$. The same Equation (3.8) then also implies $\mu_{\mathcal{G}}=\phi_*\mu_{\mathcal{H}}$.

Remark

In the case of groups, the same type of classification was first proven by Wendel in [MR0049910], when considering the whole $L^1$-algebras of locally compact Hausdorff groups instead of only algebras of compactly supported continuous functions. Further generalizations of Wendel's Theorem were proven in [MR0177058] and [MR0193531], and closely results in [MR0160846] and [MR0361622].

3.9. $(I,\ra)$-Groupoid convolution algebras

In the next result we will again use the convolution algebras of topological groupoids, however now we will consider another norm, which was already defined in the work of Hahn ([MR496797]) and played an important role in Renault's work ([MR584266]). A locally compact Hausdorff groupoid is étale if the range map $\ra\colon\mathcal{G}\to\mathcal{G}^{(0)}$ is a local homeomorphism. From this, it follows that $\mathcal{G}^{(0)}$ is open in $\mathcal{G}$, that the product map is open and that $\mathcal{G}^x$ is discrete for all $x\in\mathcal{G}^{(0)}$ (see [MR2304314]).

Let $\mathcal{G}$ be a locally compact étale Hausdorff groupoid, $\mathbb{K}=\mathbb{R}$ or $\mathbb{C}$ and $\theta=0$, and let $\lambda$ be a Haar system for $\mathcal{G}$. Again, we will consider the convolution algebra $C_c(\mathcal{G})=C_c(\mathcal{G},\mathbb{K})$ as defined in the previous subsection.

Every left Haar system on an étale groupoid is essentially the counting measure ([MR584266, 2.7]), in the sense that for all $x,y\in\mathcal{G}^{(0)}$, the map $a\mapsto \lambda^{\ra(a)}(\left\{a\right\})$ is constant on set $\mathcal{G}_x^y$.

We define the $(I,\ra)$-norm on $C_c(\mathcal{G})$ as

\begin{equation*} \Vert f\Vert_{I,\ra}=\sup_{x\in\mathcal{G}^{(0)}}\int |f|d\lambda^x. \end{equation*}

As $\mathcal{G}$ is Hausdorff, the unit space $\mathcal{G}^{(0)}$ of $\mathcal{G}$ is a closed subgroupoid of $\mathcal{G}$, hence (trivially) étale, Hausdorff and locally compact itself. The convolution product on $C_c(\mathcal{G}^{(0)})$ coincides with the pointwise product, and the $(I,\ra)$-norm is the uniform one: $\Vert f\Vert_{I,\ra}=\Vert f\Vert_\infty=\sup_{x\in\mathcal{G}^{(0)}}|f(x)|$.

Moreover, $\mathcal{G}^{(0)}$ is also open in $\mathcal{G}$ (because $\mathcal{G}$ is étale), so we can identify $C_c(\mathcal{G}^{(0)})$ with the subalgebra $\left\{f\in C_c(\mathcal{G}^{(0)}:\supp(f)\subseteq\mathcal{G}^{(0)}\right\}$ of $C_c(\mathcal{G})$.

Definition 3.26.

The algebra $C_c(\mathcal{G}^{(0)})$, identified as a subalgebra of $C_c(\mathcal{G})$, is called the diagonal subalgebra of $C_c(\mathcal{G})$. If $\mathcal{G}$ and $\mathcal{H}$ are locally compact étale Hausdorff groupoids, an isomorphism $T\colon C_c(\mathcal{G})\to C_c(\mathcal{H})$ is called diagonal-preserving if $T(C_c(\mathcal{G}^{(0)}))=C_c(\mathcal{H}^{(0)})$.

Theorem 3.27.

Let $\mathcal{G}$ and $\mathcal{H}$ be locally compact Hausdorff étale groupoids with continuous fully supported left Haar systems $\lambda_G$ and $\lambda_H$, respectively, and $T\colon C_c(\mathcal{G})\to C_c(\mathcal{H})$ a diagonal-preserving algebra isomorphism, isometric with respect to the $(I,\ra)$-norms. Then there is a (unique) topological groupoid isomorphism $\phi\colon\mathcal{H}\to\mathcal{G}$ and a continuous morphism $p\colon\mathcal{H}\to\mathbb{S}^1$ such that

\begin{equation*} Tf(h)=p(h)D(\phi(h))f(\phi(h)) \end{equation*}
where $D$ is a continuous instance of the Radon‒Nikodym derivative
\begin{equation*} D(a)=\frac{d\lambda_{\mathcal{G}}^{\ra(a)}}{d(\phi_*\lambda_{\mathcal{H}}^{\phi^{-1}(\ra(a))})}(a). \end{equation*}

Proof.

By the Banach‒Stone Theorem (3.18), there is a homeomorphism $\phi\colon\mathcal{H}^{(0)}\to\mathcal{G}^{(0)}$ and a continuous function $P\colon\mathcal{H}^{(0)}\to\mathbb{S}^1$ such that $Tf(y)=P(y)f(\phi(y))$ for all $f\in C_c(\mathcal{G}^{(0)})$ and $y\in\mathcal{H}^{(0)}$. Since $T$ is multiplicative we obtain $P=1$. (The same conclusion can be obtained in a similar manner by Milgram's or Jarosz' Theorem.)

For each $x\in\mathcal{G}^{(0)}$, let $\left\{a^x_i:i\in I_x\right\}$ be a net of functions in $C_c(\mathcal{G}^{(0)})$ satisfying:

  1. $0\leq a^x_i\leq 1$, and $a^x_i(x)=1$;

  2. $\bigcap_i\supp(a^x_i)=\left\{x\right\}$;

  3. If $j\geq i$ then $[a^x_j\neq 0]\subseteq[a^x_i\neq 0]$.

Items (ii)-(iii) and compactness of each $\supp(a^x_i)$ imply that $\left\{[a^x_i\neq 0]:i\in I_x\right\}$ is a neighbourhood basis at $x$. For $y\in\mathcal{H}^{(0)}$, let $a^y_i=T(a^{\phi(y)}_i)=a^{\phi(y)}_i\circ\phi$, so that the net $\left\{a^y_i:i\in I_{\phi(y)}\right\}$ satisfies (i)-(iii) as well.

Continuity of $\lambda_G$ implies that for all $x\in \mathcal{G}^{(0)}$ and $f\in C_c(\mathcal{G})$, $\lim_{i\in I_x}\Vert a^x_i f\Vert_{I,\ra}=\int |f(\gamma)|d\lambda^x(\gamma)$, and similarly on $\mathcal{H}$.

Given $f,g\in C_c(\mathcal{G})$, we use Lemma 3.19 to obtain

\begin{align*} f\perp g&\iff\forall x\left(f|_{\mathcal{G}^x}\perp g|_{\mathcal{G}^x}\text{ in }C(\mathcal{G}^x)\right)\\ &\iff\forall x\forall A,B(\lim\Vert a^x_i(Af+Bg)\Vert_{I,\ra}=\lim(|A|\Vert a^x_i f\Vert_{I,\ra}+|B|\Vert a^x_i g\Vert_{I,\ra} \end{align*}
and the last condition is preserved by $T$, so by Jarosz' Theorem $T$ is of the form $Tf(\alpha)=\widetilde{P}(\alpha)f(\widetilde{\phi}(\alpha))$ for a certain homeomorphism $\widetilde{\phi}\colon\mathcal{H}\to\mathcal{G}$ and a non-vanishing continuous scalar function $\widetilde{P}$. We can readily see that $\widetilde{\phi}$ and $\widetilde{P}$ are extensions of $\phi$ and $P$, respectively, so instead let us simply denote $\widetilde{\phi}=\phi$ and $\widetilde{P}=P$.

The proof that $\phi$ is a groupoid isomorphism, and that $P$ can be decomposed as $P=(D\circ\phi)p$ for the (continuous) Radon‒Nikodym derivative $D$ and some continuous morphism $p\colon\mathcal{H}\to\mathbb{C}\setminus\left\{0\right\}$ is the same as in Theorem 3.25, but the verification that $|p|=1$ is different.

Given $y\in\mathcal{H}^{(0)}$ and $f\in C_c(\mathcal{G})$, using the definition of $D$ as a Radon‒Nikodym derivative,

\begin{equation*} \int_{\mathcal{H}^y}|Tf|d\lambda_{\mathcal{H}}^y=\int_{\mathcal{H}^y} |p|(D\circ\phi)|f\circ\phi|d\lambda_{\mathcal{H}}^y=\int_{\mathcal{G}^{\phi^{-1}(y)}} |p\circ\phi^{-1}||f|d\lambda_{\mathcal{G}}^{\phi(y)}. \end{equation*}
Considering again the functions $a^{\phi(y)}_i$ and $a^y_i$, and the fact that $T$ is isometric we obtain
\begin{align*} \int_{\mathcal{G}^{\phi^{-1}(y)}} |p\circ\phi^{-1}||f|d\lambda_{\mathcal{G}}^{\phi(y)}&=\lim_{i\in I_{\phi(y)}}\Vert a_i^y Tf\Vert_{I,\ra}=\lim_{i\in I_{\phi(y)}}\Vert a_i^{\phi(y)} f\Vert_{I,\ra}= \int |f|d\lambda_{\mathcal{G}}^{\phi(y)} \end{align*}
for all $f\in C_c(\mathcal{G})$, which implies that $|p|=1$ $\lambda_{\mathcal{H}}^y$-a.e. Since $p$ is continuous and $\lambda_{\mathcal{H}}$ is fully supported, we conclude that $|p|=1$ on $\mathcal{H}$.

3.10. Steinberg Algebras

Steinberg algebras were independently introduced in [MR2565546] and [MR3274831], as algebraic analogues of groupoid C*-algebras, and are generalizations of Leavitt path algebras and universal inverse semigroup algebras. We refer to [MR3743184] and [carlsenrout2017] for more details.

A locally compact, zero-dimensional étale groupoid is called ample. A bisection of a groupoid $\mathcal{G}$ is a subset $A\subseteq\mathcal{G}$ such that the source and range maps are injective on $A$. If $\mathcal{G}$ is an ample Hausdorff groupoid, we denote by $\KB(\mathcal{G})$ the semigroup of compact-open bisections of $\mathcal{G}$, which forms a basis for the topology of $\mathcal{G}$.

In this section, $R$ is a fixed commutative ring with unit. Given an ample Hausdorff groupoid $\mathcal{G}$, we denote by $R^{\mathcal{G}}$ the $R$-module of $R$-valued functions on $\mathcal{G}$. Given $A\subseteq\mathcal{G}$, we define $1_A$ as the characteristic function of $A$ (with values in $R$).

Steinberg algebras were The goal of this section is to prove that the Steinberg algebra of an ample Hausdorff groupoid $\mathcal{G}$ together with its diagonal algebra completely characterize $\mathcal{G}$. Although the main theorem of this subsection (Theorem 3.41) is partially stated and proven (for more general graded Steinberg algebras) in [carlsenrout2017, Corollary 3.14], we can obtain a precise classification of the diagonal-preserving isomorphisms of Steinberg algebras, as described in Theorem 3.41 and Corollary 3.42

We will need to recover the bisections of $\mathcal{G}$ from $A_R(\mathcal{G})$, and in particular the compact-open subsets of $\mathcal{G}^{(0)}$. The main idea is, again, to identify subsets of $\mathcal{G}^{(0)}$ with their characteristic functions, and these are precisely the functions which attain only the values $0$ and $1$. We thus need to assume an extra condition on the ring $R$.

Definition 3.28 ([MR1878556, X.7]).

A (nontrivial) commutative unital ring $R$ is indecomposable if its only idempotents are $0$ and $1$. Equivalently, $R$ is indecomposable if it cannot be written as a direct sum $R\simeq R_1\oplus R_2$, where $R_1$ and $R_2$ are nontrivial rings.

A subset $A$ of a groupoid $\mathcal{G}$ is a bisection if and only if $AA^{-1}\cup A^{-1}A\subseteq \mathcal{G}^{(0)}$. A similar type of condition will be used to recover an ample Hausdorff groupoid $\mathcal{G}$ from the pair $(A_R(\mathcal{G}),D_R(\mathcal{G}))$.

Definition 3.29.

A normalizer of $D_R(\mathcal{G})$ is an element $f\in A_R(\mathcal{G})$ for which there exists $g\in A_R(\mathcal{G})$ such that

  1. $fD_R(\mathcal{G})g\subseteq D_R(\mathcal{G})$ and $gD_R(\mathcal{G})f\subseteq D_R(\mathcal{G})$;

  2. $fgf=f$ and $gfg=g$.

We denote by $N_R(\mathcal{G})$ the set of normalizers of $D_R(\mathcal{G})$. An element $g$ satisfying (i) and (ii) above will be called and inverse of $f$ relative to $D_R(\mathcal{G})$.

It can be verified that $N_R(\mathcal{G})$ is a multiplicative subsemigroup of $A_R(\mathcal{G})$, which is moreover an inverse semigroup. In particular, the inverse relative to $D_R(\mathcal{G})$ of an element $f\in N_R(\mathcal{G})$ is unique. However, we will not necessitate these results.

Example 3.30.

If $A\in\KB(\mathcal{G})$ then $1_A\in N_R(\mathcal{G})$. More generally, if $\lambda_1,\ldots,\lambda_n$ are invertible elements in $R$ and $U_1,\ldots,U_n$ are compatible disjoint compact-open bisections (that is, $\bigcup_i U_i$ is also a bisection), then $f=\sum_i \lambda_i1_{U_i}$ is a normalizer of $D_R(\mathcal{G})$. The unique inverse of $f$ relative to $D_R(\mathcal{G})$ is given by $f^*=\sum_i\lambda_i^{-1}1_{U_i^{-1}}$, that is, $f^*(a)=f(a^{-1})^{-1}$ for all $a\in\supp(f)^{-1}$.

In order to recover $\mathcal{G}$ from $(A_R(\mathcal{G}),D_R(\mathcal{G}))$, we need that all normalizers of $D_R(\mathcal{G})$ have the form described in the example above, so additional conditions will have to be assumed on the groupoids we consider.

The following property was considered in [arxiv1711.01903v2], when working on the same recovery problem.

Definition 3.31.

If $\mathcal{G}$ is an ample Hausdorff groupoid and $R$ is an indecomposable (commutative, unital) ring, we say that $(\mathcal{G},R)$ satisfies the local bisection hypothesis if $\supp(f)$ is a bisection for all $f\in N_R(\mathcal{G})$.

Lemma 3.32.

Suppose that $(\mathcal{G},R)$ satisfies the local bisection hypothesis and $f\in N_r(\mathcal{G})$. Then for all $a\in\supp(f)$, $f(a)$ is invertible in $R$.

Proof.

Let $g$ be an inverse of $f$ relatively to $D_R(\mathcal{G})$. First note that $fg=f1_{\so(\supp(f))}g\in D_R(\mathcal{G})$.

Let $a\in\supp(f)$. Since $fg$ is an idempotent in $D_R(\mathcal{G})$, the product in $D_R(\mathcal{G})$ is pointwise and $R$ is indecomposable, then $fg(\ra(a))\in\left\{0,1\right\}$. Moreover, as $\supp(f)$ is a bisection we have $f(a)=fgf(a)=(fg)(\ra(a))f(a)$, so $(fg)(\ra(a))=1$.

Again using that $\supp(f)$ is bisection, we obtain

\begin{equation*} 1=fg(\ra(a))=f(a)g(a^{-1}) \end{equation*}
so $f(a)$ is invertible in $R$.

The following stronger condition was considered in [carlsenrout2017], and is more easily checked than the one above. Recall that if $R$ is a ring and $G$ is a group, a trivial unit of the group-ring $RG$ is an element of the form $ug$, where $u$ is invertible in $R$ and $g\in G$.

Definition 3.33.

If $\mathcal{G}$ is an ample Hausdorff groupoid and $R$ is an indecomposable (commutative, unital) ring, we say that $(\mathcal{G},R)$ satisfies condition (S) if the set of all $x\in\mathcal{G}^{(0)}$ such that the group ring $R\mathcal{G}_x^x$ has only trivial units is dense in $\mathcal{G}^{(0)}$.

The property of a group-ring $RG$ (where $G$ is a group and $R$ is a ring) having only trivial units has been studied, for example, in [MR0002137]. A group $G$ is indexed if there exists a non-trivial group morphism from $G$ to $\mathbb{Z}$, and indicable throughout if every nontrivial finitely generated subgroup of $G$ is indexed. (Note that if $G$ is indicable throughout then $G$ is torsion-free.)

Theorem 3.34 ([MR0002137, Theorem 13]).

If $G$ is indicable throughout and $R$ is an integral domain, then $RG$ has only trivial units.

Every free group, and every torsion-free abelian group is indicable throughout. The class of indicable throughout groups is closed under products, free products and extensions (see [MR0002137]).

The following result from [carlsenrout2017] provides a large class of groupoids satisfying the local bisection hypothesis. Although in [carlsenrout2017] the authors assume stronger hypotheses (namely, that $R$ is an integral domain and $R\mathcal{G}_x^x$ does not have zero divisors for all $x$ in a dense subset of $\mathcal{G}^{(0)}$), their proof works under the weaker assumptions we adopt.

Lemma 3.35 ([carlsenrout2017, Lemma 3.5(2)]).

Suppose that $\mathcal{G}$ is an ample Hausdorff groupoid, $R$ is an indecomposable ring and that $(\mathcal{G},R)$ satisfies condition (S). Then $(\mathcal{G},R)$ satisfies the local bisection hypothesis.

An important class of groupoids consists of the topologically principal ones, whose associated algebras have been extensively studied (see, for example, [MR3189105], MR2745642], MR1681679], MR2590626]). In fact it is possible to classify C*-algebras which come from them (see [MR2460017]).

Definition 3.36.

A topological groupoid $\mathcal{G}$ is topologically principal if the set of all $x\in X$ whose isotropy group $\mathcal{G}_x^x$ is trivial is dense in $\mathcal{G}^{(0)}$.

It follows that if $\mathcal{G}$ is an ample Hausdorff topologically principal groupoid and $R$ is an indecomposable ring, then $(\mathcal{G},R)$ satisfies the local bisection hypothesis.

We are ready to classify diagonal-preserving isomorphisms of Steinberg algebras of groupoids and rings satisfying the local bisection hypothesis. For this, let us first define the class of maps of interest:

Definition 3.37.

Let $R$ and $S$ be rings and $\mathcal{G}$ be a groupoid. Denote by $\operatorname{Iso}_+(R,S)$ the set of additive isomorphisms from $R$ to $S$. A map $\chi\colon\mathcal{G}\to\operatorname{Iso}_+(R,S)$ satisfying $\chi(ab)(rs)=\chi(a)(r)\chi(b)(s)$ for all $(a,b)\in\mathcal{G}^{(2)}$ and $r,s\in R$ will be called a cocycle.

Example 3.38.

Consider $C_2=\left\{1,g\right\}$, the group of order 2, acting on itself by left multiplication and consider the transformation groupoid $\mathcal{G}=C_2\ltimes C_2$. Let $R=S=\mathbb{Z}$. If we define $\chi\colon\mathcal{G}\to\operatorname{Iso}_+(R,S)$ by $\chi(1,y)(r)=r$ and $\chi(g,r)=-r$, then $\chi$ is a cocycle. Note that $\chi(g,1)$ is not a ring isomorphism.

Example 3.39.

Suppose $R$ is a unital ring and $\chi\colon\mathcal{G}\to\operatorname{Iso}_+(R,R)$ is a cocycle. Then $\chi$ is a morphism from the groupoid $\mathcal{G}$ to the group (under composition) $\operatorname{Iso}_+(R,R)$ if, and only if, $\chi(x)=\id_R$ for all $x\in \mathcal{G}^{(0)}$.

Proposition 3.40.

Let $R$ and $S$ be commutative unital rings, $\mathcal{G}$ a groupoid and $\chi\colon\mathcal{G}\to\operatorname{Iso}_+(R,S)$ a cocycle. Then

  1. For all $x\in\mathcal{G}^{(0)}$, $\chi(x)$ is a ring isomorphism;

  2. For all $a\in\mathcal{G}$, if $u\in R$ is invertible then $\chi(a)(u)$ is invertible in $S$, and $\chi(a)(u)^{-1}=\chi(a^{-1})(u)$.

  3. For all $a\in\mathcal{G}$, $\chi(\so(a))=\chi(\ra(a))$. In other words, the restriction of $\chi$ to $\mathcal{G}^{(0)}$ is invariant.

Proof.

The cocycle condition states that $\chi(ab)(rs)=\chi(a)(r)\chi(b)(s)$ for all $a,b,r,s$. Taking $a=b=x$ yields (a). Taking $b=a^{-1}$, $r=u$ and $s=u^{-1}$ yields (b), and for item (c) we use commutativity of $S$:

\begin{equation*} \chi(\so(a))(r)=\chi(a^{-1}a)(1r)=\chi(a^{-1})(1)\chi(a)(r)=\chi(a)(r)\chi(a^{-1})(1)=\chi(\ra(a))(r). \end{equation*}

We endow $\operatorname{Iso}_+(R,S)$ with the topology of pointwise convergence, so that a map $\chi$ from a topological space $X$ to $\operatorname{Iso}_+(R,S)$ is continuous if and only if for every $r\in R$, the map $X\ni x\mapsto\chi(x)(r)\in S$ is continuous, that is, locally constant.

Theorem 3.41.

Let $\mathcal{G}$ and $\mathcal{H}$ be ample Hausdorff groupoids. Let $R$ and $S$ be two indecomposable (commutative, unital) rings such that $(\mathcal{G},R)$ and $(\mathcal{H},S)$ satisfy the local bisection hypothesis. Let $T\colon A_R(\mathcal{G})\to A_S(\mathcal{H})$ be a diagonal-preserving ring isomorphism, that is, $T(D_R(\mathcal{G}))=D_S(\mathcal{H})$.

Then there exists a unique topological groupoid isomorphism $\phi\colon\mathcal{H}\to\mathcal{G}$ and a continuous cocycle $\chi\colon\mathcal{H}\to\operatorname{Iso}_+(R,S)$ such that $Tf(a)=\chi(a)(f(\phi(a)))$ for all $a\in\mathcal{H}$ and $f\in A_R(\mathcal{G})$.

Proof.

Since $T$ preserves the respective diagonal algebras, it also preserves their normalizers, i.e., $T(N_R(\mathcal{G}))=N_S(\mathcal{H})$. Let us describe disjointness first for elements in $N_R(\mathcal{G})$. The local bisection hypothesis implies, by Lemma 3.32, that an element $f$ of $N_R(\mathcal{G})$ has the form

\begin{equation*} f=\sum_{i=1}^n\lambda_i1_{U_i} \end{equation*}
where $\lambda_1,\ldots,\lambda_n$ are invertible elements in $R$ and $U_1,\ldots,U_n$ are disjoint compact-open bisections of $\mathcal{G}$ such that $\bigcup_{i=1}^n U_i=\supp(f)$ is also a compact-open bisection. A similar statement holds for $N_S(\mathcal{H})$.

If $f,g\in N_R(\mathcal{G})$, then $f\subseteq g$ if and only if $f=g p$ for some $p\in D_R(\mathcal{G})$: Indeed,

Therefore $T$ preserves inclusion of normalizers. Since $N_R(\mathcal{G})$ contains $\left\{1_U:U\in\KB(\mathcal{G})\right\}$ then it is regular (Definition 1.5), because $\KB(\mathcal{G})$ is a basis for the topology of $\mathcal{G}$. Hence $T$ also preserver disjointness of normalizers (Proposition 1.8).

To prove that $T$ preserves disjointness in all of $A_R(\mathcal{G})$, we decompose elements of $A_R(\mathcal{G})$ in terms of elements of $N_R(\mathcal{G})$ and $D_R(\mathcal{G})$: if $f,g\in A_R(\mathcal{G})$, then $f\perp g$ if and only if there are finite collections of normalizers $f_i,g_j\in N_R(\mathcal{G})$ and elements $\widetilde{f_i},\widetilde{g_j}\in D_R(\mathcal{G})$ ($1\leq i\leq n$, $1\leq j\leq m$) such that

\begin{equation*} f=\sum_i f_i\widetilde{f_i},\quad g=\sum_j g_j\widetilde{g_j}\text{ and }f_i\perp g_j\text{ for all }i,j \end{equation*}
Indeed, if there are such $f_i,g_j,\widetilde{f_i},\widetilde{g_j}$ then $\supp(f)\subseteq\bigcup_i\supp(f_i)$ and $\supp(g)\subseteq\bigcup_j\supp(g_j)$, and the latter sets are disjoint.

Conversely, we write $f=\sum_i\lambda_i1_{A_i}$, where the $A_i$ are pairwise disjoint compact-open bisections and $\lambda_i\neq 0$, and take $f_i=1_{A_i}$ and $\widetilde{f_i}=\lambda_i1_{\so(A_i)}$, so that $\supp(f)=\bigcup_{i=1}^n\supp(f_i)$. Similarly, writing $g=\sum_j\widetilde{g_j}g_j$ where $g_j\in N_R(\mathcal{G})$ and $\supp(g)=\bigcup_j\supp(g_j)$, then $f\perp g$ implies $f_i\perp g_j$ for all $i$ and $j$.

Therefore, $T$ is a $\perp$-isomorphism. Note that $\perpp$ and $\perp$ coincide in $A_R(\mathcal{G})$, since its elements are locally constant (similarly to Example 1.4). Then $T$ is a $\perpp$-isomorphism, so let $\phi\colon\mathcal{H}\to\mathcal{G}$ be the $T$-homeomorphism. The verification that $\phi$ is a groupoid isomorphism is similar to that of Theorem 3.25.

Since elements of $A_R(\mathcal{G})$ (and $A_S(\mathcal{H})$) are locally constant, then for all $f\in A_R(\mathcal{G})$,

\begin{equation*} f(\phi(a))=0\iff \phi(a)\in Z(f)\iff x\in Z(Tf)\iff Tf(a)=0. \end{equation*}
and therefore $T$ is basic (by additivity of $T$ and Proposition 2.8). Let $\chi$ be the $T$-transform. Since $T$ is additive with the pointwise operations, each section $\chi(\alpha)=\chi(\alpha,\cdot)$ is additive (by Proposition 2.7). This yields a map $\chi\colon\mathcal{G}\to\operatorname{Iso}_+(R,S)$, and we need now to verify that $\chi$ is a cocycle.

If $(a,b)\in\mathcal{H}^{(2)}$ and $r,s\in R$, choose compact-open bisections $U,V$ of $\mathcal{G}$ containing $\phi(a)$ and $\phi(b)$, respectively. Then using multiplicativity of $T$ we obtain

\begin{align*} \chi(ab)(rs)&=\chi(ab)\big((r1_U)(s1_V)(\phi(ab))\big)=T\big((r1_U)(s1_V)\big)(ab)\\ &=\big(T(r1_U)T(s1_V)\big)(ab)=\sum_{cd=ab}T(r1_U)(c)T(s1_V)(d)\\ &=\sum_{cd=ab}\chi(c)\big(r1_U(\phi(c))\big)\chi(d)\big(s1_V(\phi(d))\big) \end{align*}
If $cd=ab$ is such that the last term above is nonzero, then $\ra(c)=\ra(a)$ and $\phi(c)\in U$, so since $U$ is a bisection we obtain $a=c$. Similarly, $d=b$, therefore $\chi(ab)(rs)=\chi(a)(r)\chi(b)(s)$, and $\chi$ is a cocycle.

It remains only to prove that $\chi$ is continuous: Let $r\in R$ be fixed, $a\in\mathcal{H}$ and $U$ any compact-open bisection containing $\phi(a)$. For all $b\in\phi^{-1}(U)$,

\begin{equation*} \chi(b)(r)=\chi(b)(r1_U(\phi(b)))=T(r1_U)(b) \end{equation*}
which means that the map $b\mapsto\chi(b)(r)$ coincides with $T(r1_U)$ on $\phi^{-1}(U)$ and thus it is continuous.

We should note that according to [arxiv1711.01903v2], the local bisection hypothesis is preserved by diagonal-preserving isomorphisms, so the same result is valid if we assume, in principle, that only $(\mathcal{G},R)$ satisfies this condition.

From this we can immediately classify the group of diagonal-preserving automorphisms of Steinberg algebras satisfying the local bisection hypothesis. Let $\mathcal{G}$ be a groupoid and $R$ a ring. Denote by $\operatorname{Coc}(\mathcal{G},R)$ the set of all continuous cocycles $\chi\colon\mathcal{G}\to\operatorname{Iso}_+(R,R)$, which is a group with the canonical (pointwise) structure: $(\chi\rho)(a)=\chi(a)\circ\rho(a)$ for all $\chi,\rho\in C(\mathcal{G},R)$ and $a\in\mathcal{G}$, where $\circ$ denotes composition.

Let $\operatorname{Aut}(\mathcal{G})$ be the group of topological groupoid automorphisms of $\mathcal{G}$. Then $\operatorname{Aut}(\mathcal{G})$ acts on $\operatorname{Coc}(\mathcal{G},R)$ in the usual (dual) manner: for $\phi\in\operatorname{Aut}(\mathcal{G})$, $\chi\in\operatorname{Coc}(\mathcal{G},R)$ and $a\in\mathcal{G}$ set $(\phi\chi)(a)=\chi(\phi^{-1}a)$.

Denote by $\operatorname{Aut}(A_R(\mathcal{G}),D_R(\mathcal{G}))$ the group of diagonal-preserving ring automorphisms of $A_R(\mathcal{G})$. From Theorem 3.41 we immediately obtain:

Corollary 3.42.

If $(\mathcal{G},R)$ satisfies the local bisection hypothesis, then the group $\operatorname{Aut}(A_R(\mathcal{G}),D_R(\mathcal{G}))$ is isomorphic to the semidirect product $C(\mathcal{G},R)\rtimes\operatorname{Aut}(\mathcal{G})$.

3.11. Groups of circle-valued functions

A natural question in C*-algebra theory is whether we can extend isomorphisms of unitary groups of C*-algebras to isomorphisms (or anti/conjugate-isomorphisms) of the whole C*-algebras. Dye proved in [MR0066568] that this is always possible for continuous von Neumann factors, however this is not true in the general C*-algebraic case, even in the commutative case. Therefore we should consider isomorphisms between unitary groups which preserve more structure than just the product, such as an analogue to that of Theorem 3.15. Recall that the unitary group of a commutative C*-algebra $C(X)$, where $X$ is compact Hausdorff, is $C(X,\mathbb{S}^1)$.

Theorem 3.43.

Let $X$ and $Y$ be two Stone (zero-dimensional, compact Hausdorff) spaces. Suppose that $T\colon C(X,\mathbb{S}^1)\to C(Y,\mathbb{S}^1)$ is a group isomorphism such that $1\in f(X)\iff 1\in Tf(X)$. Then there exist a homeomorphism $\phi\colon Y\to X$, a finite isolated subset $F\subseteq Y$ and a continuous function $p\colon Y\setminus F\to\{\pm 1\}$ satisfying $Tf(y)=f(\phi(y))^{p(y)}$ for all $y\in Y\setminus F$.

In particular, if $X$ (and/or $Y$) do not have isolated points then $F=\varnothing$.

The following lemma, based on [MR3162258], will be crucial to the proof of the theorem.

Lemma 3.44.

Suppose that $X$ is a Stone space. For every pair of continuous functions $f,g\colon X\to\mathbb{S}^1$ and for every finite subset $F\subseteq X$ such that $f(F)\cup g(F)$ does not contain $1$, there exists $h\in C(X,\mathbb{S}^1)$ such that

\begin{equation*} h(x)\not\in\left\{f(x),g(x)\right\}\text{ for all }x\text{ and }h(F)=\{1\}. \end{equation*}

Proof.

For every point $y\in F$, choose a clopen set $U_y$ containing $y$ such that $f(U_y)\cup g(U_y)$ does not contain $1$. For every other point $x\in X':=X\setminus\bigcup_{y\in F}U_y$, there is a clopen set $U\subseteq X'$ such that $f(U)\cup g(U)\neq\mathbb{S}^1$. Using compactness of $X'$ and taking complements and intersections if necessary we can find a clopen partition $U_1,\ldots,U_n$ of $X'$ such that $f(U_i)\cup g(U_i)\neq\mathbb{S}^1$ for all $i$. Simply choose $z_i\in \mathbb{S}^1\setminus(f(U_i)\cup g(U_i))$ and define $h=z_i$ on $U_i$, and $h=1$ on $\bigcup_{f\in F}U_f$.

Proof of Theorem 3.43.

For the notion of support we will use (Definition 1.1), we take $\theta=1$, the constant function at $1$, so regularity of $C(X,\mathbb{S}^1)$ is immediate.

Suppose that $f\perp g$ but that $Tf$ and $Tg$ are not disjoint. By Lemma 3.44, there exists $H\in C(Y,\mathbb{S}^1)$ such that $H\neq Tf,Tg$ everywhere, but that $1\in H(Y)$. Let $h=T^{-1}H$. Then $T(f^{-1}h)$ and $T(g^{-1}h)$ do not attain $1$, which implies that $f^{-1}h$ and $g^{-1}h$ do not attain $1$ as well. Thus

\begin{align*} h^{-1}(1)&=X\cap h^{-1}(1)=(f^{-1}(1)\cup g^{-1}(1))\cap h^{-1}(1)\\ &=(g^{-1}(1)\cap h^{-1}(1))\cup(f^{-1}(1)\cap h^{-1}(1))\subseteq (g^{-1}h)^{-1}(1)\cup (f^{-1}h)^{-1}(1)=\varnothing. \end{align*}
But $(Th)^{-1}(1)=H^{-1}(1)$ is nonempty, contradicting the given property of $T$.

Therefore $f\perp g$ implies $Tf\perp Tg$, and the same argument yields the opposite implication, so $T$ is a $\perp$-isomorphism. Let $\mathcal{A}(X)$ and $\mathcal{A}(Y)$ be the subgroups of order-$2$ elements of $C(X,\mathbb{S}^1)$ and $C(Y,\mathbb{S}^1)$, respectively (i.e., the groups of continuous functions with values in $\left\{-1,1\right\}$).

$\mathcal{A}(X)$ and $\mathcal{A}(Y)$ are also regular, since $X$ and $Y$ are zero-dimensional, and the restriction $T|_{\mathcal{A}(X)}\colon\mathcal{A}(X)\to \mathcal{A}(Y)$ is a $\perpp$-isomorphism, because $\perp$ and $\perpp$ coincide on $\mathcal{A}(X)$ and $\mathcal{A}(Y)$. Let $\phi\colon Y\to X$ be the corresponding $T|_{\mathcal{A}(X)}$-homeomorphism.

Let $h\in C(X,\mathbb{S}^1)$ be arbitrary. Since $\sigma(h)=\bigcup_{a\in \mathcal{A}(X),a\subseteq h}\sigma(a)$ and $T$ is a $\perp$-isomorphism, we obtain by Theorem 1.16 that, for all $h\in C(X,\mathbb{S}^1)$,

\begin{equation*} \phi(\sigma(Th))=\bigcup_{\substack{a\in \mathcal{A}(X)\\a\subseteq h}}\phi(\sigma(Ta))=\bigcup_{\substack{a\in \mathcal{A}(X)\\a\subseteq h}}\sigma(a)=\sigma(h) \end{equation*}
Since $\phi$ is a homeomorphism it preserves closures, from which it follows that $T$ is also a $\perpp$-isomorphism, and $\phi$ is also the $T$-homeomorphism.

Claim: $f(\phi(y))=1\iff Tf(y)=1$.

Suppose $f(\phi(y))\neq 1$. Choose a function $g\in C(X,\mathbb{S}^1)$ which coincides with $f$ on a neighbourhood of $\phi(y)$ and such that $1\not\in g(X)$. Then $1\not\in Tg(Y)$ and since $Tf$ coincides with $Tg$ on a neighbourhood of $y$ then $Tf(\phi(y))=Tg(\phi(y))\neq 1$. The other direction is analogous, and thus we have proved the claim.

Therefore $T$ is basic. Let $\chi$ be the $T$-transform, so that each section $\chi(y,\cdot)$ is an automorphism of the circle. If $\chi(y,\cdot)$ is continuous then it has the form $\chi(y,z)=z^{p(y)}$ where $p(y)\in\left\{\pm1\right\}$. Let us prove that for all except finitely many $y\in Y$, the section $\chi(y,\cdot)$ is continuous.

Let $F=\left\{y\in Y:\chi(y,\cdot)\text{ is discontinuous}\right\}$, and suppose that $F$ were infinite. By Proposition 2.9, there are countably infinitely many distinct points $y_n\in F$ ($n\in\mathbb{N}$), such that no $y_n$ lies in the closure of the other ones. We can choose a sequence $z_n\to 1$ such that $\chi(y_n,z_n)$ lies in the second quadrant of the circle. Define $f(\phi(y_n))=z_n$, $f=1$ on the boundary of $\{\phi(y_n):n\in\mathbb{N}\}$ and extend $f$ continuously to all of $X$. Let $y$ be a cluster point of $\{y_n\}_n$, so that in particular $f(\phi(y))=1$. Then

\begin{equation*} Tf(y)=\chi(y,1),\quad Tf(y_n)=\chi(y_n,z_n) \end{equation*}
But $y$ is an accumulation point of the $y_n$, and $Tf(y_n)$ lies in the second quadrant while $Tf(y)=1$, a contradiction to the continuity of $Tf$.

Therefore $F$ is finite, so now we show that it is open in order to conclude that its points are isolated in $Y$. Let $y\in Y$ and choose $z_0\in\mathbb{S}^1$ of the form $z_0=e^{it}$ where $-\pi/4\leq t\leq \pi/4$, but such that $\chi(y,z_0)$ is in the second or third quadrant, so in particular it is not $z_0$ nor $z_0^{-1}$. Denote by $z_0$ the constant function at $z_0$, we that

\begin{equation*} T(z_0)(y)=\chi(y,z_0)\neq z_0,z_0^{-1}. \end{equation*}
Since $T(z_0)$ is continuous, there is a neighbourhood $U$ of $y$ such that $\chi(x,z_0)\neq z_0,z_0^{-1}$ for all $x\in U$, so $x\in F$.

Therefore $Y'=Y\setminus F$ is also compact, and we already constructed the function $p\colon Y'\to\{\pm 1\}$ with the desired property. To see that $p$ is continuous, denote by $i$ the constant function $x\mapsto i$ and note that

\begin{equation*} p^{-1}(1)=\left\{y\in Y':\chi(y,i)=i\right\}=\left\{y\in Y':T(i)(y)=i\right\}=T(i)^{-1}(i)\cap Y' \end{equation*}
and similarly $p^{-1}(-1)=T(i)^{-1}(-i)\cap Y'$, so these two sets, which are complementary in $Y'$, are closed and hence clopen.

Example 3.45.

As an easy example where the subset $F\subseteq Y$ in the previous theorem is nonempty, let $X=Y=\left\{\ast\right\}$ be (equal) singletons, and let $t\colon\mathbb{S}^1\to\mathbb{S}^1$ be a discontinuous automorphism of $\mathbb{S}^1$.

Consider the map $T\colon C(X,\mathbb{S}^1)\to C(Y,\mathbb{S}^1)$, $T(f)(\ast)=t(f(\ast))$ (in other words, $T$ is the function obtained from $t$ by identifying $C(X,\mathbb{S}^1)$ and $C(Y,\mathbb{S}^1)$ with $\mathbb{S}^1$). Then $T$ satisfies the hypotheses of the previous theorem but $F=Y$.

We now endow $C(X,\mathbb{S}^1)$ with the uniform metric:

\begin{equation*} d_\infty(f,g)=\sup_{x\in X}|f(x)-g(x)| \end{equation*}
(which is the metric coming from the C*-algebra $C(X,\mathbb{C})$).

Theorem 3.46.

If $X$ and $Y$ are as above and $T\colon C(X,\mathbb{S}^1)\to C(Y,\mathbb{S}^1)$ is an isometric isomorphism, then there is a homeomorphism $\phi\colon Y\to X$ and a continuous function $p\colon Y\to\left\{\pm 1\right\}$ such that $Tf(y)=f(\phi(y))^{p(y)}$ for all $y\in Y$.

Proof.

We identify each $\lambda\in\mathbb{S}^1$ with the corresponding constant map on $X$ or $Y$. The constant function $-1$ is characterized by the following two properties:

Thus $T(-1)=-1$. A function $f$ does not attain $1$ if and only if $d_\infty(-1,f)\smallerthan 1$, so $T$ preserves functions not attaining $1$, and we apply Theorem 3.43 (or more precisely its proof) in order to obtain a homeomorphism $\phi\colon Y\to X$, a function $\chi\colon Y\times\mathbb{S}^1\to\mathbb{S}^1$ and a continuous function $p\colon Y'\to\left\{-1,1\right\}$, where $Y'=\left\{y\in Y:\chi(y,\cdot)\text{ is continuous}\right\}$, such that

\begin{equation*} Tf(y)=\chi(y,f(\phi(y))\qquad\text{and}\qquad\chi(y',t)=t^{p(y')} \end{equation*}
for all $y\in Y$, $y'\in Y'$ and $f\in C(X,\mathbb{S}^1)$. It remains only to prove that $Y'=Y$, i.e., every section $\chi(y,\cdot)$ is continuous.

If $\lambda_i\to\lambda$ in $\mathbb{S}^1$ then we also have uniform convergence of the corresponding constant functions, so

\begin{equation*} \chi(y,\lambda_i)=T(\lambda_i)y\to T(\lambda)y=\chi(y,\lambda) \end{equation*}
thus $\chi(y,\cdot)$ is continuous for all $y$.