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

The logical strength of minimal bad arrays

Anton Freund Fedor Pakhomov  and  Giovanni Soldà Anton Freund, University of Würzburg, Institute of Mathematics, Emil-Fischer-Straße 40, 97074 Würzburg, Germany [email protected] Fedor Pakhomov and Giovanni Soldà, Department of Mathematics: Analysis, Logic and Discrete Mathematics, Ghent University, Krijgslaan 281 S8, 9000 Ghent, Belgium [email protected], [email protected]
Abstract.

This paper studies logical aspects of the notion of better quasi order, which has been introduced by C. Nash-Williams (Mathematical Proceedings of the Cambridge Philosophical Society 1965 & 1968). A central tool in the theory of better quasi orders is the minimal bad array lemma. We show that this lemma is exceptionally strong from the viewpoint of reverse mathematics, a framework from mathematical logic. Specifically, it is equivalent to the set existence principle of Π21\Pi^{1}_{2}-comprehension, over the base theory 𝖠𝖳𝖱𝟢\mathsf{ATR_{0}}.

Key words and phrases:
Better quasi order, minimal bad array, Π21\Pi^{1}_{2}-comprehension, reverse mathematics
2020 Mathematics Subject Classification:
06A06, 03B30, 03F35
The work of Anton Freund has been funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project number 460597863. The work of Fedor Pakhomov and Giovanni Soldà has been funded by the FWO grant G0F8421N

1. Introduction

A quasi order QQ with order relation Q\leq_{Q} is a well quasi order if any infinite sequence q0,q1,q_{0},q_{1},\ldots in QQ admits i<ji<j with qiQqjq_{i}\leq_{Q}q_{j}. There is a deep theory of well quasi orders [4] with the graph minor theorem [16] as a towering achievement. A limitation of well quasi orders is the lack of closure properties under infinitary operations. In particular, R. Rado [14] has exhibited a well quasi order QQ such that the power set 𝒫(Q)\mathcal{P}(Q) is no well quasi order when we put

X𝒫(Q)Y:for any pX there is a qY with pQq.X\leq_{\mathcal{P}(Q)}Y\quad:\Leftrightarrow\quad\text{for any\leavevmode\nobreak\ $p\in X$ there is a\leavevmode\nobreak\ $q\in Y$ with $p\leq_{Q}q$}.

To secure such closure properties, C. Nash-Williams [11, 12] has introduced the more restrictive notion of better quasi order, which we recall below. As a famous application, we mention R. Laver’s proof [5] of Fraïssé’s conjecture that the embeddability relation on σ\sigma-scattered linear orders is a well and even a better quasi order. A central tool in the theory of better quasi orders is the minimal bad array lemma, which is also known as the forerunning principle. In the present paper, we use methods from logic to show that any proof of that principle must use exceptionally strong set existence axioms. A precise formulation of this result is given below.

Let us introduce some notation that is needed to define the notion of better quasi order. The collections of finite and countably infinite subsets of a set VV are denoted by [V]<ω[V]^{<\omega} and [V]ω[V]^{\omega}, respectively. Sets from []<ω[]ω[\mathbb{N}]^{<\omega}\cup[\mathbb{N}]^{\omega} are identified with the sequences that enumerate them in increasing order. For such sequences we write sts\sqsubset t to assert that ss is a proper initial segment of tt. Note that this forces ss to be finite. For X[]ωX\in[\mathbb{N}]^{\omega} with least element xx, we put X:=X\{x}X^{-}:=X\backslash\{x\}. In the following we use infix notation pRqp\,R\,q to express (p,q)R(p,q)\in R.

Definition 1.1.

Consider a set QQ with a binary relation RR. By a QQ-array we mean a map f:[V]ωQf:[V]^{\omega}\to Q with V[]ωV\in[\mathbb{N}]^{\omega} that is continuous in the sense that each X[V]ωX\in[V]^{\omega} admits an sXs\sqsubset X such that ff is constant on {Y[V]ω|sY}\{Y\in[V]^{\omega}\,|\,s\sqsubset Y\}. Such an ff is called RR-bad if there is no X[V]ωX\in[V]^{\omega} with f(X)Rf(X)f(X)\,R\,f(X^{-}). We say that RR is a better relation on QQ if no QQ-array is RR-bad. A better quasi order is a quasi order such that the order relation is a better relation on the underlying set.

It is instructive to show that any better quasi order is a well quasi order. In the proof of Proposition 2.2, we will implicitly confirm that Q𝒫(Q)Q\mapsto\mathcal{P}(Q) preserves better quasi orders. The case of better relations that are not quasi orders has not been studied widely, even though it is mentioned in [13, 17]. We will consider it in an intermediate argument. Let us note that any better relation is reflexive.

The given definition of better quasi orders does essentially coincide with the original definition in terms of blocks. We recall that a block with base V[]ωV\in[\mathbb{N}]^{\omega} is a set B[V]<ωB\subseteq[V]^{<\omega} such that any X[V]ωX\in[V]^{\omega} has a unique initial segment in BB, which may not be empty. The latter ensures that the base can be recovered as B\bigcup B. Any function f0:BQf_{0}:B\to Q on a block BB with base VV induces a QQ-array f:[V]ωQf:[V]^{\omega}\to Q given by f(X)=f0(s)f(X)=f_{0}(s) for sBs\in B with sXs\sqsubset X. Continuity entails that any array is induced in this way. In the following, we implicitly assume that all arrays are given as functions on blocks. This has the advantage that arrays become countable objects, which will be relevant for our logical analysis. Later, we will briefly discuss another equivalent definition of better quasi orders in terms of special blocks, which are called barriers. There is a further equivalent definition due to S. Simpson [18], who admits all Borel measurable functions as arrays. In our logical analysis, we will not consider this last definition by Simpson.

Now that we have discussed the definition of better quasi order, we introduce the notions that appear in the minimal bad array lemma.

Definition 1.2.

A partial ranking of a reflexive binary relation RR on a set QQ is a well-founded partial order \leq^{\prime} on QQ such that pRqrp\,R\,q\leq^{\prime}r entails pRrp\,R\,r. Given such a ranking and QQ-arrays f:[V]ωQf:[V]^{\omega}\to Q and g:[W]ωQg:[W]^{\omega}\to Q, we write fgf\leq^{\prime}g to express that we have VWV\subseteq W as well as f(X)g(X)f(X)\leq^{\prime}g(X) for all X[V]ωX\in[V]^{\omega}. In case we even have f(X)<g(X)f(X)<^{\prime}g(X) for all X[V]ωX\in[V]^{\omega}, we write f<gf<^{\prime}g. A QQ-array gg is said to be \leq^{\prime}-minimal RR-bad if it is RR-bad and there is no RR-bad QQ-array f<gf<^{\prime}g.

Let us note that qrq\leq^{\prime}r implies qRrq\,R\,r when \leq^{\prime} is a partial ranking of RR, given that the latter is reflexive. Conversely, if RR is transitive and qrq\leq^{\prime}r implies qRrq\,R\,r, then pRqrp\,R\,q\leq^{\prime}r implies pRrp\,R\,r. So in the case of quasi orders, our notion of partial ranking coincides with the usual one. The proofs of Proposition 2.2 and Theorem 3.5 suggest that the given definition is pertinent for relations that are not transitive. We can now present one of the central tools from the theory of better quasi orders.

Minimal bad array lemma (‘Simpson’s version’).

Consider a quasi order QQ and a partial ranking \leq^{\prime} of the order relation Q\leq_{Q} on QQ. Whenever f0f_{0} is a Q\leq_{Q}-bad QQ-array, there is a QQ-array ff0f\leq^{\prime}f_{0} that is \leq^{\prime}-minimal Q\leq_{Q}-bad.

The given version coincides with the one of Simpson [18], except that the latter works with a larger class of Borel measurable arrays, as mentioned above. A different definition of the relation <<^{\prime} between arrays has been given by Laver [6]. It leads to a different variant of the minimal bad array lemma, which we will later introduce as Laver’s version. In both cases, we have a version for quasi orders and an obvious generalization to reflexive relations. We will see that our results apply to all indicated versions.

Our logical analysis takes place within the framework of reverse mathematics. For a comprehensive introduction, we refer to the founding paper by H. Friedman [3] and the textbook by S. Simpson [19]. To give a rather informal account of a central idea, we note that each class 𝒞\mathcal{C} of properties (more precisely: formulas of second order arithmetic) gives rise to an axiomatic principle called 𝒞\mathcal{C}-comprehension, which asserts that the set {n|φ(n)}\{n\in\mathbb{N}\,|\,\varphi(n)\} exists for any φ\varphi from 𝒞\mathcal{C}. One writes 𝒞-CA0\mathcal{C}\textsf{-CA}_{0} for the theory that is axiomatized by 𝒞\mathcal{C}-comprehension and some basic facts about the natural numbers. Important cases include the class Δ10\Delta^{0}_{1} of properties that are algorithmically decidable, the class Π0\Pi^{0}_{\infty} of arithmetical properties (given by formulas that quantify over finite but not over infinite objects), the class Π11\Pi^{1}_{1} of coanalytic properties, and the class Π21\Pi^{1}_{2} of properties that have the form X:ψ(n,X)\forall X\subseteq\mathbb{N}:\psi(n,X) for analytic ψ\psi. The theories Δ10-CA0\Delta^{0}_{1}\textsf{-CA}_{0} and Π0-CA0\Pi^{0}_{\infty}\text{-CA}_{0} are usually denoted by 𝖱𝖢𝖠𝟢\mathsf{RCA_{0}} (recursive comprehension) and 𝖠𝖢𝖠𝟢\mathsf{ACA_{0}} (arithmetical comprehension), respectively. We will also encounter the theory 𝖠𝖳𝖱𝟢\mathsf{ATR_{0}} of arithmetic transfinite recursion, which permits to iterate arithmetical comprehension along arbitrary well orders.

To analyze a given theorem in reverse mathematics, one will often determine a class 𝒞\mathcal{C} such that the principle of 𝒞\mathcal{C}-comprehension is equivalent to the theorem in question. Such an equivalence is informative if one proves it over a sufficiently weak base theory. A common choice is 𝖱𝖢𝖠𝟢\mathsf{RCA_{0}}, but we will mostly work with 𝖠𝖳𝖱𝟢\mathsf{ATR_{0}} as base theory. This makes sense because Simpson’s version of the minimal bad array lemma is typically used in conjunction with the open Ramsey theorem, which is equivalent to arithmetic transfinite recursion over 𝖱𝖢𝖠𝟢\mathsf{RCA_{0}}.

Now that both the mathematical notions and the logical framework have been fixed, we can formulate the main result of the present paper.

Theorem 1.3.

The following are equivalent over 𝖠𝖳𝖱𝟢\mathsf{ATR_{0}}:

  1. (i)

    the principle of Π21\Pi^{1}_{2}-comprehension,

  2. (ii)

    Simpson’s version of the minimal bad array lemma (as stated above),

  3. (iii)

    Laver’s version of the minimal bad array lemma (as stated in Section 3).

Let us now put our result into context. In his textbook on reverse mathematics, Simpson [19] states “that a great many ordinary mathematical theorems are provable in 𝖠𝖢𝖠𝟢\mathsf{ACA_{0}}, and that of the exceptions, most are provable in Π11-CA0\Pi^{1}_{1}\textsf{-CA}_{0}.” In [8] it is shown that Π11\Pi^{1}_{1}-comprehension is equivalent to the minimal bad sequence lemma from the theory of well quasi orders. It may not be entirely unexpected that minimal bad arrays correspond to Π21\Pi^{1}_{2}-comprehension, since A. Marcone [8] has shown that the notion of better quasi order is Π21\Pi^{1}_{2}-complete. Nevertheless, Marcone himself has formulated the much weaker conjecture that the minimal bad array lemma is unprovable in Π11-CA0\Pi^{1}_{1}\textsf{-CA}_{0} (see Problem 2.19 from [9]). The strength of Π21\Pi^{1}_{2}-comprehension dwarfs the one of Π11\Pi^{1}_{1}-comprehension, as expressed by M. Rathjen [15]. In their analysis of a theorem from topology, C. Mummert and S. Simpson [10] assert that they have found “the first example of a theorem of core mathematics which is provable in second order arithmetic and implies Π21\Pi^{1}_{2} comprehension.” So it seems justified to conclude that the minimal bad array lemma has exceptional axiomatic strength.

2. From minimal bad arrays to Π21\Pi^{1}_{2}-comprehension

In this section, we derive Π21\Pi^{1}_{2}-comprehension from Simpson’s version of the minimal bad array lemma, which yields the first implication in Theorem 1.3 from the introduction.

We will make crucial use of Marcone’s result that the notion of better quasi order is Π21\Pi^{1}_{2}-complete. While Marcone does not specify a base theory, it is straightforward to check that his proof yields the following.

Proposition 2.1 ([7]).

Consider a Π21\Pi^{1}_{2}-formula φ(n)\varphi(n) in the language of second order arithmetic, possibly with further parameters. The theory 𝖠𝖢𝖠𝟢\mathsf{ACA_{0}} proves that, for any values of the parameters, there is a family of sets QnQ_{n} and reflexive binary relations RnR_{n} such that φ(n)\varphi(n) holds precisely if RnR_{n} is a better relation on QnQ_{n}.

Marcone also shows that the relations RnR_{n} in the proposition can be turned into quasi orders, based on a construction of M. Pouzet [13]. However, the latter makes use of Π11\Pi^{1}_{1}-comprehension in the form of local minimal bad arrays. In any case, we will later construct a relation that is not transitive even when each RnR_{n} is. To reduce to quasi orders, we will then use the following result. It strengthens the result by Pouzet, which only covers the case where \leq^{\prime} is the identity relation. Since the minimal bad array in (ii) is given, we do not need a stronger base theory.

Proposition 2.2 (𝖠𝖳𝖱𝟢\mathsf{ATR_{0}}).

When \leq^{\prime} is a partial ranking of a reflexive relation RR on a set QQ, there is a well-founded partial order Q\leq_{Q} on QQ that validates the following:

  1. (i)

    We have QR{\leq^{\prime}}\subseteq{\leq_{Q}}\subseteq R.

  2. (ii)

    For any Q\leq_{Q}-minimal Q\leq_{Q}-bad array f:[V]ωQf:[V]^{\omega}\to Q, there is a set W[V]ωW\in[V]^{\omega} such that the restriction of ff to [W]ω[W]^{\omega} is RR-bad.

Proof.

Given that partial rankings are well-founded, we find a well order (α,)(\alpha,\preceq) and a bijection o:Qαo:Q\to\alpha such that pqp\leq^{\prime}q entails o(p)o(q)o(p)\preceq o(q). Let us note that the original proof by Pouzet, in which \leq^{\prime} is the identity relation, is based on an arbitrary enumeration with α=ω\alpha=\omega. Following Pouzet, we stipulate

p<Qq:{we have o(p)o(q) and pRqand r<Qp implies r<Qq for all rQ,p<_{Q}q\quad:\Leftrightarrow\quad\begin{cases}\text{we have $o(p)\prec o(q)$ and $p\,R\,q$}\\ \text{and $r<_{Q}p$ implies $r<_{Q}q$ for all $r\in Q$,}\end{cases}

which amounts to a recursion over o(p)o(p). It is immediate that this yields a well-founded partial order that is contained in RR. Aiming at (i), we assume p<qp<^{\prime}q. Given that <<^{\prime} is a partial ranking of RR, we have pRqp\,R\,q and also o(p)o(q)o(p)\prec o(q). To get p<Qqp<_{Q}q, we show that r<Qpr<_{Q}p entails r<Qqr<_{Q}q. We use induction on o(r)o(r). Given r<Qpr<_{Q}p, we have o(r)o(p)o(q)o(r)\prec o(p)\prec o(q) as well as rRp<qr\,R\,p\,<^{\prime}q. Crucially, the latter entails rRqr\,R\,q, due to our definition of partial ranking for relations that are not quasi orders. For r<Qr<Qpr^{\prime}<_{Q}r<_{Q}p we get r<Qqr^{\prime}<_{Q}q by induction hypothesis, as needed to conclude r<Qqr<_{Q}q. Let us now consider f:[V]ωQf:[V]^{\omega}\to Q as in (ii). By the open Ramsey theorem, we first find V[V]ωV^{\prime}\in[V]^{\omega} with o(f(X))o(f(X))o(f(X))\preceq o(f(X^{-})) for all X[V]ωX\in[V^{\prime}]^{\omega} (see Theorem 4.9 in conjunction with Lemma 3.1 of [9]). Consider the powerset 𝒫(Q)\mathcal{P}(Q) with the order that is induced by Q\leq_{Q} as in the introduction. We look at the array

h:[V]ω𝒫(Q)withh(X)={rQ|r<Qf(X)}.h:[V^{\prime}]^{\omega}\to\mathcal{P}(Q)\quad\text{with}\quad h(X)=\{r\in Q\,|\,r<_{Q}f(X)\}.

Let us note that the sets h(X)h(X) are finite when we have α=ω\alpha=\omega, as in the original proof by Pouzet. For W[V]ωW\in[V^{\prime}]^{\omega}, the restriction of hh to [W]ω[W]^{\omega} cannot be bad. If it was, we could choose values g(X)h(X)g(X)\in h(X) with g(X)Qrg(X)\not\leq_{Q}r for all rh(X)r\in h(X^{-}), by the definition of the order on 𝒫(Q)\mathcal{P}(Q). In view of g(X)h(X)g(X^{-})\in h(X^{-}), the resulting array g:[W]ωQg:[W]^{\omega}\to Q would be Q\leq_{Q}-bad. But we would also have g(X)<Qf(X)g(X)<_{Q}f(X) due to g(X)h(X)g(X)\in h(X), against the minimality of ff. Again by the open Ramsey theorem, we thus find a W[V]ωW\in[V^{\prime}]^{\omega} such that h(X)𝒫(Q)h(X)h(X)\leq_{\mathcal{P}(Q)}h(X^{-}) holds for all X[W]ωX\in[W]^{\omega}. Given r<Qf(X)r<_{Q}f(X) with X[W]ωX\in[W]^{\omega}, we can now argue that rh(X)r\in h(X) yields rQrr\leq_{Q}r^{\prime} for some rh(X)r^{\prime}\in h(X^{-}), where we have r<Qf(X)r^{\prime}<_{Q}f(X^{-}) and hence r<Qf(X)r<_{Q}f(X^{-}). We can conclude that f(X)Qf(X)f(X)\leq_{Q}f(X^{-}) is equivalent to f(X)Rf(X)f(X)\,R\,f(X^{-}) for X[W]ωX\in[W]^{\omega}. Given that ff is Q\leq_{Q}-bad, it must thus be RR-bad on [W]ω[W]^{\omega}, as desired. ∎

The following corollary shows that the minimal bad array lemma for quasi orders entails a version of the same lemma for reflexive relations.

Corollary 2.3 (𝖠𝖳𝖱𝟢\mathsf{ATR_{0}}).

Assume Simpson’s version of the minimal bad array lemma for quasi orders. Consider a partial ranking \leq^{\prime} of a reflexive relation RR on a set QQ. If RR is no better relation on QQ, there is a QQ-array that is \leq^{\prime}-minimal RR-bad.

Proof.

Consider QR{\leq_{Q}}\subseteq R as in the previous proposition. Given that RR is no better relation, we get a Q\leq_{Q}-bad array f0f_{0}. The minimal bad array lemma for quasi orders yields an array fQf0f^{\prime}\leq_{Q}f_{0} that is Q\leq_{Q}-minimal Q\leq_{Q}-bad. By the previous proposition, we can restrict ff^{\prime} to an array ff that is RR-bad. The latter must indeed be \leq^{\prime}-minimal RR-bad, as an RR-bad g<fg<^{\prime}f would be Q\leq_{Q}-bad with g<QfQfg<_{Q}f\leq_{Q}f^{\prime}. ∎

Let us note that the previous proof yields fQf0f\leq_{Q}f_{0} but not necessarily ff0f\leq^{\prime}f_{0}, so that we obtain an apparently weaker version of the minimal bad array lemma for reflexive relations. A direct proof of the full version is given in the next section. We now deduce the first direction of our main result.

Theorem 2.4 (𝖠𝖳𝖱𝟢\mathsf{ATR_{0}}).

The principle of Π21\Pi^{1}_{2}-comprehension follows from Simpson’s version of the minimal bad array lemma.

Proof.

Given a Π21\Pi^{1}_{2}-formula φ(n)\varphi(n), possibly with parameters, we consider sets QnQ_{n} and reflexive relations RnR_{n} as in Proposition 2.1. We need to form the set

X={n|Rn is a better relation on Qn}.X=\{n\in\mathbb{N}\,|\,R_{n}\text{ is a better relation on }Q_{n}\}.

Let ω\omega^{*} be the order with underlying set \mathbb{N} and order relation ={(m,n)|mn}{\leq^{*}}=\{(m,n)\,|\,m\geq n\}. For each nn, we consider the disjoint union Qn=QnωQ_{n}^{*}=Q_{n}\cup\omega^{*} with the relation

Rn=Rn{(q,m)|qQn and mω}.R_{n}^{*}=R_{n}\cup{\leq^{*}}\cup\{(q,m)\,|\,q\in Q_{n}\text{ and }m\in\omega^{*}\}.

Now define QQ as the set of sequences σ=σ0,,σl(σ)1\sigma=\langle\sigma_{0},\ldots,\sigma_{l(\sigma)-1}\rangle with length l(σ)l(\sigma) and entries σiQi\sigma_{i}\in Q_{i}^{*}. To obtain a reflexive relation RR on QQ, we stipulate

σRτeither l(σ)l(τ) or there is i<l(σ)<l(τ) with σiRiτi.\sigma\,R\,\tau\quad\Leftrightarrow\quad\text{either $l(\sigma)\geq l(\tau)$ or there is $i<l(\sigma)<l(\tau)$ with $\sigma_{i}\,R_{i}^{*}\tau_{i}$}.

Then RR is no better relation. Indeed, we obtain RiR_{i}^{*}-bad arrays fi:[]ωQif_{i}:[\mathbb{N}]^{\omega}\to Q_{i}^{*} by setting fi(X)=min(X)ωf_{i}(X)=\min(X)\in\omega^{*}. An RR-bad QQ-array is given by

Xf0(X),,fn(X)forn=min(X).X\mapsto\langle f_{0}(X),\ldots,f_{n}(X)\rangle\quad\text{for}\quad n=\min(X).

For σ,τQ\sigma,\tau\in Q we declare that στ\sigma\leq^{\prime}\tau holds if we have l(σ)=l(τ)l(\sigma)=l(\tau) and either σi=τi\sigma_{i}=\tau_{i} or σiQi\sigma_{i}\in Q_{i} and τiω\tau_{i}\in\omega^{*} for each i<l(σ)i<l(\sigma). It is straightforward to see that this yields a partial ranking of RR. In view of the previous corollary, we may now consider an array f:[V]ωQf:[V]^{\omega}\to Q that is \leq^{\prime}-minimal RR-bad. As noted in the introduction, we can identify this array with a map f:BQf:B\to Q on a block BB with base VV. For s,tBs,t\in B, one writes sts\vartriangleleft t if there is an X[V]ωX\in[V]^{\omega} with sXs\sqsubset X and tXt\sqsubset X^{-}. That ff is bad means that f(s)Rf(t)f(s)\,R\,f(t) fails for any sts\vartriangleleft t. We show

Ri is a better relation on Qif(s)iω for all sB with i<l(f(s)).\text{$R_{i}$ is a better relation on\leavevmode\nobreak\ $Q_{i}$}\quad\Leftrightarrow\quad\text{$f(s)_{i}\in\omega^{*}$ for all $s\in B$ with $i<l(f(s))$}.

Here f(s)if(s)_{i} refers to the ii-th entry of the sequence f(s)Qf(s)\in Q. Once the equivalence is established, the aforementioned set XX can be formed by arithmetic comprehension. Let us first assume that RiR_{i} is a better relation on QiQ_{i}. Towards a contradiction, we assume that f(s)iQif(s)_{i}\in Q_{i} holds for some sBs\in B with i<l(f(s))i<l(f(s)). Let B/sB/s be the block of all tBt\in B such that the largest number in ss lies strictly below the smallest number in tt. Given tB/st\in B/s, we find riBr^{i}\in B with s=r0rn=ts=r^{0}\vartriangleleft\ldots\vartriangleleft r^{n}=t (consider some X[V]ωX\in[V]^{\omega} with stXs\cup t\sqsubset X). Since f(rj)Rf(rj+1)f(r^{j})\,R\,f(r^{j+1}) must fail for all j<nj<n, we obtain l(f(rj))<l(f(rj+1))l(f(r^{j}))<l(f(r^{j+1})) and inductively f(rj)iQif(r^{j})_{i}\in Q_{i}. Hence we have i<l(f(s))<l(f(t))i<l(f(s))<l(f(t)) and f(t)iQif(t)_{i}\in Q_{i}. But then the array B/stf(t)iQiB/s\ni t\mapsto f(t)_{i}\in Q_{i} is RiR_{i}-bad, against our assumption. For the other direction of our equivalence, we assume f(s)iωf(s)_{i}\in\omega^{*} holds for all sBs\in B with i<l(f(s))i<l(f(s)). Pick r0rir^{0}\vartriangleleft\ldots\vartriangleleft r^{i} in BB. As above, we inductively get jl(f(rj))j\leq l(f(r^{j})) and hence i<l(f(s))i<l(f(s)) for all sB/ris\in B/r^{i}. In other words, if V[V]ωV^{\prime}\in[V]^{\omega} is the base of B/riB/r^{i}, we have i<l(f(X))i<l(f(X)) and f(X)iωf(X)_{i}\in\omega^{*} for all X[V]ωX\in[V^{\prime}]^{\omega}. Towards a contradiction, we now assume that RiR_{i} is not better. Let h:[W]ωQih:[W]^{\omega}\to Q_{i} be an RiR_{i}-bad array. We may assume WVW\subseteq V^{\prime}. In order to define g:[W]ωQg:[W]^{\omega}\to Q, we stipulate that we have l(g(X))=l(f(X))l(g(X))=l(f(X)) as well as

g(X)j={h(X)if j=i,f(X)jif ij<l(g(X)).g(X)_{j}=\begin{cases}h(X)&\text{if $j=i$},\\ f(X)_{j}&\text{if $i\neq j<l(g(X))$}.\end{cases}

This yields an RR-bad array g<fg<^{\prime}f, against minimality. ∎

3. From Π21\Pi^{1}_{2}-comprehension to minimal bad arrays

In this section, we first show that Simpson’s version of the minimal bad array lemma follows from another version that has been isolated by Laver [6]. We then prove that Laver’s version is a consequence of Π21\Pi^{1}_{2}-comprehension. Together with Theorem 2.4, this completes the proof of Theorem 1.3 from the introduction.

So far, we have mostly viewed arrays as continuous functions with domain [V]ω[V]^{\omega} for some infinite VV\subseteq\mathbb{N}, even though it was agreed that these were represented by functions on blocks. In other words, while different blocks can be used to represent the same continuous function, most previous considerations did not depend on the choice of representative. This is in marked contrast with the following, where the blocks on which arrays are defined form a crucial part of the data. We note that the following definitions are taken from [6], where \ensurestackMath\stackengine.5exUcFFS\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}} is written \mathrel{\text{\scalebox{1.2}{$\sqsupseteq$}}}.

Definition 3.1.

For blocks BB and CC, we write B\ensurestackMath\stackengine.5exUcFFSCB\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}C if we have BC\textstyle\bigcup B\subseteq\textstyle\bigcup C and each tBt\in B admits an sCs\in C with sts\sqsubseteq t. If this holds and we have BCB\not\subseteq C, then we write BCB\lessdot C. Given arrays f:BQf:B\to Q and g:CQg:C\to Q as well as a partial order \leq^{\prime} on the set QQ, we write f\ensurestackMath\stackengine.5exUcFFSgf\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}g if we have B\ensurestackMath\stackengine.5exUcFFSCB\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}C as well as f(t)=g(t)f(t)=g(t) for tBCt\in B\cap C and f(t)<g(s)f(t)<^{\prime}g(s) for CstBC\ni s\sqsubset t\in B. If we additionally have BCB\lessdot C, we write fgf\lessdot^{\prime}g.

It is immediate that \ensurestackMath\stackengine.5exUcFFS\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}} is transitive on blocks and not hard to derive that \ensurestackMath\stackengine.5exUcFFS\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime} is transitive on arrays. One can also show that BDB\lessdot D follows from BC\ensurestackMath\stackengine.5exUcFFSDB\lessdot C\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}D (but not from B\ensurestackMath\stackengine.5exUcFFSCDB\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}C\lessdot D), which yields the analogous fact for arrays.

Minimal bad array lemma (‘Laver’s version’).

Consider a partial ranking \leq^{\prime} of a reflexive relation RR on a set QQ. For any RR-bad QQ-array f0f_{0} (given as a function on a block), there is an RR-bad QQ-array f\ensurestackMath\stackengine.5exUcFFSf0f\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{0} that admits no RR-bad QQ-array gfg\lessdot^{\prime}f.

The given version does indeed coincide with the one by Laver [6], except that the latter requires RR to be a quasi order and works with barriers at the place of blocks. To see that our formulation in terms of blocks entails the one in terms of barriers, we recall that any block BB admits a barrier BBB^{\prime}\subseteq B, provably in 𝖠𝖳𝖱0\mathsf{ATR}_{0} (see [8] and compare [1] for a related result over the much weaker theory 𝖶𝖪𝖫0\mathsf{WKL}_{0}). Given a minimal bad array f\ensurestackMath\stackengine.5exUcFFSf0f\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{0} on a block, we thus find a restriction f\ensurestackMath\stackengine.5exUcFFSff^{\prime}\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f that is defined on a barrier. By the observations above, we get f\ensurestackMath\stackengine.5exUcFFSf0f^{\prime}\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{0} while a bad gfg\lessdot^{\prime}f^{\prime} would validate gfg\lessdot^{\prime}f, against the assumption that ff is minimal. Similarly, the following proposition remains valid when Laver’s version of the minimal bad array lemma is formulated in terms of barriers rather than blocks. Once we have proved that Π21\Pi^{1}_{2}-comprehension entails Laver’s version for blocks, the resulting circle of implications will ensure that the versions for blocks and barriers are equivalent. In the same way, we learn that the version for quasi orders is equivalent to the one for general reflexive relations. Let us note that the base theory 𝖱𝖢𝖠𝟢\mathsf{RCA_{0}} suffices to get the following result for blocks. We have chosen a stronger base theory to accommodate the aforementioned connection with barriers.

Proposition 3.2 (𝖠𝖳𝖱0\mathsf{ATR}_{0}).

Laver’s version of the minimal bad array lemma entails Simpson’s version.

Proof.

Given a bad QQ-array f0f_{0}, Laver’s version yields a bad f\ensurestackMath\stackengine.5exUcFFSf0f\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{0} that admits no bad gfg\lessdot^{\prime}f. One readily derives ff0f\leq^{\prime}f_{0} according to Definition 1.2. To conclude that Simpson’s version holds, we assume that g<fg<^{\prime}f is bad and derive a contradiction. Write f:BQf:B\to Q and g:CQg:C\to Q for blocks BB and CC, where we have CB\bigcup C\subseteq\bigcup B. Let us note that we do not immediately get the contradictory gfg\lessdot^{\prime}f, unless each tCt\in C admits an sBs\in B with sts\sqsubset t. For XX\subseteq\mathbb{N} and s[]<ωs\in[\mathbb{N}]^{<\omega}, let X/sX/s be the set of all numbers in XX that are larger than all elements of ss. We write sn=s{n}s^{\frown}n=s\cup\{n\} when we have n/sn\in\mathbb{N}/s. Let us now put

B0\displaystyle B_{0} ={sB|there is tC with tsC},\displaystyle=\{s\in B\,|\,\text{there is $t\in C$ with $t\sqsubseteq s\subset\textstyle\bigcup C$}\},
C\displaystyle C^{\prime} ={tC|there is sB with st}{sn|sB0 and nC/s}.\displaystyle=\{t\in C\,|\,\text{there is $s\in B$ with $s\sqsubset t$}\}\cup\{s^{\frown}n\,|\,s\in B_{0}\text{ and }n\in{\textstyle\bigcup C}/s\}.

It is straightforward to see that CC^{\prime} is a block with C=C\bigcup C^{\prime}=\bigcup C. We consider

g:CQwithg(t)={g(t)when there is sB with st,g(t)when t=sn with Cts.g^{\prime}:C^{\prime}\to Q\quad\text{with}\quad g^{\prime}(t)=\begin{cases}g(t)&\text{when there is $s\in B$ with $s\sqsubset t$},\\ g(t^{\prime})&\text{when $t=s^{\frown}n$ with $C\ni t^{\prime}\sqsubseteq s$}.\end{cases}

Let us note that gg and gg^{\prime} induce the same continuous function from [C]ω[{\bigcup C}]^{\omega} to QQ. In particular, gg^{\prime} is bad. To reach the desired contradiction, we verify gfg^{\prime}\lessdot^{\prime}f. First note that we have CBC^{\prime}\lessdot B as well as BC=B\cap C^{\prime}=\emptyset. Given BstCB\ni s\sqsubset t\in C^{\prime}, we pick a set X[C]ωX\in[{\bigcup C}]^{\omega} with tXt\sqsubset X. By considering the induced continuous functions, we see that g<fg<^{\prime}f entails

g(t)=g(X)=g(X)<f(X)=f(s),g^{\prime}(t)=g^{\prime}(X)=g(X)<^{\prime}f(X)=f(s),

just as Definition 3.1 requires. ∎

In the rest of this section, we show that Π21\Pi^{1}_{2}-comprehension suffices to carry out the proof of the minimal bad array lemma that has been given by R. Fraïssé [2].

Definition 3.3.

Let BB and CC be blocks with C\ensurestackMath\stackengine.5exUcFFSBC\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}B. We put

E(C,B,n)\displaystyle E(C,B,n) ={tB|t[0,n]C but tC}for n,\displaystyle=\{t\in B\,|\,t\subseteq[0,n]\cup\textstyle\bigcup C\text{ but }t\not\subseteq\textstyle\bigcup C\}\quad\text{for $n\in\mathbb{N}$},
M(C,B)\displaystyle M(C,B) ={n|there is tB with tC but tC and maxt=n}.\displaystyle=\{n\in\mathbb{N}\,|\,\text{there is }t\in B\text{ with }t\subseteq\textstyle\bigcup C\text{ but }t\not\in C\text{ and }\max t=n\}.

If we have M(C,B)M(C,B)\neq\emptyset or equivalently CBC\lessdot B, we put m(C,B)=minM(C,B)m(C,B)=\min M(C,B).

The following basic construction will be needed below. Let us point out that we have CE(C,B,n)=C\cap E(C,B,n)=\emptyset, as tE(C,B,n)t\in E(C,B,n) entails tCt\not\subseteq\bigcup C and hence tCt\notin C. The given proof is very close to the one by Fraïssé [2], except that we work with blocks rather than barriers. We have included it in order to keep our paper self-contained.

Lemma 3.4 (𝖱𝖢𝖠𝟢\mathsf{RCA_{0}}).

Assume that CBC\lessdot B are blocks and that we have nm(C,B)n\leq m(C,B). Then D:=CE(C,B,n)D:=C\cup E(C,B,n) is a block with C\ensurestackMath\stackengine.5exUcFFSDBC\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}D\lessdot B and m(C,B)=m(D,B)m(C,B)=m(D,B).

Proof.

We start by showing that DD is a block. It is clear that DC\bigcup D\supseteq\bigcup C is infinite. Towards a contradiction, we assume that we have tst\sqsubset s with s,tDs,t\in D. Then ss and tt are not both from CC and not both from E(C,B,n)BE(C,B,n)\subseteq B. For tE(C,B,n)t\in E(C,B,n) we would get stCs\supset t\not\subseteq\bigcup C and hence sCs\notin C. In the remaining case, we have tCt\in C and sE(C,B,n)s\in E(C,B,n). Due to CBC\lessdot B, we then find an sBs^{\prime}\in B with stsBs^{\prime}\sqsubseteq t\sqsubset s\in B, against the assumption that BB is a block. Thus DD is a \sqsubseteq-antichain. Now consider an infinite set XDX\subseteq\bigcup D. We want to prove that XX has an initial segment in DD. In view of XBX\subseteq\textstyle\bigcup B, we may pick a tBt\in B that validates tXt\sqsubset X. First assume that we have tCt\not\subseteq\bigcup C. By the definition of E(C,B,n)E(C,B,n), all elements of D\C\bigcup D\backslash\bigcup C are bounded by nn. Hence we have t[0,n]Ct\subseteq[0,n]\cup\bigcup C, so that we get tE(C,B,n)Dt\in E(C,B,n)\subseteq D. Now assume we have tCt\subseteq\bigcup C. If we have tCDt\in C\subseteq D, then we are done. So let us assume tCt\notin C. The latter entails nm(C,B)maxtn\leq m(C,B)\leq\max t. Due to D[0,n]C\bigcup D\subseteq[0,n]\cup\bigcup C, we get XtCCX\subseteq t\cup\bigcup C\subseteq\bigcup C. As CC is a block, we find an sCDs\in C\subseteq D with sXs\sqsubset X.

It is straightforward to see that CBC\lessdot B entails C\ensurestackMath\stackengine.5exUcFFSDBC\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}D\lessdot B. To complete the proof, we show M(C,B)=M(D,B)M(C,B)=M(D,B). For maxtM(C,B)\max t\in M(C,B) with tB\Ct\in B\backslash C but tCt\subseteq\bigcup C, we obtain tE(C,B,n)t\notin E(C,B,n) and hence tDt\notin D but tDt\subseteq\bigcup D, which yields maxtM(D,B)\max t\in M(D,B). Conversely, consider maxsM(D,B)\max s\in M(D,B) with sB\Ds\in B\backslash D but sD[0,n]Cs\subseteq\bigcup D\subseteq[0,n]\cup\bigcup C. As sCs\not\subseteq\bigcup C would entail sE(C,B,n)Ds\in E(C,B,n)\subseteq D, we must have sCs\subseteq\bigcup C. Since we also have sDCs\not\in D\supseteq C, we obtain maxsM(C,B)\max s\in M(C,B). ∎

We conclude our paper with the following result. Together with Theorem 2.4 and Proposition 3.2, it yields a circle of implications that establishes Theorem 1.3.

Theorem 3.5 (𝖱𝖢𝖠0\mathsf{RCA}_{0}).

The principle of Π21\Pi^{1}_{2}-comprehension entails Laver’s version of the minimal bad array lemma.

Proof.

Recall that Π21\Pi^{1}_{2}-comprehension is equivalent to the principle that any subset of \mathbb{N} is contained in a countable coded β2\beta_{2}-model (see Theorems VII.6.9 and VII.7.4 of [19]). Here a β2\beta_{2}-model is an ω\omega-model for which Π21\Pi^{1}_{2}-statements are absolute. With the help of β2\beta_{2}-models, it is straightforward to formalize the proof of the minimal bad array lemma that is given by Fraïssé [2, 7.3.5]. We provide details for the convenience of the reader and also to verify that the argument applies to general reflexive relations rather than just to quasi orders.

For a partial ranking \leq^{\prime} of a reflexive binary relation RR on a set QQ, we consider an RR-bad array f0:B0Qf_{0}:B_{0}\to Q on a block. Let us pick a β2\beta_{2}-model \mathcal{M} that contains f0f_{0}. The statement that a given array gg is minimal RR-bad has complexity Π21\Pi^{1}_{2}, as it takes a Π11\Pi^{1}_{1}-formula to assert that the domain of a potential fgf\lessdot^{\prime}g is a block. Assuming that the minimal bad array lemma fails, any RR-bad array gg\in\mathcal{M} with g\ensurestackMath\stackengine.5exUcFFSf0g\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{0} will thus admit an RR-bad array fgf\lessdot^{\prime}g with ff\in\mathcal{M}. We intend to build a sequence of RR-bad arrays fi\ensurestackMath\stackengine.5exUcFFSf0f_{i}\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{0} with fi+1fif_{i+1}\lessdot^{\prime}f_{i} such that the following additional properties hold, where we write BiB_{i} for the block on which fif_{i} is defined:

  1. (i)

    for any nBin\in\bigcup B_{i} with nm(Bi+1,Bi)n\leq m(B_{i+1},B_{i}) we have nBi+1n\in\bigcup B_{i+1},

  2. (ii)

    for any RR-bad g:CQg:C\to Q with gfig\lessdot^{\prime}f_{i} we have m(Bi+1,Bi)m(C,Bi)m(B_{i+1},B_{i})\leq m(C,B_{i}).

The indicated relation between fif_{i} and fi+1f_{i+1} has complexity Π21\Pi^{1}_{2}, by the same consideration as above. Now Π21\Pi^{1}_{2}-comprehension entails Σ21\Sigma^{1}_{2}-dependent choice but not Π21\Pi^{1}_{2}-choice (see Theorem VII.6.9 and Remark VII.6.3 of [19]). We will soon show that any fif_{i}\in\mathcal{M} admits an fi+1f_{i+1}\in\mathcal{M} with the given properties. Once this is achieved, we can lower the quantifier complexity of our Π21\Pi^{1}_{2}-condition by evaluating it in the β2\beta_{2}-model \mathcal{M}. Then Σ21\Sigma^{1}_{2}-dependent choice is more than sufficient to obtain the desired sequence of arrays fif_{i}. To be more economical, one can pick the smallest indices that represent suitable fif_{i} in the coded model \mathcal{M}.

As promised, we now show that each RR-bad fi\ensurestackMath\stackengine.5exUcFFSf0f_{i}\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{0} in \mathcal{M} admits an fi+1f_{i+1}\in\mathcal{M} with the properties above. We have already observed that \mathcal{M} contains an RR-bad array f:BQf:B\to Q with ffif\lessdot^{\prime}f_{i}, at least if there is no minimal RR-bad array below f0f_{0}. We pick such an ff with m(B,Bi)m(B,B_{i}) as small as possible, in the sense that \mathcal{M} contains no RR-bad array g:CQg:C\to Q with gfig\lessdot^{\prime}f_{i} and m(C,Bi)<m(B,Bi)m(C,B_{i})<m(B,B_{i}). This minimality condition extends to gg that do not lie in \mathcal{M}, by Π21\Pi^{1}_{2}-absoluteness. So ff validates (ii) but possibly not (i). To satisfy the latter, we modify ff. The challenge is that (ii) needs to be preserved. We will see that this holds for fi+1:Bi+1Qf_{i+1}:B_{i+1}\to Q with

Bi+1\displaystyle B_{i+1} =BE(B,Bi,m(B,Bi)),\displaystyle=B\cup E(B,B_{i},m(B,B_{i})),
fi+1(t)\displaystyle f_{i+1}(t) ={f(t)if tB,fi(t)if tE(B,Bi,m(B,Bi)).\displaystyle=\begin{cases}f(t)&\text{if }t\in B,\\ f_{i}(t)&\text{if }t\in E(B,B_{i},m(B,B_{i})).\end{cases}

Let us first note that fi+1f_{i+1} lies in \mathcal{M} by absoluteness, since it is arithmetically definable from ff and fif_{i}. By the previous lemma, Bi+1B_{i+1} is a block with B\ensurestackMath\stackengine.5exUcFFSBi+1BiB\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}B_{i+1}\lessdot B_{i} and m:=m(Bi+1,Bi)=m(B,Bi)m:=m(B_{i+1},B_{i})=m(B,B_{i}). Due to this last equality, (ii) remains valid. To see that we get fi+1fif_{i+1}\lessdot^{\prime}f_{i}, we note that BistBi+1B_{i}\ni s\sqsubset t\in B_{i+1} forces tBt\in B, since we have E(B,Bi,m)BiE(B,B_{i},m)\subseteq B_{i}. It remains to show that fi+1f_{i+1} is RR-bad and validates (i).

To show that fi+1f_{i+1} is RR-bad, we argue by contradiction. Assume that we have s,tBi+1s,t\in B_{i+1} with sts\vartriangleleft t and fi+1(s)Rfi+1(t)f_{i+1}(s)\,R\,f_{i+1}(t). Since ff and fif_{i} are bad, we cannot have s,tBs,t\in B or s,tBis,t\in B_{i}. Concerning the two other cases, we first assume sBs\in B and tE(B,Bi,m)Bit\in E(B,B_{i},m)\subseteq B_{i}. Given that we have BBiB\lessdot B_{i} and BE(B,Bi,m)=B\cap E(B,B_{i},m)=\emptyset, we find an sBis^{\prime}\in B_{i} with sss^{\prime}\sqsubset s. We can conclude mmaxsm\leq\max s^{\prime} due to the definition of m=m(B,Bi)m=m(B,B_{i}). From t[0,m]Bt\subseteq[0,m]\cup\bigcup B and ssBs^{\prime}\subseteq s\subseteq\bigcup B we now get tBt\subseteq\bigcup B and hence tE(B,Bi,m)t\notin E(B,B_{i},m), against our assumption. Let us now assume that we have sE(B,Bi,m)Bis\in E(B,B_{i},m)\subseteq B_{i} and tBt\in B. We find a tBit^{\prime}\in B_{i} with ttt^{\prime}\sqsubseteq t and thus sts\vartriangleleft t^{\prime}. In view of ffif\lessdot^{\prime}f_{i}, we obtain

fi(s)=fi+1(s)Rfi+1(t)=f(t)fi(t).f_{i}(s)=f_{i+1}(s)\,R\,f_{i+1}(t)=f(t)\leq^{\prime}f_{i}(t^{\prime}).

Crucially, we get fi(s)Rfi(t)f_{i}(s)\,R\,f_{i}(t^{\prime}) by our definition of partial rankings for relations that need not be transitive. We now have a contradiction since fif_{i} is bad.

In order to establish (i), we consider an nBin\in\bigcup B_{i} with nm=m(Bi+1,Bi)n\leq m=m(B_{i+1},B_{i}). Due to BBi+1B\subseteq B_{i+1}, we are done unless we have nBn\notin\bigcup B. In the latter case, we can find a tBit\in B_{i} with nt[0,m]Bn\in t\subseteq[0,m]\cup\bigcup B. We thus have tE(B,Bi,m)Bi+1t\in E(B,B_{i},m)\subseteq B_{i+1} and in particular nBi+1n\in\bigcup B_{i+1}, as desired.

Under the assumption that no minimal RR-bad array below f0f_{0} exists, we have shown that there is a sequence of RR-bad arrays fi+1fi\ensurestackMath\stackengine.5exUcFFSf0f_{i+1}\lessdot^{\prime}f_{i}\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{0} that validate (i) and (ii) above. With the help of this sequence, we will now show that there is a minimal RR-bad array below f0f_{0} after all. Let us first prove some properties of the function p:p:\mathbb{N}\to\mathbb{N} with p(i):=m(Bi+1,Bi)p(i):=m(B_{i+1},B_{i}).

Claim.

For every nn\in\mathbb{N} we have |p1({n})|2n|p^{-1}(\{n\})|\leq 2^{n}.

Proof of the claim.

As there are only 2n2^{n} sets t[0,n]t\subseteq[0,n] with maxt=n\max t=n, we need only show that the same tt cannot witness both p(i)=np(i)=n and p(j)=np(j)=n for i<ji<j. Assume p(i)=np(i)=n is witnessed by tBit\in B_{i} with tBi+1t\subseteq\bigcup B_{i+1} but tBi+1t\notin B_{i+1}. Since Bi+1B_{i+1} is a block, we find a tBi+1t^{\prime}\in B_{i+1} that is \sqsubset-comparable with tt. In view of Bi+1BiB_{i+1}\lessdot^{\prime}B_{i}, we may consider an sBis\in B_{i} with sts\sqsubseteq t^{\prime}. This forces s=ts=t, so that we get ttt\sqsubset t^{\prime}. Now if tt did witness p(j)=np(j)=n, we would have tBjt\in B_{j}. Due to Bj\ensurestackMath\stackengine.5exUcFFSBi+1B_{j}\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}B_{i+1}, we could pick an sBi+1s^{\prime}\in B_{i+1} with sttBi+1s^{\prime}\sqsubseteq t\sqsubset t^{\prime}\in B_{i+1}, against the fact that Bi+1B_{i+1} is a block. ∎

We will also need the following.

Claim.

The function pp is non-decreasing.

Proof of the claim.

Write p(i+1)=maxsp(i+1)=\max s with sBi+1\Bi+2s\in B_{i+1}\backslash B_{i+2} and sBi+2s\subseteq\bigcup B_{i+2}. In view of Bi+1BiB_{i+1}\lessdot B_{i}, we find a tBit\in B_{i} with tst\sqsubseteq s. We show how to conclude in either of two cases. First assume tst\sqsubset s. We then have tsBi+1t\subseteq s\subseteq\bigcup B_{i+1} but tBi+1t\notin B_{i+1}, which yields maxtM(Bi+1,Bi)\max t\in M(B_{i+1},B_{i}) and hence

p(i)=m(Bi+1,Bi)maxt<maxs=p(i+1).p(i)=m(B_{i+1},B_{i})\leq\max t<\max s=p(i+1).

Now assume that we have t=st=s. This yields tBi\Bi+2t\in B_{i}\backslash B_{i+2} and tBi+2t\subseteq\bigcup B_{i+2}, so that we obtain maxtM(Bi+2,Bi)\max t\in M(B_{i+2},B_{i}) and thus

p(i)=m(Bi+1,Bi)m(Bi+2,Bi)maxtmaxs=p(i+1),p(i)=m(B_{i+1},B_{i})\leq m(B_{i+2},B_{i})\leq\max t\leq\max s=p(i+1),

where the first inequality holds due to (ii) above. ∎

Let HH be the range of pp. The first claim above implies that HH is infinite. Using the second claim, we establish HBiH\subseteq\bigcup B_{i} for any ii\in\mathbb{N}. In view of Bi+1BiB_{i+1}\lessdot B_{i}, it suffices to show that p(j)Hp(j)\in H is contained in Bi\bigcup B_{i} for iji\geq j. First note that we have p(j)=m(Bj+1,Bj)=maxtp(j)=m(B_{j+1},B_{j})=\max t for some tBjt\in B_{j}, which yields p(j)tBjp(j)\in t\subseteq\bigcup B_{j}. Inductively, we assume p(j)Bip(j)\in\bigcup B_{i} with iji\geq j. Here the latter ensures that we have p(j)p(i)=m(Bi+1,Bi)p(j)\leq p(i)=m(B_{i+1},B_{i}), so that we get p(j)Bi+1p(j)\in\bigcup B_{i+1} by (i) above. Let us now consider the set

C={t[H]<ω|there is i with tBj for all ji}.C=\{t\in[H]^{<\omega}\,|\,\text{there is $i\in\mathbb{N}$ with $t\in B_{j}$ for all\leavevmode\nobreak\ $j\geq i$}\}.

Since all the BjB_{j} are blocks, it is immediate that CC is a \sqsubset-antichain. To show that it is a block, we consider an arbitrary set X[H]ωX\in[H]^{\omega}. Due to HBiH\subseteq\bigcup B_{i}, we may pick an siBis_{i}\in B_{i} with siXs_{i}\sqsubset X for each ii\in\mathbb{N}. Since we have fi+1fif_{i+1}\lessdot^{\prime}f_{i}, we get si+1sis_{i+1}\sqsubseteq s_{i} and fi+1(si+1)fi(si)f_{i+1}(s_{i+1})\leq^{\prime}f_{i}(s_{i}), where the latter is an equality precisely for si=si+1s_{i}=s_{i+1}. Given that \leq^{\prime} is well-founded, the values fi(si)f_{i}(s_{i}) must indeed stabilize. So we find an ii\in\mathbb{N} such that jij\geq i entails si=sjBjs_{i}=s_{j}\in B_{j}. This yields siCs_{i}\in C, as needed to conclude that CC is a block.

Now consider g:CQg:C\to Q with g(t)=fi(t)g(t)=f_{i}(t) for tBiCt\in B_{i}\cap C, which is well-defined by the previous paragraph. It is straightforward to show that gg is RR-bad and that we have g\ensurestackMath\stackengine.5exUcFFSfig\mathrel{\ensurestackMath{\stackengine{-.5ex}{\lessdot}{-}{U}{c}{F}{F}{S}}}^{\prime}f_{i} for any ii\in\mathbb{N} (in fact we get gfig\lessdot^{\prime}f_{i}). To complete the proof, we show that gg is minimal. Aiming at a contradiction, we assume that h:DQh:D\to Q is RR-bad with hgh\lessdot^{\prime}g. By the paragraph after Definition 3.1, we get hfih\lessdot^{\prime}f_{i} for any ii\in\mathbb{N}. Write m(D,C)=maxtm(D,C)=\max t with tCt\in C and tDt\subseteq\bigcup D but tDt\notin D. When jj\in\mathbb{N} is sufficiently large, we have both tBjt\in B_{j} and m(D,C)<p(j)m(D,C)<p(j). We get

m(D,Bj)maxt=m(D,C)<p(j)=m(Bj+1,Bj),m(D,B_{j})\leq\max t=m(D,C)<p(j)=m(B_{j+1},B_{j}),

which yields a contradiction with property (ii) from above. ∎

References

  • [1] Peter Cholak, Alberto Marcone, and Reed Solomon, Reverse mathematics and the equivalence of definitions for well and better quasi-orders, The Journal of Symbolic Logic 69 (2004), no. 3, 683–712.
  • [2] Roland Fraïssé, Theory of relations, North-Holland, Amsterdam, 1986.
  • [3] Harvey Friedman, Some systems of second order arithmetic and their use, Proceedings of the International Congress of Mathematicians, Vancouver 1974 (Ralph Duncan James, ed.), vol. 1, Canadian Mathematical Congress, 1975, pp. 235–242.
  • [4] Joseph Kruskal, The theory of well-quasi-ordering: A frequently discovered concept, Journal of Combinatorial Theory, Series A 13 (1972), no. 3, 297–305.
  • [5] Richard Laver, On Fraïssé’s order type conjecture, Annals of Mathematics 93 (1971), no. 1, 89–111.
  • [6] by same author, Better-quasi-orderings and a class of trees, Studies in Foundations and Combinatorics (Gian-Carlo Rota, ed.), Advances in Mathematics and Supplementary Studies, vol. 1, Academic Press, 1978, pp. 31–48.
  • [7] Alberto Marcone, Foundations of BQO theory, Transactions of the American Mathematical Society 345 (1994), no. 2, 641–660.
  • [8] by same author, On the logical strength of Nash-Williams’ theorem on transfinite sequences, Logic: From Foundations to Applications (W. Hodges, M. Hyland, C.Steinhorn, and J.Truss, eds.), Oxford University Press, 1996, pp. 327–351.
  • [9] by same author, WQO and BQO theory in subsystems of second order arithmetic, Reverse Mathematics 2001 (Stephen Simpson, ed.), Lecture Notes in Logic, vol. 21, Cambridge University Press, 2005, pp. 303–330.
  • [10] Carl Mummert and Stephen Simpson, Reverse Mathematics and Π21\Pi^{1}_{2} Comprehension, The Bulletin of Symbolic Logic 11 (2005), no. 4, 526–533.
  • [11] Crispin St. J. A. Nash-Williams, On well-quasi-ordering infinite trees, Mathematical Proceedings of the Cambridge Philosophical Society 61 (1965), 697–720.
  • [12] by same author, On better-quasi-ordering transfinite sequences, Mathematical Proceedings of the Cambridge Philosophical Society 64 (1968), 273–290.
  • [13] Maurice Pouzet, Graphs and posets with no infinite independent set, Finite and Infinite Combinatorics in Sets and Logic (N. Sauer, R. Woodrow, and B. Sands, eds.), NATO ASI Series C, vol. 411, Springer, 1993, pp. 313–335.
  • [14] Richard Rado, Partial well-ordering of sets of vectors, Mathematika 1 (1954), 89–95.
  • [15] Michael Rathjen, The art of ordinal analysis, Proceedings of the International Congress of Mathematicians, Madrid 2006 (Marta Sanz-Solé, Javier Soria, Juan Luis Varona, and Joan Verdera, eds.), vol. 2, European Mathematical Society, 2006, pp. 45–69.
  • [16] Neil Robertson and Paul D. Seymour, Graph minors. XX. Wagner’s conjecture, Journal of Combinatorial Theory, Series B 92 (2004), no. 2, 325–357.
  • [17] Saharon Shelah, Better quasi-orders for uncountable cardinals, Israel Journal of Mathematics 42 (1982), no. 3, 177–226.
  • [18] Stephen Simpson, Bqo theory and Fraïssé’s conjecture, chapter in the book ‘Recursive Aspects of Recursive Set Theory’ by R. Mansfield and G. Weitkamp, Oxford University Press, 1985, pp. 124–138.
  • [19] Stephen G. Simpson, Subsystems of second order arithmetic, Perspectives in Logic, Cambridge University Press, 2009.