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

Complexity classes of Polishable subgroups

Martino Lupini School of Mathematics and Statistics
Victoria University of Wellington
PO Box 600, 6140 Wellington, New Zealand
martino.lupini@vuw.ac.nz
Abstract.

In this paper we further develop the theory of canonical approximations of Polishable subgroups of Polish groups, building on previous work of Solecki and Farah–Solecki. In particular, we obtain a characterization of such canonical approximations in terms of their Borel complexity class. As an application we provide a complete list of all the possible Borel complexity classes of Polishable subgroups of Polish groups or, equivalently, of the ranges of continuous group homomorphisms between Polish groups. We also provide a complete list of all the possible Borel complexity classes of the ranges of: continuous group homomorphisms between non-Archimedean Polish groups; continuous linear maps between separable Fréchet spaces; continuous linear maps between separable Banach spaces.

Key words and phrases:
Polish group, Polishable subgroup, Borel complexity class, Solecki subgroup, non-Archimedean Polish group, Fréchet space, Banach space
2000 Mathematics Subject Classification:
Primary 54H05, 22A05; Secondary 46B99, 46A04
The author was partially supported by the Marsden Fund Fast-Start Grant VUW1816 and by a Rutherford Discovery Fellowship from the Royal Society of New Zealand

1. Introduction

The goal of this paper is to exactly pin down the possible Borel complexity classes of Polishable subgroups of Polish groups. Most of the equivalence relations studied in the context of Borel complexity theory (and mathematics in general) arise as orbit equivalence relations associated with continuous actions of Polish groups on Polish spaces. Many of these actions can be seen as the (left) translation action associated with a continuous group homomorphism φ:HG\varphi:H\rightarrow G between Polish groups. In such a case, the image φ(H)\varphi\left(H\right) of φ\varphi inside of GG is Borel, and the potential complexity class (in the sense of Louveau [11]) of the orbit equivalence relation associated with the translation action of HH on GG is essentially the same as the Borel complexity class of φ(H)\varphi\left(H\right) of φ\varphi inside of GG; see Theorem 3.3 for a precise statement. It is thus an interesting problem to determine what are the possible values for the Borel complexity class of such a subgroup.

The study of Polishable subgroups of Polish groups, which are precisely the ranges of continuous homomorphisms between Polish groups, has been undertaken by several authors over a number of years. The problem of determining their complexity has been considered as early as the 1970s, when Saint-Raymond proved that there exist Polishable subgroups of \mathbb{R}^{\mathbb{N}} that are arbitrarily high in the Borel hierarchy [18]. A construction of arbitrarily complex non-Archimedean Polishable subgroups of 2\mathbb{Z}_{2}^{\mathbb{N}} was presented by Hjorth, Kechris, and Louveau in [8]. Hjorth constructed in [7] arbitrarily complex Polishable subgroups of any uncountable abelian Polish groups. Farah and Solecki in [5], building on previous work of Sain-Raymond in the context of separable Fréchet spaces [18], related the least multiplicative Borel class containing a given Polishable subgroup to the length of the canonical approximation of that Polishable subgroup as in [16, 17].

In this paper, we refine the analysis from [5] by considering not only the multiplicative classes in the Borel hierarchy, but also the additive and difference classes. By relating the Borel complexity class of a Polishable subgroup to its canonical approximation, we completely characterize the possible Borel complexity classes of Polishable subgroups of Polish groups.

Theorem 1.1.

If HH is a Polishable subgroup of a Polish group GG, then the Borel complexity class of HH is one of the following: 𝚷1+λ0\boldsymbol{\Pi}_{1+\lambda}^{0}, 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0}, D(𝚷1+λ+n+10)D(\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}), 𝚷1+λ+n+20\boldsymbol{\Pi}_{1+\lambda+n+2}^{0} for λ<ω1\lambda<\omega_{1} either zero or a limit ordinal, and n<ωn<\omega. Furthermore, each of these classes is the Borel complexity class of a Polishable subgroup of \mathbb{Z}^{\mathbb{N}}.

Theorem 4.1 from [8, Section 5] shows that the complexity class D(𝚷1+λ+10)D(\boldsymbol{\Pi}_{1+\lambda+1}^{0}) where λ\lambda is either zero or a countable limit ordinal cannot arise in the context of Theorem 1.1 if one demands HH to be non-Archimedean. In this case, we have the following characterization.

Theorem 1.2.

If HH is a non-Archimedean Polishable subgroup of a Polish group GG, then the Borel complexity class of HH in GG is one of the following: 𝚷1+λ0\boldsymbol{\Pi}_{1+\lambda}^{0}, 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0}, D(𝚷1+λ+n+20)D(\boldsymbol{\Pi}_{1+\lambda+n+2}^{0}), 𝚷1+λ+n+20\boldsymbol{\Pi}_{1+\lambda+n+2}^{0} for λ<ω1\lambda<\omega_{1} either zero or a limit ordinal, and n<ωn<\omega. Furthermore, each of these classes is the complexity class of a non-Archimedean Polishable subgroup of \mathbb{Z}^{\mathbb{N}}.

The existence assertions in Theorem 1.1 and Theorem 1.2 are proved by providing a unified approach to the constructions in [18] and [8, Section 5] of arbitrarily complex Polishable subgroups, together with a careful analysis of their canonical approximations in the sense of Solecki [16, 17].

Theorem 1.1 entails in particular a negative answer to a Question 6.3(1) from [4]. Let XX be a separable Banach space with a Schauder basis (xn)n\left(x_{n}\right)_{n\in\mathbb{N}}. Then the collection coef(X,(xn))\mathrm{coef}(X,(x_{n})) of (λn)n(\lambda_{n})_{n\in\mathbb{N}}\in\mathbb{R}^{\mathbb{N}} such that nλnxn\sum_{n\in\mathbb{N}}\lambda_{n}x_{n} converges in XX is a Polishable subgroup of \mathbb{R}^{\mathbb{N}}. Question 6.3(1) asks whether there is an example of such a Polishable subgroup that is 𝚫30\boldsymbol{\Delta}_{3}^{0} and not D(𝚺20)D(\boldsymbol{\Sigma}_{2}^{0}). By Theorem 1.1, a 𝚫30\boldsymbol{\Delta}_{3}^{0} Polishable subgroup of a Polish group must be D(𝚺20)D(\boldsymbol{\Sigma}_{2}^{0}).

We also apply the techniques of this paper to provide a complete characterization of the Borel complexity classes of the ranges of continuous homomorphisms between separable Fréchet spaces and between separable Banach spaces.

Theorem 1.3.

The complexity classes in Theorem 1.1 form a complete list of all the Borel complexity classes of the ranges of continuous linear maps between separable Fréchet spaces.

Theorem 1.4.

The following is a complete list of all the Borel complexity classes of the ranges of continuous linear maps between separable Banach spaces: 𝚷10\boldsymbol{\Pi}_{1}^{0}, 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0}, D(𝚷1+λ+n+10)D(\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}), 𝚷1+λ+n+20\boldsymbol{\Pi}_{1+\lambda+n+2}^{0} for λ<ω1\lambda<\omega_{1} either zero or a limit ordinal, and n<ωn<\omega.

Continuous linear maps with arbitrarily complex range, with a fixed separable Banach space or separable Frechet space as target, were constructed in [3, 12].

The rest of this paper is organized as follows. In Section 2 and Section 3 we recall some definitions and known results concerning Polishable subgroups and their Borel complexity class. In Section 4 we recall the definition of the canonical approximation of a Polishable subgroup, whose elements we call Solecki subgroups as they were originally described by Solecki in [16]. In Section 5, building on the work of Farah and Solecki, we refine the analysis from [5] to characterize the Solecki subgroups in terms of their Borel complexity class. This is then applied in Section 6 to obtain the characterization of complexity classes of Polishable subgroups as in Theorem 1.1. Section 7 shows that the length of the canonical approximation, called the Polishable rank in [5], coincides with a notion of rank originally considered by Saint-Raymond in [18] in the context of separable Fréchet spaces. The existence assertions in Theorem 1.1 and Theorem 1.2 are proved in Section 8. Finally, Section 9 and Section 10 contain a proof of Theorem 1.3 and Theorem 1.4, respectively.

Notation.

In this paper, we use \mathbb{N} to denote the set of positive integers excluding zero. As usual, we let ω\omega be the first infinite ordinal, which can also be seen as the set of positive integers including zero.

Acknowledgments

We are grateful to Su Gao, Alexander Kechris, André Nies, and Sławomir Solecki for useful comments and remarks on a preliminary version of this manuscript.

2. Polishable subgroups

A Polish space is a second countable topological space whose topology is induced by a complete metric. A Polish group is a group in the category of Polish spaces, namely a Polish space that is endowed with a continuous group operation such that the function that maps each element to its inverse is also continuous (in fact, the latter requirement holds automatically; see the remark after [9, Corollary 9.15]). A subgroup HH of a Polish group GG is Polishable if it is Borel and there exists a Polish group topology on HH whose open sets are Borel in GG. Notice that such a Polish topology on HH, if it exists, it is unique by [9, Theorem 9.10]. In the following, we will regard HH as a Polish group with respect to its unique Polish group topology, which is in general finer than the subspace topology induced from GG. Equivalently, HH is a Polishable subgroup of GG if and only if there exists a Polish group H~\tilde{H} and a continuous group homomorphism φ:H~G\varphi:\tilde{H}\rightarrow G with image equal to HH. Noticing that one can assume without loss of generality that φ\varphi is an injection, the equivalence of the two definitions follows from [9, Theorem 9.10] and the fact that if f:XYf:X\rightarrow Y is an injective Borel function between standard Borel spaces, then f(A)f(A) is a Borel subset of YY and f|Af|_{A} is a Borel isomorphism between AA and f(A)f(A) [9, Theorem 15.1]. If GG is a Polish group and HH\ is a Polishable subgroup of GG, then GG is a Polish HH-space with respect to the left translation action of HH on GG [2, Section 2.2]. We will denote by EHGE_{H}^{G} the corresponding orbit equivalence relation. Recall that a Polish group GG is non-Archimedean if it admits a basis of neighborhoods of the identity consisting of open subgroups; see [6, Theorem 2.4.1] for equivalent characterizations.

Lemma 2.1.

Suppose that GG is a Polish group. Let (Gn)nω\left(G_{n}\right)_{n\in\omega} be a sequence of Polishable subgroups of GG. Then Gω:=nωGnG_{\omega}:=\bigcap_{n\in\omega}G_{n} is a Polishable subgroup of GG. If GnG_{n} is non-Archimedean for every nωn\in\omega, then GωG_{\omega} is non-Archimedean as well. If AGωA\subseteq G_{\omega} is such that AA is dense in GnG_{n} for every nωn\in\omega, then AA is dense in GωG_{\omega}.

Proof.

We have that GωG_{\omega} is the image of the Polish group

Z:={(xn)nωnωGn:nωxn=xn+1}nωGnZ:=\left\{\left(x_{n}\right)_{n\in\omega}\in\prod_{n\in\omega}G_{n}:\forall n\in\omega\text{, }x_{n}=x_{n+1}\right\}\subseteq\prod_{n\in\omega}G_{n}

under the continuous injective group homomorphism ZGZ\rightarrow G, (xn)nωx0\left(x_{n}\right)_{n\in\omega}\mapsto x_{0}. This shows that GωG_{\omega} is Polishable. If GnG_{n} is non-Archimedean for every nωn\in\omega, then ZZ is non-Archimedean, and hence GωG_{\omega} is non-Archimedean as well. By the above, the sets of the form WGωW\cap G_{\omega}, where WW is a neighborhood of the identity in GnG_{n} for some nωn\in\omega, form a basis of neighborhoods of the identity in GωG_{\omega}. Thus, if AA is dense in GnG_{n} for every nωn\in\omega, then AA is dense in GωG_{\omega}. ∎

3. Potential complexity

A complexity class Γ\Gamma is a function XΓ(X)X\mapsto\Gamma\left(X\right) that assigns to each Polish space XX a collection Γ(X)\Gamma\left(X\right) of Borel subsets, such that if X,YX,Y are Polish spaces and f:XYf:X\rightarrow Y is a continuous function, then f1(A)Γ(X)f^{-1}(A)\in\Gamma\left(X\right) for every AΓ(Y)A\in\Gamma\left(Y\right). For a complexity class Γ\Gamma, we let D(Γ)D\left(\Gamma\right) be the complexity class consisting of differences between sets in Γ\Gamma; see [9, Section 22.E] where it is denoted by D2(Γ)D_{2}\left(\Gamma\right). We let Γˇ\check{\Gamma} be the dual complexity class of Γ\Gamma, such that Γˇ(X)\check{\Gamma}\left(X\right) comprises the complements of the elements of Γ(X)\Gamma\left(X\right). We say that Γ\Gamma is self-dual if Γ=Γˇ\Gamma=\check{\Gamma}. If Γ\Gamma is a complexity class that is not self-dual, then we say that Γ\Gamma is the complexity class of AXA\subseteq X if AΓ(X)A\in\Gamma\left(X\right) and AΓˇ(X)A\notin\check{\Gamma}\left(X\right). We will be mainly interested in the complexity classes 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0}, 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0}, 𝚫α0\boldsymbol{\Delta}_{\alpha}^{0}, and D(𝚷α0)D(\boldsymbol{\Pi}_{\alpha}^{0}) for αω1\alpha\in\omega_{1}; see [9, Section 11.B].

If XX is a standard Borel space and EE is an equivalence relation on XX, then EE has potential complexity Γ\Gamma if there exists a Polish topology τ\tau on XX that induces the Borel structure of XX such that EΓ(τ×τ)E\in\Gamma\left(\tau\times\tau\right) [11]. This is equivalent to the assertion that there exists a Borel equivalence relation FF on a Polish space YY such that FΓ(Y×Y)F\in\Gamma\left(Y\times Y\right) and EE is Borel reducible to FF; see [6, Lemma 12.5.4]. The following result is essentially proved in [8, Section 5].

Proposition 3.1 (Hjorth–Kechris–Louveau).

Suppose that GG is a Polish group, and XX is a Polish GG-space. For xXx\in X, denote by [x]\left[x\right] the corresponding GG-orbit. Let Γ\Gamma be a complexity class, and assume that the orbit equivalence relation EGXE_{G}^{X} is potentially Γ\Gamma. Suppose that α\alpha is a countable ordinal.

  1. (1)

    If Γ\Gamma is the class 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0} for α2\alpha\geq 2, 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0} for α3\alpha\geq 3, or D(𝚷α0)D(\boldsymbol{\Pi}_{\alpha}^{0}) for α2\alpha\geq 2, then {xX:[x]Γ}\left\{x\in X:\left[x\right]\in\Gamma\right\} is comeager in XX.

  2. (2)

    If Γ\Gamma is the class Dˇ(𝚷α0)\check{D}(\boldsymbol{\Pi}_{\alpha}^{0}) for α3\alpha\geq 3, then {xX:[x]\{x\in X:\left[x\right] is either 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0} or 𝚺α0}\boldsymbol{\Sigma}_{\alpha}^{0}\} is comeager in XX.

Proof.

Fix a countable open basis {Ui:iω}\left\{U_{i}:i\in\omega\right\} of GG. Below we adopt the Vaught transform notation as in [6, Section 3.2]. By [9, Theorem 8.38], there exists a dense GδG_{\delta} set WXW\subseteq X such that EGX(W×W)Γ(W×W)E_{G}^{X}\cap\left(W\times W\right)\in\Gamma\left(W\times W\right). Notice that WW^{\ast} is also a dense GδG_{\delta} subset of XX. Fix xWWx\in W\cap W^{\ast}. Thus, we have that [x]WΓ(W)\left[x\right]\cap W\in\Gamma(W). If Γ=𝚺α0\Gamma=\boldsymbol{\Sigma}_{\alpha}^{0} for α3\alpha\geq 3, then [x]W=AW\left[x\right]\cap W=A\cap W for some A𝚺α0(X)A\in\boldsymbol{\Sigma}_{\alpha}^{0}\left(X\right). Then we have that [x]=([x]W)Δ=(AW)Δ\left[x\right]=\left(\left[x\right]\cap W\right)^{\Delta}=\left(A\cap W\right)^{\Delta} is 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0} in XX. If Γ=𝚷α0\Gamma=\boldsymbol{\Pi}_{\alpha}^{0} for α2\alpha\geq 2, then [x]W=BW\left[x\right]\cap W=B\cap W for some B𝚷α0(X)B\in\boldsymbol{\Pi}_{\alpha}^{0}\left(X\right). Then [x]=(W[x])=(BW)\left[x\right]=\left(W\cap\left[x\right]\right)^{\ast}=\left(B\cap W\right)^{\ast} is 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0} in XX. If Γ=D(𝚷α0)\Gamma=D(\boldsymbol{\Pi}_{\alpha}^{0}) for α2\alpha\geq 2, then W[x]=ABWW\cap\left[x\right]=A\cap B\cap W where A𝚺α0(X)A\in\boldsymbol{\Sigma}_{\alpha}^{0}(X) and B𝚷α0(X)B\in\boldsymbol{\Pi}_{\alpha}^{0}(X). Thus, [x]=AΔ(BW)D(𝚷α0)\left[x\right]=A^{\Delta}\cap\left(B\cap W\right)^{\ast}\in D(\boldsymbol{\Pi}_{\alpha}^{0}).

If Γ=Dˇ(𝚷α0)\Gamma=\check{D}(\boldsymbol{\Pi}_{\alpha}^{0}) for α3\alpha\geq 3, then W[x]=(AW)(BW)W\cap\left[x\right]=\left(A\cap W\right)\cup\left(B\cap W\right), where A𝚺α0(X)A\in\boldsymbol{\Sigma}_{\alpha}^{0}\left(X\right) and B𝚷α0(X)B\in\boldsymbol{\Pi}_{\alpha}^{0}\left(X\right). Thus, [x]=(AW)Δ\left[x\right]=\left(A\cap W\right)^{\Delta} or [x]=(BW)\left[x\right]=\left(B\cap W\right)^{\ast}. Hence, either [x]\left[x\right] is 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0} or [x]\left[x\right] is 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0}. If Γ=Dˇ(𝚷20)\Gamma=\check{D}(\boldsymbol{\Pi}_{2}^{0}), then AWA\cap W as above is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) in XX, and hence [x]\left[x\right] is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) in XX. ∎

A similar proof as Proposition 3.1 gives the following.

Lemma 3.2.

Suppose that GG is a Polish group, and HH is a Polishable subgroup of GG. Let α\alpha be a countable ordinal. If HH is Dˇ(𝚷α0)\check{D}(\boldsymbol{\Pi}_{\alpha}^{0}), then HH is either 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0} or 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0}.

Proof.

Adopt the notation of the Vaught transform with respect to the left translation action of HH on GG. We have that H=ABH=A\cup B where AA is 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0} and BB is 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0}. If xHx\in H, then we have that either xAΔx\in A^{\Delta} or xBx\in B^{\ast}. Since AΔA^{\Delta} and BB^{\ast} are HH-invariant, we have that either HAΔH\subseteq A^{\Delta} or HBH\subseteq B^{\ast}. Since AΔA^{\Delta} and BB^{\ast} are contained in HH, we have that either H=AΔH=A^{\Delta} or H=BH=B^{\ast}. This concludes the proof. ∎

Applying Proposition 3.1 to the left translation action associated with a Polishable subgroup of a Polish group, we obtain Items (1) and (3) of the following result. The proof of Item (2) is postponed to Section 6.

Theorem 3.3.

Suppose that GG is a Polish group, and HH is a Polishable subgroup of GG. Denote by EHGE_{H}^{G} the corresponding coset equivalence relation.

  1. (1)

    EHGE_{H}^{G} is potentially 𝚷20\boldsymbol{\Pi}_{2}^{0} if and only if HH is closed GG.

  2. (2)

    EHGE_{H}^{G} is potentially 𝚺20\boldsymbol{\Sigma}_{2}^{0} if and only if HH is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) in GG.

  3. (3)

    Let Γ\Gamma be one of the following complexity classes: 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0} for α2\alpha\neq 2, 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0}, and D(𝚷α0)D\left(\boldsymbol{\Pi}_{\alpha}^{0}\right). Then EHGE_{H}^{G} is potentially Γ\Gamma in GG if and only if HH is Γ\Gamma in GG.

Proof.

(1): Suppose that EHGE_{H}^{G} is potentially 𝚷20\boldsymbol{\Pi}_{2}^{0}. By [6, Lemma 12.5.3] we have that EHGE_{H}^{G} is smooth. Thus, HH is closed by [17, page 574].

(2): The forward implication is a particular instance of Proposition 3.1, while the converse implication follows from Lemma 6.5 in Section 6.

(3): Only the forward implication requires a proof. If Γ=𝚺10\Gamma=\boldsymbol{\Sigma}_{1}^{0} then EHGE_{H}^{G} has countably many classes by [6, Lemma 12.5.2]. Thus, HH has countable index in GG, and hence it is nonmeger. Therefore, HH is open by [6, Theorem 2.3.2]. If Γ\Gamma is 𝚷10\boldsymbol{\Pi}_{1}^{0} or 𝚷20\boldsymbol{\Pi}_{2}^{0} or D(𝚷10)D\left(\boldsymbol{\Pi}_{1}^{0}\right), then H𝚷10(G)Γ(G)H\in\boldsymbol{\Pi}_{1}^{0}\left(G\right)\subseteq\Gamma\left(G\right) by Part (1). If Γ\Gamma is 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0} or 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0} for α3\alpha\geq 3, or D(𝚷α0)D\left(\boldsymbol{\Pi}_{\alpha}^{0}\right) for α2\alpha\geq 2, the conclusion follows from Proposition 3.1. ∎

We now recall some results concerning the possible complexity classes of Polishable subgroups. The following proposition is a reformulation of [5, Corollary 3.4].

Proposition 3.4 (Farah–Solecki).

Suppose that GG is a Polish group, and HH is a Polishable subgroup of GG. If λ<ω1\lambda<\omega_{1} is either zero or a limit ordinal, and HH is 𝚷1+λ+10\boldsymbol{\Pi}_{1+\lambda+1}^{0} in GG, then HH is 𝚷1+λ0\boldsymbol{\Pi}_{1+\lambda}^{0} in GG.

The following proposition is a consequence of [8, Theorem 4.1] and Proposition 3.1.

Proposition 3.5 (Hjorth–Kechris–Louveau).

Suppose that GG is a Polish group, and HH is a non-Archimedean Polishable subgroup of GG. Suppose that λ<ω1\lambda<\omega_{1} is either zero or limit. If HH is 𝚺1+λ+20\boldsymbol{\Sigma}_{1+\lambda+2}^{0}, then HH is 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0}.

4. Solecki subgroups

Suppose that GG is a Polish group, and HH is a Polishable subgroup of GG. Then HH admits a canonical approximation by Polishable subgroups indexed by countable ordinals. As these were originally described by Solecki in [16], we call them Solecki subgroups of GG associated with HH. They have also been considered in [17, 5].

Lemma 2.3 from [16, Lemma 2.3] implies that GG has a smallest 𝚷30\boldsymbol{\Pi}_{3}^{0} Polishable subgroup containing HH, which we denote by s1H(G)s_{1}^{H}(G). One can explicitly describe s1H(G)s_{1}^{H}(G) as the 𝚷30\boldsymbol{\Pi}_{3}^{0} subgroup of GG defined by

Vz0,z1Hz0V¯GV¯Gz1\bigcap_{V}\bigcup_{z_{0},z_{1}\in H}z_{0}\overline{V}^{G}\cap\overline{V}^{G}z_{1}

where VV ranges among the open neighborhoods of the identity in HH, and V¯G\overline{V}^{G} is the closure of VV inside of GG. It is proved in [16, Lemma 2.3] that s1H(G)s_{1}^{H}(G) satisfies the following properties—see also [19, Lemma 4.5] and [5, Section 3]:

  • HH is dense in s1H(G)s_{1}^{H}(G);

  • a neighborhood basis of xs1H(G)x\in s_{1}^{H}(G) consists of sets of the form Wx¯Gs1H(G)\overline{Wx}^{G}\cap s_{1}^{H}(G) where WW is an open neighborhood of the identity in HH;

  • if AGA\subseteq G is 𝚷30\boldsymbol{\Pi}_{3}^{0} and contains HH, then As1H(G)A\cap s_{1}^{H}(G) is comeager in the Polish group topology of s1H(G)s_{1}^{H}(G).

Lemma 4.1.

Suppose that GG is a Polish group, and HH is a non-Archimedean Polishable subgroup of GG. Then a neighborhood basis of the identity in s1H(G)s_{1}^{H}(G) consists of the sets of the form W¯Gs1H(G)\overline{W}^{G}\cap s_{1}^{H}(G) where WW is an open subgroup of HH. In particular, s1H(G)s_{1}^{H}(G) is non-Archimedean.

Proof.

Since HH is non-Archimedean, the first assertion follows from the remarks above. If WW is an open subgroup of HH, then W¯Gs1H(G)\overline{W}^{G}\cap s_{1}^{H}(G) is a subgroup of s1H(G)s_{1}^{H}(G) with nonempty interior, whence it is an open subgroup by Pettis’ Theorem [15, Corollary 3.1]. Therefore, the second assertion follows from the first one. ∎

A similar argument as in the proof of [16, Lemma 2.3] gives the following.

Lemma 4.2.

Suppose tha GG is a Polish group, and NN is a Polishable subgroup of GG. Let HH be a Polishable subgroup of GG such that:

  1. (1)

    NHN\subseteq H and NN is dense in the Polish group topology of HH;

  2. (2)

    for every open neighborhood VV of the identity in NN, V¯GH\overline{V}^{G}\cap H contains an open neighborhood of the identity in HH.

If AGA\subseteq G is 𝚷30\boldsymbol{\Pi}_{3}^{0} and contains NN, then AHA\cap H is comeager in HH. In particular, Hs1N(G)H\subseteq s_{1}^{N}(G). If HH is furthermore 𝚷30\boldsymbol{\Pi}_{3}^{0}, then H=s1N(G)H=s_{1}^{N}(G).

Proof.

It suffices to consider the case when AA is 𝚺20\boldsymbol{\Sigma}_{2}^{0}. In this case, there exist closed subsets LkL_{k} of GG for kωk\in\omega such that A=kωLkA=\bigcup_{k\in\omega}L_{k}. Suppose that UHU\subseteq H is a nonempty open set. Since NN is dense in HH, UNU\cap N is a nonempty open subset of NN. By the Baire Category Theorem, there exists k0ωk_{0}\in\omega such that Lk0UNL_{k_{0}}\cap U\cap N is not meager in NN. Thus, there exist xNx\in N and an open neighborhood VV of the identity in NN such that

VxLk0UN.Vx\subseteq L_{k_{0}}\cap U\cap N\text{.}

Since Lk0L_{k_{0}} is closed in GG, we have that Vx¯GLk0\overline{Vx}^{G}\subseteq L_{k_{0}}. By (2), there is an open neighborhood WW of the identity in HH such that WxVx¯GLk0Wx\subseteq\overline{Vx}^{G}\subseteq L_{k_{0}}. This shows that xUx\in U is in the interior of Lk0HAHL_{k_{0}}\cap H\subseteq A\cap H. Since this holds for every nonempty open subset of HH, we have that AHA\cap H contains a dense open subset of HH, and hence it is comeager in HH. This concludes the proof. ∎

The sequence of Solecki subgroups sαH(G)s_{\alpha}^{H}(G) for α<ω1\alpha<\omega_{1} of GG associated with HH is defined recursively by setting:

  • s0H(G)=H¯Gs_{0}^{H}(G)=\overline{H}^{G};

  • sα+1H(G)=s1H(sαH(G))s_{\alpha+1}^{H}(G)=s_{1}^{H}\left(s_{\alpha}^{H}(G)\right) for α<ω1\alpha<\omega_{1};

  • sλH(G)=β<λsβH(G)s_{\lambda}^{H}(G)=\bigcap_{\beta<\lambda}s_{\beta}^{H}(G) for a limit ordinal λ<ω1\lambda<\omega_{1}.

Using Lemma 2.1 at the limit stage, one can prove by induction on α<ω1\alpha<\omega_{1} that sαH(G)s_{\alpha}^{H}(G) is a Polishable subgroup of GG, and HH is dense in sαH(G)s_{\alpha}^{H}(G). Furthermore, by Lemma 4.1, if HH is non-Archimedean, then sαH(G)s_{\alpha}^{H}(G) is non-Archimedean for every 1α<ω11\leq\alpha<\omega_{1}. It is proved in [16, Theorem 2.1] that there exists α<ω1\alpha<\omega_{1} such that sαH(G)=Hs_{\alpha}^{H}(G)=H. We call the least countable ordinal α\alpha such that sαH(G)=Hs_{\alpha}^{H}(G)=H the Solecki rank of HH in GG.

One can define the Polish groups sαH(G)s_{\alpha}^{H}(G) solely in terms of HH endowed with the subspace topology inherited from GG. Indeed, s0H(G)s_{0}^{H}(G) can be seen as the completion of HH with respect to a suitable metric that induces the subspace topology inherited from GG; see [16, Section 2.1]. Using Lemma 4.2 one can describe the Solecki subgroups of products, as follows.

Lemma 4.3.

Suppose that, for every nn\in\mathbb{N}, GnG_{n} is a Polish group, and NnN_{n} is a Polishable subgroup. Define G=nωGnG=\prod_{n\in\omega}G_{n} and N=nωNnN=\prod_{n\in\omega}N_{n}. Then we have that

sγH(G)=nωsγNn(Gn)s_{\gamma}^{H}(G)=\prod_{n\in\omega}s_{\gamma}^{N_{n}}\left(G_{n}\right)

for every γ<ω1\gamma<\omega_{1}.

Proof.

It suffices to consider the case when γ=1\gamma=1. In this case, set

Hn:=s1Nn(Gn)H_{n}:=s_{1}^{N_{n}}\left(G_{n}\right)

for nωn\in\omega, and

H:=nωHn.H:=\prod_{n\in\omega}H_{n}\text{.}

Then we have that HH is a 𝚷30\boldsymbol{\Pi}_{3}^{0} Polishable subgroup of GG, NHN\subseteq H, and NN is dense in HH. If VV is an open neighborhood of the identity in NN, then there exist nωn\in\omega and open neighborhoods ViV_{i} of the identity in NiN_{i} for i<ni<n such that VV contains

i<nVi={xN:i<nxiVi}.\prod_{i<n}V_{i}=\{x\in N:\forall i<n\text{, }x_{i}\in V_{i}\}\text{.}

For i<ni<n, we have that V¯iGiHi\overline{V}_{i}^{G_{i}}\cap H_{i} contains an open neighborhood WiW_{i} of the identity in HiH_{i}. Therefore, we have that V¯GH\overline{V}^{G}\cap H contains

i<nWi={xH:i<nxiWi},\prod_{i<n}W_{i}=\{x\in H:\forall i<n\text{, }x_{i}\in W_{i}\}\text{,}

which is an open neighborhood of the identity in HH. The conclusion thus follows from Lemma 4.2. ∎

5. Complexity of Solecki subgroups

Suppose that GG is a Polish group, and HH is a Polishable subgroup of GG. For a complexity class Γ\Gamma, we define Γ(G)|H\Gamma(G)|_{H} to be the collection of sets of the form AHA\cap H for AΓ(G)A\in\Gamma(G). The following results are essentially established in [5]. In the statements and proofs, we adopt the Vaught transform notation in reference to the action of HH on GG by left translation; see [6, Section 3.2].

Lemma 5.1.

Suppose that GG is a Polish group, HH is a Polishable subgroup of GG, and α,β<ω1\alpha,\beta<\omega_{1} are ordinals. Then

𝚺1+β0(sαH(G))𝚺1+α+β0(G)|sαH(G)\boldsymbol{\Sigma}_{1+\beta}^{0}\left(s_{\alpha}^{H}(G)\right)\subseteq\boldsymbol{\Sigma}_{1+\alpha+\beta}^{0}(G)|_{s_{\alpha}^{H}(G)}

and

𝚷1+β0(sαH(G))𝚷1+α+β0(G)|sαH(G).\boldsymbol{\Pi}_{1+\beta}^{0}\left(s_{\alpha}^{H}(G)\right)\subseteq\boldsymbol{\Pi}_{1+\alpha+\beta}^{0}(G)|_{s_{\alpha}^{H}(G)}\text{.}
Proof.

It is proved in [5, Theorem 3.1] by induction on α\alpha that 𝚺10(sαH(G))𝚺1+α0(G)|sαH(G)\boldsymbol{\Sigma}_{1}^{0}\left(s_{\alpha}^{H}(G)\right)\subseteq\boldsymbol{\Sigma}_{1+\alpha}^{0}(G)|_{s_{\alpha}^{H}(G)}. By taking complements, we have that 𝚷10(sαH(G))𝚷1+α0(G)|sαH(G)\boldsymbol{\Pi}_{1}^{0}\left(s_{\alpha}^{H}(G)\right)\subseteq\boldsymbol{\Pi}_{1+\alpha}^{0}(G)|_{s_{\alpha}^{H}(G)}. This is the case β=0\beta=0 of the statement above. The rest follows by induction on β\beta. ∎

Lemma 5.2.

Suppose that GG\ is a Polish group, HH\ is a Polishable subgroup of GG, α,β<ω1\alpha,\beta<\omega_{1}, and UHU\subseteq H is open in HH. If A𝚺1+α+β0(G)A\in\boldsymbol{\Sigma}_{1+\alpha+\beta}^{0}(G) and B𝚷1+α+β0(G)B\in\boldsymbol{\Pi}_{1+\alpha+\beta}^{0}(G), then AΔUsαH(G)𝚺1+β0(sαH(G))A^{\Delta U}\cap s_{\alpha}^{H}(G)\in\boldsymbol{\Sigma}_{1+\beta}^{0}\left(s_{\alpha}^{H}(G)\right), and BUsαH(G)𝚷1+β0(sαH(G))B^{\ast U}\cap s_{\alpha}^{H}(G)\in\boldsymbol{\Pi}_{1+\beta}^{0}\left(s_{\alpha}^{H}(G)\right).

Proof.

When β=0\beta=0, the assertion about AA is the content of Claim 3.3 in the proof of [5, Theorem 3.1]. The assertion about BB follows by taking complements. This concludes the proof when β=0\beta=0. The case of an arbitrary β\beta is established by induction on β\beta using the properties of the Vaught transform; see [6, Proposition 3.2.5]. ∎

Corollary 5.3.

Suppose that GG\ is a Polish group, HH is a Polishable subgroup of GG, and α,β<ω1\alpha,\beta<\omega_{1}. Let LL be a Polishable subgroup of GG containing HH. If L𝚺1+α+β0(G)L\in\boldsymbol{\Sigma}_{1+\alpha+\beta}^{0}(G), then LsαH(G)𝚺1+β0(sαH(G))L\cap s_{\alpha}^{H}(G)\in\boldsymbol{\Sigma}_{1+\beta}^{0}\left(s_{\alpha}^{H}(G)\right). If L𝚷1+α+β0(G)L\in\boldsymbol{\Pi}_{1+\alpha+\beta}^{0}(G), then LsαH(G)𝚷1+β0(sαH(G))L\cap s_{\alpha}^{H}(G)\in\boldsymbol{\Pi}_{1+\beta}^{0}\left(s_{\alpha}^{H}(G)\right). If LD(𝚷1+α+β0)(G)L\in D(\boldsymbol{\Pi}_{1+\alpha+\beta}^{0})\left(G\right), then LsαH(G)D(𝚷1+β0)(sαH(G))L\cap s_{\alpha}^{H}(G)\in D(\boldsymbol{\Pi}_{1+\beta}^{0})\left(s_{\alpha}^{H}(G)\right).

Proof.

Observe that L=L=LΔL=L^{\ast}=L^{\Delta}. Thus, the first two assertions follow immediately from Lemma 5.2. If L=ABL=A\cap B where AA is 𝚺1+β+10\boldsymbol{\Sigma}_{1+\beta+1}^{0} in GG and BB is 𝚷1+β+10\boldsymbol{\Pi}_{1+\beta+1}^{0} in GG, then we have that LsαH(G)=AΔBsαH(G)L\cap s_{\alpha}^{H}(G)=A^{\Delta}\cap B^{\ast}\cap s_{\alpha}^{H}(G) where AΔsαH(G)𝚺1+β0(sαH(G))A^{\Delta}\cap s_{\alpha}^{H}(G)\in\boldsymbol{\Sigma}_{1+\beta}^{0}\left(s_{\alpha}^{H}(G)\right) and BsαH(G)𝚷1+β0(sαH(G))B\cap s_{\alpha}^{H}(G)\in\boldsymbol{\Pi}_{1+\beta}^{0}\left(s_{\alpha}^{H}(G)\right) by Lemma 5.2. Hence, LsαH(G)D(𝚷1+β0)(sαH(G))L\cap s_{\alpha}^{H}(G)\in D(\boldsymbol{\Pi}_{1+\beta}^{0})\left(s_{\alpha}^{H}(G)\right). ∎

Recall that, by Proposition 3.4, if α\alpha is either zero or a countable limit ordinal, and HH is a 𝚷1+α+10\boldsymbol{\Pi}_{1+\alpha+1}^{0} Polishable subgroup of a Polish group, then HH is 𝚷1+α0\boldsymbol{\Pi}_{1+\alpha}^{0}.

Theorem 5.4.

Suppose that GG is a Polish group, HH is a Polishable subgroup of HH, and α<ω1\alpha<\omega_{1}. Then sαH(G)s_{\alpha}^{H}(G) is the smallest 𝚷1+α+10\boldsymbol{\Pi}_{1+\alpha+1}^{0} Polishable subgroup of GG containing HH.

Proof.

It is established in the proof of [5, Theorem 3.1] that sα(G)s_{\alpha}(G) is a 𝚷1+α+10\boldsymbol{\Pi}_{1+\alpha+1}^{0} Polishable subgroup of GG. We now prove the minimality assertion by induction on α\alpha. For α=0\alpha=0 this follows from the fact that s0H(G)=H¯Gs_{0}^{H}(G)=\overline{H}^{G}. Suppose that the conclusion holds for α\alpha. We now prove that it holds for α+1\alpha+1. Let LL be a 𝚷1+α+20\boldsymbol{\Pi}_{1+\alpha+2}^{0} Polishable subgroup of GG containing HH. Thus, LsαH(G)L\cap s_{\alpha}^{H}(G) is a 𝚷1+α+20\boldsymbol{\Pi}_{1+\alpha+2}^{0} Polishable subgroup of sαH(G)s_{\alpha}^{H}(G). Then by Corollary 5.3 we have that LsαH(G)𝚷30(sα(H))L\cap s_{\alpha}^{H}(G)\in\boldsymbol{\Pi}_{3}^{0}\left(s_{\alpha}(H)\right). As sα+1H(G)=s1H(sαH(G))s_{\alpha+1}^{H}(G)=s_{1}^{H}\left(s_{\alpha}^{H}(G)\right) is the smallest 𝚷30(sαH(G))\boldsymbol{\Pi}_{3}^{0}\left(s_{\alpha}^{H}(G)\right) Polishable subgroup of sαH(G)s_{\alpha}^{H}(G), this implies that sα+1H(G)LsαH(G)Ls_{\alpha+1}^{H}(G)\subseteq L\cap s_{\alpha}^{H}(G)\subseteq L.

Suppose that α\alpha is a limit ordinal and the conclusion holds for every β<α\beta<\alpha. Fix an increasing sequence (αn)\left(\alpha_{n}\right) in α\alpha such that α=supnαn\alpha=\mathrm{sup}_{n}\alpha_{n}. Suppose that LL is a 𝚷1+α0\boldsymbol{\Pi}_{1+\alpha}^{0} Polishable subgroup of GG containing HH. Since LL is 𝚷1+α0\boldsymbol{\Pi}_{1+\alpha}^{0} in GG, we can write L=nωAnL=\bigcap_{n\in\omega}A_{n} where, for every nωn\in\omega, An𝚷1+αn0(G)A_{n}\in\boldsymbol{\Pi}_{1+\alpha_{n}}^{0}(G). Then by Lemma 5.2 we have that

AnsαnH(G)𝚷10(sαn(G)).A_{n}^{\ast}\cap s_{\alpha_{n}}^{H}(G)\in\boldsymbol{\Pi}_{1}^{0}\left(s_{\alpha_{n}}(G)\right)\text{.}

Since HLAnH\subseteq L\subseteq A_{n} we have that HAnsαnH(G)H\subseteq A_{n}^{\ast}\cap s_{\alpha_{n}}^{H}(G). Since HH is dense in sαn(H)s_{\alpha_{n}}(H), this implies that sαnH(G)Ans_{\alpha_{n}}^{H}(G)\subseteq A_{n}^{\ast}. Therefore, we have that

sαH(G)=nωsαnH(G)nωAn=L=L.s_{\alpha}^{H}(G)=\bigcap_{n\in\omega}s_{\alpha_{n}}^{H}(G)\subseteq\bigcap_{n\in\omega}A_{n}^{\ast}=L^{\ast}=L\text{.}

This shows that sαH(G)Ls_{\alpha}^{H}(G)\subseteq L, concluding the proof. ∎

Lemma 5.5.

Suppose that GG is a Polish group, HH is a Polishable subgroup of GG, and α<ω1\alpha<\omega_{1}. Let LL be a Polishable subgroup of sαH(G)s_{\alpha}^{H}(G).

  1. (1)

    If L𝚷30(sαH(G))L\in\boldsymbol{\Pi}_{3}^{0}\left(s_{\alpha}^{H}(G)\right), then L𝚷1+α+20(G)L\in\boldsymbol{\Pi}_{1+\alpha+2}^{0}(G).

  2. (2)

    If LD(𝚷20)(sαH(G))L\in D(\boldsymbol{\Pi}_{2}^{0})(s_{\alpha}^{H}(G)), then LD(𝚷1+α+10)(G)L\in D(\boldsymbol{\Pi}_{1+\alpha+1}^{0})(G).

  3. (3)

    If L𝚺20(sαH(G))L\in\boldsymbol{\Sigma}_{2}^{0}\left(s_{\alpha}^{H}(G)\right) and α\alpha is either zero or limit, then L𝚺1+α+10(G)L\in\boldsymbol{\Sigma}_{1+\alpha+1}^{0}(G).

Proof.

By Lemma 5.1 we have that

𝚷30(sαH(G))𝚷1+α+20(G)|sαH(G)\boldsymbol{\Pi}_{3}^{0}\left(s_{\alpha}^{H}(G)\right)\subseteq\boldsymbol{\Pi}_{1+\alpha+2}^{0}(G)|_{s_{\alpha}^{H}(G)}

and

D(𝚷20)(sαH(G))D(𝚷1+α+10)(G)|sαH(G).D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\alpha}^{H}(G)\right)\subseteq D\left(\boldsymbol{\Pi}_{1+\alpha+1}^{0}\right)(G)|_{s_{\alpha}^{H}(G)}\text{.}

Furthermore, sαH(G)𝚷1+α+10(G)s_{\alpha}^{H}(G)\in\boldsymbol{\Pi}_{1+\alpha+1}^{0}(G) by Theorem 5.4. Therefore, we have that

𝚷30(sαH(G))𝚷1+α+20(G)|sαH(G)𝚷1+α+20(G)\boldsymbol{\Pi}_{3}^{0}\left(s_{\alpha}^{H}(G)\right)\subseteq\boldsymbol{\Pi}_{1+\alpha+2}^{0}(G)|_{s_{\alpha}^{H}(G)}\subseteq\boldsymbol{\Pi}_{1+\alpha+2}^{0}(G)

and

D(𝚷20)(sαH(G))D(𝚷1+α+10)(G)|sαH(G)D(𝚷1+α+10)(G).D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\alpha}^{H}(G)\right)\subseteq D\left(\boldsymbol{\Pi}_{1+\alpha+1}^{0}\right)(G)|_{s_{\alpha}^{H}(G)}\subseteq D\left(\boldsymbol{\Pi}_{1+\alpha+1}^{0}\right)(G)\text{.}

This concludes the proof of (1) and (2).

When α\alpha is either zero or limit, we have by Theorem 5.4 and Proposition 3.4 that sαH(G)𝚷1+α0(G)𝚺1+α+10(G)s_{\alpha}^{H}(G)\in\boldsymbol{\Pi}_{1+\alpha}^{0}(G)\subseteq\boldsymbol{\Sigma}_{1+\alpha+1}^{0}(G). Therefore, in this case we have that

𝚺20(sαH(G))𝚺1+α+10(G)|sαH(G)𝚺1+α+10(G).\boldsymbol{\Sigma}_{2}^{0}\left(s_{\alpha}^{H}(G)\right)\subseteq\boldsymbol{\Sigma}_{1+\alpha+1}^{0}(G)|_{s_{\alpha}^{H}(G)}\subseteq\boldsymbol{\Sigma}_{1+\alpha+1}^{0}(G)\text{.}

This concludes the proof of (3). ∎

Lemma 5.6.

Suppose that, for every kωk\in\omega, GkG_{k} is a Polish group and HkH_{k} is a Polishable subgroup of GkG_{k}. Define G=kωGkG=\prod_{k\in\omega}G_{k} and H=kωHkH=\prod_{k\in\omega}H_{k}. Assume that, for every kωk\in\omega, HkH_{k} is 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0} in GkG_{k}, and for every β<α\beta<\alpha there exist infinitely many kωk\in\omega such that HkH_{k} is not 𝚷β0\boldsymbol{\Pi}_{\beta}^{0} in GkG_{k}. Then 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0} is the complexity class of HH in GG.

Proof.

Write H=kωHkkωGkH=\prod_{k\in\omega}H_{k}\subseteq\prod_{k\in\omega}G_{k}. Clearly, HH is 𝚷α0\boldsymbol{\Pi}_{\alpha}^{0}. By [9, Theorem 22.10], for every kωk\in\omega and β<α\beta<\alpha such that HkH_{k} is not 𝚷β0\boldsymbol{\Pi}_{\beta}^{0}, HkH_{k} is 𝚺β0\boldsymbol{\Sigma}_{\beta}^{0}-hard [9, Definition 22.9]. Therefore, HH is 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0}-hard, and hence HH is not 𝚺α0\boldsymbol{\Sigma}_{\alpha}^{0} by [9, Theorem 22.10] again. ∎

Lemma 5.7.

Suppose that, for every kωk\in\omega, GkG_{k} is a Polish group, HkH_{k} is a Polishable subgroup of GkG_{k}, and α<ω1\alpha<\omega_{1}. Define G=kωGkG=\prod_{k\in\omega}G_{k} and H=kωHkH=\prod_{k\in\omega}H_{k}. If HkH_{k} has Solecki rank α\alpha in GkG_{k} for every kωk\in\omega, then 𝚷1+α+10\boldsymbol{\Pi}_{1+\alpha+1}^{0} is the complexity class of HH in GG if α\alpha is a successor ordinal, and 𝚷1+α0\boldsymbol{\Pi}_{1+\alpha}^{0} is the complexity class of HH in GG if α\alpha is either zero or a limit ordinal.

Proof.

Define λ=1+α+1\lambda=1+\alpha+1 if α\alpha is a successor ordinal, and λ=1+α\lambda=1+\alpha if α\alpha is either zero or a limit ordinal. By Lemma 4.2, for every kωk\in\omega, HkH_{k} is 𝚷λ0\boldsymbol{\Pi}_{\lambda}^{0} but not 𝚷β0\boldsymbol{\Pi}_{\beta}^{0} for β<λ\beta<\lambda in GkG_{k}. Therefore, by Lemma 5.6, 𝚷λ0\boldsymbol{\Pi}_{\lambda}^{0} is the complexity of HH in GG. ∎

6. Complexity of Polishable subgroups

The goal of this section is to establish the following theorem, characterizing the possible values of the complexity class of a Polishable subgroup of a Polish group.

Theorem 6.1.

Suppose that GG is a Polish group, and HH is a Polishable subgroup of GG. Let α=λ+n\alpha=\lambda+n be the Solecki rank of HH in GG, where λ<ω1\lambda<\omega_{1} is either zero or a limit ordinal and n<ωn<\omega.

  1. (1)

    Suppose that n=0n=0. Then 𝚷1+λ0\boldsymbol{\Pi}_{1+\lambda}^{0} is the complexity class of HH in GG.

  2. (2)

    Suppose that n1n\geq 1. Then:

    1. (a)

      if H𝚷30(sλ+n1H(G))H\in\boldsymbol{\Pi}_{3}^{0}\left(s_{\lambda+n-1}^{H}(G)\right) and HD(𝚷20)(sλ+n1H(G))H\notin D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\lambda+n-1}^{H}(G)\right), then 𝚷1+λ+n+10\boldsymbol{\Pi}_{1+\lambda+n+1}^{0} is the complexity class of HH in GG;

    2. (b)

      if n2n\geq 2 and HD(𝚷20)(sλ+n1H(G))H\in D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\lambda+n-1}^{H}(G)\right), then D(𝚷1+λ+n0)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0}) is the complexity class of HH in GG;

    3. (c)

      if n=1n=1, HD(𝚷20)(sλH(G))H\in D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\lambda}^{H}(G)\right), and H𝚺20(sλH(G))H\notin\boldsymbol{\Sigma}_{2}^{0}\left(s_{\lambda}^{H}(G)\right), then D(𝚷1+λ+10)D\left(\boldsymbol{\Pi}_{1+\lambda+1}^{0}\right) is the complexity class of HH in GG;

    4. (d)

      if n=1n=1 and H𝚺20(sλH(G))H\in\boldsymbol{\Sigma}_{2}^{0}\left(s_{\lambda}^{H}(G)\right), then 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0} is the complexity class of HH in GG.

Furthermore, if HH is non-Archimedean then the case (2c) is excluded.

Theorem 1.1 and Theorem 1.2 are immediate consequences of Theorem 6.1. We will obtain Theorem 6.1 as a consequence of a number of complexity reduction lemmas. We fix a Polish group GG and a Polishable subgroup HH of GG. We adopt the Vaught transform notation, in reference to the left translation action of HH on GG.

Lemma 6.2.

If HH is 𝚫30\boldsymbol{\Delta}_{3}^{0}, then HH is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}).

Proof.

Since HH is 𝚷30\boldsymbol{\Pi}_{3}^{0} in GG, we have that H=s1H(G)H=s_{1}^{H}(G). Thus, HH has a countable basis {Vn:nω}\left\{V_{n}:n\in\omega\right\} of neighborhoods of the identity such that V¯nGH=Vn\overline{V}_{n}^{G}\cap H=V_{n} for every nωn\in\omega. Indeed, if {Wn:nω}\left\{W_{n}:n\in\omega\right\} is a countable basis of open neighborhood of the identity in HH, then {W¯nGH:nω}\{\overline{W}_{n}^{G}\cap H:n\in\omega\} is a countable basis of neighborhoods of the identity in HH. If Vn=W¯nGHV_{n}=\overline{W}_{n}^{G}\cap H, then we have that WnVnW¯nGW_{n}\subseteq V_{n}\subseteq\overline{W}_{n}^{G} and hence V¯nG=W¯nG\overline{V}_{n}^{G}=\overline{W}_{n}^{G} and Vn=V¯nGHV_{n}=\overline{V}_{n}^{G}\cap H.

Let also {U:ω}\left\{U_{\ell}:\ell\in\omega\right\} be a countable basis for the Polish group topology of HH. Since HH is 𝚺30\boldsymbol{\Sigma}_{3}^{0}, we can write H=kωFkH=\bigcup_{k\in\omega}F_{k} where FkF_{k} is 𝚷20\boldsymbol{\Pi}_{2}^{0} in GG. Thus, we have that H=k,ωFkUH=\bigcup_{k,\ell\in\omega}F_{k}^{\ast U_{\ell}} where, by Lemma 5.2, FkUF_{k}^{\ast U_{\ell}} is closed in HH and 𝚷20\boldsymbol{\Pi}_{2}^{0} in GG. Hence, without loss of generality we can assume that FkF_{k} is closed in HH for every kωk\in\omega. By the Baire Category Theorem, we can assume without loss of generality that V0F0V_{0}\subseteq F_{0}. Fix a countable dense subset {zm:mω}\left\{z_{m}:m\in\omega\right\} of HH. Since H=s1H(G)H=s_{1}^{H}(G), we have that, for xGx\in G, xHx\in H if and only if for every kωk\in\omega there exist m0,m1ωm_{0},m_{1}\in\omega such that xzm0V¯kGxz_{m_{0}}\in\overline{V}_{k}^{G} and zm1xV¯kGz_{m_{1}}x\in\overline{V}_{k}^{G}.

We claim that, for xGx\in G, xHx\in H if and only (1) there exists mωm\in\omega such that xzmV¯0Gxz_{m}\in\overline{V}_{0}^{G}, and (2) for all mωm\in\omega, if xzmV¯0Gxz_{m}\in\overline{V}_{0}^{G}, then xzmF0xz_{m}\in F_{0}. This will witness that HH is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) in GG.

Indeed, since {zm:mω}\left\{z_{m}:m\in\omega\right\} is dense in HH, if xHx\in H, then we have that there exists mωm\in\omega such that xzmV0V¯0Gxz_{m}\in V_{0}\subseteq\overline{V}_{0}^{G}. Furthermore, if mωm\in\omega is such that xzmV¯0Gxz_{m}\in\overline{V}_{0}^{G}, then we have xzmV¯0GH=V0F0xz_{m}\subseteq\overline{V}_{0}^{G}\cap H=V_{0}\subseteq F_{0}. Conversely suppose that there exists m0ωm_{0}\in\omega such that xzm0V¯0Gxz_{m_{0}}\in\overline{V}_{0}^{G}, and for all mωm\in\omega, if xzmV¯0Gxz_{m}\in\overline{V}_{0}^{G} then xF0x\in F_{0}. Then we have that xzm0F0Hxz_{m_{0}}\in F_{0}\subseteq H and hence xHx\in H. ∎

Lemma 6.3.

If HH is 𝚷40\boldsymbol{\Pi}_{4}^{0} in GG, then HH has a countable basis of neighborhoods of the identity consisting of sets that are in 𝚷20(G)|H\boldsymbol{\Pi}_{2}^{0}\left(G\right)|_{H}.

Proof.

Define H~=s1H(G)\tilde{H}=s_{1}^{H}\left(G\right). By Theorem 5.4 we have that H=s2H(G)=s1H(H~)H=s_{2}^{H}\left(G\right)=s_{1}^{H}(\tilde{H}). Thus a neighborhood basis of the identity in HH consists of sets of the form Wx¯H~H\overline{Wx}^{\tilde{H}}\cap H where WW is an open neighborhood of the identity in HH. We have that, by Lemma 5.1,

Wx¯H~𝚷10(s1H(G))𝚷20(G)|s1H(G).\overline{Wx}^{\tilde{H}}\in\boldsymbol{\Pi}_{1}^{0}\left(s_{1}^{H}(G)\right)\subseteq\boldsymbol{\Pi}_{2}^{0}\left(G\right)|_{s_{1}^{H}\left(G\right)}\text{.}

Therefore,

Wx¯H~H𝚷20(G)|H.\overline{Wx}^{\tilde{H}}\cap H\in\boldsymbol{\Pi}_{2}^{0}\left(G\right)|_{H}\text{.}

This concludes the proof. ∎

Lemma 6.4.

Suppose that HH is 𝚺30\boldsymbol{\Sigma}_{3}^{0} in GG, and define H~=s1H(G)\tilde{H}=s_{1}^{H}(G). Then we have that:

  1. (1)

    H=s1H(H~)H=s_{1}^{H}(\tilde{H});

  2. (2)

    We can write HH as the union of an increasing sequence (Fk)kω\left(F_{k}\right)_{k\in\omega} such that FkF_{k} is 𝚷20\boldsymbol{\Pi}_{2}^{0} in GG and closed in H~\tilde{H} for every kωk\in\omega;

  3. (3)

    HH has a countable family of neighborhoods of the identity consisting of sets that are in 𝚷20(G)𝚷10(H~)\boldsymbol{\Pi}_{2}^{0}\left(G\right)\cap\boldsymbol{\Pi}_{1}^{0}(\tilde{H}).

Proof.

(1): This is a consequence of Theorem 5.4.

(2): We can write H=kωFkH=\bigcup_{k\in\omega}F_{k} where FkF_{k} is 𝚷20\boldsymbol{\Pi}_{2}^{0} in GG. Fix a countable basis {Vn:nω}\left\{V_{n}:n\in\omega\right\} for the topology of HH. Let also {zm:nω}\left\{z_{m}:n\in\omega\right\} be a countable dense subset of HH. Then we have that H=n,kωFkVnH=\bigcup_{n,k\in\omega}F_{k}^{\ast V_{n}} where, by Lemma 5.2, FkVnF_{k}^{\ast V_{n}} is closed in H~\tilde{H} and 𝚷20\boldsymbol{\Pi}_{2}^{0} in GG. Thus, without loss of generality we can assume that FkF_{k} is closed in H~\tilde{H} for every kωk\in\omega.

(3): Let (Fk)kω\left(F_{k}\right)_{k\in\omega} be as in (2). By the Baire Category Theorem, we can assume without loss of generality that there exists an open neighborhood VV of the identity in HH such that VF0V\subseteq F_{0}.

Fix an open neighborhood WW of the identity in HH contained in VV. By Lemma 6.3, there exists a neighborhood UU of the identity in HH that belongs to 𝚷20(G)|H\boldsymbol{\Pi}_{2}^{0}\left(G\right)|_{H} and such that UWU\subseteq W. Since UWVF0U\subseteq W\subseteq V\subseteq F_{0} and F0𝚷20(G)F_{0}\in\boldsymbol{\Pi}_{2}^{0}\left(G\right), we have that U𝚷20(G)U\in\boldsymbol{\Pi}_{2}^{0}\left(G\right).

Let now U1UU_{1}\subseteq U be an open neighborhood of the identity in HH such that U1U1UU_{1}U_{1}\subseteq U and U1𝚷20(G)U_{1}\in\boldsymbol{\Pi}_{2}^{0}\left(G\right). Then we have that U1UU1U_{1}\subseteq U^{\ast U_{1}} and UU1𝚷20(G)𝚷10(H~)U^{\ast U_{1}}\in\boldsymbol{\Pi}_{2}^{0}\left(G\right)\cap\boldsymbol{\Pi}_{1}^{0}(\tilde{H}) by Lemma 5.2. This concludes the proof that HH has a countable basis of neighborhoods of the identity consisting of sets that are in 𝚷20(G)𝚷10(H~)\boldsymbol{\Pi}_{2}^{0}\left(G\right)\cap\boldsymbol{\Pi}_{1}^{0}(\tilde{H}). ∎

Lemma 6.5.

If HH is 𝚺30\boldsymbol{\Sigma}_{3}^{0} in GG, then the coset equivalence relation EHGE_{H}^{G} is potentially 𝚺20\boldsymbol{\Sigma}_{2}^{0}, and HH is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) in GG.

Proof.

By Lemma 6.4, we can fix a countable basis {Vm:mω}\left\{V_{m}:m\in\omega\right\} of neighborhoods of the identity in HH that are 𝚷20\boldsymbol{\Pi}_{2}^{0} in GG and closed in HH. Fix also a countable dense subset {hk:kω}\left\{h_{k}:k\in\omega\right\} of HH. Let (Un)nω\left(U_{n}\right)_{n\in\omega} be an enumeration of the countable set {Vmhk:m,kω}\left\{V_{m}h_{k}:m,k\in\omega\right\}. We have that for every nonempty open subset WW of HH there exists nωn\in\omega such that UnWU_{n}\subseteq W.

Let H×GH\times G be the product Polish group. By [2, Theorem 5.1.8] applied to the continuous action a:H×GGa:H\times G\curvearrowright G defined by (h,g)x=hxg1\left(h,g\right)\cdot x=hxg^{-1}, together with [2, Theorem 5.1.3], there exists a Polish topology tt of GG such that the action a:H×G(G,t)a:H\times G\curvearrowright\left(G,t\right) is continuous, UnU_{n} is tt-closed for every nωn\in\omega, tt is finer than the Polish group topology of GG, and it generates the same Borel structure as the Polish group topology of GG. Since the action a:H×G(G,t)a:H\times G\curvearrowright\left(G,t\right) is continuous, we have that the left translation action H(G,t)H\curvearrowright\left(G,t\right) and the right translation action (G,t)G\left(G,t\right)\curvearrowleft G are continuous.

Fix a metric dd on GG compatible with tt. For a closed subset CC of GG and xGx\in G we define

d(x,C)=inf{d(x,c):cC}.d\left(x,C\right)=\mathrm{\inf}\left\{d\left(x,c\right):c\in C\right\}\text{.}

Let K(G,t)K\left(G,t\right) be the space of tt-closed subsets of GG. We regard K(G,t)K\left(G,t\right) as endowed with the Wijsman topology, which is obtained by declaring a net (Ci)\left(C_{i}\right) to converge to CC if and only if, for every xXx\in X, (d(Ci,x))\left(d\left(C_{i},x\right)\right) converges to d(C,x)d\left(C,x\right) in \mathbb{R}. This turns K(G,t)K\left(G,t\right) into a Polish space; see [1, Theorem 4.3]. The Borel σ\sigma-algebra on K(G,t)K\left(G,t\right) is the σ\sigma-algebra generated by sets of the form

{CK(G,t):CW},\left\{C\in K\left(G,t\right):C\cap W\neq\varnothing\right\}\text{,}

where WW is some tt-open subset of GG [9, Section 12.C]. The relation C0C1C_{0}\subseteq C_{1} for closed subsets C0,C1C_{0},C_{1} of GG is closed in K(G,t)K\left(G,t\right), since we have that C0C1C_{0}\subseteq C_{1} if and only if d(C1,x)d(C0,x)d\left(C_{1},x\right)\leq d\left(C_{0},x\right) for every xGx\in G.

Define the Borel function GK(G,t)ωG\rightarrow K\left(G,t\right)^{\omega}

x(Unx)nω.x\mapsto(U_{n}x)_{n\in\omega}\text{.}

Notice that this function is indeed Borel: if WGW\subseteq G is tt-open, then

{xG:UnxW}=uUnu1W\left\{x\in G:U_{n}x\cap W\neq\varnothing\right\}=\bigcup_{u\in U_{n}}u^{-1}W

is tt-open, and hence Borel, for every nωn\in\omega.

We have that, for x,yGx,y\in G, xEHGyxE_{H}^{G}y if and only if ω\exists\ell\in\omega, UxU0yU_{\ell}x\subseteq U_{0}y. Indeed, if xEHGyxE_{H}^{G}y then we have that Hx=HyHx=Hy. Thus, U0yx1HU_{0}yx^{-1}\subseteq H is closed and nonmeager in the Polish topology of HH, and hence there exists ω\ell\in\omega such that UU0yx1U_{\ell}\subseteq U_{0}yx^{-1}. Conversely if there exists ω\ell\in\omega such that UxU0yU_{\ell}x\subseteq U_{0}y then we have that HxHyHx\cap Hy\neq\varnothing and xEHGyxE_{H}^{G}y.

This shows that EHGE_{H}^{G} is potentially 𝚺20\boldsymbol{\Sigma}_{2}^{0} and in particular potentially D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}). It now follows from Proposition 3.1 that HH is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) in GG. ∎

Lemma 6.6.

If λ\lambda is a limit ordinal and HH is 𝚺λ0\boldsymbol{\Sigma}_{\lambda}^{0} in GG, then there exists μ<λ\mu<\lambda such that HH is 𝚷μ0\boldsymbol{\Pi}_{\mu}^{0} in GG.

Proof.

Let α\alpha be the Solecki rank of HH in GG. If α<λ\alpha<\lambda then HH is 𝚷1+α+10\boldsymbol{\Pi}_{1+\alpha+1}^{0} in GG and hence the conclusion holds. Suppose that αλ\alpha\geq\lambda. We have that H=kωFkH=\bigcup_{k\in\omega}F_{k} where, for kωk\in\omega, FkF_{k} is 𝚺μk0\boldsymbol{\Sigma}_{\mu_{k}}^{0} in GG for some μk<λ\mu_{k}<\lambda. In this case, as in the proof of Lemma 6.2, by Lemma 5.2 we can assume without loss of generality that FkF_{k} is closed in HH for every kωk\in\omega. By the Baire Category Theorem, without loss of generality we can assume that F0F_{0} is nonmeager in HH. Thus, we have that H=F0ΔH=F_{0}^{\Delta} is 𝚺μ00\boldsymbol{\Sigma}_{\mu_{0}}^{0} and in particular 𝚷μ0+10\boldsymbol{\Pi}_{\mu_{0}+1}^{0} in GG. ∎

Lemma 6.7.

If HH is 𝚫1+λ+n+10\boldsymbol{\Delta}_{1+\lambda+n+1}^{0} in GG for some 1n<ω1\leq n<\omega and λ<ω1\lambda<\omega_{1} either zero or limit, then HH is D(𝚷1+λ+n0)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0}) in GG.

Proof.

Fix a countable open basis {Vn:nω}\left\{V_{n}:n\in\omega\right\} for HH. We have that

H=sλ+nH(G)=s1H(sλ+n1H(G))𝚷30(sλ+n1H(G)).H=s_{\lambda+n}^{H}(G)=s_{1}^{H}\left(s_{\lambda+n-1}^{H}(G)\right)\in\boldsymbol{\Pi}_{3}^{0}\left(s_{\lambda+n-1}^{H}(G)\right)\text{.}

Furthermore, we can write H=kωFkH=\bigcup_{k\in\omega}F_{k} where FkF_{k} is 𝚷1+λ+n0\boldsymbol{\Pi}_{1+\lambda+n}^{0} in GG for every kωk\in\omega. Thus, we have that H=k,ωFkVH=\bigcup_{k,\ell\in\omega}F_{k}^{\ast V_{\ell}}, where FkV𝚷20(sλ+n1H(G))F_{k}^{\ast V_{\ell}}\in\boldsymbol{\Pi}_{2}^{0}\left(s_{\lambda+n-1}^{H}(G)\right) by Lemma 5.2. Thus, we have that H𝚺30(sλ+n1H(G))H\in\boldsymbol{\Sigma}_{3}^{0}\left(s_{\lambda+n-1}^{H}(G)\right). Hence, by Lemma 6.5 we have that HD(𝚷20)(sλ+n1H(G))H\in D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\lambda+n-1}^{H}(G)\right). Furthermore, D(𝚷20)(sλ+n1H(G))D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\lambda+n-1}^{H}(G)\right) is contained in D(𝚷1+λ+n0)(G)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0})(G) by Lemma 5.5, concluding the proof. ∎

Lemma 6.8.

If HH is 𝚺1+λ+n+10\boldsymbol{\Sigma}_{1+\lambda+n+1}^{0} in GG for some 1n<ω1\leq n<\omega and λ<ω1\lambda<\omega_{1} either zero or limit, then HH is D(𝚷1+λ+n0)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0}) in GG.

Proof.

Fix a countable open basis {Vn:nω}\left\{V_{n}:n\in\omega\right\} for H.H. Let α\alpha be the Solecki rank of HH in GG. Since HH is 𝚷1+λ+n+20\boldsymbol{\Pi}_{1+\lambda+n+2}^{0} we have that αλ+n+1\alpha\leq\lambda+n+1 by Theorem 5.4. If αλ+n\alpha\leq\lambda+n then we have that HH is 𝚫1+λ+n+10\boldsymbol{\Delta}_{1+\lambda+n+1}^{0} and hence HH is D(𝚷1+λ+n0)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0}) by Lemma 6.7. Suppose that α=λ+n+1\alpha=\lambda+n+1. Thus, we have that H=s1H(sλ+nH(G))H=s_{1}^{H}\left(s_{\lambda+n}^{H}(G)\right). We can write H=kωFkH=\bigcup_{k\in\omega}F_{k} where FkF_{k} is 𝚷1+λ+n0\boldsymbol{\Pi}_{1+\lambda+n}^{0} in GG for every kωk\in\omega. Thus we have that H=k,ωFkVH=\bigcup_{k,\ell\in\omega}F_{k}^{\ast V_{\ell}}, where, by Lemma 5.2, FkVF_{k}^{\ast V_{\ell}} is 𝚷20\boldsymbol{\Pi}_{2}^{0} in sλ+n1H(G)s_{\lambda+n-1}^{H}(G). Thus, H𝚺30(sλ+n1H(G))H\in\boldsymbol{\Sigma}_{3}^{0}\left(s_{\lambda+n-1}^{H}(G)\right). By Lemma 6.5, this implies that HD(𝚷20)(sλ+n1H(G))H\in D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\lambda+n-1}^{H}(G)\right). By Lemma 5.5, we have that D(𝚷20)(sλ+n1H(G))D(𝚷1+λ+n0)(G)D(\boldsymbol{\Pi}_{2}^{0})\left(s_{\lambda+n-1}^{H}(G)\right)\subseteq D(\boldsymbol{\Pi}_{1+\lambda+n}^{0})(G), concluding the proof. ∎

We have now all the ingredients to present a proof of Theorem 6.1.

Proof of Theorem 6.1.

(1) We have that HH is closed if and only if its Solecki rank is zero. Suppose now that λ\lambda is a limit ordinal and n=0n=0. By Theorem 5.4 we have that HH is 𝚷λ0\boldsymbol{\Pi}_{\lambda}^{0} and not 𝚷μ0\boldsymbol{\Pi}_{\mu}^{0} for μ<λ\mu<\lambda. Thus, HH is not 𝚺λ0\boldsymbol{\Sigma}_{\lambda}^{0} by Lemma 6.6.

(2a) By Lemma 5.5 we have that HH is 𝚷1+λ+n+10\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}. It remains to prove that HH is not 𝚺1+λ+n+10\boldsymbol{\Sigma}_{1+\lambda+n+1}^{0}. Suppose that HH is 𝚺1+λ+n+10\boldsymbol{\Sigma}_{1+\lambda+n+1}^{0}. Then by Lemma 6.8 we have that HH is D(𝚷1+λ+n0)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0}). Thus, by Corollary 5.3, HH\in D(𝚷20)(sλ+n1H(G))D(\boldsymbol{\Pi}_{2}^{0})(s_{\lambda+n-1}^{H}(G)), contradicting the hypothesis.

(2b) By Lemma 5.5 we have that HH is D(𝚷1+λ+n0)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0}). It remains to prove that HH is not Dˇ(𝚷1+λ+n0)\check{D}(\boldsymbol{\Pi}_{1+\lambda+n}^{0}). Suppose by contradiction that HH is Dˇ(𝚷1+λ+n0)\check{D}(\boldsymbol{\Pi}_{1+\lambda+n}^{0}). Then by Lemma 3.2 we have that HH is either 𝚷1+λ+n0\boldsymbol{\Pi}_{1+\lambda+n}^{0} or 𝚺1+λ+n0\boldsymbol{\Sigma}_{1+\lambda+n}^{0}. If HH is 𝚷1+λ+n0\boldsymbol{\Pi}_{1+\lambda+n}^{0} then by Theorem 5.4 and Proposition 3.4 we have that λ+n1\lambda+n-1 is the Solecki rank of HH in GG, contradicting the hypothesis. If HH is 𝚺1+λ+n0\boldsymbol{\Sigma}_{1+\lambda+n}^{0}, then H𝚺20(sλ+n1H(G))H\in\boldsymbol{\Sigma}_{2}^{0}(s_{\lambda+n-1}^{H}(G)) by Corollary 5.3, contradicting the hypothesis.

(2c) By Lemma 5.5 we have that HH is D(𝚷1+λ+n0)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0}). The same proof as (2b) shows that HH is not Dˇ(𝚷1+λ+n0)\check{D}(\boldsymbol{\Pi}_{1+\lambda+n}^{0}).

(2d) By Lemma 5.5 we have that HH is 𝚺1+λ+n0\boldsymbol{\Sigma}_{1+\lambda+n}^{0}. The same proof as (2b) shows that HH is not 𝚷1+λ+n0\boldsymbol{\Pi}_{1+\lambda+n}^{0}.

When HH is non-Archimedean, the case (2c) is excluded by Proposition 3.5. ∎

7. The Saint-Raymond rank

Saint-Raymond introduced in [18, Definition 18] a notion of rank (therein called degree) for Fréchetable subspaces of Fréchet spaces; see Section 9. We consider in this section the natural generalization of such a notion to Polishable subgroups of Polish groups. Recall that for a complexity class Γ\Gamma, and a Polishable subgroup HH of a Polish group GG, we define Γ(G)|H\Gamma(G)|_{H} to be the collection of sets of the form AHA\cap H for AΓ(G)A\in\Gamma(G).

Definition 7.1.

Suppose that GG is a Polish group, and HGH\subseteq G is a Polishable subgroup. The Saint-Raymond rank of HH is the least countable ordinal α\alpha such that every open subset in the Polish group topology of HH belongs to 𝚺1+α0(G)|H\boldsymbol{\Sigma}_{1+\alpha}^{0}(G)|_{H}.

Suppose that X,YX,Y are Polish spaces, and α\alpha is a countable ordinal. As in [18, page 216], one can define α(X,Y)\mathcal{B}_{\alpha}\left(X,Y\right) to be the set of Borel functions that have class α\alpha as in [10, Section 31], namely are 𝚺1+α0\boldsymbol{\Sigma}_{1+\alpha}^{0}-measurable; see [9, Definition 24.2]. By definition, the Saint-Raymond rank of HH is the least countable ordinal α\alpha such that the identity function of HH belongs to α(X,Y)\mathcal{B}_{\alpha}\left(X,Y\right), where XX is equal to HH endowed with the subspace topology inherited from GG, and YY is equal to HH endowed with its Polish group topology. Adapting an argument of Tsankov from [19], we now show that the Saint-Raymond rank and the Solecki rank of a Polishable subgroup of GG coincide.

Theorem 7.2.

Suppose that GG is a Polish group, and HGH\subseteq G is a Polishable subgroup. Then the Saint-Raymond rank is equal to the Solecki rank.

Proof.

By Lemma 5.1 we have that the Saint-Raymond rank is less than or equal to the Solecki rank. We prove the converse inequality as in the proof of [19, Proposition 4.6]. If HH has Saint-Raymond rank α\alpha, then every open set in HH belongs to 𝚺1+α0(G)|H\boldsymbol{\Sigma}_{1+\alpha}^{0}(G)|_{H}. Suppose that UU is an open neighborhood of the identity in HH, and let VV be an open neighborhood of the identity in HH such that V1VUV^{-1}V\subseteq U. Then there exists A𝚺1+α0(G)A\in\boldsymbol{\Sigma}_{1+\alpha}^{0}(G) such that AH=VA\cap H=V. Thus, 1AΔV1\in A^{\Delta V}, where AΔVsαH(G)A^{\Delta V}\cap s_{\alpha}^{H}(G) is open in sαH(G)s_{\alpha}^{H}(G) by Lemma 5.2. Furthermore, we have that AΔVHV1VUA^{\Delta V}\cap H\subseteq V^{-1}V\subseteq U. This shows that UU contains a neighborhood of the identity with respect to the topology on HH induced by sαH(G)s_{\alpha}^{H}(G). This shows that the Polish topology on HH is the subspace topology induced by sαH(G)s_{\alpha}^{H}(G), whence HH is closed in sαH(G)s_{\alpha}^{H}(G). As HH is dense in sαH(G)s_{\alpha}^{H}(G), we have that H=sαH(G)H=s_{\alpha}^{H}(G). ∎

8. Polishable subgroups in each complexity class

The goal of this section is to prove the following theorem. Recall that a Polish group is CLI if it admits a compatible complete left-invariant metric or, equivalently, its left uniformity is complete [13].

Theorem 8.1.

Let Γ\Gamma be one of the possible complexity classes of Polishable subgroups from Theorem 1.1. Suppose that GG is a nontrivial CLI Polish group. Then there exists a CLI Polishable subgroup of GG^{\mathbb{N}} whose complexity class is Γ\Gamma.

Remark 8.2.

After replacing GG with GG^{\mathbb{N}}, we can assume that GG is not discrete. We will assume that GG is not discrete in the rest of this section.

Recall that a pseudo-length function on a group HH is a function L:H[0,+)L:H\rightarrow[0,+\infty) such that, for h,hHh,h^{\prime}\in H:

  • L(1H)=0L\left(1_{H}\right)=0;

  • L(h1)=L(h)L\left(h^{-1}\right)=L\left(h\right);

  • L(hh)L(h)+L(h)L\left(hh^{\prime}\right)\leq L\left(h\right)+L\left(h^{\prime}\right).

A length function is a pseudo-length function LL such that L(h)=0h=1HL\left(h\right)=0\Rightarrow h=1_{H} for hHh\in H. A (pseudo-)length function LL gives rise to a left-invariant (pseudo-)metric dd defined by setting d(h,h)=L(h1h)d\left(h,h^{\prime}\right)=L\left(h^{-1}h^{\prime}\right), and every left-invariant metric arises in this fashion.

Suppose that GG is a CLI Polish group, and let LGL_{G} be a length function on GG that induces the Polish topology on GG. We define the length functions L1L_{1} and LL_{\infty} on GG^{\mathbb{N}}, with corresponding left-invariant metrics d1d_{1} and dd_{\infty}, by setting

L1((gn)n):=nLG(gn)L_{1}(\left(g_{n}\right)_{n\in\mathbb{N}}):=\sum_{n\in\mathbb{N}}L_{G}\left(g_{n}\right)

and

L((gn)n):=supnLG(gn).L_{\infty}(\left(g_{n}\right)_{n\in\mathbb{N}}):=\mathrm{sup}_{n\in\mathbb{N}}L_{G}\left(g_{n}\right)\text{.}

for a sequence (gn)nG\left(g_{n}\right)_{n\in\mathbb{N}}\in G^{\mathbb{N}}. We say that (gn)n\left(g_{n}\right)_{n\in\mathbb{N}} is LGL_{G}-summable if L1((gn)n)<L_{1}(\left(g_{n}\right)_{n\in\mathbb{N}})<\infty, and has bounded (left) LGL_{G}-variation if kLG(gn+11gn)<\sum_{k\in\mathbb{N}}L_{G}(g_{n+1}^{-1}g_{n})<\infty. We let 1(G,LG)G\ell_{1}\left(G,L_{G}\right)\subseteq G^{\mathbb{N}} to be the CLI Polishable subgroup of LGL_{G}-summable sequences, bv0(G,LG)G\mathrm{bv}_{0}\left(G,L_{G}\right)\subseteq G^{\mathbb{N}} to be the CLI Polishable subgroup of vanishing sequences of bounded LGL_{G}-variation, and c(G)G\mathrm{c}\left(G\right)\subseteq G^{\mathbb{N}} be the CLI Polishable subgroup of convergent sequences.

Fix, for each limit ordinal λ<ω1\lambda<\omega_{1}, an increasing cofinal sequence (λi)i\left(\lambda_{i}\right)_{i\in\mathbb{N}} in λ\lambda. If γ=δ+1\gamma=\delta+1 is a successor ordinal, define γi=δ\gamma_{i}=\delta for every ii\in\mathbb{N}. Define by recursion on α<ω1\alpha<\omega_{1}, I00={(,)}I_{0}^{0}=\left\{\left(\varnothing,\varnothing\right)\right\} where \varnothing is the empty tuple., and I0αI_{0}^{\alpha} to be the set of tuples (n0,,nd;β0,,βd)\left(n_{0},\ldots,n_{d};\beta_{0},\ldots,\beta_{d}\right) for dωd\in\omega, n0,,ndn_{0},\ldots,n_{d}\in\mathbb{N}, 0=β0<<βd=αnd0=\beta_{0}<\cdots<\beta_{d}=\alpha_{n_{d}}, and (n0,,nd1;β0,,βd1)I0αnd\left(n_{0},\ldots,n_{d-1};\beta_{0},\ldots,\beta_{d-1}\right)\in I_{0}^{\alpha_{n_{d}}}.

Similarly for a fixed γ<ω1\gamma<\omega_{1} we define IγαI_{\gamma}^{\alpha} by recursion on αγ\alpha\geq\gamma, by setting Iγγ={(,)}I_{\gamma}^{\gamma}=\left\{\left(\varnothing,\varnothing\right)\right\}, and IγαI_{\gamma}^{\alpha} to be the set of tuples (n0,,nd;β0,,βd)\left(n_{0},\ldots,n_{d};\beta_{0},\ldots,\beta_{d}\right) for dωd\in\omega, n0,,ndn_{0},\ldots,n_{d}\in\mathbb{N}, γ=β0<<βd=αnd\gamma=\beta_{0}<\cdots<\beta_{d}=\alpha_{n_{d}}, and (n0,,nd1;β0,,βd1)Iγαnd\left(n_{0},\ldots,n_{d-1};\beta_{0},\ldots,\beta_{d-1}\right)\in I_{\gamma}^{\alpha_{n_{d}}}. Notice that, by definition, if (n0,,nd;β0,,βd)Iγα\left(n_{0},\ldots,n_{d};\beta_{0},\ldots,\beta_{d}\right)\in I_{\gamma}^{\alpha} for some γ1\gamma\geq 1, then (m,n0,,nd;γm,β0,,βd)Iγmα\left(m,n_{0},\ldots,n_{d};\gamma_{m},\beta_{0},\ldots,\beta_{d}\right)\in I_{\gamma_{m}}^{\alpha} for every mm\in\mathbb{N}.

Thus, for example we have that, for 1k<ω1\leq k<\omega, Iγγ+kI_{\gamma}^{\gamma+k} is the set of tuples

(n0,,nk1;γ,γ+1,,γ+k1)\left(n_{0},\ldots,n_{k-1};\gamma,\gamma+1,\ldots,\gamma+k-1\right)

for n0,,nk1n_{0},\ldots,n_{k-1}\in\mathbb{N}, and Iγγ+ωI_{\gamma}^{\gamma+\omega} is the set of tuples

(n0,,n(γ+ω)d;γ,γ+1,,(γ+ω)d)(n_{0},\ldots,n_{\left(\gamma+\omega\right)_{d}};\gamma,\gamma+1,\ldots,\left(\gamma+\omega\right)_{d})

for dωd\in\omega such that (γ+ω)dγ\left(\gamma+\omega\right)_{d}\geq\gamma, and n0,,n(γ+ω)dn_{0},\ldots,n_{\left(\gamma+\omega\right)_{d}}\in\mathbb{N}.

We also define Iαα={(;)}I_{\alpha}^{\alpha}=\left\{\left(\varnothing;\varnothing\right)\right\}. We denote by IαI^{\alpha} the union of IγαI_{\gamma}^{\alpha} for γα\gamma\leq\alpha. If γα\gamma\leq\alpha, we denote by IγαI_{\leq\gamma}^{\alpha} the union of IδαI_{\delta}^{\alpha} for δγ\delta\leq\gamma, and by I<γαI_{<\gamma}^{\alpha} to be the union of IδαI_{\delta}^{\alpha} for δ<γ\delta<\gamma. For (n;β)\left(n;\beta\right) and (m;τ)\left(m;\tau\right) in IαI^{\alpha} we define (n;β)(m;τ)\left(n;\beta\right)\leq\left(m;\tau\right) if and only if there exist γ0γ1α\gamma_{0}\leq\gamma_{1}\leq\alpha such that (n;β)Iγ0α\left(n;\beta\right)\in I_{\gamma_{0}}^{\alpha}, (m;τ)Iγ1α\left(m;\tau\right)\in I_{\gamma_{1}}^{\alpha}, mm is a tail of nn, and τ\tau is a tail of β\beta, i.e. we have that, for some d<ω\ell\leq d<\omega, (n;β)=(n0,,nd;β0,,βd)\left(n;\beta\right)=\left(n_{0},\ldots,n_{d};\beta_{0},\ldots,\beta_{d}\right), (m;τ)=(m0,,m;τ0,,τ)\left(m;\tau\right)=\left(m_{0},\ldots,m_{\ell};\tau_{0},\ldots,\tau_{\ell}\right), and for 0i0\leq i\leq\ell, mi=ni+dm_{i}=n_{i+d-\ell} and τi=βi+d\tau_{i}=\beta_{i+d-\ell} . We regard IαI^{\alpha} as an ordered set with respect to this order relation. Observe that IγαI_{\leq\gamma}^{\alpha} and I<γαI_{<\gamma}^{\alpha} are downward-closed. For a subset FF of IαI^{\alpha}, we denote by FF_{\downarrow} its downward closure. Notice also that if FIαF\subseteq I^{\alpha} is finite, and (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha} for some γ1\gamma\geq 1 is such that (k,n;γk,β)F\left(k,n;\gamma_{k},\beta\right)\in F_{\downarrow} for infinitely many kk\in\mathbb{N}, then (n;β)F\left(n;\beta\right)\in F_{\downarrow}.

Fix a countable ordinal α\alpha. We define, by recursion on γ<α\gamma<\alpha, a decreasing sequence (Pγ)γ<α\left(P_{\gamma}\right)_{\gamma<\alpha} of CLI Polishable subgroups of GI0αG^{I_{0}^{\alpha}}. Furthermore, for xPγx\in P_{\gamma}, we define the values x(n;β)Gx\left(n;\beta\right)\in G for (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}. If γ1\gamma\geq 1 and (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}, then we let 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right) be the convergent sequence (x(k,n;γk,β))kω\left(x\left(k,n;\gamma_{k},\beta\right)\right)_{k\in\omega} in GG with limit x(n;β)x\left(n;\beta\right). If (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha}, then we let 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right) be the sequence constantly equal to x(n;β)x\left(n;\beta\right).

Define P0P_{0} to be GI0αG^{I_{0}^{\alpha}}. This is a CLI Polish group with topology induced by the pseudo-length functions

L0(n;β)(x)=LG(x(n;β))L_{0}^{\left(n;\beta\right)}\left(x\right)=L_{G}(x\left(n;\beta\right))

for (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha}. Suppose that 1γα1\leq\gamma\leq\alpha, and that PδP_{\delta} has been defined for every δ<γ\delta<\gamma. Define P<γ=δ<γPδP_{<\gamma}=\bigcap_{\delta<\gamma}P_{\delta}, and PγP<γP_{\gamma}\subseteq P_{<\gamma} to contain those xP<γx\in P_{<\gamma} such that, for every (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}, the sequence 𝒙(n;β):=(x(i,n;γi,β))i\boldsymbol{x}\left(n;\beta\right):=\left(x\left(i,n;\gamma_{i},\beta\right)\right)_{i\in\mathbb{N}} is convergent. For xPγx\in P_{\gamma} and (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}, we define x(n;β)x\left(n;\beta\right) to be the limit of 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right). Then we have that the Polish topology on PγP_{\gamma} is induced by the restriction to PγP_{\gamma} of the continuous pseudo-length functions on PδP_{\delta} for δ<γ\delta<\gamma, together with the pseudo-length functions Lγ(n;β)(x)=L(𝒙(n;β))L_{\gamma}^{\left(n;\beta\right)}\left(x\right)=L_{\infty}(\boldsymbol{x}\left(n;\beta\right)) for (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}. This concludes the recursive definition of the CLI Polishable subgroups PγP_{\gamma} of GI0αG^{I_{0}^{\alpha}} for γα\gamma\leq\alpha. Notice that in particular PαP_{\alpha} contains the elements xP<αx\in P_{<\alpha} such that the sequence 𝒙(α):=(x(n,αn))n\boldsymbol{x}\left(\alpha\right):=\left(x\left(n,\alpha_{n}\right)\right)_{n\in\mathbb{N}} belongs to c(G)\mathrm{c}\left(G\right). We also define SαS_{\alpha} and DαD_{\alpha} to be the subgroups of PαP_{\alpha} containing the elements xP<αx\in P_{<\alpha} such that the sequence 𝒙(α)\boldsymbol{x}\left(\alpha\right) belongs to 1(G,LG)\ell_{1}\left(G,L_{G}\right) and bv0(G,LG)\mathrm{bv}_{0}\left(G,L_{G}\right), respectively. Theorem 8.1 will be a consequence of the following.

Theorem 8.3.

Fix α=1+λ+n<ω1\alpha=1+\lambda+n<\omega_{1}, where λ\lambda is a limit ordinal or zero and n<ωn<\omega:

  1. (1)

    if n=0n=0 and λ\lambda is limit, then P<λP_{<\lambda} has Solecki rank λ\lambda in GI0αG^{I_{0}^{\alpha}}, and complexity class 𝚷λ0\boldsymbol{\Pi}_{\lambda}^{0};

  2. (2)

    if n=0n=0, then S1+λS_{1+\lambda}, D1+λD_{1+\lambda}, and P1+λP_{1+\lambda} have Solecki rank λ+1\lambda+1 in GI0αG^{I_{0}^{\alpha}}, and complexity class 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0}, D(𝚷1+λ+10)D(\boldsymbol{\Pi}_{1+\lambda+1}^{0}), and 𝚷1+λ+10\boldsymbol{\Pi}_{1+\lambda+1}^{0} respectively;

  3. (3)

    if n1n\geq 1, then S1+λ+nS_{1+\lambda+n}, D1+λ+nD_{1+\lambda+n}, and P1+λ+nP_{1+\lambda+n} have Solecki rank λ+n+1\lambda+n+1 in GI0αG^{I_{0}^{\alpha}}, and complexity class D(𝚷1+λ+n+10)D(\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}), D(𝚷1+λ+n+10)D(\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}), and 𝚷1+λ+n+10\boldsymbol{\Pi}_{1+\lambda+n+1}^{0} respectively.

We will obtain Theorem 8.3 as a consequence of a number of lemmas.

Lemma 8.4.

Suppose that γ<α\gamma<\alpha, FF is a finite subset of IγαI_{\leq\gamma}^{\alpha}, and xPγx\in P_{\gamma}. Define yGI0αy\in G^{I_{0}^{\alpha}} by setting, for (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha},

y(n;β):={x(n;β)if (n;β)F;1Gotherwise.y\left(n;\beta\right):=\left\{\begin{array}[]{ll}x\left(n;\beta\right)&\text{if }\left(n;\beta\right)\in F_{\downarrow}\text{;}\\ 1_{G}&\text{otherwise.}\end{array}\text{}\right. (1)

Then we have that ySαy\in S_{\alpha}, and (1) holds for every (n;β)Iα\left(n;\beta\right)\in I^{\alpha}.

Proof.

We prove by induction on σ<α\sigma<\alpha that yPσy\in P_{\sigma}, and that (1) holds for every (n;β)Iσα\left(n;\beta\right)\in I_{\sigma}^{\alpha}. For σ=0\sigma=0, this holds by definition. Suppose that the conclusion holds for every δ<σ\delta<\sigma. Fix (n;β)Iσα\left(n;\beta\right)\in I_{\sigma}^{\alpha}. If (n;β)F\left(n;\beta\right)\in F_{\downarrow} then necessarily σγ\sigma\leq\gamma, and for every kk\in\mathbb{N}, (k,n;σk,β)IσkαF\left(k,n;\sigma_{k},\beta\right)\in I_{\sigma_{k}}^{\alpha}\cap F_{\downarrow} and hence by the inductive hypothesis, we have that y(k,n;σk,β)=x(k,n;σk,β)y\left(k,n;\sigma_{k},\beta\right)=x\left(k,n;\sigma_{k},\beta\right). Since xPγPσkx\in P_{\gamma}\subseteq P_{\sigma_{k}}, we have that the sequence 𝒚(n;β)=𝒙(n;β)\boldsymbol{y}\left(n;\beta\right)=\boldsymbol{x}\left(n;\beta\right) converges to x(n;β)x\left(n;\beta\right). Thus, y(n;β)=x(n;β)y\left(n;\beta\right)=x\left(n;\beta\right). If (n;β)F\left(n;\beta\right)\notin F_{\downarrow} then we have that there exists k0k_{0} such that, for all kk0k\geq k_{0}, (k,n;σk,β)F\left(k,n;\sigma_{k},\beta\right)\notin F_{\downarrow}. Therefore, we have that the sequence 𝒚(n;β)\boldsymbol{y}\left(n;\beta\right) is eventually equal to 1G1_{G}, and thus y(n;β)=1Gy\left(n;\beta\right)=1_{G}. This shows that yPσy\in P_{\sigma}. This concludes the proof by induction.

By the above, we have that yP<αy\in P_{<\alpha}. For kk\in\mathbb{N} such that αk>γ\alpha_{k}>\gamma we have that y(k,αk)=1Gy\left(k,\alpha_{k}\right)=1_{G} and hence ySαy\in S_{\alpha}. ∎

Lemma 8.5.

For every γ<α\gamma<\alpha, SαS_{\alpha} is dense in PγP_{\gamma}.

Proof.

Suppose that xPγx\in P_{\gamma}, and let VV be a neighborhood of xx in PγP_{\gamma}. Then we have that there exist ε>0\varepsilon>0 and a finite subset FF of IγαI_{\leq\gamma}^{\alpha} such that

(n;β)F{zPγ:d(𝒙(n;β),𝒛(n;β))<ε}\bigcap_{\left(n;\beta\right)\in F}\left\{z\in P_{\gamma}:d_{\infty}(\boldsymbol{x}\left(n;\beta\right),\boldsymbol{z}\left(n;\beta\right))<\varepsilon\right\}

is contained in VV. Define zGI0αz\in G^{I_{0}^{\alpha}} by setting, for (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha},

z(n;β)={x(n;β)if (n;β)F1Gotherwise.z\left(n;\beta\right)=\left\{\begin{array}[]{ll}x\left(n;\beta\right)&\text{if }\left(n;\beta\right)\in F_{\downarrow}\\ 1_{G}&\text{otherwise.}\end{array}\right.

Then by Lemma 8.4, we have that zSαz\in S_{\alpha} and, for (n;β)Iα\left(n;\beta\right)\in I^{\alpha} we have that

z(n;β)={x(n;β)if (n;β)F;1Gotherwise.z\left(n;\beta\right)=\left\{\begin{array}[]{ll}x\left(n;\beta\right)&\text{if }\left(n;\beta\right)\in F_{\downarrow}\text{;}\\ 1_{G}&\text{otherwise.}\end{array}\right.

In particular, we have that zVz\in V. ∎

Lemma 8.6.

For every γ<α\gamma<\alpha, for every open neighborhood VV of the identity in SαS_{\alpha}, V¯P<γPγ\overline{V}^{P_{<\gamma}}\cap P_{\gamma} contains an open neighborhood of the identity in PγP_{\gamma}.

Proof.

Let VV be a neighborhood of the identity in SαS_{\alpha}. Fix a finite subset FF of IαI^{\alpha} and ε>0\varepsilon>0 such that

{wSα:L1(𝒘(α))<ε}(n;β)F{wSα:L(𝒘(n;β))<ε}\left\{w\in S_{\alpha}:L_{1}(\boldsymbol{w}\left(\alpha\right))<\varepsilon\right\}\cap\bigcap_{\left(n;\beta\right)\in F}\left\{w\in S_{\alpha}:L_{\infty}(\boldsymbol{w}\left(n;\beta\right))<\varepsilon\right\}

is contained in VV. Define

N=maxγ<δαFIδαmax{n:δnγ}N=\max_{\begin{subarray}{c}\gamma<\delta\leq\alpha\\ F\cap I_{\delta}^{\alpha}\neq\varnothing\end{subarray}}\max\left\{n\in\mathbb{N}:\delta_{n}\leq\gamma\right\}

Define also the finite subset

B=(FIγα){(k,n;δk,β):γ<δα,kN,(n;β)FIδα}B=\left(F\cap I_{\leq\gamma}^{\alpha}\right)\cup\left\{\left(k,n;\delta_{k},\beta\right):\gamma<\delta\leq\alpha,k\leq N,\left(n;\beta\right)\in F\cap I_{\delta}^{\alpha}\right\}

of IγαI_{\leq\gamma}^{\alpha}. Consider the open neighborhood WW of the identity in PγP_{\gamma} defined by

W={xPγ:nNLG(x(n;αn))<ε}(n;β)B{xPγ:L(𝒙(n;β))<ε}.W=\left\{x\in P_{\gamma}:\sum_{n\leq N}L_{G}(x\left(n;\alpha_{n}\right))<\varepsilon\right\}\cap\bigcap_{\left(n;\beta\right)\in B}\left\{x\in P_{\gamma}:L_{\infty}(\boldsymbol{x}\left(n;\beta\right))<\varepsilon\right\}\text{.}

We claim that WV¯P<γPγW\subseteq\overline{V}^{P_{<\gamma}}\cap P_{\gamma}. Suppose that xWx\in W. Let UU be an open neighborhood of xx in P<γP_{<\gamma}. Then there exist a finite subset AA of I<γαI_{<\gamma}^{\alpha} containing BI<γαB\cap I_{<\gamma}^{\alpha} and ε1>0\varepsilon_{1}>0 such that

(n;β)A{zP<γ:d(𝒙(n;β),𝒛(n;β))<ε1}\bigcap_{\left(n;\beta\right)\in A}\left\{z\in P_{<\gamma}:d_{\infty}(\boldsymbol{x}\left(n;\beta\right),\boldsymbol{z}\left(n;\beta\right))<\varepsilon_{1}\right\}

is contained in UU. We need to prove that UVU\cap V\neq\varnothing.

We define zGI0αz\in G^{I_{0}^{\alpha}} by setting, for (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha},

z(n;β):={x(n;β)if (n;β)A;1Gotherwise.z\left(n;\beta\right):=\left\{\begin{array}[]{ll}x\left(n;\beta\right)&\text{if }\left(n;\beta\right)\in A_{\downarrow}\text{;}\\ 1_{G}&\text{otherwise.}\end{array}\right.

Then by Lemma 8.4, we have that zSαz\in S_{\alpha} and, for (n;β)Iα\left(n;\beta\right)\in I^{\alpha}, we have that

z(n;β)={x(n;β)if (n;β)A;1Gotherwise.z\left(n;\beta\right)=\left\{\begin{array}[]{ll}x\left(n;\beta\right)&\text{if }\left(n;\beta\right)\in A_{\downarrow}\text{;}\\ 1_{G}&\text{otherwise.}\end{array}\right.

In particular, we have that zUz\in U. We need to show that zVz\in V, i.e., that L1(𝒛(α))<εL_{1}(\boldsymbol{z}\left(\alpha\right))<\varepsilon and if (n,β)F\left(n,\beta\right)\in F, then L(𝒛(n;β))<εL_{\infty}(\boldsymbol{z}\left(n;\beta\right))<\varepsilon. We have that

L1(𝒛(α))1=nLG(z(n;αn))nNLG(x(n;αn))<ε.L_{1}(\boldsymbol{z}\left(\alpha\right))_{1}=\sum_{n\in\mathbb{N}}L_{G}(z\left(n;\alpha_{n}\right))\leq\sum_{n\leq N}L_{G}(x\left(n;\alpha_{n}\right))<\varepsilon\text{.}

If (n,β)FI<γα\left(n,\beta\right)\in F\cap I_{<\gamma}^{\alpha}, then 𝒛(n;β)=𝒙(n;β)\boldsymbol{z}\left(n;\beta\right)=\boldsymbol{x}\left(n;\beta\right). As (n;β)B\left(n;\beta\right)\in B and xWx\in W, this implies that L(𝒛(n;β))=L(𝒙(m;β))<εL_{\infty}(\boldsymbol{z}\left(n;\beta\right))=L_{\infty}(\boldsymbol{x}\left(m;\beta\right))<\varepsilon. If (n,β)FIγα\left(n,\beta\right)\in F\cap I_{\gamma}^{\alpha}, then we have that

L(𝒛(n;β))\displaystyle L_{\infty}(\boldsymbol{z}\left(n;\beta\right)) =\displaystyle= supkLG(z(k,n;γk,β))\displaystyle\mathrm{sup}_{k\in\mathbb{N}}L_{G}(z\left(k,n;\gamma_{k},\beta\right))
\displaystyle\leq supkLG(x(k,n;γk,β))=L(𝒙(n;β))<ε\displaystyle\mathrm{sup}_{k\in\mathbb{N}}L_{G}(x\left(k,n;\gamma_{k},\beta\right))=L_{\infty}(\boldsymbol{x}\left(n;\beta\right))<\varepsilon

since (n;β)B\left(n;\beta\right)\in B and xWx\in W. If (n;β)FIδα\left(n;\beta\right)\in F\cap I_{\delta}^{\alpha} for some δ>γ\delta>\gamma, then

L(𝒛(n;β))\displaystyle L_{\infty}(\boldsymbol{z}\left(n;\beta\right)) =\displaystyle= supkLG(z(k,n;δk,β))\displaystyle\mathrm{sup}_{k\in\mathbb{N}}L_{G}(z\left(k,n;\delta_{k},\beta\right))
\displaystyle\leq maxkNLG(x(k,n;δk,β))maxkNL(𝒙(k,n;δk,β))<ε\displaystyle\max_{k\leq N}L_{G}(x\left(k,n;\delta_{k},\beta\right))\leq\max_{k\leq N}L_{\infty}(\boldsymbol{x}\left(k,n;\delta_{k},\beta\right))<\varepsilon

since (k,n;δk,β)B\left(k,n;\delta_{k},\beta\right)\in B for kNk\leq N, and xWx\in W. This shows that zVz\in V, concluding the proof. ∎

Proposition 8.7.

For γ<α\gamma<\alpha we have that

sγSα(GI0α)=sγDα(GI0α)=sγPα(GI0α)=sγP<α(GI0α)=P<(1+γ)s_{\gamma}^{S_{\alpha}}(G^{I_{0}^{\alpha}})=s_{\gamma}^{D_{\alpha}}(G^{I_{0}^{\alpha}})=s_{\gamma}^{P_{\alpha}}(G^{I_{0}^{\alpha}})=s_{\gamma}^{P_{<\alpha}}(G^{I_{0}^{\alpha}})=P_{<\left(1+\gamma\right)}
Proof.

Since SαDαPαP<(1+γ)S_{\alpha}\subseteq D_{\alpha}\subseteq P_{\alpha}\subseteq P_{<\left(1+\gamma\right)}, it suffices to prove that sγSα(GI0α)=P<(1+γ)s_{\gamma}^{S_{\alpha}}(G^{I_{0}^{\alpha}})=P_{<\left(1+\gamma\right)}. We do this by induction on γ<α\gamma<\alpha. For γ=0\gamma=0, we have that SαS_{\alpha} is dense in GI0α=P0G^{I_{0}^{\alpha}}=P_{0} by Lemma 8.5, and hence s0Sα(GI0α)=P0s_{0}^{S_{\alpha}}(G^{I_{0}^{\alpha}})=P_{0}. Suppose that the conclusion holds for every δ<γ\delta<\gamma. If γ\gamma is limit, then we have that

sγSα(GI0α)=δ<γsδ+1Sα(GI0α)=δ<γP1+δ=P<γ=P<(1+γ).s_{\gamma}^{S_{\alpha}}(G^{I_{0}^{\alpha}})=\bigcap_{\delta<\gamma}s_{\delta+1}^{S_{\alpha}}(G^{I_{0}^{\alpha}})=\bigcap_{\delta<\gamma}P_{1+\delta}=P_{<\gamma}=P_{<\left(1+\gamma\right)}\text{.}

Suppose that γ=δ+1\gamma=\delta+1 is a successor ordinal. Then we have that, by the inductive hypothesis

sγSα(GI0α)=sδ+1Sα(GI0α)=s1Sα(sδSα(GI0α))=s1Sα(P<(1+δ)).s_{\gamma}^{S_{\alpha}}(G^{I_{0}^{\alpha}})=s_{\delta+1}^{S_{\alpha}}(G^{I_{0}^{\alpha}})=s_{1}^{S_{\alpha}}(s_{\delta}^{S_{\alpha}}(G^{I_{0}^{\alpha}}))=s_{1}^{S_{\alpha}}\left(P_{<(1+\delta)}\right)\text{.}

Thus, it remains to prove that

s1Sα(P<(1+δ))=P1+δ.s_{1}^{S_{\alpha}}\left(P_{<\left(1+\delta\right)}\right)=P_{1+\delta}\text{.}

Notice that P1+δP_{1+\delta} is a 𝚷30\boldsymbol{\Pi}_{3}^{0} subspace of P<(1+δ)P_{<\left(1+\delta\right)}. Thus, the conclusion follows from Lemma 4.2, in view of Lemma 8.5 and Lemma 8.6. ∎

Lemma 8.8.

For every γα\gamma\leq\alpha, there exists a continuous group homomorphism Φ:GI<γαP<γ\Phi:G^{I_{<\gamma}^{\alpha}}\rightarrow P_{<\gamma} such that Φ(z)(k,n;γk,β)=z(k,n;γk,β)\Phi\left(z\right)\left(k,n;\gamma_{k},\beta\right)=z\left(k,n;\gamma_{k},\beta\right) for every zGI<γαz\in G^{I_{<\gamma}^{\alpha}}, (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}, and kk\in\mathbb{N}.

Proof.

For zGI<γαz\in G^{I_{<\gamma}^{\alpha}}, define Φ(z):=xGI0α\Phi\left(z\right):=x\in G^{I_{0}^{\alpha}} by setting, for (m;τ)I0α\left(m;\tau\right)\in I_{0}^{\alpha},

x(m;τ):={z(k,n;γk,β)if (m,τ)(k,n;γk,β) for some k and (n;β)Iγα;1Gotherwise.x\left(m;\tau\right):=\left\{\begin{array}[]{ll}z\left(k,n;\gamma_{k},\beta\right)&\text{if }\left(m,\tau\right)\leq\left(k,n;\gamma_{k},\beta\right)\text{ for some }k\in\mathbb{N}\text{ and }\left(n;\beta\right)\in I_{\gamma}^{\alpha}\text{;}\\ 1_{G}&\text{otherwise.}\end{array}\right.

It is clear that Φ:GI<γαGI0α\Phi:G^{I_{<\gamma}^{\alpha}}\rightarrow G^{I_{0}^{\alpha}} is a continuous group homomorphism. One can prove by induction on δ<γ\delta<\gamma that xPδx\in P_{\delta}, and for (m,τ)Iδα\left(m,\tau\right)\in I_{\delta}^{\alpha},

x(m;τ)={z(k,n;γk,β)if (m,τ)z(k,n;γk,β) for some k and (n;β)Iγα;1Gotherwise.x\left(m;\tau\right)=\left\{\begin{array}[]{ll}z\left(k,n;\gamma_{k},\beta\right)&\text{if }\left(m,\tau\right)\leq z\left(k,n;\gamma_{k},\beta\right)\text{ for some }k\in\mathbb{N}\text{ and }\left(n;\beta\right)\in I_{\gamma}^{\alpha}\text{;}\\ 1_{G}&\text{otherwise.}\end{array}\right.

This concludes the proof. ∎

Lemma 8.9.

We have that:

  1. (1)

    𝚺20\boldsymbol{\Sigma}_{2}^{0} is the complexity class of 1(G,LG)\ell_{1}\left(G,L_{G}\right) in GG^{\mathbb{N}};

  2. (2)

    D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) is the complexity class of bv0(G,LG)\mathrm{bv}_{0}\left(G,L_{G}\right) in GG^{\mathbb{N}};

  3. (3)

    𝚷30\boldsymbol{\Pi}_{3}^{0} is the complexity class of c(G)\mathrm{c}\left(G\right) in GG^{\mathbb{N}}.

Proof.

(1): It is clear that 1(G,LG)\ell_{1}\left(G,L_{G}\right) is 𝚺20\boldsymbol{\Sigma}_{2}^{0} in GG^{\mathbb{N}}. Since 1(G,LG)\ell_{1}\left(G,L_{G}\right) is a dense, proper subgroup of GG^{\mathbb{N}}, it is not closed.

(2): We have that (gn)nbv0(G,LG)\left(g_{n}\right)_{n\in\mathbb{N}}\in\mathrm{bv}_{0}\left(G,L_{G}\right) if and only if

nLG(gn+11gn)<+\sum_{n\in\mathbb{N}}L_{G}\left(g_{n+1}^{-1}g_{n}\right)<+\infty

and for all ε>0\varepsilon>0 and n0n_{0}\in\mathbb{N} there exists nn0n\geq n_{0} such that LG(gn)<εL_{G}\left(g_{n}\right)<\varepsilon. This shows that bv0(G,LG)\mathrm{bv}_{0}\left(G,L_{G}\right) is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) in GG^{\mathbb{N}}. It remains to prove that it is not 𝚺20\boldsymbol{\Sigma}_{2}^{0}. Suppose by contradiction that

bv0(G,LG)=kFk\mathrm{bv}_{0}\left(G,L_{G}\right)=\bigcup_{k\in\mathbb{N}}F_{k}

where FkGF_{k}\subseteq G^{\mathbb{N}} is closed for every kk\in\mathbb{N}. By the Baire Category Theorem, we can assume without loss of generality that F0F_{0} contains a neighborhood of the identity. Thus, there exists ε>0\varepsilon>0 such that

{(gn)nbv0(G,LG):nLG(gn+11gn)<ε and supnLG(gn)<ε}F0.\left\{\left(g_{n}\right)_{n\in\mathbb{N}}\in\mathrm{bv}_{0}\left(G,L_{G}\right):\sum_{n\in\mathbb{N}}L_{G}\left(g_{n+1}^{-1}g_{n}\right)<\varepsilon\text{ and }\mathrm{\mathrm{sup}}_{n\in\mathbb{N}}L_{G}\left(g_{n}\right)<\varepsilon\right\}\subseteq F_{0}\text{.}

Since we are assuming that GG is not discrete—see Remark 8.2—there exists gGg\in G such that 0<LG(g)<ε0<L_{G}(g)<\varepsilon. Define for NN\in\mathbb{N}, x(N)bv0(G,LG)x^{\left(N\right)}\in\mathrm{bv}_{0}\left(G,L_{G}\right) by setting

xk(N)={gif kN;1Gotherwise.x_{k}^{\left(N\right)}=\left\{\begin{array}[]{cc}g&\text{if }k\leq N\text{;}\\ 1_{G}&\text{otherwise.}\end{array}\right.

Then we have that x(N)F0x^{\left(N\right)}\in F_{0} for every NN\in\mathbb{N}. The sequence (x(N))N\left(x^{\left(N\right)}\right)_{N\in\mathbb{N}} converges in GG^{\mathbb{N}} to the element xGx\in G^{\mathbb{N}} that is the sequence constantly equal to gg. Since F0F_{0} is closed in GG^{\mathbb{N}}, we have that xbv0(G,LG)x\in\mathrm{bv}_{0}\left(G,L_{G}\right), which is a contradiction to the fact that LG(g)>0L_{G}\left(g\right)>0.

(3): By definition, we have that c(G)\mathrm{c}\left(G\right) is 𝚷30\boldsymbol{\Pi}_{3}^{0} in GG^{\mathbb{N}}. By Theorem 3.3, it suffices to prove that c(G)\mathrm{c}\left(G\right) is not potentially 𝚺20\boldsymbol{\Sigma}_{2}^{0}. Let E0E_{0} be the relation of tail equivalence in 22^{\mathbb{N}}, and let E0E_{0}^{\mathbb{N}} be the corresponding product equivalence relation on (2)=2×\left(2^{\mathbb{N}}\right)^{\mathbb{N}}=2^{\mathbb{N}\times\mathbb{N}}. Then we have that 𝚷30\boldsymbol{\Pi}_{3}^{0} is the potential complexity class of E0E_{0}^{\mathbb{N}}, for example by Lemma 5.7 and Theorem 3.3.

Thus, it suffices to define a Borel function 2×G2^{\mathbb{N}\times\mathbb{N}}\rightarrow G^{\mathbb{N}} that is a Borel reduction from E0E_{0}^{\mathbb{N}} to the coset relation of c(G)\mathrm{c}\left(G\right) inside GG^{\mathbb{N}}. We argue as in [6, Lemma 8.5.3]. Fix a bijection ,:×ω\left\langle\cdot,\cdot\right\rangle:\mathbb{N}\times\mathbb{N}\rightarrow\omega such that, if nnn\leq n^{\prime} and mmm\leq m^{\prime}, then n,mn,m\left\langle n,m\right\rangle\leq\left\langle n^{\prime},m^{\prime}\right\rangle. Fix also a sequence (gn)n\left(g_{n}\right)_{n\in\mathbb{N}} in GG such that 0<LG(gn)<2(n+1)0<L_{G}\left(g_{n}\right)<2^{-(n+1)} for every nn\in\mathbb{N}. Define Ξ:2×Gω\Xi:2^{\mathbb{N}\times\mathbb{N}}\rightarrow G^{\omega}, φa\varphi\mapsto a by setting

an,m={gnφ(n,m)=1;1Gφ(n,m)=0.a_{\left\langle n,m\right\rangle}=\left\{\begin{array}[]{cc}g_{n}&\varphi\left(n,m\right)=1\text{;}\\ 1_{G}&\varphi\left(n,m\right)=0\text{.}\end{array}\right.

Fix φ,ψ2×\varphi,\psi\in 2^{\mathbb{N}\times\mathbb{N}}. Define Ξ(φ)=a\Xi\left(\varphi\right)=a and Ξ(ψ)=b\Xi\left(\psi\right)=b.

Suppose that φE0ψ\varphi E_{0}^{\mathbb{N}}\psi. Thus we have that for every nn\in\mathbb{N} there exists MnM_{n}\in\mathbb{N} such that φ(n,m)=ψ(n,m)\varphi\left(n,m\right)=\psi\left(n,m\right) for mMnm\geq M_{n}. Fix ε>0\varepsilon>0 and fix NN\in\mathbb{N} such that 2N<ε2^{-N}<\varepsilon. Define then

M=max{Mn:n<N}.M=\max\left\{M_{n}:n<N\right\}\text{.}

We claim that for kN,Mk\geq\left\langle N,M\right\rangle we have that LG(ak1bk)<εL_{G}\left(a_{k}^{-1}b_{k}\right)<\varepsilon. Indeed, suppose that kN,Mk\geq\left\langle N,M\right\rangle. Then k=n,mk=\left\langle n,m\right\rangle for some n,mn,m\in\mathbb{N}. If nNn\geq N then we have that

LG(ak1bk)\displaystyle L_{G}\left(a_{k}^{-1}b_{k}\right) \displaystyle\leq LG(ak)+LG(bk)\displaystyle L_{G}\left(a_{k}\right)+L_{G}\left(b_{k}\right)
\displaystyle\leq 2LG(gn)<2n2N<ε.\displaystyle 2L_{G}\left(g_{n}\right)<2^{-n}\leq 2^{-N}<\varepsilon\text{.}

If n<Nn<N then we must have that mMMnm\geq M\geq M_{n}, and hence φ(n,m)=ψ(n,m)\varphi\left(n,m\right)=\psi\left(n,m\right) and ak=bka_{k}=b_{k}. This shows that a1bc(G)a^{-1}b\in\mathrm{c}\left(G\right).

Conversely, suppose that a1bc(G)a^{-1}b\in\mathrm{c}\left(G\right). Fix n0n_{0}\in\mathbb{N}. Then we have that there exists k0k_{0}\in\mathbb{N} such that for kk0k\geq k_{0}, LG(ak1bk)<LG(gn0)L_{G}\left(a_{k}^{-1}b_{k}\right)<L_{G}\left(g_{n_{0}}\right). Thus, for kk0k\geq k_{0}, if k=n0,mk=\left\langle n_{0},m\right\rangle for some mm\in\mathbb{N}, we must have ak=bka_{k}=b_{k} and φ(n0,m)=ψ(n0,m)\varphi\left(n_{0},m\right)=\psi\left(n_{0},m\right). Thus, if m0m_{0}\in\mathbb{N} is such that n0,m0k0\left\langle n_{0},m_{0}\right\rangle\geq k_{0}, we must have that φ(n0,m)=ψ(n0,m)\varphi\left(n_{0},m\right)=\psi\left(n_{0},m\right) for all mm0m\geq m_{0}. As this holds for every n0n_{0}\in\mathbb{N}, φE0ψ\varphi E_{0}^{\mathbb{N}}\psi, concluding the proof. ∎

Corollary 8.10.

For every γ<α\gamma<\alpha, PγP_{\gamma} is a proper subgroup of P<γP_{<\gamma}. The complexity class of SαS_{\alpha}, DαD_{\alpha}, PαP_{\alpha}, respectively, inside P<αP_{<\alpha} is 𝚺20\boldsymbol{\Sigma}_{2}^{0}, D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}), and 𝚷30\boldsymbol{\Pi}_{3}^{0}, respectively.

Proof.

Fix γ<α\gamma<\alpha and (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}. By Lemma 8.8 there exists xP<γx\in P_{<\gamma} such that 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right) is not convergent. Such an xx does not belong to PγP_{\gamma}, thus showing that PγP_{\gamma} is a proper subgroup of P<γP_{<\gamma}.

We now prove the assertion about PαP_{\alpha}, as the other assertions are proved in a similar fashion. Recall that Iαα={(;)}I_{\alpha}^{\alpha}=\left\{\left(\varnothing;\varnothing\right)\right\}. Define

H={xGIα:(x(k;αk))kc(G)}.H=\left\{x\in G^{I^{\alpha}}:\left(x\left(k;\alpha_{k}\right)\right)_{k\in\mathbb{N}}\in\mathrm{c}\left(G\right)\right\}\text{.}

By Lemma 8.9 we have that 𝚷30\boldsymbol{\Pi}_{3}^{0} is the complexity class of HH in GIαG^{I^{\alpha}} and of c(G)\mathrm{c}\left(G\right) in GG^{\mathbb{N}}.

By Lemma 8.8 there exists a continuous group homomorphism Φ:GIαP<α\Phi:G^{I^{\alpha}}\rightarrow P_{<\alpha} such that Φ(x)(k;αk)=x(k;αk)\Phi\left(x\right)\left(k;\alpha_{k}\right)=x\left(k;\alpha_{k}\right) for every kk\in\mathbb{N}, and hence Φ1(Pα)=H\Phi^{-1}\left(P_{\alpha}\right)=H. Similarly, the function Ψ:P<αG\Psi:P_{<\alpha}\rightarrow G^{\mathbb{N}}, x𝒙(α)x\mapsto\boldsymbol{x}\left(\alpha\right) is a continuous group homomorphism such that Ψ1(c(G))=Pα\Psi^{-1}\left(\mathrm{c}\left(G\right)\right)=P_{\alpha}. Thus, 𝚷30\boldsymbol{\Pi}_{3}^{0} is the complexity class of PαP_{\alpha} in P<αP_{<\alpha}. ∎

We are now in position to present the proof of Theorem 8.3.

Proof of Theorem 8.3.

By the first assertion in Corollary 8.10 and Proposition 8.7, we have that SαS_{\alpha}, DαD_{\alpha}, PαP_{\alpha} have Solecki rank α+1\alpha+1 in P0P_{0}, and P<αP_{<\alpha} has Solecki rank α\alpha in P0P_{0} if α\alpha is limit. The conclusion now follows by applying Theorem 6.1 and the second assertion in Corollary 8.10. ∎

Recall that a (pseudo-)ultralength function on a group HH is a (pseudo-)length function LL such that L(hh)max{L(h),L(h)}L\left(hh^{\prime}\right)\leq\max\left\{L\left(h\right),L\left(h^{\prime}\right)\right\} for h,hHh,h^{\prime}\in H. A Polish group GG is non-Archimedean if and only if it admits a compatible ultralength function [6, Theorem 2.4.1]. In a similar fashion as above, one can prove the following statement; see also [8, Section 5].

Theorem 8.11.

Let Γ\Gamma be one of the possible complexity classes of non-Archimedean Polishable subgroups from Theorem 1.2. Suppose that GG is a countable discrete group. Then there exists a non-Archimedean CLI Polishable subgroup of GG^{\mathbb{N}} whose complexity class is Γ\Gamma.

Define H:=GH:=G^{\mathbb{N}}. This is a non-Archimedean CLI group. The topology on HH is induced by the ultralength function

LH((gn)n)=exp(min{n:gn1G}).L_{H}\left(\left(g_{n}\right)_{n\in\mathbb{N}}\right)=\exp\left(-\min\left\{n\in\mathbb{N}:g_{n}\neq 1_{G}\right\}\right)\text{.}

Notice that the subgroup c(H)\mathrm{c}\left(H\right) of HH^{\mathbb{N}} convergent sequences is a non-Archimedean CLI Polishable subgroup of HH^{\mathbb{N}} of complexity class 𝚷30\boldsymbol{\Pi}_{3}^{0} with topology induced by the ultralength function

L((hn)n)=max{LH(hn):n}.L_{\infty}\left(\left(h_{n}\right)_{n\in\mathbb{N}}\right)=\max\left\{L_{H}\left(h_{n}\right):n\in\mathbb{N}\right\}\text{.}

The subgroup σ(H)\sigma\left(H\right) of HH^{\mathbb{N}} consisting of sequences (hn)n\left(h_{n}\right)_{n\in\mathbb{N}} such that the sequence (hn(0))n\left(h_{n}\left(0\right)\right)_{n\in\mathbb{N}} in GG is eventually equal to 1G1_{G} is a non-Archimedean CLI Polishable subgroup of HH^{\mathbb{N}} of complexity class 𝚺20\boldsymbol{\Sigma}_{2}^{0}.

Fix α<ω1\alpha<\omega_{1}. We define by recursion on γα\gamma\leq\alpha a decreasing sequence (Fγ)γ<α\left(F_{\gamma}\right)_{\gamma<\alpha} of non-Archimedean Polishable subgroups of HI0αH^{I_{0}^{\alpha}}. We also recursively define, for xFγx\in F_{\gamma} and (n;β)I0β\left(n;\beta\right)\in I_{0}^{\beta}, the values x(n;β)Hx\left(n;\beta\right)\in H. We set F0=HI0αF_{0}=H^{I_{0}^{\alpha}}. If FδF_{\delta} has been defined for every δ<γ\delta<\gamma, define F<γ=δ<γFδF_{<\gamma}=\bigcap_{\delta<\gamma}F_{\delta}, FγF_{\gamma} to contain those xF<γx\in F_{<\gamma} such that, for every (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}, the sequence 𝒙(n;β):=(x(k,n;γk,β))kω\boldsymbol{x}\left(n;\beta\right):=\left(x\left(k,n;\gamma_{k},\beta\right)\right)_{k\in\omega} is convergent in HH. For xFγx\in F_{\gamma} and (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}, we define x(n;β)x\left(n;\beta\right) to be the limit of 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right). Then we have that the non-Archimedean Polish group topology on FγF_{\gamma} is induced by the restriction of the continuous pseudo-ultralength functions on FδF_{\delta} for δ<γ\delta<\gamma together with the pseudo-ultralength function

Lγ(n;β)(x)=L(𝒙(n;β))L_{\gamma}^{\left(n;\beta\right)}\left(x\right)=L_{\infty}\left(\boldsymbol{x}\left(n;\beta\right)\right)

for (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}. This concludes the recursive definition of the non-Archimedean Polishable subgroups FγF_{\gamma} of HI0αH^{I_{0}^{\alpha}} for γα\gamma\leq\alpha. Notice that in particular FαF_{\alpha} contains the elements xF<αx\in F_{<\alpha} such that 𝒙(α):=(x(n,αn))n\boldsymbol{x}\left(\alpha\right):=\left(x\left(n,\alpha_{n}\right)\right)_{n\in\mathbb{N}} belongs to c(H)c\left(H\right). Define ZαZ_{\alpha} to contain those elements xF<αx\in F_{<\alpha} such that 𝒙(α)\boldsymbol{x}\left(\alpha\right) belongs to σ(H)\sigma\left(H\right). The same argument as above, gives the following.

Theorem 8.12.

Adopt the notations above. Suppose that α=1+λ+n\alpha=1+\lambda+n where λ<ω1\lambda<\omega_{1} is either limit or zero and n<ωn<\omega.

  1. (1)

    If n=0n=0 and λ\lambda is limit, then F<λF_{<\lambda} has Solecki rank λ\lambda and complexity class 𝚷λ0\boldsymbol{\Pi}_{\lambda}^{0} in HI0λH^{I_{0}^{\lambda}}.

  2. (2)

    If n=0n=0, then Z1+λZ_{1+\lambda} and F1+λF_{1+\lambda} have Solecki rank λ+1\lambda+1 in HI0αH^{I_{0}^{\alpha}}, and complexity class 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0} and 𝚷1+λ+20\boldsymbol{\Pi}_{1+\lambda+2}^{0}, respectively;

  3. (3)

    if n1n\geq 1, then Z1+λZ_{1+\lambda} and F1+λF_{1+\lambda} have Solecki rank λ+n+1\lambda+n+1 in HI0αH^{I_{0}^{\alpha}}, and complexity class D(𝚷1+λ+n+10)D(\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}) and 𝚷1+λ+n+20\boldsymbol{\Pi}_{1+\lambda+n+2}^{0}, respectively.

9. Fréchetable subspaces

In this and the following section, we assume all the vector spaces to be over the reals. Similar considerations apply to complex vector spaces. Recall that a Fréchet space is a locally convex topological vector space whose topology is given by a complete, translation-invariant metric. Thus, the additive group of a separable Fréchet space is a Polish group. In analogy with the notion of Polishable subgroup of a Polish group, we consider the notion of Fréchetable subspace of a separable Fréchet space.

Definition 9.1.

Suppose that XX is a separable Fréchet space, and YY is a subspace of XX. Then we say that YY is Fréchetable if it is Borel, and there exists a separable Fréchet space topology on YY whose open sets are Borel in XX.

This notion was considered by Saint-Raymond in [18]: a subspace YY of XX is Fréchetable if and only if it has a separable model according to [18, Definition 1]. Notice that a Fréchetable subspace of XX is, in particular, a Polishable subgroup of the additive group of XX. Thus, if it exists, the separable Fréchet space topology on YY as in Definition 9.1, is unique; see also [14, Corollary 4.38]. A subspace YY of a separable Fréchet space XX is Fréchetable if and only if there exists a separable Fréchet space ZZ and a continuous linear map φ:ZX\varphi:Z\rightarrow X with image equal to YY [18, Proposition 4]. If YY is a Fréchetable subspace of XX, then the separable Fréchet space topology on YY is the finest locally convex topological vector space topology on YY that makes all the Borel linear functionals on YY continuous [18, Theoreme 9]. Furthermore, we have a subspace YY of XX is Fréchetable if and only if it is a Polishable subgroup of the additive group of XX, and the Polish topology on YY has a basis of neighborhoods of zero consisting of convex, balanced sets; see [14, Proposition 3.33 and Corollary 3.36]

Lemma 9.2.

Suppose that XX is a separable Fréchet space, and YY a Fréchetable subspace of XX. The first Solecki subgroup s1Y(X)s_{1}^{Y}\left(X\right) of XX relative to YY, where XX and YY are regarded as additive groups, is a Fréchetable subspace of XX.

Proof.

By definition, we have that, for xXx\in X, xYx\in Y if and only if for every open neighborhood VV of zero in YY there exists zYz\in Y such that x+zV¯Gx+z\in\overline{V}^{G}. If xYx\in Y, λ\lambda\in\mathbb{R} is nonzero, and VV is an open neighborhood of zero in YY, then there exists zYz\in Y such that x+zλ1V¯Gx+z\in\overline{\lambda^{-1}V}^{G}, whence λx+λzV¯G\lambda x+\lambda z\in\overline{V}^{G}. This shows that λxs1Y(X)\lambda x\in s_{1}^{Y}\left(X\right), whence s1Y(X)s_{1}^{Y}\left(X\right) is a subspace of XX.

We now show that s1(Y)s_{1}\left(Y\right) is Fréchetable. Since YY is a separable Fréchet space, by the remarks above it has a basis (Vn)nω\left(V_{n}\right)_{n\in\omega} of neighborhoods of zero consisting of convex, balanced sets. Thus, (V¯nGs1Y(X))nω(\overline{V}_{n}^{G}\cap s_{1}^{Y}\left(X\right))_{n\in\omega} is a basis of neighborhoods of zero in s1(Y)s_{1}\left(Y\right) consisting of convex, balanced sets. Thus, s1Y(X)s_{1}^{Y}\left(X\right) is a Fréchetable subspace of XX by the remarks above again. ∎

As an immediate consequence of Lemma 9.2 and Theorem 5.4 by induction on α<ω1\alpha<\omega_{1} we have the following.

Theorem 9.3.

Suppose that XX is a separable Fréchet space, YY is a Fréchetable subspace of XX, and α<ω1\alpha<\omega_{1}. Then the α\alpha-th Solecki subgroup sαY(X)s_{\alpha}^{Y}\left(X\right) of XX relative to YY, where XX and YY are regarded as additive groups, is the smallest 𝚷1+α+10\boldsymbol{\Pi}_{1+\alpha+1}^{0} Fréchetable subspace of XX containing YY.

A similar proof as Theorem 8.1 gives the following.

Theorem 9.4.

Let Γ\Gamma be one of the possible complexity classes of Polishable subgroups from Theorem 1.1. Suppose that XX is a nontrivial separable Fréchet space. Then there exists a Fréchetable subspace of XX^{\mathbb{N}} whose complexity class is Γ\Gamma.

10. Banachable subspaces

Let VV be a separable Fréchet space. A subspace XVX\subseteq V is Banachable if it is the image of a continuous linear map T:ZVT:Z\rightarrow V for some separable Banach space ZZ. Equivalently, XX is a Borel subspace of VV that is also a separable Banach space such that the inclusion map YVY\rightarrow V is continuous. We have that the Solecki subgroups whose index is a successor associated with a Banachable subspace of a separable Fréchet space are also Banachable.

Proposition 10.1.

Suppose that VV is a separable Fréchet space, XVX\subseteq V is Banachable. Then sα+1X(V)Vs_{\alpha+1}^{X}\left(V\right)\subseteq V is Banachable for every α<ω1\alpha<\omega_{1}.

Proof.

It suffices to consider the case α=0\alpha=0. Suppose that X\left\|\cdot\right\|_{X} is a compatible norm on XX and BB is the corresponding unit ball. Define C:=B¯Vs1X(V)C:=\overline{B}^{V}\cap s_{1}^{X}\left(V\right). As (2nB)nω\left(2^{-n}B\right)_{n\in\omega} is a basis of neighborhoods of the identity in XX, (2nC)nω\left(2^{-n}C\right)_{n\in\omega} is a basis of neighborhoods of the identity in s1X(V)s_{1}^{X}\left(V\right). Thus s1X(V)s_{1}^{X}\left(V\right) is a normed space, and hence Banach space (being complete). ∎

In this section we will prove using the methods from Section 8 and Theorem 8.3 the following characterization of the possible complexity class of Banachable subspaces.

Theorem 10.2.

The following is a complete list of all the possible complexity classes of Banachable subspaces of separable Fréchet spaces: 𝚷10\boldsymbol{\Pi}_{1}^{0}, 𝚷1+λ+n+10\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}, D(𝚷1+λ+n0)D(\boldsymbol{\Pi}_{1+\lambda+n}^{0}), and 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0} for λ<ω1\lambda<\omega_{1} either zero or limit and 1n<ω1\leq n<\omega. Furthermore, for every complexity class Γ\Gamma in this list and nontrivial separable Banach space ZZ, there exists a Banachable subspace of c0(,Z)\mathrm{c}_{0}\left(\mathbb{N},Z\right) that has complexity class Γ\Gamma.

We begin with showing that a Banachable subspace of a separable Fréchet space cannot have complexity class 𝚷λ0\boldsymbol{\Pi}_{\lambda}^{0} for some countable limit ordinal λ\lambda.

Proposition 10.3.

Suppose that VV is a separable Fréchet space, and XVX\subseteq V is a Fréchetable subspace. Suppose that sαX(V)s_{\alpha}^{X}\left(V\right) is Banachable for some limit ordinal α\alpha. Then XX has Solecki rank less than α\alpha.

Proof.

Without loss of generality, we can assume that X=sαX(V)X=s_{\alpha}^{X}\left(V\right). Since XX is Banachable, there exists BXB\subseteq X such that (2nB)nω\left(2^{-n}B\right)_{n\in\omega} forms a basis of neighorhoods of zero in XX. Since X=β<αsβX(V)X=\bigcap_{\beta<\alpha}s_{\beta}^{X}\left(V\right), there exists β<α\beta<\alpha and a neighborhood CC of 0 in sβX(V)s_{\beta}^{X}\left(V\right) such that B=CsαX(V)B=C\cap s_{\alpha}^{X}\left(V\right). Thus, XX is endowed with the subspace topology inherited from sβX(V)s_{\beta}^{X}\left(V\right). Whence, XX is closed in sβX(V)s_{\beta}^{X}\left(V\right). Since XX is also dense in sβX(V)s_{\beta}^{X}\left(V\right), we have that X=sβX(V)X=s_{\beta}^{X}\left(V\right). Hence, XX has Solecki rank at most β\beta. ∎

Corollary 10.4.

Suppose that VV is a separable Fréchet space, and XVX\subseteq V is a Banachable subspace. If XX is 𝚷λ0\boldsymbol{\Pi}_{\lambda}^{0} for some limit ordinal λ<ω1\lambda<\omega_{1}, then XX is 𝚷β0\boldsymbol{\Pi}_{\beta}^{0} for some β<λ\beta<\lambda.

Proof.

By Theorem 6.1 we have that X=sλX(V)X=s_{\lambda}^{X}\left(V\right) is Banachable. Thus, by Proposition 10.3 we have that XX has Solecki rank β\beta for some β<λ\beta<\lambda, and hence XX is 𝚷1+β+10\boldsymbol{\Pi}_{1+\beta+1}^{0} by Theorem 6.1 again. ∎

In order to conclude the proof of Theorem 10.2 it remains to prove that all the complexity classes from the statement of Theorem 10.2 can arise. Fix a countable ordinal α\alpha. We adopt the notation from Section 8. We regard IγαI_{\gamma}^{\alpha} as a set fibred over α\alpha, with respect to the map IααI^{\alpha}\rightarrow\alpha, (n;β)γ\left(n;\beta\right)\mapsto\gamma such that (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}. For γ<α\gamma<\alpha we define

Jγα={(k,σ)×(α+1):σk<γ<σα}J_{\gamma}^{\alpha}=\left\{\left(k,\sigma\right)\in\mathbb{N}\times(\alpha+1):\sigma_{k}<\gamma<\sigma\leq\alpha\right\}

We also regard JγαJ_{\gamma}^{\alpha} as a set fibred over α\alpha with respect to the function JγααJ_{\gamma}^{\alpha}\rightarrow\alpha, (k,σ)σ\left(k,\sigma\right)\mapsto\sigma. We then define the fibred product

JγαIα={((k,σ),(n;β)):(k,σ)Jγα,(n;β)Iσα}.J_{\gamma}^{\alpha}\ast I^{\alpha}=\left\{\left(\left(k,\sigma\right),\left(n;\beta\right)\right):\left(k,\sigma\right)\in J_{\gamma}^{\alpha},\left(n;\beta\right)\in I_{\sigma}^{\alpha}\right\}\text{.}

Notice that the projection map JσαIαIαJ_{\sigma}^{\alpha}\ast I^{\alpha}\rightarrow I^{\alpha} is finite-to-one. Indeed, suppose that ((k,σ),(n;β))JγαIα\left(\left(k,\sigma\right),\left(n;\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha}. Then we have that γ<σ\gamma<\sigma, and hence {k:σk<γ}\left\{k\in\mathbb{N}:\sigma_{k}<\gamma\right\} is finite.

Fix a nontrivial separable Banach space ZZ. We denote the norm of zZz\in Z by |z|\left|z\right|. We consider the Banach spaces

1(Z)={(xn)Z:n|xn|<+}\ell_{1}\left(Z\right)=\left\{\left(x_{n}\right)\in Z^{\mathbb{N}}:\sum_{n\in\mathbb{N}}\left|x_{n}\right|<+\infty\right\}

and

bv0(Z)={(xn)Z:n|znzn+1|<+ and (zn)n is vanishing}.\text{{bv}}_{0}\left(Z\right)=\left\{\left(x_{n}\right)\in Z^{\mathbb{N}}:\sum_{n\in\mathbb{N}}\left|z_{n}-z_{n+1}\right|<+\infty\text{ and }\left(z_{n}\right)_{n\in\mathbb{N}}\text{ is vanishing}\right\}\text{.}

Define X0=c0(I0α,Z)X_{0}=\mathrm{c}_{0}\left(I_{0}^{\alpha},Z\right). We now define by recursion on γα\gamma\leq\alpha, Banachable subspaces XγX_{\gamma} and Fréchetable subspaces X<γX_{<\gamma} of X0X_{0} such that XγX<γXδX_{\gamma}\subseteq X_{<\gamma}\subseteq X_{\delta} for δ<γα\delta<\gamma\leq\alpha. Furthermore, for xXγx\in X_{\gamma}, we define the values x(n;β)Zx\left(n;\beta\right)\in Z for (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}, such that the linear functional xx(n;β)x\mapsto x\left(n;\beta\right) on XγX_{\gamma} is continuous. If γ1\gamma\geq 1 and (n;β)Iγα\left(n;\beta\right)\in I_{\gamma}^{\alpha}, then we let 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right) be the convergent sequence (kx(k,n;γk,β))kω\left(kx\left(k,n;\gamma_{k},\beta\right)\right)_{k\in\omega} with limit x(n;β)x\left(n;\beta\right). If (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha}, then we let 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right) be the sequence constantly equal to x(n;β)x\left(n;\beta\right). Suppose that 1γα1\leq\gamma\leq\alpha, and that XδX_{\delta} has been defined for δ<γ\delta<\gamma, in such a way that XδX_{\delta} is a separable Banach space with norm Xδ\left\|\cdot\right\|_{X_{\delta}}.

Define X<γX_{<\gamma} to be the intersection of XδX_{\delta} for δ<γ\delta<\gamma. Consider the continuous linear map

Tγ0:X<γ(Z)IγαT_{\gamma}^{0}:X_{<\gamma}\rightarrow\left(Z^{\mathbb{N}}\right)^{I_{\leq\gamma}^{\alpha}}

defined by

Tγ0(x)=(𝒙(n;β))(n;β)Iγα.T_{\gamma}^{0}\left(x\right)=\left(\boldsymbol{x}\left(n;\beta\right)\right)_{\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}}\text{.}

Consider also the continuous linear map

Tγ1:X<γZJγαIαT_{\gamma}^{1}:X_{<\gamma}\rightarrow Z^{J_{\gamma}^{\alpha}\ast I^{\alpha}}

defined by

Tγ1(x)=(kx(k,n;σk,β))((k,σ),(n;β))JγαIα.T_{\gamma}^{1}\left(x\right)=\left(kx\left(k,n;\sigma_{k},\beta\right)\right)_{\left(\left(k,\sigma\right),\left(n;\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha}}\text{.}

Define XγX<γX_{\gamma}\subseteq X_{<\gamma} to be the intersection of the preimage of

c0(Iγα,c(,Z))(Z)Iγα\mathrm{c}_{0}\left(I_{\leq\gamma}^{\alpha},\mathrm{c}\left(\mathbb{N},Z\right)\right)\subseteq\left(Z^{\mathbb{N}}\right)^{I_{\gamma}^{\alpha}}

under Tγ0T_{\gamma}^{0} and the preimage of

c0(JγαIα,Z)ZJγαIα\mathrm{c}_{0}\left(J_{\gamma}^{\alpha}\ast I^{\alpha},Z\right)\subseteq Z^{J_{\gamma}^{\alpha}\ast I^{\alpha}}

under Tγ1T_{\gamma}^{1}. It follows from Lemma 10.6 below that XγX_{\gamma} is a separable Banach space with respect to the norm

xXγ=max{Tγ0(x)c0(Iγα,c(,Z)),Tγ1(x)c0(JγαIα,Z)}\left\|x\right\|_{X_{\gamma}}=\max\left\{\left\|T_{\gamma}^{0}\left(x\right)\right\|_{\mathrm{c}_{0}(I_{\leq\gamma}^{\alpha},\mathrm{c}\left(\mathbb{N},Z\right))},\left\|T_{\gamma}^{1}\left(x\right)\right\|_{\mathrm{c}_{0}(J_{\gamma}^{\alpha}\ast I^{\alpha},Z)}\right\}

for xXγx\in X_{\gamma}. Observe that in particular

Xα={xX<α:𝒙(α)c(,Z)}X_{\alpha}=\left\{x\in X_{<\alpha}:\boldsymbol{x}\left(\alpha\right)\in\mathrm{c}\left(\mathbb{N},Z\right)\right\}

and

xXα=max{supγ<αxXγ,𝒙(α)c(,Z)}\left\|x\right\|_{X_{\alpha}}=\max\left\{\mathrm{sup}_{\gamma<\alpha}\left\|x\right\|_{X_{\gamma}},\left\|\boldsymbol{x}\left(\alpha\right)\right\|_{\mathrm{c}\left(\mathbb{N},Z\right)}\right\}

for xXαx\in X_{\alpha}, where 𝒙(α):=(kx(k;αk))k\boldsymbol{x}\left(\alpha\right):=\left(kx\left(k;\alpha_{k}\right)\right)_{k\in\mathbb{N}}. Define also SαDαXαS_{\alpha}\subseteq D_{\alpha}\subseteq X_{\alpha} by setting

Sα={xX<α:supγ<αxXγ<+ and 𝒙(α)1(Z)}S_{\alpha}=\left\{x\in X_{<\alpha}:\mathrm{sup}_{\gamma<\alpha}\left\|x\right\|_{X_{\gamma}}<+\infty\text{ and }\boldsymbol{x}\left(\alpha\right)\in\ell_{1}\left(Z\right)\right\}

and

Dα={xX<α:supγ<αxXγ<+ and 𝒙(α)bv0(Z)}D_{\alpha}=\left\{x\in X_{<\alpha}:\mathrm{sup}_{\gamma<\alpha}\left\|x\right\|_{X_{\gamma}}<+\infty\text{ and }\boldsymbol{x}\left(\alpha\right)\in\mathrm{bv}_{0}\left(Z\right)\right\}

where

𝒙(α)=(x(k;αk))k.\boldsymbol{x}\left(\alpha\right)=\left(x\left(k;\alpha_{k}\right)\right)_{k\in\mathbb{N}}\text{.}

Then we have that SαS_{\alpha} is a separable Banach space with respect to the norm

xSα=max{xXα,𝒙(α)1(Z)}\left\|x\right\|_{S_{\alpha}}=\max\left\{\left\|x\right\|_{X_{\alpha}},\left\|\boldsymbol{x}\left(\alpha\right)\right\|_{\ell_{1}\left(Z\right)}\right\}

and DαD_{\alpha} is a separable Banach space with respect to the norm

xDα=max{xXα,𝒙(α)bv0(Z)}.\left\|x\right\|_{D_{\alpha}}=\max\left\{\left\|x\right\|_{X_{\alpha}},\left\|\boldsymbol{x}\left(\alpha\right)\right\|_{\mathrm{bv}_{0}\left(Z\right)}\right\}\text{.}

The existence statement in Theorem 10.2 will be a consequence of the following result.

Theorem 10.5.

Fix α=1+λ+n<ω1\alpha=1+\lambda+n<\omega_{1}, where λ\lambda is a limit ordinal or zero and n<ωn<\omega:

  1. (1)

    if n=0n=0 and λ\lambda is limit, then X<λX_{<\lambda} has Solecki rank λ\lambda in X0X_{0}, and complexity class 𝚷λ0\boldsymbol{\Pi}_{\lambda}^{0};

  2. (2)

    if n=0n=0, then S1+λS_{1+\lambda}, D1+λD_{1+\lambda}, and X1+λX_{1+\lambda} have Solecki rank λ+1\lambda+1 in X0X_{0}, and complexity class 𝚺1+λ+10\boldsymbol{\Sigma}_{1+\lambda+1}^{0}, D(𝚷1+λ+10)D(\boldsymbol{\Pi}_{1+\lambda+1}^{0}), and 𝚷1+λ+10\boldsymbol{\Pi}_{1+\lambda+1}^{0} respectively;

  3. (3)

    if n1n\geq 1, then S1+λ+nS_{1+\lambda+n}, D1+λ+nD_{1+\lambda+n}, and X1+λ+nX_{1+\lambda+n} have Solecki rank λ+n+1\lambda+n+1 in X0X_{0}, and complexity class D(𝚷1+λ+n+10)D(\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}), D(𝚷1+λ+n+10)D(\boldsymbol{\Pi}_{1+\lambda+n+1}^{0}), and 𝚷1+λ+n+10\boldsymbol{\Pi}_{1+\lambda+n+1}^{0} respectively.

The rest of this section contains the proof of Theorem 10.5.

Lemma 10.6.

We have that

xXδxXγ\left\|x\right\|_{X_{\delta}}\leq\left\|x\right\|_{X_{\gamma}}

for δ<γα\delta<\gamma\leq\alpha and xXγx\in X_{\gamma}.

Proof.

It suffices to prove that

Tδ1(x)c0(JδαIα,Z)xXγ.\left\|T_{\delta}^{1}\left(x\right)\right\|_{\mathrm{c}_{0}\left(J_{\delta}^{\alpha}\ast I^{\alpha},Z\right)}\leq\left\|x\right\|_{X_{\gamma}}\text{.}

Suppose that ((k,σ),(n;β))JδαIα\left(\left(k,\sigma\right),\left(n;\beta\right)\right)\in J_{\delta}^{\alpha}\ast I^{\alpha}. Then we have that σk<δ<σ\sigma_{k}<\delta<\sigma and (n;β)Iσα\left(n;\beta\right)\in I_{\sigma}^{\alpha}. Suppose initially that σγ\sigma\leq\gamma. Then we have that (n;β)Iγα\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha} and hence

|kx(k,n;σk,β)|𝒙(n;β)xXγ.\left|kx\left(k,n;\sigma_{k},\beta\right)\right|\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\left\|x\right\|_{X_{\gamma}}\text{.}

Suppose now that γ<σ\gamma<\sigma. Then we have that ((k,σ),(n;β))JγαIα\left(\left(k,\sigma\right),\left(n;\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha} and hence

|kx(k,n;σk,β)|Tγ1(x)c0(JγαIα,)xXγ.\left|kx\left(k,n;\sigma_{k},\beta\right)\right|\leq\left\|T_{\gamma}^{1}\left(x\right)\right\|_{\mathrm{c}_{0}\left(J_{\gamma}^{\alpha}\ast I^{\alpha},\mathbb{R}\right)}\leq\left\|x\right\|_{X_{\gamma}}\text{.}

This concludes the proof. ∎

Lemma 10.7.

Fix γ<α\gamma<\alpha and xXγx\in X_{\gamma}. Let FIγαF\subseteq I_{\leq\gamma}^{\alpha} be a finite set. Define zZI0αz\in Z^{I_{0}^{\alpha}} by setting, for (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha},

z(n;β)={x(n;β)if (n;β)F;0otherwise.z\left(n;\beta\right)=\left\{\begin{array}[]{ll}x\left(n;\beta\right)&\text{if }\left(n;\beta\right)\in F_{\downarrow}\text{;}\\ 0&\text{otherwise.}\end{array}\right.\text{} (2)

Then we have that zSαz\in S_{\alpha} and (2) holds for every (n;β)Iα\left(n;\beta\right)\in I^{\alpha}.

Proof.

We prove by induction on σα\sigma\leq\alpha that zXσz\in X_{\sigma} and (2) holds for every (n;β)Iσα\left(n;\beta\right)\in I_{\sigma}^{\alpha}. Define

F~={(m;γ)Iα:(n;β)F,(n;β)(m;γ)}.\tilde{F}=\left\{\left(m;\gamma\right)\in I^{\alpha}:\exists\left(n;\beta\right)\in F,\left(n;\beta\right)\leq\left(m;\gamma\right)\right\}\text{.}

Case σ=0\sigma=0: We have that (2) holds for (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha} by definition of zz. As xXγx\in X_{\gamma}, for every ε>0\varepsilon>0 there exists a finite subset EE of IγαI_{\leq\gamma}^{\alpha} such that 𝒙(n;β)<ε\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}<\varepsilon for (n;β)IγαE\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E. Thus, if (n;β)I0αE\left(n;\beta\right)\in I_{0}^{\alpha}\setminus E, then

𝒛(n;β)=|z(n;β)||x(n;β)|<ε.\left\|\boldsymbol{z}\left(n;\beta\right)\right\|_{\infty}=\left|z\left(n;\beta\right)\right|\leq\left|x\left(n;\beta\right)\right|<\varepsilon\text{.}

Case 1σγ1\leq\sigma\leq\gamma: Fix (n;β)Iσα\left(n;\beta\right)\in I_{\sigma}^{\alpha}. If (n;β)F\left(n;\beta\right)\in F_{\downarrow} then (k,n;σk,β)F\left(k,n;\sigma_{k},\beta\right)\in F_{\downarrow} for every kk\in\mathbb{N}. Thus, by the inductive hypothesis,

kz(k,n;σk,β)=kx(k,n;σk,β)kz\left(k,n;\sigma_{k},\beta\right)=kx\left(k,n;\sigma_{k},\beta\right)

for every kk\in\mathbb{N}, and hence

𝒛(n;β)=𝒙(n;β).\boldsymbol{z}\left(n;\beta\right)=\boldsymbol{x}\left(n;\beta\right)\text{.}

Since by assumption 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right) is a convergent sequence with limit x(n;β)x\left(n;\beta\right), we have that 𝒛(n;β)\boldsymbol{z}\left(n;\beta\right) is a convergent sequence with limit z(n;β)=x(n;β)z\left(n;\beta\right)=x\left(n;\beta\right). If (n;β)F\left(n;\beta\right)\notin F_{\downarrow} then there exists NN\in\mathbb{N} such that for k>Nk>N, (k,n;σk,β)F\left(k,n;\sigma_{k},\beta\right)\notin F_{\downarrow}. By the inductive hypothesis, we have that kz(k,n;σk,β)=0kz\left(k,n;\sigma_{k},\beta\right)=0 for k>Nk>N. Thus, 𝒛(n;β)\boldsymbol{z}\left(n;\beta\right) is a sequence eventually zero with limit z(n;β)=0z\left(n;\beta\right)=0.

Fix ε>0\varepsilon>0. Since xXγx\in X_{\gamma}, there exist a finite subset EIγαE\subseteq I_{\leq\gamma}^{\alpha} such that 𝒙(n;β)<ε\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}<\varepsilon for (n;β)IγαE\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E, and a finite subset EJγαIαE^{\prime}\subseteq J_{\gamma}^{\alpha}\ast I^{\alpha} such that |kx(k,n;τk,β)|<ε\left|kx\left(k,n;\tau_{k},\beta\right)\right|<\varepsilon for ((k,τ),(n;β))(JγαIα)E\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\gamma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime}. Define

E′′=E{((k,τ),(n;β))(JσαIα):(n;β)E}E^{\prime\prime}=E^{\prime}\cup\left\{\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\sigma}^{\alpha}\ast I^{\alpha}\right):\left(n;\beta\right)\in E\right\}

If (n;β)IσαE\left(n;\beta\right)\in I_{\leq\sigma}^{\alpha}\setminus E then we have that

𝒛(n;β)𝒙(n;β)ε.\left\|\boldsymbol{z}\left(n;\beta\right)\right\|_{\infty}\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\varepsilon\text{.}

If ((k,τ),(n;β))(JσαIα)E′′\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\sigma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime\prime} then we have that τk<σ<τ\tau_{k}<\sigma<\tau and (n;β)Iτα\left(n;\beta\right)\in I_{\tau}^{\alpha}. If τγ\tau\leq\gamma then we have that (n;β)IγαE\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E and hence

|kz(k,n;τk,β)|𝒙(n;β)ε.\left|kz\left(k,n;\tau_{k},\beta\right)\right|\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\varepsilon\text{.}

If γ<τ\gamma<\tau then ((k,τ),(n;β))(JγαI)E\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\gamma}^{\alpha}\ast I\right)\setminus E^{\prime} and hence

|kz(k,n;τk,β)||kx(k,n;τk,β)|ε.\left|kz\left(k,n;\tau_{k},\beta\right)\right|\leq\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\varepsilon\text{.}

Case σ>γ\sigma>\gamma: Fix (n;β)Iσα\left(n;\beta\right)\in I_{\sigma}^{\alpha}. Then by the inductive assumption we have that z(k,n;σk,β)=0z\left(k,n;\sigma_{k},\beta\right)=0 for k>Nk>N. Thus, the sequence 𝒛(n;β)\boldsymbol{z}\left(n;\beta\right) is eventually zero, and z(n;β)=0z\left(n;\beta\right)=0.

Fix ε>0\varepsilon>0. Since xXγx\in X_{\gamma}, there exist a finite set EIγαE\subseteq I_{\leq\gamma}^{\alpha} such that

𝒙(n;β)<ε\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}<\varepsilon

for (n;β)IγαE\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E, and a finite set E(JγαIα)E^{\prime}\subseteq\left(J_{\gamma}^{\alpha}\ast I^{\alpha}\right) such that

|kx(k,n;τk,β)|<ε\left|kx\left(k,n;\tau_{k},\beta\right)\right|<\varepsilon

for ((k,τ),(n;β))(JγαIα)E\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\gamma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime}. Define

E~=EF~\tilde{E}=E\cup\tilde{F}
E~=F~E{((k,τ),(n;β))JσαIα:(n;β)E~}.\tilde{E}^{\prime}=\tilde{F}\cup E^{\prime}\cup\left\{\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in J_{\sigma}^{\alpha}\ast I^{\alpha}:\left(n;\beta\right)\in\tilde{E}\right\}\text{.}

Fix (n;β)IσαE~\left(n;\beta\right)\in I_{\leq\sigma}^{\alpha}\setminus\tilde{E}. Fix δσ\delta\leq\sigma such that (n;β)Iδα\left(n;\beta\right)\in I_{\delta}^{\alpha}. If δγ\delta\leq\gamma, then we have that (m;β)IγαE\left(m;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E and hence

𝒛(n;β)𝒙(n;β)ε.\left\|\boldsymbol{z}\left(n;\beta\right)\right\|_{\infty}\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\varepsilon\text{.}

Suppose that δ>γ\delta>\gamma, and fix kk\in\mathbb{N}. If z(k,n;δk,β)0z\left(k,n;\delta_{k},\beta\right)\neq 0 then we have that (k,n;δk,β)F\left(k,n;\delta_{k},\beta\right)\in F, whence (n;β)F~E~\left(n;\beta\right)\in\tilde{F}\subseteq\tilde{E}, contradicting the hypothesis. Thus, 𝒛(n;β)\boldsymbol{z}\left(n;\beta\right) is the sequence constantly equal to zero. Fix ((k,τ),(n;β))(JσαIα)E~\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\sigma}^{\alpha}\ast I^{\alpha}\right)\setminus\tilde{E}^{\prime}. Thus, τk<σ<τ\tau_{k}<\sigma<\tau and (n;β)Iτα\left(n;\beta\right)\in I_{\tau}^{\alpha}. If τk<γ\tau_{k}<\gamma, then ((k,τ),(n;β))(JγαIα)E\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\gamma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime} and hence

|kz(k,n;τk,β)||kx(k,n;τk,β)|ε.\left|kz\left(k,n;\tau_{k},\beta\right)\right|\leq\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\varepsilon\text{.}

Suppose that τk=γ\tau_{k}=\gamma. If z(k,n;τk,β)z\left(k,n;\tau_{k},\beta\right) is nonzero, then (k,n;τk,β)F\left(k,n;\tau_{k},\beta\right)\in F and hence (n;β)F~E~\left(n;\beta\right)\in\tilde{F}\subseteq\tilde{E}, contradicting the assumption that (k,n;τk,β)E~\left(k,n;\tau_{k},\beta\right)\notin\tilde{E}^{\prime}. If τk>γ\tau_{k}>\gamma then we have that (k,n;τk,β)F\left(k,n;\tau_{k},\beta\right)\notin F_{\downarrow} and hence z(k,n;τk,β)=0z\left(k,n;\tau_{k},\beta\right)=0. This concludes the inductive proof.

Finally, to see that zSαz\in S_{\alpha} observe that if N=max{k:αkγ}N=\max\left\{k\in\mathbb{N}:\alpha_{k}\leq\gamma\right\} then we have that

k|z(k;αk)|kN|x(k;αk)|<+.\sum_{k\in\mathbb{N}}\left|z\left(k;\alpha_{k}\right)\right|\leq\sum_{k\leq N}\left|x\left(k;\alpha_{k}\right)\right|<+\infty\text{.}

The next lemma is similar to the previous one, with the difference that the finite set FF is supposed to be a subset of I<γαI_{<\gamma}^{\alpha} instead of IγαI_{\leq\gamma}^{\alpha}.

Lemma 10.8.

Fix γ<α\gamma<\alpha and xXγx\in X_{\gamma}. Let FI<γαF\subseteq I_{<\gamma}^{\alpha} be a finite set. Define zZI0αz\in Z^{I_{0}^{\alpha}} by setting, for (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha},

z(n;β)={x(n;β)if (n;β)F;0otherwise.z\left(n;\beta\right)=\left\{\begin{array}[]{ll}x\left(n;\beta\right)&\text{if }\left(n;\beta\right)\in F_{\downarrow}\text{;}\\ 0&\text{otherwise.}\end{array}\text{}\right. (3)

Then we have that zSαz\in S_{\alpha}, zXαxXγ\left\|z\right\|_{X_{\alpha}}\leq\left\|x\right\|_{X_{\gamma}}, and furthermore (3) holds for every (n;β)Iα\left(n;\beta\right)\in I^{\alpha}. Furthermore

zSαmax{kNx(k;αk),xXγ}\left\|z\right\|_{S_{\alpha}}\leq\max\left\{\sum_{k\leq N}\left\|x\left(k;\alpha_{k}\right)\right\|,\left\|x\right\|_{X_{\gamma}}\right\}

where N=max{k:αk<γ}N=\max\left\{k\in\mathbb{N}:\alpha_{k}<\gamma\right\}.

Proof.

It follows from Lemma 10.7 that zSαz\in S_{\alpha} and (3) holds for every (n;β)Iσα\left(n;\beta\right)\in I_{\sigma}^{\alpha}. We now prove by induction on σα\sigma\leq\alpha that zXσxXγ\left\|z\right\|_{X_{\sigma}}\leq\left\|x\right\|_{X_{\gamma}}. Suppose that the conclusion holds for every δ<σ\delta<\sigma.

Case σ=0\sigma=0: If (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha}, then we have that |z(n;β)||x(n;β)|\left|z\left(n;\beta\right)\right|\leq\left|x\left(n;\beta\right)\right|. This shows that zX0xX0xXγ\left\|z\right\|_{X_{0}}\leq\left\|x\right\|_{X_{0}}\leq\left\|x\right\|_{X_{\gamma}}.

Case 1σγ1\leq\sigma\leq\gamma: For (n;β)Iσα\left(n;\beta\right)\in I_{\leq\sigma}^{\alpha}, we have

𝒛(n;β)𝒙(n;β)xXγ.\left\|\boldsymbol{z}\left(n;\beta\right)\right\|_{\infty}\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\left\|x\right\|_{X_{\gamma}}\text{.}

Fix ((k,τ),(n;β))JσαIα\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in J_{\sigma}^{\alpha}\ast I^{\alpha}. Thus, we have that τk<σ<τ\tau_{k}<\sigma<\tau and (n;β)Iτα\left(n;\beta\right)\in I_{\tau}^{\alpha}. If τγ\tau\leq\gamma, then (n;β)Iγα\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha} and hence

|kz(k,n;τk,β)||kx(k,n;τk,β)|𝒙(n;β)xXγ.\left|kz\left(k,n;\tau_{k},\beta\right)\right|\leq\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\left\|x\right\|_{X_{\gamma}}\text{.}

If γ<τ\gamma<\tau, then ((k,τ),(n;β))JγαIα\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha} and hence

|kz(k,n;τk,β)||kx(k,n;τk,β)|xXγ.\left|kz\left(k,n;\tau_{k},\beta\right)\right|\leq\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\left\|x\right\|_{X_{\gamma}}\text{.}

Case σ>γ\sigma>\gamma. Suppose that (n;β)Iδα\left(n;\beta\right)\in I_{\delta}^{\alpha} for some δσ\delta\leq\sigma. If δγ\delta\leq\gamma, then

𝒛(n;β)𝒙(n;β)xXγ.\left\|\boldsymbol{z}\left(n;\beta\right)\right\|_{\infty}\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\left\|x\right\|_{X_{\gamma}}\text{.}

If γ<δ\gamma<\delta then for every kk\in\mathbb{N} we have that either δk<γ\delta_{k}<\gamma, in which case ((k,δ),(n;β))JγαIα\left(\left(k,\delta\right),\left(n;\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha} and hence

|kz(k,n;δk,β)||kx(k,n;δk,β)|xXγ\left|kz\left(k,n;\delta_{k},\beta\right)\right|\leq\left|kx\left(k,n;\delta_{k},\beta\right)\right|\leq\left\|x\right\|_{X_{\gamma}}

or γδk\gamma\leq\delta_{k}, in which case (k,n;δk,β)F\left(k,n;\delta_{k},\beta\right)\notin F_{\downarrow} and

z(k,n;δk,β)=0.z\left(k,n;\delta_{k},\beta\right)=0\text{.}

Thus

𝒛(n;β)xXγ.\left\|\boldsymbol{z}\left(n;\beta\right)\right\|_{\infty}\leq\left\|x\right\|_{X_{\gamma}}\text{.}

Suppose now that ((k,τ),(n,β))JσαIα\left(\left(k,\tau\right),\left(n,\beta\right)\right)\in J_{\sigma}^{\alpha}\ast I^{\alpha}. Thus τk<σ<τ\tau_{k}<\sigma<\tau. If τk<γ\tau_{k}<\gamma then we have that ((k,τ),(n,β))JγαIα\left(\left(k,\tau\right),\left(n,\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha} and hence

|kz(k,n;τk,β)||kx(k,n;τk,β)|xXγ\left|kz\left(k,n;\tau_{k},\beta\right)\right|\leq\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\left\|x\right\|_{X_{\gamma}}

If γτk\gamma\leq\tau_{k} then (k,n;τk,β)F\left(k,n;\tau_{k},\beta\right)\notin F_{\downarrow} and hence

z(k,n;τk,β)=0.z\left(k,n;\tau_{k},\beta\right)=0\text{.}

This concludes the inductive proof that zXσxXγ\left\|z\right\|_{X_{\sigma}}\leq\left\|x\right\|_{X_{\gamma}} for every σα\sigma\leq\alpha.

Finally, we have that

k|z(k;αk)|kN|x(k;αk)|.\sum_{k\in\mathbb{N}}\left|z\left(k;\alpha_{k}\right)\right|\leq\sum_{k\leq N}\left|x\left(k;\alpha_{k}\right)\right|\text{.}

This shows that zSαz\in S_{\alpha} and

zSα=max{zXα,k|z(k;αk)|}max{xXα,kN|x(k;αk)|}.\left\|z\right\|_{S_{\alpha}}=\max\left\{\left\|z\right\|_{X_{\alpha}},\sum_{k\in\mathbb{N}}\left|z\left(k;\alpha_{k}\right)\right|\right\}\leq\max\left\{\left\|x\right\|_{X_{\alpha}},\sum_{k\leq N}\left|x\left(k;\alpha_{k}\right)\right|\right\}\text{.}

This concludes the proof. ∎

Lemma 10.9.

For every γ<α\gamma<\alpha, SαS_{\alpha} is dense in XγX_{\gamma}.

Proof.

Suppose that xXγx\in X_{\gamma}, and ε>0\varepsilon>0. We need to prove that there exists zSαz\in S_{\alpha} such that xzXγε\left\|x-z\right\|_{X_{\gamma}}\leq\varepsilon. Since xXγx\in X_{\gamma}, there exists a finite subset EIγαE\subseteq I_{\leq\gamma}^{\alpha} such that 𝒙(n;β)<ε\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}<\varepsilon for (n;β)IγαE\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E, and a finite subset EJγαIαE^{\prime}\subseteq J_{\gamma}^{\alpha}\ast I^{\alpha} such that |kx(k,n;τk,β)|<ε\left|kx\left(k,n;\tau_{k},\beta\right)\right|<\varepsilon for ((k,τ),(n;β))(JγαIα)E\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\gamma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime}.

Suppose that zXαz\in X_{\alpha} is obtained from xXγx\in X_{\gamma} and

F=(EIγα){(k,n;γk,β):((k,τ),(n;β))E}F=\left(E\cap I_{\leq\gamma}^{\alpha}\right)\cup\left\{\left(k,n;\gamma_{k},\beta\right):\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in E^{\prime}\right\}

as in Lemma 10.7.

Fix (n;β)Iγα\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}. If (n;β)F\left(n;\beta\right)\in F_{\downarrow} then 𝒛(n;β)=𝒙(n;β)\boldsymbol{z}\left(n;\beta\right)=\boldsymbol{x}\left(n;\beta\right), while if (n;β)F\left(n;\beta\right)\notin F_{\downarrow}, then

𝒛(n;β)𝒙(n;β)𝒙(n;β)ε.\left\|\boldsymbol{z}\left(n;\beta\right)-\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\varepsilon\text{.}

Consider ((k,τ),(n;β))JγαIα\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha}. Thus τk<γ<τ\tau_{k}<\gamma<\tau and (n;β)Iτα\left(n;\beta\right)\in I_{\tau}^{\alpha}. If (k,n;τk,β)F\left(k,n;\tau_{k},\beta\right)\in F then kz(k,n;τk,β)=kx(k,n;τk,β)kz\left(k,n;\tau_{k},\beta\right)=kx\left(k,n;\tau_{k},\beta\right). If (k,n;τk,β)F\left(k,n;\tau_{k},\beta\right)\notin F, then (k,n;τk,β)(JγαIα)E\left(k,n;\tau_{k},\beta\right)\in\left(J_{\gamma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime} and hence

|kz(k,n;τk,β)kx(k,n;τk,β)||kx(k,n;τk,β)|ε.\left|kz\left(k,n;\tau_{k},\beta\right)-kx\left(k,n;\tau_{k},\beta\right)\right|\leq\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\varepsilon\text{.}

This concludes the proof that zxXγε\left\|z-x\right\|_{X_{\gamma}}\leq\varepsilon. ∎

Lemma 10.10.

Fix γ<α\gamma<\alpha. If VV is a neighborhood of zero in SαS_{\alpha}, then V¯X<γXγ\overline{V}^{X_{<\gamma}}\cap X_{\gamma} contains an open neighborhood of zero in XγX_{\gamma}.

Proof.

Define

N=max{k:αkγ}.N=\max\left\{k\in\mathbb{N}:\alpha_{k}\leq\gamma\right\}\text{.}

Suppose that VV is a neighborhood of zero in XαX_{\alpha}. Without loss of generality, we can assume that

V={zSα:zXαε,k|z(k;αk)|ε}.V=\left\{z\in S_{\alpha}:\left\|z\right\|_{X_{\alpha}}\leq\varepsilon,\sum_{k\in\mathbb{N}}\left|z\left(k;\alpha_{k}\right)\right|\leq\varepsilon\right\}\text{.}

We claim that V¯X<γXγ\overline{V}^{X_{<\gamma}}\cap X_{\gamma} contains

W:={xXγ:xXγε,kN|x(k;αk)|ε}.W:=\left\{x\in X_{\gamma}:\left\|x\right\|_{X_{\gamma}}\leq\varepsilon,\sum_{k\leq N}\left|x\left(k;\alpha_{k}\right)\right|\leq\varepsilon\right\}\text{.}

Indeed, suppose that xWx\in W. Let UU be an open neighborhood of xx in X<γX_{<\gamma}. Without loss of generality, we can assume that

U={zX<γ:xzXδε1}U=\left\{z\in X_{<\gamma}:\left\|x-z\right\|_{X_{\delta}}\leq\varepsilon_{1}\right\}

for some δ<γ\delta<\gamma and ε1>0\varepsilon_{1}>0. We need to prove that UVU\cap V\neq\varnothing.

Since xXγx\in X_{\gamma}, there exists a finite subset EE of XγX_{\gamma} such that 𝒙(n;β)ε1\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\varepsilon_{1} for (n;β)IγαE\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E, and a finite subset EE^{\prime} of JγαIαJ_{\gamma}^{\alpha}\ast I^{\alpha} such that |kx(k,n;τk,β)|ε1\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\varepsilon_{1} for ((k,τ),(n;β))(JγαIα)E\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\gamma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime}. Define

E′′=E{((k,τ),(n;β))JγαIα:(n;β)E}E^{\prime\prime}=E^{\prime}\cup\left\{\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha}:\left(n;\beta\right)\in E\right\}

Let zXαz\in X_{\alpha} be obtained from xx and

F:=(EIδα){(k,n;τk,β):((k,τ),(n;β))E′′}{(k;αk):kN}F:=\left(E\cap I_{\leq\delta}^{\alpha}\right)\cup\left\{\left(k,n;\tau_{k},\beta\right):\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in E^{\prime\prime}\right\}\cup\left\{\left(k;\alpha_{k}\right):k\leq N\right\}

as in Lemma 10.8. Then we have that zSαz\in S_{\alpha},

zXαxXγε,\left\|z\right\|_{X_{\alpha}}\leq\left\|x\right\|_{X_{\gamma}}\leq\varepsilon\text{,}

and

kN|z(k;αk)|kN|x(k;αk)|ε\sum_{k\leq N}\left|z\left(k;\alpha_{k}\right)\right|\leq\sum_{k\leq N}\left|x\left(k;\alpha_{k}\right)\right|\leq\varepsilon

and hence zVz\in V. It remains to prove that zxXδε1\left\|z-x\right\|_{X_{\delta}}\leq\varepsilon_{1}. For (n;β)Iδα\left(n;\beta\right)\in I_{\leq\delta}^{\alpha}, if (n;β)F\left(n;\beta\right)\in F then

𝒛(n;β)=𝒙(n;β)\boldsymbol{z}\left(n;\beta\right)=\boldsymbol{x}\left(n;\beta\right)

while if (n;β)F\left(n;\beta\right)\notin F, then (n;β)IγαE\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E and we have that

𝒛(n;β)𝒙(n;β)𝒙(n;β)ε1\left\|\boldsymbol{z}\left(n;\beta\right)-\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\varepsilon_{1}

by the choice of EE.

For ((k,τ),(n;β))JδαIα\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in J_{\delta}^{\alpha}\ast I^{\alpha}, we have that τk<δ<τ\tau_{k}<\delta<\tau and (n;β)Iτα\left(n;\beta\right)\in I_{\tau}^{\alpha}. Suppose that γ<τ\gamma<\tau, in which case we have that ((k,τ),(n;β))JγαIα\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in J_{\gamma}^{\alpha}\ast I^{\alpha}. If ((k,τ),(n;β))F\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in F, then

kz(k,n;τk,β)=kx(k,n;τk,β);kz\left(k,n;\tau_{k},\beta\right)=kx\left(k,n;\tau_{k},\beta\right)\text{;}

if ((k,τ),(n;β))F\left(\left(k,\tau\right),\left(n;\beta\right)\right)\notin F then ((k,τ),(n;β))(JγαIα)E\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in\left(J_{\gamma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime} and hence

|kz(k,n;τk,β)kx(k,n;τk,β)|=|kx(k,n;τk,β)|ε1.\left|kz\left(k,n;\tau_{k},\beta\right)-kx\left(k,n;\tau_{k},\beta\right)\right|=\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\varepsilon_{1}\text{.}

Suppose now that τγ\tau\leq\gamma, in which case τk<δ<τγ\tau_{k}<\delta<\tau\leq\gamma. If ((k,τ),(n;β))F\left(\left(k,\tau\right),\left(n;\beta\right)\right)\in F, then

kz(k,n;τk,β)=kx(k,n;τk,β);kz\left(k,n;\tau_{k},\beta\right)=kx\left(k,n;\tau_{k},\beta\right)\text{;}

while if ((k,τ),(n;β))F\left(\left(k,\tau\right),\left(n;\beta\right)\right)\notin F, then we have that (n;β)IγαE\left(n;\beta\right)\in I_{\leq\gamma}^{\alpha}\setminus E and hence

|kz(k,n;τk,β)kx(k,n;τk,β)||kx(k,n;τk,β)|𝒙(n;β)ε1.\left|kz\left(k,n;\tau_{k},\beta\right)-kx\left(k,n;\tau_{k},\beta\right)\right|\leq\left|kx\left(k,n;\tau_{k},\beta\right)\right|\leq\left\|\boldsymbol{x}\left(n;\beta\right)\right\|_{\infty}\leq\varepsilon_{1}\text{.}

This concludes the proof that zxXδε1\left\|z-x\right\|_{X_{\delta}}\leq\varepsilon_{1}. ∎

Using Lemma 10.10 and Lemma 10.9, one can prove Proposition 10.11, similarly as Proposition 8.7 is proved from Lemma 8.6 and Lemma 8.5

Proposition 10.11.

For γ<α\gamma<\alpha we have that

sγSα(X0)=sγDα(X0)=sγXα(X0)=sγX<α(X0)=X<(1+γ)s_{\gamma}^{S_{\alpha}}(X_{0})=s_{\gamma}^{D_{\alpha}}(X_{0})=s_{\gamma}^{X_{\alpha}}(X_{0})=s_{\gamma}^{X_{<\alpha}}(X_{0})=X_{<\left(1+\gamma\right)}

Recall that, for (n;β)\left(n;\beta\right) and (m;τ)\left(m;\tau\right) in IαI^{\alpha}, we define (n;β)(m;τ)\left(n;\beta\right)\leq\left(m;\tau\right) if and only if there exist γ0γ1α\gamma_{0}\leq\gamma_{1}\leq\alpha such that (n;β)Iγ0α\left(n;\beta\right)\in I_{\gamma_{0}}^{\alpha}, (m;τ)Iγ1α\left(m;\tau\right)\in I_{\gamma_{1}}^{\alpha}, mm is a tail of nn, and τ\tau is a tail of β\beta, i.e. we have that, for some d<ω\ell\leq d<\omega, (n;β)=(n0,,nd;β0,,βd)\left(n;\beta\right)=\left(n_{0},\ldots,n_{d};\beta_{0},\ldots,\beta_{d}\right), and (m;τ)=(nd,,nd;βd,,βd)\left(m;\tau\right)=\left(n_{d-\ell},\ldots,n_{d};\beta_{d-\ell},\ldots,\beta_{d}\right). In this case, we set

π(m;τ)(n;β):=1n0nd1.\pi_{\left(m;\tau\right)}^{\left(n;\beta\right)}:=\frac{1}{n_{0}\cdots n_{d-\ell-1}}\text{.}
Lemma 10.12.

Fix γα\gamma\leq\alpha and (m;τ)Iγα\left(m;\tau\right)\in I_{\gamma}^{\alpha}. There exists a continuous group homomorphism Φ:c0(,Z)X<γ\Phi:\mathrm{c}_{0}\left(\mathbb{N},Z\right)\rightarrow X_{<\gamma} such that Φ(t)(k,m;γk,τ)=tk\Phi\left(t\right)\left(k,m;\gamma_{k},\tau\right)=t_{k} for every tc0(,Z)t\in\mathrm{c}_{0}\left(\mathbb{N},Z\right) and kk\in\mathbb{N}.

Proof.

For ε>0\varepsilon>0, let KεK_{\varepsilon}\in\mathbb{N} be such that, for k>Kεk>K_{\varepsilon} one has that |tk|ε\left|t_{k}\right|\leq\varepsilon. For t(Z)Iγαt\in\left(Z^{\mathbb{N}}\right)^{I_{\gamma}^{\alpha}}, define Φ(t):=xZI0α\Phi\left(t\right):=x\in Z^{I_{0}^{\alpha}} by setting, for (n;β)I0α\left(n;\beta\right)\in I_{0}^{\alpha},

x(n;β):={π(n;β)(k,m;γk,τ)tkif (n;β)(k,m;γk,τ) for some k;0otherwise.x\left(n;\beta\right):=\left\{\begin{array}[]{ll}\pi_{\left(n;\beta\right)}^{\left(k,m;\gamma_{k},\tau\right)}t_{k}&\text{if }\left(n;\beta\right)\leq\left(k,m;\gamma_{k},\tau\right)\text{ for some }k\in\mathbb{N}\text{;}\\ 0&\text{otherwise.}\end{array}\right.\text{} (4)

It is clear that Φ:ZZI0α\Phi:Z^{\mathbb{N}}\rightarrow Z^{I_{0}^{\alpha}} is a continuous group homomorphism. We now prove by induction on σ<γ\sigma<\gamma that xXσx\in X_{\sigma}, and that (4) holds for every (n;β)Iα\left(n;\beta\right)\in I^{\alpha}. Suppose that the conclusion holds for all δ<σ\delta<\sigma.

Case σ=0\sigma=0: We need to prove that xc0(I0α,Z)x\in\mathrm{c}_{0}\left(I_{0}^{\alpha},Z\right). Fix ε>0\varepsilon>0. Consider

F={(n;β)I0α:(n;β)(k,m;γk,τ) for some kKε}F=\left\{\left(n;\beta\right)\in I_{0}^{\alpha}:\left(n;\beta\right)\leq\left(k,m;\gamma_{k},\tau\right)\text{ for some }k\leq K_{\varepsilon}\right\}

Then FI0αF\subseteq I_{0}^{\alpha} is finite and for (n;β)I0αF\left(n;\beta\right)\in I_{0}^{\alpha}\setminus F one has that either

x(n;β)=0x\left(n;\beta\right)=0

or (n;β)(k,m;γk,τ)\left(n;\beta\right)\leq\left(k,m;\gamma_{k},\tau\right) for some k>Kεk>K_{\varepsilon}, in which case

|x(n;β)|=|π(n;β)(k,m;γk,τ)tk||tk|ε\left|x\left(n;\beta\right)\right|=\left|\pi_{\left(n;\beta\right)}^{\left(k,m;\gamma_{k},\tau\right)}t_{k}\right|\leq\left|t_{k}\right|\leq\varepsilon

Case 1σ<γ1\leq\sigma<\gamma: Fix (n;β)Iδα\left(n;\beta\right)\in I_{\delta}^{\alpha} for some δσ\delta\leq\sigma. If (n;β)(k,m;γk,τ)\left(n;\beta\right)\leq\left(k,m;\gamma_{k},\tau\right) for some kk\in\mathbb{N}, then we have that for every \ell\in\mathbb{N}, (,n;δ,β)(k,m;γk,τ)\left(\ell,n;\delta_{\ell},\beta\right)\leq\left(k,m;\gamma_{k},\tau\right). Thus,

x(,n;δ,β)=π(,n;δ,β)(k,m;γk,τ)tkx\left(\ell,n;\delta_{\ell},\beta\right)=\pi_{\left(\ell,n;\delta_{\ell},\beta\right)}^{\left(k,m;\gamma_{k},\tau\right)}t_{k}

and

x(,n;δ,β)=π(n;β)(k,m;γk,τ)tk\ell x\left(\ell,n;\delta_{\ell},\beta\right)=\pi_{\left(n;\beta\right)}^{\left(k,m;\gamma_{k},\tau\right)}t_{k}

Thus, the sequence 𝒙(n;β)\boldsymbol{x}\left(n;\beta\right) is constantly equal to π(n;β)(k,m;γk,τ)tk\pi_{\left(n;\beta\right)}^{\left(k,m;\gamma_{k},\tau\right)}t_{k}. This shows that

x(n;β)=π(n;β)(k,m;γk,τ)tk.x\left(n;\beta\right)=\pi_{\left(n;\beta\right)}^{\left(k,m;\gamma_{k},\tau\right)}t_{k}\text{.}

Suppose that there does not exist kk\in\mathbb{N} such that (n;β)(k,m;γk,τ)\left(n;\beta\right)\leq\left(k,m;\gamma_{k},\tau\right). Fix \ell\in\mathbb{N}. If (,n;δ,β)(k,m;γk,τ)\left(\ell,n;\delta_{\ell},\beta\right)\leq\left(k,m;\gamma_{k},\tau\right) for some kk\in\mathbb{N}, then we have that (k,m)\left(k,m\right) is a tail of (,n)\left(\ell,n\right) and (γk,τ)\left(\gamma_{k},\tau\right) is a tail of (δ,β)\left(\delta_{\ell},\beta\right). If the length of (k,m)\left(k,m\right) is strictly less than the length of (,n)\left(\ell,n\right), then mm is a tail of nn and τ\tau is a tail of β\beta, and hence (n;β)(k,m;γk,τ)\left(n;\beta\right)\leq\left(k,m;\gamma_{k},\tau\right), contradicting the assumption. Therefore, we have that (,n;δ,β)=(k,m;γk,τ)\left(\ell,n;\delta_{\ell},\beta\right)=\left(k,m;\gamma_{k},\tau\right). In particular, we have that (n;β)=(m;τ)Iγα\left(n;\beta\right)=\left(m;\tau\right)\in I_{\gamma}^{\alpha} contradicting the assumption that (n;β)Iδα\left(n;\beta\right)\in I_{\delta}^{\alpha} and δσ<γ\delta\leq\sigma<\gamma. Thus, the sequence 𝒛(n;β)\boldsymbol{z}\left(n;\beta\right) is constantly zero, and hence z(n;β)=0z\left(n;\beta\right)=0.

We now prove that xXσx\in X_{\sigma}. Fix ε>0\varepsilon>0. Define

N=max{Kε,max{k:γkσ}}.N=\max\{K_{\varepsilon},\mathrm{\max}\left\{k\in\mathbb{N}:\gamma_{k}\leq\sigma\right\}\}\text{.}

Consider

E={(n;β)Iσα:(n;β)(k,m;γk,τ) for some kN}.E=\left\{\left(n;\beta\right)\in I_{\leq\sigma}^{\alpha}:\left(n;\beta\right)\leq\left(k,m;\gamma_{k},\tau\right)\text{ for some }k\leq N\right\}\text{.}

If (n;β)IσαE\left(n;\beta\right)\in I_{\leq\sigma}^{\alpha}\setminus E and x(n;β)0x\left(n;\beta\right)\neq 0 then, by the argument above, (n;β)(k,m;γk,τ)\left(n;\beta\right)\leq\left(k,m;\gamma_{k},\tau\right) for some k>NKεk>N\geq K_{\varepsilon} and hence

|x(n;β)||tk|ε.\left|x\left(n;\beta\right)\right|\leq\left|t_{k}\right|\leq\varepsilon\text{.}

Consider the finite set

E={((,ρ);(n;β))JσαIα:(,n;ρ,β)E}E^{\prime}=\left\{\left(\left(\ell,\rho\right);\left(n;\beta\right)\right)\in J_{\sigma}^{\alpha}\ast I^{\alpha}:\left(\ell,n;\rho_{\ell},\beta\right)\in E\right\}

If (,ρ;(n;β))(JσαIα)E\left(\ell,\rho;\left(n;\beta\right)\right)\in\left(J_{\sigma}^{\alpha}\ast I^{\alpha}\right)\setminus E^{\prime} and x(,n;ρ,β)0x\left(\ell,n;\rho_{\ell},\beta\right)\neq 0, then ρ<σ<ρ\rho_{\ell}<\sigma<\rho and

(,n;ρ,β)(k,m;γk,τ)\left(\ell,n;\rho_{\ell},\beta\right)\leq\left(k,m;\gamma_{k},\tau\right)

for some kk\in\mathbb{N}. Since (,ρ;(n;β))E\left(\ell,\rho;\left(n;\beta\right)\right)\notin E^{\prime}, we have that k>Nk>N and hence γk>σ>ρ\gamma_{k}>\sigma>\rho_{\ell} and

π(,n;ρ,β)(k,m;γk,τ)1.\pi_{\left(\ell,n;\rho_{\ell},\beta\right)}^{\left(k,m;\gamma_{k},\tau\right)}\leq\frac{1}{\ell}\text{.}

Thus,

|x(,n;ρ,β)||π(n;β)(k,m;γk,τ)tk||tk|ε\left|\ell x\left(\ell,n;\rho_{\ell},\beta\right)\right|\leq\left|\ell\pi_{\left(n;\beta\right)}^{\left(k,m;\gamma_{k},\tau\right)}t_{k}\right|\leq\left|t_{k}\right|\leq\varepsilon

since k>NNεk>N\geq N_{\varepsilon}. This concludes the proof. ∎

Lemma 10.13.

Consider the continuous function c0(,Z)Z\mathrm{c}_{0}(\mathbb{N},Z)\rightarrow Z^{\mathbb{N}}, (xn)n(nxn)n\left(x_{n}\right)_{n\in\mathbb{N}}\mapsto\left(nx_{n}\right)_{n\in\mathbb{N}}. We have that:

  • 𝚺20\boldsymbol{\Sigma}_{2}^{0} is the complexity class of τ1(1(Z))\tau^{-1}\left(\ell_{1}\left(Z\right)\right) in c0(,Z)\mathrm{c}_{0}(\mathbb{N},Z);

  • D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) is the complexity class of τ1(bv0(Z))\tau^{-1}\left(\mathrm{bv}_{0}(Z)\right) in c0(Z)\mathrm{c}_{0}\left(Z\right);

  • 𝚷30\boldsymbol{\Pi}_{3}^{0} is the complexity class of τ1(c0(,Z))\tau^{-1}\left(\mathrm{c}_{0}\left(\mathbb{N},Z\right)\right) and of τ1(c(,Z))\tau^{-1}\left(\mathrm{c}\left(\mathbb{N},Z\right)\right) in c0(,Z)\mathrm{c}_{0}(\mathbb{N},Z).

Proof.

(1) Since 1(Z)\ell_{1}\left(Z\right) is 𝚺20\boldsymbol{\Sigma}_{2}^{0} in ZZ^{\mathbb{N}}, we have that τ1(1)\tau^{-1}\left(\ell_{1}\right) is a 𝚺20\boldsymbol{\Sigma}_{2}^{0} Polishable subgroup of c0(,Z)\mathrm{c}_{0}(\mathbb{N},Z) that is not closed. Thus, 𝚺20\boldsymbol{\Sigma}_{2}^{0} is the complexity class of τ1(1(Z))\tau^{-1}\left(\ell_{1}\left(Z\right)\right).

(2) Since bv0(Z)\mathrm{bv}_{0}\left(Z\right) is D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}) in ZZ^{\mathbb{N}}, and τ1(bv0(Z))\tau^{-1}\left(\mathrm{bv}_{0}\left(Z\right)\right) is a Polishable subgroup of c0(,Z)\mathrm{c}_{0}\left(\mathbb{N},Z\right), by Theorem 3.3 it suffices to prove that τ1(bv0(Z))\tau^{-1}\left(\mathrm{bv}_{0}\left(Z\right)\right) is not 𝚺20\boldsymbol{\Sigma}_{2}^{0} in c0(,Z)\mathrm{c}_{0}\left(\mathbb{N},Z\right). Suppose by contradiction that τ1(bv0(Z))=kωFk\tau^{-1}\left(\mathrm{bv}_{0}\left(Z\right)\right)=\bigcup_{k\in\omega}F_{k} where Fkc0(,Z)F_{k}\subseteq\mathrm{c}_{0}\left(\mathbb{N},Z\right) is closed. Observe that a compatible norm on bv0(Z)\mathrm{bv}_{0}\left(Z\right) is given by

xbv0(Z)=n|xn+1xn|+supn|xn|.\left\|x\right\|_{\mathrm{bv}_{0}\left(Z\right)}=\sum_{n\in\mathbb{N}}\left|x_{n+1}-x_{n}\right|+\mathrm{\mathrm{sup}}_{n\in\mathbb{N}}\left|x_{n}\right|\text{.}

By the Baire category theorem without loss of generality we can assume that

{𝒂τ1(bv0(Z)):τ((an))bv02}F0.\{\boldsymbol{a}\in\tau^{-1}\left(\mathrm{bv}_{0}\left(Z\right)\right):\left\|\tau\left(\left(a_{n}\right)\right)\right\|_{\mathrm{bv}_{0}}\leq 2\}\subseteq F_{0}.

Define then for NN\in\mathbb{N}, 𝒂(N)τ1(bv0(Z))\boldsymbol{a}^{\left(N\right)}\in\tau^{-1}\left(\mathrm{bv}_{0}\left(Z\right)\right) by setting

an(N)={1nif nN;0otherwise.a_{n}^{\left(N\right)}=\left\{\begin{array}[]{ll}\frac{1}{n}&\text{if }n\leq N\text{;}\\ 0&\text{otherwise.}\end{array}\right.

Then we have that τ(𝒂(N))bv0(Z)2\left\|\tau\left(\boldsymbol{a}^{(N)}\right)\right\|_{\mathrm{bv}_{0}\left(Z\right)}\leq 2 and 𝒂(N)F0\boldsymbol{a}^{\left(N\right)}\in F_{0} for every NN\in\mathbb{N}. Furthermore, the sequence (𝒂(N))N(\boldsymbol{a}^{\left(N\right)})_{N\in\mathbb{N}} converges in c0(,Z)\mathrm{c}_{0}\left(\mathbb{N},Z\right) to the sequence 𝒂\boldsymbol{a} defined by an=1na_{n}=\frac{1}{n} for every nn\in\mathbb{N}. Since F0F_{0} is closed in c0(,Z)\mathrm{c}_{0}\left(\mathbb{N},Z\right), we must have that 𝒂F0τ1(bv0(Z))\boldsymbol{a}\in F_{0}\subseteq\tau^{-1}\left(\mathrm{bv}_{0}\left(Z\right)\right). However, τ(𝒂)\tau\left(\boldsymbol{a}\right) is not vanishing, and so τ(𝒂)bv0(Z)\tau\left(\boldsymbol{a}\right)\notin\mathrm{bv}_{0}\left(Z\right).

(3) Since c(,Z)\mathrm{c}\left(\mathbb{N},Z\right) is 𝚷30\boldsymbol{\Pi}_{3}^{0} in ZZ^{\mathbb{N}}, and τ1(c(,Z))\tau^{-1}\left(\mathrm{c}\left(\mathbb{N},Z\right)\right) is a Polishable subgroup of c0\mathrm{c}_{0}, by Theorem 3.3 it suffices to prove that τ1(c(,Z))\tau^{-1}\left(\mathrm{c}\left(\mathbb{N},Z\right)\right) is not potentially 𝚺20\boldsymbol{\Sigma}_{2}^{0}. Let E0E_{0} be the relation of tail equivalence in 22^{\mathbb{N}}, and let E0E_{0}^{\mathbb{N}} be the corresponding product equivalence relation on (2)=2×\left(2^{\mathbb{N}}\right)^{\mathbb{N}}=2^{\mathbb{N}\times\mathbb{N}}. Then we have that 𝚷30\boldsymbol{\Pi}_{3}^{0} is the potential complexity class of E0E_{0}^{\mathbb{N}}, for example by Lemma 5.7 and Theorem 3.3.

Thus, it suffices to define a Borel function 2×c0(,Z)2^{\mathbb{N}\times\mathbb{N}}\rightarrow\mathrm{c}_{0}\left(\mathbb{N},Z\right) that is a Borel reduction from E0E_{0}^{\mathbb{N}} to the coset relation of τ1(c(,Z))\tau^{-1}\left(\mathrm{c}\left(\mathbb{N},Z\right)\right) inside c0(,Z)\mathrm{c}_{0}\left(\mathbb{N},Z\right). Fix a bijection ,:×\left\langle\cdot,\cdot\right\rangle:\mathbb{N}\times\mathbb{N}\rightarrow\mathbb{N} such that, if nnn\leq n^{\prime} and mmm\leq m^{\prime}, then n,mn,m\left\langle n,m\right\rangle\leq\left\langle n^{\prime},m^{\prime}\right\rangle. Define 2×Z2^{\mathbb{N}\times\mathbb{N}}\rightarrow Z^{\mathbb{N}}, xax\mapsto a by setting an,m=1n,m2nxn,ma_{\left\langle n,m\right\rangle}=\frac{1}{\left\langle n,m\right\rangle}2^{-n}x_{n,m}. Then the argument in [6, Lemma 8.5.3] shows that xE0xxE_{0}^{\mathbb{N}}x^{\prime} if and only if τ(𝒂)τ(𝒂)=τ(𝒂𝒂)c(,Z)\tau\left(\boldsymbol{a}\right)-\tau\left(\boldsymbol{a}^{\prime}\right)=\tau\left(\boldsymbol{a}-\boldsymbol{a}^{\prime}\right)\in\mathrm{c}\left(\mathbb{N},Z\right), if and only if 𝒂𝒂τ1(c(,Z))\boldsymbol{a}-\boldsymbol{a}^{\prime}\in\tau^{-1}\left(\mathrm{c}\left(\mathbb{N},Z\right)\right).

The same argument shows that 𝚷30\boldsymbol{\Pi}_{3}^{0} is the complexity class of τ1(c0(,Z))\tau^{-1}\left(\mathrm{c}_{0}\left(\mathbb{N},Z\right)\right) in c0(,Z)\mathrm{c}_{0}\left(\mathbb{N},Z\right). ∎

The same proof as Corollary 8.10, where Lemma 10.13 replaces Lemma 8.9, gives the proof of Corollary 10.14 below.

Corollary 10.14.

For every γ<α\gamma<\alpha, XγX_{\gamma} is a proper subspace of X<γX_{<\gamma}. The complexity class inside X<αX_{<\alpha} of SαS_{\alpha}, DαD_{\alpha}, XαX_{\alpha}, respectively, is 𝚺20\boldsymbol{\Sigma}_{2}^{0}, D(𝚷20)D(\boldsymbol{\Pi}_{2}^{0}), and 𝚷30\boldsymbol{\Pi}_{3}^{0}, respectively.

Finally, Theorem 10.5 is proved using Corollary 10.14 and Proposition 10.11 similarly as Theorem 8.3 is proved using Corollary 8.10 and Proposition 8.7.

References

  • Bee [91] Gerald Beer, A Polish topology for the closed subsets of a Polish space, Proceedings of the American Mathematical Society 113 (1991), no. 4, 1123–1133.
  • BK [96] Howard Becker and Alexander S. Kechris, The descriptive set theory of Polish group actions, London Mathematical Society Lecture Note Series, vol. 232, Cambridge University Press, Cambridge, 1996.
  • DG [08] Longyun Ding and Su Gao, On separable Banach subspaces, Journal of Mathematical Analysis and Applications 340 (2008), no. 1, 746–751.
  • Din [17] Longyun Ding, On equivalence relations generated by Schauder bases, The Journal of Symbolic Logic 82 (2017), no. 4, 1459–1481.
  • FS [06] Ilijas Farah and Sławomir Solecki, Borel subgroups of Polish groups, Advances in Mathematics 199 (2006), no. 2, 499–541.
  • Gao [09] Su Gao, Invariant Descriptive Set Theory, Pure and Applied Mathematics, vol. 293, CRC Press, Boca Raton, FL, 2009.
  • Hjo [06] Greg Hjorth, Subgroups of abelian Polish groups, Set theory, Trends Math., Birkhäuser, Basel, 2006, pp. 297–308.
  • HKL [98] Greg Hjorth, Alexander S. Kechris, and Alain Louveau, Borel equivalence relations induced by actions of the symmetric group, Annals of Pure and Applied Logic 92 (1998), no. 1, 63–112.
  • Kec [95] Alexander S. Kechris, Classical descriptive set theory, Graduate Texts in Mathematics, vol. 156, Springer-Verlag, New York, 1995.
  • Kur [66] Casimir Kuratowski, Topology. Vol. I, New edition, revised and augmented. Translated from the French by J. Jaworowski, Academic Press, New York-London; Państwowe Wydawnictwo Naukowe, Warsaw, 1966.
  • Lou [94] Alain Louveau, On the reducibility order between Borel equivalence relations, Logic, methodology and philosophy of science, IX (Uppsala, 1991), Stud. Logic Found. Math., vol. 134, North-Holland, Amsterdam, 1994, pp. 151–155.
  • Mal [08] Maciej Malicki, Polishable subspaces of infinite-dimensional separable Banach spaces, Real Analysis Exchange 33 (2008), no. 2, 317–322.
  • Mal [11] by same author, On Polish groups admitting a compatible complete left-invariant metric, Journal of Symbolic Logic 76 (2011), no. 2, 437–447.
  • Osb [14] M. Scott Osborne, Locally convex spaces, Graduate Texts in Mathematics, vol. 269, Springer, Cham, 2014.
  • Pet [50] Billy J. Pettis, On continuity and openness of homomorphisms in topological groups, Annals of Mathematics 52 (1950), no. 2, 293–308, Publisher: Annals of Mathematics.
  • Sol [99] Sławomir Solecki, Polish group topologies, Sets and proofs (Leeds, 1997), London Math. Soc. Lecture Note Ser., vol. 258, Cambridge Univ. Press, Cambridge, 1999, pp. 339–364.
  • Sol [09] by same author, The coset equivalence relation and topologies on subgroups, American Journal of Mathematics 131 (2009), no. 3, 571–605.
  • SR [76] Jean Saint-Raymond, Espaces à modèle séparable, Université de Grenoble. Annales de l’Institut Fourier 26 (1976), no. 3, xi, 211–256.
  • Tsa [06] Todor Tsankov, Compactifications of \mathbb{N} and Polishable subgroups of SS_{\infty}, Fundamenta Mathematicae 189 (2006), no. 3, 269–284.