Hostname: page-component-857557d7f7-qdphv Total loading time: 0 Render date: 2025-11-26T02:53:30.722Z Has data issue: false hasContentIssue false

THE WEAKNESS OF FINDING DESCENDING SEQUENCES IN ILL-FOUNDED LINEAR ORDERS

Published online by Cambridge University Press:  29 October 2025

JUN LE GOH
Affiliation:
DEPARTMENT OF MATHEMATICS NATIONAL UNIVERSITY OF SINGAPORE SINGAPORE E-mail: gohjunle@nus.edu.sg
ARNO PAULY
Affiliation:
DEPARTMENT OF COMPUTER SCIENCE SWANSEA UNIVERSITY UNITED KINGDOM E-mail: Arno.M.Pauly@gmail.com
MANLIO VALENTI*
Affiliation:
DEPARTMENT OF COMPUTER SCIENCE SWANSEA UNIVERSITY UNITED KINGDOM E-mail: Arno.M.Pauly@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

We explore the Weihrauch degree of the problems “find a bad sequence in a non-well quasi order” ($\mathsf {BS}$) and “find a descending sequence in an ill-founded linear order” ($\mathsf {DS}$). We prove that $\mathsf {DS}$ is strictly Weihrauch reducible to $\mathsf {BS}$, correcting our mistaken claim in [18]. This is done by separating their respective first-order parts. On the other hand, we show that $\mathsf {BS}$ and $\mathsf {DS}$ have the same finitary and deterministic parts, confirming that $\mathsf {BS}$ and $\mathsf {DS}$ have very similar uniform computational strength. We prove that König’s lemma $\mathsf {KL}$ and the problem $\mathsf {wList}_{{2^{\mathbb {N}}},\leq \omega }$ of enumerating a given non-empty countable closed subset of ${2^{\mathbb {N}}}$ are not Weihrauch reducible to $\mathsf {DS}$ or $\mathsf {BS}$, resolving two main open questions raised in [18]. We also answer the question, raised in [12], on the existence of a “parallel quotient” operator, and study the behavior of $\mathsf {BS}$ and $\mathsf {DS}$ under the quotient with some known problems.

Information

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of The Association for Symbolic Logic

1 Introduction

Linear and quasi-orders are ubiquitous structures that play an important role in all areas of mathematics. Computability theory has been successfully applied to highlight the difference between classical and effective properties of orders. A well-known example is that the simple fact that every infinite linear order contains an infinite ascending sequence or an infinite descending sequence is not computably true. A more extreme example is the existence of computable ill-founded linear orders with no hyperarithmetic descending sequence (see, e.g., [Reference Sacks26, Lemma III.2.1]). These orders have been extensively used and studied in reverse mathematics (under the name of pseudo-well-orders) [Reference Simpson27]. We refer the reader to [Reference Downey, Ershov, Goncharov, Nerode, Remmel and Marek11] for a more comprehensive presentation of computability-theoretical results on linear and partial orders.

The natural generalization of well-orders in the context of quasi-orders is the notion of well quasi-orders: formally, a quasi-order $(Q,\preceq )$ is called well quasi-order (abbreviated wqo) if, for every infinite sequence $(q_n)_{{n\in \mathbb {N}}}$ of elements of Q, there are $i,j$ with $i<j$ such that $q_i\preceq _Q q_j$ . This can be restated by saying that a quasi-order is a wqo if it contains no infinite bad sequences, where a (possibly finite) sequence $(q_n)_{{n}}$ is called bad if $q_i \not \preceq _Q q_j$ for every $i<j$ . Equivalently, wqos can be defined as quasi-orders that contain no infinite descending sequence and no infinite antichain. There is an extensive literature on the theory of wqos. For an overview, we refer the reader to [Reference Marcone22].

In this article, we continue our investigation (started in [Reference Goh, Pauly and Valenti18]) of the uniform computational strength of the problems:

  • given a countable ill-founded linear order, find an infinite Descending Sequence in it ( $\mathsf {DS}$ ), and

  • given a countable non-well quasi-order, find a Bad Sequence in it ( $\mathsf {BS}$ ).

A suitable framework for this is given by Weihrauch reducibility (see [Reference Brattka, Gherardi, Pauly, Brattka and Hertling3] for a self-contained introduction). Several results on the Weihrauch degree of $\mathsf {DS}$ were proved in our previous work, where we showed, e.g., that, despite the fact that $\mathsf {DS}$ is not a hyperarithmetic problem, it is rather weak from the uniform point of view, as it only computes the limit computable functions. In the language of Weihrauch reducibility, this can be stated as $\operatorname {Det}\!\left (\mathsf {DS}\right )\equiv _{\mathrm {W}}\mathsf {lim}$ . We also explored how the uniform computational strength changes when working with $\boldsymbol {\Gamma }$ -presented orders, where $\boldsymbol {\Gamma }$ is a Borel pointclass, or $\boldsymbol {\Delta }^1_1$ , $\boldsymbol {\Sigma }^1_1$ , or $\boldsymbol {\Pi }^1_1$ . Other results on the Weihrauch degree of principles related to well-orders and wqo are obtained in [Reference Carlucci, Mainardi and Zdanowski7]. The computational strength of descending sequences was also independently explored in [Reference Calvert, Franklin and Turetsky6, Section 4].

This article is organized as follows: in Section 2, we introduce the notation and provide a brief overview of the necessary background notions on Weihrauch reducibility.

In Section 3, we refute the claim, falsely stated in [Reference Goh, Pauly and Valenti18, Proposition 4.5], that $\mathsf {DS}\equiv _{\mathrm {W}}\mathsf {BS}$ . In fact, we obtain the separation $\mathsf {DS}<_{\mathrm {W}}\mathsf {BS}$ (Theorem 3.1) by proving a separation between the respective first-order parts.

The following is a list of the results in [Reference Goh, Pauly and Valenti18] that are affected:

We are not aware whether all the above-mentioned claims admit a counterexample.

All the other results in [Reference Goh, Pauly and Valenti18] do not use [Reference Goh, Pauly and Valenti18, Proposition 4.5]; they either deal only with $\mathsf {DS}$ or are standalone results about $\mathsf {BS}$ which are not affected by the above error. In particular, [Reference Goh, Pauly and Valenti18, Theorem 5.3], [Reference Goh, Pauly and Valenti18, Theorem 5.14], [Reference Goh, Pauly and Valenti18, Proposition 5.15], [Reference Goh, Pauly and Valenti18, Theorem 5.23] and [Reference Goh, Pauly and Valenti18, Corollary 5.24] are correct to the best of our knowledge. This list of errata is also available in the arxiv version of [Reference Goh, Pauly and Valenti18].

The rest of the article is devoted to better understanding the Weihrauch degrees of $\mathsf {DS}$ and $\mathsf {BS}$ . More precisely, in Section 4, we answer (negatively) two main open questions raised in [Reference Goh, Pauly and Valenti18, Questions 6.1 and 6.2], namely whether $\mathsf {KL}$ and $\mathsf {wList}_{2^{\mathbb {N}},\leq \omega }$ are Weihrauch reducible to $\mathsf {DS}$ (Corollaries 4.5 and 4.6). Both results follow from a more general characterization (Theorem 4.3), stating that

$$\begin{align*}\mathsf{lim} \equiv_{\mathrm{W}} \max_{\leq_{\mathrm{W}}}\{f {}\,:\,{} \widehat{\mathsf{ACC}_{\mathbb{N}}} \times f \leq_{\mathrm{W}} \mathsf{DS}\}\equiv_{\mathrm{W}} \max_{\leq_{\mathrm{W}}}\{f {}\,:\,{} \widehat{\mathsf{ACC}_{\mathbb{N}}} \times f \leq_{\mathrm{W}} \mathsf{BS}\}. \end{align*}$$

In other words, even though $\widehat {\mathsf {ACC}_{\mathbb {N}}}$ is fairly weak (in particular it is below $\mathsf {lim}$ , $\mathsf {KL}$ , and $\mathsf {DS}$ ), neither $\mathsf {DS}$ nor $\mathsf {BS}$ can compute $\widehat {\mathsf {ACC}_{\mathbb {N}}} \times f$ if $f \nleq _{\mathrm {W}} \mathsf {lim}$ .

In Section 5, we show that, despite $\mathsf {DS}$ and $\mathsf {BS}$ not being Weihrauch equivalent, their finitary and deterministic parts are in fact the same (Theorem 5.3 and Corollary 5.5). In other words, it is necessary to consider a non-finitary, non-deterministic problem (such as ${}^1\mathsf {BS}$ ) in order to separate $\mathsf {BS}$ and $\mathsf {DS}$ .

Section 6 contains a short detour where we answer another question that was left open in our previous paper [Reference Goh, Pauly and Valenti18, Question 3.5], namely whether the first-order part and the deterministic part commute. We answer this question by constructing an explicit counterexample for which the two operations do not commute (Theorem 6.6).

In Sections 7 and 8, we analyze the following question: Is the restriction of $\mathsf {BS}$ to trees stronger than $\mathsf {DS}$ ? This question is very natural when considering that, when proving that $\mathsf {BS}\not \leq _{\mathrm {W}} \mathsf {DS}$ , we construct a “tree-like structure” (a partial order admitting a tree decomposition, see Definition 3.3). In Section 7 we obtain some technical results on the degrees of the problems $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ and $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ . In Section 8, we show that the restriction $\mathsf {BS}|_{\operatorname {Tree}}$ of $\mathsf {BS}$ to trees with infinite width is much weaker than $\mathsf {BS}$ by proving that $\operatorname {Fin}_{}\!\left (\mathsf {BS}|_{\operatorname {Tree}}\right ) \equiv _{\mathrm {W}} \operatorname {Det}\!\left (\mathsf {BS}|_{\operatorname {Tree}}\right ) \equiv _{\mathrm {W}} \mathrm {id}$ (Corollary 8.3) and that ${}^1\mathsf {BS}|_{\operatorname {Tree}}$ is equivalent to the problem that maps a tree with infinite width to some $v\in T$ that belongs to some infinite antichain in T (Proposition 8.5).

Finally, in Section 9 we observe that one of our main technical tools (Theorem 4.3) provides an example of a “parallel quotient” [Reference Dzhafarov, Goh, Hirschfeldt, Patey and Pauly12, Remark 1], namely a specific case where

$$\begin{align*}\max_{\leq_{\mathrm{W}}} \{h {}\,:\,{} h \times g \leq_{\mathrm{W}} f\} \end{align*}$$

is defined. Whether this $\max $ exists for all choices of f and g was asked in [Reference Dzhafarov, Goh, Hirschfeldt, Patey and Pauly12, Remark 3.11]. We answer their question by showing that this operator is always defined when $g\neq \emptyset $ (Definition 9.1 and Proposition 9.9). We conclude the article by analyzing the behavior of $\mathsf {BS}$ and $\mathsf {DS}$ under the quotient with some known problems.

2 Background

We now briefly introduce the relevant notions in Weihrauch complexity. For a more thorough presentation, we refer the reader to [Reference Brattka, Gherardi, Pauly, Brattka and Hertling3]. We end this section with the observation that $\mathsf {BS}$ is equivalent to its restriction to partial orders (Proposition 2.2).

A represented space $\mathbf {X} = (X,\delta _{\mathbf {X}})$ consists of a set X and a (possibly partial) surjection $\delta _{\mathbf {X}}:\subseteq \mathbb {N}^{\mathbb {N}} \to X$ . Many spaces of interest can be represented in standard ways, such as $\mathbb {N}^{\mathbb {N}}$ , $\mathbb {N}$ , and $\mathbb {N}^{<\mathbb {N}}$ , initial segments of $\mathbb {N}$ , the set of binary relations on $\mathbb {N}$ , and the set of $\boldsymbol {\Gamma }$ -definable subsets of $\mathbb {N}$ where $\boldsymbol {\Gamma }$ is a pointclass in the arithmetic or projective hierarchy, countable Cartesian products, and countable disjoint unions of represented spaces. For the formal definitions of the representation maps of these spaces, we refer the reader to [Reference Brattka, Gherardi, Pauly, Brattka and Hertling3], or to our previous paper [Reference Goh, Pauly and Valenti18].

A problem f is a (possibly partial) multivalued function between represented spaces $\mathbf {X}$ and $\mathbf {Y}$ , denoted $f: \subseteq \mathbf {X} \rightrightarrows \mathbf {Y}$ . For each $x \in X$ , $f(x)$ denotes the set of possible outputs (i.e., f-solutions) corresponding to the input x. The domain $\operatorname {dom}(f)$ is the set of all $x \in X$ such that $f(x)$ is non-empty. Such an x is called an f-instance. If $f(x)$ is a singleton for all $x \in \operatorname {dom}(f)$ , we say f is single-valued and write $f: \subseteq \mathbf {X} \to \mathbf {Y}$ . In this case, if y is the f-solution to x, we write $f(x) = y$ instead of (the formally correct) $f(x) = \{y\}$ . A function $F: \subseteq \mathbb {N}^{\mathbb {N}} \to \mathbb {N}^{\mathbb {N}}$ is a realizer for a problem f if whenever p is a name for some $x \in \operatorname {dom}(f)$ , then $F(p)$ is a name for an f-solution to x. We say a problem is computable (resp., continuous) if it has a computable (resp., continuous) realizer.

A problem f is Weihrauch reducible to a problem g, written $f \leq _{\mathrm {W}} g$ , if there are computable maps $\Phi , \Psi : \subseteq \mathbb {N}^{\mathbb {N}} \to \mathbb {N}^{\mathbb {N}}$ such that if p is a name for some $x \in \operatorname {dom}(f)$ , then

  1. (1) $\Phi (p)$ is a name for some $y \in \operatorname {dom}(g)$ , and

  2. (2) if q is a name for some g-solution of y, then $\Psi (p,q)$ is a name for some f-solution of x.

If $\Phi $ and $\Psi $ satisfy the above, we say that $f \leq _{\mathrm {W}} g$ via $\Phi $ , $\Psi $ . We sometimes refer to $\Phi $ and $\Psi $ as the forward and backward functionals, respectively. We say that f is strongly Weihrauch reducible to g, written $f \leq _{\mathrm {sW}} g$ , if the above holds but with $\Psi (p,q)$ replaced by $\Psi (q)$ in (2). For us, strong Weihrauch reducibility is only of occasional technical interest.

Weihrauch reducibility forms a preorder on problems. We say f and g are Weihrauch equivalent, written $f \equiv _{\mathrm {W}} g$ , if $f \leq _{\mathrm {W}} g$ and $g \leq _{\mathrm {W}} f$ . The $\equiv _{\mathrm {W}}$ -equivalence classes (Weihrauch degrees) are partially ordered by $\leq _{\mathrm {W}}$ . Among the numerous algebraic operations in the Weihrauch degrees, we consider:

  • for problems $f_i: \subseteq \mathbf {X}_i \rightrightarrows \mathbf {Y}_i$ , the parallel product

    $$\begin{align*}f_0 \times f_1: \subseteq \mathbf{X}_0 \times \mathbf{X}_1 \rightrightarrows \mathbf{Y}_0 \times \mathbf{Y}_1 \text{ defined by } (x_0,x_1) \mapsto f_0(x_0) \times f_1(x_1), \end{align*}$$
    i.e., given an $f_0$ -instance and an $f_1$ -instance, solve both;
  • for a problem $f: \subseteq \mathbf {X} \rightrightarrows \mathbf {Y}$ , the (infinite) parallelization

    $$\begin{align*}\widehat{f}: \subseteq \mathbf{X}^{\mathbb{N}} \rightrightarrows \mathbf{Y}^{\mathbb{N}} \text{ defined by } (x_i)_i \mapsto \prod_i f(x_i), \end{align*}$$
    i.e., given a countable sequence of f-instances, solve all of them.

These operations are defined on problems, but they all lift to the Weihrauch degrees. Parallelization even forms a closure operator, i.e., $f \leq _{\mathrm {W}} \widehat {f}$ , $f \leq _{\mathrm {W}} g$ implies $\widehat {f} \leq _{\mathrm {W}} \widehat {g}$ , and $\widehat {\widehat {f}} \equiv _{\mathrm {W}} \widehat {f}$ . We also briefly mention two further operations on Weihrauch degrees, respectively, called compositional product and implication, characterized by $f * g = \max _{\leq _{\mathrm {W}}} \{F \circ G {}\,:\,{} F \leq _{\mathrm {W}} f \wedge G \leq _{\mathrm {W}} g\}$ and $f \rightarrow g = \min _{\leq _{\mathrm {W}}} \{h {}\,:\,{} g \leq _{\mathrm {W}} f * h\}$ . Both are total operations. The compositional product can be equivalently described as follows: let $(\Phi _w)_{{w\in \mathbb {N}^{\mathbb {N}}}}$ be an effective enumeration of all partial continuous functionals $\mathbb {N}^{\mathbb {N}}\rightarrow \mathbb {N}^{\mathbb {N}}$ with $G_\delta $ domain. For every $f:\subseteq \mathbf {X} \rightrightarrows \mathbf {Y}$ , $g:\subseteq \mathbf {U} \rightrightarrows \mathbf {V}$ , the domain of $f* g$ is the set

$$\begin{align*}\{ (w,u) \in \mathbb{N}^{\mathbb{N}} \times \operatorname{dom}(g) {}\,:\,{} (\forall v \in g(u))(\forall p_v\in \delta_{\mathbf{V}}^{-1}(v))(\delta_{\mathbf{X}}\Phi_w(p_v)\in \operatorname{dom}(f))\}.\end{align*}$$

Given as input a pair $(w,u)$ , $f* g$ produces a pair $(y,v)$ with $v\in g(u)$ and ${y\in f\delta _{\mathbf {X}}\Phi _w(p_v) }$ for some $p_v\in \delta _{\mathbf {V}}^{-1}(v)$ . If $f,g$ are problems on the Baire space, $f* g$ is simply the problem that maps $(w,u)$ to a pair $(y,v)$ with $v\in g(u)$ and $y\in f(v)$ . For details, we refer to [Reference Brattka and Pauly4, Reference Westrick29].

The Weihrauch degrees also support a number of interior operators, which have been used to separate degrees of interest (see, e.g., [Reference Soldà and Valenti28, Section 3.1]). For any problem f and any represented space $\mathbf {X}$ , the problem

$$\begin{align*}\mathrm{Det}_{ \mathbf{X} }(f) := \max\nolimits _{\leq_{\mathrm{W}}} \{g \leq_{\mathrm{W}} f {}\,:\,{} g \text{ has codomain }\mathbf{X}\text{ and is single-valued}\} \end{align*}$$

exists [Reference Goh, Pauly and Valenti18, Theorem 3.2]. We call $\mathrm {Det}_{ \mathbf {\mathbb {N}^{\mathbb {N}}} }(f)$ the deterministic part of f and denote it by $\operatorname {Det}\!\left (f\right )$ for short. We say that a problem is deterministic if $f\equiv _{\mathrm {W}} \operatorname {Det}\!\left (f\right )$ , i.e., if f is Weihrauch equivalent to a single-valued function on Baire space. This terminology is motivated by the observation that the treatment of multi-valued functions in computable analysis inherently corresponds to a non-deterministic computation point of view.

Observe that [Reference Goh, Pauly and Valenti18, Proposition 3.6] can be generalized slightly.

Proposition 2.1. $\operatorname {Det}\!\left (f\right ) \leq _{\mathrm {W}} \widehat {\mathrm {Det}_{ \mathbf {2} }(f)}$ . ( $\mathbf {2}$ is the two-point space with the discrete topology.)

Proof. Suppose g is single-valued, has codomain $\mathbb {N}^{\mathbb {N}}$ , and $g \leq _{\mathrm {W}} f$ . Define a single-valued problem h with codomain $\textbf {2}$ as follows: Given $n,m \in \mathbb {N}$ and a g-instance x, produce $1$ if $g(x)(n) \geq m$ , otherwise produce $0$ . It is easy to see that $g \leq _{\mathrm {W}} \widehat {h}$ and $h \leq _{\mathrm {W}} f$ . The latter implies $h \leq _{\mathrm {W}} \mathrm {Det}_{ \mathbf {2} }(f)$ and so $g \leq _{\mathrm {W}} \widehat {h} \leq _{\mathrm {W}} \widehat {\mathrm {Det}_{ \mathbf {2} }(f)}$ .

For any problem f and $\mathbf {X} = \mathbb {N}$ or $\mathbf {k}$ , it is also known that

$$\begin{align*}\max\nolimits _{\leq_{\mathrm{W}}} \{g \leq_{\mathrm{W}} f {}\,:\,{} g \text{ has codomain }\mathbf{X}\} \end{align*}$$

exists. For $\mathbf {X} = \mathbb {N}$ we call it the first-order part of f [Reference Dzhafarov, Solomon and Yokoyama13, Theorem 2.2], denoted by ${}^1f$ , while for $\mathbf {X} = \mathbf {k}$ we call it the $\mathbf {k}$ -finitary part of f [Reference Cipriani and Pauly10, Proposition 2.9], denoted by $\operatorname {Fin}_{k}\!\left (f\right )$ . We say that a problem is first-order if $f\equiv _{\mathrm {W}} {}^1f$ . It is immediate from the definitions that $\mathrm {Det}_{ \mathbf {\mathbb {N}} }(f) \leq _{\mathrm {W}} {}^1f$ and $\mathrm {Det}_{ \mathbf {k} }(f) \leq _{\mathrm {W}} \operatorname {Fin}_{k}\!\left (f\right ) \leq _{\mathrm {W}} {}^1f$ .

Before providing examples of problems of interest, we shall specify our notation for sequences and trees. A string or sequence is a function $\sigma $ from an initial segment of $\mathbb {N}$ to $\mathbb {N}$ , with length denoted by $|\sigma |$ . The set of finite (resp., infinite) strings of natural numbers is denoted by $\mathbb {N}^{<\mathbb {N}}$ (resp., $\mathbb {N}^{\mathbb {N}}$ ). Likewise, we use $n^{<\mathbb {N}}$ and $n^{\mathbb {N}}$ , respectively, for the sets of finite or infinite strings with range in . In particular, the sets of finite and infinite binary strings are denoted $2^{<\mathbb {N}}$ and ${2^{\mathbb {N}}}$ , respectively.

The constant sequence with value n and length i is denoted by $n^i$ . For $i = 0$ , this is the empty string $\varepsilon $ . We use $n^\omega $ for the constant infinite string with value n. The concatenation of strings $\sigma $ and $\tau $ is usually denoted by . For the sake of readability, we omit the concatenation symbol whenever we are concatenating constant strings (e.g., we use $0^i1^j$ in place of ). The prefix relation on strings is denoted by $\sqsubseteq $ .

For each $p \in \mathbb {N}^{\mathbb {N}}$ and $i \in \mathbb {N}$ , $p[i]$ denotes the string obtained by restricting p to i. A subtree of $\mathbb {N}^{<\mathbb {N}}$ (resp., $2^{<\mathbb {N}}$ ) is a subset $T \subseteq \mathbb {N}^{<\mathbb {N}}$ (resp., $2^{<\mathbb {N}}$ ) which is closed downwards under prefix. We picture trees growing upwards, i.e., the root $\varepsilon $ lies at the bottom.

We now list some well-studied computational problems that will be useful.

  • $\mathsf {LPO}$ : Given $p \in {2^{\mathbb {N}}}$ , produce $1$ if there is $k\in \mathbb {N}$ such that $p(k) = 1$ , otherwise produce $0$ .

  • $\mathsf {C}_{k}$ : Given $p\in (k+1)^{\mathbb {N}}$ such that $(\exists n<k)(n+1\notin \operatorname {ran}(p))$ , find $n<k$ such that $n+1\notin \operatorname {ran}(p)$ .

  • $\mathsf {C}_{\mathbb {N}}$ : Given $p\in \mathbb {N}^{\mathbb {N}}$ such that $(\exists n)(n+1\notin \operatorname {ran}(p))$ , find n such that $n+1\notin \operatorname {ran}(p)$ .

  • $\mathsf {lim}$ : Given a convergent sequence $(p_n)_{{n\in \mathbb {N}}}$ in $\mathbb {N}^{\mathbb {N}}$ , find its limit.

  • $\mathsf {KL}$ : Given an infinite finitely branching subtree of $\mathbb {N}^{<\mathbb {N}}$ , find an infinite path through it.

The problems $\mathsf {C}_{k}$ and $\mathsf {C}_{\mathbb {N}}$ are examples of choice problems, as they can be rephrased as “given a non-empty co-c.e. closed subset of $\mathbf {k}$ (resp., $\mathbb {N}$ ), find a point in it”. Choice problems are pivotal in the study of Weihrauch reducibility, as they provide a useful benchmark to describe the Weihrauch degrees of other problems. For this reason, many variations of choice problems have been introduced in the literature. For example, we denote with $\mathsf {UC}_{\mathbb {N}}$ the restriction of $\mathsf {C}_{\mathbb {N}}$ to co-c.e. closed singletons (or, equivalently, to sequences p such that there is exactly one n such that $n+1\notin \operatorname {ran}(p)$ ). It is known that $\mathsf {C}_{\mathbb {N}}\equiv _{\mathrm {W}} \mathsf {UC}_{\mathbb {N}}$ [Reference Brattka, Gherardi and Marcone2, Theorem 3.8], hence $\mathsf {C}_{\mathbb {N}}$ is a deterministic, first-order problem. Another variant that plays an important role in this article is all or co-unique choice: if $\mathbf {X}$ is $\mathbf {k}$ or $\mathbb {N}$ , $\mathsf {ACC}_{\mathbf {X}}$ is the problem “Given an enumeration of a set $A \subseteq \mathbf {X}$ of size at most $1$ , find a number not in A”. It is known that $\widehat {\mathsf {ACC}_{\mathbb {N}}} <_{\mathrm {W}} \mathsf {lim} <_{\mathrm {W}} \mathsf {KL}$ (see, e.g., [Reference Brattka, Gherardi, Pauly, Brattka and Hertling3]).

Another important family of problems comes from Ramsey’s theorem for n-tuples and k-colors. In this article, we only deal with colorings of the natural numbers. We define $\mathsf {RT}^{{1}}_{{k}}$ , a.k.a. the pigeonhole principle, as the total multi-valued function that maps a coloring to the set of all infinite c-homogeneous sets, i.e., the set of all infinite $H\subseteq \mathbb {N}$ such that c is constant on H. Many well-known facts about the Weihrauch degrees of Ramsey principles can be found in [Reference Brattka and Rakotoniaina5]. We only mention that (in terms of Weihrauch reducibility) we can equivalently think of $\mathsf {RT}^{{1}}_{{k}}$ as the problem that, given c, produces the color of a c-homogeneous solution. It is known that $j<k$ implies $\mathsf {RT}^{{1}}_{{j}}<_{\mathrm {W}} \mathsf {RT}^{{1}}_{{k}}$ and that $\widehat {\mathsf {RT}^{{1}}_{{2}}}\equiv _{\mathrm {W}} \mathsf {KL}$ . In particular, $\mathsf {KL}$ is parallelizable (as the parallelization is a closure operator).

To study the problems $\mathsf {DS}$ and $\mathsf {BS}$ from the point of view of Weihrauch reducibility, we need to introduce the represented spaces of linear orders and quasi-orders. We only work with countable linear orders/quasi-orders with domain contained in $\mathbb {N}$ . We represent a linear order $(L,\le _L)$ with the characteristic function of the set $\{ \langle n,m \rangle {}\,:\,{} n\le _L m\}$ . Likewise, we represent a quasi-order $(Q,\preceq _Q)$ with the characteristic function of the set $\{ \langle n,m \rangle {}\,:\,{} n\preceq _Q m\}$ .

We conclude this section by observing a fact about $\mathsf {BS}$ which was implicit in [Reference Goh, Pauly and Valenti18].

Proposition 2.2. $\mathsf {BS}$ is Weihrauch equivalent to its restriction to partial orders.

Proof. Given a non-well quasi-order $(Q,\preceq _Q)$ where $Q \subseteq \mathbb {N}$ , compute the set $S := \{a \in Q {}\,:\,{} (\forall b <_{\mathbb {N}} a)(a \not \preceq _Q b \text { or }b \not \preceq _Q a)\}$ . The restriction $(S,\preceq _Q)$ is a non-well partial order because it is isomorphic to the partial order of $\preceq _Q$ -equivalence classes.

Henceforth we will use Proposition 2.2 without mention.

3 Separating $\mathsf {BS}$ and $\mathsf {DS}$

We shall separate $\mathsf {BS}$ and $\mathsf {DS}$ by separating their first-order parts.

Theorem 3.1. ${}^1\mathsf {BS} \nleq _{\mathrm {W}} {}^1\mathsf {DS}$ and so $\mathsf {DS} <_{\mathrm {W}} \mathsf {BS}$ .

Recall from [Reference Goh, Pauly and Valenti18, Theorem 4.10] that ${}^1\mathsf {DS} \equiv _{\mathrm {W}} \boldsymbol {\Pi }^1_1\mathsf {-Bound}$ , which is the problem of producing an upper bound for a finite subset of $\mathbb {N}$ (given via a $\boldsymbol {\Pi }^1_1$ -code). Observe that $\boldsymbol {\Pi }^1_1\mathsf {-Bound}$ is upwards closed, i.e., if $n \in g(x)$ then $m \in g(x)$ for all $m> n$ .

Lemma 3.2. Let f be a problem with codomain $\mathbb {N}$ . The following are equivalent:

  1. (1) there exists an upwards closed problem g with codomain $\mathbb {N}$ such that $f\leq _{\mathrm {W}} g$ ;

  2. (2) there is a computable procedure which takes as input any $x \in \operatorname {dom}(f)$ and produces a sequence $p_x\in \mathbb {N}^{\mathbb {N}} $ of guesses for f-solutions to x which is correct cofinitely often.

Proof. For $(1) \Rightarrow (2)$ , let g be upwards closed with codomain $\mathbb {N}$ and assume $f \leq _{\mathrm {W}} g$ via $\Phi $ and $\Psi $ . Given $x \in \operatorname {dom}(f)$ , run the computations $(\Psi (x,m))_{m \in \mathbb {N}}$ in parallel. Once some $\Psi (x,m)$ halts, we output its result and cancel $\Psi (x,n)$ for all $n < m$ . This produces a sequence of numbers. The fact that g is upwards closed guarantees that cofinitely many elements of this sequence are elements of $f(x)$ .

For the converse direction, for every $x\in \operatorname {dom}(f)$ , let $p_x\in \mathbb {N}^{\mathbb {N}}$ be as in the hypothesis. Define $M_x:=\max \{m {}\,:\,{} p_x(m)\notin f(x)\}$ and let $g(x):=\{n {}\,:\,{} n>M_x\}$ . Clearly g is upwards closed. The fact that $f\leq _{\mathrm {W}} g$ follows from the fact that $x\mapsto p_x$ is computable.

Given a non-well quasi-order $(Q,\preceq _Q)$ , we say that a finite sequence $\sigma $ is extendible to an infinite $\preceq _Q$ -bad sequence (or, more compactly, $\sigma $ is $\preceq _Q$ -extendible) if there is a $\preceq _Q$ -bad sequence $(q_n)_{{n\in \mathbb {N}}}$ such that $(\forall i<|{\sigma }|)(\sigma (i)=q_i)$ . We omit the subscript whenever there is no ambiguity.

Observe that ${}^1\mathsf {BS}$ can compute the problem “given a non-well partial order ${(P,\le _P)}$ , produce an element of P that is extendible to an infinite bad sequence”. In light of Lemma 3.2, to prove Theorem 3.1 it suffices to show that one cannot computably “guess” solutions for $\mathsf {BS}$ . In other words, given a computable procedure which tries to guess extendible elements in a non-wqo, we want to construct a non-wqo P on which the procedure outputs a non-extendible element infinitely often. This would imply that ${}^1\mathsf {BS} \not \leq _{\mathrm {W}} \boldsymbol {\Pi }^1_1\mathsf {-Bound}$ . The non-wqos P we construct will be “tree-like” in the following sense.

Definition 3.3. A tree decomposition of a partial order $(P,\le _P)$ consists of a tree $T \subseteq 2^{<\mathbb {N}}$ and a function such that:

  1. (1) If $w_1,w_2 \in T$ and $w_1$ is a proper prefix of $w_2$ (written $w_1 \sqsubset w_2$ ), then $\iota (w_1) <_P \iota (w_2)$ .

  2. (2) P is partitioned into finite P-intervals, where each interval has the form

    for some vertex (with final entry b), or $(\varepsilon ] = \{\iota (\varepsilon )\}$ (where $\varepsilon $ denotes the root of $2^{<\mathbb {N}}$ ). For $v \in P$ let $\lceil v \rceil \in T$ be uniquely defined by $v \in (\lceil v \rceil ]$ .
  3. (3) If $w_1, w_2 \in T$ are incompatible, so are $\iota (w_1)$ and $\iota (w_2)$ (i.e., they have no common upper bound in P).

The following lemma is straightforward.

Lemma 3.4. If is a tree decomposition, then P has no infinite descending sequences. Moreover, T is wqo (i.e., it has finite width) if and only if P is wqo. In other words, T has an infinite antichain iff so does P.

Proof. The fact that every partial order that admits a tree decomposition does not have an infinite descending sequence follows from the fact that if ( $v_n)_{n \in \mathbb {N}}$ is an infinite descending sequence in P, then since every interval $(\lceil v_n \rceil ]$ is finite, up to removing duplicates, the sequence $(\lceil v_n\rceil )_{n\in \mathbb {N}}$ would be an infinite descending sequence in T.

If $(w_n)_{n \in \mathbb {N}}$ is an infinite antichain in T then, by definition of tree decomposition, $(\iota (w_n))_{n \in \mathbb {N}}$ is an infinite antichain in P. Conversely, if $(v_n)_{n \in \mathbb {N}}$ is an infinite antichain in P, then for every n, for all but finitely many m, $\lceil v_n \rceil $ is $\sqsubseteq $ -incomparable with $\lceil v_m \rceil $ . In particular, we can obtain an infinite antichain in T by choosing a subsequence $(v_{n_i})_{i \in \mathbb {N}}$ such that, for every $i\neq j$ , $\lceil v_{n_i} \rceil $ and $\lceil v_{n_j} \rceil $ are $\sqsubseteq $ -incomparable.

Lemma 3.5. There is no computable procedure that, given as input a non-well partial order which admits a tree decomposition, outputs an infinite sequence of elements of that partial order such that cofinitely many elements in the output are extendible to a bad sequence.

We point out a subtle yet important aspect regarding Lemma 3.5: The procedure only has access to the partial order, not to a tree decomposition of it.

Proof. Fix a computable “guessing” procedure g that receives as input a partial order (admitting a tree decomposition) and outputs an infinite sequence of elements in that partial order. We shall build a partial order P together with a tree decomposition in stages such that, infinitely often, g outputs an element of P that does not extend to an infinite bad sequence.

Start with $T_0 = \{\varepsilon \}$ and $P_0$ having a single element $v_\varepsilon $ , with $\iota _0(\varepsilon ) = v_\varepsilon $ . In stage s, we have built a finite tree decomposition and wish to extend it to some . The tree $T_{s+1}$ will always be obtained by giving each leaf in $T_s$ a single child, and then adding two children to exactly one of the new leaves. To decide which leaf gets two children, say a finite extension Q of $P_s$ is suitable for if for every $v \in Q \setminus P_s$ , there is exactly one leaf $w \in T_s$ such that $\iota _s(w) <_Q v$ . Pick the left-most leaf $\sigma $ of $T_s$ with the following property:

There is some suitable extension Q of $P_s$ such that, when given Q, the guessing procedure g would guess an element of Q which is comparable with $\iota _s(\sigma )$ .

To see that such $\sigma $ must exist, consider extending $P_s$ by adding an “infinite comb” (i.e., a copy of $\{0^n1^i {}\,:\,{} n \in \mathbb {N}, i \in \{0,1\}\}$ ) above the $\iota _s$ -image of a single leaf in $T_s$ . The resulting partial order $\overline {Q}$ is non-wqo, admits a tree decomposition (obtained by extending $T_s$ and $\iota _s$ in the obvious way), and its finite approximations (extending $P_s$ ) are suitable for $\iota _s$ . Hence, by hypothesis, g eventually guesses some element, which must be comparable with $\iota _s(\sigma )$ for some leaf $\sigma \in T_s$ (because all elements of $\overline {Q}$ are).

Having identified $\sigma $ , we fix any corresponding suitable extension Q of $P_s$ . In order to extend $\iota _s$ , we further extend Q to $Q'$ by adding a new maximal element $v_w$ to Q for each leaf $w \in T_s$ as follows: $v_w$ lies above all $v \in Q\setminus P_s$ such that $\iota _s(w) <_P v$ , and is incomparable with all other elements (including the other new maximal elements $v_{w'}$ ). To extend $T_s$ , we add a new leaf to $T_s$ for each leaf $\tau $ , obtaining a tree $T'$ . We extend $\iota _s$ to yield a tree decomposition in the obvious way.

Finally, we add two children to in $T'$ , i.e., define . We also add two children $v_1$ , $v_2$ to in $Q'$ to obtain $P_{s+1}$ , and extend $\iota '$ to $\iota _{s+1}$ by setting . This concludes stage s.

Figure 1 Schematic representation of the construction used in the proof of Lemma 3.5. The dashed box contains the partial construction up to stage s. The gray boxes contain intervals in the partial order P. For simplicity, each interval up to stage s only contains one point (i.e., the partial order $P_s$ is isomorphic to a tree), but this need not be the case in general. The black nodes are those in the range of the (partial) tree decomposition $\iota _s$ . The square gray node is the node guessed by g at stage $s+1$ (which identifies $\iota (\sigma )$ and ).

It is clear from the construction that $\iota : T \to P$ is a tree decomposition. Let us discuss the shape of the tree T. In stage s, we introduced a bifurcation above a leaf $\sigma _s$ of $T_s$ . These are the only bifurcations in T. Observe that, whenever $s' < s$ , $\sigma _s$ is either above or to the right of $\sigma _{s'}$ , because every suitable extension of $P_s$ is also a suitable extension of $P_{s'}$ and, at stage $s'$ , the chosen leaf was the left-most. This implies that the sequence $(\sigma _s)_s$ converges to a path $p\in [T]$ , i.e., for every n there is a stage s such that for every $s'>s$ , $p[n]\sqsubseteq \sigma _{s'}$ . This also implies that p is the unique non-isolated path (if $\tau \not \sqsubseteq p$ then there are only finitely many bifurcations above $\tau $ ). Observe also that a vertex w in T is extendible to an infinite antichain in T if and only if it does not belong to p.

We may now apply Lemma 3.4 to analyze P. First, since T is not wqo, neither is P. Second, we claim that if $v <_P \iota (\sigma )$ for some vertex $\sigma $ on p, then v is not extendible to an infinite bad sequence. To prove this, suppose v is extendible. Then so is $\iota (\sigma )$ . The proof of Lemma 3.4 implies that $\lceil \iota (\sigma ) \rceil = \sigma $ is extendible to an infinite antichain in T. So $\sigma $ cannot lie on p, proving our claim.

To complete the proof, observe that our construction of $\iota $ ensures that for each s, $g(P)$ eventually outputs a guess which is below . Whenever lies along p (which holds for infinitely many s), this guess is wrong by the above claim (Figure 1).

We may now complete the proof of Theorem 3.1.

Proof of Theorem 3.1.

Suppose towards a contradiction that ${}^1\mathsf {BS} \leq _{\mathrm {W}} \boldsymbol {\Pi }^1_1\mathsf {-Bound}$ . Since the problem of finding an element in a non-wqo which extends to an infinite bad sequence is first-order, it is Weihrauch reducible to $\boldsymbol {\Pi }^1_1\mathsf {-Bound}$ as well. Now, $\boldsymbol {\Pi }^1_1\mathsf {-Bound}$ is upwards closed, so there is a computable guessing procedure for this problem (Lemma 3.2). However such a procedure cannot exist, even for partial orders which admit a tree decomposition (Lemma 3.5).

4 Separating $\mathsf {KL}$ and $\mathsf {BS}$

In this section, we answer two of the main questions that were left open in [Reference Goh, Pauly and Valenti18, Questions 6.1 and 6.2], namely whether the problems $\mathsf {KL}$ and $\mathsf {wList}_{2^{\mathbb {N}},\leq \omega }$ are Weihrauch reducible to $\mathsf {DS}$ . We already introduced $\mathsf {KL}$ in Section 2, while $\mathsf {wList}_{2^{\mathbb {N}},\leq \omega }$ is the problem of enumerating all elements (possibly with repetition) of a given non-empty countable closed subset of $2^{\mathbb {N}}$ . This problem was introduced in [Reference Kihara, Marcone and Pauly17, Section 6] and was also studied in [Reference Cipriani, Marcone and Valenti9].

In fact, we prove something stronger, namely that neither $\mathsf {KL}$ nor $\mathsf {wList}_{2^{\mathbb {N}},\leq \omega }$ are Weihrauch reducible to $\mathsf {BS}$ . The core of the proofs rests on the following two technical results.

Definition 4.1. Given a fixed partial order $(P,\leq _P)$ , we define the following quasi-order on the (finite or infinite) $\leq _P$ -bad sequences:

$$\begin{align*}\alpha \trianglelefteq^P \beta :\hspace{-1mm}\iff \alpha = \beta \text{ or } (\exists i<|{\alpha}|)(\forall j<|{\beta}|)(\alpha(i) \leq_P \beta(j)). \end{align*}$$

We just write $\trianglelefteq $ when the partial order is clear from the context.

Lemma 4.2. Let $(P,\le _P)$ be a non-well partial order and let $\alpha ,\beta $ be finite $\le _P$ -bad sequences. If $\alpha \trianglelefteq \beta $ and $\alpha $ is extendible to an infinite $\le _P$ -bad sequence, then so is $\beta $ . If $\alpha $ is not extendible then there is an infinite $\le _P$ -bad sequence $B\in \mathbb {N}^{\mathbb {N}}$ such that $\alpha \trianglelefteq B$ . (Hence $\alpha \trianglelefteq \beta $ for every initial segment $\beta $ of B.)

Proof. To prove the first part of the theorem, fix $\alpha \trianglelefteq \beta $ and let $A\in \mathbb {N}^{\mathbb {N}}$ be an infinite $\le _P$ -bad sequence extending $\alpha $ . Let also $i<|{\alpha }|$ be a witness for $\alpha \trianglelefteq \beta $ . For every $j>i$ and every $k<|{\beta }|$ , $\beta (k)\not \le _P A(j)$ (otherwise $A(i) = \alpha (i) \le _P\beta (k)\le _P A(j)$ would contradict the fact that A is a $\le _P$ -bad sequence), which implies that $\beta $ is extendible.

Assume now that $\alpha $ is non-extendible and let $F \in \mathbb {N}^{\mathbb {N}}$ be a $\le _P$ -bad sequence. We show that there is $i<|{\alpha }|$ and infinitely many k such that $\alpha (i) <_P F(k)$ . This is enough to conclude the proof, as we could take B as any subsequence of F with $\alpha (i) <_P B(k)$ for every k (i.e., $\alpha \trianglelefteq B$ ).

Assume that, for every $i<|{\alpha }|$ there is $k_i$ such that for every $k\ge k_i$ , $\alpha (i)\not \le _P F(k)$ (since P is a partial order, there can be at most one k such that $\alpha (i)=F(k)$ ). Since $\alpha $ is finite, we can take $k:=\max _{i<|{\alpha }|} k_i$ and consider the sequence which extends $\alpha $ by the tail $F(k+1), F(k+2), \dots $ of F. We have now reached a contradiction as this is an infinite $\le _P$ -bad sequence extending $\alpha $ .

Theorem 4.3. Let f be a problem. The following are equivalent:

  1. (1) $f \leq _{\mathrm {W}} \mathsf {lim;}$

  2. (2) $\widehat {\mathsf {ACC}_{\mathbb {N}}} \times f \leq _{\mathrm {W}} \mathsf {lim;}$

  3. (3) $\widehat {\mathsf {ACC}_{\mathbb {N}}} \times f \leq _{\mathrm {W}} \mathsf {DS;}$

  4. (4) $\widehat {\mathsf {ACC}_{\mathbb {N}}} \times f \leq _{\mathrm {W}} \mathsf {BS.}$

This theorem provides another example of a “parallel quotient” (cf. [Reference Dzhafarov, Goh, Hirschfeldt, Patey and Pauly12, Remark 3.11]). See Section 9 for details.

Proof. The implication from (1) to (2) holds because $\widehat {\mathsf {ACC}_{\mathbb {N}}} \times \mathsf {lim} \equiv _{\mathrm {W}} \mathsf {lim}$ . The implications from (2) to (3) and from (3) to (4) hold because $\mathsf {lim} \leq _{\mathrm {W}} \mathsf {DS} \leq _{\mathrm {W}} \mathsf {BS}$ [Reference Goh, Pauly and Valenti18, Theorem 4.16]. For the implication from (4) to (1), we consider a name x for an input to f together with witnesses $\Phi ,\Psi $ for the reduction to $\mathsf {BS}$ . We show that, from them, we can uniformly compute an input q to $\widehat {\mathsf {ACC}_{\mathbb {N}}}$ together with an enumeration of the set W of all finite sequences $\sigma $ in the (non-well) partial order P built by $\Phi $ on $(q,x)$ such that $\sigma $ does not extend to an infinite bad sequence. We can then use $\mathsf {lim}$ to obtain the characteristic function of W. Having access to this lets us find an infinite bad sequence in P greedily by avoiding sequences in W. From such a bad sequence (and $(q,x)$ ), $\Psi $ then computes a solution to f for x.

It remains to construct $q = (q_\sigma )_{{\sigma \in \mathbb {N}}}$ and W to achieve the above. At the beginning, W is empty, and we extend each $q_\sigma $ in a way that removes no solution from its $\mathsf {ACC}_{\mathbb {N}}$ -instance. As we do so, for each $\sigma \notin W$ (in parallel), we monitor whether the following condition has occurred:

$\sigma $ is a bad sequence in P (as computed by the finite prefix of $(q,x)$ built/observed thus far), and there is some (finite) bad sequence $\tau $ in P such that:

  • $\sigma \trianglelefteq ^P \tau $ ;

  • the functional $\Psi $ , upon reading the current prefix of $(q,x)$ and $\tau $ , produces some output m for the $\mathsf {ACC}_{\mathbb {N}}$ -instance indexed by $\sigma $ .

Once the above occurs for $\sigma $ (if ever), we remove m as a valid solution to $q_\sigma $ by enumerating it. This ensures that $\tau $ , and hence $\sigma $ by Lemma 4.2, cannot extend to an infinite bad sequence in P. We shall then enumerate $\sigma $ into W. This completes our action for $\sigma $ , after which we return to monitoring the above condition for sequences not in W.

It is clear that each $q_\sigma $ is an $\mathsf {ACC}_{\mathbb {N}}$ -instance (with solution set $\mathbb {N}$ if the condition is never triggered, otherwise with solution set $\mathbb {N}\setminus \{m\}$ ). Hence $P:=\Phi (q,x)$ is a non-well partial order. As argued above, no $\sigma \in W$ extends to an infinite bad sequence in P. Conversely, suppose $\sigma $ is a bad sequence in P which does not extend to an infinite bad sequence. Since P is non-wqo, by Lemma 4.2 there is an infinite bad sequence r such that $\sigma \trianglelefteq ^P r$ . Then $\Psi $ has to produce all $\mathsf {ACC}_{\mathbb {N}}$ answers upon receiving $(q,x)$ and r, including an answer to $q_\sigma $ . By continuity, this answer is determined by finite prefixes only. In particular, after having constructed a sufficiently long prefix of q, some finite prefix $\tau $ of r will trigger the condition for $\sigma $ (unless something else triggered it previously), which ensures that $\sigma $ gets placed into W. This shows that W contains exactly the non-extendible finite bad sequences in P, thereby concluding the proof.

In other words, this theorem can be restated saying that $\mathsf {lim}$ is the strongest problem such that its parallel product with $\widehat {\mathsf {ACC}_{\mathbb {N}}}$ is reducible to either $\mathsf {lim}$ , $\mathsf {DS}$ , or $\mathsf {BS}$ . In symbols,

$$ \begin{align*} \mathsf{lim} & \equiv_{\mathrm{W}} \max\nolimits _{\leq_{\mathrm{W}}} \{ h {}\,:\,{} \widehat{\mathsf{ACC}_{\mathbb{N}}} \times h \leq_{\mathrm{W}} \mathsf{lim}\}\\ & \equiv_{\mathrm{W}}\max\nolimits _{\leq_{\mathrm{W}}} \{ h {}\,:\,{} \widehat{\mathsf{ACC}_{\mathbb{N}}} \times h \leq_{\mathrm{W}} \mathsf{DS}\} \\ & \equiv_{\mathrm{W}} \max\nolimits _{\leq_{\mathrm{W}}} \{ h {}\,:\,{} \widehat{\mathsf{ACC}_{\mathbb{N}}} \times h \leq_{\mathrm{W}} \mathsf{BS}\}. \end{align*} $$

This characterizes the “parallel quotient” of $\mathsf {DS}$ and $\mathsf {BS}$ over $\widehat {\mathsf {ACC}_{\mathbb {N}}}$ . The parallel quotient will be discussed more extensively in Section 9.

Corollary 4.4. If f is a parallelizable problem (i.e., $f \equiv _{\mathrm {W}} \widehat {f}$ ) with $\mathsf {ACC}_{\mathbb {N}} \leq _{\mathrm {W}} f \leq _{\mathrm {W}} \mathsf {BS}$ , then $f \leq _{\mathrm {W}} \mathsf {lim}$ .

Proof. Since $\mathsf {ACC}_{\mathbb {N}} \leq _{\mathrm {W}} f \leq _{\mathrm {W}} \mathsf {BS}$ and f is parallelizable, we have $\widehat {\mathsf {ACC}_{\mathbb {N}}} \times f \leq _{\mathrm {W}} f \leq _{\mathrm {W}} \mathsf {BS}$ . By the previous theorem, $f \leq _{\mathrm {W}} \mathsf {lim}$ .

The negative answer to [Reference Goh, Pauly and Valenti18, Question 6.1] is an immediate consequence of the following result.

Corollary 4.5. $\mathsf {KL} \nleq _{\mathrm {W}} \mathsf {BS}$ .

Proof. Recall that $\mathsf {KL}$ is parallelizable, $\mathsf {ACC}_{\mathbb {N}} \leq _{\mathrm {W}} \mathsf {KL}$ , yet $\mathsf {KL} \nleq _{\mathrm {W}} \mathsf {lim}$ , hence the claim follows from Corollary 4.4.

Similarly, the negative answer to [Reference Goh, Pauly and Valenti18, Question 6.2] can be obtained as follows.

Corollary 4.6. $\mathsf {wList}_{2^{\mathbb {N}},\leq \omega } \nleq _{\mathrm {W}} \mathsf {BS}$ .

Proof. As proved in [Reference Kihara, Marcone and Pauly17, Propositions 6.12 and 6.13], $\mathsf {wList}_{2^{\mathbb {N}},\leq \omega }$ is parallelizable and $\mathsf {ACC}_{\mathbb {N}} \leq _{\mathrm {W}} \mathsf {lim} \leq _{\mathrm {W}} \mathsf {wList}_{2^{\mathbb {N}},\leq \omega }$ . On the other hand, $\mathsf {wList}_{2^{\mathbb {N}},\leq \omega } \nleq _{\mathrm {W}} \mathsf {lim}$ [Reference Kihara, Marcone and Pauly17, Corollary 6.16], hence the claim follows from Corollary 4.4.

Continuous Weihrauch reducibility ( $\leq _{\mathrm {W}}^*$ ) is a variant of Weihrauch reducibility defined via continuous functions in place of computable functions. Using a recent result of Pauly and Soldà [Reference Pauly, Soldà, Patey, Pimentel, Galeotti and Manea25], we can characterize the parallelizations of first-order problems which are reducible to $\mathsf {DS}$ up to continuous Weihrauch reducibility.

Corollary 4.7. If $\widehat {f} \leq _{\mathrm {W}}^* \mathsf {DS}$ , then ${}^1f \leq _{\mathrm {W}}^* \mathsf {C}_{\mathbb {N}}$ . Therefore, for any first-order f,

$$\begin{align*}\widehat{f} \leq_{\mathrm{W}}^* \mathsf{DS} \qquad \text{if and only if} \qquad f \leq_{\mathrm{W}}^* \mathsf{C}_{\mathbb{N}}. \end{align*}$$

Proof. If ${}^1f$ is continuous, the conclusion of the first statement is satisfied. Otherwise, $\mathsf {ACC}_{\mathbb {N}} \leq _{\mathrm {W}}^* f$ by [Reference Pauly, Soldà, Patey, Pimentel, Galeotti and Manea25, Theorem 1]. The relativization of Theorem 4.3 then implies $\widehat {f} \leq _{\mathrm {W}}^* \mathsf {lim}$ . We conclude ${}^1f \leq _{\mathrm {W}}^* {}^1\mathsf {lim} \equiv _{\mathrm {W}} \mathsf {C}_{\mathbb {N}}$ . The second statement then follows from $\widehat {\mathsf {C}_{\mathbb {N}}} \equiv _{\mathrm {W}} \mathsf {lim} \leq _{\mathrm {W}} \mathsf {DS}$ .

5 The finitary part and deterministic part of $\mathsf {BS}$

In this section, we show that $\mathsf {BS}$ and $\mathsf {DS}$ cannot be separated by looking at their respective finitary or deterministic parts. Recall from [Reference Goh, Pauly and Valenti18, Theorems 4.16 and 4.31] that $\operatorname {Det}\!\left (\mathsf {DS}\right ) \equiv _{\mathrm {W}} \mathsf {lim}$ and $\operatorname {Fin}_{k}\!\left (\mathsf {DS}\right ) \equiv _{\mathrm {W}} \mathsf {RT}^{{1}}_{{k}}$ . Since both the deterministic and the finitary parts are monotone, this implies that $\mathsf {lim} \leq _{\mathrm {W}} \operatorname {Det}\!\left (\mathsf {BS}\right )$ and $\mathsf {RT}^{{1}}_{{k}}\leq _{\mathrm {W}} \operatorname {Fin}_{k}\!\left (\mathsf {BS}\right )$ , so we only need to show that the converse reductions hold.

Let $(Q,\preceq _Q)$ be a countable quasi-order. We call a subset $A \subseteq Q$ dense if for every $w \in Q$ there is some $u \succeq _Q w$ with $u \in A$ . We call it upwards-closed if $w \in A$ and $w \preceq _Q u$ implies $u \in A$ . By $\boldsymbol {\Sigma }^1_1\mathsf {-DUCC}$ we denote the following problem: Given a (non-empty) quasi-order $(Q,\preceq _Q)$ and a $\boldsymbol {\Sigma }^1_1$ -code for a dense upwards-closed subset $A \subseteq Q$ , find some element of A. We can think of a $\boldsymbol {\Sigma }^1_1$ -code for A as a sequence $(T_n)_{{n\in Q}}$ of subtrees of $\mathbb {N}^{<\mathbb {N}}$ such that $T_n$ is ill-founded if and only if $n \in A$ .

Proposition 5.1. ${}^1\mathsf {BS} \leq _{\mathrm {W}} \boldsymbol {\Sigma }^1_1\mathsf {-DUCC}$ .

Proof. Let f be a problem with codomain $\mathbb {N}$ and assume $f\leq _{\mathrm {W}} \mathsf {BS}$ via $\Phi ,\Psi $ . Fix $x\in \operatorname {dom}(f)$ . Let $(P,\leq _P)$ denote the non-well partial order defined by $\Phi (x)$ . We say that a finite $\leq _P$ -bad sequence $\beta $ is sufficiently long if $\Psi (x, \beta )$ returns a natural number in at most $|{\beta }|$ steps.

Consider the non-empty $\Sigma ^{1,x}_1$ set of sufficiently long finite extendible bad sequences. To show that $f\leq _{\mathrm {W}} \boldsymbol {\Sigma }^1_1\mathsf {-DUCC}$ , it is enough to notice that Lemma 4.2 implies that the aforementioned set is dense and upwards-closed with respect to the quasi-order $\trianglelefteq ^P$ (Definition 4.1).

Lemma 5.2. $\boldsymbol {\Sigma }^1_1\mathsf {-DUCC} \equiv _{\mathrm {W}} \boldsymbol {\Sigma }^1_1\mathsf {-DUCC}(2^{<\mathbb {N}}, \cdot )$ , where the latter denotes the restriction of the former to $2^{<\mathbb {N}}$ with the prefix ordering $\sqsubseteq $ .

Proof. Clearly, we only need to show that $\boldsymbol {\Sigma }^1_1\mathsf {-DUCC} \leq _{\mathrm {W}} \boldsymbol {\Sigma }^1_1\mathsf {-DUCC}(2^{<\mathbb {N}}, \cdot )$ . Suppose we are given as input a countable quasi-order $(Q,\preceq _Q)$ and a $\boldsymbol {\Sigma }^1_1$ -code for a set $A \subseteq Q$ which is dense and upwards-closed. We shall define a labelling which is computable uniformly in $(Q,\preceq _Q)$ , such that $\lambda ^{-1}(A)$ is dense and upwards-closed. This suffices to prove the claimed reduction, as the preimage of A via $\lambda $ is uniformly $\Sigma ^1_1$ relative to the input and, given $\sigma \in \lambda ^{-1}(A)$ , we may use the input $(Q,\preceq _Q)$ to compute $\lambda (\sigma )\in A$ .

The labelling $\lambda : 2^{<\mathbb {N}} \to Q$ is defined via an auxiliary function (useful to describe $\lambda $ in a recursive manner). First, for every $i \in \mathbb {N}$ , $\overline {\lambda }(x, 0^i):=\overline {\lambda }(x, 0^i1):= x$ . To evaluate $\overline {\lambda }$ on other strings, we fix a standard Q-computable enumeration $(x_n)_{{n}}$ of Q and define, recursively:

We then define $\lambda (\sigma ):=\overline {\lambda }(x_0,\sigma )$ . It is clear that $\lambda $ is total and it is uniformly computable in $(Q,\preceq _Q)$ . Rephrasing the definition, we can observe that, for every x and $\sigma $ , . Besides, . In particular,

Intuitively, $\lambda $ guides the construction of an ascending $\preceq _Q$ -sequence from $x_0$ . Indeed, we will show that $\lambda $ is weakly monotone, and therefore each $\sigma $ induces a finite ascending $\preceq _Q$ -sequence from $x_0$ to $\lambda (\sigma )$ . At the same time, the labelling is designed to be highly redundant, which we use to prove that $\lambda ^{-1}(A)$ is dense.

To show that $\lambda ^{-1}(A)$ is upwards-closed, we claim that for every $\sigma ,\tau \in 2^{<\mathbb {N}}$ , . The claim is proved by first using structural induction on $\sigma $ to show that for all $x \in Q$ , we have $x \preceq _Q \overline {\lambda }(x,\sigma )$ . Using this fact, one can perform another structural induction on $\sigma $ to show that for all x and $\tau $ , we have . Taking $x = x_0$ yields . Since A is upwards-closed in Q, it follows that $\lambda ^{-1}(A)$ is upwards-closed in the prefix ordering.

To prove that $\lambda ^{-1}(A)$ is dense, fix $\rho \in 2^{<\mathbb {N}}$ and assume $\lambda (\rho )\notin A$ . Since A is dense, there is some n such that $\lambda (\rho )\preceq _Q x_n \in A$ . Observe that $\overline {\lambda }(\lambda (\rho ), 0^n11) = \overline {\lambda }(x_n, \varepsilon ) = x_n$ , so it suffices to find some $\rho ' \sqsupseteq \rho $ such that $\lambda (\rho ') = \overline {\lambda }(\lambda (\rho ), 0^n11)$ . The construction of $\rho '$ depends on the form of $\rho $ (all the following claims can be proven by structural induction on $\rho $ ):

  • if $\rho $ ends with $0$ , then , so ;

  • if $\rho $ ends with an odd number of $1$ s, then , so ;

  • otherwise, $\rho $ is either $\varepsilon $ or ends with a positive even number of $1$ s. Then , so .

In all cases, we have found an extension of $\rho $ which maps to $x_n$ , as desired.

Theorem 5.3. $\operatorname {Fin}_{k}\!\left (\mathsf {BS}\right ) \equiv _{\mathrm {W}} \operatorname {Fin}_{k}\!\left (\boldsymbol {\Sigma }^1_1\mathsf {-DUCC}\right ) \equiv _{\mathrm {W}} \operatorname {Fin}_{k}\!\left (\mathsf {DS}\right ) \equiv _{\mathrm {W}} \mathsf {RT}^{{1}}_{{k}}$ for every k.

Proof. We have $\mathsf {RT}^{{1}}_{{k}} \equiv _{\mathrm {W}} \operatorname {Fin}_{k}\!\left (\mathsf {DS}\right ) \leq _{\mathrm {W}} \operatorname {Fin}_{k}\!\left (\mathsf {BS}\right ) \leq _{\mathrm {W}} \operatorname {Fin}_{k}\!\left (\boldsymbol {\Sigma }^1_1\mathsf {-DUCC}\right )$ by [Reference Goh, Pauly and Valenti18, Theorem 4.31] and Proposition 5.1. It remains to show that $\operatorname {Fin}_{k}\!\left (\boldsymbol {\Sigma }^1_1\mathsf {-DUCC}\right ) \leq _{\mathrm {W}} \mathsf {RT}^{{1}}_{{k}}$ . By Lemma 5.2, it is enough to show that $\operatorname {Fin}_{k}\!\left (\boldsymbol {\Sigma }^1_1\mathsf {-DUCC}(2^{<\mathbb {N}}, \cdot )\right ) \leq _{\mathrm {W}} \mathsf {RT}^{{1}}_{{k}}$ .

Let f be a problem with codomain k and assume $f\leq _{\mathrm {W}}\boldsymbol {\Sigma }^1_1\mathsf {-DUCC}(2^{<\mathbb {N}}, \cdot )$ via $\Phi ,\Psi $ . Observe that every $x\in \operatorname {dom}(f)$ induces a coloring as follows: run $\Psi (x,\sigma )$ in parallel on every $\sigma \in 2^{<\mathbb {N}}$ . Whenever we see that $\Psi (x,\sigma )$ returns a number less than k, we define $c(\tau ):=\Psi (x,\sigma )$ for every $\tau \sqsubseteq \sigma $ such that $c(\tau )$ is not defined yet. By density of the set coded by $\Phi (x)$ , c is total.

By the Chubb–Hirst–McNicholl tree theorem [Reference Chubb, Hirst and McNicholl8], there is some $\rho \in 2^{<\mathbb {N}}$ and some color $i<k$ such that i appears densely above $\rho $ . We claim that such i would be an f-solution to x. To prove this, fix (by density) some $\rho ' \sqsupseteq \rho $ which lies in the set coded by $\Phi (x)$ . Then fix some $\tau \sqsupseteq \rho '$ with color i. Finally, fix some $\sigma \sqsupseteq \tau $ such that $c(\tau )$ was defined to be $\Psi (x,\sigma )$ . Now $\sigma $ lies in the set coded by $\Phi (x)$ (by upwards-closure), so $\Psi (x,\sigma )$ is an f-solution to x. Since $\Psi (x,\sigma ) = c(\tau ) = i$ , we conclude that i is an f-solution to x.

This implies that f reduces to the problem “given a k-coloring of $2^{<\mathbb {N}}$ , find i such that, for some $\sigma $ , i appears densely above $\sigma $ ”. We claim that the latter can be solved using $\mathsf {RT}^{{1}}_{{k}}$ . This follows from the fact that, as shown in [Reference Pauly, Pradic and Soldà24, Corollary 42], $\mathsf {RT}^{{1}}_{{k}}$ can solve the problem $(\mathsf {c}(\eta )^1_{<\infty })_k$ defined as “given a k-coloring , produce $i<k$ such that, for some interval I with rational endpoints, $\alpha ^{-1}(i)$ is dense in I”.

We introduce the order $\prec $ on $2^{<\mathbb {N}}$ defined by setting for all $\sigma , \tau , \rho \in 2^{<\mathbb {N}}$ . This is a computable dense linear order with no endpoints, therefore there is a computable order isomorphism $\varphi $ between $(2^{<\mathbb {N}}, \prec )$ and ${(\mathbb {Q},<)}$ .

Every open interval in $(\mathbb {Q},<)$ contains the $\varphi $ -image of an upper cone in $(2^{<\mathbb {N}}, \prec )$ and vice versa. Specifically, the interval $(\varphi (\sigma ), \varphi (\tau ))$ contains the image of the cone above if is not a prefix of $\tau $ , and the image of the cone above if is a prefix of $\tau $ . Thus, by translating the coloring along $\varphi $ to $\mathbb {Q}$ , we find that a color which appears densely in an interval in $\mathbb {Q}$ also appears densely in a cone in $2^{<\mathbb {N}}$ . This means we can use $(\mathsf {c}(\eta )^1_{<\infty })_k$ to find a color appearing densely in a cone.

It is immediate from the previous theorem that the finitary parts of $\mathsf {BS}$ and $\mathsf {DS}$ in the sense of Cipriani and Pauly [Reference Cipriani and Pauly10, Definition 2.10] agree as well. Finally, we shall prove that the deterministic parts of $\mathsf {BS}$ and $\mathsf {DS}$ agree.

Lemma 5.4. If $\operatorname {Fin}_{2}\!\left (f\right ) \leq _{\mathrm {W}} \mathsf {RT}^{{1}}_{{2}}$ , then $\operatorname {Det}\!\left (f\right ) \leq _{\mathrm {W}} \mathsf {lim}$ .

Proof. By algebraic properties of $\operatorname {Det}\!\left (\cdot \right )$ , $\operatorname {Fin}_{k}\!\left (\cdot \right )$ , and the parallelization, we have

$$\begin{align*}\operatorname{Det}\!\left(f\right) \leq_{\mathrm{W}} \widehat{\mathrm{Det}_{ \mathbf{2} }(f)} \leq_{\mathrm{W}} \widehat{\operatorname{Fin}_{2}\!\left(f\right)} \leq_{\mathrm{W}} \widehat{\mathsf{RT}^{{1}}_{{2}}}. \end{align*}$$

Since $\operatorname {Det}\!\left (f\right )$ is deterministic, it remains to show that $\operatorname {Det}\!\left (\widehat {\mathsf {RT}^{{1}}_{{2}}}\right ) \leq _{\mathrm {W}} \mathsf {lim}$ . This is known, but for completeness we give a proof using results from the survey of Brattka, Gherardi, and Pauly [Reference Brattka, Gherardi, Pauly, Brattka and Hertling3]:

$$ \begin{align*} \widehat{\mathsf{RT}^{{1}}_{{2}}} &\equiv_{\mathrm{sW}} \widehat{\mathsf{C}_{2}'} & [3, Theorem 11.8.11, Proposition 11.6.10] \\ &\equiv_{\mathrm{sW}} \left(\widehat{\mathsf{C}_{2}}\right)' & [3, Proposition 11.6.12] \\ &\equiv_{\mathrm{sW}} \mathsf{C}_{{2^{\mathbb{N}}}}' & [3, Theorem 11.7.23] \\ &\equiv_{\mathrm{W}} \mathsf{C}_{{2^{\mathbb{N}}}} * \mathsf{lim} & [3, Proposition 11.7.6, Proposition 11.6.14]. \end{align*} $$

By choice elimination (see [Reference Brattka, Gherardi, Pauly, Brattka and Hertling3, Theorem 11.7.25] or [Reference Goh, Pauly and Valenti18, Theorem 3.9]), $\operatorname {Det}\!\left (\mathsf {C}_{{2^{\mathbb {N}}}} * \mathsf {lim}\right ) \leq _{\mathrm {W}} \mathsf {lim}$ . Therefore $\operatorname {Det}\!\left (\widehat {\mathsf {RT}^{{1}}_{{2}}}\right ) \leq _{\mathrm {W}} \mathsf {lim}$ .

Since $\operatorname {Det}\!\left (\mathsf {BS}\right ) \geq _{\mathrm {W}} \operatorname {Det}\!\left (\mathsf {DS}\right ) \equiv _{\mathrm {W}} \mathsf {lim}$ [Reference Goh, Pauly and Valenti18, Theorem 4.16], we conclude that:

Corollary 5.5. $\operatorname {Det}\!\left (\mathsf {BS}\right )\equiv _{\mathrm {W}} \operatorname {Det}\!\left (\mathsf {DS}\right )\equiv _{\mathrm {W}} \mathsf {lim}$ .

Corollary 5.6. $\mathrm {Det}_{ \mathbf {\mathbb {N}} }(\mathsf {BS}) \equiv _{\mathrm {W}} \mathsf {C}_{\mathbb {N}}$ .

Proof. Since $\mathbb {N}$ computably embeds in $\mathbb {N}^{\mathbb {N}}$ , for every problem f we have $\mathrm {Det}_{ \mathbf {\mathbb {N}} }(f)\leq _{\mathrm {W}} \operatorname {Det}\!\left (f\right )$ . In particular, by Corollary 5.5, $\mathrm {Det}_{ \mathbf {\mathbb {N}} }(\mathsf {BS}) \leq _{\mathrm {W}}\operatorname {Det}\!\left (\mathsf {BS}\right ) \equiv _{\mathrm {W}} \mathsf {lim}$ . Since ${}^1\mathsf {lim}\equiv _{\mathrm {W}} \mathsf {C}_{\mathbb {N}}$ ([Reference Brattka, Gherardi and Marcone2, Proposition 13.10], see also [Reference Soldà and Valenti28, Theorem 7.2]), this implies $\mathrm {Det}_{ \mathbf {\mathbb {N}} }(\mathsf {BS})\leq _{\mathrm {W}} \mathsf {C}_{\mathbb {N}}$ . The converse reduction follows from the fact that $\mathsf {C}_{\mathbb {N}}\equiv _{\mathrm {W}} \mathrm {Det}_{ \mathbf {\mathbb {N}} }(\mathsf {DS})$ [Reference Goh, Pauly and Valenti18, Proposition 4.14].

We remark that for establishing $\operatorname {Fin}_{k}\!\left (\boldsymbol {\Sigma }^1_1\mathsf {-DUCC}\right ) \leq _{\mathrm {W}} \mathsf {RT}^{{1}}_{{k}}$ in Theorem 5.3 it was immaterial that the set of correct solutions was provided as a $\boldsymbol {\Sigma }^1_1$ -set. If we consider any other represented point class $\boldsymbol {\Gamma }$ which is effectively closed under taking preimages under computable functions, and define $\boldsymbol {\Gamma }\mathsf {-DUCC}$ in the obvious way, we can obtain the following.

Corollary 5.7. $\operatorname {Fin}_{k}\!\left (\boldsymbol {\Gamma }\mathsf {-DUCC}\right ) \leq _{\mathrm {W}} \mathsf {RT}^{{1}}_{{k}}$ .

This observation could be useful, e.g., for exploring the Weihrauch degree of finding bad arrays in non-better-quasi-orders (cf. [Reference Freund, Pakhomov and Soldà14]).

Proposition 5.8. $\mathsf {C}_{\mathbb {N}} \equiv _{\mathrm {W}} \operatorname {Det} (\boldsymbol {\Pi }^1_1\mathsf {-Bound}) \equiv _{\mathrm {W}} \operatorname {Det}({}^1\mathsf {BS}) \equiv _{\mathrm {W}} \operatorname {Det}(\boldsymbol {\Sigma }^1_1\mathsf {-DUCC})$ .

Proof. The reductions $\mathsf {C}_{\mathbb {N}} \leq _{\mathrm {W}} \operatorname {Det} (\boldsymbol {\Pi }^1_1\mathsf {-Bound}) \leq _{\mathrm {W}} \operatorname {Det} ({}^1\mathsf {BS} )\leq _{\mathrm {W}} \operatorname {Det} (\boldsymbol {\Sigma }^1_1 \mathsf {-DUCC} )$ are straightforward: the first one follows from $\mathsf {C}_{\mathbb {N}}\equiv _{\mathrm {W}} \mathsf {UC}_{\mathbb {N}}$ and $\mathsf {C}_{\mathbb {N}}\leq _{\mathrm {W}} \boldsymbol {\Pi }^1_1\mathsf {-Bound}$ , the second one follows from $\boldsymbol {\Pi }^1_1\mathsf {-Bound}\leq _{\mathrm {W}} {}^1\mathsf {BS}$ , and the last one follows from Proposition 5.1. It remains to show that $\operatorname {Det} (\boldsymbol {\Sigma }^1_1\mathsf {-DUCC} )\leq _{\mathrm {W}} \mathsf {C}_{\mathbb {N}}$ . In light of Lemma 5.2, it suffices to show that if $f:\subseteq \mathbb {N}^{\mathbb {N}} \to \mathbb {N}^{\mathbb {N}}$ is such that $f\leq _{\mathrm {W}} \boldsymbol {\Sigma }^1_1\mathsf {-DUCC}(2^{<\mathbb {N}}, \cdot )$ via $\Phi ,\Psi $ then $f\leq _{\mathrm {W}}\mathsf {C}_{\mathbb {N}}$ . Given $x\in \operatorname {dom}(f)$ , we can compute the set

$$ \begin{align*} X:=\{ \sigma \in 2^{<\mathbb{N}} {}\,:\,{} & (\forall \tau_0,\tau_1\sqsupseteq \sigma)(\forall n\in\mathbb{N}) \\ & ((\Psi(x,\tau_0)(n)\downarrow \land \Psi(x,\tau_1)(n)\downarrow)\rightarrow \Psi(x,\tau_0)(n)=\Psi(x,\tau_1)(n)) \}. \end{align*} $$

Observe that this is a $\Pi ^{0,x}_1$ subset of $2^{<\mathbb {N}}$ and it is non-empty: indeed, letting $A\subseteq 2^{<\mathbb {N}}$ be the set described by $\Phi (x)$ , the fact that f is deterministic implies that $A\subseteq X$ . We can therefore use $\mathsf {C}_{\mathbb {N}}$ to compute some $\sigma \in X$ . Observe that, by the density of A, there is $\tau \sqsupseteq \sigma $ such that $\tau \in A$ . In particular, $\Psi (x,\sigma )=\Psi (x,\tau )\in f(x)$ .

Corollary 5.5 and Proposition 5.5 imply that ${}^1\operatorname {Det}\!\left (\mathsf {BS}\right ) \equiv _{\mathrm {W}} {}^1\mathsf {lim} \equiv _{\mathrm {W}} \mathsf {C}_{\mathbb {N}} \equiv _{\mathrm {W}} \operatorname {Det}\!\left ({}^1\mathsf {BS}\right )$ . In general, ${}^1\operatorname {Det}\!\left (f\right ) \leq _{\mathrm {W}} \operatorname {Det}\!\left ({}^1f\right )$ [Reference Goh, Pauly and Valenti18, Proposition 3.4]. Goh, Pauly, and Valenti [Reference Goh, Pauly and Valenti18, Question 3.5] asked if there is some problem f such that ${}^1\operatorname {Det}\!\left (f\right ) <_{\mathrm {W}} \operatorname {Det}\!\left ({}^1f\right )$ . The next section is devoted to an affirmative answer (Theorem 6.6).

6 Gödelization

In this section, we show that the first-order part operator and the deterministic part operator do not commute, answering our earlier question [Reference Goh, Pauly and Valenti18, Question 3.5]. For this, we introduce a dual to the first-order part, the Gödelization operator, defined below. Our starting point is Dzhafarov, Solomon, and Yokoyama [Reference Dzhafarov, Solomon and Yokoyama13, Theorem 1.6] which asserts that a problem is computably true (i.e., for every instance p there exists a solution q with $q \leq _{\mathrm {T}} p$ ) if and only if it is Weihrauch reducible to some first-order problem. The construction in their proof shows even more:

Theorem 6.1. If g is a problem which is computably true, then

$$\begin{align*}\min_{\leq_{\mathrm{W}}}\{f \geq_{\mathrm{W}} g {}\,:\,{} f \text{ is first-order}\} \end{align*}$$

exists and is represented by

$$\begin{align*}g^{\mathsf{G}}(p) = \{e \in \mathbb{N} {}\,:\,{} \Phi_e(p) \in g(p)\}. \end{align*}$$

Proof. Notice $\operatorname {dom}(g^{\mathsf {G}}) = \operatorname {dom}(g)$ since g is computably true. It is clear that $g \leq _{\mathrm {W}} g^{\mathsf {G}}$ . Next, suppose h has codomain $\mathbb {N}$ and $g \leq _{\mathrm {W}} h$ via $\Gamma $ and $\Delta $ . To reduce $g^{\mathsf {G}}$ to h, observe that if $p \in \operatorname {dom}(g^{\mathsf {G}})$ , then $\Gamma (p) \in \operatorname {dom}(h)$ , and if $n \in h(\Gamma (p))$ , then $\Delta (p,n) \in g(p)$ . Using an index for $\Delta $ , we can define a functional $\Psi $ such that $\Phi _{\Psi (p,n)}(p) = \Delta (p,n)$ for every $p \in \mathbb {N}^{\mathbb {N}}$ and $n \in \mathbb {N}$ . Then if $n \in h(\Gamma (p))$ , we have $\Psi (p,n) \in g^{\mathsf {G}}(p)$ as desired. (Indeed $\Psi $ does not even need access to p, so $g^{\mathsf {G}} \leq _{\mathrm {sW}} h$ .)

Therefore, every problem g which is computably true but not first-order lies in an interval with no first-order problems, bounded below by its first-order part ${}^1g$ and above by $g^{\mathsf {G}}$ . We choose to call $g^{\mathsf {G}}$ the Gödelization of g, inspired by the Gödel numbering problem $\mathsf {G}$ studied by Brattka [Reference Brattka, Vedova, Dundua, Lempp and Manea1].

Proposition 6.2. If g is computably true, single-valued, and $g^{\mathsf {G}}$ is not deterministic (i.e., Weihrauch equivalent to a single-valued function with codomain $\mathbb {N}^{\mathbb {N}}$ ), then ${}^1\operatorname {Det}\!\left (g^{\mathsf {G}}\right ) <_{\mathrm {W}} \operatorname {Det}\!\left (g^{\mathsf {G}}\right )$ .

Proof. Since g is single-valued and $g \leq _{\mathrm {W}} g^{\mathsf {G}}$ , we have $g \leq _{\mathrm {W}} \operatorname {Det}\!\left (g^{\mathsf {G}}\right ) <_{\mathrm {W}} g^{\mathsf {G}}$ . So $\operatorname {Det}\!\left (g^{\mathsf {G}}\right )$ cannot be first-order, by Theorem 6.1. We conclude that ${}^1\operatorname {Det}\!\left (g^{\mathsf {G}}\right ) <_{\mathrm {W}} \operatorname {Det}\!\left (g^{\mathsf {G}}\right )$ .

In Proposition 6.5, we shall construct such g using the following two ingredients.

Proposition 6.3. If f is deterministic, then it is Weihrauch equivalent to one of its realizers.

Proof. Suppose $h : \subseteq \mathbb {N}^{\mathbb {N}} \to \mathbb {N}^{\mathbb {N}}$ satisfies $f \equiv _{\mathrm {W}} h$ . Fix $\Phi $ and $\Psi $ witnessing that $f \leq _{\mathrm {W}} h$ . We define $r : \subseteq \mathbb {N}^{\mathbb {N}} \to \mathbb {N}^{\mathbb {N}}$ by $r(x) = \Psi (x,h(\Phi (x)))$ . By construction, r is a realizer of f, thus it in particular holds that $f \leq _{\mathrm {W}} r$ . Also by construction, $r \leq _{\mathrm {W}} h$ . Since $h \equiv _{\mathrm {W}} f$ , it follows that $r \equiv _{\mathrm {W}} f$ .

The second ingredient is the following Rice-like result, which forms the recursion-theoretic core of our construction. For each $e \in \mathbb {N}$ , let $\phi _e$ denote the e-th partial computable function on $\mathbb {N}$ . Let $\mathrm {Tot}$ denote the set of $e \in \mathbb {N}$ such that $\phi _e$ is total.

Lemma 6.4. There is no computable $F: \subseteq \mathbb {N} \to \mathbb {N}$ such that whenever $e,e' \in \mathrm {Tot}$ , then $F(e) = F(e')$ if and only if $\phi _e = \phi _{e'}$ . In fact, fix $e_0 \in \mathrm {Tot}$ and define $A := \{e \in \mathbb {N} {}\,:\,{} \phi _e = \phi _{e_0}\}$ . Then no c.e. set containing A is disjoint from $\mathrm {Tot}\setminus A$ .

Proof. Suppose towards a contradiction that $U \supseteq A$ is c.e. and disjoint from $\mathrm {Tot}\setminus A$ . Then we can compute halting as follows. Given e, compute an index $f(e)$ such that $\phi _{f(e)}$ copies $\phi _{e_0}$ until $\phi _e(e)$ halts (if ever), after which $\phi _{f(e)}$ flips each output of $\phi _{e_0}$ . Notice if $\phi _e(e)$ halts, then $f(e) \in \mathrm {Tot}\setminus A$ , else $f(e) \in A \subseteq U$ . So we can compute whether $\phi _e(e)$ halts by waiting for it to halt, or for $f(e)$ to appear in U.

Proposition 6.5. There is a problem which is computably true and single-valued, such that $g^{\mathsf {G}}$ is not deterministic. (In particular, g cannot be first-order, else $g^{\mathsf {G}} \equiv _{\mathrm {W}} g$ is deterministic.)

Proof. We shall construct a sequence of computable reals $(g_n)_{n \in \mathbb {N}}$ such that for every pair of Turing functionals $\Gamma $ and $\Delta $ (which form a potential Weihrauch reduction from some realizer of $g^{\mathsf {G}}$ to $g^{\mathsf {G}}$ itself), one of the following holds:

  1. (1) some $\Gamma (n)$ is not a name for a natural number (i.e., an instance of g as defined below);

  2. (2) there are n, e, $e' \in \mathbb {N}$ such that $\Phi _e(\Gamma (n)) = g_{\Gamma (n)} = \Phi _{e'}(\Gamma (n))$ but $\Delta (n,e) \neq \Delta (n,e');$

  3. (3) there are n, $e \in \mathbb {N}$ such that $\Phi _e(\Gamma (n)) = g_{\Gamma (n)}$ but $\Phi _{\Delta (n,e)}(n) \neq g_n$ .

If we then define by $g(n) := g_n$ , we see that g is single-valued, computably true, and $g^{\mathsf {G}}$ is not Weihrauch equivalent to any of its realizers (hence not deterministic, by Proposition 6.3).

We shall construct $(g_n)_{n \in \mathbb {N}}$ in stages. At each stage, we handle a single pair $\Gamma $ and $\Delta $ by defining at most two new computable reals $g_n$ . We assume that for every $n \in \mathbb {N}$ , $\Gamma (n)$ names a natural number (otherwise no action is needed). Consider the following cases.

Case 1. For some $n \in \mathbb {N}$ such that $\Gamma (n) \neq n$ , $g_n$ has not been defined. By defining $g_{\Gamma (n)}$ if necessary, we may fix some e such that $\Phi _e(\Gamma (n)) = g_{\Gamma (n)}$ . Since $\Gamma (n) \neq n$ , we may then define $g_n$ to be a computable real which differs from $\Phi _{\Delta (n,e)}(n)$ (regardless of whether the latter is total), ensuring (3).

Case 2. Fix some $n \in \mathbb {N}$ such that $g_n$ is not defined. Since Case 1 fails, $\Gamma (n) = n$ .

Case 2a. There are $e,e' \in \mathbb {N}$ such that $\Phi _e(\Gamma (n)) = \Phi _{e'}(\Gamma (n))$ and $\Delta (n,e) \neq \Delta (n,e')$ (this includes the case where either or both sides diverge). We define $g_{\Gamma (n)} = \Phi _e(\Gamma (n))$ , ensuring (2).

Case 2b. If there is some $e \in \mathbb {N}$ such that $\Phi _e(\Gamma (n))$ is total and not equal to $\Phi _{\Delta (n,e)}(n)$ , we define $g_n = \Phi _e(\Gamma (n))$ . This ensures (3).

To complete the current stage of the construction, we shall show that the above cases encompass all possibilities. Suppose otherwise. Since Case 2b fails, for all $e \in \mathbb {N}$ , whenever $\Phi _e(\Gamma (n))$ is total, so is $\Phi _{\Delta (n,e)}(n)$ and they must be equal. Since Case 2a fails, whenever $\Phi _e(\Gamma (n)) = \Phi _{e'}(\Gamma (n))$ , we have $\Delta (n,e) = \Delta (n,e')$ . We deduce that whenever $\Phi _e(\Gamma (n))$ and $\Phi _{e'}(\Gamma (n))$ are total, $\Delta (n,e) = \Delta (n,e')$ if and only if $\Phi _e(\Gamma (n)) = \Phi _{e'}(\Gamma (n))$ . This contradicts Lemma 6.4: Define $F: \subseteq \mathbb {N} \to \mathbb {N}$ by $F(d) = \Delta (n,e)$ , where $\Phi _e$ upon input $\Gamma (n)$ merely simulates $\phi _d$ .

This completes the current stage of the construction. Observe we have ensured that one of (1)–(3) hold for $\Gamma $ and $\Delta $ .

We may now answer [Reference Goh, Pauly and Valenti18, Question 3.5] in the affirmative.

Theorem 6.6. There is a problem f such that ${}^1\operatorname {Det}\!\left (f\right ) <_{\mathrm {W}} \operatorname {Det}\!\left ({}^1f\right )$ .

Proof. By Proposition 6.5, fix a problem g which is computably true and single-valued, such that $g^{\mathsf {G}}$ is not deterministic. By Proposition 6.2, ${}^1\operatorname {Det}\!\left (g^{\mathsf {G}}\right ) <_{\mathrm {W}} \operatorname {Det}\!\left (g^{\mathsf {G}}\right )$ . Let $f = g^{\mathsf {G}}$ . Then ${}^1\operatorname {Det}\!\left (f\right ) \equiv _{\mathrm {W}} {}^1\operatorname {Det}\!\left (g^{\mathsf {G}}\right ) <_{\mathrm {W}} \operatorname {Det}\!\left (g^{\mathsf {G}}\right ) \equiv _{\mathrm {W}} \operatorname {Det}\!\left ({}^1f\right )$ , where the last equivalence holds because f is first-order and $\operatorname {Det}\!\left (\cdot \right )$ is degree-theoretic.

7 $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ and $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$

A crucial role in the separation between $\mathsf {BS}$ and $\mathsf {DS}$ in Section 3 is played by partial orders that admit a tree decomposition. It is therefore natural to ask what is the strength of the restriction $\mathsf {BS}|_{\operatorname {Tree}}$ of $\mathsf {BS}$ to trees. In order to present our results about the first-order and finitary parts of $\mathsf {BS}|_{\operatorname {Tree}}$ we first prove some results about $\boldsymbol {\Pi }^0_2$ all or co-unique choice, $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ . This problem has also been studied in [Reference Pauly, Soldà, Patey, Pimentel, Galeotti and Manea25] in the context of continuous Weihrauch reducibility.

Definition 7.1. The problem $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ receives as input a $\boldsymbol {\Pi }^0_2$ -subset $A \subseteq \mathbf {k}$ with $|A| \geq k-1$ and returns some $i \in A$ . The problem $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ receives as input a $\boldsymbol {\Pi }^0_2$ -subset $A \subseteq \mathbb {N}$ with $|\mathbb {N} \setminus A| \leq 1$ and returns some $i \in A$ .

The relevance of these problems to $\mathsf {BS}|_{\operatorname {Tree}}$ will be explained in Section 8.

Rather than having to reason directly with $\boldsymbol {\Pi }^0_2$ -subsets, we can use the following more concrete characterization of the problem.

Proposition 7.2. Let $\mathbf {X} \in \{k, \mathbb {N}\}$ for $k \geq 2$ . Then $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbf {X}}$ is equivalent to the problem “Given some $p \in \mathbf {X}^{\mathbb {N}}$ , output some $n \in \mathbf {X}$ such that $n \neq \lim _{i \to \infty } p_i$ (which we understand to be true if the limit does not exist)”.

Proof. Given some $p \in \mathbf {X}^{\mathbb {N}}$ , we observe that $\{n \in \mathbf {X} {}\,:\,{} n \neq \lim _{i \to \infty } p_i\}$ is a $\Pi ^{0,p}_2$ -set omitting at most one element. That shows that $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbf {X}}$ can solve the problem of finding a non-limit point.

Conversely, let A be a $\boldsymbol {\Pi }^0_2$ -subset of $\mathbf {X}$ with $|\mathbf {X}\setminus A|\le 1$ . We can A-compute such that, for each $n\in \mathbf {X}$ , $q(n,\cdot )$ contains infinitely many $1$ s iff $n \in A$ . We define a sequence $p \in \mathbf {X}^{\mathbb {N}}$ as follows: for each $n\in \mathbf {X}$ and $k\in \mathbb {N}$ , let $M_{n,k}:= |\{ j \le k {}\,:\,{} q(n,j)=0 \}|$ .

For the sake of readability, we distinguish the cases where $\mathbf {X}$ is finite or not. If $\mathbf {X}=k$ , we define

$$\begin{align*}p_i:= \min \{ n < k {}\,:\,{} (\forall m < k)(M_{n,i}\ge M_{m,i})\}. \end{align*}$$

If $\mathbf {X}=\mathbb {N}$ , at each stage i, for each $n\le i$ we only see $j(n)$ -many elements of the sequence $(q(n,j))_{{j\in \mathbb {N}}}$ , where $j(n)$ is least such that $\langle n,j(n) \rangle>i$ . In this case, we define

$$\begin{align*}p_i:= \min \{ n \le i {}\,:\,{} (\forall m\le i)(M_{n,j(n)}\ge M_{m,j(m)})\}. \end{align*}$$

Observe that, in both cases, if $A = \mathbf {X} \setminus \{m\}$ for some m, then $\lim _{i\to \infty } p_i = m$ , hence any $n\neq \lim _{i\to \infty } p_i$ is a valid solution for $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbf {X}}(A)$ .

Observe that $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ can be equivalently thought of as the restriction of $\mathsf {RT}^{{1}}_{{k}}$ to colourings where at most one color does not appear infinitely often. In particular, $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{2} \equiv _{\mathrm {W}} \mathsf {RT}^{{1}}_{{2}}$ .

While in some sense $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ is a very weak principle – after all, there is just one single incorrect answer amongst all of the natural numbers – it is at the same time not particularly easy to solve, as evidenced by the following result.

Proposition 7.3. $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}} \nleq _{\mathrm {W}} \mathsf {lim}$ .

Proof. As $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ is a first-order problem, $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}} \leq _{\mathrm {W}} \mathsf {lim}$ would already imply $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}} \leq _{\mathrm {W}} \mathsf {C}_{\mathbb {N}}$ . The characterization in Proposition 7.2 shows that $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ is a closed fractal. By the absorption theorem for closed fractals [Reference Le Roux and Pauly20, Theorem 2.4], this means that $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}} \leq _{\mathrm {W}} \mathsf {C}_{\mathbb {N}}$ in turn would imply that $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ is computable, which is readily seen to be false.

The diamond operator [Reference Neumann and Pauly23, Definition 9] roughly captures the possibility of using a multi-valued function as oracle an arbitrary but finite number of times during a computation (with the additional requirement of having to declare, at some finite stage, that no more oracle calls will be made). It can be equivalently defined in terms of the following reduction game.

Definition 7.4 [Reference Hirschfeldt and Jockusch16, Definitions 4.1 and 4.3], see also [Reference Soldà and Valenti28, Definition 6.1].

Let $f,g:\subseteq \mathbb {N}^{\mathbb {N}} \rightrightarrows \mathbb {N}^{\mathbb {N}}$ be two partial multi-valued functions. We define the reduction game $G(f\to g)$ as the following two-player game: on the first move, Player 1 plays $x_0\in \operatorname {dom}(g)$ , and Player 2 either plays an $x_0$ -computable $y_0\in g(x_0)$ and declares victory, or responds with an $x_0$ -computable instance $z_1$ of f.

For $n> 1$ , on the n-th move (if the game has not yet ended), Player 1 plays a solution $x_{n-1}$ to the input $z_{n-1}\in \operatorname {dom}(f)$ . Then Player 2 either plays a -computable solution to $x_0$ and declares victory, or plays a -computable instance $z_n$ of f.

If at any point one of the players does not have a legal move, then the game ends with a victory for the other player. Player 2 wins if it ever declares victory (or if Player 1 has no legal move at some point in the game). Otherwise, Player 1 wins.

We define $f^\diamond :\subseteq \mathbb {N}\times \mathbb {N}^{\mathbb {N}} \rightrightarrows (\mathbb {N}^{\mathbb {N}})^{<\mathbb {N}}$ as the following problem:

  • $\operatorname {dom}(f^\diamond )$ is the set of pairs $(e,d)$ s.t. Player 2 wins the game $G(f\to \mathrm {id})$ when Player 1 plays d as his first move, and $\Phi _e$ is a winning strategy for Player 2;

  • a solution is the list of moves of Player $1$ for a run of the game.

Observe that $g \leq _{\mathrm {W}} f^\diamond $ iff Player 2 has a computable winning strategy for the game $G(f\to g)$ , i.e., if there is a Turing functional $\Phi $ such that Player 2 always plays , and wins independently of the strategy of Player 1.

We described the game assuming that $f,g$ have domain and codomain $\mathbb {N}^{\mathbb {N}}$ . The definition can be extended to arbitrary multi-valued functions, and the moves of the players are names for the instances/solutions.

Proposition 7.5. $\operatorname {Fin}_{k}\!\left (\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+2}^\diamond \right ) \equiv _{\mathrm {W}} \mathrm {id}$ for every k.

Proof. Let $g:\subseteq \mathbb {N}^{\mathbb {N}} \rightrightarrows k$ be such that $g\leq _{\mathrm {W}}\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+2}^\diamond $ and let $\Phi $ be the computable functional witnessing the winning strategy for Player 2 for the game $G= G(\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+2}\to g)$ . For each input $x\in \operatorname {dom}(g)$ , we can build the c.e. tree T of all possible runs of the game G where Player 1 starts by playing x. More precisely, for $\sigma \in (k+2)^{<\mathbb {N}}$ and $i<k+2$ , we enumerate if $\sigma \in T$ and Player 2 does not declare victory on the $(|{\sigma }|+1)$ -th move. Moreover, if Player 2 declares victory on the $(|{\sigma }|+1)$ -th move, we enumerate in T and commit to never extend this branch further. Intuitively, if Player 2 declares victory then we reached a leaf of the tree, and the leaf has a special label that contains the last Player 2’s move, i.e., the solution to $g(x)$ obtained using $\Phi $ when the oracle answers are .

Observe that, each non-leaf $\sigma \in T$ has exactly $k+2$ -many children, and at least $k+1$ of them correspond to valid Player 1 moves (i.e., possible $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+2}$ -answers). Observe also that the tree T need not be well-founded and not all the leaves are necessarily labelled as such. Indeed, the hypothesis that Player 2 has a winning strategy only guarantees that the subtree $S\subseteq T$ corresponding to valid runs of the game is well-founded (and it is computable to tell if a string in S is a label).

By unbounded search, we can compute a well-founded tree $S\subseteq T$ such that, $\varepsilon \in S$ , each non-leaf $\sigma \in S$ has exactly $(k+1)$ -many children, and each leaf of S is of the form for some $\tau \in (k+2)^{<\mathbb {N}}$ and $\Phi (x, \tau )<k$ . To compute a solution for $g(x)$ we proceed recursively on the (finite) rank of S: each leaf of S is simply labelled with $\Phi (x, \tau )$ . To compute the label of a non-leaf $\sigma $ , observe that, by the pigeonhole principle, we can choose $i<k$ such that at least two of the $k+1$ children of $\sigma $ are labelled with i. We can then label $\sigma $ with i and move to the next stage.

We claim that the label of $\varepsilon $ is a valid solution for $g(x)$ . Indeed, assume $\tau \in S$ corresponds to a valid run of the game, namely for each $i<|{\tau }|$ , $\tau (i)$ is a valid $(n+2)$ -th move for Player 1 when Player 2 plays according to $\Phi $ and the first $(n+1)$ moves of Player 1 are . By induction, any such $\tau $ receives a label $i<k$ such that $i\in g(x)$ . The statement follows from the fact that $\varepsilon $ always represents a valid run of the game.

Note that the previous proof only uses that $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+2}$ is a problem with codomain $k+2$ having at most one incorrect answer for every input. Therefore, for every pointclass $\boldsymbol {\Gamma }\in \{\boldsymbol {\Sigma }^0_k,\boldsymbol {\Pi }^0_k,\boldsymbol {\Delta }^0_k, \boldsymbol {\Sigma }^1_1,\boldsymbol {\Pi }^1_1,\boldsymbol {\Delta }^1_1 \}$ , we immediately obtain the following.

Corollary 7.6. $\operatorname {Fin}_{k}\!\left (\widehat {\boldsymbol {\Gamma }\mathsf {-ACC}_{k+2}}\right ) \equiv _{\mathrm {W}} \mathrm {id}$ for every $k\ge 1$ .

Proof. This can be proved observing that, for every f, $\operatorname {Fin}_{k}\!\left (f\right ) \leq _{\mathrm {W}} {}^1f$ , and therefore $\operatorname {Fin}_{k}\!\left (f\right ) \equiv _{\mathrm {W}} \operatorname {Fin}_{k}\!\left ({}^1f\right )$ , as $\operatorname {Fin}_{k}\!\left (f\right )$ is the strongest problem with codomain k that is reducible to f. This implies that

$$\begin{align*}\operatorname{Fin}_{k}\!\left(\widehat{\boldsymbol{\Gamma}\mathsf{-ACC}_{k+2}}\right) \equiv_{\mathrm{W}} \operatorname{Fin}_{k}\!\left({}^1\left(\widehat{\boldsymbol{\Gamma}\mathsf{-ACC}_{k+2}} \right) \right) \leq_{\mathrm{W}} \operatorname{Fin}_{k}\!\left(\boldsymbol{\Gamma}\mathsf{-ACC}_{k+2}^\diamond \right)\leq_{\mathrm{W}} \mathrm{id}, \end{align*}$$

where the second inequality follows from the fact, for every first-order f, ${}^1\widehat {f}\leq _{\mathrm {W}} f^\diamond $ (see [Reference Soldà and Valenti28, Theorem 5.7 and Proposition 6.3]). The last inequality is the statement of Proposition 7.5.

Corollary 7.7. $\operatorname {Det}\!\left (\widehat {\boldsymbol {\Gamma }\mathsf {-ACC}_{k+2}}\right ) \equiv _{\mathrm {W}} \mathrm {id}$ for every $k\ge 2$ .

Proof. Proposition 2.1 and the definitions of deterministic/k-finitary part imply that, for every f and every $k\ge 2$ ,

$$\begin{align*}\operatorname{Det}\!\left(f\right) \leq_{\mathrm{W}} \widehat{\mathrm{Det}_{ \mathbf{k} }(f)} \leq_{\mathrm{W}} \widehat{\operatorname{Fin}_{k}\!\left(f\right)}. \end{align*}$$

Letting $f= \widehat {\boldsymbol {\Gamma }\mathsf {-ACC}_{k+2}}$ , the claim follows from Corollary 7.6.

We observe that Proposition 7.5 cannot be improved to show that $\operatorname {Fin}_{k}\!\left (\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}^\diamond \right )$ is computable.

Proposition 7.8. $\mathsf {ACC}_{k} \leq _{\mathrm {W}} \operatorname {Fin}_{k}\!\left (\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}\right )$ for every $k\ge 2$ .

Proof. We use the equivalent formulation of $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}$ introduced in Proposition 7.2. For every $A\in \operatorname {dom}(\mathsf {ACC}_k)$ , we uniformly compute a sequence $p\in \mathbb {N}^{\mathbb {N}}$ as follows:

$$\begin{align*}p(s):=\begin{cases} i & \text{if } i \text{ is enumerated outside of }A \text{ by stage }s \\ k & \text{otherwise.} \end{cases} \end{align*}$$

Let $n<k+1$ be such that $n\neq \lim _{s\to \infty } p(s)$ . If $n=k$ then $A\neq k$ , therefore we can computably find (by unbounded search) the unique $m \in k\setminus A$ . Conversely, if $n\neq k$ then it follows by the definition of p that $n\in A$ .

The following proposition shows that for $k> 2$ , the degree $\operatorname {Fin}_{k}\!\left (\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}\right )$ escapes a “nice” characterization. The exceptional case of $\operatorname {Fin}_{2}\!\left (\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{3}\right )$ is covered in Corollary 7.10 below.

Proposition 7.9. Let $\mathbb {N}_\bot := \mathbb {N}\cup \{\bot \}$ be the represented space where $\bot $ is represented by $0^\omega $ and every $n\in \mathbb {N}$ is represented by any $p\neq 0^\omega $ such that $n+1$ is the first non-zero element of p.

For every $k\ge 2$ , $\operatorname {Fin}_{k}\!\left (\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}\right )$ is equivalent to the problem F defined as follows: given $A\in \operatorname {dom}(\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k})$ and $n\in \mathbb {N}_\bot $ with the assumption that:

  • $n=\bot $ iff $|A| = k;$

  • if $n\in \mathbb {N}$ then $n\in A$ ,

find $m\in \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}(A)$ .

Proof. We shall use the characterization of $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ introduced in Proposition 7.2, i.e., we can assume that a name for $A\in \operatorname {dom}(\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k})$ is a sequence $p\in k^{\mathbb {N}}$ such that if $\lim _n p(n)$ exists then it is the unique $i<k$ not in A.

Let us first show that $F\leq _{\mathrm {W}} \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}$ . Let $\langle p, q \rangle $ be a name for $(A,n)\in \operatorname {dom}(F)$ . We define a sequence $r\in (k+1)^{\mathbb {N}}$ as follows: as long as $q(i)=0$ , let $r(i):=k$ . If $q(i)\neq 0$ for some i, then, for every $j\ge i$ , $r(j):=p(j)$ .

Clearly, r is a (name for a) valid input for $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}$ . Observe that if q is not constantly $0$ then $|A|<k$ , hence the sequence p has a limit $l<k$ . In particular, r always has a limit. We claim that, given $m\neq \lim _{i\to \infty } r(i)$ , we can uniformly compute a valid solution for $F(A,n)$ . Indeed, if $m<k$ , then $m\in A$ . This follows from the fact that if q is constantly $0$ then $|A|=k$ . Conversely, if q is not constantly $0$ then $m\neq \lim _i p(i)$ , and therefore $m\in A$ . On the other hand, if $m=k$ then q is not constantly $0$ , so, by hypothesis, if $q(i)$ is the first non-zero element of q then $q(i)-1\in A$ . This shows that $F\leq _{\mathrm {W}} \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}$ .

Assume now that $f:\subseteq \mathbb {N}^{\mathbb {N}} \rightrightarrows k$ is such that $f\leq _{\mathrm {W}} \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}$ via $\Phi ,\Psi $ . For the sake of readability, we can assume that $\Psi $ has codomain $\mathbb {N}$ . To reduce f to F, suppose we are given some $x \in \operatorname {dom}(f)$ . In parallel, we can simulate the computations $\Psi (x,i)$ for every $i<k+1$ . We can uniformly compute two sequences $p, q\in k^{\mathbb {N}}$ as follows: we wait until we either find $i<j<k+1$ such that $\Psi (x,i)=\Psi (x,j)=n$ or we see that $|\{ \Psi (x,i) {}\,:\,{} i<k+1\}|=k$ (by the pigeonhole principle, at least one of two cases must happen). If we first find $i<j<k+1$ such that ${\Psi (x,i)=\Psi (x,j)=n}$ , we define $p,q$ so that $0^s (n+1)\sqsubseteq q$ for some s and p is any convergent sequence with $\lim p \neq n$ . On the other hand, if we first see that $|\{ \Psi (x,i) {}\,:\,{} i<k+1\}|=k$ , we proceed as follows: we keep searching for $i<j$ as above. If they are never found, we let $q:=0^\omega $ and p be any non-convergent sequence. Conversely, if they are found, we define q so that $0^s (n+1)\sqsubseteq q$ for some s. Moreover, we let p be a convergent sequence with $\lim p\neq n$ such that if $\lim _t \Psi (x,\Phi (x)(t))$ exists and is different from n then $\lim p = \lim _t \Psi (x,\Phi (x)(t))$ . This can be done by computing $\Psi (x,\Phi (x)(t))$ and replacing each occurrence of n with some $m\neq n$ .

Observe that $\langle p,q \rangle $ is a valid name for an input $(A,n_\bot )$ of F. Indeed, q is constantly $0$ iff p is not convergent, i.e., iff p is a name for the input A of $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ with $|A|=k$ . Moreover, if q is not constantly $0$ then it is a name for some $n\in \mathbb {N}$ with $n\neq \lim p$ . Given $m\in F(A,n_\bot )$ , we can uniformly compute a solution for $f(x)$ as follows: we wait until we see that q is the name for some $n\neq \bot $ or that $|\{ \Psi (x,i) {}\,:\,{} i<k+1\}|=k$ . If the former is observed first, then $n\in f(x)$ , as there are $i<j$ with $\Psi (x,i)=\Psi (x,j)=n$ , and either i or j is a valid solution for $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}(\Phi (x))$ . Conversely, if we first see that $|\{ \Psi (x,i) {}\,:\,{} i<k+1\}|=k$ , then any $m\in F(A,n_\bot )$ is a valid solution for $f(x)$ . Indeed, if $|f(x)|<k$ then the (unique) wrong solution is $\lim \Psi (x,\Phi (x))$ (as $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+1}(\Phi (x))$ has at most one wrong solution). By construction, $m\neq \lim p = \lim \Psi (x,\Phi (x))$ , and therefore $m\in f(x)$ . Observe that, in this case, q is not the constantly $0$ sequence (if q is constantly $0$ then every $i<k+1$ is valid solution for $f(x)$ ).

Corollary 7.10. $\operatorname {Fin}_{2}\!\left (\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{3}\right )\equiv _{\mathrm {W}} \mathsf {C}_2$ .

Proof. As $\mathsf {ACC}_2 \equiv _{\mathrm {W}} \mathsf {C}_2$ , one direction follows from Proposition 7.8. That $\operatorname {Fin}_{2}\!\left (\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{3}\right ) \leq _{\mathrm {W}} \mathsf {C}_2$ follows from Proposition 7.9 by observing $\mathsf {C}_2$ can solve the task “Given $n \in \mathbb {N}_\bot $ , compute some $b \in \{0,1\}$ such that if $n \in \{0,1\}$ , then $b = n$ ”.

8 Finding antichains in trees

We turn our attention to trees as a special case of partial orders, where we identify a tree with its prefix relation. We observe that a tree is a wqo iff it has finite width; in particular, being wqo is merely a $\Sigma ^0_2$ -property for trees rather than $\Pi ^1_1$ as for general quasi-orders. While trees can only avoid being wqo by having an infinite antichain, being a bad sequence is a weaker notion than being an antichain. However, a given infinite bad sequence in a tree can be computably thinned out to yield an infinite antichain. We let $\mathsf {BS}|_{\operatorname {Tree}}$ denote the restriction of $\mathsf {BS}$ to subtrees of $\mathbb {N}^{<\mathbb {N}}$ , and think of it as the problem that receives as input a tree of infinite width and returns an infinite antichain in it. How $\mathsf {BS}|_{\operatorname {Tree}}$ and its first-order part relate to other principles, in particular those discussed in the previous section, is depicted in Figure 2.

Figure 2 The figure shows all reductions between the depicted principles up to transitivity, with the potential exception that we have not ruled out being $\mathsf {BS}|_{\operatorname {Tree}} \leq _{\mathrm {W}} \widehat {\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}}$ . The diagram can be thought of as a cube whose vertical edges have decreasing chains parameterized by $k \geq 2$ on them, and the bottom corners are labelled by degrees which are below the ones on the edges for all k.

Even though the instances of $\mathsf {BS}|_{\operatorname {Tree}}$ are arbitrary subtrees of $\mathbb {N}^{<\mathbb {N}}$ of infinite width, we show below that its strength comes already from the pruned subtrees of $2^{<\mathbb {N}}$ (a tree is pruned if it has no leaves). Compare this with the problem $\mathsf {C}_{\mathbb {N}^{\mathbb {N}}}$ of finding an infinite path through a given ill-founded subtree of $\mathbb {N}^{<\mathbb {N}}$ : Its restriction to ill-founded pruned trees is of course computable, and its restriction $\mathsf {C}_{2^{\mathbb {N}}}$ to subtrees of $2^{<\mathbb {N}}$ is known to be much weaker (see [Reference Brattka, Gherardi, Pauly, Brattka and Hertling3, Figure 11.2]). This is unsurprising given the difference in the complexity between “T is wqo” and “T is well-founded”.

Proposition 8.1. $\mathsf {BS}|_{\operatorname {Tree}}$ is Weihrauch equivalent to its restriction to pruned subtrees of $2^{<\mathbb {N}}$ .

Proof. Given a tree $T \subseteq \mathbb {N}^{<\mathbb {N}}$ of infinite width, we shall uniformly compute a pruned tree $S \subseteq 2^{<\mathbb {N}}$ , also of infinite width, together with a surjective labelling function , such that if $\ell (v) \sqsubseteq \ell (w)$ then $v \sqsubseteq w$ . The image of any infinite S-antichain under $\ell $ would then be an infinite T-antichain.

Our construction proceeds in phases, where at the end of phase n, we commit to $S \cap 2^{<n}$ and the corresponding restriction of $\ell $ . Throughout the construction we will respect the constraint that $\ell (v) \sqsubseteq \ell (w)$ implies $v \sqsubseteq w$ . At the end of phase n, all leaves in S will have height at least n, so S will be ultimately be a pruned tree.

During phase $0$ , we simply add the root to S and label it with the root of T. Fix an ordering of the vertices of T (with order-type $\omega $ ) which agrees with $\sqsubseteq $ . In phase n, start by considering the first vertex $u \in T$ that is not yet enumerated in $\operatorname {ran}(\ell )$ . Our constraint on $\ell $ implies that the set $\{v' \in S: \ell (v') \sqsubseteq u\}$ forms a path in S, say with final vertex $v'$ . (Observe that $\ell (v')$ is the parent of u.) Let v be the first vertex of height at least $n-1$ such that $v' \sqsubseteq v$ and either or . Such v exists because all leaves in S have height at least $n-1$ . We then add a child to v (i.e., put in S if , else put in S) and label the child with u.

If we now have both and in S, we conclude phase n by adding a child to each vertex in S which was in S at the beginning of phase n and is currently a leaf in S. Each such child is labelled the same way as its parent in S. This ensures every leaf in S has height at least n. Notice that this action maintains our constraint on $\ell $ .

On the other hand, if only one of and lie in S, we consider the next $u \in T$ that is not in $\operatorname {ran}(\ell )$ and repeat the above steps. To see why this process terminates, observe that if it did not, then there must be some $u \in T$ first considered in phase n which has multiple children. Suppose we acted for u by adding to S and defining . Let $u'$ denote the first child of u to be considered in phase n. Then would be the first vertex of height at least $n-1$ which extends every $v'$ with $\ell (v') \sqsubseteq u'$ , because the highest vertex $v'$ with $\ell (v') \sqsubseteq u'$ is precisely , which has height at least n. Therefore we would add to S and define . Similarly, if $u"$ denotes the second child of u to be considered in phase n, then we would add to S and define . We would then conclude phase n, contrary to assumption.

This completes the description of our construction. We have ensured that S contains infinitely many vertices v with , so S has infinite width. The other requirements are directly enforced by the construction.

Next, we shall characterize the finitary parts and deterministic part of $\mathsf {BS}|_{\operatorname {Tree}}$ . If $(T,\sqsubseteq )$ is a tree and v is a vertex in T, define $[v]_T = \{w \in T: v \sqsubseteq w\}$ . We omit the subscript T if the tree in question is clear from context.

Proposition 8.2. $\mathsf {BS}|_{\operatorname {Tree}} \leq _{\mathrm {W}} \widehat {\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}}$ for every $k \geq 2$ .

Proof. Given a tree $T \subseteq \mathbb {N}^{<\mathbb {N}}$ with infinite width, we consider all antichains $(v_0,v_1,\ldots ,v_{k-1})$ of size k in it. For each such antichain, at most one vertex $v_i$ must be avoided when building an infinite antichain, which happens if $T\setminus [v_i]$ has finite width. We can use $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ to select a suitable choice amongst the $v_i$ . We can then greedily build an infinite antichain from the selected vertices.

Corollary 8.3. $\operatorname {Fin}_{k}\!\left (\mathsf {BS}|_{\operatorname {Tree}}\right ) \equiv _{\mathrm {W}} \mathrm {id}$ for every $k\ge 1$ . Therefore, $\operatorname {Det}\!\left (\mathsf {BS}|_{\operatorname {Tree}}\right ) \equiv _{\mathrm {W}} \mathrm {id}$ .

Proof. For the first part of the statement, fix $k\ge 1$ . By Proposition 8.2 and Corollary 7.6,

$$\begin{align*}\operatorname{Fin}_{k}\!\left(\mathsf{BS}|_{\operatorname{Tree}}\right)\leq_{\mathrm{W}} \operatorname{Fin}_{k}\!\left(\widehat{\boldsymbol{\Pi}^0_2\mathsf{-ACC}_{k+2}}\right) \equiv_{\mathrm{W}} \mathrm{id}. \end{align*}$$

Similarly, the second part of the statement then follows from $\operatorname {Det}\!\left (\widehat {\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k+2}}\right ) \equiv _{\mathrm {W}} \mathrm {id}$ (Corollary 7.7).

We now turn our attention to the first-order part of $\mathsf {BS}|_{\operatorname {Tree}}$ . The following first-order problem clearly reduces to $\mathsf {BS}|_{\operatorname {Tree}}$ .

Definition 8.4. Let $\operatorname {ExtVer}$ denote the problem taking a tree T of infinite width, and returning a vertex $v \in T$ such that $T \setminus [v]$ still has infinite width, i.e., a vertex that is extendible to an infinite antichain.

The construction employed to prove Proposition 8.1 also shows that for $\operatorname {ExtVer}$ it is immaterial whether we consider countably-branching or binary trees, and whether we assume the trees to be pruned or not.

Proposition 8.5. $\operatorname {ExtVer} \equiv _{\mathrm {W}} {}^1\mathsf {BS}|_{\operatorname {Tree}}$ .

Proof. We only need to show that ${}^1\mathsf {BS}|_{\operatorname {Tree}} \leq _{\mathrm {W}} \operatorname {ExtVer}$ . An instance of ${}^1\mathsf {BS}|_{\operatorname {Tree}}$ is a tree T of infinite width together with a notion of “sufficiently large finite antichain” (more formally, the latter is a computable functional which, when given T and an infinite T-antichain, must halt on some initial segment thereof). A solution is a sufficiently large finite antichain which is extendible to an infinite one. We shall compute a tree $T' \subseteq \mathbb {N}^{<\mathbb {N}}$ of infinite width such that each vertex $v \in T'$ (other than the root) is labelled with a sufficiently large finite antichain $A_v$ in T, and if v extends to an infinite $T'$ -antichain, then $A_v$ extends to an infinite T-antichain. For clarity, we shall use Greek letters ( $\sigma $ , $\tau $ ) to denote the vertices of T, and Roman letters (v, w) to denote the vertices of $T'$ .

The tree $T'$ is defined recursively as follows. Start by putting the root in $T'$ . The children of the root will each be labelled with a sufficiently large finite antichain $A_i$ in T. Specifically, the i-th child of the root is defined recursively as follows: First search for $A_0$ and $A_1$ in T, disjoint and each sufficiently large, such that $A_0 \cup A_1$ forms an antichain. Since T has infinite width, this search is successful and we label the first child with $A_0$ and the second child with $A_1$ . Then search for $A_2$ sufficiently large and disjoint from $A_0 \cup A_1$ such that $A_0 \cup A_1 \cup A_2$ forms an antichain. If found, we label the third child with $A_2$ . Proceed to search for $A_3$ , and so on. Note that $(A_i)_{{i}}$ may not be infinite. The reason why we start by searching for two antichains $A_0$ and $A_1$ , and then continue by searching one extra antichain at a time (first $A_2$ , then $A_3$ , and so on) is to guarantee that if a node has a child, then it has at least two incomparable children. More generally, if $v \in T'$ is not the root, then we define the children of v by searching as above among the sufficiently large antichains B such that $B \trianglerighteq ^T A_v$ (i.e., there is some $\sigma \in A_v$ such that $\sigma \sqsubseteq \tau $ for all $\tau \in B$ ). Note that for certain v, this search may turn up empty. This completes the construction of $T'$ , excepting a final modification to ensure that $T'$ is computable (uniformly in T): Each child of a vertex in $T'$ shall encode the stage at which its corresponding T-antichain was found.

We claim that $T'$ has infinite width. If any vertex has infinitely many children, we are done. Otherwise, we shall prove by induction that at each level of $T'$ (beyond the root), there is some vertex v and some $\sigma \in A_v$ such that $[\sigma ]_T$ has infinite width. This would imply that v has at least two children, allowing us to conclude that $T'$ has infinite width. The inductive step proceeds as follows: Suppose $v \in T'$ and $\sigma \in A_v$ are such that $[\sigma ]_T$ has infinite width. Say the children of v are labelled by the antichains $A_0,A_1,\dots ,A_k$ . Then $A_0 \cup \dots \cup A_k$ is not extendible to an infinite antichain in $\bigcup _{\rho \in A_v} [\rho ]_T$ (else v would have additional children), but the latter contains $[\sigma ]_T$ and thus has infinite width. So there is some i and some $\rho \in A_i$ such that $[\rho ]_T$ has infinite width. The base case proceeds similarly; simply consider the children of the root in $T'$ and use the assumption that T has infinite width.

Next, we claim that if $v_0,v_1 \in T'$ are incomparable, then any $\sigma _0 \in A_{v_0}$ and any $\sigma _1 \in A_{v_1}$ are T-incomparable. This would imply $A_{v_0}$ and $A_{v_1}$ are disjoint and $A_{v_0} \cup A_{v_1}$ is a T-antichain. To prove the claim, let $w \in T'$ denote the longest common ancestor of $v_0$ and $v_1$ . Let $w_0 \sqsubseteq v_0$ and $w_1 \sqsubseteq v_1$ be children of w. By construction, $A_{w_0}$ and $A_{w_1}$ are disjoint and $A_{w_0} \cup A_{w_1}$ is a T-antichain. Also by construction, for $i =0,1$ , there is some $\tau _i \in A_{w_i}$ such that $\tau _i \sqsubseteq \sigma _i$ . Since $\tau _0$ and $\tau _1$ are T-incomparable, so are $\sigma _0$ and $\sigma _1$ .

The above claim implies that if $v \in T'$ is extendible to an infinite $T'$ -antichain $(v_i)_{{i}}$ , then $A_v$ is extendible to an infinite T-antichain, namely $\bigcup _i A_{v_i}$ . Since every $A_v$ is sufficiently large, this yields a Weihrauch reduction from ${}^1\mathsf {BS}|_{\operatorname {Tree}}$ to $\operatorname {ExtVer}$ .

Similar to Proposition 8.2, we shall obtain an upper bound for ${}^1\mathsf {BS}|_{\operatorname {Tree}}$ . Let denote the following problem: Given a sequence $(p_k)_{{k \in \mathbb {N}}}$ where each $p_k$ is an $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ -instance, find $(k,i) \in \mathbb {N}^2$ such that $i \in \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}(p_k)$ . Observe that for every $\ell $ . The operation is a generalization of the binary meet to countably many problems. However, it is known that the Weihrauch degrees are not a $\aleph _0$ -complete meet semilattice ([Reference Higuchi and Pauly15], see also [Reference Lempp, Marcone and Valenti21]).

Proposition 8.6. .

Proof. First, we prove the reduction. Given a tree T of infinite width, search for antichains of each finite size. Within each antichain, at most one vertex is not extendible to an infinite antichain in T. Since extendibility is a $\Pi ^0_2$ property (relative to T), for each k we can build an input for $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ to identify one vertex in the antichain of size k which is extendible. Then a solution for any of the $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ -instances yields an $\operatorname {ExtVer}$ -solution to T.

To show that the reduction is strict, assume towards a contradiction that . Apply the reduction to the “neutral” input to , i.e., where every $j < k$ is a solution. If the resulting tree T contains incomparable vertices $v_0$ and $v_1$ which yield the same output $j < k$ under the reduction, we may fix a finite prefix which determines this fact, then render j wrong in the $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}$ -instance by enumerating it. This leads to a contradiction because either $v_0$ or $v_1$ must be a valid $\operatorname {ExtVer}$ -solution. On the other hand, if incomparable vertices in T always yield different outputs under the reduction, by the pigeonhole principle, there must be incomparable vertices $v_0$ and $v_1$ which yield outputs for different k. Say $v_i$ determines the output $j_i < k_i$ . Once again this fact is determined by a finite prefix of the neutral input. We then render $j_i$ wrong in the $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k_i}$ -instance for $i=0,1$ after said finite prefix. Since either $v_0$ or $v_1$ must be an $\operatorname {ExtVer}$ -solution, we reach a contradiction.

It turns out that $\operatorname {ExtVer}$ is sandwiched between $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ and .

Proposition 8.7. $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}} \leq _{\mathrm {W}} \operatorname {ExtVer}$ .

Proof. Given some $p \in \mathbb {N}^{\mathbb {N}}$ , we shall build some $T \subseteq 2^{<\mathbb {N}}$ of infinite width such that from any vertex in T which is extendible to an infinite antichain, we can compute some n such that p does not end in $n^\omega $ (i.e., if $\lim _i p_i$ exists, it is not n). We can assume without loss of generality that $p_i \leq i$ . The tree is constructed level by level. After each level is constructed, each newly added vertex is labelled by the least natural number which has yet to be used as a label, in order from left to right. When building level s, first consider the leftmost leaf which extends the vertex with label $p_s$ . Such a vertex must exist because T already has more than $s \geq p_s$ vertices. Add both children of said leaf to T. Then, for every other leaf in T (excluding the children just added), add its left child to T.

As we have a bifurcation on every level and no dead-ends, T has infinite width. Moreover, if p ends in $n^\omega $ , then the subtree of T not extending the vertex with label n has finite width, i.e., the vertex with label n is not extendible.

The following proposition is stated for $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ but can be easily generalized to any pointclass $\boldsymbol {\Gamma }$ .

Proposition 8.8. $\operatorname {ExtVer} \nleq _{\mathrm {W}} \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ .

Proof. Since $\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}} \nleq _{\mathrm {W}} \mathsf {lim}$ (Proposition 7.3), it suffices (by Proposition 8.7) to prove that if $\operatorname {ExtVer} \leq _{\mathrm {W}} \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ , then $\operatorname {ExtVer} \leq _{\mathrm {W}} \mathsf {lim}$ . We will in fact reduce $\operatorname {ExtVer}$ to $\mathsf {LPO}$ .

Observe that the forward functional of a putative reduction $\operatorname {ExtVer} \leq _{\mathrm {W}} \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}$ constitutes a computable procedure for producing a sequence of vertices $(v_i)_{i \in \mathbb {N}}$ in a given tree of infinite width such that there is at most one vertex $v_i$ which is not extendible (to an infinite antichain). To reduce $\operatorname {ExtVer}$ to $\mathsf {LPO}$ , ask $\mathsf {LPO}$ if there are $j, k$ with $v_j \sqsubset v_k$ . If so, then $v_k$ must be extendible because (1) at most one $v_i$ is not extendible and (2) if $v_j$ is extendible, so is $v_k$ . On the other hand, if $(v_i)_{i \in \mathbb {N}}$ is already an infinite antichain, then $v_0$ is extendible.

To summarize:

Corollary 8.9. .

Corollary 8.10. $\mathsf {ACC}_{\mathbb {N}} \times \mathsf {ACC}_{\mathbb {N}}\not \leq _{\mathrm {W}} \mathsf {BS}|_{\operatorname {Tree}}$ . In particular, ${}^1\mathsf {BS}|_{\operatorname {Tree}} <_{\mathrm {W}} {}^1\mathsf {BS}|_{\operatorname {Tree}} \times {}^1\mathsf {BS}|_{\operatorname {Tree}}$ and $\mathsf {BS}|_{\operatorname {Tree}} <_{\mathrm {W}} \mathsf {BS}|_{\operatorname {Tree}} \times \mathsf {BS}|_{\operatorname {Tree}}$ .

Proof. By Proposition 8.6, $\mathsf {ACC}_{\mathbb {N}} \times \mathsf {ACC}_{\mathbb {N}}\leq _{\mathrm {W}} \mathsf {BS}|_{\operatorname {Tree}}$ implies $\mathsf {ACC}_{\mathbb {N}} \times \mathsf {ACC}_{\mathbb {N}} \leq _{\mathrm {W}} \boldsymbol {\Pi }^0_2\mathsf {-ACC}_{2} \equiv _{\mathrm {W}} \mathsf {RT}^{{1}}_{{2}}$ , which is easily seen to be false. The rest of the statement follows easily from the fact that $\mathsf {BS}|_{\operatorname {Tree}} \times \mathsf {BS}|_{\operatorname {Tree}}\leq _{\mathrm {W}} \mathsf {BS}|_{\operatorname {Tree}}$ implies ${}^1\mathsf {BS}|_{\operatorname {Tree}} \times {}^1\mathsf {BS}|_{\operatorname {Tree}}\leq _{\mathrm {W}} {}^1\mathsf {BS}|_{\operatorname {Tree}}$ (see, e.g., [Reference Soldà and Valenti28, Proposition 4.1]), which, in turn, implies $\mathsf {ACC}_{\mathbb {N}} \times \mathsf {ACC}_{\mathbb {N}} \leq _{\mathrm {W}} \mathsf {BS}|_{\operatorname {Tree}}$ .

We end this section by drawing attention to the following open question.

Open Question 8.11. Does $\mathsf {BS}|_{\operatorname {Tree}} \leq _{\mathrm {W}} \mathsf {DS}$ ?

All lower bounds for $\mathsf {BS}|_{\operatorname {Tree}}$ known to us are also below $\mathsf {DS}$ ; this in particular applies to the finitary, deterministic, and first-order parts of $\mathsf {BS}|_{\operatorname {Tree}}$ . On the other hand, the best upper bound for $\mathsf {BS}|_{\operatorname {Tree}}$ we have is $\widehat {\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{k}}$ (Proposition 8.2), which in turn does not reduce to $\mathsf {DS}$ (or even $\mathsf {BS}$ ) by Corollary 4.4. One natural attempt to reduce $\mathsf {BS}|_{\operatorname {Tree}}$ to $\mathsf {DS}$ is blocked by the following observation.

Proposition 8.12. There is no computable procedure that receives as input any tree $T \subseteq 2^{<\mathbb {N}}$ and returns a linear extension $(T,\prec )$ of the prefix relation on T such that whenever T has infinite width, then $\prec $ is ill-founded.

Proof. For the sake of a contradiction, assume there exists such a procedure. We shall define a pruned binary tree T of infinite width on which the procedure produces a well-ordering. At each stage, we define one new level of T as follows. Start by putting the root, $0$ , and $1$ into T. Wait for the procedure to decide whether $0 \prec 1$ or $1 \prec 0$ . While waiting, we extend $0$ and $1$ by $0$ s. Eventually, say at stage $s_0$ , the procedure decides $w_0 \in \{0,1\}$ is the $\prec $ -max of $0$ and $1$ . We then add a split above $w_0$ , i.e., put and into T. In general, if we added a split into T at stage $s_n$ , let $A_n$ denote the T-antichain of vertices which are currently T-maximal. We wait for the procedure to decide $\max _\prec A_n$ . While waiting, we extend each T-maximal element by $0$ s. Once $w_n$ has been decided as $\max _\prec A_n$ , we add a split above where k is largest such that . We end the stage by extending all other T-maximal elements by $0$ .

The construction adds infinitely many splits, so T has infinite width. To show that $(T,\prec )$ is a well-ordering, first observe that $(w_n)_n$ forms an increasing cofinal sequence in $(T,\prec )$ . It therefore suffices to show that each $w_n$ lies in the $\prec $ -well-founded part W. Suppose towards a contradiction that there is some infinite $\prec $ -descending sequence S which begins with $w_n$ . There must be some $a \in A_n$ , different from $w_n$ , such that infinitely many elements of S are T-above a. This is because if $x\prec b$ for every $b\in A_n$ , then x is a prefix of some of them. Then the subtree of T above a must contain splits; furthermore, some $s \in S$ lies T-above such a split (otherwise the subsequence of S which is T-above a forms a T-chain, which must have $\prec $ -order type $\omega $ ). However, by construction of T, such a split occurs above some $w_m$ where $m> n$ , so we have $s \succ w_m \succ w_n$ . This contradicts the fact that S begins with $w_n$ .

9 The quotient relative to the parallel product

Theorem 4.3 implies that

$$\begin{align*}\max\nolimits _{\leq_{\mathrm{W}}} \{h {}\,:\,{} h\times \widehat{\mathsf{ACC}_{\mathbb{N}}} \leq_{\mathrm{W}} \mathsf{BS}\} \end{align*}$$

exists and is equal to (the Weihrauch degree of) $\mathsf {lim}$ . The existence of such a maximum (with different problems in place of $\widehat {\mathsf {ACC}_{\mathbb {N}}}$ and $\mathsf {BS}$ ) was discussed in [Reference Dzhafarov, Goh, Hirschfeldt, Patey and Pauly12, Theorem 3.7], which prompted them to ask about the extent to which the “parallel quotient” operator

$$\begin{align*}f/g \equiv_{\mathrm{W}} \max\nolimits _{\leq_{\mathrm{W}}} \{h {}\,:\,{} h \times g \leq_{\mathrm{W}} f\} \end{align*}$$

is defined. While [Reference Dzhafarov, Goh, Hirschfeldt, Patey and Pauly12, Remark 3.11] says “We have no reason to think this operator is total [ $\ldots $ ]”, we address their question by showing that the operator in question is in fact total.

Definition 9.1. Given $f,g:\subseteq \mathbb {N}^{\mathbb {N}} \rightrightarrows \mathbb {N}^{\mathbb {N}}$ with g different from $0$ , we define their parallel quotient $f/g:\subseteq \mathbb {N}\times \mathbb {N}\times \mathbb {N}^{\mathbb {N}} \rightrightarrows \mathbb {N}^{\mathbb {N}}$ as follows:

$$ \begin{align*} \operatorname{dom}(f/g) := \{(e,i,p){}\,:\,{} (\forall q \in \operatorname{dom}(g))[ &\Phi_e(p,q) \in \operatorname{dom}(f) \text{ and } \\ &(\forall r \in f(\Phi_e(p,q)))(\Phi_i(p,q,r) \in g(q))]\} \end{align*} $$
$$\begin{align*}f/g(e,i,p) := \{\langle q,r \rangle{}\,:\,{} q \in \operatorname{dom}(g) \ \land \ r \in f(\Phi_e(p,q))\}.\end{align*}$$

Proposition 9.2. $f/g \equiv _{\mathrm {W}} \max _{\leq _{\mathrm {W}}} \{h{}\,:\,{} h \times g \leq _{\mathrm {W}} f\}$ .

Proof. Given a g-instance q and an $f/g$ -instance $(e,i,p)$ , consider the f-instance $\Phi _e(p,q)$ . Given an $r \in f(\Phi _e(p,q))$ , we may compute (using q and $e,i,p$ ) $\Phi _i(p,q,r)$ , which is a g-solution of q, and $\langle q,r \rangle $ , which is an $f/g$ -solution of $(e,i,p)$ . This proves $f/g \times g \leq _{\mathrm {W}} f$ .

Suppose h is a problem such that $h \times g \leq _{\mathrm {W}} f$ via $\Phi _e$ , $\Psi $ . We want to show that $h \leq _{\mathrm {W}} f/g$ . Let p be an h-instance. Observe that, in order to compute a solution for $h(p)$ , we only need a g-input q and any solution $r\in f(\Phi _e(p,q))$ .

Let i be an index for the functional which takes in input $(p,q,r)$ and produces the projection of $\Psi (p, q, r)$ to the second coordinate. Given the h-instance p, we uniformly compute the $f/g$ -instance $(e,i,p)$ . For every $\langle q,r \rangle \in f/g(e,i,p)$ , the projection of $\Psi (p,q,r)$ to the first coordinate is an h-solution to p.

It follows that $f/g$ is well-defined on all (non-zero) Weihrauch degrees, so we may extend Definition 9.1 to all (non-zero) problems. The following proposition is a straightforward consequence of the definitions.

Proposition 9.3.

  1. (1) If g is pointed, then $f/g \leq _{\mathrm {W}} f$ .

  2. (2) If $f_0 \leq _{\mathrm {W}} f_1$ and $g_0 \leq _{\mathrm {W}} g_1$ , then $f_0/g_1 \leq _{\mathrm {W}} f_1/g_0$ .

The parallel quotient is useful to describe to what extent a problem f is stable under parallel product. If f is pointed and closed under product, then $f/f \equiv _{\mathrm {W}} f$ . At the same time, if f is not closed under product, the quotient $f/f$ gives a quantitative estimate of “the lack of closure of f under parallel product”.

The parallel quotient is an example of a residual operator. In general, for a fixed operator $\cdot $ , it is interesting to investigate the existence of the following degrees: $\max \{h {}\,:\,{} f \cdot h \leq _{\mathrm {W}} g\}$ , $\max \{h {}\,:\,{} h \cdot f \leq _{\mathrm {W}} g\}$ , $\min \{h {}\,:\,{} g \leq _{\mathrm {W}} f \cdot h \}$ , and $\min \{h {}\,:\,{} g \leq _{\mathrm {W}} h \cdot f\}$ . A residual operator for $\cdot $ is one that witnesses the existence of one (or more) of these maxima/minima. The existence of residual operators for the join $\sqcup $ and the meet $\sqcap $ in the Weihrauch degrees has been studied in [Reference Higuchi and Pauly15]. Besides, the existence of $\min \{h {}\,:\,{} g \leq _{\mathrm {W}} f * h\}$ has been proven in [Reference Brattka and Pauly4] (resulting in the implication operator $\to $ ). The parallel quotient is a residual operator for $\times $ . A more thorough study of the existence of residual operators for other choices of $\cdot $ is currently under investigation. Further properties of the parallel quotient will be discussed in an upcoming paper.

As already observed, Theorem 4.3 can be rewritten as

$$\begin{align*}\mathsf{DS}/\widehat{\mathsf{ACC}_{\mathbb{N}}}\equiv_{\mathrm{W}} \mathsf{BS}/\widehat{\mathsf{ACC}_{\mathbb{N}}} \equiv_{\mathrm{W}} \mathsf{lim}. \end{align*}$$

We now discuss the parallel quotient of $\mathsf {DS}$ and $\mathsf {BS}$ over other well-known problems.

Proposition 9.4. $\mathsf {DS} {\kern-1pt}\equiv _{\mathrm {W}}{\kern-1pt} \mathsf {C}_{\mathbb {N}} {\kern-1pt}\times{\kern-1pt} \mathsf {DS} {\kern-1pt}\equiv _{\mathrm {W}}{\kern-1pt} \mathsf {DS} * \mathsf {C}_{\mathbb {N}}$ and ${\mathsf {BS} \equiv _{\mathrm {W}} \mathsf {C}_{\mathbb {N}} \times \mathsf {BS} \equiv _{\mathrm {W}} \mathsf {BS}{\kern-1pt} * {\kern-1pt}\mathsf {C}_{\mathbb {N}}}$ .

Proof. The reductions $\mathsf {DS} \leq _{\mathrm {W}} \mathsf {C}_{\mathbb {N}} \times \mathsf {DS} \leq _{\mathrm {W}} \mathsf {DS} * \mathsf {C}_{\mathbb {N}}$ are trivial. To show that $\mathsf {DS}* \mathsf {C}_{\mathbb {N}} \leq _{\mathrm {W}} \mathsf {DS}$ , let $(w,p)\in \operatorname {dom}(\mathsf {DS}* \mathsf {C}_{\mathbb {N}})$ . Without loss of generality, we may view the (possibly finite) string produced by $\Phi _w(n)$ as a linear order $L_n$ . Clearly $L_n$ need not be ill-founded if $n\notin \mathsf {C}_{\mathbb {N}}(p)$ .

We shall uniformly compute a linear order L as follows: the elements of L will be of the form $(n,x,s)$ , where

$$\begin{align*}(n,x,s) \le_L (m,y,s') :\hspace{-1mm}\iff n<m \text{ or } (n=m \text{ and } x\le_{L_n} y).\end{align*}$$

The third component plays no role in the order and is only used to guarantee that the resulting linear order is computable (and not just c.e.). The construction below will guarantee that no two triples share the same third component.

Starting from $n = 0$ , as long as n is not enumerated outside of $\mathsf {C}_{\mathbb {N}}(p)$ , we build an isomorphic copy of $L_n$ at the top of L. More precisely we add $(n,x,s)$ to L, where s is the current step of the construction and x is the $\mathbb {N}$ -least element of $L_n$ which has not been added.

If we ever see $n \notin \mathsf {C}_{\mathbb {N}}(p)$ , we continue the above with $n+1$ instead of n.

This completes the construction of L. Observe that if $n = \min \mathsf {C}_{\mathbb {N}}(p)$ then, modulo a finite initial segment, L is isomorphic to $L_n$ . In particular, L is ill-founded and, given any infinite descending sequence through L, we can uniformly compute n and a descending sequence through $L_n$ .

The very same argument works for $\mathsf {BS}$ in place of $\mathsf {DS}$ : just replace “linear order”, “ill-founded”, and “descending sequence” with “quasi-order”, “a non-well quasi order”, and “bad sequence”, respectively.

Lemma 9.5. Let g be computably true (i.e., every g-instance has a g-computable solution). Suppose $g\leq _{\mathrm {W}} f\left. \hspace {-1mm} \right |{}_{X}$ , where $f\left. \hspace {-1mm} \right |{}_{X}$ is such that, for every $x\in X$ and every $y\in f(x)$ , $y\not \le _{\mathrm {T}} x$ . Then $f/g\not \leq _{\mathrm {W}} g$ .

Proof. Let $(e,i,p)\in \operatorname {dom}(f/g)$ be such that p is computable and the maps $q\mapsto \Phi _e(p, q)$ and $(q,r)\mapsto \Phi _i(p,q,r)$ witness the reduction $g\leq _{\mathrm {W}} f\left. \hspace {-1mm} \right |{}_{X}$ . Observe that, by definition of $f/g$ and $(e,i,p)$ , every solution of $f/g(e,i,p)$ computes a solution for $f(\Phi _e(p,q))$ for some input $q\in \operatorname {dom}(g)$ . In particular, since $\Phi _e(p,q)\in X$ , $f/g(e,i,p)$ does not have any computable solution.

If $f/g \leq _{\mathrm {W}} g$ as witnessed by $\Phi ,\Psi $ , then for every $t\in g\Phi (e,i,p)$ , $\Psi ( (e,i,p), t) \in f/g(e,i,p)$ . This is a contradiction as g is computably true and therefore $g\Phi (e,i,p)$ always contains a computable point.

Proposition 9.6. For $f\in \{\mathsf {DS},\mathsf {BS}\}$ , $\boldsymbol {\Pi }^1_1\mathsf {-Bound}<_{\mathrm {W}} f / \boldsymbol {\Pi }^1_1\mathsf {-Bound}\leq _{\mathrm {W}} f / \mathsf {RT}^{{1}}_{{2}} <_{\mathrm {W}} f$ .

Proof. The reduction $\boldsymbol {\Pi }^1_1\mathsf {-Bound}\leq _{\mathrm {W}} \mathsf {DS} / \boldsymbol {\Pi }^1_1\mathsf {-Bound}$ follows from the fact that $\boldsymbol {\Pi }^1_1\mathsf {-Bound}$ is closed under product (see, e.g., [Reference Soldà and Valenti28, Theorem 7.16]) and $\boldsymbol {\Pi }^1_1\mathsf {-Bound} \leq _{\mathrm {W}} \mathsf {DS}$ . To show that this reduction is strict, we shall use Lemma 9.5. Let X be the set of ill-founded linear orders L with no L-computable descending sequence. One can modify the reduction [Reference Goh, Pauly and Valenti18, Proposition 4.8] from $\boldsymbol {\Pi }^1_1\mathsf {-Bound}$ to $\mathsf {DS}$ to show $\boldsymbol {\Pi }^1_1\mathsf {-Bound} \leq _{\mathrm {W}} \mathsf {DS}|_X$ : Given a $\boldsymbol {\Pi }^1_1\mathsf {-Bound}$ -instance (thought of as a sequence of trees $(T_n)_n$ ), define a $(T_n)_n$ -computable ill-founded tree S with no $(T_n)_n$ -computable path (this can be done in a manner which is uniformly computable from $(T_n)_n$ ). Then define a linear order $L = \bigcup _n \{n\} \times \mathrm {KB}(T_n \times S)$ . Note that if $T_n$ is well-founded, then so is $T_n \times S$ , while if $T_n$ is ill-founded, then $T_n \times S$ is ill-founded with no $(T_n)_n$ -computable path. So L is ill-founded with no L-computable descending sequence. The backward functional works as in [Reference Goh, Pauly and Valenti18, Proposition 4.8]. It follows from Lemma 9.5 that $\mathsf {DS}/\boldsymbol {\Pi }^1_1\mathsf {-Bound} \nleq _{\mathrm {W}} \boldsymbol {\Pi }^1_1\mathsf {-Bound}$ .

The reduction $\mathsf {DS} / \boldsymbol {\Pi }^1_1\mathsf {-Bound}\leq _{\mathrm {W}} \mathsf {DS} / \mathsf {RT}^{{1}}_{{2}}$ follows by the antimonotonicity of the quotient operator. The reduction $ \mathsf {DS} / \mathsf {RT}^{{1}}_{{2}} \leq _{\mathrm {W}} \mathsf {DS}$ is immediate as $\mathsf {RT}^{{1}}_{{2}}$ is pointed. The fact that it is strict follows from the fact that if $\mathsf {DS} \leq _{\mathrm {W}} \mathsf {DS} /\mathsf {RT}^{{1}}_{{2}}$ then, by definition, $\mathsf {RT}^{{1}}_{{2}}\times \mathsf {DS} \leq _{\mathrm {W}} \mathsf {DS}$ . This is impossible as, by Theorem 4.3, $\mathsf {RT}^{{1}}_{{2}}\times \widehat {\mathsf {ACC}_{\mathbb {N}}}\not \leq _{\mathrm {W}} \mathsf {DS}$ .

Replacing $\mathsf {DS}$ with $\mathsf {BS}$ in the above argument shows that the statement holds for $\mathsf {BS}$ as well.

Notice that, in the proof of the previous result, we used the fact that both $\mathsf {DS}/\boldsymbol {\Pi }^1_1\mathsf {-Bound}$ and $\mathsf {BS}/\boldsymbol {\Pi }^1_1\mathsf {-Bound}$ (and hence $\mathsf {DS}/\mathsf {RT}^{{1}}_{{2}}$ and $\mathsf {BS}/\mathsf {RT}^{{1}}_{{2}}$ ) have computable instances with no computable solutions. In particular, this implies that none of them is a first-order problem. The following result provides a lower bound for $\mathsf {DS}/\mathsf {RT}^{{1}}_{{2}}$ and $\mathsf {BS}/\mathsf {RT}^{{1}}_{{2}}$ .

Let us denote with $\mathsf {DS}^{\mathsf {FE}}$ (resp., $\mathsf {BS}^{\mathsf {FE}}$ ) the problem “given an ill-founded linear order (resp., non-well quasi-order) X, compute a sequence that, cofinitely, is an X-descending sequence (resp., X-bad sequence)”. We mention that $\mathsf {DS}^{\mathsf {FE}}$ and $\mathsf {BS}^{\mathsf {FE}}$ can be characterized as $\mathsf {C}_{\mathbb {N}} \rightarrow \mathsf {DS}$ and $\mathsf {C}_{\mathbb {N}} \rightarrow \mathsf {BS,}$ respectively.

Proposition 9.7. $\mathsf {RT}^{{1}}_{{2}} \times \mathsf {DS}^{\mathsf {FE}} \leq _{\mathrm {W}} \mathsf {DS}$ and $\mathsf {RT}^{{1}}_{{2}} \times \mathsf {BS}^{\mathsf {FE}} \leq _{\mathrm {W}} \mathsf {BS}$ .

Proof. We only prove that $\mathsf {RT}^{{1}}_{{2}} \times \mathsf {DS}^{\mathsf {FE}} \leq _{\mathrm {W}} \mathsf {DS}$ ; the proof for $\mathsf {BS}$ can be obtained analogously by adapting the following argument. Let $(c,L)$ be an instance of ${\mathsf {RT}^{{1}}_{{2}} \times \mathsf {DS}^{\mathsf {FE}}}$ , where is a $2$ -coloring and $L=(x_n)_{{n\in \mathbb {N}}}$ is an ill-founded linear order. We build a linear order M in stages. Intuitively, we want to build a copy of L until we see that the current color changes. When this happens, we start building a fresh copy of L at the top of the current linear order. However, to guarantee that the resulting order will be ill-founded, we also build another copy of L at the bottom, which gets expanded every time two color changes are observed.

Formally, we construct the order M as follows: elements are of the form $(n,x,b,s)$ and are ordered as

$$\begin{align*}(n,x,b,s) \le_M (m,y,b',s') :\hspace{-1mm}\iff n<m \text{ or } (n=m \text{ and } x\le_L y). \end{align*}$$

The third and fourth components play no role in the order M. The construction below will guarantee that this is a linear order (rather than a quasi-order).

At stage $0$ , we add $(1,x_0,c(0),0)$ to M. We also define $p_0:=-1$ , $q_0:=0$ , and $h_0:=1$ . Intuitively, $p_s$ stores the last position in the bottom copy of L and $q_s$ stores the last position in the top copy of L.

At stage $s+1$ , we distinguish the following cases:

  • If $c(s+1)=c(s)=b$ , we add $(h_s,x_{q_s+1},b, s+1)$ to M. We define $q_{s+1}:=q_s+1$ , $p_{s+1}:=p_s$ , $h_{s+1}:=h_s$ and go to the next stage.

  • If $c(s+1)=1$ and $c(s)=0$ , we add $(h_s+1,x_{0},1,s+1)$ to M. We define $q_{s+1}:=0$ , $p_{s+1}:=p_s$ , $h_{s+1}:=h_s+1$ and go to the next stage.

  • If $c(s+1)=0$ and $c(s)=1$ , we add $(h_s+1,x_{0},0, s+1)$ to M. We also add $(0,x_{p_s+1},2, s+1)$ to M. We define $q_{s+1}:=0$ , $p_{s+1}:=p_s+1$ , $h_{s+1}:=h_s+1$ and go to the next stage.

This concludes the construction.

Observe that if the coloring c is eventually constant and $\lim _n c(n)=b$ , then $M \cong N+L$ for some $N\in \mathbb {N}$ (in particular it is ill-founded) and every descending sequence through M is of the form $((n,y_i,b,s_i))_{{i\in \mathbb {N}}}$ for some n, where $(y_i)_{{i\in \mathbb {N}}}$ is an L-descending sequence. In particular, we can compute a solution for $(\mathsf {RT}^{{1}}_{{2}} \times \mathsf {DS}^{\mathsf {FE}})(c,L)$ via projections.

On the other hand, if c is not eventually constant, then $M \cong L + \omega $ . Every descending sequence through M must eventually list elements of the initial segment of M isomorphic to L. In particular, if $((n_i,y_i,b_i,s_i))_{{i\in \mathbb {N}}}$ is an M-descending sequence then $b_0\in \mathsf {RT}^{{1}}_{{2}}(c)$ (trivially) and $(y_i)_{{i\in \mathbb {N}}} \in \mathsf {DS}^{\mathsf {FE}}(L)$ .

Let $\operatorname {Sort}_2: {2^{\mathbb {N}}} \to {2^{\mathbb {N}}}$ be defined by $\operatorname {Sort}_2(p) = 0^i1^\omega $ if there are exactly i-many n such that $p(n)=0$ , otherwise $\operatorname {Sort}_2(p) = 0^\omega $ . While $\operatorname {Sort}_2$ and $\mathsf {RT}^{{1}}_{{2}}$ are quite weak on their own, their parallel product does not reduce to $\mathsf {BS}$ .

Proposition 9.8. $\operatorname {Sort}_2 \times \mathsf {RT}^{{1}}_{{2}}\not \leq _{\mathrm {W}} \mathsf {BS}$ .

Proof. Observe first of all that $\mathsf {BS}$ is a cylinder, i.e., $\mathrm {id} \times \mathsf {BS} \leq _{\mathrm {sW}} \mathsf {BS}$ (this can be easily proved by adapting the proof of [Reference Goh, Pauly and Valenti18, Proposition 4.6]). This means that it suffices to prove $\operatorname {Sort}_2 \times \mathsf {RT}^{{1}}_{{2}} \not \leq _{\mathrm {sW}} \mathsf {BS}$ (see [Reference Brattka, Gherardi, Pauly, Brattka and Hertling3, Proposition 11.3.5]).

Assume towards a contradiction that the functionals $\Phi ,\Psi _0,\Psi _1$ witness the strong reduction $\operatorname {Sort}_2 \times \mathsf {RT}^{{1}}_{{2}}\leq _{\mathrm {sW}} \mathsf {BS}$ (where $\Psi _0$ and $\Psi _1$ produce, respectively, the $\operatorname {Sort}_2$ and the $\mathsf {RT}^{{1}}_{{2}}$ answers). We build an instance $(p,c)$ of $\operatorname {Sort}_2 \times \mathsf {RT}^{{1}}_{{2}}$ in stages as follows. At each stage s, we have a finite string $\rho _s$ , a finite coloring $\gamma _s$ , and a set $D_s$ of finite bad sequences in the quasi-order $\preceq _s$ produced by $\Phi (\rho _s,\gamma _s)$ in s steps. Intuitively, the set $D_s$ corresponds to the set of finite bad sequences that have already been forced to be non-extendible. Recall the quasi-order $\trianglelefteq $ (Definition 4.1) on bad sequences.

We start with $\rho _0 := \gamma _0 := \varepsilon $ and $D_0:=\emptyset $ . At stage $s+1=2n+1$ , we consider the set $M_{s+1}$ of all the $\trianglelefteq $ -maximal bad sequences $\beta $ in $\preceq _{s+1}$ that are not in $D_s$ and such that:

  • $\Psi _1(\beta )(0)\downarrow =k<2;$

  • $\Psi _0(\beta )$ did not commit to a finite number of zeroes for the $\operatorname {Sort}_2$ solution (i.e., the string produced by $\Psi _0(\beta )$ in $s+1$ steps is not of the form for any i or $\sigma $ ).

For each such $\beta $ , we search for a sufficiently large $m\in \mathbb {N}$ such that there exists a finite bad sequence $\alpha $ in such that $\beta \trianglelefteq \alpha $ and there is i such that $0^i1 \sqsubseteq \Psi _0(\alpha )$ within $s+1+m$ steps. In other words, we extend our finite input of $\operatorname {Sort}_2$ and finite coloring $\gamma _s$ so as to force the forward functional to produce some finite bad sequence $\alpha \trianglerighteq \beta $ such that $\Psi _0(\alpha )$ commits to a finite number of zeroes for the $\operatorname {Sort}_2$ solution. Observe that this needs to happen because $\beta $ cannot be extendible in (it is producing the wrong $\mathsf {RT}^{{1}}_{{2}}$ answer) and hence, by Lemma 4.2, there is a finite bad sequence $\alpha $ in as desired. We diagonalize against $\alpha $ by extending with $(i+1)$ -many zeroes. We then add $\alpha $ and every bad sequence $\trianglelefteq $ -below $\alpha $ to $D_s$ . This concludes the action needed for $\beta $ . Once all the bad sequences in $M_{s+1}$ have received attention, we go to the next stage.

At stage $s+1=2n+2$ , we consider the set $M_{s+1}$ of all the $\trianglelefteq $ -maximal bad sequences $\beta $ in $\preceq _{s+1}$ that are not in $D_s$ and such that, for some i, $0^i1 \sqsubseteq \Psi _0(\beta )$ . Let

$$\begin{align*}j:= \max \{ i {}\,:\,{} (\exists \beta \in M_{s+1})(0^i1\sqsubseteq \Psi_0(\beta)) \}+1.\end{align*}$$

We can uniformly compute such j because $M_{s+1}$ is finite. We define and . We also add the $\trianglelefteq $ -downward closure of $M_{s+1}$ into $D_s$ and go to the next stage.

This concludes the construction. It is apparent that the strings $\rho _s$ and $\gamma _s$ are extended infinitely often, and therefore the pair $(p,c)$ with $p:= \bigcup _s \rho _s$ and $c:=\bigcup _s \gamma _s$ is a valid instance for $\operatorname {Sort}_2\times \mathsf {RT}^{{1}}_{{2}}$ . Let $\preceq $ be the non-well quasi-order defined by $\Phi (p,c)$ and let $B\in \mathsf {BS}(\Phi (p,c))$ . By continuity, let n be such that $\Psi _1(B[n])\downarrow $ and let $s+1$ be the first stage by which every element of $B[n]$ enters $\preceq $ . We show that there is a finite bad sequence $\alpha \trianglerighteq B[n]$ which produces a wrong answer for $\operatorname {Sort}_2\times \mathsf {RT}^{{1}}_{{2}}(p,c)$ . This suffices to reach a contradiction, as, by Lemma 4.2, $B[n]$ being extendible implies that $\alpha $ is extendible.

Assume first that $\Psi _0(B[n])_{s+1}$ commits to finitely many zeroes. Let $t\ge s+1$ be even. Let $\alpha \trianglerighteq B[n]$ be $\trianglelefteq $ -maximal in $\preceq _{t}$ such that, for some i, $0^i1\sqsubseteq \Psi _0(\alpha )_{t}$ . By construction, at stage t we are adding more than i-many zeroes in p, therefore $\Psi _0(\alpha )$ is not the prefix of $\operatorname {Sort}_2(p)$ .

If $\Psi _0(B[n])_{s+1}$ does not commit to finitely many zeroes, let $r\ge s+1$ be odd. Let $\beta \trianglerighteq B[n]$ be $\trianglelefteq $ -maximal in $\preceq _{r}$ such that $\beta \in M_r$ . At stage r, we forced the forward functional $\Phi $ to produce some $\alpha \trianglerighteq \beta $ such that, for some i, $0^i1\sqsubseteq \Psi _0(\alpha )$ , and then extended $\rho _r$ by adding more than i-many zeroes. In other words, $\Psi _0(\alpha )$ is not a prefix of $\operatorname {Sort}_2(p)$ , and this concludes the proof.

Observe that $\operatorname {Sort}_2$ is one of the weakest “natural” problems that require infinitely many mind-changes to be solved. The previous result shows that, roughly speaking, while the quotient of $\mathsf {DS}$ (resp., $\mathsf {BS}$ ) over $\mathsf {RT}^{{1}}_{{2}}$ is capable of solving every problem that can be solved with finitely-many mind changes (essentially because $\mathsf {RT}^{{1}}_{{2}}\times \mathsf {RT}^{{1}}_{{\mathbb {N}}}\equiv _{\mathrm {W}} \mathsf {RT}^{{1}}_{{\mathbb {N}}}\leq _{\mathrm {W}} \mathsf {DS}$ , see [Reference Goh, Pauly and Valenti18, Proposition 4.24]), allowing infinitely-many mind changes constitutes a critical obstacle. This leads to the following open problem.

Open Question 9.9. Characterize $\mathsf {DS}/\mathsf {RT}^{{1}}_{{2}}$ and $\mathsf {BS}/\mathsf {RT}^{{1}}_{{2}}$ .

In an attempt to better understand the stability of $\mathsf {DS}$ and $\mathsf {BS}$ under parallel products, we also highlight the following open question.

Open Question 9.10. What are the degrees of $\mathsf {DS}/\mathsf {DS}$ and $\mathsf {BS}/\mathsf {BS}$ ?

Observe that, as a consequence of Proposition 9.4 and Proposition 9.8, $\mathsf {C}_{\mathbb {N}}\leq _{\mathrm {W}} \mathsf {DS}/\mathsf {DS} <_{\mathrm {W}} \mathsf {lim}$ . It is therefore natural to ask whether $\mathsf {DS}/\mathsf {DS} \leq _{\mathrm {W}} \mathsf {C}_{\mathbb {N}}$ . A possible way to separate $\mathsf {DS}/\mathsf {DS}$ and $\mathsf {C}_{\mathbb {N}}$ would be showing that $\mathsf {DS}$ is equivalent to its restriction to inputs L with no L-computable solution. This can be rephrased as asking whether $\mathsf {NON}\times \mathsf {DS} \leq _{\mathrm {W}} \mathsf {DS}$ , where $\mathsf {NON}(p):=\{q\in \mathbb {N}^{\mathbb {N}} {}\,:\,{} q\not \le _{\mathrm {T}} p\}$ . The same observations can be made by substituting $\mathsf {DS}$ with $\mathsf {BS}$ .

Acknowledgements

We are grateful to Takayuki Kihara for pointing out the mistake in our previous article. We also thank Cécilia Pradic for comments which greatly improved the presentation of some results. We thank the anonymous reviewers for their careful reading of the article. This article extends the conference paper [Reference Goh, Pauly, Valenti, Patey, Pimentel, Galeotti and Manea19].

References

Brattka, V., On the complexity of learning programs , Unity of Logic and Computation (Vedova, G. D., Dundua, B., Lempp, S., and Manea, F., editors), Lecture Notes in Computer Science, 13967, Springer, Cham, 2023, pp. 166177.Google Scholar
Brattka, V., Gherardi, G., and Marcone, A., The Bolzano-Weierstrass theorem is the jump of Weak König’s lemma . Annals of Pure and Applied Logic , vol. 163 (2012), no. 6, pp. 623655.Google Scholar
Brattka, V., Gherardi, G., and Pauly, A., Weihrauch complexity in computable analysis , Handbook of Computability and Complexity in Analysis (Brattka, V. and Hertling, P., editors), Springer International Publishing, Springer, Cham, 2021, pp. 367417.Google Scholar
Brattka, V. and Pauly, A., On the algebraic structure of Weihrauch degrees . Logical Methods in Computer Science , vol. 14 (2018), no. 4, pp. 136.Google Scholar
Brattka, V. and Rakotoniaina, T., On the uniform computational content of Ramsey’s theorem . The Journal of Symbolic Logic , vol. 82 (2017), no. 4, pp. 12781316.Google Scholar
Calvert, W., Franklin, J. N. Y., and Turetsky, D., Structural highness notions . The Journal of Symbolic Logic , vol. 88 (2023), no. 4, pp. 16921724.Google Scholar
Carlucci, L., Mainardi, L., and Zdanowski, K., Reductions of well-ordering principles to combinatorial theorems, preprint, 2024.Google Scholar
Chubb, J., Hirst, J. L., and McNicholl, T. H., Reverse mathematics, computability, and partitions of trees . The Journal of Symbolic Logic , vol. 74 (2009), no. 1, pp. 201215.Google Scholar
Cipriani, V., Marcone, A., and Valenti, M., The Weihrauch lattice at the level of ${\varPi}_1^1{-}{\mathsf{CA}}_0$ : The Cantor-Bendixson theorem . The Journal of Symbolic Logic , vol. 90 (2025), no. 2, pp. 752790.Google Scholar
Cipriani, V. and Pauly, A., Embeddability of graphs and Weihrauch degrees, Journal of Mathematical Logic , to appear. published online.Google Scholar
Downey, R. G., Computability theory and linear orderings , Handbook of Recursive Mathematics , vol. 2, (Ershov, Y. L., Goncharov, S. S., Nerode, A., Remmel, J. B., and Marek, V. W., editors), Studies in Logic and the Foundations of Mathematics, 139, North-Holland, Amsterdam, 1998, pp. 823976.Google Scholar
Dzhafarov, D. D., Goh, J. L., Hirschfeldt, D. R., Patey, L., and Pauly, A., Ramsey’s theorem and products in the Weihrauch degrees . Computability , vol. 9 (2020), no. 2, pp. 85110.Google Scholar
Dzhafarov, D. D., Solomon, R., and Yokoyama, K., On the first-order parts of problems in the Weihrauch degrees . Computability , vol. 13 (2024), nos. 3–4, pp. 363375.Google Scholar
Freund, A., Pakhomov, F., and Soldà, G., The logical strength of minimal bad arrays . Proceedings of the American Mathematical Society , vol. 152 (2024), no. 12, pp. 49935005.Google Scholar
Higuchi, K. and Pauly, A., The degree structure of Weihrauch reducibility . Logical Methods in Computer Science , vol. 9 (2013), no. 2, pp. 117.Google Scholar
Hirschfeldt, D. R. and Jockusch, C. G. Jr., On notions of computability-theoretic reduction between ${\varPi}_2^1$ principles . Journal of Mathematical Logic , vol. 16 (2016), no. 1, Article no. 1650002, 59 pp.Google Scholar
Kihara, T., Marcone, A., and Pauly, A., Searching for an analogue of $AT{R}_0$ in the Weihrauch lattice . The Journal of Symbolic Logic , vol. 85 (2020), no. 3, pp. 10061043.Google Scholar
Goh, J. L., Pauly, A., and Valenti, M., Finding descending sequences through ill-founded linear orders . The Journal of Symbolic Logic , vol. 86 (2021), no. 2, pp. 817854. https://arxiv.org/abs/2010.03840.Google Scholar
Goh, J. L., Pauly, A., and Valenti, M., The weakness of finding descending sequences in ill-founded linear orders , Twenty Years of Theoretical and Practical Synergies (Patey, L. L., Pimentel, E., Galeotti, L., and Manea, F., editors), Lecture Notes in Computer Science, Springer Nature Switzerland, Cham, 2024, pp. 339350.Google Scholar
Le Roux, S. and Pauly, A., Finite choice, convex choice and finding roots . Logical Methods in Computer Science , vol. 11 (2015), no. 4, p. 4:6, 30 pp.Google Scholar
Lempp, S., Marcone, A., and Valenti, M., Chains and antichains in the Weihrauch lattice, preprint, 2024.Google Scholar
Marcone, A., WQO and BQO Theory in Subsystems of Second Order Arithmetic , Lecture Notes in Logic, Cambridge University Press, Assoc. Symbol. Logic, La Jolla, CA, 2005, pp. 303330.Google Scholar
Neumann, E. and Pauly, A., A topological view on algebraic computation models . Journal of Complexity , vol. 44 (2018), pp. 122.Google Scholar
Pauly, A., Pradic, C., and Soldà, G., On the Weihrauch degree of the additive Ramsey theorem , Computability , vol. 13 (2024), pp. 459483.Google Scholar
Pauly, A. and Soldà, G., Sequential discontinuity and first-order problems , Twenty Years of Theoretical and Practical Synergies (Patey, L. L., Pimentel, E., Galeotti, L., and Manea, F., editors), Springer Nature Switzerland, Cham, 2024, pp. 351365.Google Scholar
Sacks, G. E., Higher Recursion Theory , first ed., Perspectives in Mathematical Logic, Springer-Verlag, Berlin, 1990.Google Scholar
Simpson, S. G., Subsystems of Second Order Arithmetic , second ed., Perspectives in Logic, Cambridge University Press, Cambridge; Association for Symbolic Logic, Poughkeepsie, 2009.Google Scholar
Soldà, G. and Valenti, M., Algebraic properties of the first-order part of a problem . Annals of Pure and Applied Logic , vol. 174 (2023), no. 7, Article no. 103270, 41 pp.Google Scholar
Westrick, L. B., A note on the diamond operator . Computability , vol. 10 (2021), no. 2, pp. 107110.Google Scholar
Figure 0

Figure 1 Schematic representation of the construction used in the proof of Lemma 3.5. The dashed box contains the partial construction up to stage s. The gray boxes contain intervals in the partial order P. For simplicity, each interval up to stage s only contains one point (i.e., the partial order $P_s$ is isomorphic to a tree), but this need not be the case in general. The black nodes are those in the range of the (partial) tree decomposition $\iota _s$. The square gray node is the node guessed by g at stage $s+1$ (which identifies $\iota (\sigma )$ and ).

Figure 1

Figure 2 The figure shows all reductions between the depicted principles up to transitivity, with the potential exception that we have not ruled out being $\mathsf {BS}|_{\operatorname {Tree}} \leq _{\mathrm {W}} \widehat {\boldsymbol {\Pi }^0_2\mathsf {-ACC}_{\mathbb {N}}}$. The diagram can be thought of as a cube whose vertical edges have decreasing chains parameterized by $k \geq 2$ on them, and the bottom corners are labelled by degrees which are below the ones on the edges for all k.