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

Absolute \mathcal{F}-Borel classes

Vojtěch Kovařík
E-mail: kovarikv@karlin.mff.cuni.cz

Charles University
Faculty of Mathematics and Physics
Department of Mathematical Analysis

Sokolovská 83, Karlín
Praha 8, 186 00
Czech Republic
2010 Mathematics Subject Classification: Primary 54H05; Secondary 54G20.Key words and phrases: compactification, Borel set, F-Borel class, absolute complexity.

We investigate and compare \mathcal{F}-Borel classes and absolute \mathcal{F}-Borel classes. We provide precise examples distinguishing these two hierarchies. We also show that for separable metrizable spaces, \mathcal{F}-Borel classes are automatically absolute.

1 Introduction

Borel sets (in some topological space XX) are the smallest family containing the open sets, which is closed under the operations of taking countable intersections, countable unions and complements (in XX). The family of \mathcal{F}-Borel (resp. 𝒢\mathcal{G}-Borel) sets is the smallest family containing all closed (resp. open) sets, which is closed under the operations of taking countable intersections and countable unions of its elements. In metrizable spaces the families of Borel, \mathcal{F}-Borel and 𝒢\mathcal{G}-Borel sets coincide. In non-metrizable spaces, open set might not necessarily be a countable union of closed sets, so we need to make a distinction between Borel, \mathcal{F}-Borel and 𝒢\mathcal{G}-Borel sets.

In the present paper, we investigate absolute \mathcal{F}-Borel classes. While Borel classes are absolute by [HS] (see Proposition 1.8 below), by [Tal] it is not the case for \mathcal{F}-Borel classes (see Theorem 1.7 below). We develop a method of estimating absolute Borel classes from above, which enables us to compute the exact complexity of Talagrand’s examples and to provide further examples by modifying them. Our main results are Theorem 5.14 and Corollary 5.15.

The paper is organized as follows: In the rest of the introductory section, we define the Borel, \mathcal{F}-Borel and 𝒢\mathcal{G}-Borel hierarchies and recall some basic results. In Section 2, we recall the definitions and basic results concerning compactifications and their ordering. In Theorem 2.3, we show that the complexity of separable metrizable spaces coincides with their absolute complexity (in any of the hierarchies). Section 3 is devoted to showing that absolute complexity is inherited by closed subsets. In Section 4, we study special sets of sequences of integers – trees, sets which extend to closed discrete subsets of ωω\omega^{\omega} and the ‘broom sets’ introduced by Talagrand. We study in detail the hierarchy of these sets using the notion of rank. In Section 5, we introduce the class of examples of spaces used by Talagrand. We investigate in detail their absolute complexity and in Section 5.3, we prove our main results.

Let us start by defining the basic notions. Throughout the paper, all the spaces will be Tychonoff. For a family of sets 𝒞\mathcal{C}, we will denote by 𝒞σ\mathcal{C}_{\sigma} the collection of all countable unions of elements of 𝒞\mathcal{C} and by 𝒞δ\mathcal{C}_{\delta} the collection of all countable intersections of elements of 𝒞\mathcal{C}.

Definition 1.1 (Borel classes).

Let XX be a topological space. We define the Borel multiplicative classes α(X)\mathcal{M}_{\alpha}(X) and Borel additive classes 𝒜α(X)\mathcal{A}_{\alpha}(X), α<ω1\alpha<\omega_{1}, as follows:

  • 0(X)=𝒜0(X):=\mathcal{M}_{0}(X)=\mathcal{A}_{0}(X):= the algebra generated by open subsets of XX,

  • α(X):=(β<α(β(X)𝒜β(X)))δ\mathcal{M}_{\alpha}(X):=\left(\underset{\beta<\alpha}{\bigcup}\left(\mathcal{M}_{\beta}(X)\cup\mathcal{A}_{\beta}(X)\right)\right)_{\delta} for 1α<ω11\leq\alpha<\omega_{1},

  • 𝒜α(X):=(β<α(β(X)𝒜β(X)))σ\mathcal{A}_{\alpha}(X):=\left(\underset{\beta<\alpha}{\bigcup}\left(\mathcal{M}_{\beta}(X)\cup\mathcal{A}_{\beta}(X)\right)\right)_{\sigma} for 1α<ω11\leq\alpha<\omega_{1}.

In any topological space XX, we have the families of closed sets and open sets, denoted by F(X)F(X) and G(X)G(X), and we can continue with classes Fσ(X)F_{\sigma}(X), Gδ(X)G_{\delta}(X), Fσδ(X)F_{\sigma\delta}(X) and so on. However, this notation quickly gets impractical, so we use the following notation.

Definition 1.2 (\mathcal{F}-Borel and 𝒢\mathcal{G}-Borel classes).

We define the hierarchy of \mathcal{F}-Borel sets on a topological space XX as follows:

  • 1(X):=\mathcal{F}_{1}(X):= closed subsets of XX,

  • α(X):=(β<αβ(X))σ\mathcal{F}_{\alpha}(X):=\left(\underset{\beta<\alpha}{\bigcup}\mathcal{F}_{\beta}(X)\right)_{\sigma} for 2α<ω12\leq\alpha<\omega_{1} even,

  • α:=(β<αβ(X))δ\mathcal{F}_{\alpha}:=\left(\underset{\beta<\alpha}{\bigcup}\mathcal{F}_{\beta}(X)\right)_{\delta} for 3α<ω13\leq\alpha<\omega_{1} odd.

The sets of α\alpha-th 𝒢\mathcal{G}-Borel class, 𝒢α(X)\mathcal{G}_{\alpha}(X), are the complements of α(X)\mathcal{F}_{\alpha}(X) sets.

By a descriptive class of sets, we will always understand one of the Borel, \mathcal{F}-Borel or 𝒢\mathcal{G}-Borel classes.

Remark 1.3.

(i)(i) In any topological space XX, we have

1(X)=F(X)0(X).\mathcal{F}_{1}(X)=F(X)\subset\mathcal{M}_{0}(X).

It follows that each class n\mathcal{F}_{n}, nn\in\mathbb{N}, is contained in 𝒜n1\mathcal{A}_{n-1} or n1\mathcal{M}_{n-1} (depending on the parity). The same holds for classes α\mathcal{F}_{\alpha}, αω\alpha\geq\omega, except that the difference between ranks disappears.

(ii)(ii) If XX is metrizable, the Borel, \mathcal{F}-Borel and 𝒢\mathcal{G}-Borel classes are related to the standard Borel hierarchy [Kech, ch.11]. In particular, each α(X)\mathcal{F}_{\alpha}(X) is equal to either Σα0(X)\Sigma^{0}_{\alpha}(X) or Πα0(X)\Pi^{0}_{\alpha}(X) (depending on the parity of α\alpha) and we have

F(X)G(X)0(X)=𝒜0(X)Fσ(X)Gδ(X).F(X)\cup G(X)\subset\mathcal{M}_{0}(X)=\mathcal{A}_{0}(X)\subset F_{\sigma}(X)\cap G_{\delta}(X).

The relations between our descriptive classes can then be summarized as:

  • n(X)=Πn0(X)=n1(X)\mathcal{F}_{n}(X)=\Pi^{0}_{n}(X)=\mathcal{M}_{n-1}(X) holds if 1n<ω1\leq n<\omega is odd,

  • n(X)=Σn0(X)=𝒜n1(X)\mathcal{F}_{n}(X)=\Sigma^{0}_{n}(X)=\mathcal{A}_{n-1}(X) holds if 2n<ω2\leq n<\omega is even,

  • α(X)=Σα0(X)=𝒜α(X)\mathcal{F}_{\alpha}(X)=\Sigma^{0}_{\alpha}(X)=\mathcal{A}_{\alpha}(X) holds if ωα<ω1\omega\leq\alpha<\omega_{1} is even,

  • α(X)=Πα0(X)=α(X)\mathcal{F}_{\alpha}(X)=\Pi^{0}_{\alpha}(X)=\mathcal{M}_{\alpha}(X) holds if ωα<ω1\omega\leq\alpha<\omega_{1} is odd.

Clearly, the 𝒢\mathcal{G}-Borel classes satisfy the dual version of (i)(i) and (ii)(ii).

Note also that in compact spaces, closed sets are compact, so in this context the \mathcal{F}-Borel sets are sometimes called 𝒦\mathcal{K}-Borel, FσF_{\sigma} sets are called KσK_{\sigma} and so on.

Let us define two notions central to the topic of our paper.

Definition 1.4.

Let XX be a (Tychonoff) topological space and 𝒞\mathcal{C} be a descriptive class of sets. We say that XX is a 𝒞\mathcal{C} space if there exists a compactification cXcX of XX, such that X𝒞(cX)X\in\mathcal{C}(cX).

If X𝒞(Y)X\in\mathcal{C}(Y) holds for any Tychonoff topological space YY in which XX is densely embedded, we say that XX is an absolute 𝒞\mathcal{C} space We call the class 𝒞\mathcal{C} absolute if every 𝒞\mathcal{C} space is an absolute 𝒞\mathcal{C} space.

The basic relation between complexity and absolute complexity are noted in the following remark.

Remark 1.5.

Consider the following statements:

  1. (i)

    X𝒞(cX)X\in\mathcal{C}(cX) holds for some compactification cXcX;

  2. (ii)

    X𝒞(βX)X\in\mathcal{C}(\beta X) holds for the Čech-Stone compactification;

  3. (iii)

    X𝒞(cX)X\in\mathcal{C}(cX) holds for every compactification cXcX;

  4. (iv)

    X𝒞(Y)X\in\mathcal{C}(Y) holds for every Tychonoff space where YY is densely embedded;

  5. (v)

    X𝒞(Y)X\in\mathcal{C}(Y) holds for every Tychonoff space where YY is embedded.

Clearly, the implications (v)(iv)(iii)(ii)(i)(v)\implies(iv)\implies(iii)\implies(ii)\implies(i) always hold. In the opposite direction, we always have (i)(ii)(i)\implies(ii) (this is standard, see Remark 2.1) and (iii)(iv)(iii)\implies(iv). For Borel and \mathcal{F}-Borel classes, (iv)(iv) is equivalent to (v)(v) (since these classes are closed under taking intersections with closed sets). For 𝒢\mathcal{G}-Borel classes, (iv)(iv) is never equivalent to (v)(v) (just take YY so large that X¯Y\overline{X}^{Y} is not GδG_{\delta} in YY).

The interesting part is therefore the relation between (i)(i) and (iii)(iii) from Remark 1.5. For the first two levels of \mathcal{F}-Borel and 𝒢\mathcal{G}-Borel hierarchies, (i)(i) is equivalent to (iii)(iii), that is, these classes are absolute. A more precise formulation is given in the following remark.

Remark 1.6.

For a topological space XX we have

  1. (i)

    XX is an 1\mathcal{F}_{1} space \iff XX is absolutely 1\mathcal{F}_{1} \iff XX is compact;

  2. (ii)

    XX is an 2\mathcal{F}_{2} space \iff XX is absolutely 2\mathcal{F}_{2} \iff XX is σ\sigma-compact;

  3. (iii)

    XX is a 𝒢1\mathcal{G}_{1} space \iff XX is absolutely 𝒢1\mathcal{G}_{1} \iff XX is locally compact;

  4. (iv)

    XX is a 𝒢2\mathcal{G}_{2} space \iff XX is absolutely 𝒢2\mathcal{G}_{2} \iff XX is Čech-complete

(for the proofs of (iii)(iii) and (iv)(iv) see [Eng, Theorem 3.5.8 and Theorem 3.9.1]).

The first counterexample (and the only one known to the author) was found by Talagrand, who showed that already the class of FσδF_{\sigma\delta} sets is not absolute:

Theorem 1.7 ([Tal]).

There exists an 3\mathcal{F}_{3} space TT and its compactification KK, such that TT is not \mathcal{F}-Borel in KK.

This not only shows that none of the classes α\mathcal{F}_{\alpha} for α4\alpha\geq 4 are absolute, but also that the difference between absolute and non-absolute complexity can be large. Indeed, the ‘non-absolute complexity’ of TT is ‘3\mathcal{F}_{3}’, but its ‘absolute complexity’ is ‘𝒦\mathcal{K}-analytic’. In [KK], the authors give a sufficient condition for an 3\mathcal{F}_{3} space to be absolutely 3\mathcal{F}_{3}, but a characterization of absolutely α\mathcal{F}_{\alpha} spaces is still unknown for all α3\alpha\geq 3.

In [HS] and [Raj], the authors studied (among other things) absoluteness of Borel classes. In particular, the following result is relevant to our topic:

Proposition 1.8 ([HS, Corollary 14]).

For every 1α<ω11\leq\alpha<\omega_{1}, the classes 𝒜α\mathcal{A}_{\alpha} and α\mathcal{M}_{\alpha} are absolute.

Another approach is followed in [MP] and [JK], where the absoluteness is investigated in the metric setting (that is, for spaces which are of some class 𝒞\mathcal{C} when embedded into any metric space). In [JK] a characterization of ‘metric-absolute’ 3\mathcal{F}_{3} spaces is given in terms of existence complete sequences of covers (a classical notion used by Frolík for characterization of Čech-complete spaces, see [Fro1]) which are σ\sigma-discrete. Unfortunatelly, this is not applicable to the topological version of absoluteness, because every countable cover is (trivially) σ\sigma-discrete, and any 3\mathcal{F}_{3} space does have such a cover by [Fro2] – even the Talagrand’s non-absolute space.

2 Compactifications and their ordering

By a compactification of a topological space XX we understand a pair (cX,φ)(cX,\varphi), where cXcX is a compact space and φ\varphi is a homeomorphic embedding of XX onto a dense subspace of cXcX. Symbols cXcX, dXdX and so on will always denote compactifications of XX.

Compactification (cX,φ)(cX,\varphi) is said to be larger than (dX,ψ)(dX,\psi), if there exists a continuous mapping f:cXdXf:cX\rightarrow dX, such that ψ=fφ\psi=f\circ\varphi. We denote this as cXdXcX\succeq dX. Recall that for a given T31/2T_{3\,^{1}\!/_{2}} topological space XX, its compactifications are partially ordered by \succeq and Stone-Čech compactification βX\beta X is the largest one.

Often, we encounter a situation where XcXX\subset cX and the corresponding embedding is identity. In this case, we will simply write cXcX instead of (cX,id|X)(cX,\textrm{id}|_{X}).

Much more about this topic can be found in many books, see for example [Fre]. The basic relation between the complexity of a space XX and the ordering of compactifications is the following:

Remark 2.1.

If 𝒞\mathcal{C} is a descriptive classes of sets, we have

X𝒞(dX),cXdXX𝒞(cX).X\in\mathcal{C}(dX),\ cX\succeq dX\implies X\in\mathcal{C}(cX).

In particular, XX is a 𝒞\mathcal{C} space if and only if X𝒞(βX)X\in\mathcal{C}(\beta X).

We will also make use of the following result about existence of small metrizable compactifications.

Proposition 2.2.

Let XX be a separable metrizable space and cXcX its compactification. Then XX has some metrizable compactification dXdX, such that dXcXdX\preceq cX.

This proposition is an exercise, so we only include a sketch of its proof:

Proof.

We can assume that cX[0,1]κcX\subset[0,1]^{\kappa} for some κ\kappa. Since XX has a countable base, there is a countable set of coordinates IκI\subset\kappa such that the family of all VXV\cap X, where VV is a basic open set V[0,1]κV\subset[0,1]^{\kappa} depending on coordinates in II only is a base of XX. Then we can take dX:=πI(cX)dX:=\pi_{I}(cX), where πI\pi_{I} denotes the projection. ∎

We put together several known results to obtain the following theorem (which we have not found anywhere in the literature):

Theorem 2.3.

Let 𝒞\mathcal{C} be a descriptive class of sets. Then any separable metrizable 𝒞\mathcal{C} space is an absolute 𝒞\mathcal{C} space (so the first four conditions of Remark 1.5 are equivalent).

If 𝒞\mathcal{C} is one of the Borel or \mathcal{F}-Borel classes, this is further equivalent to XX being in 𝒞(Y)\mathcal{C}(Y) for every Tychonoff space in which YY is embedded.

Proof.

The statement for the classes 𝒜α\mathcal{A}_{\alpha} and α\mathcal{M}_{\alpha} follows from the general result of [HS] (see Proposition 1.8 above). Let us continue by proving the result for the \mathcal{F}-Borel classes. For the first two levels, it follows from Remark 1.6. Suppose that Xα(β(X))X\in\mathcal{F}_{\alpha}(\beta(X)) for some α3\alpha\geq 3 and let cXcX be any compactification of XX.

By Proposition 2.2 we can find a smaller metrizable compactification dXdX. Let 𝒟\mathcal{D} be the Borel class corresponding to α\mathcal{F}_{\alpha} (see Remark 1.3). Then X𝒟(βX)X\in\mathcal{D}(\beta X). Since 𝒟\mathcal{D} is absolute by Proposition 1.8, we get X𝒟(dX)X\in\mathcal{D}(dX). Since dXdX is metrizable, we deduce that Xα(dX)X\in\mathcal{F}_{\alpha}(dX), so Xα(cX)X\in\mathcal{F}_{\alpha}(cX) holds by Remark 2.1.

The proof for 𝒢\mathcal{G}-Borel classes is analogous.

3 Hereditarity of absolute complexity

In this section, we show that absolute complexity is hereditary with respect to closed subspaces. To do this, we first need to be able to extend compactifications of subspaces to compactifications of the original space. We start with a topological lemma:

Lemma 3.1.

Suppose that KK is compact, FKF\subset K is closed and f:FLf:F\rightarrow L is a continuous surjective mapping. Define a mapping q:KL(KF)q:K\to L\cup(K\setminus F) as q(x)=xq(x)=x for xKFx\in K\setminus F and q(x)=f(x)q(x)=f(x) for xFx\in F. Then (KF)L\left(K\setminus F\right)\cup L equipped with the quotient topology (induced by qq) is a compact space.

Proof.

Since ff is surjective, the space (KF)L(K\setminus F)\cup L coincides with the adjunction space KfLK\cup_{f}L determined by KK, LL and ff (for definition, see [Eng, below Ex. 2.4.12]). Since KfLK\cup_{f}L is a quotient of a compact space KLK\oplus L with respect to a closed equivalence relation ([Eng, below Ex. 2.4.12]), we get by Alexandroff theorem ([Eng, Thm 3.2.11]) that the space KfLK\cup_{f}L is compact (in particular, it is Hausdorff). ∎

This gives us a way to extend a compactification, provided that we already know how to extend some bigger compactification:

Proposition 3.2.

Let YY be a topological space and suppose that for a closed XYX\subset Y we have cXX¯dYcX\preceq\overline{X}^{dY} for some compactifications cXcX and dYdY. Then cXcX is equivalent to X¯cX\overline{X}^{cX} for some compactification cYcY of YY.

Proof.

Let XX, YY, cXcX and dYdY be as above and denote by f:X¯dYcXf:\overline{X}^{dY}\rightarrow cX the mapping which witnesses that cXcX is smaller than X¯dY\overline{X}^{dY}. We will assume that XcXX\subset cX and YdYY\subset dY, which means that f|X=idXf|_{X}=\textrm{id}_{X}.

By Lemma 3.1 (with K:=dYK:=dY, F:=X¯dYF:=\overline{X}^{dY} and L:=cXL:=cX), the space cYcY, defined by the formula

cY:=(dYX¯dY)cX=(KF)L,cY:=\left(dY\setminus\overline{X}^{dY}\right)\cup cX=\left(K\setminus F\right)\cup L,

is compact. To show that cYcY is a compactification of YY, we need to prove that the mapping q:dYcYq:dY\rightarrow cY, defined as ff extended to dYX¯dYdY\setminus\overline{X}^{dY} by identity, is a homeomorphism when restricted to YY. The continuity of q|Yq|_{Y} follows from the continuity of qq. The restriction is injective, because ψ\psi is injective on (dYX¯dY)X\left(dY\setminus\overline{X}^{dY}\right)\cup X (and XX is closed in YY). The definition of qq and the fact that XX is closed in YY imply that the inverse mapping is continuous on YXY\setminus X. To get the continuity of (q|Y)1(q|_{Y})^{-1} at the points of XX, let xXx\in X and UU be a neighborhood of xx in KK. We want to prove that q(UY)q(U\cap Y) is a neighborhood of q(x)q(x) in q(Y)q(Y). To see this, observe that C:=dYUC:=dY\setminus U is compact, therefore q(C)q(C) is closed, its complement V:=cYq(C)V:=cY\setminus q(C) is open and it satisfies Vq(Y)=q(UY)V\cap q(Y)=q(U\cap Y).

Lastly, we note that for ycYq(X)¯cYy\in cY\setminus\overline{q(X)}^{cY}, q1(y)q^{-1}(y) is in dYX¯dYdY\setminus\overline{X}^{dY}. This gives the second identity on the following line, and completes the proof:

X¯cY=q(X)¯cY=q(X¯dY)=f(X¯dY)=cX.\overline{X}^{cY}=\overline{q(X)}^{cY}=q\left(\overline{X}^{dY}\right)=f\left(\overline{X}^{dY}\right)=cX.

Recall that a subspace XX of YY is said to be CC^{\star}-embedded in YY if any bounded continuous real function on XX can be continuously extended to YY.

Corollary 3.3.

If a closed set XX is CC^{\star}-embedded in YY, then each compactification cXcX is of the form cX=X¯cYcX=\overline{X}^{cY} for some compactification cYcY.

Proof.

Recall that the Čech-Stone compactification βX\beta X is characterized by the property that each bounded continuous function from XX has a unique continuous extension to βX\beta X. Using this characterization, it is a standard exercise to show that XX is CC^{\star}-embedded in YY if and only if X¯βY=βX\overline{X}^{\beta Y}=\beta X. It follows that for XX and YY as above, any compactification cXcX is smaller than X¯βY\overline{X}^{\beta Y}, and the result follows from Proposition 3.2. ∎

We use these results to get the following corollary:

Proposition 3.4.

Let 𝒞\mathcal{C} be one of the classes α\mathcal{F}_{\alpha}, α\mathcal{M}_{\alpha} or 𝒜α\mathcal{A}_{\alpha} for some α<ω1\alpha<\omega_{1}.

  1. (i)

    Any closed subspace of a 𝒞\mathcal{C} space is a 𝒞\mathcal{C} space;

  2. (ii)

    Any closed subspace of an absolute 𝒞\mathcal{C} space is an absolute 𝒞\mathcal{C} space.

Proof.

(i)(i): This is trivial, since if XX is a closed subspace of YY and YY is of the class 𝒞\mathcal{C} in some compactification cYcY, then XX is of the same class in the compactification X¯cY\overline{X}^{cY}.

(ii)(ii): For Borel classes, which are absolute, the result follows from the first part – therefore, we only need to prove the statement for the \mathcal{F}-Borel classes. Let α<ω1\alpha<\omega_{1} and assume that XX is a closed subspace of some α\mathcal{F}_{\alpha} space YY.

Firstly, if YY is an α\mathcal{F}_{\alpha} subset of βY\beta Y, it is 𝒦\mathcal{K}-analytic and, in particular, Lindelöf. Since we also assume that YY is Hausdorff, we get that YY is normal [KKLP, Propositions 3.4 and 3.3]. By Tietze’s theorem, XX is CC^{\star}-embedded in YY, which means that every compactification of XX can be extended to a compactification of YY (Corollary 3.3). We conclude the proof as in (i)(i). ∎

Note that the second part of Proposition 3.4 does not hold, in a very strong sense, if we replace the closed subspaces by FσδF_{\sigma\delta} subspaces. Indeed, the space TT from Theorem 1.7 is an FσδF_{\sigma\delta} subspace of a compact space, but it is not even absolutely \mathcal{F}-Borel (that is, \mathcal{F}-Borel in every compactification). Whether anything positive can be said about FσF_{\sigma} subspaces of absolute α\mathcal{F}_{\alpha} spaces is unknown.

4 Special sets of sequences

In this chapter, we define ‘broom sets’ – a special type sets of finite sequences, which have a special ‘tree’ structure and all of their infinite extensions are closed and discrete. As in [Tal], these sets will then be used to construct spaces with special complexity properties.

4.1 Trees and a rank on them

First, we introduce the (mostly standard) notation which will be used in the sequel.

Notation 4.1 (Integer sequences of finite and infinite length).

We denote

  • ωω:=\omega^{\omega}:= infinite sequences of non-negative integers :={σ:ωω}:=\left\{\sigma:\omega\rightarrow\omega\right\},

  • ω<ω:=\omega^{<\omega}:= finite sequences of non-neg. integers :={s:nω|nω}:=\left\{s:n\rightarrow\omega|\ n\in\omega\right\}.

Suppose that sω<ωs\in\omega^{<\omega} and σωω\sigma\in\omega^{\omega}. We can represent σ\sigma as (σ(0),σ(1),)(\sigma(0),\sigma(1),\dots) and ss as (s(0),s(1),,s(n1))(s(0),s(1),\dots,s(n-1)) for some nωn\in\omega. We denote the length of ss as |s|=dom(s)=n|s|=\textrm{dom}(s)=n, and set |σ|=ω|\sigma|=\omega. If for some tω<ωωωt\in\omega^{<\omega}\cup\omega^{\omega} we have |t||s||t|\geq|s| and t||s|=st|_{|s|}=s, we say that uu extends ss, denoted as ttt\sqsubset t. We say that u,vω<ωu,v\in\omega^{<\omega} are non-comparable, denoting as uvu\perp v, when neither uvu\sqsubset v nor uvu\sqsupset v holds.

Unless we say otherwise, ωω\omega^{\omega} will be endowed with the standard product topology, whose basis consists of sets 𝒩(s):={σs|σωω}\mathcal{N}(s):=\left\{\sigma\sqsupset s|\ \sigma\in\omega^{\omega}\right\}, sω<ωs\in\omega^{<\omega}.

For nωn\in\omega, we denote by (n)(n) the corresponding sequence of length 11. By s^ts\hat{\ }t we denote the concatenation of a sequences sω<ωs\in\omega^{<\omega} and tω<ωωωt\in\omega^{<\omega}\cup\omega^{\omega}. We will also use this notation to extend finite sequences by integers and sets, using the convention s^k:=s^(k)s\hat{\ }k:=s\hat{\ }(k) for kωk\in\omega and s^T:={s^t|tT}s\hat{\ }T:=\left\{s\hat{\ }t\ |\ t\in T\right\} for Tω<ωωωT\subset\omega^{<\omega}\cup\omega^{\omega}.

Definition 4.2 (Trees on ω\omega).

A tree (on ω\omega) is a set Tω<ωT\subset\omega^{<\omega} which satisfies

(s,tω<ω):st&tTsT.(\forall s,t\in\omega^{<\omega}):s\sqsubset t\ \&\ t\in T\implies s\in T.

By Tr we denote the space of all trees on ω\omega. For Sω<ωS\subset\omega^{<\omega} we denote by clTr(S):={uω<ω|sS:su}\mathrm{cl}_{\mathrm{Tr}}(S):=\left\{u\in\omega^{<\omega}|\ \exists s\in S:s\sqsupset u\right\} the smallest tree containing SS. Recall that the empty sequence \emptyset can be thought of as the ‘root’ of each tree, since it is contained in any nonempty tree.

If each of the initial segments σ|n\sigma|n, nωn\in\omega, of some σωω\sigma\in\omega^{\omega} belongs to TT, we say that σ\sigma is an infinite branch of TT. By WF we denote the space of all trees which have no infinite branches (the ‘well founded’ trees).

We define a variant of the standard derivative on trees and the corresponding Π11\Pi^{1}_{1}-rank. We do not actually use the fact that the rank introduced in Definition 4.3 is a Π11\Pi^{1}_{1} rank, but more details about such ranks and derivatives can be found in [Kech, ch. 34 D,E].

Definition 4.3.

Let TTrT\in\textrm{Tr}. We define its ‘infinite branching’ derivative Di(T)D_{i}(T) as

Di(T):={tT|ts holds for infinitely many sT}.D_{i}(T):=\left\{t\in T|\ t\sqsubset s\textrm{ holds for infinitely many }s\in T\right\}.

This operation can be iterated in the standard way:

Di0(T)\displaystyle D^{0}_{i}(T) :=T,\displaystyle:=T,
Diα+1(T)\displaystyle D^{\alpha+1}_{i}\left(T\right) :=Di(Diα(T)) for successor ordinals,\displaystyle:=D_{i}\left(D_{i}^{\alpha}\left(T\right)\right)\textrm{ for successor ordinals},
Diλ(T)\displaystyle D_{i}^{\lambda}\left(T\right) :=α<λDiα(T) for limit ordinals.\displaystyle:=\underset{\alpha<\lambda}{\bigcap}D_{i}^{\alpha}\left(T\right)\textrm{ for limit ordinals}.

This allows us to define a rank rir_{i} on Tr as ri():=1r_{i}(\emptyset):=-1, ri(T):=0r_{i}(T):=0 for finite trees and

ri(T):=min{α<ω1|Diα+1(T)=}.r_{i}(T):=\min\{\alpha<\omega_{1}|\ D_{i}^{\alpha+1}\left(T\right)=\emptyset\}.

If no such α<ω1\alpha<\omega_{1} exists, we set ri(T):=ω1r_{i}(T):=\omega_{1}.

Note that on well founded trees, the derivative introduced in Definition 4.3 behaves the same way as the derivative from [Kech, Exercise 21.24], but it leaves any infinite branches untouched. It follows from the definition that the rank rir_{i} of a tree TT is countable if and only if TT is well founded. We consider this approach better suited for our setting.

While this definition of derivative and rank is the most natural for trees, we can extend it to all subsets SS of ω<ω\omega^{<\omega} by setting Di(S):=Di(clTr(S))D_{i}(S):=D_{i}(\mathrm{cl}_{\mathrm{Tr}}(S)) and defining the rest of the notions using this extended ‘derivative’. Clearly, if STrS\in Tr, this coincides with the original definition.

Lemma 4.4.

No set Sω<ωS\subset\omega^{<\omega} which satisfies ri(S)<ω1r_{i}(S)<\omega_{1} can be covered by finitely many sets of rank strictly smaller than ri(S)r_{i}(S).

Proof.

Note that if T1T_{1} and T2T_{2} are trees, T1T2T_{1}\cup T_{2} is a tree as well, and we have Diα(T1T2)=Diα(T1)Diα(T2)D_{i}^{\alpha}(T_{1}\cup T_{2})=D_{i}^{\alpha}(T_{1})\cup D_{i}^{\alpha}(T_{2}) for any α<ω1\alpha<\omega_{1}. Moreover, for any S1,S2ω<ωS_{1},S_{2}\subset\omega^{<\omega} we have clTr(S1S2)=clTr(S1)clTr(S2)\mathrm{cl}_{\mathrm{Tr}}(S_{1}\cup S_{2})=\mathrm{cl}_{\mathrm{Tr}}(S_{1})\cup\mathrm{cl}_{\mathrm{Tr}}(S_{2}). This yields the formula

ri(S1Sk)=max{ri(S1),,ri(Sk)}r_{i}(S_{1}\cup\dots\cup S_{k})=\max\{r_{i}(S_{1}),\dots,r_{i}(S_{k})\}

for any Siω<ωS_{i}\subset\omega^{<\omega}, which gives the result. ∎

4.2 Sets which extend into closed discrete subsets of ωω\omega^{\omega}

Definition 4.5.

A set AωωA\subset\omega^{\omega} is said to be an infinite extension of Bω<ωB\subset\omega^{<\omega}, if there exists a bijection φ:BA\varphi:B\rightarrow A, such that (sB):φ(s)s(\forall s\in B):\varphi(s)\sqsupset s. If A~\tilde{A} has the same property as AA, except that we have A~ω<ω\tilde{A}\subset\omega^{<\omega}, it is said to be a (finite) extension of BB.

By 𝒟𝒫(ω<ω)\mathcal{D}\subset\mathcal{P}\left(\omega^{<\omega}\right) we denote the system of those sets Dω<ωD\subset\omega^{<\omega} which satisfy

  • (i)

    (s,tD)(\forall s,t\in D) : ststs\neq t\implies s\perp t;

  • (ii)

    every infinite extension of DD is closed and discrete in ωω\omega^{\omega}.

Example 4.6 (Broom sets of the first class).

Let hω<ωh\in\omega^{<\omega} and snω<ωs_{n}\in\omega^{<\omega} for nωn\in\omega be finite sequences and suppose that (fn)nω(f_{n})_{n\in\omega} is an injective sequence of elements of ω\omega. Using these sequences, we define a ‘broom’ set BB as

B:={h^fn^sn|nω}.B:=\left\{h\hat{\ }f_{n}\hat{\ }s_{n}|\ n\in\omega\right\}.

The intuition is that hh is the ‘handle’ of the broom, sns_{n} are the ‘bristles’, and the sequence (fn)n(f_{n})_{n} causes these bristles to ‘fork out’ from the end of the handle. Note that the injectivity of (fn)n(f_{n})_{n} guarantees that B𝒟B\in\mathcal{D} (see Proposition 4.7).

Proposition 4.7 (Properties of 𝒟\mathcal{D}).

The system 𝒟\mathcal{D} has the following properties (where DD is an arbitrary element of 𝒟\mathcal{D}):

  1. (i)

    {(n)|nω}𝒟\{(n)|\ n\in\omega\}\in\mathcal{D};

  2. (ii)

    (Eω<ω):EDE𝒟(\forall E\subset\omega^{<\omega}):E\subset D\implies E\in\mathcal{D};

  3. (iii)

    (Eω<ω):E(\forall E\subset\omega^{<\omega}):E is a finite extension of DE𝒟D\implies E\in\mathcal{D};

  4. (iv)

    (hω<ω):h^D𝒟\left(\forall h\in\omega^{<\omega}\right):h\hat{\ }D\in\mathcal{D};

  5. (v)

    For each sequence (Dn)n\left(D_{n}\right)_{n} of elements of 𝒟\mathcal{D} and each one-to-one enumeration {dn|nω}\left\{d_{n}|\ n\in\omega\right\} of DD, we have nωdn^Dn𝒟\bigcup_{n\in\omega}d_{n}\hat{\ }D_{n}\in\mathcal{D}.

Proof.

The first four properties are obvious. To prove 5.5., let DD, DnD_{n}, E:=nωdn^DnE:=\bigcup_{n\in\omega}d_{n}\hat{\ }D_{n} be as above. The show that EE satisfies (i) from Definition 4.5, suppose that e=dn^e,f=dm^fDe=d_{n}\hat{\ }e^{\prime},\ f=d_{m}\hat{\ }f^{\prime}\in D are comparable. Then dnd_{n} and dmd_{m} are also comparable, which means that m=nm=n and e,fDn=Dme^{\prime},f^{\prime}\in D_{n}=D_{m} are comparable. It follows that e=fe^{\prime}=f^{\prime} and hence e=fe=f.

For Aω<ωA\subset\omega^{<\omega}, the condition (ii) from Definition 4.5 is equivalent to

(σωω)(mω):{sA|s is comparable with σ|m} is finite.(\forall\sigma\in\omega^{\omega})(\exists m\in\omega):\ \left\{s\in A|\ s\textrm{ is comparable with }\sigma|m\right\}\textrm{ is finite}. (4.1)

Indeed, suppose (4.1) holds and let HH be an infinite extension of AA. For σωω\sigma\in\omega^{\omega} let mωm\in\omega be as in (4.1). Any μH𝒩(σ|m)\mu\in H\cap\mathcal{N}(\sigma|m) is an extension of some sAs\in A, which must be comparable with σ|m\sigma|m. It follows that the neighborhood 𝒩(σ|m)\mathcal{N}(\sigma|m) of σ\sigma only contains finitely many elements of HH, which means that σ\sigma is not a cluster point of HH and HH is discrete. In the opposite direction, suppose that the condition does not hold. Then we can find a one-to-one sequence (sm)mω(s_{m})_{m\in\omega} of elements of AA, such that smσ|ms_{m}\sqsupset\sigma|m for each mωm\in\omega. Then σ\sigma is a cluster point of (any) infinite extension of {sm|mω}\{s_{m}|\ m\in\omega\}.

If AA also satisfies (i) from Definition 4.5, (4.1) is clearly equivalent to

(σωω)(mω):σ|m is comparable with at most one sA.(\forall\sigma\in\omega^{\omega})(\exists m\in\omega):\ \sigma|m\textrm{ is comparable with at most one }s\in A. (4.2)

To see that EE satisfies (4.2), let σωω\sigma\in\omega^{\omega}. Since D𝒟D\in\mathcal{D}, let mm be the integer obtained by applying (4.2) to this set. If no dDd\in D is comparable with σ|m\sigma|m, then neither is any eEe\in E and we are done.

Otherwise, let dn0d_{n_{0}} be the only element of DD comparable with σ|m\sigma|m and let m0m_{0} be the integer obtained by applying (4.2) to dn0^Dn0d_{n_{0}}\hat{\ }D_{n_{0}}. Then m:=max{m0,m}m^{\prime}:=\max\{m_{0},m\} witnesses that EE satisfies (4.2): Indeed, any e=dn^ee=d_{n}\hat{\ }e^{\prime} comparable with σ|m\sigma|{m^{\prime}} must have dnd_{n} comparable with σ|m\sigma|{m^{\prime}}. In particular, dnd_{n} is comparable with σ|m\sigma|m. It follows that dn=dn0d_{n}=d_{n_{0}} and edn0^Dn0e\in d_{n_{0}}\hat{\ }D_{n_{0}}. Therefore, ee must be comparable with σ|m0\sigma|{m_{0}} and since there is only one such element, the proof is finished. ∎

4.3 Broom sets

We are now ready to define the broom sets and give some of their properties.

Definition 4.8 (Broom sets of higher classes).

We define the hierarchy of broom sets α\mathcal{B}_{\alpha}, α<ω1\alpha<\omega_{1}. We set 0:={{s}|sω<ω}\mathcal{B}_{0}:=\left\{\left\{s\right\}|\ s\in\omega^{<\omega}\right\} and, for 1<α<ω11<\alpha<\omega_{1} we define α\mathcal{B}_{\alpha} as the union of β<αβ\bigcup_{\beta<\alpha}\mathcal{B}_{\beta} and the collection of all sets Bω<ωB\subset\omega^{<\omega} of the form

B=nωh^fn^BnB=\underset{n\in\omega}{\bigcup}h\hat{\ }f_{n}\hat{\ }B_{n}

for some finite sequence hω<ωh\in\omega^{<\omega} (‘handle’), broom sets Bnβ<αβB_{n}\in\bigcup_{\beta<\alpha}\mathcal{B}_{\beta} (‘bristles’) and one-to-one sequence (fn)nωωω(f_{n})_{n\in\omega}\in\omega^{\omega} (‘forking sequence’). We also denote as ω1:=α<ω1α\mathcal{B}_{\omega_{1}}:=\underset{\alpha<\omega_{1}}{\bigcup}\mathcal{B}_{\alpha} the collection of all these broom classes.

In Section 5, we will need a result of the type ‘if BB is an element of α+1α\mathcal{B}_{\alpha+1}\setminus\mathcal{B}_{\alpha}, then BB cannot be covered by finitely many brooms of class α\alpha’. While this is, in general, not true, we can restrict ourselves to the following subcollection ~α+1\widetilde{\mathcal{B}}_{\alpha+1} of α+1α\mathcal{B}_{\alpha+1}\setminus\mathcal{B}_{\alpha} of ‘(α+1)(\alpha+1)-brooms whose bristles are of the highest allowed class’, where the result holds (as we will see later):

Definition 4.9.

We set ~0:=0\widetilde{\mathcal{B}}_{0}:=\mathcal{B}_{0}. For successor ordinals we define ~α+1\widetilde{\mathcal{B}}_{\alpha+1} as the collection of all sets Bω<ωB\subset\omega^{<\omega} of the form

B=nωh^fn^Bn,B=\underset{n\in\omega}{\bigcup}h\hat{\ }f_{n}\hat{\ }B_{n},

where hω<ωh\in\omega^{<\omega}, Bn~αB_{n}\in\widetilde{\mathcal{B}}_{\alpha} and (fn)n(f_{n})_{n} is the (1-to-1) forking sequence. For a limit ordinal α<ω1\alpha<\omega_{1}, the family ~α\widetilde{\mathcal{B}}_{\alpha} is defined in the same way, except that the bristles BnB_{n} belong to ~βn\widetilde{\mathcal{B}}_{\beta_{n}} for some βn<α\beta_{n}<\alpha, where supnβn=α\sup_{n}\beta_{n}=\alpha.

Remark 4.10 (Basic properties of broom sets).

It is clear that for every broom BαB\in\mathcal{B}_{\alpha}, any finite extension of BB is also a broom of class α\alpha, and so is the set h^Bh\hat{\ }B for any hω<ωh\in\omega^{<\omega}. Moreover, it follows from Proposition 4.7 that ω1𝒟\mathcal{B}_{\omega_{1}}\subset\mathcal{D}. These properties also hold for collections ~α\widetilde{\mathcal{B}}_{\alpha}, α<ω1\alpha<\omega_{1}.

Actually, broom sets from Definition 4.8 are exactly those elements of 𝒟\mathcal{D} for which every ‘branching point’ is a ‘point of infinite branching’. However, we do not need this description of broom sets, so we will not give its proof here.

Broom sets are connected with rank rir_{i} in the following way:

Proposition 4.11.

For Bω1B\in\mathcal{B}_{\omega_{1}} and α<ω1\alpha<\omega_{1}, the following holds:

  1. (i)

    Bαri(B)αB\in\mathcal{B}_{\alpha}\implies r_{i}(B)\leq\alpha,

  2. (ii)

    B~αri(B)=αB\in\widetilde{\mathcal{B}}_{\alpha}\implies r_{i}(B)=\alpha.

Proof.

Firstly, we make the following two observations:

(Bω<ω):ri(B)=1B=,(Bω1):ri(B)=0B0=~0.\begin{split}(\forall B\subset\omega^{<\omega}):r_{i}(B)=-1&\iff B=\emptyset,\\ (\forall B\in\mathcal{B}_{\omega_{1}}):r_{i}(B)=0&\iff B\in\mathcal{B}_{0}=\widetilde{\mathcal{B}}_{0}.\end{split}

We use this as a starting point for transfinite induction. We now prove (i)(i) and (ii)(ii) separately, distinguishing also between the cases of successor α\alpha and limit α\alpha.

(i)(i): Suppose that the statement holds for α<ω1\alpha<\omega_{1} and let Bα+1B\in{\mathcal{B}}_{\alpha+1}. By definition, B=nh^fn^BnB=\bigcup_{n}h\hat{\ }f_{n}\hat{\ }B_{n} holds for some handle hω<ωh\in\omega^{<\omega}, forking sequence (fn)n(f_{n})_{n} and BnαB_{n}\in\mathcal{B}_{\alpha}. For each nωn\in\omega we have ri(Bn)αr_{i}(B_{n})\leq\alpha, which gives Diα+1(Bn)=D_{i}^{\alpha+1}(B_{n})=\emptyset and hence Diα+1(fn^Bn)=D_{i}^{\alpha+1}(f_{n}\hat{\ }B_{n})=\emptyset. It follows that

Diα+1(nh^fn^Bn)clTr(h),D_{i}^{\alpha+1}(\bigcup_{n}h\hat{\ }f_{n}\hat{\ }B_{n})\subset\mathrm{cl}_{\mathrm{Tr}}(h),

and

Diα+2(nh^fn^Bn)=,D_{i}^{\alpha+2}(\bigcup_{n}h\hat{\ }f_{n}\hat{\ }B_{n})=\emptyset,

which, by definition or rir_{i}, means that ri(B)α+1r_{i}(B)\leq\alpha+1.

For the case of limit ordinal α<ω1\alpha<\omega_{1}, assume that the statement holds for every β<α\beta<\alpha, and let BαB\in{\mathcal{B}}_{\alpha}. As above, B=nh^fn^BnB=\bigcup_{n}h\hat{\ }f_{n}\hat{\ }B_{n} holds for some hh, (fn)n(f_{n})_{n} and BnB_{n}, where BnβnB_{n}\in\mathcal{B}_{\beta_{n}}, βn+1<α\beta_{n}+1<\alpha. We successively deduce that ri(Bn)r_{i}(B_{n}) is at most βn\beta_{n}, hence Diα(Bn)Diβn+1(Bn)D_{i}^{\alpha}(B_{n})\subset D_{i}^{\beta_{n}+1}(B_{n}) holds and we have Diα(fn^Bn)=D_{i}^{\alpha}(f_{n}\hat{\ }B_{n})=\emptyset. Since (fn)n(f_{n})_{n} is a forking sequence, then for any sfn0^Bn0s\in f_{n_{0}}\hat{\ }B_{n_{0}} which is different from the empty sequence, we get the following equivalence:

h^sDiα(nh^fn^Bn)(n0ω):sDiα(fn0^Bn0).h\hat{\ }s\in D_{i}^{\alpha}(\bigcup_{n}h\hat{\ }f_{n}\hat{\ }B_{n})\iff\left(\exists n_{0}\in\omega\right):s\in D_{i}^{\alpha}(f_{n_{0}}\hat{\ }B_{n_{0}}). (4.3)

From (4.3) we conclude that Diα(B)clTr({h})D_{i}^{\alpha}(B)\subset\mathrm{cl}_{\mathrm{Tr}}(\{h\}) is finite, which means that ri(B)αr_{i}(B)\leq\alpha.

(ii)(ii): Suppose that the second part of the proposition holds for some α<ω\alpha<\omega and let B=nh^fn^Bn~α+1B=\bigcup_{n}h\hat{\ }f_{n}\hat{\ }B_{n}\in\widetilde{\mathcal{B}}_{\alpha+1}. For each nωn\in\omega, BnB_{n} is an element of ~α\widetilde{\mathcal{B}}_{\alpha}, which by the induction hypothesis means that ri(Bn)=αr_{i}(B_{n})=\alpha, and thus there exists some sequence snDiα(B)s_{n}\in D_{i}^{\alpha}(B). In particular, we have h^fn^snDiα(h^fn^Bn)h\hat{\ }f_{n}\hat{\ }s_{n}\in D_{i}^{\alpha}(h\hat{\ }f_{n}\hat{\ }B_{n}). We conclude that hDiα+1(B)h\in D_{i}^{\alpha+1}(B) and ri(B)α+1r_{i}(B)\geq\alpha+1.

Assume now that α<ω1\alpha<\omega_{1} is a limit ordinal and that (ii)(ii) holds for all β<α\beta<\alpha. Let B=nh^fn^Bn~αB=\bigcup_{n}h\hat{\ }f_{n}\hat{\ }B_{n}\in\widetilde{\mathcal{B}}_{\alpha}, where Bn~βnB_{n}\in\widetilde{\mathcal{B}}_{\beta_{n}}, supnβn=α\sup_{n}\beta_{n}=\alpha. For each β<α\beta<\alpha, there is some nωn\in\omega with βnβ\beta_{n}\geq\beta. Since Diβn(Bn)D_{i}^{\beta_{n}}(B_{n})\neq\emptyset, we have

Diβ(B)Diβ(h^fn^Bn)Diβn(h^fn^Bn)h^fn^Diβn(Bn)h^fn^shD_{i}^{\beta}(B)\supset D_{i}^{\beta}(h\hat{\ }f_{n}\hat{\ }B_{n})\supset D_{i}^{\beta_{n}}(h\hat{\ }f_{n}\hat{\ }B_{n})\supset h\hat{\ }f_{n}\hat{\ }D_{i}^{\beta_{n}}(B_{n})\ni h\hat{\ }f_{n}\hat{\ }s\sqsupset h

for some sω<ωs\in\omega^{<\omega}. This implies that hDiβ(B)h\in D_{i}^{\beta}(B) holds for each β<α\beta<\alpha, which gives hβ<αDiβ(B)=Diα(B)h\in\underset{\beta<\alpha}{\bigcap}D_{i}^{\beta}(B)=D_{i}^{\alpha}(B) and ri(B)αr_{i}(B)\geq\alpha. ∎

From Proposition 4.11, it follows that no ~\widetilde{\mathcal{B}} set can be covered by finitely many sets from β<αβ\bigcup_{\beta<\alpha}\mathcal{B}_{\beta}. We will prove a slightly stronger result later (Lemma 5.12).

5 Broom spaces and their α\mathcal{F}_{\alpha} parametrization

We will study a certain collection of one-point Lindelöfications of the discrete uncountable space which give the space a special structure. These spaces were used by (among others) Talagrand, who constructed a space which is non-absolutely FσδF_{\sigma\delta} ([Tal]). Our goal is to compute the absolute complexity of spaces of this type.

For an arbitrary system 𝒫(ωω)\mathcal{E}\subset\mathcal{P}\left(\omega^{\omega}\right) which contains only countable sets, we define the space XX_{\mathcal{E}} as the set ωω{}\omega^{\omega}\cup\{\infty\} with the following topology: each σωω\sigma\in\omega^{\omega} is an isolated point and a neighborhood subbasis of \infty consists of all sets of the form {}(ωωE)\{\infty\}\cup(\omega^{\omega}\setminus E) for some EE\in\mathcal{E}.

The outline of Section 5 is the following: We note in Example 5.4 that if X=:XcXX_{\mathcal{E}}=:X\subset cX, then there is a natural formula for an FσδF_{\sigma\delta} set Y1cXY_{1}\subset cX containing XX. Then in Definition 5.7, we generalize the formula and define sets YαY_{\alpha}. These are useful from two reasons: the first is that their definition is ‘absolute’ (it does not depend on the choice of cXcX). The other is that it is easy to compute (an upper bound on) their complexity, which we do in Lemma 5.9.

Our goal is to show that X=YαX_{\mathcal{E}}=Y_{\alpha} holds for a suitable α\alpha. In Section 5.2, we prepare the tools for establishing a connection between the complexity of \mathcal{E} and the value of ‘the α\alpha that works’. In Proposition 5.13, we use these results to prove that X=YαX_{\mathcal{E}}=Y_{\alpha} holds in every compactification. Applying Lemma 5.9, we obtain the absolute complexity of XX. We finish by giving two corollaries of this proposition.

Rather than working with a general 𝒫(ωω)\mathcal{E}\subset\mathcal{P}(\omega^{\omega}), we will be interested in the following systems:

Definition 5.1 (Infinite broom sets111In [Tal] these systems are denoted as 𝒜α\mathcal{A}_{\alpha}, but we have to use a different letter to distinguish between infinite brooms and additive Borel classes. The letter \mathcal{E} is supposed to stand for “extension” (of \mathcal{B}). Note also that Talagrand’s collections 𝒜n\mathcal{A}_{n}, nωn\in\omega, correspond to our n+1\mathcal{E}_{n+1} (for αω\alpha\geq\omega the enumeration is the same).).

For αω1\alpha\leq\omega_{1} we define

α:={Eωω|E is an infinite extension of some Bα}.\mathcal{E}_{\alpha}:=\left\{E\subset\omega^{\omega}|\ E\textrm{ is an infinite extension of some }B\in\mathcal{B}_{\alpha}\right\}.

Since ω1𝒟\mathcal{B}_{\omega_{1}}\subset\mathcal{D}, every Eω1E\in\mathcal{E}_{\omega_{1}} is closed and discrete in ωω\omega^{\omega}, and the following result holds:

Proposition 5.2 ([Tal]).

If each EE\in\mathcal{E} is closed discrete in ωω\omega^{\omega}, then XFσδ(βX)X_{\mathcal{E}}\in F_{\sigma\delta}\left(\beta X_{\mathcal{E}}\right).

In the remainder of Section 5, XX will stand for the space XX_{\mathcal{E}} for some family ω1\mathcal{E}\subset\mathcal{E}_{\omega_{1}} and cXcX will be a fixed compactification of XX. All closures will automatically be taken in cXcX. The choice of \mathcal{E} is important, because it gives the space XX the following property:

Lemma 5.3.

For every xcXXx\in cX\setminus X there exists HXH\subset X satisfying

  1. (i)

    HH can be covered by finite union of elements of \mathcal{E};

  2. (ii)

    (FX):xF¯xFH¯\left(\forall F\subset X\right):x\in\overline{F}\implies x\in\overline{F\cap H}.

Proof.

Since xx\neq\infty, there is some U𝒰(x)U\in\mathcal{U}(x) with U¯\infty\notin\overline{U}. Because UU is a neighborhood of xx, we get

(FX):xF¯xFU¯.\left(\forall F\subset X\right):x\in\overline{F}\implies x\in\overline{F\cap U}.

Therefore, we define HH as H:=UXH:=U\cap X. Clearly, HH satisfies the second condition. Moreover, (i)(i) is also satisfied, because by definition of XX_{\mathcal{E}}, sets of the form {}(ωωE)\{\infty\}\cup(\omega^{\omega}\setminus E), EE\in\mathcal{E} form a neighborhood subbasis of \infty. ∎

5.1 \mathcal{F}-Borel sets YαY_{\alpha} containing XX_{\mathcal{E}}

First, we start with motivation for our definitions:

Example 5.4 (FσδF_{\sigma\delta} and FσδσδF_{\sigma\delta\sigma\delta} sets containing XX_{\mathcal{E}}).

When trying to compute the complexity of XX in cXcX, the following canonical KσδK_{\sigma\delta} candidate comes to mind333Recall that the notions related to sequences are defined in Notation 4.1. In particular, 𝒩(s)\mathcal{N}(s) denotes the standard Baire interval.

X{}nωsωn𝒩(s)¯=:Y1.X\subset\{\infty\}\cup\underset{n\in\omega}{\bigcap}\underset{s\in\omega^{n}}{\bigcup}\overline{\mathcal{N}(s)}=:Y_{1}.

The two sets might not necessarily be identical, but the inclusion above always holds. At the cost of increasing the complexity of the right hand side to FσδσδF_{\sigma\delta\sigma\delta}, we can define a smaller set, which will still contain XX:

X{}nωsωnkωtωk𝒩(s^t)¯=:Y2.X\subset\{\infty\}\cup\underset{n\in\omega}{\bigcap}\ \underset{s\in\omega^{n}}{\bigcup}\ \underset{k\in\omega}{\bigcap}\ \underset{t\in\omega^{k}}{\bigcup}\ \overline{\mathcal{N}(s\hat{\ }t)}=:Y_{2}.

In the same manner as in Example 5.4, we could define sets Y1Y2Y3XY_{1}\supset Y_{2}\supset Y_{3}\supset\dots\supset X. Unfortunately, the notation from Example 5.4 is impractical if we continue any further, and moreover, it is unclear how to extend the definition to infinite ordinals. We solve this problem by introducing ‘admissible mappings’ and a more general definition below.

Definition 5.5 (Admissible mappings).

Let TTrT\in\textrm{Tr} be a tree and φ:Tω<ω\varphi:T\rightarrow\omega^{<\omega} a mapping from TT to the space of finite sequences on ω\omega. We will say that φ\varphi is admissible, if it satisfies

  1. (i)

    (s,tT):stφ(s)φ(t)(\forall s,t\in T):s\sqsubset t\implies\varphi(s)\sqsubset\varphi(t)

  2. (ii)

    (tT):|φ(t)|=t(0)+t(1)++t(|t|1)(\forall t\in T):\left|\varphi(t)\right|=t(0)+t(1)+\dots+t(|t|-1).

For STS\subset T we denote by φ~(S):=clTr(φ(S))\widetilde{\varphi}(S):=\mathrm{cl}_{\mathrm{Tr}}(\varphi(S)) the tree generated by φ(S)\varphi(S).

Notation 5.6.

For each limit ordinal α<ω1\alpha<\omega_{1} we fix a bijection πα:ωα\pi_{\alpha}:\omega\rightarrow\alpha. If α=β+1\alpha=\beta+1 is a successor ordinal, we set πα(n):=β\pi_{\alpha}(n):=\beta for each nωn\in\omega. For α=0\alpha=0, we define T0:={}T_{0}:=\{\emptyset\} to be the tree which only contains the root. For α1\alpha\geq 1, we define TαT_{\alpha} (‘maximal trees of height α\alpha’) as

Tα:={}nωn^Tπα(n).T_{\alpha}:=\{\emptyset\}\cup\underset{n\in\omega}{\bigcup}n\hat{\ }T_{\pi_{\alpha}(n)}.
Definition 5.7.

For α<ω1\alpha<\omega_{1} we define

Yα:={xcX|(φ:Tαω<ω adm.)(tTα):x𝒩(φ(t))¯}{}.Y_{\alpha}:=\left\{x\in cX|\ \left(\exists\varphi:T_{\alpha}\rightarrow\omega^{<\omega}\textrm{ adm.}\right)\left(\forall t\in T_{\alpha}\right):x\in\overline{\mathcal{N}\left(\varphi(t)\right)}\right\}\cup\left\{\infty\right\}.

For α=1\alpha=1, it is not hard to see that this definition coincides with the one given in Example 5.4. This follows from the equivalence

xnωsωn𝒩(s)¯(nω)(snω<ω,|sn|=n):x𝒩(sn)¯()(φ:T1ω<ω adm.)(tT1):x𝒩(φ(t))¯defxY1,\begin{split}x\in\underset{n\in\omega}{\bigcap}\underset{s\in\omega^{n}}{\bigcup}\overline{\mathcal{N}(s)}\iff&(\forall n\in\omega)(\exists s_{n}\in\omega^{<\omega},\ |s_{n}|=n):x\in\overline{\mathcal{N}(s_{n})}\\ \overset{(\star)}{\iff}&(\exists\varphi:T_{1}\rightarrow\omega^{<\omega}\textrm{ adm.})(\forall t\in T_{1}):x\in\overline{\mathcal{N}\left(\varphi(t)\right)}\\ \overset{\textrm{def}}{\iff}&x\in Y_{1},\end{split}

where ()(\star) holds because the formula φ():=\varphi(\emptyset):=\emptyset, φ((n)):=sn\varphi((n)):=s_{n} defines an admissible mapping on T1={}{(n)|nω}T_{1}=\{\emptyset\}\cup\{(n)|\ n\in\omega\} and also any admissible mapping on T1T_{1} is defined in this way for some sequences sns_{n}, nωn\in\omega. The definition also coincides with the one given above for α=2\alpha=2 – this follows from the proof of Lemma 5.9. We now state the main properties of sets YαY_{\alpha}.

Lemma 5.8.

YαXY_{\alpha}\supset X holds for every α<ω1\alpha<\omega_{1}.

Proof.

Let σωω\sigma\in\omega^{\omega}. For tTαt\in T_{\alpha} we set φ(t):=σ|t(0)++t(|t|1)\varphi(t):=\sigma|_{t(0)+\dots+t(|t|-1)}, and observe that this is an admissible mapping which witnesses that σYα\sigma\in Y_{\alpha}. ∎

Lemma 5.9 (Complexity of YαY_{\alpha}).

For any α=λ+m<ω1\alpha=\lambda+m<\omega_{1} (where mωm\in\omega and λ\lambda is either zero or a limit ordinal), YαY_{\alpha} belongs to λ+2m+1(cX)\mathcal{F}_{\lambda+2m+1}(cX).

Proof.

For each hω<ωh\in\omega^{<\omega} and α<ω1\alpha<\omega_{1}, we will show that the set YαhY^{h}_{\alpha} defined as

Yαh:={xcX|(φ:Tαω<ω adm.)(tTα):x𝒩(h^φ(t)¯}Y^{h}_{\alpha}:=\left\{x\in cX|\ (\exists\varphi:T_{\alpha}\rightarrow\omega^{<\omega}\textrm{ adm.})(\forall t\in T_{\alpha}):x\in\overline{\mathcal{N}(h\hat{\ }\varphi(t)}\right\}

belongs to λ+2m+1(cX)\mathcal{F}_{\lambda+2m+1}(cX). The only admissible mapping from T0T_{0} is the mapping φ:\varphi:\emptyset\mapsto\emptyset, which means that Yαh=𝒩(h)¯1Y^{h}_{\alpha}=\overline{\mathcal{N}(h)}\in\mathcal{F}_{1}. In particular, the claim holds for α=0\alpha=0.

Let 1α<ω11\leq\alpha<\omega_{1}. First, we prove the following series of equivalences

xYαh(a)(φ:Tαω<ω adm.)(tTα):x𝒩(h^φ(t))¯(b)(φ:Tαω<ω adm.)(nω)(tTπα(n)):x𝒩(h^φ(n^t))¯(c)(nω)(snωn)(φn:Tπα(n)ω<ω adm.)(tTπα(n)):x𝒩(h^sn^φn(t))¯(d)xnωsnωnYπα(n)h^sn,\begin{split}x\in Y^{h}_{\alpha}\overset{(a)}{\iff}&(\exists\varphi:T_{\alpha}\rightarrow\omega^{<\omega}\textrm{ adm.})(\forall t\in T_{\alpha}):\\ &x\in\overline{\mathcal{N}\left(h\hat{\ }\varphi(t)\right)}\\ \overset{(b)}{\iff}&(\exists\varphi:T_{\alpha}\rightarrow\omega^{<\omega}\textrm{ adm.})(\forall n\in\omega)(\forall t\in T_{\pi_{\alpha}(n)}):\\ &x\in\overline{\mathcal{N}\left(h\hat{\ }\varphi(n\hat{\ }t)\right)}\\ \overset{(c)}{\iff}&(\forall n\in\omega)(\exists s_{n}\in\omega^{n})(\exists\varphi_{n}:T_{\pi_{\alpha}(n)}\rightarrow\omega^{<\omega}\textrm{ adm.})(\forall t\in T_{\pi_{\alpha}(n)}):\\ &x\in\overline{\mathcal{N}\left(h\hat{\ }s_{n}\hat{\ }\varphi_{n}(t)\right)}\\ \overset{(d)}{\iff}&x\in\bigcap_{n\in\omega}\bigcup_{s_{n}\in\omega^{n}}Y^{h\hat{\ }s_{n}}_{\pi_{\alpha}(n)},\end{split}

from which the equation (5.1) follows:

Yαh=nωsnωnYπα(n)h^sn.Y^{h}_{\alpha}=\bigcap_{n\in\omega}\bigcup_{s_{n}\in\omega^{n}}Y^{h\hat{\ }s_{n}}_{\pi_{\alpha}(n)}. (5.1)

The equivalences (a)(a) and (d)(d) are simply the definition of YαhY^{h}_{\alpha}, and (b)(b) follows from the definition of trees TαT_{\alpha}. The nontrivial part is (c)(c). To prove the implication ‘\implies’, observe that if φ\varphi is admissible, then

(nω)(tTπα(n))(φn(t)ω<ω):φ(n^t)=φ((n))^φn(t),(\forall n\in\omega)(\forall t\in T_{\pi_{\alpha}(n)})(\exists\varphi_{n}(t)\in\omega^{<\omega}):\varphi(n\hat{\ }t)=\varphi\left((n)\right)\hat{\ }\varphi_{n}(t),

the mappings φn:Tπα(n)ω<ω\varphi_{n}:T_{\pi_{\alpha}(n)}\rightarrow\omega^{<\omega} defined by this formula are admissible and |φ((n))|=n|\varphi((n))|=n. The implication from right to left follows from the fact that whenever φn:Tπα(n)ω<ω\varphi_{n}:T_{\pi_{\alpha}(n)}\rightarrow\omega^{<\omega} are admissible mappings and snωns_{n}\in\omega^{n}, the mapping φ\varphi defined by formula φ():=\varphi(\emptyset):=\emptyset, φ(n^t):=sn^φn(t)\varphi(n\hat{\ }t):=s_{n}\hat{\ }\varphi_{n}(t) is admissible as well.

We now finish the proof. If α\alpha is a successor ordinal, then πα(n)=α1\pi_{\alpha}(n)=\alpha-1 holds for all nωn\in\omega. Therefore, we can rewrite (5.1) as

Yαh=nωsnωnYπα(n)h^sn=nωsnωnYα1h^snY^{h}_{\alpha}=\bigcap_{n\in\omega}\bigcup_{s_{n}\in\omega^{n}}Y^{h\hat{\ }s_{n}}_{\pi_{\alpha}(n)}=\bigcap_{n\in\omega}\bigcup_{s_{n}\in\omega^{n}}Y^{h\hat{\ }s_{n}}_{\alpha-1}

and observe that each ‘successor’ step increases the complexity by ()σδ(\cdot)_{\sigma\delta}. Lastly, assume that α\alpha is a limit ordinal and Yαsβ<αβY^{s}_{\alpha^{\prime}}\in\underset{\beta<\alpha}{\bigcup}\mathcal{F}_{\beta} holds for all sω<ωs\in\omega^{<\omega} and α<α\alpha^{\prime}<\alpha. Since for limit α\alpha, we have πα(n)<α\pi_{\alpha}(n)<\alpha for each nωn\in\omega, (5.1) gives

Yαh=nωsnωnYπα(n)h^sn(β<αβ)σδ=α+1,Y^{h}_{\alpha}=\bigcap_{n\in\omega}\bigcup_{s_{n}\in\omega^{n}}Y^{h\hat{\ }s_{n}}_{\pi_{\alpha}(n)}\in\left(\underset{\beta<\alpha}{\bigcup}\mathcal{F}_{\beta}\right)_{\sigma\delta}=\mathcal{F}_{\alpha+1},

which is what we wanted to prove. ∎

5.2 Some auxiliary results

In this section, we give a few tools which will be required to obtain our main result. Later, we will need to show that for a suitable α\alpha, YαXY_{\alpha}\setminus X is empty. In order to do this, we first explore some properties of those xYαx\in Y_{\alpha} which are in XX, and of those xYαx\in Y_{\alpha} which do not belong to XX. We address these two possibilities separately in Lemma 5.10 and Lemma 5.11 and show that they are related with the properties of the corresponding admissible mappings444By such a mapping we mean an admissible mapping which witnesses that xYαx\in Y_{\alpha} (for details, see Definitions 5.5 and 5.7)..

Lemma 5.10 (The non-WF case).

For any σωω\sigma\in\omega^{\omega} and any increasing sequence of integers (nk)kω(n_{k})_{k\in\omega}, we have k𝒩(σ|nk)¯cX={σ}\bigcap_{k}\overline{\mathcal{N}(\sigma|n_{k})}^{cX}=\{\sigma\}. In particular, we have

kω𝒩(σ|nk)¯cXX.\bigcap_{k\in\omega}\overline{\mathcal{N}(\sigma|n_{k})}^{cX}\subset X.
Proof.

Suppose that σ\sigma and nkn_{k} are as above and the intersection k𝒩(σ|nk)¯\bigcap_{k}\overline{\mathcal{N}(\sigma|n_{k})} contains some xσx\neq\sigma. Clearly, we have xcXXx\in cX\setminus X. By Lemma 5.3, there is some discrete (in ωω\omega^{\omega}) set HωωH\subset\omega^{\omega}, such that for each kωk\in\omega, xx belongs to 𝒩(σ|nk)H¯\overline{\mathcal{N}(\sigma|n_{k})\cap H}. Since the sets 𝒩(σ|nk)\mathcal{N}(\sigma|n_{k}) form a neighborhood basis of σ\sigma, there exists n0ωn_{0}\in\omega, such that 𝒩(sn0)H\mathcal{N}(s_{n_{0}})\cap H is a singleton. In particular, this means that 𝒩(sn0)H\mathcal{N}(s_{n_{0}})\cap H is closed in cXcX, and we have x𝒩(sn0)H¯=𝒩(sn0)HHXx\in\overline{\mathcal{N}(s_{n_{0}})\cap H}=\mathcal{N}(s_{n_{0}})\cap H\subset H\subset X – a contradiction. ∎

On the other hand, Lemma 5.10 shows that if the assumptions of Lemma 5.10 are not satisfied, we can find a broom set555Recall that the collections B~α\widetilde{B}_{\alpha} were introduced in Definition 4.9. of class α\alpha which corresponds to xx. We remark that Lemma 5.11 is a key step on the way to Proposition 5.13 – in particular, it is the step in which the admissible mappings are an extremely useful tool.

Lemma 5.11 (Well founded case).

Let xYαx\in Y_{\alpha} and suppose that φ\varphi is any admissible mapping witnessing this fact. If φ~(Tα)WF\widetilde{\varphi}(T_{\alpha})\in\textrm{WF}, then there exists B~αB\in\widetilde{\mathcal{B}}_{\alpha} with Bφ~(Tα)B\subset\widetilde{\varphi}(T_{\alpha}).

Proof.

We prove the statement by induction over α\alpha. For α=0\alpha=0 we have T0={}T_{0}=\{\emptyset\} and φ()ω<ω=~0\varphi(\emptyset)\in\omega^{<\omega}=\widetilde{\mathcal{B}}_{0}.

Consider α>0\alpha>0 and assume that the statement holds for every β<α\beta<\alpha. We have {(n)|nω}Tα\{(n)|\ n\in\omega\}\subset T_{\alpha}. By the second defining property of admissible mappings, we get φ((m))φ((n))\varphi((m))\neq\varphi((n)) for distinct m,nωm,n\in\omega, which means that the tree T:=φ~({(n)|nω})T:=\widetilde{\varphi}(\{(n)|\ n\in\omega\}) is infinite. Since Tφ~(Tα)WFT\subset\widetilde{\varphi}(T_{\alpha})\in\textrm{WF}, we use König’s lemma to deduce that TT contains some hω<ωh\in\omega^{<\omega} with infinite set of successors S:=succT(h)S:=\textrm{succ}_{T}(h).

For each sSs\in S we choose one nsωn_{s}\in\omega with φ(ns)s\varphi(n_{s})\sqsupset s. In the previous paragraph, we have observed that (s(|h|))sS\left(s(|h|)\right)_{s\in S} is a forking sequence. From the first property of admissible mappings, we get

(sS)(tTα s.t. t(0)=ns):φ(t)s.(\forall s\in S)(\forall t\in T_{\alpha}\textrm{ s.t. }t(0)=n_{s}):\varphi(t)\sqsupset s. (5.2)

In particular, for each sSs\in S we have

φ({tTα|t(0)=ns})=s^Ss for some Ssω<ω.\varphi(\{t\in T_{\alpha}|\ t(0)=n_{s}\})=s\hat{\ }S_{s}\textrm{ for some }S_{s}\subset\omega^{<\omega}.

By definition666Recall that TαT_{\alpha} is an ω\omega-ary tree of height α\alpha, defined in Notation 5.6. of TαT_{\alpha}, we have {tω<ω|ns^tTα}=Tπα(ns)\{t\in\omega^{<\omega}|\ n_{s}\hat{\ }t\in T_{\alpha}\}=T_{\pi_{\alpha}(n_{s})}. By (5.2), each φ(ns^t)\varphi(n_{s}\hat{\ }t) for tTπα(ns)t\in T_{\pi_{\alpha}(n_{s})} is of the form φ(ns^t)=φ(ns)^φs(t)\varphi(n_{s}\hat{\ }t)=\varphi(n_{s})\hat{\ }\varphi_{s}(t) for some φs(t)ω<ω\varphi_{s}(t)\in\omega^{<\omega}, and, since φ\varphi is admissible, the mapping φs(t):Tπα(ns)ω<ω\varphi_{s}(t):T_{\pi_{\alpha}(n_{s})}\rightarrow\omega^{<\omega} defined by this formula is admissible as well.

It follows from the induction hypothesis that clTr(φs(Tπα(n)))\mathrm{cl}_{\mathrm{Tr}}\left(\varphi_{s}(T_{\pi_{\alpha}(n)})\right) contains some ~πα(n)\widetilde{\mathcal{B}}_{\pi_{\alpha}(n)}-set DsD_{s}. Let Csφs(Tπα(n))C_{s}\subset\varphi_{s}(T_{\pi_{\alpha}(n)}) be some finite extension of DsD_{s}. By Remark 4.10, CsC_{s} belongs to ~πα(n)\widetilde{\mathcal{B}}_{\pi_{\alpha}(n)}. Finally, since

s^Ss=φ(ns)^φs(Tπα(ns))φ(ns)^Cs~πα(n),s\hat{\ }S_{s}=\varphi(n_{s})\hat{\ }\varphi_{s}(T_{\pi_{\alpha}(n_{s})})\supset\varphi(n_{s})\hat{\ }C_{s}\in\widetilde{\mathcal{B}}_{\pi_{\alpha}(n)},

there also exists some BsSsB_{s}\subset S_{s} with Bs~πα(n)B_{s}\in\widetilde{\mathcal{B}}_{\pi_{\alpha}(n)} (the set BsB_{s} can have the same bristles and forking sequence as CsC_{s}, but ss might be shorter than φ(ns)\varphi(n_{s}), so BsB_{s} might have a longer handle than CsC_{s}). We finish the proof by observing that

φ~(Tα)φ(sS{tTα|t(0)=ns})sSs^SssSh^s(|h|)^Bs.\widetilde{\varphi}(T_{\alpha})\supset\varphi\left(\underset{s\in S}{\bigcup}\left\{t\in T_{\alpha}|\ t(0)=n_{s}\right\}\right)\supset\underset{s\in S}{\bigcup}s\hat{\ }S_{s}\supset\underset{s\in S}{\bigcup}h\hat{\ }s(|h|)\hat{\ }B_{s}. (5.3)

Because the set SS is infinite, the rightmost set in (5.3) is, by definition, an element of ~α\widetilde{\mathcal{B}}_{\alpha}. ∎

We will eventually want to show that the assumptions of Lemma 5.11 can only be satisfied if the family \mathcal{E} is sufficiently rich. To this end, we need to extend Lemma 4.4 to a situation where a set HωωH\subset\omega^{\omega} cannot be covered by finitely many ‘infinite broom sets’777That is, the infinite extensions of broom sets, introduced in Definition 5.1.:

Lemma 5.12.

Suppose that a set HωωH\subset\omega^{\omega} contains an infinite extension of some Bω<ωB\subset\omega^{<\omega} which satisfies ri(B)αr_{i}(B)\geq\alpha.888Recall that rir_{i} denotes the ‘infinite branching rank’ introduced in Definition 4.3. Then HH cannot be covered by finitely many elements of β<αβ\bigcup_{\beta<\alpha}\mathcal{E}_{\beta}.

Proof.

Let α<ω1\alpha<\omega_{1}, BB and HH be as above. For contradiction, assume that Hj=0kEjH\subset\bigcup_{j=0}^{k}E_{j} holds for some EjβjE_{j}\in\mathcal{E}_{\beta_{j}}, βj<α\beta_{j}<\alpha.

By definition of βj\mathcal{E}_{\beta_{j}}, each EjE_{j} is an infinite extension of some BjβjB_{j}\in\mathcal{B}_{\beta_{j}} – in other words, there are bijections ψj:BjEj\psi_{j}:B_{j}\rightarrow E_{j} satisfying ψj(s)s\psi_{j}(s)\sqsupset s for each sBjs\in B_{j}. Similarly, HH contains some infinite extension of BB, which means there exists an injective mapping ψ:BH\psi:B\rightarrow H satisfying ψ(s)s\psi(s)\sqsupset s for each sBs\in B. Clearly, we have

ψ(B)Hj=0kEj=j=0kψj(Bj).\psi(B)\subset H\subset\bigcup_{j=0}^{k}E_{j}=\bigcup_{j=0}^{k}\psi_{j}(B_{j}). (5.4)

Firstly, observe that if tBt\in B and sBjs\in B_{j} satisfy ψj(s)=ψ(t)\psi_{j}(s)=\psi(t), the sequences ss and tt must be comparable. This means that for each jkj\leq k, the following formula correctly defines a mapping φj:Bjω<ω\varphi_{j}:B_{j}\rightarrow\omega^{<\omega}:

φj(s):={s, for ψj(s)ψ(B),ψ1(ψj(s)), for sψ1(ψj(s)),s, for sψ1(ψj(s)),\varphi_{j}(s):=\begin{cases}s,&\textrm{ for }\psi_{j}(s)\notin\psi(B),\\ \psi^{-1}(\psi_{j}(s)),&\textrm{ for }s\sqsubset\psi^{-1}(\psi_{j}(s)),\\ s,&\textrm{ for }s\sqsupset\psi^{-1}(\psi_{j}(s)),\end{cases}

which satisfies φj(s)s\varphi_{j}(s)\sqsupset s for each sBjs\in B_{j}. Since no two elements of BjB_{j} are comparable, it also follows that φj\varphi_{j} is injective and thus φ(Bj)\varphi(B_{j}) is an extension of BjB_{j}. By Remark 4.10, this implies that φj(Bj)βj\varphi_{j}(B_{j})\in\mathcal{B}_{\beta_{j}}.

By (5.4), for tBt\in B there exists some jkj\leq k and sBjs\in B_{j}, such that ψj(s)=ψ(t)\psi_{j}(s)=\psi(t). It follows from the definition of ψj\psi_{j} that

t=ψ1(ψ(t))=ψ1(ψj(s))φj(s).t=\psi^{-1}(\psi(t))=\psi^{-1}(\psi_{j}(s))\sqsubset\varphi_{j}(s).

In other words, we get

BclTr(j=0kφj(Bj)).B\subset\mathrm{cl}_{\mathrm{Tr}}\left(\bigcup_{j=0}^{k}\varphi_{j}(B_{j})\right).

We observe that this leads to a contradiction:

α\displaystyle\alpha =ri(B)ri(clTr(j=0kφj(Bj)))=ri(j=0kφj(Bj))=\displaystyle=r_{i}(B)\leq r_{i}\left(\mathrm{cl}_{\mathrm{Tr}}\left(\bigcup_{j=0}^{k}\varphi_{j}(B_{j})\right)\right)=r_{i}\left(\bigcup_{j=0}^{k}\varphi_{j}(B_{j})\right)=
=maxjri(φj(Bj))maxjβj<α.\displaystyle=\max_{j}\ r_{i}\left(\varphi_{j}(B_{j})\right)\leq\max_{j}\beta_{j}<\alpha.

5.3 Absolute complexity of XX_{\mathcal{E}}

In this section, we prove an upper bound on the absolute complexity of spaces XX_{\mathcal{E}} for any collection of infinite broom sets \mathcal{E}. Talagrand’s earlier result will then imply that, in some cases, this bound is sharp.

Proposition 5.13 (Complexity of XX_{\mathcal{E}}).

For any integer mωm\in\omega and limit ordinal λ<ω1\lambda<\omega_{1}, we have

  1. 1)

    m\mathcal{E}\subset\mathcal{E}_{m} \implies XX_{\mathcal{E}} is absolutely 2m+1\mathcal{F}_{2m+1};

  2. 2)

    β<λβ\mathcal{E}\subset\underset{\beta<\lambda}{\bigcup}\mathcal{E}_{\beta} \implies XX_{\mathcal{E}} is absolutely λ+1\mathcal{F}_{\lambda+1};

  3. 3)

    λ+m\mathcal{E}\subset\mathcal{E}_{\lambda+m} \implies XX_{\mathcal{E}} is absolutely λ+2m+3\mathcal{F}_{\lambda+2m+3}.

Proof.

Suppose that λ\lambda and mm are as above, \mathcal{E} is a family of discrete subsets of ωω\omega^{\omega} and cXcX is a compactification of the space X:=XX:=X_{\mathcal{E}}. Denote α1:=m\alpha_{1}:=m, α2:=λ\alpha_{2}:=\lambda and α3:=λ+m+1\alpha_{3}:=\lambda+m+1. Note that we are only going to use the following property of \mathcal{E} and αi\alpha_{i}:

 satisfies the hypothesis of i) for i=1β<αi+1β, satisfies the hypothesis of i) for i{2,3}β<αiβ.\begin{split}&\textrm{$\mathcal{E}$ satisfies the hypothesis of }i)\textrm{ for }i=1\ \ \ \ \ \implies\mathcal{E}\subset\bigcup_{\beta<{\alpha_{i}+1}}\mathcal{E}_{\beta},\\ &\textrm{$\mathcal{E}$ satisfies the hypothesis of }i)\textrm{ for }i\in\{2,3\}\implies\mathcal{E}\subset\bigcup_{\beta<{\alpha_{i}}}\mathcal{E}_{\beta}.\end{split} (5.5)

We will show that X=YαiX=Y_{\alpha_{i}} holds in each of these cases999Where YαY_{\alpha} is given by Definition 5.7.. Once we have this identity, the conclusion immediately follows from Lemma 5.9.

Let i{1,2,3}i\in\{1,2,3\}. Suppose that there exists xYαiXx\in Y_{\alpha_{i}}\setminus X and let φ:Tαiω<ω\varphi:T_{\alpha_{i}}\rightarrow\omega^{<\omega} be an admissible mapping101010Recall that admissible mappings were introduced in Definition 5.5. The ‘tilde version’ is defined as φ~(S):=clTr(φ(S))\widetilde{\varphi}(S):=\mathrm{cl}_{\mathrm{Tr}}(\varphi(S)). witnessing that xYαix\in Y_{\alpha_{i}}. Moreover, let HωωH\subset\omega^{\omega} be a set satisfying the conclusion of Lemma 5.3.

Recall that by definition of YαiY_{\alpha_{i}}, we have

(tTαi):x𝒩(φ(t))¯.\left(\forall t\in T_{\alpha_{i}}\right):x\in\ \overline{\mathcal{N}(\varphi(t))}. (5.6)

By Lemma 5.10, φ~(Tαi)\widetilde{\varphi}(T_{\alpha_{i}}) contains no infinite branches. Therefore, we can apply Lemma 5.11 to obtain a ~αi\widetilde{\mathcal{B}}_{\alpha_{i}}-set111111~α\widetilde{\mathcal{B}}_{\alpha} is the ‘disjoint version’ of broom collection α\mathcal{B}_{\alpha} (see Definitions 4.8 and 4.9). Bφ~(Tαi)B\subset\widetilde{\varphi}(T_{\alpha_{i}}). Since, by (5.6), xx belongs to 𝒩(s)¯\overline{\mathcal{N}(s)} for each sBs\in B, we conclude that xH𝒩(s)¯x\in\overline{H\cap\mathcal{N}(s)} holds for every sBs\in B (Lemma 5.3, (ii)(ii)). In particular, all the intersections H𝒩(s)H\cap\mathcal{N}(s) must be infinite. Note that for the next part of the proof, we will only need the intersections to be non-empty. The fact that they are infinite will be used in the last part of the proof.

We can now conclude the proof of 2)2) and 3)3). Let i{2,3}i\in\{2,3\} and assume that the hypothesis of i)i) holds. Since H𝒩(s)H\cap\mathcal{N}(s) is non-empty for each sBs\in B, it follows that HH contains an infinite extension of BB. Since ri(B)=αir_{i}(B)=\alpha_{i} holds by Lemma 4.11, Lemma 5.12 yields that HH cannot be covered by finitely many elements of β<αiβ\bigcup_{\beta<{\alpha_{i}}}\mathcal{E}_{\beta}. By (5.5), we have β<αiβ\mathcal{E}\subset\bigcup_{\beta<{\alpha_{i}}}\mathcal{E}_{\beta}. This contradicts the first part of Lemma 5.3 (which claims that HH can be covered by finitely many elements of \mathcal{E}).

For the conclusion of the proof of 1)1), let i=1i=1 and assume by (5.5) that β<αi+1β\mathcal{E}\subset\bigcup_{\beta<\alpha_{i}+1}\mathcal{E}_{\beta} holds. Compare this situation with the setting from the previous paragraph – it is clear that if we show that HH in fact contains an infinite extension of some B~\widetilde{B} with ri(B~)αi+1r_{i}(\widetilde{B})\geq\alpha_{i}+1, we can replace BB by B~\widetilde{B}. We can then apply the same proof which worked for 2)2) and 3)3).

To find B~\widetilde{B}, enumerate BB as B={sn|nω}B=\{s_{n}|\ n\in\omega\}. Since each H𝒩(sn)H\cap\mathcal{N}(s_{n}) is infinite, there exist distinct sequences σnkH\sigma^{k}_{n}\in H, kωk\in\omega, satisfying snσnks_{n}\sqsubset\sigma^{k}_{n}. We use these sequences to obtain snkω<ωs_{n}^{k}\in\omega^{<\omega} for n,kωn,k\in\omega, which satisfy snsnkσnks_{n}\sqsubset s^{k}_{n}\sqsubset\sigma^{k}_{n} and klk\neq l \implies snksnls^{k}_{n}\neq s^{l}_{n}. We denote B~:={snk|n,kω}\widetilde{B}:=\{s^{k}_{n}|\ n,k\in\omega\}. Clearly, this set satisfies Di(B~)BD_{i}(\widetilde{B})\supset B and HH contains some infinite extension of B~\widetilde{B}. Because ri(B)=αir_{i}(B)=\alpha_{i} is finite, we have ri(B~)1+αi=αi<ωαi+1>αi=ri(B)r_{i}(\widetilde{B})\geq 1+\alpha_{i}\overset{\alpha_{i}<\omega}{=}\alpha_{i}+1>\alpha_{i}=r_{i}(B), which completes the proof. ∎

Applying Proposition 5.13 to the construction from [Tal] immediately yields the following corollary:

Theorem 5.14.

For every even 4α<ω14\leq\alpha<\omega_{1}, there exists a space TαT_{\alpha} with properties

  1. (i)

    TαT_{\alpha} is an FσδF_{\sigma\delta} space;

  2. (ii)

    TαT_{\alpha} is not an absolute α\mathcal{F}_{\alpha} space;

  3. (iii)

    TαT_{\alpha} is an absolute α+1\mathcal{F}_{\alpha+1} space.

Proof.

The existence of spaces TαT_{\alpha} which satisfy (i)(i) and (ii)(ii) follows from [Tal]. In this paper, Talagrand constructed a family 𝒜Tω1\mathcal{A}_{T}\subset\mathcal{E}_{\omega_{1}}, which is both ‘rich enough’ and almost-disjoint (that is, the intersection of any two families is finite). This is accomplished by a smart choice of bristles and forking sequences on the ‘highest level of each broom’.

He then shows that for a suitable α~\widetilde{\alpha}, any space XX_{\mathcal{E}} with 𝒜Tα~\mathcal{E}\supset\mathcal{A}_{T}\cap\mathcal{E}_{\widetilde{\alpha}}, where \mathcal{E} is almost-disjoint, satisfies (i)(i) and (ii)(ii). In particular, this holds for Tα:=X𝒜Tα~T_{\alpha}:=X_{\mathcal{A}_{T}\cap\mathcal{E}_{\widetilde{\alpha}}}. Note however that Talagrand was interested in the ‘maximal’ version of the construction, which is the space T:=X𝒜TT:=X_{\mathcal{A}_{T}}. We, on the other hand, will use the ‘intermediate steps’ of his construction.

For us, the details of the construction are not relevant – the only properties we need are (i)(i), (ii)(ii) and the correspondence between α\alpha and α~\widetilde{\alpha}.121212Note that in [Tal], the author uses slightly different (but equivalent) definition of broom families α\mathcal{E}_{\alpha}, which shifts their numbering for finite α\alpha by 11. This correspondence is such that the broom class only increases with countable intersections (that is, odd steps of α\mathcal{F}_{\alpha}), which translates to our notation as follows:

  • α=2nω\alpha=2n\in\omega, α4\alpha\geq 4 Tα=X\implies T_{\alpha}=X_{\mathcal{E}} for some n{\mathcal{E}}\subset\mathcal{E}_{n};

  • α<ω1\alpha<\omega_{1} is a limit ordinal Tα=X\implies T_{\alpha}=X_{\mathcal{E}} for some β<αβ{\mathcal{E}}\subset\underset{\beta<\alpha}{\bigcup}\mathcal{E}_{\beta};

  • α=λ+2n+2\alpha=\lambda+2n+2 for nωn\in\omega and limit λ\lambda Tα=X\implies T_{\alpha}=X_{\mathcal{E}} for some λ+n{\mathcal{E}}\subset\mathcal{E}_{\lambda+n}.

In all the cases, it follows from Proposition 5.13 that XX_{\mathcal{E}} is absolutely α+1\mathcal{F}_{\alpha+1}. ∎

If β<ω1\beta<\omega_{1} is the least ordinal for which some XX is an (absolute) β\mathcal{F}_{\beta} space, we say that the (absolute) complexity of XX is β\mathcal{F}_{\beta}. Using this notation, Theorem 5.14 can be rephrased as “for every odd 5α<ω15\leq\alpha<\omega_{1}, there exists an FσδF_{\sigma\delta} space whose absolute complexity is α\mathcal{F}_{\alpha}”. By modifying the spaces from Theorem 5.14, we obtain the following result:

Corollary 5.15.

For every two countable ordinals αβ3\alpha\geq\beta\geq 3, α\alpha odd, there exists a space Xα,βX_{\alpha,\beta}, such that

  1. (i)

    the complexity of Xα,βX_{\alpha,\beta} is β\mathcal{F}_{\beta};

  2. (ii)

    the absolute complexity of Xα,βX_{\alpha,\beta} is α\mathcal{F}_{\alpha}.

Proof.

If α=β\alpha=\beta, this is an immediate consequence of absoluteness of Borel classes. To make this claim more precise, let PP be some uncountable Polish space. By Remark 1.3 (ii)(ii), β(P)\mathcal{F}_{\beta}(P) corresponds to some Borel class 𝒞(P)\mathcal{C}(P). We define ZβZ_{\beta} as one of the subspaces of PP, which are of the class 𝒞(P)\mathcal{C}(P), but not of the ‘dual’ class – that is, if 𝒞(P)=Σγ0(P)\mathcal{C}(P)=\Sigma^{0}_{\gamma}(P) for some γ<ω1\gamma<\omega_{1}, then ZβΣγ0(P)Πγ0(P)Z_{\beta}\in\Sigma^{0}_{\gamma}(P)\setminus\Pi^{0}_{\gamma}(P) (and vice versa for 𝒞(P)=Πγ0(P)\mathcal{C}(P)=\Pi^{0}_{\gamma}(P)). For the existence of such a set, see for example Corollary 3.6.8 in [Sri].

Since ZβZ_{\beta} is separable and metrizable, Theorem 2.3 guarantees that it is an absolute β\mathcal{F}_{\beta} space. However, it is not of the class β\mathcal{F}_{\beta^{\prime}} for any β<β\beta^{\prime}<\beta, because that would by Remark 1.3 imply that ZβZ_{\beta} belongs to β(P)Πγ0(P)\mathcal{F}_{\beta^{\prime}}(P)\subset\Pi^{0}_{\gamma}(P). This shows that if β=α\beta=\alpha, the space Xα,β:=ZβX_{\alpha,\beta}:=Z_{\beta} has the desired properties.

If β<α\beta<\alpha, we define XX as the topological sum of ZβZ_{\beta} and the space Tα1T_{\alpha-1} from Theorem 5.14. Since Tα1Fσδ(βTα1)=3(βTα1)T_{\alpha-1}\in F_{\sigma\delta}(\beta T_{\alpha-1})=\mathcal{F}_{3}(\beta T_{\alpha-1}) and β3\beta\geq 3, we get that Xα,βX_{\alpha,\beta} is of the class β\mathcal{F}_{\beta} in the topological sum βXα,β=βZββTα1\beta X_{\alpha,\beta}=\beta Z_{\beta}\oplus\beta T_{\alpha-1}. By previous paragraph, it is of no lower class in βXα,β\beta X_{\alpha,\beta}. By Theorem 5.14 (and the previous paragraph), Xα,βX_{\alpha,\beta} also has the correct absolute complexity. ∎

To get a complete picture of possible combinations of complexity and absolute complexity for \mathcal{F}-Borel spaces, it remains to answer the following questions:

Problem 5.16.
  1. (i)

    Let 4α<ω14\leq\alpha<\omega_{1} be an even ordinal number. Does there exists a space XαX_{\alpha}, whose absolute complexity is α\mathcal{F}_{\alpha} and complexity is β\mathcal{F}_{\beta} for some β<α\beta<\alpha? What is the lowest possible value of β\beta?

  2. (ii)

    If a space XX is \mathcal{F}-Borel in every compactification, is it necessarily absolutely α\mathcal{F}_{\alpha} for some α<ω1\alpha<\omega_{1}?

One could expect that Corollary 5.15 might also hold for even α\alpha, so that we should be able to find an FσδF_{\sigma\delta} space answering the first part of Problem 5.16 in positive.

Acknowledgment

I would like to thank my supervisor, Ondřej Kalenda, for numerous very helpful suggestions and fruitful consultations regarding this paper. This work was supported by the research grant GAUK No. 915.

References

  • [Eng] Ryszard Engelking. General topology, Sigma series in pure mathematics, vol. 6, 1989.
  • [Fre] Ronald C. Freiwald. An introduction to set theory and topology. 2014.
  • [Fro1] Zdeněk Frolík. Generalizations of the Gδ-property of complete metric spaces. Czechoslovak Mathematical Journal, 10(3):359–379, 1960.
  • [Fro2] Zdeněk Frolík. On the descriptive theory of sets. Czechoslovak Mathematical Journal, 13(3):335–359, 1963.
  • [HS] Petr Holický and Jiří Spurný. Perfect images of absolute Souslin and absolute Borel Tychonoff spaces. Topology and its Applications, 131(3):281–294, 2003.
  • [JK] Heikki J.K. Junnila and Hans-Peter A. Küunzi. Characterizations of absolute Fσδ-sets. Czechoslovak Mathematical Journal, 48(1):55–64, 1998.
  • [Kech] Alexander Kechris. Classical descriptive set theory, volume 156. Springer Science & Business Media, 2012.
  • [KK] Ondřej F.K. Kalenda and Vojtěch Kovařík. Absolute Fσδ spaces. arXiv preprint arXiv:1703.03066, 2017.
  • [KKLP] Jerzy Kakol, Wiesław Kubiś, and Manuel López-Pellicer. Descriptive topology in selected topics of functional analysis, volume 24. Springer Science & Business Media, 2011.
  • [MP] Witold Marciszewski and Jan Pelant. Absolute Borel sets and function spaces. Transactions of the American Mathematical Society, 349(9):3585–3596, 1997.
  • [Raj] Matias Raja. On some class of Borel measurable maps and absolute Borel topological spaces. Topology and its Applications, 123(2):267–282, 2002.
  • [Sri] Sashi M. Srivastava. A course on Borel sets, volume 180. Springer Science & Business Media, 2008.
  • [Tal] Michel Talagrand. Choquet simplexes whose set of extreme points is K-analytic. In Annales de l’institut Fourier, volume 35, pages 195–206, 1985.