This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Equicontinuity criteria for metric-valued sets of continuous functions

Marita Ferrer Universitat Jaume I, Instituto de Matemáticas de Castellón, Campus de Riu Sec, 12071 Castellón, Spain. mferrer@mat.uji.es Salvador Hernández Universitat Jaume I, INIT and Departamento de Matemáticas, Campus de Riu Sec, 12071 Castellón, Spain. hernande@mat.uji.es  and  Luis Tárrega Universitat Jaume I, IMAC and Departamento de Matemáticas, Campus de Riu Sec, 12071 Castellón, Spain. ltarrega@uji.es
(Date: July 26, 2025)
Abstract.

Combining ideas of Troallic [20] and Cascales, Namioka, and Vera [3], we prove several characterizations of almost equicontinuity and hereditarily almost equicontinuity for subsets of metric-valued continuous functions when they are defined on a Čech-complete space. We also obtain some applications of these results to topological groups and dynamical systems.

Research Partially supported Universitat Jaume I, grant P1-1B2015-77. The second author acknowledges partial support by Generalitat Valenciana, grant code: PROMETEO/2014/062, and the third author also acknowledges partial support of the Spanish Ministerio de Economía y Competitividad grant MTM 2013-42486-P
2010 Mathematics Subject Classification. Primary: 46A50, 54C35. Secondary: 22A05, 37B05, 54H11, 54H20.
Key Words and Phrases: Almost equicontinuous, Čech-completeness, dynamical system, fragmentability, pointwise convergence topology, topological group.

1. Introduction

Let XX and (M,d)(M,d) be a Hausdorff, completely regular space and a metric space, respectively, and let C(X,M)C(X,M) denote the set of all continuous functions from XX to MM. A subset GC(X,M)G\subseteq C(X,M) is said to be almost equicontinuous if GG is equicontinuous on a dense subset of XX. If GG is almost equicontinuous for every closed nonempty subset of XX, then it is said that GG is hereditarily almost equicontinuous. The main goal of this paper is to extend to arbitrary topological spaces these two important notions, which were introduced in the setting of topological dynamics studying the enveloping semigroup of a flow [1, 10, 11].

In addition to their intrinsic academic interest, it turns out that these two concepts have found application in other different settings as it will be made clear in the sequel. First, we shall provide some basic notions and terminology.

Given FXF\subseteq X, the symbol tp(F)t_{p}(F) (resp. t(F)t_{\infty}(F)) will denote the topology, on C(X,M)C(X,M), of pointwise convergence (resp. uniform convergence) on FF. For a set GG of functions from XX to MM and ZXZ\subseteq X, the symbol G|ZG|_{Z} will denote the set {g|Z:gG}\{g|_{Z}:g\in G\}. We denote by G¯MX\overline{G}^{M^{X}} the closure of GG in the Tychonoff product space MXM^{X}.
The symbolism (F,tp(G¯MX))(F,t_{p}(\overline{G}^{M^{X}})) will denote the set FF equipped with the weak topology generated by the functions in G¯MX|F\overline{G}^{M^{X}}|_{F}. In like manner, the symbol [A]ω[A]^{\leq\omega} will denote the set of all countable subsets of AA. A topological space XX is said to be Čech-complete if it is a GδG_{\delta}-subset of its Stone-Čech compatification. The family of Čech-complete spaces includes all complete metric spaces and all locally compact spaces. Several quotient spaces are used along the paper. For the reader’s sake, a detailed description of them is presented at the Appendix.

We now formulate our main results.

Theorem A.

Let XX and (M,d)(M,d) be a Čech-complete space and a separable metric space, respectively, and let GC(X,M)G\subseteq C(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. Consider the following three properties:

  1. (a)

    GG is almost equicontinuous.

  2. (b)

    There exists a dense Baire subset FXF\subseteq X such that (G¯MX)|F(\overline{G}^{M^{X}})|_{F} is metrizable.

  3. (c)

    There exists a dense GδG_{\delta} subset FXF\subseteq X such that (F,tp(G¯MX))(F,t_{p}(\overline{G}^{M^{X}})) is Lindelöf.

Then (b)(c)(a)(b)\Rightarrow(c)\Rightarrow(a). If XX is also a hereditarily Lindelöf space, then all conditions are equivalent.

Next result characterizes hereditarily almost equicontinuous families of functions defined on a Čech-complete space (this question has been studied in detail in [19] for compact spaces).

Theorem B.

Let XX and (M,d)(M,d) be a Čech-complete space and a metric space, respectively, and let GC(X,M)G\subseteq C(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. Then the following conditions are equivalent:

  1. (a)

    GG is hereditarily almost equicontinuous.

  2. (b)

    LL is hereditarily almost equicontinuous on FF, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

  3. (c)

    (L¯MX)|F(\overline{L}^{M^{X}})|_{F} is metrizable, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

  4. (d)

    (F,tp(L¯MX))(F,t_{p}(\overline{L}^{M^{X}})) is Lindelöf, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

Remark 1.3.

If GG is a subset of C(X,M)C(X,M) such that K=defG¯MXK\buildrel\rm def\over{=}\overline{G}^{M^{X}} is contained in C(X,M)C(X,M), then the implication (c)(a)(c)\Rightarrow(a) in Theorem B provides a different proof of the celebrated Namioka Theorem [14, Theorem 2.3]. Indeed, given any L[G]ωL\in[G]^{\leq\omega} and any separable compact subset FF of XX, since KC(X,M)K\subseteq C(X,M) and FF is separable, it follows that ((L¯MX)|F,tp(F))((\overline{L}^{M^{X}})|_{F},t_{p}(F)) is metrizable. Thus GG (and therefore KK) is hereditarily almost equicontinuous.

Corollary 1.4.

With the same hypothesis of Theorem B, consider the following three properties:

  1. (a)

    GG is hereditarily almost equicontinuous.

  2. (b)

    GG is hereditarily almost equicontinuous on FF, for all FF a separable and compact subset of XX.

  3. (c)

    (F,tp(G¯MX))(F,t_{p}(\overline{G}^{M^{X}})) is Lindelöf, for all FF a separable and compact subset of XX.

Then (a)(b)(c)(a)\Leftrightarrow(b)\Leftarrow(c).

2. Applications

The results formulated in the previous section have consequences in different settings. First, we consider an application to fragmentability.

A topological space XX is said to be fragmented by a pseudometric ρ\rho if for each nonempty subset AA of XX and for each ϵ>0\epsilon>0 there exists a nonempty open subset UU of XX such that UAU\cap A\neq\emptyset and ρ\rho-diam(UA)ϵdiam(U\cap A)\leq\epsilon. This notion was introduced by Jayne and Rogers in [12]. Further work has been done by many workers. It will suffice to mention here the contribution by Namioka [15] and Ribarska [17].

Let XX be a topological space, (M,d)(M,d) a metric space and GMXG\subseteq M^{X} a family of functions. Whenever feasible, for example if G¯MX\overline{G}^{M^{X}} is compact, we will consider the pseudometric ρG,d\rho_{G,d}, defined as follows:

ρG,d(x,y)=defsupgGd(g(x),g(y)),x,yX.\rho_{G,d}(x,y)\buildrel\rm def\over{=}\sup\limits_{g\in G}d(g(x),g(y)),\quad\forall x,y\in X.

Therefore, taking into account Definition 3.1 and Lemma 3.2, we have the following proposition.

Proposition 2.1.

Let XX and (M,d)(M,d) be a topological space and a metric space, respectively, and let GC(X,M)G\subseteq C(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. Consider the following two properties:

  1. (a)

    GG is hereditarily almost equicontinuous.

  2. (b)

    XX is fragmented by ρG,d\rho_{G,d}.

Then (a)(a) implies (b)(b). If XX is a hereditarily Baire space, then (a)(a) and (b)(b) are equivalent.

As a consequence, we have the following corollary of Theorem B.

Corollary 2.2.

Let XX and (M,d)(M,d) be a Čech-complete space and a metric space, respectively, and let GC(X,M)G\subseteq C(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. Then the following conditions are equivalent:

  1. (a)

    XX is fragmented by ρG,d\rho_{G,d}.

  2. (b)

    FF is fragmented by ρL,d\rho_{L,d}, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

  3. (c)

    ((L¯MX)|F,tp(F))((\overline{L}^{M^{X}})|_{F},t_{p}(F)) is metrizable, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

  4. (d)

    (F,tp(L¯MX))(F,t_{p}(\overline{L}^{M^{X}})) is Lindelöf, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

It is easy to check that, in the context of topological groups, the notion of almost equicontinuity is equivalent to equicontinuity. This fact allows us to characterize equicontinuous subsets of group homomorphisms using Theorem A.

From here on, if XX and MM are topological groups, the symbol CHom(X,M)CHom(X,M) will denote the set of continuous homomorphisms of XX into MM. Recall that a topological group GG is said to be ω\omega-narrow if for every neighborhood VV of the neutral element, there exists a countable subset EE of GG such that G=EVG=EV.

Corollary 2.3.

Let XX and (M,d)(M,d) be a Čech-complete topological group and a metric separable group, respectively, and let GG be a subset of CHom(X,M)CHom(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. Consider the following three properties:

  1. (a)

    GG is equicontinuous.

  2. (b)

    GG is relatively compact in CHom(X,M)CHom(X,M) with respect to the compact open topology.

  3. (c)

    There exists a dense Baire subset FXF\subseteq X such that (G¯MX)|F(\overline{G}^{M^{X}})|_{F} is metrizable.

  4. (d)

    There exists a dense GδG_{\delta} subset FXF\subseteq X such that (F,tp(G¯MX))(F,t_{p}(\overline{G}^{M^{X}})) is Lindelöf.

Then (c)(d)(a)(b)(c)\Rightarrow(d)\Rightarrow(a)\Leftrightarrow(b). If XX is also ω\omega-narrow, then all conditions are equivalent. Furthermore (c)(c) and (d)(d) are also true for F=XF=X.

Proof.

The equivalence (a)(b)(a)\Leftrightarrow(b) follows from Ascoli Theorem. So, after Theorem A, it will suffice to show the implication (a)(c)(a)\Rightarrow(c) for an ω\omega-narrow XX. Now, assuming that GG is equicontinuous, it follows that K=defG¯MXCHom(X,M)K\buildrel\rm def\over{=}\overline{G}^{M^{X}}\subseteq CHom(X,M). Thus KK is an equicontinuous compact subset of continuous group homomorphisms. As a consequence, it is known that KK is metrizable. (see [7, Cor. 3.5]). ∎

Extending a result given by Troallic in [20, Corollary 3.2], we can reduce the verification of hereditarily almost equicontinuity to countable subsets. The equivalence (a)(b)(a)\Leftrightarrow(b) bellow is due to Troallic (op. cit.).

Corollary 2.4.

Let XX and (M,d)(M,d) be a Čech-complete topological group and a metric group, respectively, and let GG be a subset of CHom(X,M)CHom(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. Then the following conditions are equivalent:

  1. (a)

    GG is equicontinuous.

  2. (b)

    LL is equicontinuous on FF, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

  3. (c)

    ((L¯MX)|F,tp(F))((\overline{L}^{M^{X}})|_{F},t_{p}(F)) is metrizable, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

  4. (d)

    (F,tp(L¯MX))(F,t_{p}(\overline{L}^{M^{X}})) is Lindelöf, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

For a function f:X×YMf:X\times Y\rightarrow M let fx:YMf_{x}:Y\rightarrow M (fy:XMf^{y}:X\rightarrow M ) be f(x,)f(x,\cdot) for a fixed xXx\in X (f(,y)f(\cdot,y) for a fixed yYy\in Y, resp.).

A variation of the celebrated Namioka Theorem [14] is also obtained as a corollary of Theorems A and B (cf. [13, 18, 16]).

Corollary 2.5.

Let XX, HH, and (M,d)(M,d) be a Čech-complete space, a compact space, and a metric space, respectively, and let f:X×HMf:X\times H\rightarrow M be a map satisfying that fxC(H,M)f_{x}\in C(H,M) for every xXx\in X and there is a dense subset GG of HH such that fgC(X,M)f^{g}\in C(X,M) for every gGg\in G. Suppose that any of the two following equivalent conditions holds.

  1. (a)

    There exists a dense Baire subset FXF\subseteq X such that (G¯MX)|F(\overline{G}^{M^{X}})|_{F} is metrizable.

  2. (b)

    There exists a dense GδG_{\delta} subset FXF\subseteq X such that (F,tp(G¯MX))(F,t_{p}(\overline{G}^{M^{X}})) is Lindelöf.

Then there exists a GδG_{\delta} and dense subset FF in XX such that ff is jointly continuous at each point of F×HF\times H.

Finally, we obtain some applications to dynamical systems [10, 9, 11]. Recall that a dynamical system, or a GG-space, is a Hausdorff space XX on which a topological group GG acts continuously. We denote such a system by (G,X)(G,X). For each gGg\in G we have the self-homeomorphism xgxx\mapsto gx of XX that we call gg-translation.

Corollary 2.6.

Let XX be a Polish GG-space such that G¯XX\overline{G}^{X^{X}} is compact. The following properties are equivalent:

  1. (a)

    GG is almost equicontinuous.

  2. (b)

    There exists a dense Baire subset FXF\subseteq X such that (G¯XX)|F(\overline{G}^{X^{X}})|_{F} is metrizable.

  3. (c)

    There exists a dense GδG_{\delta} subset FXF\subseteq X such that (F,tp(G¯XX))(F,t_{p}(\overline{G}^{X^{X}})) is Lindelöf.

Corollary 2.7.

Let XX be a completely metrizable GG-space such that G¯XX\overline{G}^{X^{X}} is compact. Then the following conditions are equivalent:

  1. (a)

    GG is hereditarily almost equicontinuous.

  2. (b)

    LL is hereditarily almost equicontinuous on FF, for all L[G]ωL\in[G]^{\leq\omega} and FF a compact subset of XX.

  3. (c)

    ((L¯MX)|F,tp(F))((\overline{L}^{M^{X}})|_{F},t_{p}(F)) is metrizable, for all L[G]ωL\in[G]^{\leq\omega} and FF a compact subset of XX.

  4. (d)

    (F,tp(L¯MX))(F,t_{p}(\overline{L}^{M^{X}})) is Lindelöf, for all L[G]ωL\in[G]^{\leq\omega} and FF a compact subset of XX.

In [2, Problem 28], Arkhangel’skii raises the following question: Let XX be a Lindelöf space and let KK be a compact subset of (C(X),tp(X))(C(X),t_{p}(X)). Is it true that the tightness of KK is countable? As far as we know, this question is still open in ZFC. Here we provide a partial answer to Arkhangel’skii’s question.

Corollary 2.8.

Let XX be a Lindelöf space and let KK be a compact subspace of (C(X),tp(X))(C(X),t_{p}(X)). If there is a a dense subset GKG\subseteq K such that (X,tp(G))(X,t_{p}(G)) is Čech-complete and hereditarily Lindelöf, then KK is metrizable.

Proof.

The proof of this result is consequence of Theorem B. Indeed, remark that, if FF is a subset of XX that is closed in the tp(G)t_{p}(G)-topology, then FF will be Čech-complete and hereditarily Lindelöf as well. Moreover, since GKG\subseteq K, it follows that FF is also closed in the tp(K)t_{p}(K)-topology and, as a consequence, Lindelöf. Applying Corollary 1.4 to the (compact) space KK, which is equipped with the tp(X)t_{p}(X)-topology, it follows that GG is hereditarily almost equicontinuous on XX. Since (X,tp(G))(X,t_{p}(G)) is Čech-complete and hereditarily Lindelöf, Proposition 4.6 yields the metrizability of K=G¯XK=\overline{G}^{\mathbb{R}^{X}}. ∎

3. Basic results

Within the setting of dynamical systems, the following definitions appear in [1].

Definition 3.1.

Let XX and (M,d)(M,d) be a topological space and a metric space respectively, and let GC(X,M)G\subseteq C(X,M). According to [1], we say that a point xXx\in X is an equicontinuity point of GG when for every ϵ>0{\epsilon}>0 there is a neighborhood UU of xx such that diam(g(U))<ϵdiam(g(U))<{\epsilon} for all gGg\in G. We say that GG is almost equicontinuous when the subset of equicontinuity points of GG is dense in XX. Furthermore, it is said that GG is hereditarily almost equicontinuous if G|AG|_{A} is almost equicontinuous for every nonempty closed subset AA of XX.

The proof of the following lemma is known. However it is very useful in order to obtain subsets of continuous functions that are not almost equicontinuous. We include its proof here for completeness sake.

Lemma 3.2.

Let XX and (M,d)(M,d) be a topological space and a metric space respectively, and let GC(X,M)G\subseteq C(X,M). Consider the following two properties:

  1. (a)

    GG is almost equicontinuous.

  2. (b)

    For every nonempty open subset UU of XX and ϵ>0\epsilon>0, there exists a nonempty open subset VUV\subseteq U such that diam(g(V))<ϵdiam(g(V))<\epsilon for all gGg\in G.

Then (a) implies (b). If X is a Baire space, then (a) and (b) are equivalent. Furthermore, in this case, the subset of equicontinuity points of GG is a dense GδG_{\delta}-set in XX.

Proof.

That (a) implies (b) is obvious. Assume that XX is a Baire space and (b) holds. Given ϵ>0\epsilon>0 arbitrary, we consider the open set Oϵ=def{UX:UO_{\epsilon}\buildrel\rm def\over{=}\bigcup\{U\subseteq X:\quad U is a nonempty open subset \wedge diam(g(U))<ϵgG}diam(g(U))<\epsilon\quad\forall g\in G\}. By (b), we have that OϵO_{\epsilon} is nonempty and dense in XX. Since XX is Baire, taking W=defn<ωO1nW\buildrel\rm def\over{=}\bigcap\limits_{n<\omega}O_{\frac{1}{n}}, we obtain a dense GδG_{\delta} subset which is the subset of equicontinuity points of GG. ∎

Remark 3.3.

As a consequence of assertion (b) in Lemma 3.2, it follows that, when XX is a Baire space, a subset of functions GG is hereditarily almost equicontinuous if, and only if, G|AG|_{A} is almost equicontinuous for every nonempty (non necessarily closed) subset AA of XX. Since we mostly work with Baire spaces here, we will make use of this fact in some parts along the paper.

Note that the set of equicontinuity points of a subset of functions GG is a GδG_{\delta}-set. Next corollary is a straightforward consequence of Lemma 3.2.

Corollary 3.4.

Let XX and (M,d)(M,d) be a topological space and a metric space respectively, and let GC(X,M)G\subseteq C(X,M). Suppose there is an open basis 𝒱\mathcal{V} in XX and ϵ>0{\epsilon}>0 such that for every V𝒱V\in\mathcal{V}, there is gVGg_{V}\in G with diam(gV(V))ϵdiam(g_{V}(V))\geq{\epsilon}. Then GG is not almost equicontinuous.

Let 2ω2^{\omega} be the Cantor space and let 2(ω)2^{(\omega)} denote the set of finite sequences of 0’s and 11’s. For a t2(ω)t\in 2^{(\omega)}, we designate by |t||t| the length of tt. For σ2ω\sigma\in 2^{\omega} and n>0n>0 we write σ|n\sigma|n to denote (σ(0),,σ(n1))2(ω)(\sigma(0),\ldots,\sigma(n-1))\in 2^{(\omega)}. If n=0n=0 then σ|0=def\sigma|0\buildrel\rm def\over{=}\emptyset.

Applying Corollary 3.4, it is easy to obtain subsets of continuous functions that are not almost equicontinuous.

Example 3.5.

Let X=2ωX=2^{\omega} be the Cantor space and let G={πn}n<ωG=\left\{\pi_{n}\right\}_{n<{\omega}} be the set of all projections of XX onto {0,1}\{0,1\}. Then GG is not almost equicontinuous.

Proof.

Let UU\neq\emptyset be an open subset in XX. Then, for some index n<ωn<{\omega} we have πn(U)={0,1}\pi_{n}(U)=\{0,1\}, which implies diam(πn(U))>1/2diam(\pi_{n}(U))>1/2. Therefore GG is not almost equicontinuous by Corollary 3.4. ∎

The precedent result can be generalized in order to obtain a more general example of non-almost equicontinuous set of functions. It turns out that this example is universal in a sense that will become clear along the paper.

Example 3.6.

Let X=2ωX=2^{\omega} be the Cantor space and let (M,d)(M,d) be a metric space. Let {Ut:t2(ω)}\{U_{t}:t\in 2^{({\omega})}\} be the canonical open basis of XX. If G={gt}t2(ω)G=\{g_{t}\}_{t\in 2^{({\omega})}} is a set of continuous functions on XX into MM satisfying that diam(gt(Ut))ϵdiam(g_{t}(U_{t}))\geq{\epsilon} for some fixed ϵ>0{\epsilon}>0 and all t2(ω)t\in 2^{({\omega})}, then GG is not almost equicontinuous.

Next result gives a sufficient condition for the equicontinuity of a family of functions. It extends a well known result by Corson and Glicksberg [5]. However, we remark that the subset FF found in the lemma below can become the empty set if ZZ is a first category subset of XX.

Lemma 3.7.

Let XX and (M,d)(M,d) be a topological space and a separable metric space, respectively. If GC(X,M)G\subseteq C(X,M) and (G¯MX)|Z(\overline{G}^{M^{X}})|_{Z} is metrizable and compact for some dense subset ZZ of XX, then there is a residual subset FF in ZZ such that GG is equicontinuous at every point in FF. In case ZZ is of second category in XX, it follows that FF will be necessarily nonempty.

Proof.

Set H=defG¯MXH\buildrel\rm def\over{=}\overline{G}^{M^{X}} and consider the map eval:XC(H,M)eval:X\rightarrow C(H,M), xevalxx\mapsto eval_{x}; defined by evalx(f)=deff(x)eval_{x}(f)\buildrel\rm def\over{=}f(x) for all xXx\in X and fHf\in H.

For simplicity’s sake, the symbols Ctp(G)(H|Z,M)C_{t_{p}(G)}(H|_{Z},M) and C(H|Z,M)C_{\infty}(H|_{Z},M) will denote the space C(H|Z,M)C(H|_{Z},M) equipped with the pointwise convergence tp(G)t_{p}(G) and the uniform convergence topology, respectively.

Now set Φ\Phi such that the following diagram commutes

Z\textstyle{Z\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}eval\scriptstyle{eval}Φ\scriptstyle{\Phi}Ctp(G)(H|Z,M)\textstyle{C_{t_{p}(G)}(H|_{Z},M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{id}C(H|Z,M)\textstyle{C_{\infty}(H|_{Z},M)}

Remark that the evaluation map, evaleval, is continuous because GC(X,M)G\subseteq C(X,M). Since H|ZH|_{Z} is tp(Z)t_{p}(Z)-compact and metrizable and ZZ is dense in XX, it follows that C(H|Z,M)C_{\infty}(H|_{Z},M) is separable and metrizable (see [6, Cor. 4.2.18]). Therefore, for every n<ωn<{\omega}, there is a sequence of closed balls {B¯(ui(n),1/n):i<ω}\{\overline{B}(u_{i}^{(n)},1/n):i<{\omega}\} that covers C(H|Z,M)C_{\infty}(H|_{Z},M). Furthermore, since GG is dense in HH, we have that each B¯(ui(n),1/n)\overline{B}(u_{i}^{(n)},1/n) is also closed in Ctp(G)(H|Z,M)C_{t_{p}(G)}(H|_{Z},M). As a consequence K(i,n)=defΦ1(B¯(ui(n),1/n))=eval1(B¯(ui(n),1/n))K_{(i,n)}\buildrel\rm def\over{=}\Phi^{-1}(\overline{B}(u_{i}^{(n)},1/n))=eval^{-1}(\overline{B}(u_{i}^{(n)},1/n)) is closed in ZZ for all i,n<ωi,n<\omega, because evaleval is continuous.

We have that Zi<ωK(i,n)Z\subseteq\bigcup\limits_{i<{\omega}}K_{(i,n)} for every n<ωn<\omega, so Zn<ωi<ωK(i,n)Z\subseteq\bigcap\limits_{n<{\omega}}\bigcup\limits_{i<{\omega}}K_{(i,n)}. Observe that n<ωi<ω(K(i,n)intZ(K(i,n)))\bigcup\limits_{n<{\omega}}\bigcup\limits_{i<{\omega}}(K_{(i,n)}\setminus int_{Z}(K_{(i,n)})) is a set of first category in ZZ. As a consequence

F=defZn<ωi<ω(K(i,n)intZ(K(i,n)))F\buildrel\rm def\over{=}Z\setminus\bigcup\limits_{n<{\omega}}\bigcup\limits_{i<{\omega}}(K_{(i,n)}\setminus int_{Z}(K_{(i,n)}))

is a residual set in ZZ.

We now verify that GG is equicontinuous at each point zFz\in F. Let zFz\in F and ϵ>0{\epsilon}>0 arbitrary. Take n0<ωn_{0}<{\omega} such that 2/n0<ϵ2/n_{0}<{\epsilon}. Since zFn<ωi<ωK(i,n)i<ωK(i,n0)z\in F\subseteq\bigcap\limits_{n<{\omega}}\bigcup\limits_{i<{\omega}}K_{(i,n)}\subseteq\bigcup\limits_{i<{\omega}}K_{(i,n_{0})} there is i0<ωi_{0}<{\omega} such that zK(i0,n0)z\in K_{(i_{0},n_{0})}. We claim that zintZ(K(i0,n0))z\in int_{Z}(K_{(i_{0},n_{0})}). Indeed, if we assume that zintZ(K(i0,n0))z\not\in int_{Z}(K_{(i_{0},n_{0})}), then zK(i0,n0)intZ(K(i0,n0))z\in K_{(i_{0},n_{0})}\setminus int_{Z}(K_{(i_{0},n_{0})}). Therefore, zn<ωi<ω(K(i,n)intZ(K(i,n)))z\in\bigcup\limits_{n<{\omega}}\bigcup\limits_{i<{\omega}}(K_{(i,n)}\setminus int_{Z}(K_{(i,n)})) and zFz\not\in F, which is a contradiction.

Since zintZ(K(i0,n0))z\in int_{Z}(K_{(i_{0},n_{0})}) there is a nonempty open set AA in XX such that intZ(K(i0,n0))=AZint_{Z}(K_{(i_{0},n_{0})})=A\cap Z. Note that AZA\cap Z is dense on AA because ZZ is dense in XX. So, zA=AZ¯AAZ¯Xz\in A=\overline{A\cap Z}^{A}\subseteq\overline{A\cap Z}^{X}.

Let a,bAZa,b\in A\cap Z. Then Φ(a)=evala,Φ(b)=evalbB¯(ui0(n0),1/n0)\Phi(a)=eval_{a},\Phi(b)=eval_{b}\in\overline{B}(u_{i_{0}}^{(n_{0})},1/n_{0}). Consequently, d(g(a),g(b))2/n0d(g(a),g(b))\leq 2/n_{0} for every gGg\in G. So, given x,yAAZ¯Xx,y\in A\subseteq\overline{A\cap Z}^{X} we have that d(g(x),g(y))2/n0d(g(x),g(y))\leq 2/n_{0} for every gGg\in G. Then diam(g(A))2/n0<ϵdiam(g(A))\leq 2/n_{0}<\epsilon for all gGg\in G. ∎

Remark 3.8.

Let XX be a topological space, (M,d)(M,d) be a metric space and GG be a subset of C(X,M)C(X,M) that we consider equipped with the pointwise convergence topology tp(X)t_{p}(X) in the sequel, unless otherwise stated.

Set

K=def{α:M[1,1]:|α(m1)α(m2)|d(m1,m2),m1,m2M}.K\buildrel\rm def\over{=}\{\alpha:M\rightarrow[-1,1]:|\alpha(m_{1})-\alpha(m_{2})|\leq d(m_{1},m_{2}),\quad\forall m_{1},m_{2}\in M\}.

It is readily seen that KK is a compact subspace of [1,1]M[-1,1]^{M}.

Consider the evaluation map φ:X×GM\varphi:X\times G\rightarrow M defined by φ(x,g)=defg(x)\varphi(x,g)\buildrel\rm def\over{=}g(x) for all (x,g)X×G(x,g)\in X\times G, which is clearly separately continuous. The map φ\varphi has associated a separately continuous map f:X×(G×K)[1,1]f:X\times(G\times K)\rightarrow[-1,1] defined by f(x,(g,α))=defα(g(x))f(x,(g,\alpha))\buildrel\rm def\over{=}\alpha(g(x)) for all (x,(g,α))X×(G×K)(x,(g,\alpha))\in X\times(G\times K).

Set

ν:G¯MX×K[1,1]X\nu:\overline{G}^{M^{X}}\times K\rightarrow[-1,1]^{X}

defined by

ν(h,α)=defαhfor allhG¯MXandαK.\nu(h,\alpha)\buildrel\rm def\over{=}\alpha\circ h\ \hbox{for all}\ h\in\overline{G}^{M^{X}}\ \hbox{and}\ \alpha\in K.

We claim that ν\nu is continuous. Indeed, let {(hδ,αδ)}δΔG¯MX×K\{(h_{\delta},\alpha_{\delta})\}_{\delta\in\Delta}\subseteq\overline{G}^{M^{X}}\times K be a net that converges to (h,α)G¯MX×K(h,\alpha)\in\overline{G}^{M^{X}}\times K. Given ϵ>0\epsilon>0 and xXx\in X, then there exists δ0Δ\delta_{0}\in\Delta such that d(hδ(x),h0(x))<ϵ/2d(h_{\delta}(x),h_{0}(x))<\epsilon/2 and |αδ(h0(x))α0(h0(x))|<ϵ/2|\alpha_{\delta}(h_{0}(x))-\alpha_{0}(h_{0}(x))|<\epsilon/2 for all δ>δ0\delta>\delta_{0}. Therefore, we have that |ν(h0,α0)(x)ν(hδ,αδ)(x)|=|α0(h0(x))αδ(hδ(x))||α0(h0(x))αδ(h0(x))|+|αδ(h0(x))αδ(hδ(x))||α0(h0(x))αδ(h0(x))|+d(h0(x),hδ(x))<ϵ|\nu(h_{0},\alpha_{0})(x)-\nu(h_{\delta},\alpha_{\delta})(x)|=|\alpha_{0}(h_{0}(x))-\alpha_{\delta}(h_{\delta}(x))|\leq|\alpha_{0}(h_{0}(x))-\alpha_{\delta}(h_{0}(x))|+|\alpha_{\delta}(h_{0}(x))-\alpha_{\delta}(h_{\delta}(x))|\leq|\alpha_{0}(h_{0}(x))-\alpha_{\delta}(h_{0}(x))|+d(h_{0}(x),h_{\delta}(x))<\epsilon for all δ>δ0\delta>\delta_{0}.

Since GC(X,M)G\subseteq C(X,M), we have that ν(G×K)C(X,[1,1])\nu(G\times K)\subseteq C(X,[-1,1]).

For m0Mm_{0}\in M, define αm0[1,1]M\alpha_{m_{0}}\in[-1,1]^{M} by αm0(m)=defd(m,m0)\alpha_{m_{0}}(m)\buildrel\rm def\over{=}d(m,{m_{0}}) for all mMm\in M. It is easy to check that αm0K\alpha_{m_{0}}\in K.

Lemma 3.9.

Let XX be a topological space, (M,d)(M,d) a metric space and GG a subset of C(X,M)C(X,M). Let KK and ν\nu be the space and the map defined in Remark 3.8. Then, for every subset FF of XX, the identity map id:(F,tp(G¯MX))(F,tp(ν(G¯MX×K)))id:(F,t_{p}(\overline{G}^{M^{X}}))\rightarrow(F,t_{p}(\nu(\overline{G}^{M^{X}}\times K))) is a homeomorphism.

Proof.

Let {xδ}δΔF\{x_{\delta}\}_{\delta\in\Delta}\subseteq F be a net that tp(G¯MX)t_{p}(\overline{G}^{M^{X}})-converges to xx. Since α\alpha is continuous, for any (h,α)G¯MX×K(h,\alpha)\in\overline{G}^{M^{X}}\times K, we have limδΔν(h,α)(xδ)=limδΔα(h(xδ))=α(h(x))=ν(h,α)(x)\lim\limits_{\delta\in\Delta}\nu(h,\alpha)(x_{\delta})=\lim\limits_{\delta\in\Delta}\alpha(h(x_{\delta}))=\alpha(h(x))=\nu(h,\alpha)(x). So, idid is continuous. Conversely, let {xδ}δΔF\{x_{\delta}\}_{\delta\in\Delta}\subseteq F be a net that tp(ν(G¯MX×K))t_{p}(\nu(\overline{G}^{M^{X}}\times K))-converges to x0Fx_{0}\in F. Given hG¯MXh\in\overline{G}^{M^{X}} arbitrary, take αh(x0)K\alpha_{h(x_{0})}\in K. So, fixed ϵ>0\epsilon>0, there is δ0Δ\delta_{0}\in\Delta such that ϵ>|ν(h,αh(x0))(xδ)ν(h,αh(x0))(x0)|=|d(h(xδ),h(x0))d(h(x0),h(x0))|=d(h(xδ),h(x0))\epsilon>|\nu(h,\alpha_{h(x_{0})})(x_{\delta})-\nu(h,\alpha_{h(x_{0})})(x_{0})|=|d(h(x_{\delta}),h(x_{0}))-d(h(x_{0}),h(x_{0}))|=d(h(x_{\delta}),h(x_{0})) for every δ>δ0\delta>\delta_{0}. That is, the net {xδ}δΔ\{x_{\delta}\}_{\delta\in\Delta} converges to x0x_{0} in tp(G¯MX)t_{p}(\overline{G}^{M^{X}}), which completes the proof. ∎

It is well known that the metric d¯:M×M\bar{d}:M\times M\rightarrow\mathbb{R} defined by d¯(m1,m2)=defmin{d(m1,m2),1}\bar{d}(m_{1},m_{2})\buildrel\rm def\over{=}\min\{d(m_{1},m_{2}),1\} for all m1,m2Mm_{1},m_{2}\in M induces the same topology as dd. So, without loss of generality, we work with this metric from here on.

The following lemma reduces many questions related to a general metric space MM to the interval [1,1][-1,1] (cf. [4]).

Lemma 3.10.

Let XX and (M,d)(M,d) be a topological and a metric space, respectively. If GG is a subset of C(X,M)C(X,M), then GG is equicontinuous at a point x0Xx_{0}\in X if and only if ν(G×K)\nu(G\times K) is equicontinuous at it.

Proof.

Assume that GG is equicontinuous at x0x_{0}. Given ϵ>0\epsilon>0, there is an open neighbouhood UU of x0x_{0} such that d(g(x0),g(x))<ϵd(g(x_{0}),g(x))<\epsilon for all xUx\in U and gGg\in G. Let αK\alpha\in K, xUx\in U and gGg\in G, then we have

|ν(g,α)(x0)ν(g,α)(x)|=|α(g(x0))α(g(x))|d(g(x0),g(x))<ϵ.|\nu(g,\alpha)(x_{0})-\nu(g,\alpha)(x)|=|\alpha(g(x_{0}))-\alpha(g(x))|\leq d(g(x_{0}),g(x))<\epsilon.

Conversely, assume that ν(G×K)\nu(G\times K) is equicontinuous in x0x_{0}. Given ϵ>0\epsilon>0, there is an open neighbouhood UU of x0x_{0} such that |ν(g,α)(x0)ν(g,α)(x)|<ϵ|\nu(g,\alpha)(x_{0})-\nu(g,\alpha)(x)|<\epsilon for all xUx\in U, gGg\in G and αK\alpha\in K.

For gGg\in G, consider the map αg(x0)K\alpha_{g(x_{0})}\in K. In order to finish the proof, it will suffice to observe that

|αg(x0)(g(x0))αg(x0)(g(x))|=d(g(x),g(x0))|\alpha_{g(x_{0})}(g(x_{0}))-\alpha_{g(x_{0})}(g(x))|=d(g(x),g(x_{0}))

for all xUx\in U and gGg\in G. ∎

Corollary 3.11.

Let XX and (M,d)(M,d) be a topological and a metric space, respectively, and let GG be an arbitrary subset of C(X,M)C(X,M). Then GG is (hereditarily) almost equicontinuous if and only if ν(G×K)\nu(G\times K) is (hereditarily) almost equicontinuous.

4. Proof of main results

The following technical lemma is essential in most results along this paper. The construction of the proof is based on an idea that appears in [18] and [3]. We recall that a topological space is hemicompact if it has a sequence of compact subsets such that every compact subset of the space lies inside some compact set in the sequence. Every compact space or every locally compact and Lindelöf space is hemicompact.

Lemma 4.1.

Let XX and (M,d)(M,d) be a Čech-complete space and a hemicompact metric space, respectively, and let GG be a subset of C(X,M)C(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. If GG is not almost equicontinuous, then for every GδG_{\delta} and dense subset FF of XX there exists a countable subset LL in GG, a compact separable subset CFFC_{F}\subseteq F, a compact subset NMN\subseteq M and a continuous and surjective map Ψ\Psi of CFC_{F} onto the Cantor set 2ω2^{\omega} such that for every lLl\in L there exists a continuous map l:2ωNl^{*}:2^{\omega}\rightarrow N satisfying that the following diagram is commutative

Diagram 1:

CF\textstyle{C_{F}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ\scriptstyle{\Psi}l|CF\scriptstyle{l|_{C_{F}}}2ω\textstyle{2^{\omega}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l^{*}}N\textstyle{N}

Furthermore, the subset L=def{l:lL}C(2ω,N)L^{*}\buildrel\rm def\over{=}\{l^{*}:l\in L\}\subseteq C(2^{\omega},N) separate points in 2ω2^{\omega} and is not almost equicontinuous on 2ω2^{\omega}.

Proof.

Let FF be a GδG_{\delta} and dense subset of XX. Then there is a sequence {Wn}n=1\{W_{n}\}_{n=1}^{\infty} of open dense subsets of XX such that WsWrW_{s}\subseteq W_{r} if r<sr<s and F=n=1WnF=\bigcap\limits_{n=1}^{\infty}W_{n}.

Let {Mn}n<ω\{M_{n}\}_{n<\omega} be a sequence of compact subsets, that we obtain by hemicompactness such that M=n<ωMnM=\bigcup\limits_{n<\omega}M_{n} and for every compact subset KMK\subseteq M there is n<ωn<\omega such that KMnK\subseteq M_{n}.

For each n<ωn<\omega we consider the closed subset Xn={xX:g(x)MngG}X_{n}=\{x\in X:g(x)\in M_{n}\quad\forall g\in G\}. We claim that X=n<ωXnX=\bigcup\limits_{n<\omega}X_{n}. Indeed, let xXx\in X. Since G¯MXMX\overline{G}^{M^{X}}\subseteq M^{X} is compact and the xxth projection πx\pi_{x} is continuous, then πx(G¯MX)M\pi_{x}(\overline{G}^{M^{X}})\subseteq M is compact. So, there is nx<ωn_{x}<\omega such that πx(G¯MX)Mnx\pi_{x}(\overline{G}^{M^{X}})\subseteq M_{n_{x}} by hemicompactness. Therefore xXnxx\in X_{n_{x}}.

Since GG is not almost equicontinuous there exists a nonempty open subset UU of XX and ϵ>0\epsilon>0 such that for all nonempty open subset VUV\subseteq U there exists a function gVGg_{V}\in G such that diam(gV(V))2ϵ>ϵdiam(g_{V}(V))\geq 2\epsilon>{\epsilon} by Lemma 3.2.

Note that UU is Čech-complete. If we express U=nω(UXn)U=\bigcup\limits_{n\in\omega}(U\cap X_{n}), by Baire’s theorem, there is n0<ωn_{0}<\omega such that U~=defintU(UXn0)\tilde{U}\buildrel\rm def\over{=}int_{U}(U\cap X_{n_{0}})\neq\emptyset and open in XX.

Set C=U~¯Xn0C=\overline{\tilde{U}}^{X_{n_{0}}}, which is closed in XX, and On=WnU~=WnU~CO_{n}=W_{n}\cap\tilde{U}=W_{n}\cap\tilde{U}\cap C that is open and dense in CC for each n<ωn<\omega. Then OsOrO_{s}\subseteq O_{r} if r<sr<s and H=n=1OnFH=\bigcap\limits_{n=1}^{\infty}O_{n}\subseteq F is a dense GδG_{\delta} subset of CC, which is a Baire space. Remark further that g(x)Mn0g(x)\in M_{n_{0}} for all xCx\in C and gGg\in G. Since Mn0M_{n_{0}} is compact, every function fC(C,Mn0)f\in C(C,M_{n_{0}}) can be extended to a continuous function fβC(βC,Mn0)f^{\beta}\in C(\beta C,M_{n_{0}}). Set Gβ={gβ:gG}C(βC,Mn0)G^{\beta}=\{g^{\beta}:g\in G\}\subseteq C({\beta}C,M_{n_{0}}).

The space CC, being Čech-complete, is a dense GδG_{\delta} subset of its Stone-Čech compactification βC{\beta}C. Therefore, since HH is a GδG_{\delta} subset of CC, it follows that HH also is a dense GδG_{\delta} subset of βC\beta C. Consider a sequence {En}n=1\{E_{n}\}_{n=1}^{\infty} of open dense subsets of βC\beta C such that EsErE_{s}\subseteq E_{r} if r<sr<s and H=n=1EnH=\bigcap\limits_{n=1}^{\infty}E_{n}. We have that H=n=1(EnOnβ)H=\bigcap\limits_{n=1}^{\infty}(E_{n}\cap O^{\beta}_{n}), where Onβ=βC(COn)¯βCO^{\beta}_{n}={\beta}C\setminus\overline{(C\setminus O_{n})}^{{\beta}C} is open in βC{\beta}C and OnβC=OnO^{{\beta}}_{n}\cap C=O_{n}.

By induction on n=|t|n=|t| with t2(ω)t\in 2^{(\omega)}, we construct a family {Ut:t2(ω)}\{U_{t}:t\in 2^{(\omega)}\} of nonempty open subsets of βC{\beta}C and a family of countable functions L=def{gt:t2(ω)}GL\buildrel\rm def\over{=}\{g_{t}:t\in 2^{(\omega)}\}\subseteq G, satisfying the following conditions for all t2(ω)t\in 2^{(\omega)}:

  1. (i)

    UU¯βCO0β=defβC(CU~)¯βCU_{\emptyset}\subseteq\overline{U_{\emptyset}}^{{\beta}C}\subseteq O^{\beta}_{0}\buildrel\rm def\over{=}{\beta}C\setminus\overline{(C\setminus\tilde{U})}^{{\beta}C} (remark that O0βC=U~O^{\beta}_{0}\cap C=\tilde{U});

  2. (ii)

    UtiUti¯βCE|t|O|t|βUtU_{ti}\subseteq\overline{U_{ti}}^{{\beta}C}\subseteq E_{|t|}\cap O^{\beta}_{|t|}\cap U_{t} for i=0,1i=0,1 (where E0=defβCE_{0}\buildrel\rm def\over{=}\beta C);

  3. (iii)

    Ut0Ut1=U_{t0}\cap U_{t1}=\emptyset;

  4. (iv)

    d(gt(x),gt(y))>ϵd(g_{t}(x),g_{t}(y))>\epsilon, xUt0C\forall x\in U_{t0}\cap C and yUt1C\forall y\in U_{t1}\cap C;

  5. (v)

    whenever s,t2(ω)s,t\in 2^{(\omega)} and |s|<|t||s|<|t|, diam(gs(UtjC))<1|t|diam(g_{s}(U_{tj}\cap C))<\frac{1}{|t|} for j=0,1j=0,1.

Indeed, if n=0n=0, by regularity we can find UU_{\emptyset} an open set in βC{\beta}C such that UU¯βCE0O0βU_{\emptyset}\subseteq\overline{U_{\emptyset}}^{{\beta}C}\subseteq E_{0}\cap O^{\beta}_{0}. For n0n\geq 0, suppose {Ut:|t|n}\{U_{t}:|t|\leq n\} and {gt:|t|<n}\{g_{t}:|t|<n\} have been constructed satisfying (i)(v)(i)-(v). Fix a t2(ω)t\in 2^{(\omega)} with |t|=n|t|=n. Since UtU_{t} is open in βC{\beta}C, there is an open set AtA_{t} in XX such that UtC=AtCU_{t}\cap C=A_{t}\cap C. Therefore

UtC=(AtC)O0β=At(O0βC)=AtU~U_{t}\cap C=(A_{t}\cap C)\cap O^{\beta}_{0}=A_{t}\cap(O^{\beta}_{0}\cap C)=A_{t}\cap\tilde{U}

is open in XX and included in UU.

By assumption there exist gtGg_{t}\in G such that diam(gt(UtC))>ϵdiam(g_{t}(U_{t}\cap C))>\epsilon. Consequently, we can find xt,ytVtCx_{t},y_{t}\in V_{t}\cap C such that d(gt(xt),gt(yt))>ϵd(g_{t}(x_{t}),g_{t}(y_{t}))>\epsilon. By continuity, we can select two open disjoint neighbourhoods in βC{\beta}C, St0S_{t0} and St1S_{t1} of xtx_{t} and yty_{t}, respectively, satisfying conditions (iii)(iii) and (iv)(iv).

If i{0,1}i\in\{0,1\}, observe that UtStiO0βU_{t}\cap S_{ti}\cap O^{\beta}_{0} is open in βC{\beta}C and nonempty. Since E|t|O|t|βE_{|t|}\cap O^{\beta}_{|t|} is dense in βC{\beta}C then UtStiE|t|O|t|βU_{t}\cap S_{ti}\cap E_{|t|}\cap O^{\beta}_{|t|} is a nonempty open subset of βC{\beta}C. By regularity there exists a nonempty open subset UtiU_{ti} of βC{\beta}C such that UtiUti¯βCUtStiE|t|O|t|βU_{ti}\subseteq\overline{U_{ti}}^{{\beta}C}\subseteq U_{t}\bigcap S_{ti}\bigcap E_{|t|}\cap O^{\beta}_{|t|}. Therefore, Ut0U_{t0} and Ut1U_{t1} satisfies conditions (ii),(iii)(ii),(iii) and (iv)(iv) and, by continuity, we can adjust the open sets to satisfy (v)(v).

Set K=defn=0|t|=nUt¯βCK\buildrel\rm def\over{=}\bigcap\limits_{n=0}^{\infty}\bigcup\limits_{|t|=n}\overline{U_{t}}^{{\beta}C}, which is closed in βC{\beta}C and, as a consequence, also compact. Remark that we can express K=σ2ωn=0Uσ|n¯βCK=\bigcup\limits_{\sigma\in 2^{\omega}}\bigcap\limits_{n=0}^{\infty}\overline{U_{\sigma|n}}^{{\beta}C}. Therefore, for each σ2ω\sigma\in 2^{\omega}, we have n=0Uσ|n¯βC\bigcap\limits_{n=0}^{\infty}\overline{U_{\sigma|n}}^{\beta C}\neq\emptyset by the compactness of βC{\beta}C, which implies KK\neq\emptyset. Furthermore, since Kn=0(EnOnβ)=HFK\subseteq\bigcap\limits_{n=0}^{\infty}(E_{n}\cap O^{\beta}_{n})=H\subseteq F, it follows that KK is contained in FF.

Let Ψ:K2ω\Psi:K\rightarrow 2^{\omega} be the canonical map defined such that Ψ1(σ)=n=0Uσ|n¯βC\Psi^{-1}(\sigma)=\bigcap\limits_{n=0}^{\infty}\overline{U_{\sigma|n}}^{\beta C} for all σ2ω\sigma\in 2^{\omega}. Clearly Ψ\Psi is onto and continuous. Observe that for each t2(ω)t\in 2^{(\omega)} and σ2ω\sigma\in 2^{\omega}, gt(Ψ1(σ))g_{t}(\Psi^{-1}(\sigma)) is a singleton by (iv)(iv). Therefore, gtg_{t} lifts to a continuous function gtg_{t}^{*} on 2ω2^{\omega} such that gt(x)=gt(Ψ(x))g_{t}(x)=g_{t}^{*}(\Psi(x)) for all xKx\in K.

Take a countable subset DD of KK such that Ψ(D)=2(ω)\Psi(D)=2^{(\omega)} and makes Ψ|D\Psi|_{D} injective. Set CF=defD¯KC_{F}\buildrel\rm def\over{=}\overline{D}^{K}. Note that 2(ω)2^{(\omega)} is a countable dense subset of 2ω2^{\omega}.

We have that Ψ|CF:CF2ω\Psi_{|C_{F}}:C_{F}\rightarrow 2^{\omega} is an onto and continuous map. We consider the set LC(2ω,Mn0)L^{*}\subseteq C(2^{\omega},M_{n_{0}}) defined by L={l:lL|CF}L^{*}=\{l^{*}:l\in L|_{C_{F}}\} that makes the diagram 1 commutative. We claim that LL^{*} separates points in 2ω2^{\omega} and, as a consequence, defines its topology. Indeed, let σ,σ2ω\sigma,\sigma^{\prime}\in 2^{\omega} be two arbitrary points such that σσ\sigma\neq\sigma^{\prime}. Since Ψ\Psi is an onto map there exist x,yCFx,y\in C_{F} such that σ=Ψ(x)\sigma=\Psi(x) and σ=Ψ(y)\sigma^{\prime}=\Psi(y). Therefore, xn=0Uσ|n¯βCx\in\bigcap\limits_{n=0}^{\infty}\overline{U_{\sigma|n}}^{\beta C} and yn=0Uσ|n¯βCy\in\bigcap\limits_{n=0}^{\infty}\overline{U_{\sigma^{\prime}|n}}^{\beta C}. Since σσ\sigma\neq\sigma^{\prime}, there is n0ωn_{0}\in\omega such that σ|n0=σ|n0\sigma|n_{0}=\sigma^{\prime}|n_{0} and σ(n0+1)σ(n0+1)\sigma(n_{0}+1)\neq\sigma^{\prime}(n_{0}+1). Taking t=σ|n0t=\sigma|n_{0}, then by (iv)(iv) we know that d(gt(x),gt(y))>ϵd(g_{t}(x),g_{t}(y))>\epsilon. So, gt(σ)gt(σ)g_{t}^{*}(\sigma)\neq g_{t}^{*}(\sigma^{\prime}).

On the other hand, by the commutativity of Diagram 1, and taking into account how LL and LL^{*} have been defined, it is easily seen that LL^{*} is not almost equicontinuous on 2ω2^{\omega} using Example 3.6. ∎

Applying Corollary D of [3] by Cascales, Namioka and Vera and Facts 5.1, 5.2, 5.3 and 5.4, next result follows easily.

Proposition 4.2.

Let XX be a compact space, (M,d)(M,d) be a compact metric space and let GG be a subset of C(X,M)C(X,M). If (X,tp(G¯MX))(X,t_{p}(\overline{G}^{M^{X}})) is Lindelöf, then GG is hereditarily almost equicontinuous.

Using Lemma 4.1, the constraints in Proposition 4.2 can be relaxed as the following result shows.

Proposition 4.3.

Let XX be a Čech-complete space, (M,d)(M,d) be a compact metric space and let GG be a subset of C(X,M)C(X,M). If there exists a dense GδG_{\delta} subset FXF\subseteq X such that (F,tp(G¯MX))(F,t_{p}(\overline{G}^{M^{X}})) is Lindelöf, then GG is almost equicontinuous.

Proof.

Reasoning by contradiction, suppose that GG is not almost equicontinuous. By Lemma 4.1 there exists a compact separable subset CFC_{F} of FF, a continuous onto map Ψ:CF2ω\Psi:C_{F}\rightarrow 2^{\omega}, and a countable subset LL of GG such that the subset LC(2ω,M)L^{*}\subseteq C(2^{\omega},M) defined by l(Ψ(x))=l(x)l^{*}(\Psi(x))=l(x) for all xCFx\in C_{F} separate points in 2ω2^{\omega} and is not almost equicontinuous.

Let KFK_{F} be the closure of CFC_{F} in FF with respect to the initial topology generated by the maps in LL. Using a compactness argument, it follows that if pKFp\in K_{F} then there is xpCFx_{p}\in C_{F} such that l(p)=l(xp)l(p)=l(x_{p}) for all lLl\in L. Indeed, let pKFp\in K_{F}. Then there is a net {xδ}δΔCF\{x_{\delta}\}_{\delta\in\Delta}\subseteq C_{F} that tp(L)t_{p}(L)-converges to pp. Since CFC_{F} is compact there is a subnet {xγ}γΓ\{x_{\gamma}\}_{\gamma\in\Gamma} such that converges to x0CFx_{0}\in C_{F}. Given lLl\in L, we know that limγΓl(xγ)=l(x0)\lim\limits_{\gamma\in\Gamma}l(x_{\gamma})=l(x_{0}) because ll is continuous. Therefore, l(x0)=limγΓl(xγ)=l(p)l(x_{0})=\lim\limits_{\gamma\in\Gamma}l(x_{\gamma})=l(p). Consequently, we can extend Ψ\Psi to a map Φ:KF2ω\Phi:K_{F}\rightarrow 2^{\omega} by Φ(p)=Ψ(xp)\Phi(p)=\Psi(x_{p}) for all pKFp\in K_{F}.

Let’s see that Φ\Phi is well-defined. Let pKFp\in K_{F}, suppose that there are xp,x~pCFx_{p},\tilde{x}_{p}\in C_{F} such that xpx~px_{p}\neq\tilde{x}_{p} and l(p)=l(xp)=l(x~p)l(p)=l(x_{p})=l(\tilde{x}_{p}) for all lLl\in L. Since the Diagram 1 commutes, we know that l(Ψ(xp))=l(Ψ(x~p))l^{*}(\Psi(x_{p}))=l^{*}(\Psi(\tilde{x}_{p})) for all lLl^{*}\in L^{*}. So, Ψ(xp)=Ψ(x~p)\Psi(x_{p})=\Psi(\tilde{x}_{p}) because LL^{*} separates points in 2ω2^{\omega}.
Observe that the following diagram is commutative

Diagram 2:

KF\textstyle{K_{F}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ\scriptstyle{\Phi}l|KF\scriptstyle{l|_{K_{F}}}2ω\textstyle{2^{\omega}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l^{*}}M\textstyle{M}

Certainly, let pKFp\in K_{F}, then there is xpCFx_{p}\in C_{F} such that Φ(p)=Ψ(xp)\Phi(p)=\Psi(x_{p}). Given lLl\in L, we have that l(p)=l(xp)=l(Ψ(xp))=l(Φ(p))l(p)=l(x_{p})=l^{*}(\Psi(x_{p}))=l^{*}(\Phi(p)).
We claim that Φ:(KF,tp(L))(2ω,tp(L))\Phi:(K_{F},t_{p}(L))\rightarrow(2^{\omega},t_{p}(L^{*})) is also continuous. Indeed, let {hδ}δΔKF\{h_{\delta}\}_{\delta\in\Delta}\subseteq K_{F} a net that tp(L)t_{p}(L)-converges to h0KFh_{0}\in K_{F}. For each δΔ\delta\in\Delta there is xδCFx_{\delta}\in C_{F} such that Φ(hδ)=Ψ(xδ)\Phi(h_{\delta})=\Psi(x_{\delta}) and l(hδ)=l(xδ)l(h_{\delta})=l(x_{\delta}) for all lLl\in L. Analogously, there is x0CFx_{0}\in C_{F} such that Φ(h0)=Ψ(x0)\Phi(h_{0})=\Psi(x_{0}) and l(h0)=l(x0)l(h_{0})=l(x_{0}) for all lLl\in L.
Since CFC_{F} is compact there is a subnet {xγ}γΓ\{x_{\gamma}\}_{\gamma\in\Gamma} such that converges to x~CF\tilde{x}\in C_{F}. Given lLl\in L, we know that limγΓl(xγ)=l(x~)\lim\limits_{\gamma\in\Gamma}l(x_{\gamma})=l(\tilde{x}) because ll is continuous. On the other hand, we also have that limγΓl(xγ)=limγΓl(hγ)=l(h0)=l(x0)\lim\limits_{\gamma\in\Gamma}l(x_{\gamma})=\lim\limits_{\gamma\in\Gamma}l(h_{\gamma})=l(h_{0})=l(x_{0}). Therefore, l(x~)=l(x0)l(\tilde{x})=l(x_{0}) for all lLl\in L. So, Ψ(x~)=Ψ(x0)\Psi(\tilde{x})=\Psi(x_{0}) because LL^{*} separates points in 2ω2^{\omega}. The continuity follows because limγΓΦ(hγ)=limγΓΨ(xγ)=Ψ(x~)=Ψ(x0)=Φ(h0)\lim\limits_{\gamma\in\Gamma}\Phi(h_{\gamma})=\lim\limits_{\gamma\in\Gamma}\Psi(x_{\gamma})=\Psi(\tilde{x})=\Psi(x_{0})=\Phi(h_{0}).

Now, since KFK_{F} is tp(L)t_{p}(L)-closed in FF, it follows that it is also tp(G¯MX)t_{p}(\overline{G}^{M^{X}})-closed in FF.

By our initial assumption, we have that FF is tp(G¯MX)t_{p}(\overline{G}^{M^{X}})-Lindelöf, which implies that also KFK_{F} is tp(G¯MX)t_{p}(\overline{G}^{M^{X}})-Lindelöf.

We claim that (2ω,tp(L¯M2ω))(2^{\omega},t_{p}(\overline{L^{*}}^{M^{2^{\omega}}})) is also Lindelöf. Indeed, it is enough to prove that Φ\Phi is continuous on KFK_{F} when it is equipped with the tp(G¯MX)t_{p}(\overline{G}^{M^{X}})-topology and 2ω2^{\omega} is equipped with the tp(L¯M2ω)t_{p}(\overline{L^{*}}^{M^{2^{\omega}}})-topology.

Take a map kL¯M2ωk\in\overline{L^{*}}^{M^{2^{\omega}}} and let {lγ}γΓL\{l^{*}_{\gamma}\}_{\gamma\in\Gamma}\subseteq L^{*} be a net converging to kk pointwise on 2ω2^{\omega}. Since G¯MX\overline{G}^{M^{X}} is compact, we may assume wlog that {lγ}γΓL\{l_{\gamma}\}_{\gamma\in\Gamma}\subseteq L tp(X)t_{p}(X)-converges to hG¯MXh\in\overline{G}^{M^{X}}. Therefore, for each xKFx\in K_{F} we have that k(Φ(x))=limγΓlγ(Φ(x))=limγΓlγ(x)=h(x)k(\Phi(x))=\lim\limits_{\gamma\in\Gamma}l^{*}_{\gamma}(\Phi(x))=\lim\limits_{\gamma\in\Gamma}l_{\gamma}(x)=h(x). That is kΦ=hk\circ\Phi=h. Since hh is continuous on KFK_{F}, the continuity of Φ\Phi follows.

By Proposition 4.2, this implies that LL^{*} is a hereditarily almost equicontinuous family on 2ω2^{\omega}, which is a contradiction. ∎

Proposition 4.4.

Let XX be a Čech-complete space, (M,d)(M,d) be a metric space and let GG be a subset of C(X,M)C(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. If there exists a dense GδG_{\delta} subset FXF\subseteq X such that (F,tp(G¯MX))(F,t_{p}(\overline{G}^{M^{X}})) is Lindelöf, then GG is almost equicontinuous.

Proof.

Let KK and ν\nu defined as in Remark 3.8. Since ν(G¯MX×K)\nu(\overline{G}^{M^{X}}\times K) is a compact subset of [1,1]X[-1,1]^{X}, it follows that ν(G×K)¯[1,1]X=ν(G¯MX×K)\overline{\nu(G\times K)}^{[-1,1]^{X}}=\nu(\overline{G}^{M^{X}}\times K).

By Lemma 3.9 we know that (F,tp(ν(G¯MX×K)))(F,t_{p}(\nu(\overline{G}^{M^{X}}\times K))) is Lindelöf . Now, applying Proposition 4.3 to the subset ν(G×K)C(X,[1,1])\nu(G\times K)\subseteq C(X,[-1,1]), it follows that ν(G×K)\nu(G\times K) is almost equicontinuous. Therefore, GG is almost equicontinuous by Corollary 3.11. ∎

The following lemma is known. We refer to [7, Cor. 3.5] for its proof.

Lemma 4.5.

Let XX be a Lindelöf space, (M,d)(M,d) be a metric space. If GG is an equicontinuous subset of C(X,M)C(X,M), then G¯MX\overline{G}^{M^{X}} is metrizable.

We are now in position of proving Theorem A.

Proof of Theorem A.

(b)(c)(b)\Rightarrow(c) Since (G¯MX)|F(\overline{G}^{M^{X}})|_{F} is compact metric, it follows by Lemma 3.7 that there is a dense subset EE such that GG is equicontinuous at the points in EE with respect to XX. Since EE is dense in FF, which is dense in XX, it follows that EE is also be dense in XX. Moreover, if YY denotes the GδG_{\delta} subset of equicontinuity points of GG in XX, since EYE\subseteq Y, it follows that YY, the set of equicontinuity points of GG is a dense GδG_{\delta}-set in XX. Set K=def(G¯MX)K\buildrel\rm def\over{=}(\overline{G}^{M^{X}}).The equicontinuity of GG at the points in YY combined with the density of EFE\subseteq F in YY, implies that the map Θ:K|FK|Y\Theta:K|_{F}\longrightarrow K|_{Y} defined by Θ(f|F)=deff|Y\Theta(f|_{F})\buildrel\rm def\over{=}f|_{Y} is a homeomorphism of K|FK|_{F} onto K|YK|_{Y}.

By our initial assumption we have that K|FK|_{F} is compact and metrizable, which yields the metrizability of K|YK|_{Y}. Thus, the evaluation map Eval:YC(K|Y,M)Eval:Y\longrightarrow C_{\infty}(K|_{Y},M) is a well defined and continuous map. We know that C(K|Y,M)C_{\infty}(K|_{Y},M) is a separable space by [6, Cor. 4.2.18]. Therefore (Eval(Y),t(K|Y))(Eval(Y),t_{\infty}(K|_{Y})) and (Y,t(K|Y))(Y,t_{\infty}(K|_{Y})) are Lindelöf spaces. As a consequence (Y,tp(K|Y))(Y,t_{p}(K|_{Y})) must be also Lindelöf and we are done.

(c)(a)(c)\Rightarrow(a) This implication is Proposition 4.4

(a)(b)(a)\Rightarrow(b) Suppose that XX is Čech-complete and hereditarily Lindelöf. By Lemma 3.2, the subset, FF, of equicontinuity points of GG is a dense GδG_{\delta}-set in XX, which is a Lindelöf space by our initial assumption. Since GG is equicontinuous on FF, Lemma 4.5 implies that (G¯MX)|F(\overline{G}^{M^{X}})|_{F} must be metrizable. ∎

The following result can be found in [11, Prop. 2.5 and Section 5] in the setting of compact metric spaces. Notwithstanding this, the proof given there can be adapted easily for Čech-complete and hereditarily Lindelöf spaces, as it is formulated in the next proposition. A sketch of the proof is included here for completeness sake.

Proposition 4.6.

Let XX be a hereditarily Lindelöf space, (M,d)(M,d) is a metric space and GC(X,M)G\subseteq C(X,M). If H=defG¯MXH\buildrel\rm def\over{=}\overline{G}^{M^{X}} is compact and hereditarily almost equicontinuous, then HH is metrizable.

Proof.

The symbol C(H,M)C_{\infty}(H,M) denote the space C(H,M)C(H,M) equipped with the uniform convergence topology. Consider the map eval:XC(H,M)eval:X\rightarrow C_{\infty}(H,M) defined by eval(x)[h]=defh(x)eval(x)[h]\buildrel\rm def\over{=}h(x) for all xXx\in X and hHh\in H.

By Proposition 2.1 XX is fragmented by ρG,d\rho_{G,d}. Thus, for each nonempty subset AA of XX and for each ϵ>0\epsilon>0 there exists a nonempty open subset UU of XX such that UAU\cap A\neq\emptyset and diam(h(UA))ϵdiam(h(U\cap A))\leq\epsilon for all hHh\in H. Thus, dd_{\infty}-diam(eval(UA))ϵdiam(eval(U\cap A))\leq\epsilon.

We claim that eval(X)eval(X) is separable. Indeed, pick ϵ>0\epsilon>0. Let 𝒜\mathcal{A} be the collection of all open subsets OO of XX such that eval(O)eval(O) can be covered by countably many sets of diameter less than ϵ\epsilon. Since XX is hereditarily Lindelöf there is a countable subfamily \mathcal{B} of 𝒜\mathcal{A} such that A𝒜A=BB\bigcup\limits_{A\in\mathcal{A}}A=\bigcup\limits_{B\in\mathcal{B}}B. Take V=defA𝒜AV\buildrel\rm def\over{=}\bigcup\limits_{A\in\mathcal{A}}A. Observe that VV is the largest element of 𝒜\mathcal{A}. Let’s see that A=defXVA\buildrel\rm def\over{=}X\setminus V is empty. Assume that AA\neq\emptyset. Then there is a nonempty set UU of XX such that UAU\cap A\neq\emptyset and dd_{\infty}-diam(eval(UA))ϵdiam(eval(U\cap A))\leq\epsilon. Since eval(UV)=eval(UA)eval(V)eval(U\cup V)=eval(U\cap A)\cup eval(V) we know that eval(UV)eval(U\cup V) can be covered by countably many sets of diameter less than ϵ\epsilon. So, UV𝒜U\cup V\in\mathcal{A} and we arrive to a contradiction because U(XV)U\cap(X\setminus V)\neq\emptyset. Since X=V𝒜X=V\in\mathcal{A} and ϵ\epsilon was arbitrary eval(X)eval(X) is separable.

There is a dense and countable subset DD of eval(X)eval(X). We know that DD separates points of HH because eval(X)eval(X) also separates points. Let ΔD:HMD\Delta D:H\rightarrow M^{D} be the diagonal product. Since ΔD\Delta D is an embedding and MDM^{D} is metrizable we conclude that HH is metrizable. ∎

Next result is due basically to Namioka [15, Lemma 2.1]. It can also be found in [8, Lemma 6.4.], where the reference to Namioka is acknowledged. Again, we include a sketch of the proof here for completeness sake.

Lemma 4.7.

Let XX, YY and (M,d)(M,d) be two arbitrary compact spaces and a metric space, respectively, and let GG be a subset of C(Y,M)C(Y,M). Suppose that p:XYp:X\longrightarrow Y is a continuous onto map. Then Gp=def{gp:gG}C(X,M)G\circ p\buildrel\rm def\over{=}\{g\circ p:g\in G\}\subseteq C(X,M) is hereditarily almost equicontinuous if and only if GG is also hereditarily almost equicontinuous.

Proof.

In order to prove this result, we will apply Lemma 3.2. Assume that GpG\circ p is hereditarily almost equicontinuous. Let AA be a closed (and compact) subset of YY, UU be a nonempty relatively open set in A and ϵ>0\epsilon>0. By Zorn’s Lemma, there exists a minimal compact subset ZZ of XX such that p(Z)=Ap(Z)=A. Since U~=defp1(U)Z\tilde{U}\buildrel\rm def\over{=}p^{-1}(U)\cap Z is a nonempty relatively open set in ZZ and (Gp)|Z(G\circ p)|_{Z} is almost equicontinuous there is a nonempty relatively open set V~U~\tilde{V}\subseteq\tilde{U} in ZZ such that diam((gp)(V~))<ϵdiam((g\circ p)(\tilde{V}))<\epsilon for all gGg\in G. Let V=defAp(ZV~)V\buildrel\rm def\over{=}A\setminus p(Z\setminus\tilde{V}), that is relatively open set in AA. We claim that VV\neq\emptyset. Indeed, assume that V=V=\emptyset. Then A=p(ZV~)A=p(Z\setminus\tilde{V}) and this contradicts the minimality of ZZ. Since Vp(V~)V\subseteq p(\tilde{V}) we have that diam(g(V))<ϵdiam(g(V))<\epsilon for all gGg\in G.

Conversely, let ZZ be a closed subset of XX, U~\tilde{U} be a nonempty relatively open set in ZZ and ϵ>0\epsilon>0. Consider the closed subset W0=defp(U~)¯W_{0}\buildrel\rm def\over{=}\overline{p(\tilde{U})} of YY. Since G|W0G|_{W_{0}} is almost equicontinuous there is a nonempty relatively open set V0V_{0} in YY such that V0W0V_{0}\cap W_{0}\neq\emptyset and diam(g(V0W0))<ϵdiam(g(V_{0}\cap W_{0}))<\epsilon for all gGg\in G. Take V~=defp1(V0)U~\tilde{V}\buildrel\rm def\over{=}p^{-1}(V_{0})\cap\tilde{U}. Since V~\tilde{V} is a nonempty relatively open set in ZZ and p(V~)V0W0p(\tilde{V})\subseteq V_{0}\cap W_{0} we conclude that diam(g(p(V~)))<ϵdiam(g(p(\tilde{V})))<\epsilon for all gGg\in G. ∎

Remark 4.8.

If the map pp of the previous lemma is open or quasi-open we obtain the same result for almost equicontinuity. Recall that a map f:XYf:X\rightarrow Y between two topological spaces is quasi-open if for any nonempty open set UXU\subseteq X the interior of f(U)f(U) in YY is nonempty.

Proof.

Let UU be a nonempty open set of YY and ϵ>0\epsilon>0. Since GpG\circ p is almost equicontinuous and U~=p1(U)\tilde{U}=p^{-1}(U) is an open subset of XX there is a nonempty open subset V~U~\tilde{V}\subseteq\tilde{U} of XX such that diam((gp)(V~))<ϵdiam((g\circ p)(\tilde{V}))<\epsilon for all gGg\in G. Since the nonempty open set V=defint(p(V~))V\buildrel\rm def\over{=}int(p(\tilde{V})) is included in p(V~)p(\tilde{V}) we have that diam(g(V))<ϵdiam(g(V))<\epsilon for all gGg\in G.

Conversely, let U~\tilde{U} be a nonempty open set of XX and ϵ>0\epsilon>0. Take U=defint(p(U~))U\buildrel\rm def\over{=}int(p(\tilde{U}))\neq\emptyset. Since GG is almost equicontinuous there is a nonempty open subset VUV\subseteq U of YY such that diam(g(V))<ϵdiam(g(V))<\epsilon for all gGg\in G. So, taking the open subset V~=defp1(V)U~\tilde{V}\buildrel\rm def\over{=}p^{-1}(V)\cap\tilde{U}, we conclude that diam((gp)(V~))<ϵdiam((g\circ p)(\tilde{V}))<\epsilon for all gGg\in G. ∎

Proposition 4.9.

Let XX be a Čech-complete space, (M,d)(M,d) be a hemicompact metric space and GC(X,M)G\subseteq C(X,M) such that G¯MX\overline{G}^{M^{X}} is compact. Then the following conditions are equivalent:

  1. (a)

    GG is hereditarily almost equicontinuous.

  2. (b)

    LL is hereditarily almost equicontinuous on FF, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable and compact subset of XX.

Proof.

(a)(a) implies (b)(b) is trivial. To see the other implication, assume, reasoning by contradiction, that (a) does not hold. Then there must be some closed subset AXA\subseteq X such that G|AG|_{A} is not almost equicontinuous. By Lemma 4.1 there exists a compact and separable subset FF of XX, an onto and continuous map Ψ:F2ω\Psi:F\rightarrow 2^{\omega}, and a countable subset LL of GG such that the subset LC(2ω,M)L^{*}\subseteq C(2^{\omega},M) defined by l(Ψ(x))=l(x)l^{*}(\Psi(x))=l(x) for all xFx\in F is not almost equicontinuous. Therefore, LL is not hereditarily almost equicontinuous on FF by Lemma 4.7 and we arrive to a contradiction. ∎

We can now prove Theorem B.

Proof of Theorem B.

(b)(a)(b)\Rightarrow(a) is a direct consequence of Proposition 4.9 and Corollary 3.11.

(a)(c)(a)\Rightarrow(c) Let L[G]ωL\in[G]^{\leq\omega} and let FF be a separable and compact subset of XX. LL defines an equivalence relation on FF by xyx\sim y if and only if l(x)=l(y)l(x)=l(y) for all lLl\in L. If F~=F/\widetilde{F}=F/{\sim} is the compact quotient space and p:FF~p:F\rightarrow\widetilde{F} denotes the canonical quotient map, each lLl\in L has associated a map l~C(F~,M)\tilde{l}\in C(\widetilde{F},M) defined as l~(x~)=defl(x)\tilde{l}(\tilde{x})\buildrel\rm def\over{=}l(x) for any xFx\in F with p(x)=x~p(x)=\tilde{x}. Furthermore, if L~=def{l~:lL}\tilde{L}\buildrel\rm def\over{=}\{\widetilde{l}:l\in L\}, we can extend this definition to the closure of L~\tilde{L} in MF~M^{\widetilde{F}}. Thus, each lL¯MFl\in\overline{L}^{M^{F}} has associated a map l~L~¯MF~\tilde{l}\in\overline{\tilde{L}}^{M^{\widetilde{F}}} such that l~p=l\tilde{l}\circ p=l. By construction, we have that L~\tilde{L} separates the points in F~\widetilde{F}. Since L~\tilde{L} is countable it follows that (F~,tp(L~))(\widetilde{F},t_{p}(\tilde{L})) is a compact metric space. On the other hand, GG is hereditarily almost equicontinuous on XX. Applying Lemma 4.7 to FF and F~\widetilde{F}, it follows that L~\widetilde{L} is hereditarily almost equicontinuous on F~\widetilde{F}. Therefore, the space L~¯MF~\overline{\tilde{L}}^{M^{\widetilde{F}}} is metrizable by Proposition 4.6. In order to finish the proof, it suffices to remark that L¯MF\overline{L}^{M^{F}} is canonically homeomorphic to L~¯MF~\overline{\tilde{L}}^{M^{\widetilde{F}}} (see Fact 5.7).

(c)(d)(c)\Rightarrow(d) Let L[G]ωL\in[G]^{\leq\omega} and let FF be a separable and compact subset of XX. We know that H=def((L¯MX)|F,tp(F))H\buildrel\rm def\over{=}((\overline{L}^{M^{X}})|_{F},t_{p}(F)) is compact metric. Since FF is separable, we have that l(F)l(F) is a separable for every lLl\in L. Hence N=deflLl(F)¯MN\buildrel\rm def\over{=}\overline{\bigcup\limits_{l\in L}l(F)}^{M} is a separable subset of MM. Now, remark that MM can be replaced by NN without loss of generality. On the other hand, since FC(H,M)F\subseteq C(H,M) and HH is compact metric, it follows that (F,t(H))(F,t_{\infty}(H)) is separable and metrizable by [6, Cor. 4.2.18], which implies that (F,t(H))(F,t_{\infty}(H)) is Lindelöf. Since the the topology tp(H)t_{p}(H) is weaker than t(H)t_{\infty}(H), we deduce that (F,tp(H))(F,t_{p}(H)) must be Lindelöf.

(d)(b)(d)\Rightarrow(b) By Lemma 3.9, for all L[G]ωL\in[G]^{\leq\omega} and FF a separable compact subset of XX, we have that (F,tp(ν(L¯MX×K)))(F,t_{p}(\nu(\overline{L}^{M^{X}}\times K))) is Lindelöf . Applying [3, Corollary D], it follows that ν(L¯MX×K)\nu(\overline{L}^{M^{X}}\times K) is hereditarily almost equicontinuous for all L[G]ωL\in[G]^{\leq\omega} and FF a separable compact subset of XX. Thus, Corollary 3.11 yields (b)(b). ∎

5. Appendix

It is well known that for every compact metric space (M,d)(M,d), there is a canonical continuous one-to-one mapping M:M[0,1]ω\mathcal{E}_{M}:M\longrightarrow[0,1]^{\omega} that embeds MM into [0,1]ω[0,1]^{\omega} as a closed subspace. Let ρn:[1,1][0,1]\rho_{n}:[-1,1]\longrightarrow[0,1] the map defined by ρn(r)=|r|2n\rho_{n}(r)=\frac{|r|}{2^{n}} for every n<ωn<\omega. Along this paper, we will consider that [0,1]ω[0,1]^{\omega} is equipped with the metric ρ\rho defined by

ρ((xn),(yn))=n<ωρn(xnyn)\rho((x_{n}),(y_{n}))=\sum_{n<\omega}\rho_{n}(x_{n}-y_{n})

The proof of the following lemma is obtained by a standard argument of compactness, using the continuity of M1\mathcal{E}_{M}^{-1} and that every continuous map defined on a compact space is uniformly continuous. We omit its proof here.

Fact 5.1.

Let (M,d)(M,d) be a compact metric space. Let M:M[0,1]ω\mathcal{E}_{M}:M\longrightarrow[0,1]^{\omega} denote its attached embedding into [0,1]ω[0,1]^{\omega}, and let πn:[0,1]ω[0,1]\pi_{n}:[0,1]^{\omega}\rightarrow[0,1] denote the nnth canonical projection. Then, for every ϵ>0\epsilon>0, there is δ>0\delta>0 and n0<ωn_{0}<\omega such that if (x,y)M×M(x,y)\in M\times M and ρn(πn(M(x))πn(M(y)))<δ/2n0\rho_{n}(\pi_{n}(\mathcal{E}_{M}(x))-\pi_{n}(\mathcal{E}_{M}(y)))<\delta/2n_{0} for nn0n\leq n_{0} then d(x,y)<ϵd(x,y)<\epsilon.

We know recall some simple remarks that will be used along the paper.

Fact 5.2.

Let XX be a topological space and (M,d)(M,d) a compact metric space. If πn\pi_{n} is the nnth projection mapping defined above, then the following map is continuous if we consider that the two spaces have the topology of pointwise convergence.

πn:MX[0,1]X\pi_{n}^{*}:M^{X}\rightarrow[0,1]^{X}

defined by πn(f)=defπnMf\pi_{n}^{*}(f)\buildrel\rm def\over{=}\pi_{n}\circ\mathcal{E}_{M}\circ f, fMXf\in M^{X}, for each n<ωn<\omega.

For each SMXS\subseteq M^{X} and each n<ωn<\omega we define Sn=defπn(S)S_{n}\buildrel\rm def\over{=}\pi_{n}^{*}(S).

Fact 5.3.

Let XX be a Baire space, (M,d)(M,d) be a compact metric space, GC(X,M)G\subseteq C(X,M) and H=defG¯MXH\buildrel\rm def\over{=}\overline{G}^{M^{X}}. Then Hn=Gn¯[0,1]XH_{n}=\overline{G_{n}}^{[0,1]^{X}}.

Proof.

Indeed, since πn\pi_{n}^{*} is continuous we have that Hn=πn(H)=πn(G¯MX)πn(G)¯[0,1]X=Gn¯[0,1]XH_{n}=\pi_{n}^{*}(H)=\pi_{n}^{*}(\overline{G}^{M^{X}})\subseteq\overline{\pi_{n}^{*}(G)}^{[0,1]^{X}}=\overline{G_{n}}^{[0,1]^{X}}. For the reverse inclusion, remark that Gn¯[0,1]X\overline{G_{n}}^{[0,1]^{X}} is the smallest closed subset that contains GnG_{n} and GnHnG_{n}\subseteq H_{n}. ∎

Fact 5.4.

Let XX be a Baire space, (M,d)(M,d) be a compact metric space and GC(X,M)G\subseteq C(X,M). If GnG_{n} is almost equicontinuous for every n<ωn<\omega, then GG is almost equicontinuous.

Proof.

For each nωn\in\omega there exists a dense GδG_{\delta} subset DnD_{n} of XX such that GnG_{n} is equicontinuous on DnD_{n}. Since XX is a Baire space, the D=n<ωDnD=\bigcap\limits_{n<\omega}D_{n} is dense in XX. We claim that GG is equicontinuous in DD. Indeed, let x0Dx_{0}\in D and ϵ>0\epsilon>0. By Fact 5.1 we get δ>0\delta>0 and n0<ωn_{0}<\omega. Take ϵ0=δ2n0\epsilon_{0}=\frac{\delta}{2n_{0}}. For each n<n0n<n_{0}, being GnG_{n} equicontinuous in x0x_{0}, there is an open neighbourhood UnU_{n} of x0x_{0} such that |gn(x0)gn(x)|<ϵ0|g_{n}(x_{0})-g_{n}(x)|<\epsilon_{0} for all xUnx\in U_{n} and gnGng_{n}\in G_{n}. Consider the open neighbourhood U=n<n0UnU=\bigcap\limits_{n<n_{0}}U_{n} of x0x_{0}. So, let an arbitrary gGg\in G and xUx\in U, then ρn(πn(M(g(x0)))πn(M(g(x))))=ρn(πn(g)(x0)πn(g)(x))=|πn(g)(x0)πn(g)(x)|2n<ϵ02nδ2n0\rho_{n}(\pi_{n}(\mathcal{E}_{M}(g(x_{0})))-\pi_{n}(\mathcal{E}_{M}(g(x))))=\rho_{n}(\pi^{*}_{n}(g)(x_{0})-\pi^{*}_{n}(g)(x))=\frac{|\pi^{*}_{n}(g)(x_{0})-\pi^{*}_{n}(g)(x)|}{2^{n}}<\frac{\epsilon_{0}}{2^{n}}\leq\frac{\delta}{2n_{0}}. Consequently, d(g(x0),g(x))<ϵd(g(x_{0}),g(x))<\epsilon by Fact 5.1. ∎

Fact 5.5.

The diagonal map Δ:Hn<ωHn\Delta:H\rightarrow\prod\limits_{n<\omega}H_{n} defined by Δ(h)=(πnMh)n<ω\Delta(h)=(\pi_{n}\circ\mathcal{E}_{M}\circ h)_{n<\omega} for each hHh\in H, is a homeomorphism of HH onto its image.

Fact 5.6.

Given a subset LGL\subseteq G, it defines an equivalence relation on XX by xyx\sim y if and only if l(x)=l(y)l(x)=l(y) for all lLl\in L. Let X~=X/\widetilde{X}=X/{\sim} be the quotient space and let p:XX~p:X\rightarrow\widetilde{X} denote the canonical quotient map, then each lLl\in L has associated a map l~C(X~,M)\tilde{l}\in C(\widetilde{X},M) defined as l~(x~)=defl(x)\tilde{l}(\tilde{x})\buildrel\rm def\over{=}l(x) for any xXx\in X with p(x)=x~p(x)=\tilde{x}. Furthermore, if L~=def{l~:lL}\tilde{L}\buildrel\rm def\over{=}\{\widetilde{l}:l\in L\}, we can extend this definition to the closure of L~\tilde{L}. Thus, each lL¯MXl\in\overline{L}^{M^{X}} has associated a map l~L~¯MX~\tilde{l}\in\overline{\tilde{L}}^{M^{\widetilde{X}}} such that l~p=l\tilde{l}\circ p=l.

Fact 5.7.

Let LL be a countable subset of GC(X,M)G\subseteq C(X,M). We denote by XLX_{L} the topological space (X~,tp(L~))(\widetilde{X},t_{p}(\tilde{L})), which is metrizable because L~\tilde{L} is countable. Consider the map p:(MX~,tp(X~))(MX,tp(X))p^{*}:(M^{\tilde{X}},t_{p}(\tilde{X}))\rightarrow(M^{X},t_{p}(X)) defined by p(f~)=f~pp^{*}(\tilde{f})=\tilde{f}\circ p, for each f~MX~\tilde{f}\in M^{\tilde{X}}. Then pp^{*} is a homeomorphism of L~¯MX~\overline{\tilde{L}}^{M^{\tilde{X}}} onto L¯MX\overline{L}^{M^{X}}.

Proof.

We observe that pp^{*} is continuous, since a net {fα~}αA\{\tilde{f_{\alpha}}\}_{\alpha\in A} tp(X~)t_{p}(\tilde{X})-converges to f~\tilde{f} in L~¯MX~\overline{\tilde{L}}^{M^{\tilde{X}}} if and only if {fα~p}αA\{\tilde{f_{\alpha}}\circ p\}_{\alpha\in A} tp(X)t_{p}(X)-converges to f~p\tilde{f}\circ p in L¯MX\overline{L}^{M^{X}}.
Let’s see that p(L~¯MX~)=L¯MXp^{*}(\overline{\tilde{L}}^{M^{\tilde{X}}})=\overline{L}^{M^{X}}. Indeed, since pp^{*} is continuous we have that p(L~¯MX~)p(L~)¯MX=L¯MXp^{*}(\overline{\tilde{L}}^{M^{\tilde{X}}})\subseteq\overline{p^{*}(\tilde{L})}^{M^{X}}=\overline{L}^{M^{X}}. We have the other inclusion because L¯MX\overline{L}^{M^{X}} is the smaller closed set that contains LL and Lp(L~¯MX~)L\subseteq p^{*}(\overline{\tilde{L}}^{M^{\tilde{X}}}).
Let f~,g~L~¯MX~\tilde{f},\tilde{g}\in\overline{\tilde{L}}^{M^{\tilde{X}}} such that f~g~\tilde{f}\neq\tilde{g}. Then there exists x~X~\tilde{x}\in\tilde{X} such that f~(x~)g~(x~)\tilde{f}(\tilde{x})\neq\tilde{g}(\tilde{x}). Let xXx\in X an element such that x~=p(x)\tilde{x}=p(x). Thus (f~p)(x)(g~p)(x)(\tilde{f}\circ p)(x)\neq(\tilde{g}\circ p)(x). So, pp^{*} is injective because f~pg~p\tilde{f}\circ p\neq\tilde{g}\circ p.
Finally, we arrive to the conclusion that p|L~¯MX~p^{*}|_{\overline{\tilde{L}}^{M^{\tilde{X}}}} is a homeomorphism because it is defined between compact spaces. ∎

6. Acknowledgments

We are very grateful to the referee for a thorough report that helped considerably in improving the presentation of this paper.

References

  • [1] E Akin, J Auslander, and K Berg. Almost equicontinuity and the enveloping semigroup. Contemp. Math., 215:75–81, 1998.
  • [2] AV Arkhangel’skii. Problems in Cp-theory. Open Problems in Topology, North-Holland, 1990.
  • [3] B Cascales, I Namioka, and G Vera. The Lindelöf property and fragmentability. Proc. Amer. Math. Soc., 128(11):3301–3309, 2000.
  • [4] JPR Christensen. Joint Continuity of separately continuous functions. Proceedings of the American Mathematical Society, 82(3):455–461, 1981.
  • [5] HH Corson and I Glicksberg. Compactness in Hom (G, H). Canad. J. Math., 22:164–170, 1970.
  • [6] R Engelking. General topology. Heldermann Verlag, 1989.
  • [7] MV Ferrer and S Hernández. Dual topologies on non-abelian groups. Topology and its Applications, 159(9):2367–2377, 2012.
  • [8] E Glasner and M Megrelishvili. Linear representations of hereditarily non-sensitive dynamical systems. arXiv preprint math.DS/0406192, 2004.
  • [9] E Glasner and M Megrelishvili. Hereditarily non-sensitive dynamical systems and linear representations. Colloq. Math., 104(2):223–283, 2006.
  • [10] E Glasner and B Weiss. Locally equicontinuous dynamical systems. Colloq. Math, 84/85(2):345–361, 2000.
  • [11] E L I Glasner, Michael Megrelishvili, and Vladimir V Uspenskij. On metrizable enveloping semigroups. Israel Journal of Mathematics, 164(1):317–332, 2006.
  • [12] JE Jayne and CA Rogers. Borel selectors for upper semi-continuous set-valued maps. Acta Math., 155(1-2):41–79, 1985.
  • [13] P Kenderov and W Moors. Separate continuity, joint continuity and the Lindelöf property. Proceedings of the American Mathematical Society, 2006.
  • [14] I Namioka. Separate continuity and joint continuity. Pacific J. Math., 51(2):515–531, 1974.
  • [15] I Namioka. Radon Nikodym compact spaces and fragmentability. Mathematika, 34(2):258–281, 1987.
  • [16] Z Piotrowski. Separate and joint continuity. Real Anal. Exchange, 1985.
  • [17] NK Ribarska. Internal characterization of fragmentable spaces. Mathematika, 34(2):243–257, 1987.
  • [18] M Talagrand. Deux généralisations d’un théoreme de I. Namioka. Pacific J. Math, 81(1):239–251, 1979.
  • [19] L Tárrega. El teorema de Namioka y algunas generalizaciones. Master Thesis, June, 2014.
  • [20] JP Troallic. Sequential criteria for equicontinuity and uniformities on topological groups. Topology and its Applications, 68(1):83–95, 1996.