1. Introduction
It is a challenging problem to study the structure of the space $\mathcal{X}$ of isomorphism classes of metric measure spaces, where we assume a metric measure space to be a complete separable metric space with a Borel probability measure. Denote by ${\mathbb{R}}_+$ the multiplicative group of positive real numbers. We have the natural group action of ${\mathbb{R}}_+$ on $\mathcal{X}$ defined as
where tX is the space X with the t-scaled metric of X. Note that the isomorphism class of a one-point metric measure space, denoted as $*$, is the only fixed-point of this action. Let $\mathcal{X}_* := \mathcal{X} \setminus \{*\}$ and let Σ denote the quotient space $\mathcal{X}_*/{\mathbb{R}}_+$ equipped with the quotient topology.
As for the structure of the space $\mathcal{X}$, Sturm [Reference Sturm19] obtained the remarkable result that the subspace $\mathcal{X}_{pq}$ of $\mathcal{X}$ with finite L pq-size and equipped with the $L^{p,q}$-distortion metric is a non-negatively curved Alexandrov space isometric to a Euclidean cone for p = 2 and $q \in [\,1,+\infty\,)$. He also determined geodesics in $\mathcal{X}_{pq}$ for $p,q \in [\,1,+\infty\,)$ and proved that any orbit of the ${\mathbb{R}}_+$-action is a geodesic ray, which implies that $\mathcal{X}_{pq}$ is homeomorphic to the cone over $\Sigma_{pq}$ for any $p,q \in [\,1,+\infty\,)$, where $\Sigma_{pq}$ is the subspace of Σ corresponding to $\mathcal{X}_{pq}$.
Also, Ivanov and Tuzhilin [Reference Ivanov and Tuzhilin8] pointed out that the Gromov–Hausdorff space is homeomorphic to the cone over the quotient space by the ${\mathbb{R}}_+$-action.
In this article, we study the topological structure of $\mathcal{X}$ with the box and observable metrics and also of the pyramidal compactification of $\mathcal{X}$. Those metrics and the pyramidal compactification are fundamental concepts in the study of metric measure spaces, originally introduced by Gromov [Reference Gromov6] (see also [Reference Shioya18]). The box metric is closely related to the $L^{0,q}$-distortion metric (see [Reference Sturm19]) and coincides with the Gromov–Prokhorov metric (see [Reference Greven, Pfaffelhuber and Winter5, Reference Löhr12]). The observable metric is defined by the idea of the concentration of measure phenomenon established by Lévy and Milman [Reference Lévy13, Reference Milman15] (see also [Reference Ledoux14]) and is useful to study convergence of spaces with dimensions divergent to infinity. The topologies induced from the box and observable metrics are called the box and concentration topologies, respectively. In our previous article [Reference Kazukawa, Nakajima and Shioya11], we have proved that the box metric is geodesic and that $\mathcal{X}$ is locally path-connected and contractible with respect to the box and concentration topologies, which are the same properties as those of $\mathcal{X}_{pq}$ with the $L^{p,q}$-distortion metric studied by Sturm [Reference Sturm19]. However, the box and observable metrics are much different from the $L^{p,q}$-distortion metric. One of the essential differences is that any orbit of the ${\mathbb{R}}_+$-action is never a geodesic ray for the box and observable metrics. Also, there are intricately branching geodesics with respect to the box metric, and the Alexandrov curvature is not bounded neither from below nor from above (see [Reference Kazukawa, Nakajima and Shioya11, remark 6.7]). The question that arises here is:
• Is $\mathcal{X}$ with the box and/or concentration topologies homeomorphic to the cone over Σ?
Surprisingly the answer is negative.
Theorem 1.1 For neither of the box nor concentration topologies, $\mathcal{X}_* = \mathcal{X} \setminus \{*\}$ is not homeomorphic to ${\mathbb{R}}_+ \times \Sigma$, and $\mathcal{X}$ is not homeomorphic to the cone over Σ.
One of our main theorems is stated as follows.
Theorem 1.2 For the box and concentration topologies, the action of ${\mathbb{R}}_+$ on $\mathcal{X}_*$ admits the structure of a non-trivial and locally trivial principal ${\mathbb{R}}_+$-bundle over Σ.
In general, a locally trivial principal bundle with contractible fibre over a paracompact Hausdorff base space is trivial (see [Reference Dold3, corollary 2.8] and [Reference Husemoller7, 8.1 theorem of Chapter 4] for example), which is not necessarily true if the base space is not paracompact. It is remarkable that our principal bundle ${\mathbb{R}}_+ \to \mathcal{X}_* \to \Sigma$ presents such a counterexample.
The action of ${\mathbb{R}}_+$ on $\mathcal{X}$ naturally extends to the pyramidal compactification of $\mathcal{X}$, say Π (see § 2.3 for the definition of Π). Denote by $\operatorname{Fix}(\Pi)$ the set of fixed-points of the action of ${\mathbb{R}}_+$ on Π, and put $\Pi_* := \Pi \setminus \operatorname{Fix}(\Pi)$. We also have the following.
Theorem 1.3 The action of ${\mathbb{R}}_+$ on $\Pi_*$ admits the structure of a non-trivial and locally trivial principal ${\mathbb{R}}_+$-bundle over the quotient space $\Pi_*/{\mathbb{R}}_+$.
We investigate the structure of $\operatorname{Fix}(\Pi)$ and have the following theorem. Denote by $\mathcal{A}$ the set of all monotone non-increasing sequences of non-negative real numbers with total sum not greater than 1.
Theorem 1.4 The fixed-point set $\operatorname{Fix}(\Pi)$ is homeomorphic to $\mathcal{A}$ with the l 2-weak topology.
As for the topology of $\mathcal{X}_*$ and $\Pi_*$, we have the following.
(i) $\mathcal{X}_*$ is contractible with respect to the box and concentration topologies.
(ii) $\Pi_*$ is contractible.
Since the fibres of our bundles in theorems 1.2 and 1.3 are contractible, we have the following.
(i) For the box and concentration topologies, all homotopy groups of Σ vanish.
(ii) All homotopy groups of $\Pi_*/{\mathbb{R}}_+$ vanish.
Let us mention the ideas of our proofs.
A key point of the proof of theorem 1.1 is to prove that Σ is not a Urysohn space (lemma 3.2). If $\Sigma \times {\mathbb{R}}_+$ were to be homeomorphic to $\mathcal{X}_*$, then Σ would be metrizable, which is contrary to the non-Urysohn property of Σ. It is quite delicate that Σ is a Hausdorff space (proposition 3.10).
For theorems 1.2 and 1.3, the local triviality of the bundles is a core of the proof.
For theorem 1.2 with the box topology, we construct an ${\mathbb{R}}_+$-invariant open covering $\{\mathcal{X}_\Delta\}_{\Delta \in (\,0,1\,)}$ of $\mathcal{X}$ and continuous 1-homogeneous functions $r_\Delta \colon \mathcal{X}_\Delta \to {\mathbb{R}}_+$ for $\Delta \in (\,0,1\,)$. We define $\mathcal{X}_\Delta$ to be the set of mm-spaces such that any atom has measure less than Δ and define $r_\Delta(X)$, $X \in \mathcal{X}_\Delta$, as the integral of the partial diameter $\operatorname{diam}(X;s)$ with respect to the parameter $s \in [\,0,(\Delta+1)/2\,]$. The reason why we take the integral is for the sake of the continuity of $r_\Delta$. Using $r_\Delta$, we obtain a local trivialization $\mathcal{X}_\Delta \simeq \mathcal{X}_\Delta/{\mathbb{R}}_+ \times {\mathbb{R}}_+$.
Theorem 1.2 for the concentration topology is derived from theorem 1.3 just by restricting the base space $\Pi_*/{\mathbb{R}}_+$ to $\mathcal{X}_*/{\mathbb{R}}_+$.
For the proof of theorem 1.3, we construct an ${\mathbb{R}}_+$-invariant open covering of $\Pi_*$ and continuous 1-homogeneous functions on each open set in the covering. This time, for an $(N+1)$-tuple $\kappa = (\kappa_0,\dots,\kappa_N)$ of positive real numbers with $\sum_{i=0}^N \kappa_i \lt 1$, we define $\Pi_\kappa$ to be the set of all $\mathcal{P} \in \Pi$ such that $\operatorname{Sep}(\mathcal{P};\kappa_0,\dots,\kappa_N) \lt +\infty$ and $\operatorname{Sep}(\mathcal{P};\kappa_0+\delta,\dots,\kappa_N+\delta) \gt 0$ for some δ > 0, and we define $r_\kappa(\mathcal{P})$, $\mathcal{P} \in \Pi_\kappa$, to be the integral of $\operatorname{Sep}(\mathcal{P};\kappa_0+s,\dots,\kappa_N+s)$ with respect to $s \in [\,0,1\,]$ (see definition 2.17 for the definition of $\operatorname{Sep}(\cdots)$). These induce a local trivialization $\Pi_\kappa \simeq \Pi_\kappa/{\mathbb{R}}_+ \times {\mathbb{R}}_+$. However, it is not easy to prove that the union of all $\Pi_\kappa$ coincides with $\Pi_*$. For the proof, we need to investigate the structure of pyramids in $\operatorname{Fix}(\Pi)$ as follows. For $A = \{a_i\}_{i=1}^\infty \in \mathcal{A}$, we define
Theorem 1.7 For a given $\mathcal{P} \in \Pi$, the following (i)–(iv) are equivalent to each other.
(i) $\mathcal{P} \in \operatorname{Fix}(\Pi)$.
(ii) $t\mathcal{P} = \mathcal{P}$ for some $t \in {\mathbb{R}}_+$ with t ≠ 1.
(iii) For any $\kappa_0,\dots,\kappa_N \gt 0$ with $\sum_{i=0}^N \kappa_i \lt 1$, the separation distance $\operatorname{Sep}(\mathcal{P};\kappa_0,\dots,\kappa_N)$ is either 0 or $+\infty$.
(iv) There exists $A \in \mathcal{A}$ such that $\mathcal{P} = \mathcal{P}_A$.
In theorem 1.7, the implication ‘$(\mathrm {iii}) \Rightarrow (\mathrm {iv})$’ is highly non-trivial, and we need a delicate discussion to prove it. Theorem 1.7 with a little discussion implies that the union of $\Pi_\kappa$ coincides with $\Pi_*$.
Theorem 1.4 is derived from theorem 1.7 and the following.
Theorem 1.8 The map $\mathcal{A} \ni A \longmapsto \mathcal{P}_A \in \Pi$ is a topological embedding.
Theorem 1.8 is also proved by Esaki–Kazukawa–Mitsuishi [Reference Esaki, Kazukawa and Mitsuishi4] independently. Our proof is simpler than [Reference Esaki, Kazukawa and Mitsuishi4]. It is proved in [Reference Esaki, Kazukawa and Mitsuishi4] that the weak topology on $\mathcal{A}$ coincides with the $l^\infty$-topology.
The organization of this article is as follows. After the preliminaries section, we study in § 3 the scale-change action of ${\mathbb{R}}_+$ on $\mathcal{X}_*$. We prove that Σ is not Urysohn, which leads to theorem 1.1. We also prove theorems 1.2 and 1.3 with the help of theorem 1.7. In § 4, we determine the structure of pyramids in $\operatorname{Fix}(\Pi)$ and prove theorems 1.7 and 1.8 to obtain theorem 1.4. In § 5, we prove proposition 1.5. In § 6, we present several questions.
2. Preliminaries
In this section, we describe the definitions and some properties of metric measure space, the box distance, the observable distance, the pyramid, and the weak topology. We use most of these notions along [Reference Shioya18]. As for more details, we refer to [Reference Shioya18] and [Reference Gromov6, Chapter 3$\frac{1}{2}_+$].
2.1. Metric measure spaces
Let $(X, d_X)$ be a complete separable metric space and µX a Borel probability measure on X. We call the triple $(X, d_X, \mu_X)$ a metric measure space, or an mm-space for short. We sometimes say that X is an mm-space, in which case the metric and the measure of X are, respectively, indicated by dX and µX.
Definition 2.1 (mm-Isomorphism)
Two mm-spaces X and Y are said to be mm-isomorphic to each other if there exists an isometry $f \colon \operatorname{supp}{\mu_X} \to \operatorname{supp}{\mu_Y}$ such that $f_* \mu_X = \mu_Y$, where $f_* \mu_X$ is the push-forward measure of µX by f. Such an isometry f is called an mm-isomorphism. Denote by $\mathcal{X}$ the set of mm-isomorphism classes of mm-spaces.
Note that an mm-space X is mm-isomorphic to $(\operatorname{supp}{\mu_X}, d_X , \mu_X)$. We assume that an mm-space X satisfies
unless otherwise stated.
Definition 2.2 (Lipschitz order)
Let X and Y be two mm-spaces. We say that X (Lipschitz) dominates Y and write $Y \prec X$ if there exists a 1-Lipschitz map $f \colon X \to Y$ satisfying $f_* \mu_X = \mu_Y$. We call the relation $\prec$ on $\mathcal{X}$ the Lipschitz order.
The Lipschitz order $\prec$ is a partial order relation on $\mathcal{X}$.
2.2. Box distance and observable distance
For a subset A of a metric space $(X, d_X)$ and for a real number r > 0, we set
where $d_X(x, A) := \inf_{a \in A} d_X(x, a)$.
Definition 2.3 (Prokhorov distance)
The Prokhorov distance $d_{\mathrm{P}}(\mu, \nu)$ between two Borel probability measures µ and ν on a metric space X is defined to be the infimum of ɛ > 0 satisfying
for any Borel subset $A \subset X$.
The Prokhorov metric $d_{\mathrm{P}}$ is a metrization of the weak convergence of Borel probability measures on X provided that X is a separable metric space.
Definition 2.4 (Ky Fan metric)
Let $(X, \mu)$ be a measure space and $(Y, d_Y)$ a metric space. For two µ-measurable maps $f,g \colon X \to Y$, we define $d_{\mathrm{KF}}^\mu (f, g)$ to be the infimum of $\varepsilon \geq 0$ satisfying
The function $d_{\mathrm{KF}}^\mu$ is a metric on the set of µ-measurable maps from X to Y by identifying two maps if they are equal to each other µ-almost everywhere. We call $d_{\mathrm{KF}}^\mu$ the Ky Fan metric.
Definition 2.5 (Parameter)
Let $I := [0,1)$ and let X be an mm-space. A map $\varphi \colon I \to X$ is called a parameter of X if φ is a Borel measurable map such that
where $\mathcal{L}^1$ is the one-dimensional Lebesgue measure on I.
Note that any mm-space has a parameter (see [Reference Shioya18, lemma 4.2]).
Definition 2.6 (Box distance)
We define the box distance $\square(X, Y)$ between two mm-spaces X and Y to be the infimum of $\varepsilon \geq 0$ satisfying that there exist parameters $\varphi \colon I \to X$, $\psi \colon I \to Y$, and a Borel subset $I_0 \subset I$ with $\mathcal{L}^1(I_0) \geq 1 - \varepsilon$ such that
for any $s,t \in I_0$.
Theorem 2.7 ([Reference Shioya18, theorem 4.10])
The box distance function $\square$ is a complete separable metric on $\mathcal{X}$.
Various distances equivalent to the box distance are defined and studied, for example, the Gromov–Prokhorov distance introduced by Greven–Pfaffelhuber–Winter [Reference Greven, Pfaffelhuber and Winter5].
Theorem 2.8 ([Reference Löhr12, theorem 3.1], [Reference Shioya18, remark 4.16])
For any two mm-spaces X and Y, we have
where $d_{\mathrm{GP}}(X,Y)$ is the Gromov–Prokhorov metric defined to be the infimum of $d_{\mathrm{P}}(\mu_X, \mu_Y)$ for all metrics on the disjoint union of X and Y that are extensions of dX and dY. In particular,
The topology induced from the box distance has historically various names, for example, the weak-Gromov topology. However, we call it the box topology in this article.
The total variation distance is useful for estimating the box distance.
Definition 2.9 (Total variation distance)
The total variation distance $d_{\mathrm{TV}}(\mu, \nu)$ of two Borel probability measures µ and ν on a topological space X is defined by
where A runs over all Borel subsets of X.
If µ and ν are both absolutely continuous with respect to a Borel measure ω on X, then
where $\frac{d\mu}{d\omega}$ is the Radon–Nikodym derivative of µ with respect to ω.
Proposition 2.10 ([Reference Shioya18, proposition 4.12])
For any two Borel probability measures µ and ν on a complete separable metric space X, we have
Given an mm-space X and a parameter $\varphi \colon I \to X$ of X, we set
which consists of Borel measurable functions on I.
Definition 2.11 (Observable distance)
We define the observable distance $d_{\mathrm{conc}}(X, Y)$ between two mm-spaces X and Y by
where $\varphi \colon I \to X$ and $\psi \colon I \to Y$ run over all parameters of X and Y, respectively, and $d_{\mathrm{H}}$ is the Hausdorff distance with respect to the metric $d_{\mathrm{KF}}^{\mathcal{L}^1}$.
Theorem 2.12 ([Reference Shioya18, proposition 5.5 and theorem 5.13])
The observable distance function $d_{\mathrm{conc}}$ is a metric on $\mathcal{X}$. Moreover, for any two mm-spaces X and Y,
We call the topology on $\mathcal{X}$ induced from $d_{\mathrm{conc}}$ the concentration topology. We say that a sequence $\{X_n\}_{n=1}^\infty$ of mm-spaces concentrates to an mm-space X if $X_n~d_{\mathrm{conc}}$-converges to X as $n \to \infty$.
2.3. Pyramid
Definition 2.13 (Pyramid)
A subset $\mathcal{P} \subset \mathcal{X}$ is called a pyramid if it satisfies the following (i) – (iii).
(i) If $X \in \mathcal{P}$ and $Y \prec X$, then $Y \in \mathcal{P}$.
(ii) For any $Y, Y' \in \mathcal{P}$, there exists $X \in \mathcal{P}$ such that $Y \prec X$ and $Y' \prec X$.
(iii) $\mathcal{P}$ is non-empty and $\square$-closed.
We denote the set of all pyramids by Π. Note that Gromov’s definition of a pyramid is only by (i) and (ii). The condition (iii) is added in [Reference Shioya18].
For an mm-space X, we define
which is a pyramid. We call $\mathcal{P}_X$ the pyramid associated with X.
We observe that $Y \prec X$ if and only if $\mathcal{P}_Y \subset \mathcal{P}_X$. Note that $\mathcal{X}$ itself is a pyramid.
We define the weak convergence of pyramids as follows. This is exactly the Kuratowski–Painlevé convergence as closed subsets of $(\mathcal{X}, \square)$ (see [Reference Kazukawa, Nakajima and Shioya11, §8]).
Definition 2.14 (Weak convergence)
Let $\mathcal{P}$ and $\mathcal{P}_n$, $n = 1, 2, \ldots$, be pyramids. We say that $\mathcal{P}_n$ converges weakly to $\mathcal{P}$ as $n \to \infty$ if the following (i) and (ii) are both satisfied.
(i) For any mm-space $X \in \mathcal{P}$, we have
\begin{equation*} \lim_{n \to \infty} \square(X, \mathcal{P}_n) = 0. \end{equation*}(ii) For any mm-space $X \in \mathcal{X} \setminus \mathcal{P}$, we have
\begin{equation*} \liminf_{n \to \infty} \square(X, \mathcal{P}_n) \gt 0. \end{equation*}
Theorem 2.15 ([Reference Shioya18, Section 6])
There exists a metric ρ on Π such that the following (i)–(iv) hold.
(i) ρ is compatible with weak convergence.
(ii) Π is ρ-compact.
(iii) The map $\mathcal{X} \ni X \mapsto \mathcal{P}_X \in \Pi$ is a 1-Lipschitz topological embedding map with respect to $d_{\mathrm{conc}}$ and ρ.
(iv) The image of $\mathcal{X}$ is ρ-dense in Π.
In particular, $(\Pi,\rho)$ is a compactification of $(\mathcal{X},d_{\mathrm{conc}})$. We call $(\Pi,\rho)$ the pyramidal compactification of $(\mathcal{X},d_{\mathrm{conc}})$. We often identify X with $\mathcal{P}_X$, and we say that a sequence of mm-spaces converges weakly to a pyramid if the associated pyramid converges weakly.
Definition 2.16 (Approximation of a pyramid)
A sequence $\{Y_m\}_{m = 1}^\infty$ of mm-spaces is called an approximation of a pyramid $\mathcal{P}$ provided that it satisfies
In particular, $\{Y_m\}_{m=1}^\infty$ converges weakly to $\mathcal{P}$ as $m \to \infty$ and $Y_m \in \mathcal{P}$ for all m.
It is known that any pyramid $\mathcal{P}$ admits an approximation (see [Reference Shioya18, lemma 7.14]).
2.4. Separation distance
The separation distance is one of the most fundamental invariants of an mm-space and a pyramid.
Definition 2.17 (Separation distance)
Let X be an mm-space. For any real numbers $\kappa_0,\kappa_1,\ldots,\kappa_N \gt 0$ with $N \geq 1$, we define the separation distance
of X as the supremum of $\min_{i\neq j} d_X(A_i,A_j)$ over all sequences of N + 1 Borel subsets $A_0,A_1,\ldots,A_N \subset X$ satisfying $\mu_X(A_i) \geq \kappa_i$ for all $i=0,1,\ldots,N$. If $\kappa_i \gt 1$ for some i, then we define $\operatorname{Sep}(X;\kappa_0,\kappa_1,\ldots,\kappa_N) := 0$. Moreover, we define the separation distance of a pyramid $\mathcal{P}$ by
The separation distance for mm-spaces is an invariant under mm-isomorphism. Note that
for any $\kappa_0,\kappa_1,\ldots,\kappa_N \gt 0$ and that $\operatorname{Sep}(\mathcal{P};\kappa_0,\kappa_1,\ldots,\kappa_N)$ is monotone non-increasing and left-continuous in κi for each $i = 0,1,\ldots,N$, and that
Theorem 2.18 ([Reference Ozawa and Shioya16, theorem 1.1], Limit formula for separation distance)
Let $\mathcal{P}$ and $\mathcal{P}_n$, $n = 1, 2, \ldots$, be pyramids. If $\mathcal{P}_n$ converges weakly to $\mathcal{P}$ as $n \to \infty$, then
for any $\kappa_0,\kappa_1,\ldots,\kappa_N \gt 0$.
3. Scale-change action
In this section, we prove theorems 1.1–1.3.
Let ${\mathbb{R}}_+ := (0,+\infty)$ be the multiplicative group of positive real numbers. We consider the scale-change action on $\mathcal{X}$;
The one-point space $*$ is the only fixed-point of this action and the set $\mathcal{X}_* := \mathcal{X} \setminus \{*\}$ is invariant. The ${\mathbb{R}}_+$-action on $\mathcal{X}_*$ is free. Let $\Sigma := \mathcal{X}_*/{\mathbb{R}}_+$ be the quotient space of $\mathcal{X}_*$ and $\pi \colon \mathcal{X}_* \to \Sigma$ the quotient map. We denote the orbit $\pi(X)$ by $[X]$.
Simultaneously, we consider the scale-change action on Π;
which is a natural extension of the action on $\mathcal{X}$. Denote by $\operatorname{Fix}(\Pi)$ the set of fixed-points of this action, and put $\Pi_* := \Pi \setminus \operatorname{Fix}(\Pi)$. Then, ${\mathbb{R}}_+$ acts on $\Pi_*$ freely.
For the proof of theorem 1.1, we need a lemma.
Lemma 3.1. Let Yɛ be the mm-space defined by $Y_\varepsilon := (\{0,1\},|\cdot|,(1-\varepsilon)\delta_0+\varepsilon\delta_1)$ for $0 \lt \varepsilon \lt 1$. Then, for any closed subset V of Σ with non-empty interior with respect to the box topology, there exists a $\delta(V) \gt 0$ such that $[Y_\varepsilon]$ belongs to V for any ɛ with $0 \lt \varepsilon \le \delta(V)$.
Proof. Let V be a closed subset $V \subset \Sigma$ with non-empty interior. Since any mm-space can be approximated by an mm-space with finite diameter, there is an mm-space X with finite diameter such that $[X]$ is an interior point of V. Suppose that $[Y_\varepsilon]$ does not belong to V. Then, since Yɛ is an element in the open set $\pi^{-1}(\Sigma \setminus V)$, there is a large number $r_\varepsilon \gt 0$ such that the mm-space Zɛ defined by
satisfies $r_\varepsilon^{-1} Z_\varepsilon \in \pi^{-1}(\Sigma \setminus V)$. Indeed, $\square(Y_\varepsilon, r_\varepsilon^{-1} Z_\varepsilon)$ is sufficiently small since $X \subset Z_\varepsilon$ is close to a one-point by scaling down Zɛ with $r_\varepsilon^{-1}d_{Z_\varepsilon}(z, X) = 1$. On the other hand, $Z_\varepsilon~\square$-converges to X as $\varepsilon \to 0+$, which implies $[Z_\varepsilon] \in V$ for ɛ > 0 small enough. This is a contradiction. Thus, $[Y_\varepsilon]$ belongs to V for every sufficiently small ɛ > 0. This completes the proof.
Lemma 3.1 implies the following.
Lemma 3.2. For the quotient of the box topology on Σ, any two distinct points in Σ cannot be separated by any closed neighbourhoods. In particular, Σ is not a Urysohn space.
Proof. We take two distinct points $[X], [X'] \in \Sigma$ and take any closed neighbourhoods V, V ʹ of $[X]$, $[X']$, respectively. Lemma 3.1 proves that $[Y_\varepsilon]$ belongs to both V and V ʹ for $0 \lt \varepsilon \le \min\{\delta(V),\delta(V')\}$. This completes the proof.
Remark 3.3. As is proved in proposition 3.10, Σ is Hausdorff. In lemmas 3.1 and 3.2, to consider closed neighbourhoods is essential.
Corollary 3.4. For the quotient of the concentration topology, Σ is not Urysohn. Moreover, $\Pi_*/{\mathbb{R}}_+$ is not Urysohn.
Proof. Since the quotient of the concentration topology is coarser than that of the box topology on Σ, lemma 3.2 implies the first statement of the corollary. Since Σ is contained in $\Pi_*/{\mathbb{R}}_+$ as a subspace, we obtain the second. This completes the proof.
Proof of theorem 1.1
Suppose that $\mathcal{X}_*$ is homeomorphic to $\Sigma \times {\mathbb{R}}_+$. Since $\mathcal{X}_*$ is a metric space, $\Sigma \simeq \Sigma \times \{1\}$ is metrizable, which contradicts lemma 3.2 for the box topology and corollary 3.4 for the concentration topology. In the same way, $\mathcal{X}$ is not homeomorphic to the cone over Σ. This completes the proof.
In the same way as above, we see the following.
Theorem 3.5 $\Pi_*$ is not homeomorphic to $(\Pi_*/{\mathbb{R}}_+) \times {\mathbb{R}}_+$.
The rest of this section is devoted to prove theorems 1.2 and 1.3.
We first assume that $\mathcal{X}$ is equipped with the box topology and prove theorem 1.2 for the box topology.
Proposition 3.6. $\pi \colon \mathcal{X}_* \to \Sigma$ is a principal ${\mathbb{R}}_+$-bundle.
Proof. We verify that if sequences $\{X_n\}_{n=1}^\infty \subset \mathcal{X}_*$ and $\{t_n\}_{n=1}^\infty \subset {\mathbb{R}}_+$ satisfy that Xn and $t_nX_n~\square$-converge to $X \in \mathcal{X}_*$ and tX, respectively, as $n \to \infty$, then tn converges to t. Since $\mathcal{X} \in \mathcal{X}_*$, there exist real numbers $\kappa_0, \ldots, \kappa_N \gt 0$ with $\sum_{i=0}^N \kappa_i \lt 1$ such that
Suppose that $t_n \not\to t$. There exist a real number δ > 0 and a subsequence $\{t_{n_k}\}_{k=1}^\infty$ such that either $t_{n_k} \gt t+\delta$ for any k or $t_{n_k} \lt t-\delta$ for any k. By applying theorem 2.18, if $t_{n_k} \gt t+\delta$, then we have
which implies the contradiction $t \geq t+\delta$. Similarly, if $t_{n_k} \lt t-\delta$, then the contradiction $t \leq t-\delta$ holds. Thus, we obtain $t_n \to t$. The proof is completed.
Let us prove the local triviality of the principal fibre bundle $\pi \colon \mathcal{X}_* \to \Sigma$. Let Δ be a real number with $0 \lt \Delta \lt 1$ and put
We remark that $\sup_{x \in X} \mu_X(\{x\}) \lt \Delta$ if and only if $\operatorname{diam}(X; \Delta) \gt 0$, where $\operatorname{diam}(X; \alpha)$ is the partial diameter, which is a fundamental invariant of an mm-space, given by
We see that $\mathcal{X}_\Delta$ is open. We have $\mathcal{X}_\Delta \subset \mathcal{X}_{\Delta'}$ for $\Delta \leq \Delta'$, and
Since $X \in \mathcal{X}_\Delta$ implies $tX \in \mathcal{X}_\Delta$ for any t > 0, the set $\mathcal{X}_\Delta$ is invariant with respect to the ${\mathbb{R}}_+$-action. Put $\Sigma_\Delta := \mathcal{X}_\Delta/{\mathbb{R}}_+$. Then, for the local triviality, it is sufficient to prove that $\mathcal{X}_\Delta$ is homeomorphic to $\Sigma_\Delta \times {\mathbb{R}}_+$ for every $\Delta \in (0,1)$.
We define a map $r_\Delta \colon \mathcal{X}_\Delta \to {\mathbb{R}}_+$ by
Note that, for any $X \in \mathcal{X}_\Delta$,
and that $r_\Delta(tX) = t \, r_\Delta(X)$ for any t > 0 since $\operatorname{diam}(tX; s) = t \operatorname{diam}(X; s)$ for any s > 0.
Lemma 3.7. ([Reference Shioya18, lemma 5.43])
If a sequence $\{X_n\}_{n=1}^\infty$ of mm-spaces $\square$-converges to an mm-space X, then we have
for any s > 0.
Lemma 3.8. The map $r_\Delta$ is continuous on $\mathcal{X}_\Delta$.
Proof. Take any sequence $\{X_n\}_{n=1}^\infty \subset \mathcal{X}_\Delta~\square$-converging to an mm-space $X \in \mathcal{X}_\Delta$. Let
for $s \in [0, \frac{\Delta+1}{2}]$. Since f is non-decreasing on $[0, \frac{\Delta+1}{2}]$, the discontinuous points of f are at most countable. Thus, lemma 3.7 implies fn converges almost everywhere to f and
Therefore, by the dominated convergence theorem,
The proof is completed.
Proof of theorem 1.2 for the box topology
The continuous 1-homogeneous map $r_\Delta \colon \mathcal{X}_\Delta \to {\mathbb{R}}_+$ induces the homeomorphism $\Phi \colon \mathcal{X}_\Delta \to \Sigma_\Delta \times {\mathbb{R}}_+$ defined by
Indeed, the inverse map Φ−1 is given by
In other words, the map $r_\Delta$ produces the continuous section
so that $\mathcal{X}_\Delta \to \Sigma_\Delta$ is trivial. This implies the local triviality of our principal fibre bundle ${\mathbb{R}}_+ \to \mathcal{X}_* \to \Sigma$.
Theorem 1.1 proves that the fibre bundle ${\mathbb{R}}_+ \to \mathcal{X}_* \to \Sigma$ is globally non-trivial. This completes the proof of theorem 1.2 for the box topology.
The following corollary is a by-product of theorem 1.2.
Corollary 3.9. There is no continuous 1-homogeneous map $r \colon \mathcal{X}_* \to {\mathbb{R}}_+$ with respect to the box topology on $\mathcal{X}_*$.
The following proposition is compared with lemma 3.2.
Proposition 3.10. Σ is a Hausdorff space.
Proof. For any distinct two points $[X], [X'] \in \Sigma$, there exists $0 \lt \Delta \lt 1$ such that $[X], [X'] \in \Sigma_\Delta$. Since $\mathcal{X}_\Delta$ and $\Sigma_\Delta \times {\mathbb{R}}_+$ are homeomorphic, $\Sigma_\Delta$ is metrizable. Thus, the two points $[X], [X']$ can be separated by neighbourhoods in Σ since $\Sigma_\Delta$ is open in Σ. The proof is completed.
Before proving theorem 1.2 for the concentration topology, we study the scale-change action on Π and prove theorem 1.3 with the help of theorem 1.7, where the proof of theorem 1.7 is deferred to the next section. The concentration case of theorem 1.2 is obtained as a corollary of theorem 1.3.
Proposition 3.11. The quotient map $\pi \colon \Pi_* \to \Pi_*/{\mathbb{R}}_+$ is a principal ${\mathbb{R}}_+$-bundle.
Proof. Assume that $\{\mathcal{P}_n\}_{n=1}^\infty \subset \Pi_*$ and $\{t_n\}_{n=1}^\infty \subset {\mathbb{R}}_+$ satisfy that $\mathcal{P}_n$ and $t_n\mathcal{P}_n$ converge weakly to $\mathcal{P} \in \Pi_*$ and $t\mathcal{P}$, respectively, as $n \to \infty$. Since $\mathcal{P} \in \Pi_*$, there exist real numbers $\kappa_0, \ldots, \kappa_N \gt 0$ with $\sum_{i=0}^N \kappa_i \lt 1$ such that
by theorem 1.7. By the same argument as in the proof of proposition 3.6, we have $t_n \to t$ as $n\to\infty$. The proof is completed.
Proof of theorem 1.3 under assuming theorem 1.7
For any tuple $\kappa = (\kappa_0, \kappa_1, \ldots, \kappa_N)$ of positive real numbers with $\sum_{i=0}^N \kappa_i \lt 1$, we define
It is obvious that $\Pi_\kappa$ is invariant under ${\mathbb{R}}_+$-action. We prove that $\Pi_\kappa$ is open. Indeed, for any sequence $\{\mathcal{P}_n\}_{n=1}^\infty$ of pyramids convergent weakly to a pyramid $\mathcal{P}$, if $\operatorname{Sep}(\mathcal{P}_n; \kappa_0,\ldots,\kappa_N) = +\infty$, then
by theorem 2.18, and if $\operatorname{Sep}(\mathcal{P}_n; \kappa_0+\delta,\ldots,\kappa_N+\delta) = 0$ for any δ > 0, then
for any δ > 0 by theorem 2.18. These imply that $\Pi_\kappa$ is open.
We define a map $r_\kappa \colon \Pi_\kappa \to {\mathbb{R}}_+$ by
Then, the map rκ is continuous and 1-homogeneous on $\Pi_\kappa$, so that $\Pi_\kappa \to \Pi_\kappa/{\mathbb{R}}_+$ is trivial.
The rest is to prove
By theorem 1.7, the inclusion $\bigcup_\kappa \Pi_\kappa \subset \Pi_*$ is obvious. To prove the reverse inclusion $\Pi_* \subset \bigcup_\kappa \Pi_\kappa$, we take any $\mathcal{P} \in \Pi_*$. By theorem 1.7, there exist real numbers $\kappa_0, \kappa_1, \ldots, \kappa_N \gt 0$ with $\sum_{i=0}^N \kappa_i \lt 1$ such that
By the left-continuity of $\operatorname{Sep}(\mathcal{P};s_0,\ldots,s_N)$ in si, there is ɛ > 0 such that
which implies $\mathcal{P} \in \Pi_{(\kappa_0-\varepsilon, \ldots, \kappa_N-\varepsilon)}$ and then $\Pi_* \subset \bigcup_\kappa \Pi_\kappa$. Thus, we obtain the local triviality of $\pi \colon \Pi_* \to \Pi_*/{\mathbb{R}}_+$. The proof is completed.
Proof of theorem 1.2 for the concentration topology
Assume that $\mathcal{X}$ is equipped with the concentration topology and consider the ${\mathbb{R}}_+$-action. Then, $\pi \colon \mathcal{X}_* \to \mathcal{X}_*/{\mathbb{R}}_+$ is the restriction of bundle $\pi \colon \Pi_* \to \Pi_*/{\mathbb{R}}_+$. Thus, it is also a principal ${\mathbb{R}}_+$-bundle and locally trivial (see [Reference Husemoller7]). This bundle is globally non-trivial because of theorem 1.1. We finish the proof.
(i) There is no continuous 1-homogeneous map $r \colon \Pi_* \to {\mathbb{R}}_+$.
(ii) There is no continuous 1-homogeneous map $r \colon \mathcal{X}_* \to {\mathbb{R}}_+$ with respect to the concentration topology on $\mathcal{X}_*$.
We also obtain the following proposition.
Proposition 3.13. Let $\mathcal{X}_{{\mathrm{NA}}}$ be the space of all non-atomic mm-spaces. Then, $\pi \colon \mathcal{X}_{{\mathrm{NA}}} \to \mathcal{X}_{{\mathrm{NA}}}/{\mathbb{R}}_+$ is a trivial bundle with respect to the box and concentration topologies.
Proof. For the box topology, since $\mathcal{X}_{{\mathrm{NA}}}$ is an ${\mathbb{R}}_+$-invariant subspace of $\mathcal{X}_\Delta$ for any Δ, we see that $\pi \colon \mathcal{X}_{{\mathrm{NA}}} \to \mathcal{X}_{{\mathrm{NA}}}/{\mathbb{R}}_+$ is the restriction of trivial bundle $\pi \colon \mathcal{X}_{\Delta} \to \Sigma_{\Delta}$. Similarly, for the concentration topology, $\mathcal{X}_{{\mathrm{NA}}}$ is a subspace of $\Pi_\kappa$ for any κ. This implies the triviality.
4. Scale-invariant pyramids
In this section, we prove theorems 1.7 and 1.8. We recall that $\mathcal{A}$ is the set of all monotone non-increasing sequences of non-negative real numbers with total sum not greater than 1. We equip $\mathcal{A}$ with the weak topology as a closed convex subset of the space l 2. In particular, $\mathcal{A}$ is compact. For every $A = \{a_i\}_{i=1}^\infty \in \mathcal{A}$, we set
We remark that $\mathcal{P}_A$ is a pyramid a priori, where the $\square$-closedness of $\mathcal{P}_A$ follows from the argument in the proof of claim 4.2 below.
4.1. Characterization of $\operatorname{Fix}(\Pi)$
In order to prove theorem 1.7, we need a lemma.
Lemma 4.1. Let $a_1, \ldots, a_k$, $k \lt +\infty$, and ɛ be positive numbers with $\sum_{i=1}^k a_i \lt 1$ and $\min_{i=1,\ldots,k}a_i \geq \varepsilon \gt 0$. If an mm-space X admits distinct k points $\{x_i\}_{i=1}^k$ with
then there exist positive numbers $\kappa_0,\ldots,\kappa_N \gt 0$ such that $\kappa_i \leq \varepsilon$ for every i,
Proof. Take any mm-space X having distinct k points $\{x_i\}_{i=1}^k$ with
Let $\{\xi_i\}_{i=1}^\infty$ be a countable dense subset of X and let $d_i := d_X(\xi_i, \cdot \,)$. Put
for some sufficiently small $\varepsilon' \gt 0$ with $\mu_X(X\setminus X_0) - \sum_{i=1}^k \mu_X(\{x_i\}) =: \eta_0 \ll \varepsilon$. We first find ɛ-atomic points $\alpha_1, \ldots, \alpha_m$ of ${d_1}_* (\mu_X\lfloor_{X_0})$, i.e.,
if these exist. We put
There exist finitely many disjoint closed intervals $I_0, \ldots, I_{l_1}$ of ${\mathbb{R}}$ such that
We put $A_i := d_1^{-1}(I_i) \cap X_0$. Note that $\mu_X(A_i) = \kappa_i$.
We next find ɛ-atomic points $\alpha_{i1}, \ldots, \alpha_{im_i}$ of ${d_2}_* (\mu_X\lfloor_{X_i})$ for $i=1,\ldots,m$ if these exist, and we set
There exist finitely many disjoint closed intervals $I_{l_1+1}, \ldots, I_{l_1+l_2}$ such that
We put $A_{j} := d_2^{-1}(I_j) \cap X_{i(j)}$ for every $j=l_1+1,\ldots,l_1+l_2$.
Repeating this construction, we obtain a monotone sequence $X^{(1)} \supset X^{(2)} \supset \cdots$ and a disjoint family $\{A_i\}$ on X.
We prove that $X^{(n)} = \emptyset$ for some n. It is sufficient to prove $\bigcap_{n=1}^\infty X^{(n)} = \emptyset$ since this implies $\mu_X(X^{(n)}) \to 0$ as $n \to \infty$, which contradicts the fact $\mu_X(X^{(n)}) \geq \varepsilon$ if $X^{(n)} \neq \emptyset$. Suppose that there exists a point $x_0 \in \bigcap_{n=1}^\infty X^{(n)} \neq \emptyset$. Then, there exists a sequence $\{i_k\}_k$ such that $x_0 \in X_{i_1i_2\cdots i_k}$ for any k, where $i_k \in \{1, \ldots, m_{i_1i_2\cdots i_{k-1}}\}$. Then it holds that
Indeed, suppose that there exists another point $y \in \bigcap_{k=1}^\infty X_{i_1i_2\cdots i_k}$ and let $r_0 := d_X(x_0, y)/4$. There exists a sufficiently large k 0 such that $d_X(\xi_{k_0}, x_0) \leq r_0$. Since $x_0, y \in X_{i_1i_2\cdots i_{k_0}}$, we have
which implies the contradiction $4r_0 = d_X(x_0, y) \leq d_X(\xi_{k_0}, x_0) + d_X(\xi_{k_0}, y) \leq 2r_0$.
Thus, we obtain
but this contradicts the assumption of this lemma.
Therefore, we have
where $N = l_1+l_2+\cdots+l_n$, and
Therefore, taking ηi with $\sum_{i=0}^n \eta_i \leq \varepsilon - \sum_{i=1}^k (\mu_X(\{x_i\})-a_i)$, we obtain the conclusion.
Proof of theorem 1.7
Since $(\mathrm {iv}) \Rightarrow (\mathrm {i}) \Rightarrow (\mathrm {ii}) \Rightarrow (\mathrm {iii})$ is trivial, we prove $(\mathrm {iii}) \Rightarrow (\mathrm {iv})$. Assume that a pyramid $\mathcal{P}$ satisfies the condition (iii). Let $\{Y_n\}_{n=1}^\infty$ be an approximation of $\mathcal{P}$ and let
be the inverse limit of $\{Y_n\}_{n=1}^\infty$. There exists a probability measure $\mu_{Y_\infty}$ such that
for any n, where $\pi_n \colon Y_\infty \to Y_n$ is the projection (see [Reference Choksi2]). Note that $Y_\infty$ admits an extended metric that can take values in $[0, +\infty]$ and πn is 1-Lipschitz.
Let $\{y_i\}_{i=1}^M \subset Y_\infty$, $M \leq +\infty$, be the sequence of atomic points of $\mu_{Y_\infty}$ and let
for every i. By relabeling, we can assume that
and then $A := \{a_i\}_{i=1}^\infty \in \mathcal{A}$. Then, for any n, we have
Claim 4.2.
It holds that
Proof. Take any $X \in \mathcal{P}$. Then there exist Borel maps $f_n \colon Y_n \to X$ such that fn is 1-Lipschitz up to ɛn with non-exceptional domain $\tilde{Y}_n$, that is, $\mu_{Y_n}(\tilde{Y}_n) \geq 1-\varepsilon_n$ and
for any $y,y'\in\tilde{Y}_n$, and $d_{\mathrm{P}}({f_n}_* \mu_{Y_n}, \mu_X) \leq \varepsilon_n$ for some $\varepsilon_n \to 0$ (see [Reference Kazukawa10, lemma 4.6]).
We prove that, for each $i = 1,2,\ldots,M$, the sequence $\{f_n \circ \pi_n(y_i)\}_{n=1}^\infty \subset X$ has a convergent subsequence. Suppose that $\{f_n \circ \pi_n(y_i)\}_{n=1}^\infty$ has no convergent subsequence. Then, there exist a real number η > 0 and a subsequence $\{n_k\}_{k=1}^\infty$ of $\{n\}$ such that $\{B_\eta(f_{n_k} \circ \pi_{n_k}(y_i))\}_{k=1}^\infty$ is a disjoint family. Since $\pi_{n_k}(y_i) \in \tilde{Y}_{n_k}$ for sufficiently large k and
we have
for sufficiently large k, which contradicts the disjointness of $\{B_\eta(f_{n_k} \circ \pi_{n_k}(y_i))\}_{n=1}^\infty$. Thus, $\{f_n \circ \pi_n(y_i)\}_{n=1}^\infty$ has a convergent subsequence.
Let
where $\{n_k\}$ is a subsequence of $\{n\}$ such that $\{f_{n_k} \circ \pi_{n_k}(y_i)\}_{k=1}^\infty$ converges for any $i=1,2,\ldots,M$ (by the diagonal argument). We prove
For any non-negative bounded continuous function $\varphi\colon X \to [0, +\infty)$, by Fatou’s lemma, we have
which implies $\sum_{i=1}^M a_i \delta_{x_i} \leq \mu_X$. Therefore, the proof of this claim is completed.
We next prove the converse inclusion under the condition (iii). We take any ɛ > 0 and any positive integer k such that
Then, for sufficiently large n, since $\pi_n \colon \{y_1, \ldots, y_k\} \to Y_n$ is injective, we have
Indeed, if not, then the atomic part of $\mu_{Y_\infty}$ on the inverse limit $Y_\infty$ is not equal to $\sum_{i=1}^M a_i\delta_{y_i}$. Thus, by lemma 4.1, there exist $\kappa_0,\ldots,\kappa_N \gt 0$ such that $\kappa_i \leq \varepsilon$ for every i,
for some large n. Thus, we have
Combining this and the condition (iii) implies that
By the limit formula of the separation distance, we have
Now we prove the following claim.
Claim 4.3.
It holds that
Proof. We take any mm-space X admitting a sequence $\{x_i\}_{i=1}^\infty \subset X$ such that $\sum_{i=1}^\infty a_i\delta_{x_i} \leq \mu_X$. By the approximation, we can assume that X is finite. Indeed, there exist finite nets $\mathcal{N}_n$ of X and Borel maps $\xi_n \colon X \to \mathcal{N}_n$ such that $\lim_{n\to\infty}d_{\mathrm{P}}({\xi_n}_* \mu_X, \mu_X)=0$. Then, we have
so that X can be approximated keeping our assumption. Let $\{z_1, \ldots, z_m\} := X$ and
Since $\lim_{\eta \to 0+}\lim_{n \to \infty} \operatorname{Sep}(Y_n; a_1-\eta, \ldots, a_k-\eta, \kappa_0-\eta, \ldots, \kappa_N-\eta) = \infty$, there exist Borel subsets $Y_{n,1}, \ldots, Y_{n,N+k+1} \subset Y_n$ for any sufficiently large n such that
for some $\eta \lt (N+k+1)^{-1}\varepsilon$. We define a Borel map $g_n \colon Y_n \to X$ satisfying
where
Letting $\tilde{Y}_n := \bigcup_{i=1}^{j_{m}+k} Y_{n,i}$, by the definition of gn, we have
for any $l=1,\ldots,m$. In particular, $1-\mu_{Y_n}(\tilde{Y}_n) \leq 2m\varepsilon$. Moreover, for any $B \subset X$, we have
Thus, gn is 1-Lipschitz up to $2m\varepsilon$ with non-exceptional domain $\tilde{Y}_n$ and $d_{\mathrm{P}}({g_n}_* \mu_{Y_n}, \mu_X) \leq 2m\varepsilon$. Since $Y_n \in \mathcal{P}$, taking ɛ → 0, we obtain $X \in \mathcal{P}$ (see [Reference Kazukawa10, corollary 4.7]).
We finish the proof of this theorem.
4.2. Topological structure of $\operatorname{Fix}(\Pi)$
The goal here is to prove theorem 1.8.
Lemma 4.4. The map $\mathcal{A} \ni A \mapsto \mathcal{P}_A \in \operatorname{Fix}(\Pi)$ is injective.
Proof. Take any $A=\{a_i\}_{i=1}^\infty, A'=\{a'_i\}_{i=1}^\infty \in \mathcal{A}$ with $A \neq A'$. There exists a number k such that $a_i=a'_i$ for any i < k and $a_k \neq a'_k$. We can assume that $a_k \lt a'_k$. An mm-space X is defined as the unit interval $([0,1], |\cdot|)$ with probability measure
where ${\mathcal{L}}^1$ is the Lebesgue measure on $[0,1]$. Then, we have $X \in \mathcal{P}_A$ and $X \not\in \mathcal{P}_{A'}$. Indeed, if $X \in \mathcal{P}_{A'}$, then there exists a sequence $\{x_i\}_{i=1}^\infty \subset X$ such that $\sum_{i=1}^\infty a'_i\delta_{x_i} \leq \mu_X$. Since A is monotone non-increasing, we have
which is a contradiction. Therefore, we obtain $\mathcal{P}_A \neq \mathcal{P}_{A'}$. This completes the proof.
Lemma 4.5. The map $\mathcal{A} \ni A \mapsto \mathcal{P}_A \in \operatorname{Fix}(\Pi)$ is continuous.
Proof. Assume that $A_n = \{a_{ni}\}_{i=1}^\infty \in \mathcal{A}$ converges weakly to $A = \{a_i\}_{i=1}^\infty \in \mathcal{A}$. Let us prove that $\mathcal{P}_{A_n}$ converges weakly to $\mathcal{P}_A$.
We first prove that $\lim_{n\to\infty} \square(X, \mathcal{P}_{A_n}) = 0$ for any $X \in \mathcal{P}_A$. By the standard approximation, X can be assumed to be a finite mm-space. Take any ɛ > 0 and find a number k such that $a_{k+1} \lt \varepsilon$. Then, for sufficiently large n, we have
Since An is a monotone non-increasing sequence, $a_{n,k+1} \lt \varepsilon$ implies $\sup_{i \gt k} a_{ni} \lt \varepsilon$. Take such large n and fix it. Let $\{x_i\}_{i=1}^\infty$ be a sequence in X such that $\sum_{i=1}^\infty a_i\delta_{x_i} \leq \mu_X$ and let $\{y_1,\ldots,y_N\} := X$. We define
and define two maps $\varphi\colon\tilde{X}\to X$ and $\psi\colon X \to\tilde{X}$ by
We now define two probability measures $\mu_{\tilde{X}}$ and $\mu_{\tilde{X}_n}$ on $\tilde{X}$ as
and $\mu_{\tilde{X}_n}\lfloor_{\{\eta_1, \ldots, \eta_N\}}$ is determined as follows. Find finitely many real numbers $b_{n1},\ldots,b_{nM} \in [0,\varepsilon)$ with
and set
Note that $\sup_{j} c_{nj} \lt \varepsilon$. We define
for $i=1,\ldots,N$, where $j_0 := 0$, $j_N := +\infty$, and
for $i=1,\ldots,N-1$. Under $\inf{\emptyset} = +\infty$, if there exists $i_0 \lt N$ such that $j_{i_0} = \cdots =j_N = +\infty$, then we understand
Letting $i_0 := \min\left\{1\leq i \leq N \, | \, j_i=+\infty \right\}$, we have
for any $i \lt i_0$ by the definition. On the other hand, since
we have
These imply that
Hence, we have
Therefore, letting $X_n := (X, d_X, \varphi_* \mu_{\tilde{X}_n})$, we obtain $X_n \in \mathcal{P}_{A_n}$ and
which imply $\lim_{n\to\infty} \square(X, \mathcal{P}_{A_n}) = 0$.
We next prove that $\liminf_{n\to\infty} \square(X, \mathcal{P}_{A_n}) \gt 0$ for any $X \in \mathcal{X}\setminus\mathcal{P}_A$. It is sufficient to prove that if $X_n \in \mathcal{P}_{A_n}~\square$-converges to X, then we have $X \in \mathcal{P}_A$, due to considering the contraposition and extracting a subsequence. Assume that $X_n \in \mathcal{P}_{A_n}~\square$-converges to X. Let $\{x_{ni}\}_{i=1}^\infty$ be a sequence in Xn with
There exist Borel maps $f_n \colon X_n \to X$ and a sequence $\varepsilon_n \to 0$ such that fn is 1-Lipschitz up to ɛn and $d_{\mathrm{P}}({f_n}_* \mu_{X_n}, \mu_X) \leq \varepsilon_n$ (actually, fn can be assumed to be an ɛn-mm-isomorphism but this is unnecessary here). The sequence $\{f_n(x_{ni})\}_{n=1}^\infty$ has a convergent subsequence by the same argument as in the proof of claim 4.2. Let
where $\{n_k\}$ is a subsequence of $\{n\}$ such that $\{f_{n_k}(x_{n_ki})\}_{k=1}^\infty$ converges for any i. Then, since An converges weakly to A and ${f_n}_* \mu_{X_n}$ converges weakly to µX, we have
for any non-negative bounded continuous function $\varphi\colon X \to [0, +\infty)$, where the first inequality follows from Fatou’s lemma. This implies $\sum_{i=1}^\infty a_i \delta_{x_i} \leq \mu_X$, so that $X \in \mathcal{P}_A$. The proof of this lemma is now completed.
Proof of theorem 1.8
By lemmas 4.4 and 4.5, the map $\mathcal{A} \ni A \mapsto \mathcal{P}_A \in \operatorname{Fix}(\Pi)$ is a continuous bijection from the compact space $\mathcal{A}$ to the Hausdorff space $\operatorname{Fix}(\Pi)$. Thus, this is a homeomorphism. The proof is completed.
5. Contractibility of total and base spaces
In this section, we study the topology of the total space $\mathcal{X}_* = \mathcal{X} \setminus \{*\}$ and the base space $\Sigma = \mathcal{X}_*/{\mathbb{R}}_+$, and prove proposition 1.5. Since $\mathcal{X}_*$ and Σ (resp. $\Pi_*$ and $\Pi_*/{\mathbb{R}}$) are weakly homotopy equivalent to each other, proposition 1.5 implies corollary 1.6 directly.
Proof of proposition 1.5(i)
Let Z be an arbitrary mm-space with at least two different points and let $p \in [1,\infty]$. We define a map $H \colon [0,1] \times \mathcal{X}_* \to \mathcal{X}_*$ by
where $\times_p$ means the lp-product of two mm-spaces and we agree $0X := *$ for any $X \in \mathcal{X}$. We see that
and each mm-space $H(t, X)$ has at least two different points, that is, $H(t,X) \in \mathcal{X}_*$ for any $t \in [0,1]$ and $X\in\mathcal{X}_*$. Since the scaling and the lp-product are continuous operations (see [Reference Kazukawa9]), the map H is continuous with respect to both the box and concentration topologies. Thus, H is a deformation retraction of $\mathcal{X}_*$ onto $\{Z\}$. Therefore, $\mathcal{X}_*$ is contractible.
On another note, the following proposition can be obtained by the same proof as above.
Proposition 5.1. Let $\mathcal{F}$ be the family of all finite mm-spaces. Then, $\mathcal{X}\setminus \mathcal{F}$ is contractible.
Proof. Just take Z as $([0,1], |\cdot|, {\mathcal{L}})$ in the proof of proposition 1.5(i). Then, $H(t,X)$ is a deformation retraction of $\mathcal{X}\setminus\mathcal{F}$ onto $\{Z\}$ since $H(t,X)$ is not finite.
We next prove proposition 1.5(ii) for pyramids. In order to prove this, we prepare the following lemma about the continuity of the lp-product of pyramids. As for the lp-product of pyramids, we refer to [Reference Esaki, Kazukawa and Mitsuishi4, Reference Kazukawa, Nakajima and Shioya11].
Lemma 5.2. If a sequence $\{\mathcal{P}_n\}_{n=1}^\infty$ of pyramids converges weakly to a pyramid $\mathcal{P}$, then, for any mm-space Z and any $p \in [1,\infty]$, the sequence $\{\mathcal{P}_n \times_p \mathcal{P}_Z\}_{n=1}^\infty$ converges weakly to $\mathcal{P}\times_p \mathcal{P}_Z$, where the lp-product of two pyramids $\mathcal{P}$ and $\mathcal{Q}$ is given by
Proof. The outline of this proof is similar to that of [Reference Kazukawa10, theorem 1.2]. For simplicity of the proof, we assume that $p \lt +\infty$. The case of $p=+\infty$ can be proved in the same way.
We first prove that for any $Y \in \mathcal{P} \times_p \mathcal{P}_Z$,
Note that (5.1) holds if and only if there exist mm-spaces $X_n \in \mathcal{P}_n$, $n=1,2,\ldots$, such that
It is sufficient to prove (5.1) for any mm-space Y with $Y \prec X \times_p Z$ for some $X \in \mathcal{P}$. Assume that $Y \prec X\times_p Z$ for some $X \in \mathcal{P}$. Since $\mathcal{P}_n \to \mathcal{P}$, there exists a sequence $\{X_n\}_{n=1}^\infty~\square$-converging to X such that $X_n \in \mathcal{P}_n$ for every n. Then, we have
by [Reference Kazukawa9, proposition 4.1]. Since $Y \prec X \times_p Z$, we have
by [Reference Shioya18, proposition 6.10]. Therefore, we obtain (5.1) for $Y \prec X \times_p Z$, and then for any $Y \in \mathcal{P} \times_p \mathcal{P}_Z$.
We next prove that if an mm-space Y satisfies (5.1), then $Y\in \mathcal{P} \times_p \mathcal{P}_Z$ conversely, which implies that
for any $Y \not\in \mathcal{P} \times_p \mathcal{P}_Z$ by choosing an appropriate subsequence. Assume that an mm-space Y satisfies (5.1) and assume that Y has at least two different points since $* \in \mathcal{P} \times_p \mathcal{P}_Z$ is always true. Then there exist mm-spaces $X_n \in \mathcal{P}_n$, $n=1,2,\ldots$, such that
By [Reference Kazukawa10, lemma 4.6], there exist a Borel measurable map $f_n \colon X_n \times_p Z \to Y$ and a Borel measurable subset $\tilde{Y}_n \subset X_n \times_p Z$ such that $d_{\mathrm{P}}({f_n}_* (\mu_{X_n}\otimes \mu_Z), \mu_Y) \le \varepsilon_n$, $\mu_n(\tilde{Y}_n) \ge 1-\varepsilon_n$, and
for some $\varepsilon_n \to 0$.
Take any ɛ > 0 and fix it. There exist open subsets $Y_1, \ldots, Y_N \subset Y$ such that $\operatorname{diam}{Y_j} \lt \varepsilon$ and $\mu_Y(Y_j) \gt 0$ for any $j=1, \ldots, N$,
cf. [Reference Ozawa and Yokota17, lemma 42]. Let
and take $y_j \in Y_j$ for each $j=0,1,\ldots, N$. In the case of $Y_0 =\emptyset$, it is rather easy and will be omitted. We define an mm-space $\dot{Y}$ by
Let $0 \lt \eta \ll \varepsilon$ be a small real number as
Then, there exist pairwisely disjoint open subsets $Z_1, \ldots, Z_M \subset Z$ such that $\operatorname{diam}{Z_i} \lt \eta$ and $\mu_Z(Z_i) \gt 0$ for any $i=1, \ldots, M$, and
We define subsets $A_{ij}^n \subset X_n$ for every $i=1,\ldots,M$, $j=1,\ldots,N$ by
and define a 1-Lipschitz map $\Phi_n \colon X_n \to ({\mathbb{R}}^{MN}, \|\cdot\|_\infty)$ by
for $x \in X_n$. By letting
since $\operatorname{supp}{\nu_n} \subset \{w\in {\mathbb{R}}^{MN} \mid \|w\|_\infty \le \operatorname{diam}{\dot{Y}}\}$, the sequence $\{\nu_n\}_{n=1}^\infty$ is tight. Thus, we can assume that $\{\nu_n\}_{n=1}^\infty$ converges weakly to a Borel probability measure ν on ${\mathbb{R}}^{MN}$.
We now define an mm-space X by
Note that $X \in \mathcal{P}$ because $({\mathbb{R}}^{MN}, \|\cdot\|_\infty, \nu_n) \prec X_n \in \mathcal{P}_n$ for every n. Our goal is to construct a Borel map $\Psi \colon X \times_p Z \to \dot{Y}$ such that $d_{\mathrm{P}}(\Psi_*(\mu_X \otimes \mu_Z), \mu_{\dot{Y}}) \lt 2\varepsilon$ and
for some Borel subset $\tilde{Y} \subset X \times_p Z$ with $\mu_X\otimes \mu_Z(\tilde{Y}) \gt 1-2\varepsilon$. Indeed, if such a map Ψ exists, then we have $Y \in \mathcal{P}_{X \times_p Z}$ by applying [Reference Kazukawa10, lemma 4.6] to the composition $\iota \circ \Psi$ of the inclusion map $\iota\colon\dot{Y} \to Y$ and the map $\Psi \colon X \times_p Z \to \dot{Y}$. Therefore, we obtain
since $\mathcal{P}_{X \times_p Z} \subset \mathcal{P} \times_p \mathcal{P}_Z$.
For each $i \in \{1,\ldots,M\}$, we set $\mathrm{proj}_i \colon {\mathbb{R}}^{MN} \to {\mathbb{R}}^N$ the projection given by
for $(w_{kl})_{k,l} \in {\mathbb{R}}^{MN}$. Put
We see that $\{{\Phi_{n,i}}_* \mu_{X_n}\}_{n=1}^\infty$ converges weakly to νi on ${\mathbb{R}}^N$ as $n \to \infty$. For each $j \in \{1,\ldots,N\}$, we define a closed subset $W_j \subset {\mathbb{R}}^N$ by
We now prepare two claims.
Claim 5.3.
For any sufficient large n and for any $i=1,\ldots,M$, $j=1,\ldots,N$, we have
Proof. Take any $x \in A_{ij}^n$. We have
We prove that for any $j'\neq j$,
By the definition of δ, it is clear that $\delta \le \delta' \le \operatorname{diam}{\dot{Y}}$. We take any $x' \in A_{ij'}^n$. Then there exist $z, z' \in Z_i$ such that
Thus, we have
which implies that $d_{X_n}(x,x') \ge \delta$ if n is large as $\varepsilon_n \lt \delta'/2$. Therefore, we have $d_{X_n}(x, A_{ij'}^n) \ge \delta$. This completes the proof.
Claim 5.4.
For any $j = 1,\ldots, N$, we have
Proof. A straight-forward calculation implies
where the fourth inequality follows from claim 5.3.
We define a Borel map $\Psi \colon X \times_p Z \to \dot{Y}$ by
First, we prove that
For any $j=1,\ldots,N$, we have
by claim 5.4. On the other hand, we have
Therefore, we have
for any $j=1,\ldots,N$. Taking $\mu_{\dot{Y}}(\{y_0\}) \lt \varepsilon$ and $N\eta \lt \varepsilon$ into account, we obtain (5.2).
Let
We see that
Finally, we prove that
Take any $(x,z),(x',z') \in \tilde{Y}$ and find $i,i' \in \{1,\ldots,M\}$ and $j,j' \in \{1,\ldots,N\}$ such that $\mathrm{proj}_i(x) \in W_j$, $z \in Z_i$, $\mathrm{proj}_{i'}(x') \in W_{j'}$, and $z' \in Z_{i'}$. There exist $x_n \in A_{ij}^n$ and $x'_n \in A_{i'j'}^n$, $n=1,2,\ldots$, such that
as $n \to \infty$. There exists $z_n \in Z_i$ such that $(x_n, z_n) \in f_n^{-1}(Y_j) \cap \tilde{Y}_n$. For any $\tilde{x}_n \in A_{i'j'}^n$ and $\tilde{z}_n \in Z_{i'}$ with $(\tilde{x}_n, \tilde{z}_n) \in f_n^{-1}(Y_{j'}) \cap \tilde{Y}_n$, we have
which implies that
Thus, we have
This implies (5.4). Thus, the map $\Psi\colon X \times_p Z \to \dot{Y}$ satisfies (5.2), (5.3), and (5.4), so that we obtain $Y \in \mathcal{P}_{X \times_p Z} \subset \mathcal{P} \times_p \mathcal{P}_Z$. Therefore, the sequence $\{\mathcal{P}_n \times_p \mathcal{P}_Z\}_{n=1}^\infty$ converges weakly to the pyramid $\mathcal{P} \times_p \mathcal{P}_Z$.
Remark 5.5. We conjecture that if two sequences $\{\mathcal{P}_n\}_{n=1}^\infty$ and $\{\mathcal{Q}_n\}_{n=1}^\infty$ of pyramids converge weakly to pyramids $\mathcal{P}$ and $\mathcal{Q}$, respectively, then $\{\mathcal{P}_n \times_p \mathcal{Q}_n\}_{n=1}^\infty$ converges weakly to $\mathcal{P} \times_p \mathcal{Q}$ as $n\to\infty$ for all $p\in [1,\infty]$. Furthermore, this may be true for the setting that extends [Reference Kazukawa9]. However, we need some new idea/proof that does not rely on the partition of the fixed mm-space Z.
Proof of proposition 1.5(ii)
Let Z be an arbitrary mm-space with at least two different points and let $p \in [1,\infty]$. We define a map $H \colon [0,1] \times \Pi_* \to \Pi_*$ by
where $F_t(s) := \min\{s, \frac{1}{t}-1\}$ for $s \ge 0$, which is a monotone metric preserving function, and $F_t(\mathcal{P})$ for a pyramid $\mathcal{P}$ is given by $F_1(\mathcal{P}) := \mathcal{P}_*$ and
for $t \in [0,1)$. It is a known fact that $F_0(\mathcal{P}) = \mathcal{P}$ and the map $(t,\mathcal{P}) \mapsto F_t(\mathcal{P})$ is continuous on $[0,1]\times \Pi_*$. We remark that the map $(t,\mathcal{P}) \mapsto (1-t)\mathcal{P}$ is not continuous at t = 1. See [Reference Kazukawa, Nakajima and Shioya11, §5] for more details. We see that
Moreover, $H(t, \mathcal{P}) \in \Pi_*$ since $H(t, \mathcal{P}) \neq \mathcal{P}_*$ and $H(t, \mathcal{P})$ has finite observable diameter if t > 0. We prove the continuity of H. Assume that $t_n \to t$ on $[0,1]$ and $\mathcal{P}_n \to \mathcal{P}$ on $\Pi_*$. If t > 0, then we have
as $n\to\infty$ by lemma 5.2 and [Reference Kazukawa10, corollary 1.5]. If t = 0, then we have
by [Reference Kazukawa, Nakajima and Shioya11, proposition 3.4], which implies that
Thus, the map H is continuous, so that it is a deformation retraction of $\Pi_*$ onto $\{Z\}$. Therefore, $\Pi_*$ is contractible.
6. Further questions
It is asked in [Reference Banakh, Cauty, Zarichnyi and Pearl1, Question 9.1] if the Gromov–Hausdorff space is homeomorphic to l 2. In our previous article [Reference Kazukawa, Nakajima and Shioya11], we have proved that $\mathcal{X}$ is not homeomorphic to l 2 with respect to the concentration topology. The following question remains.
Questiona 6.1.
Is $\mathcal{X}$ homeomorphic to l 2 with respect to the box topology?
We also ask the following.
Questiona 6.2.
If two sequences $\{\mathcal{P}_n\}_{n=1}^\infty$ and $\{\mathcal{Q}_n\}_{n=1}^\infty$ of pyramids converge weakly to pyramids $\mathcal{P}$ and $\mathcal{Q}$, respectively, then does $\{\mathcal{P}_n \times_p \mathcal{Q}_n\}_{n=1}^\infty$ converge weakly to $\mathcal{P} \times_p \mathcal{Q}$ as $n\to\infty$ for all $p\in [1,\infty]$?
Questiona 6.3.
Are Σ and $\Pi_*/{\mathbb{R}}_+$ contractible?
Questiona 6.4.
Are the spaces $\mathcal{X}$ and $\mathcal{X}_*$ locally contractible with respect to the box and/or concentration topologies?
Questiona 6.5.
Are Π and $\Pi_*$ locally contractible?
The following is already stated in our previous article [Reference Kazukawa, Nakajima and Shioya11].
Questiona 6.6.
Is the observable metric on $\mathcal{X}$ geodesic? Is the pyramidal compactification Π of $\mathcal{X}$ geodesic?
Acknowledgements
The authors would like to thank Professors Theo Sturm, Max von Renesse, Ayato Mitsuishi, Atsushi Katsuda, Kohei Suzuki, Yosuke Morita, and Shouhei Honda for their valuable comments. The authors would also like to thank an anonymous referee for carefully reading the manuscript and for his/her comments.
This work was supported by JSPS KAKENHI Grant Numbers JP22K20338, JP22K13908, JP19K03459.